Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-m9kch Total loading time: 0 Render date: 2024-05-20T17:46:21.729Z Has data issue: false hasContentIssue false

Bibliography

Published online by Cambridge University Press:  04 November 2021

Paul F. Hudson
Affiliation:
Universiteit Leiden
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
Flooding and Management of Large Fluvial Lowlands
A Global Environmental Perspective
, pp. 298 - 325
Publisher: Cambridge University Press
Print publication year: 2021

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., Maurice-Bourgoin, L., Dunne, T., Montgomery, D. R., Nittrouer, C. A., and Guyot, J.-L. 2003. Episodic sediment accumulation on Amazonian flood plains influenced by El Niño/Southern Oscillation. Nature 425, 493497.Google Scholar
Aalto, R., Lauer, J. W., and Dietrich, W. E. 2008. Spatial and temporal dynamics of sediment accumulation and exchange along Strickland River floodplains (Papua New Guinea) over decadal to centennial timescales. Journal of Geophysical Research: Earth Surface 113, F01S04. doi: 10.1029/2006JF0006277.Google Scholar
Abernethy, B., and Rutherfurd, I. D. 2001. The distribution and strength of riparian tree roots in relation to riverbank reinforcement. Hydrological Processes 15, 6379.CrossRefGoogle Scholar
Abizaid, C. 2005. An anthropogenic meander cutoff along the Ucayali River, Peruvian Amazon. Geographical Review 95, 122135.Google Scholar
Adams, J. 1980. Active tilting of the United States mid-continent: geodetic and geomorphic evidence. Geology 8, 442446.Google Scholar
Adams, P. N., Slingerland, R. L., and Smith, N. D. 2004. Variations in natural levee morphology in anastomosed channel floodplain complexes. Geomorphology 61, 127142. doi: 10.1016/j.geomorph.2003.10.005.Google Scholar
Adamson, M. 2018. Flood risk management in Europe: the EU “Floods” directive and a case study of Ireland. International Journal of River Basin Management 16, 261272.Google Scholar
Aerts, J. C. J. H., Botzen, W. J., Clarke, K. C., et al. 2018. Integrating human behaviour dynamics into flood disaster risk assessment. Nature Climate Change 8, 193199. doi: 10.1038/s41558–018-0085-1.Google Scholar
Ahmed, A. A., and Ismail, U. H. A. E. 2008. Sediment in the Nile River System. UNESCO International Hydrological Programme, International Sediment Initiative. Khartoum, Sudan.Google Scholar
Ahn, J., Yang, C. T., Boyd, P. M., Pridal, D. B., and Remus, J. I. 2013. Numerical modeling of sediment flushing from Lewis and Clark Lake. International Journal of Sediment Research 28, 182193.Google Scholar
Alcamo, J. D., Van Vuuren, D. P., Cramer, W. et al. 2005. Changes in ecosystem services and their drivers across the scenarios. In Sinh, B. T., Hammond, A., Field, C. et al. (eds.), Ecosystems and Human Well‐Being: Scenarios. Island Press, Washington, DC, Chapter 9, pp. 297373.Google Scholar
Alexander, J., Fielding, C. R., and Pocock, G. D. 1999. Flood deposits of the Burdekin River, tropical north Queensland, Australia. In Marriott, S. G., and Alexander, J. (eds.), Floodplains: Interdisciplinary Approaches, vol. 163. Geological Society London Special Publication, London, pp. 2740.Google Scholar
Alexander, J. S., Wilson, R. C., and Green, W. R. 2012. A brief history and summary of the effects of river engineering and dams on the Mississippi River system and delta. U.S. Geological Survey Circular, 1375.Google Scholar
Alexander, J. S., Jacobson, R. B., and Rus, D. L. 2013. Sediment transport and deposition in the lower Missouri River during the 2011 flood. U.S. Geological Survey Professional Paper 1798–F. doi: 10.3133/PP1798F.Google Scholar
Alexander, M. A., Kilbourne, K. H., and Nye, J. A. 2014. Climate variability during warm and cold phases of the Atlantic Multidecadal Oscillation (AMO) 1871–2008. Journal of Marine Systems 143, 1426.Google Scholar
Allen, Y., Kimmel, K. M., and Constant, G. C. 2014. Alligator gar movement and water quality patterns on the St. Catherine Creek National Wildlife Refuge floodplain. U.S. Fish and Wildlife Service, Conservation Office Report, Baton Rouge, LA.Google Scholar
Allison, M. A., and Meselhe, E. A. 2010. The use of large water and sediment diversions in the lower Mississippi River (Louisiana) for coastal restoration. Journal of Hydrology 387, 346360.Google Scholar
Allison, M. A., Kuehl, S. A., Martin, T. C., and Hassan, A. 1998. The importance of floodplain sedimentation for river sediment budgets and terrigenous inputs to the ocean: insights from the Brahmaputra–Jamuna rivers. Geology 26, 175178.2.3.CO;2>CrossRefGoogle Scholar
Allison, M. A., Vosburg, B. M., Ramirez, M. T., and Meselhe, E. A. 2013. Mississippi River channel response to the Bonnet Carré Spillway opening in the 2011 flood and its implications for the design and operation of river diversions. Journal of Hydrology 477, 104118.Google Scholar
Allison, M. A., Törnqvist, T., Amelung, F., et al. 2016. Global risks and research priorities for coastal subsidence. Eos 97. doi: 10.1029/2016EO055013.Google Scholar
Amer, R., Kolker, A. S., and Muscietta, A. 2017. Propensity for erosion and deposition in a deltaic wetland complex: implications for river management and coastal restoration. Remote Sensing of Environment 199, 3350.Google Scholar
American Rivers. 2020. American Rivers Dam Removal Database. https://figshare.com/articles/dataset/American_Rivers_Dam_Removal_Database/5234068 (accessed May 5, 2020).Google Scholar
American Society of Civil Engineers (ASCE). 2007. The New Orleans hurricane protection system: what went wrong and why. American Society of Civil Engineers, Hurricane Katrina External Review Panel.Google Scholar
Amissah, G. J., Kiss, T., and Fiala, K. 2018. Morphological evolution of the lower Tisza River (Hungary) in the 20th century in response to human interventions. Water 10, 884.CrossRefGoogle Scholar
Amoros, C., Elger, A., Dufour, S. et al. 2005. Flood scouring and groundwater supply in rehabilitated side-channels of the Rhône River, France: Sedimentation and aquatic vegetation responses. Large Rivers, Archives für Hydrobiologie Suppl., 15, 14.Google Scholar
ANCOLD. 2020. Registrar of large dams in Australia. Australian National Committee on Large Dams. www.ancold.org.au/ (accessed May 4, 2020).Google Scholar
Andrews, J. 1938. Goyder’s line: a vanished frontier. Australian Geographer 3, 3236. doi: 10.1080/00049183808702187.CrossRefGoogle Scholar
Annandale, G. W., Morris, G. L., and Karki, P. 2016. Extending the Life of Reservoirs: Sustainable Sediment Management for Dams and Run-of-River Hydropower. Directions in Development – Energy and Mining. World Bank, Washington, DC.Google Scholar
Annual Report of the Chief of Engineers. 1890. Improvement of the navigation of Red River, Louisiana, and of certain rivers in Louisiana, Texas, Mississippi, Arkansas, and Tennessee; and water-gauges on the Mississippi and its principal tributaries. Report of J. H. Willard, Appendix W.Google Scholar
Annual Report of the Chief of Engineers. 1904. Western rivers: improvement of certain rivers and waterways in Louisiana, Texas, Arkansas, Indian Territory, and Mississippi Tributary to Mississippi River. Annual Reports for the War Department for Fiscal Year ending in June 30, 1904, part 1, volume 5, Washington, DC, pp. 382–415.Google Scholar
Anonymous. 1828. Letter to members of the Board of Internal Improvements on the importance of removing the raft on the Atchafalaya River. United States. Works Progress Administration of Louisiana, March 4, 1928, p. 3, c. 2–4.Google Scholar
Anthony, E., Brunier, G., Besset, M., et al. 2015. Linking rapid erosion of the Mekong River delta to human activities. Scientific Reports 5, 14745. doi: 10.1038/srep14745.Google Scholar
Archer, M. W., Pracheil, B. M., Otto, A. E., and Pegg, M. E. 2019. Fish community response to in-channel woody debris in a channelized river system. Journal of Freshwater Ecology 34, 351362.Google Scholar
Areneo, D., and Villalba, R. 2014. Variability in the annual cycle of the Río Atuel streamflows and its relationship with tropospheric circulation. International Journal of Climatology 35, 29482967.Google Scholar
Armstrong, C., Mohrig, D., Hess, T., George, T., and Straub, T. M. 2013. Influence of growth faults on coastal fluvial systems: examples from the late Miocene to Recent Mississippi River Delta. Sedimentary Geology 301, 120132.Google Scholar
Arnaud, F., Piégay, H., Schmitt, L., Rollet, A. J., Ferrier, V., and Béal, D. 2015. Historical geomorphic analysis (1932–2011) of a by-passed river reach in process-based restoration perspectives: the Old Rhine downstream of the Kembs diversion dam (France, Germany). Geomorphology 236, 163177.Google Scholar
Arroyave, V. J. A., and Crosato, A. 2010. Effects of river floodplain lowering and vegetation cover. Water Management 163, 457467. doi: 10.1680/wama.900023.Google Scholar
ASDSO. 2016. The cost of rehabilitating our nation’s dams: a methodology, estimate and proposed funding mechanisms. Prepared by a Task Committee of the Association of State Dam Safety Officials. https://damsafety.org/state-performance (accessed 5 April 2020).Google Scholar
Ashmore, P. J. 1991. How do gravel-bed rivers braid? Canadian Journal of Earth Sciences 28, 326341.Google Scholar
Ashmore, P. J. 1996. Mid-Channel bar growth and its relationship to local flow strength and direction. Earth Surface Processes and Landforms 21, 103123.Google Scholar
Aslan, A., and Autin, W. J. 1998. Holocene flood-plain soil formation in the southern lower Mississippi Valley: implications for interpreting alluvial paleosols. Geological Society of America Bulletin 110, 433449.Google Scholar
Aslan, A., and 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, 800815.Google Scholar
Aslan, A., and Blum, M. D. 1999. Contrasting styles of Holocene avulsion, Texas Gulf Coastal Plain, U.S.A. In Smith, N. D., and Rogers, J. (eds.), Fluvial Sedimentology VI. International Association of Sedimentologists, Special Publication 28. Blackwell, Oxford, pp. 293308.Google Scholar
Aslan, A., White, W. A., Warne, A. G., and Andguevara, E. J. 2003. Holocene evolution of the western Orinoco Delta, Venezuela. Geological Society of America Bulletin 115, 479498.2.0.CO;2>CrossRefGoogle Scholar
Aslan, A., Autin, W. J., and Blum, M. D. 2005. Late Holocene avulsion history of the Mississippi River, south Louisiana, U.S.A. Journal of Sedimentary Research 75, 648662.Google Scholar
Aspen Institute. 2002. Dam Removal: A New Option for a New Century. Program on Energy, the Environment, and the Economy.Google Scholar
Asselman, N. E. M. 1999. Suspended sediment dynamics in a large drainage basin: the River Rhine. Hydrological Processes 13, 14371450.Google Scholar
Asselman, N. E. M., and Middelkoop, H. 1998. Temporal variability of contemporary floodplain sedimentation in the Rhine–Meuse delta, The Netherlands. Earth Surface Processes and Landforms 23, 595609.Google Scholar
Asselman, N. E. M., Middelkoop, H., and Van Dijk, P. M. 2003. The impact of changes in climate and land use on soil erosion, transport and deposition of suspended sediment in the River Rhine. Hydrological Processes 17, 32253244.Google Scholar
Auerbach, D. A., Deisenroth, D. B., McShane, R. R., McCluney, K. E., and Poff, N. L. 2014. Beyond the concrete: accounting for ecosystem services from free-flowing rivers. Ecosystem Services 10, 15.Google Scholar
Auerbach, L. W., Goodbred, S. L. Jr., Mondal, D. R., et al. 2015. Flood risk of natural and embanked landscapes on the Ganges–Brahmaputra tidal delta plain. Nature Climate Change 5, 153157. doi: 10.1038/nclimate2472.Google Scholar
Autin, W. J., Burns, S. F., Miller, B. J., Saucier, R. T., and Snead, J. I. 1991. Quaternary geology of the lower Mississippi Valley. In Morrison, R. B. (ed.), Quaternary Nonglacial Geology, vol. K-2. Geological Society of America, The Geology of North America, Boulder, CO, pp. 547581.Google Scholar
Avakyan, A. B. 1998. Volga-Kama cascade reservoirs and their optimal use. Lakes and Reservoirs Research and Management 3, 113121. doi: 10.1111/j.1440-1770.1998.tb00038.x.Google Scholar
AWA. 2020. Large dams fact sheet. Australian Water Association. www.awa.asn.au/ (accessed May 7, 2020).Google Scholar
Baker, R. G., Schwert, D. P., Bettis, E. A., and Chumbly, C. A. 1993. Impact of Euro-American settlement on a riparian landscape in northeast Iowa, Midwestern USA: an integrated approach based on historical evidence, floodplain sediments, fossil pollen, plant macrofossils and insects. The Holocene 3, 314323.Google Scholar
Bakkenist, S., and Flos, S. 2015. Dijkinspectie met drones: Verkenning van de Mogeikheden voor het gebruik van drones bij inspectie en het beheer van waterkeringen, STOWA Report 2015-09, Amersfoort, NL.Google Scholar
Baldassarre, G. Di and Castellarin, A. 2009. Analysis of the effects of levee heightening on flood propagation: example of the River Po, Italy. Hydrological Science Journal 54, 10071017.CrossRefGoogle Scholar
Bank Swallow Technical Advisory Committee (BSTAC). 2013. Bank swallow (Riparia riparia) conservation strategy for the Sacramento River Watershed, California. Version 1.0. www.sacramentoriver.org/bans/ (accessed September 15, 2018).Google Scholar
Baptist, M. J., Penning, W. E., Duel, H., et al. 2004. Assessment of the effects of cyclic floodplain rejuvenation on flood levels and biodiversity along the Rhine River. River Research and Applications 20, 285297.Google Scholar
Barnett, J. F. Jr. 2017. Beyond Control: The Mississippi River’s New Channel to the Gulf of Mexico. University Press of Mississippi, Oxford.Google Scholar
Barras, J. A., Bernier, J. C., and Morton, R. A. 2008. Land area change in coastal Louisiana – a multidecadal perspective (from 1956 to 2006). U.S. Geological Survey Scientific Investigations Map 3019, scale 1:250,000, 14 p. pamphlet.Google Scholar
Barras, J. A., Beville, S., Britsch, D. et al. 2003. Historical and projected coastal Louisiana land changes: 1978–2050. U.S. Geological Survey, Open-File Report 2003–334.Google Scholar
Barry, J. A. 1997. Rising Tide: The Great Mississippi Flood of 1927 and How It Changed America. Simon and Schuster, New York.Google Scholar
Batalla, R. J. 2003. Sediment deficit in rivers caused by dams and instream gravel mining. A review with examples from NE Spain. Review Cuaternario y Geomorfologia 17, 7991.Google Scholar
Batalla, R. J., Kondolf, G. M., and Gomez, C. M. .2004. Reservoir-induced hydrological changes in the Ebro River basin, NE Spain. Journal of Hydrology 290, 117136.Google Scholar
Batalla, R. J., Vericat, D., and Martínez, T. I. 2006. River-channel changes downstream from dams in the lower Ebro River. Zeitschrift für Geomorphologie/Supplement 143, 115.Google Scholar
Bawden, G. W., Johnson, M. R., Kasmarek, M. C., Brandt, J., and Middleton, C. S. 2012. Investigation of Land Subsidence in the Houston-Galveston Region of Texas by using the Global Positioning System and Interferometric Synthetic Aperture Radar, 1993–2000. U.S. Geological Survey, Scientific Investigations Report, 2012–5211, 88pp.Google Scholar
Baxter, R. M. 1977. Environmental effects of dams and impoundments. Annual Review of Ecology and Systematics 8, 255283. doi: 10.1146/annurev.es.08.110177.001351.Google Scholar
Beach, T. 1994. The fate of eroded soil: sediment sinks and sediment budgets of agrarian landscapes in southern Minnesota, 1851–1988. Annals of the Association of American Geographers 84, 528.Google Scholar
Beach, T., Luzzadder-Beach, S., Dunning, N., and Cook, D. 2008. Human and natural impacts on fluvial and karst depressions of the Maya Lowlands. Geomorphology 101, 301331.Google Scholar
Becker, R. H., and Sultan, M. 2009. Land subsidence in the Nile Delta: inferences from radar interferometry. The Holocene 19, 949954.Google Scholar
Bednarek, A. T., and Hart, D. D. 2005. Modifying dam operations to restore rivers: ecological responses to Tennessee River dam mitigation. Ecological Applications 15, 9971008.Google Scholar
Beilfuss, R. D. 2018. The Zambezi Delta (Mozambique). In Finlayson, C. M., Everard, M., Irvine, K. et al. (eds.), The Wetland Book II: Distribution, Description, and Conservation. Springer, Dordrecht, pp. 12331242, https://doi.org/10.1007/978-94-007-4001-3_195.Google Scholar
Bellin, J. N. 1764. Suite du cours du fleuve St. Louis depuis la rivière d’Iberville jusq’à celle des Yasous, et les parties connues de la Rivière Rouge et la Rivière Noire. [Paris] [Map] Retrieved from the Library of Congress Geography and Map Division, Washington, DC. www.loc.gov/item/74693009/.Google Scholar
Belt, C. B. Jr. 1975. The 1973 flood and man’s constriction of the Mississippi River. Science 189, 681684.Google Scholar
Benedetti, M. M. 2003. Controls on overbank deposition in the Upper Mississippi River. Geomorphology 56, 271290.Google Scholar
Benito, G., and Hudson, P. F. 2010. Flood hazards: the context of fluvial geomorphology. In Alcántara-Ayala, I., and Goudie, A. (eds.), Geomorphological Hazards and Disaster Prevention. Cambridge University Press, Cambridge, pp. 111128.Google Scholar
Benito, G., and Machado, M. J. 2012. Floods in the Iberian Peninsula. In Kundzewicz, Z. W. (ed.), Changes of Flood Risk in Europe. IAHS Special Publication 10. IAHS Press and CRC Press, Balkema.Google Scholar
Benke, A. C., and Wallace, J. B. 2003. Influence of wood on invertebrate communities in streams and rivers. In Gregory, S. V., Boyer, K. L., and Gurnell, A. M. (eds.), The Ecology and Management of Wood in World Rivers. American Fisheries Society Symposium 37. American Fisheries Society, Bethesda, MD, pp. 149177.Google Scholar
Benke, A. C., Henry, R. L., Gillespie, D. M., and Hunter, R. J. 1985. Importance of snag habitat for animal production in southeastern streams. Fisheries 10, 813.Google Scholar
Benson, M. A., and Thomas, D. M. 1966. A definition of dominant discharge. Hydrological Sciences Journal 11, 7680. doi: 10.1080/02626666609493460.Google Scholar
Berendsen, H. J. A. 1982. De genese van het landschap in het zuiden van de provincie Utrecht, een fysisch-geografische studie. Dissertatie, Utrechtse Geografische Studies 25.Google Scholar
Berendsen, H. J. A. 1993. De ontwikkeling van het Nederlandse rivierengebied. Stichting Geologische Aktiviteiten 26 (2), 49–76.Google Scholar
Berendsen, H. J. A. 2008. Landschappelijk Nederland: De fysisch-geografische regio's, 4th Ed. Van Gorcum, Assen, NL.Google Scholar
Berendsen, H. J. A., and Stouthamer, E. 2001. Paleogeographic Development of the Rhine-Meuse Delta, the Netherlands. Koninklijke Van Gorcum. 3 colored maps, CD-Rom.Google Scholar
Berkowitz, J. F., Johnson, D. R., and Price, J. J. 2020. Forested wetland hydrology in a large Mississippi River tributary system. Wetlands 40, 11331148. doi: 10.1007/s13157-019-01249-5.Google Scholar
Berri, G. J., Ghietto, M. A., and García, N. O. 2002. The influence of ENSO in the flows of the Upper Paraná River of South America over the past 100 years. Journal of Hydrometeorology 2, 5765.Google Scholar
Best, J. L., and Darby, S. E. 2020. The pace of human-induced change in large rivers: stresses, resilience, and vulnerability to extreme events. One Earth 2, 510514.Google Scholar
Best, J. L., Ashworth, P. J., Sarker, M. H., and Roden, J. E. 2007. The Brahmaputra-Jamuna River, Bangladesh. In Gupta, A. (ed.), Large Rivers: Geomorphology and Management. John Wiley & Sons, Chichester, pp. 395433.Google Scholar
Bianchi, T. S., Allison, M. A., and Karl, D. M. 2009. Large-river delta-front estuaries as natural “recorders” of global environmental change. Proceedings of the National Academy of Sciences of the United States of America 106, 80858092.Google Scholar
Biedenharn, D. S., Thorne, C. R., and Watson, C. C. 2000. Recent morphological evolution of the lower Mississippi River. Geomorphology 34, 227250.Google Scholar
Biedenharn, D. S., Allison, M. A., Little, C. D. Jr., Thorne, C. R., and Watson, C. C. 2017. Large-Scale Geomorphic Change in the Mississippi River from St. Louis, MO, to Donaldsonville, LA, as revealed by Specific Gage Records. MRGandP Report No. 10. U. S. Army Corps of Engineers, Mississippi Valley Division, Vicksburg, MS. doi: 10.21079/11681/22744.Google Scholar
Biedenharn, D. S., Killgore, K. J., Little, C. D., Murphy, C. E. Jr., and Kleiss, B. A. 2018. Attributes of the lower Mississippi River Batture. MRGandP Technical Note No. 4. U.S. Army Corps of Engineers, Vicksburg, MS.Google Scholar
Biemans, H., Haddeland, I., Kabat, P., et al. 2011. Impact of reservoirs on river discharge and irrigation water supply during the 20th century (Fig. 3, Cumulative capacity of global dam storage over the 20th century). Water Resources Research 47, W03509. doi: 10.1029/2009WR008929.Google Scholar
Blamey, R. C., Ramos, A. M., Trigo, A. M., Tome, R., and Reason, C. J. C. 2017. The influence of atmospheric rivers over the South Atlantic on winter rainfall in South Africa. Journal of Hydrometeorology 11, 127142. doi: 10.1175/JHM-D-17-0111.1.Google Scholar
Blum, M. D., and Roberts, H. H. 2009. Drowning of the Mississippi Delta due to insufficient sediment supply and global sea-level rise. Nature Geoscience 2, 488491.Google Scholar
Blum, M. D., and Roberts, H. H. 2012. The Mississippi Delta region: past, present, and future. Annual Review of Earth and Planetary Sciences 40, 655683.Google Scholar
Blum, M. D., and Törnqvist, T. E. 2000. Fluvial response to climate and sea level change: a review and look forward. Sedimentology 47, 248.Google Scholar
Blum, M. D., IIIToomey, R. S., and Valastro, S. Jr. 1994. Fluvial response to Late Quaternary climatic and environmental change, Edwards Plateau, Texas. Palaeogeography, Palaeoclimatology, Palaeoecology 108, 121.Google Scholar
Boelter, D. H. 1969. Physical properties of peats as related to degree of decomposition. Soil Science Society of America Proceedings 33, 606609.Google Scholar
BoM. 2012. Record-breaking La Niña events. Bureau of Meteorology, Australian Government, August 2012. www.bom.gov.au/climate/enso/history/La-Nina-2010-12.pdf (accessed February 23, 2020).Google Scholar
BoM. 2016. What is La Niña and how does it impact Australia? Bureau of Meteorology, Australian Government, August 2016. www.bom.gov.au/climate/updates/articles/a020.shtml (accessed February 23, 2020).Google Scholar
Bonnema, M., Hossain, F., Nijssen, B., and Holtgrieve, G. 2020. Hydropower’s hidden transformation of rivers in the Mekong. Environmental Research Letters 15, 044017.Google Scholar
Boothroyd, R. J., Williams, R. D., Hoey, T. B., Barrett, B., and Prasojo, O. A. 2021. Applications of Google Earth Engine in fluvial geomorphology for detecting river channel change. WIREs Water 8, e21496. doi: 10.1002/wat2.1496.Google Scholar
Born, S. M., Genskow, K. D., Filbert, T. L., Hernandez-Mora, N., Keefer, M. L., and White, K. A. 1998. Socioeconomic and institutional dimensions of dam removals: the Wisconsin experience. Environmental Management 22, 359370.Google Scholar
Bouse, R. M., Fuller, C. C., Luoma, N., Hornberger, M., Jaffe, B. E., and Smith, R. 2010. Mercury-contaminated hydraulic mining debris in San Francisco Bay. San Francisco Estuary and Watershed Science 8. doi: 10.15447/sfews.2010v8iss1art3.Google Scholar
Bowman, M. 2002. Legal perspectives on dam removal. BioScience 52, 739747.Google Scholar
Brammer, A. J., Rodriguez del Rey, Zo., Spalding, E. A., and Poirrier, M. A. 2007. Effects of the 1997 Bonnet Carré Spillway opening on infaunal macroinvertebrates in Lake Pontchartrain, Louisiana. Journal of Coastal Research 23, 12921303.Google Scholar
Brandao, I., Mannaerts, C. M., and Saraiva, A. C. F. 2017. Seasonal variation of phytoplankton indicates small impacts of anthropic activities in a Brazilian Amazonian reserveroir. Ecohydrology and Ecobiology 17, 217226, doi: 10.1016/j.ecohyd.2017.04.001.Google Scholar
Brandt, S. A. 2000. Classification of geomorphological effects downstream of dams. Catena 40, 375401.Google Scholar
Brauneck, J., Pohl, R., and Jupner, R. 2016. Experiences of using UAVs for monitoring levee breaches. In IOP Conference Series: Earth and Environmental Science, vol. 46, 6th Digital Earth Summit, July 7–8, 2016, Beijing, China.Google Scholar
Bravard, J.-P. 2019. Sedimentary Crisis at the Global Scale 2: Deltas, a Major Environmental Crisis. John Wiley & Sons, Chichester.Google Scholar
Bravard, J.-P., and Gaydou, P. 2015. Historical development and integrated management of the Rhône River Floodplain, from the Alps to the Camargue Delta, France. In Hudson, P. F., and Middelkoop, H. (eds.), Geomorphic Approaches to Integrated Floodplain Management of Lowland Fluvial Systems in North America and Europe, Springer-Verlag New York, New York, pp. 289320.Google Scholar
Bravard, J.-P., Goichot, M., and Gaillot, S. 2013. Geography of sand and gravel mining in the lower Mekong River: first survey and impact assessment. EchoGéo 26. doi: 10.4000/echogeo.13659.Google Scholar
Breitburg, D., Levin, L. A., Oschlies, A., et al. 2018. Declining oxygen in the global ocean and coastal waters. Science 359, 7240. doi: 10.1126/science.aam7240.Google Scholar
Bridge, J. S. 2003. Rivers and Floodplains: Forms, Processes, and Sedimentary Record. John Wiley & Sons, Chichester.Google Scholar
Brierley, G. J., Ferguson, R. J., and Woolfe, K. J. 1997. What is a fluvial levee? Sedimentary Geology 114, 19.Google Scholar
Bristow, C. S., Skelly, R. L., and Ethridge, F. G. 1999. Crevasse splays from the rapidly aggrading sand-bed, braided Niobara River, Nebraska: effect of base level rise. Sedimentology 46, 10291047.Google Scholar
Brooks, G. R. 2000. Channel changes along the lower reaches of major Mackenzie River tributaries. In Dyke, L. D., and Brooks, G. R. (eds.), The Physical Environment of the Mackenzie Valley, Northwest Territories: A Base Line for the Assessment of Environmental Change. Geological Survey of Canada, Bulletin 547, pp. 159166.Google Scholar
Broothaerts, N., Verstraeten, G., Notebaert, B., et al. 2013. Human impact on floodplain geoecology. A Holocene perspective for the Dijle catchment, Central Belgium. In Abstract Volume 8th International Conference on Geomorphology (IAG), August 27–31, 2013, Paris.Google Scholar
Brouns, K., Verhouven, J. T. A., and Hefting, M. M. 2014. Short period of oxygenation releases latch on peat decomposition. Science of the Total Environment 481, 6168. doi: 10.1016/j.scitotenv.2014.02.030.Google Scholar
Brown, A. G., 2008. Geoarchaeology, the four dimensional (4D) fluvial matrix and climatic causality. Geomorphology 101, 278297.Google Scholar
Brown, A. G., Lespez, L., Sear, D. A., et al. 2018. Natural vs anthropogenic streams in Europe: history, ecology and implications for restoration, river-rewilding and riverine ecosystem services. Earth-Science Reviews 180, 185205.Google Scholar
Bucala, A. 2014. The impact of human activities on land use and land cover changes and environmental processes in the Gorce Mountains (Western Polish Carpathians) in the past 50 years. Journal of Environmental Management 138, 414.Google Scholar
Bucala-Hrabia, A. 2017. Long-term impact of socio-economic changes on agricultural land use in the Polish Carpathians. Land Use Policy 64, 391404.Google Scholar
Buglewicz, E. G., Mitchell, W. A., Scott, J. E., Smith, M., and King, W. L. 1988. A Physical description of main stem levee borrow pits along the lower Mississippi River. Lower Mississippi River Environmental Program, Report 2, US Army Corps of Engineers, Mississippi River Commission, Vicksburg, MS.Google Scholar
Buisje, T. A. D., Roozen, F. C. J. M., Grift, R. E., and van Geest, G. J. 2001. Stagnant water bodies. In Wolters, H. A., Platteeuw, M., and Schoor, M. M. (eds.), Guidelines for Rehabilitation and Management of Floodplains: Ecology and Safety Combined. Rijkswaterstaat, RIZA Report No. 2001.059, pp. 7188.Google Scholar
Buisje, T. A. D., Coops, H., Staras, M., et al. 2002. Restoration strategies for river floodplains along large lowland rivers in Europe. Freshwater Biology 47, 889907.Google Scholar
Bull, L. 1997. Magnitude and variation in the contribution of bank erosion to the suspended sediment load of the River Severn, UK. Earth Surface Processes and Landforms 22, 11091123.Google Scholar
Burnett, A. W. and Schumm, S. A. 1983. Alluvial-river response to neotectonic deformation in Louisiana and Mississippi. Science 222, 4950.Google Scholar
Burt, T. P., Bates, P. D., Steward, M. D., Claxton, A. J., Anderson, M. G., and Price, D. A. 2002. Water table fluctuations within the floodplain of the River Severn, England. Journal of Hydrology 262, 120.Google Scholar
Butzer, K. W. 1971. Recent history of an Ethiopian delta: the Omo River and the level of Lake Rudolph. Research Paper 136, Department of Geography, University of Chicago.Google Scholar
Butzer, K. W. 1976. Early Hydraulic Civilization in Egypt: A Study in Cultural Ecology. University of Chicago Press, Chicago and London, 134 pp.Google Scholar
Butzer, K. W. 2012. Collapse, environment, and society. Proceedings of the National Academy of Sciences 109, 36323639. doi: 10.1073/pnas.1114845109.Google Scholar
Butzer, K. W., and Helgren, D. M. 2005. Livestock, land cover, and environmental history: The Tablelands of New South Wales, Australia, 1820–1920. Annals of the Association of American Geographers 95, 80111.Google Scholar
Butzer, K. W., Abbott, J. T., Frederick, C., and Lehman, P. 2008. Soil-geomorphology and identification of “wet” cycles in the Holocene record of north-central Mexico. Geomorphology 101, 237277.Google Scholar
Caernarvon Delta and Diversion Study. 2014. Final report submitted to the Mississippi River Delta coalition. Science Contingency Funds, Louisiana, pp. 7–9, June 5, 2014.Google Scholar
California Department of Water Resources (California DWR). 2010. Fact sheet – Sacramento River flood control project weirs and flood relief structures. California Division of Flood Management, Sacramento, California. www.water.ca.gov/newsroom/docs/WeirsReliefStructures.pdf (accessed February 15, 2016)Google Scholar
California Statewide Flood Management Planning Program (CSFMP). 2013. California’s flood future: recommendations for managing the state’s flood risk. Final Report, November 2013. State of California, The Natural Resources Agency, Department of Water Resources.Google Scholar
Camillo, C. A. 2012. Divine Providence: The 2011 Flood in the Mississippi River and Tributaries Project. Mississippi River Commission, Vicksburg, MS.Google Scholar
Campanella, R. 2007. Geographies of New Orleans: Urban Fabrics before the Storm. Center for Louisiana Studies, University of Louisiana at Lafayette.Google Scholar
Campanella, R. 2008. Bienville’s Dilemma: A Historical Geography of New Orleans. Center for Louisiana Studies, University of Louisiana, Lafayette, Louisiana (U.S.).Google Scholar
Canadian Dam Association. 2019. Dams in Canada. Ottawa, Canada.Google Scholar
Cannatelli, K. M., and Crowe-Curran, J. 2012. Importance of hydrology on channel evolution following dam removal: case study and conceptual model. ASCE Journal of Hydraulic Engineering 138, 377390.Google Scholar
Carle, M. V., Sasser, C. E., and Roberts, H. H. 2015. Accretion and vegetation community change in the Wax Lake Delta following the historic 2011 Mississippi River flood. Journal of Coastal Research 31, 569587.Google Scholar
Carling, P., Jansen, J., and Meshkova, L. 2014. Multichannel rivers: their definition and classification. Earth Surface Processes and Landforms 39, 2637.Google Scholar
Carlson, S. D., and Guccione, M. J. 2010. Short-term uplift rates and surface deformation along the Reelfoot Fault, New Madrid Seismic Zone. Bulletin of the Seismological Society of America 100, 16591677.Google Scholar
Carminati, E., Doglioni, C., and Scrocca, D. 2005. Magnitude and causes of long-term subsidence of the Po Plain and Venetian Region. In Fletcher, C. A., Spencer, T., Da Mosto, J., and Campostrini, P. (eds.), Flooding and Environmental Challenges for Venice and Its Lagoon: State of Knowledge. Cambridge University Press, Cambridge, pp. 2128.Google Scholar
Caviedes, C. N. 2001. El Niño in History: Storming through the Ages. University Press of Florida, Gainesville.Google Scholar
Cazanacli, D., and Smith, N. D. 1997. A study of morphology and texture of natural levees, Cumberland Marshes, Saskatchewan, Canada. Geomorphology 25, 4355.Google Scholar
Cencetti, C., and Tacconi, P. 2005. The fluvial dynamics of the Arno River. Giornole di Geologia Applicata 1, 193202. doi: 10.1474/GGA.2005-01.0-19.0019.Google Scholar
Chamberlain, E. C., Törnqvist, T. E., Shen, Z., Mauz, B., and Wallinga, J. 2018. Anatomy of Mississippi Delta growth and its implications for coastal restoration. Science Advances 4. doi: 10.1126/sciadv.aar4740.Google Scholar
Chan, F., Mitchell, G., Adekola, G., and Mcdonald, A. T. 2013. Flood risk in Asia’s urban mega-deltas drivers, impacts and response. Environment and Urbanization ASIA 3, 4161.Google Scholar
Chang, H., Chiu, M., Chuang, Y., et al. 2017. Community responses to dam removal in a subtropical mountainous stream. Aquatic Sciences 79, 967983.Google Scholar
Changxing, S., Dian, Z., Lianyuan, Y., Bingyuan, L., Zulu, Z., and Ouyang, Z. 2007. Land subsidence as a result of sediment consolidation in the Yellow River Delta. Journal of Coastal Research 23, 173181.Google Scholar
Charlton, R. 2008. Fundamentals of Fluvial Geomorphology. Routledge, London, 234pp.Google Scholar
Chatanantavet, P., Lamb, M. P., and Nittouer, J. A. 2012. Backwater controls of avulsion location on deltas. Geophysical Research Letters 39, L01402. doi: 10.1029/2011GL050197.Google Scholar
Chen, C., Lin, H., Zhang, Y., and Lu, Z. 2012. Ground subsidence geo-hazards induced by rapid urbanization: implications from InSAR observation and geological analysis. Natural Hazards and Earth Systems Science 12, 935942. doi: 10.5194/nhess-12-935-2012.Google Scholar
Chen, L., Wei, W., Fu, B.-J., and Lu, Y. 2007. Soil and water conservation on the Loess Plateau in China: review and perspective. Progress in Physical Geography 31, 389403. doi: 10.1177/0309133307081290.Google Scholar
Chen, Y., Syvitski, J. P. M., Gao, S., Overeem, I., and Kettner, A. J. 2012. Socio-economic impacts on flooding: a 4000-year history of the Yellow River, China. AMBIO 41, 682698. doi: 10.1007/s13280–012-0290-5.Google Scholar
Cheng, L., Opperman, J. J., Tickner, D., Speed, R., Guo, Q., and Chen, D. 2018. Managing the three Gorges Dam to implement environmental flows in the Yangtze River. Frontiers in Environmental Science 6, 64. doi: 10.3389/fenvs.2018.00064.Google Scholar
Chiew, F. H. S., and McMahon, T. 2002. Global ENSO-streamflow teleconnection, streamflow forecasting and interannual variability. Hydrological Sciences Journal/Journal des Sciences Hydrologiques 47, 505522.Google Scholar
Chin, A., Harris, D. L., Trice, T. H., and Given, J. L. 2002. Adjustment of stream channel capacity following dam closure, Yegua Creek, Texas. Journal of the American Water Resources Association 38, 15211531. doi: 10.1111/j.1752-1688.2002.tb04362.x.Google Scholar
Church, J. A., Clark, P. U., Cazenave, A., et al. 2013. Sea level change. In Stocker, T. F., Qin, D., Plattner, G.-K., et al. (eds.), Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, 11371216.Google Scholar
Church, M. 1983. Pattern of instability in a wandering gravel bed channel. In Collinson, J. D., and Lewin, J. (eds.), Modern and Ancient Fluvial Systems. International Association of Sedimentologists, Special Publication, 6, Oxford, pp. 169180.Google Scholar
Church, M. 2002. Geomorphic thresholds in riverine landscapes. Freshwater Biology 47, 541557.Google Scholar
Church, M. 2015. Channel stability: morphodynamics and morphology of rivers. In Rowinski, P., and Radecki-Pawlik, A. (eds.), Rivers – Physical, Fluvial and Environmental Processes. Series: GeoPlanet: Earth and Planetary Sciences. Springer, Switzerland, pp. 282321.Google Scholar
CIRIA. 2013. The International Levee Handbook (C731). CIRIA, London.Google Scholar
Citterio, A., and Piégay, H. 2009. Overbank sedimentation rates in former channel lakes: characterization and control factors. Sedimentology 56, 461482.Google Scholar
City of Rotterdam. 2015. Rotterdam climate initiative. www.010duurzamestad.nl/ (accessed November 20, 2018).Google Scholar
Clifford, N. J. 1993. Formation of riffle – pool sequences: field evidence for an autogenetic process. Sedimentary Geology 85, 3951.Google Scholar
Coastal Protection and Restoration Authority of Louisiana (CPRA Louisiana). 2017. Louisiana’s Comprehensive Master Plan for a Sustainable Coast. Coastal Protection and Restoration Authority of Louisiana, Baton Rouge, LA.Google Scholar
Coe, M., Latrubesse, E., Ferreira, M., and Amsler, M. 2011. The effect of deforestation and climate variability on the stream flow of the Araguaia River. Biogeochemistry 105, 119131.Google Scholar
Cohen, E. 2012. Flooded: an auto-ethnography of the 2011 Bangkok flood. ASEAS – Austrian Journal of South-East Asian Studies 5, 316334.Google Scholar
Cohen, K. M. 2003. Differential Subsidence within a Coastal Prism: Late-Glacial Holocene Tectonics in the Rhine-Meuse Delta, The Netherlands. Netherlands Geographical Survey 316, Utrecht.Google Scholar
Cohen, K. M., Stouthamer, E., and Berendsen, H. J. A. 2002. Fluvial deposits as a record for Late Quaternary neotectonic activity in the Rhine-Meuse delta, the Netherlands. Netherlands Journal of Geosciences / Geologie en Mijnbouw 81, 389405.Google Scholar
Cohen, K. M., Stouthamer, E., and Berendsen, H. J. A. 2009. Riviersystemen in de polders van het Rivierengebied Zand. In Cohen, K. M., Stouthamer, E., Hoek, W. Z., Berendsen, H. J. A., and Kempen, H. F. J. (eds.), Banen – Zanddieptekaarten van het Rivierengebied en het IJsseldal in de provincies Gelderland en Overijssel. Arnhem, Provincie Gelderland, pp. 5568.Google Scholar
Coker, H. E, Hotchkiss, R. H., and Johnson, D. A. 2009. Conversion of a Missouri River dam and reservoir to a sustainable system: sediment management. Journal of the American Water Resources Association 45, 815827. doi: 10.1111/j.1752-1688.2009.00324.Google Scholar
Coleman, J. M. 1969. Brahmaputra River: channel processes and sedimentation. Sedimentary Geology 3, 129239.Google Scholar
Coleman, J. M. 1988. Dynamic changes and processes in the Mississippi River delta. Geological Society of America Bulletin 100, 9991015.Google Scholar
Coleman, J. M., and Gagliano, S.M. 1964. Cyclic sedimentation in the Mississippi river deltaic plain. Gulf Coast Association of Geological Societies, Transactions 14, 6780.Google Scholar
Coleman, J. M., and Roberts, H. H. 1989. Deltaic coastal wetlands. Geologie en Mijnbouw (Netherlands Journal of Geosciences) 68, 124.Google Scholar
Coleman, J. M., and Wright, L. D. 1971. Analysis of major river systems and their deltas, procedures and rationale, with two examples. Louisiana State University, Coastal Studies Institute, Technical Report, 95.Google Scholar
Coleman, J. M., and Wright, L. D. 1975. Modern river deltas, variability of processes and sand bodies. In Broussard, M. L. (ed.), Deltas, Models for Exploration. Houston Geological Society, Houston, TX, pp. 99149.Google Scholar
Coleman, J. M., Roberts, H. H., and Huh, O. K. 1986. Deltaic landforms. In Short, N. M., and Blair, R. W. Jr. (eds.), Geomorphology from Space: A Global Overview of Regional Landforms. National Aeronatuics and Space Administration, Scientific and Technical Branch, Washington, DC, pp. 317352.Google Scholar
Coleman, J. M., Roberts, H. H., and Stone, G. W. 1998. Mississippi River delta: an Overview. Journal of Coastal Research 14, 698716.Google Scholar
Coleman, J. S. M., and Budikova, D. 2013. Eastern U.S. summer streamflow during extreme phases of the North Atlantic oscillation. Journal of Geophysical Research: Atmospheres 118, 41814193.Google Scholar
Collier, M., Webb, R. H., and Schmidt, J. C. 1996. A primer on the downstream effects of dams. U.S. Geological Survey, Circular 1126.Google Scholar
Colloff, M. J. 2014. Flooded Forest and Desert Creek: Ecology and History of the River Red Gum. CSIRO Publishing, Collingwood, Australia, 312 p.Google Scholar
Comeaux, M. 1970. The Atchafalaya River raft. Louisiana Studies 9, 217227.Google Scholar
Constantine, J. A., Dunne, T., Ahmed, J., Legleiter, C., and Lazarus, E. D. 2014. Sediment supply as a driver of river meandering and floodplain evolution. Nature Geoscience 7, 899903. doi: 10.1038/NGEO2282.Google Scholar
Constantinescu, Ş., Achim, D., Rus, I., and Giosan, L. 2015. Embanking the lower Danube: from natural to engineered floodplains and back. In Hudson, P. F., and Middelkoop, H. (eds.), Geomorphic Approaches to Integrated Floodplain Management of Lowland Fluvial Systems in North America and Europe. Springer, New York, pp. 265288.Google Scholar
Coomes, O. T., Abizaid, C., and Lapointe, M. 2009. Human modification of a large meandering Amazonian river: genesis, ecological and economic consequences of The Masisea cutoff on the Central Ucayali, Peru. AMBIO: A Journal of the Human Environment 38, 130134.Google Scholar
Couvillion, B. R., Beck, H., Schoolmaster, D., and Fischer, M. 2017. Land area change in coastal Louisiana 1932 to 2016. U.S. Geological Survey Scientific Investigations Map 3381, 16 p. pamphlet.Google Scholar
Crist, E., Mora, C., and Engelman, R. 2017. The interaction of human population, food production, and biodiversity protection. Science 356, 260264.Google Scholar
Cruz, R. V., Harasawa, H., Lal, M., et al. 2007. Asia. Climate Change 2007: impacts, adaptation and vulnerability. In Parry, M. L., Canziani, O. F., Palutikof, J. P., van der Linden, P. J., and Hanson, C. E. (eds.), Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, pp. 469506.Google Scholar
Cuenca, M. C., Hannsen, R., and van Leijn, F. 2007. Subsidence due to Peat Decomposition in the Netherlands – Kinematic Observations from Radar Interferometry, TU Delft Research Project, presented in Frascati, Italy.Google Scholar
Dai, P., and Tan, B. 2017. The Nature of the Arctic Oscillation and diversity of the extreme surface weather anomalies it generates. Journal of Climate 30, 55635584. doi: 10.1175/JCLI-D-16-0467.1.Google Scholar
Dahl, T. E., Swords, J., and Bergeson, M. T. 2009. Wetland inventory of the Yazoo Backwater Area, Mississippi – Wetland status and potential changes based on an updated inventory using remotely sensed imagery. U.S. Fish and Wildlife Service, Division of Habitat and Resource Conservation, Washington, DC.Google Scholar
Dam Removal Europe. 2018. https://damremoval.eu/ (accessed January 9, 2019).Google Scholar
Daniels, J. M. 2002. Drainage network adjustment following channelization, Homochitto River Basin, Mississippi. In Steinberg, M. K., and Hudson, P. F. (eds.), Cultural and Physical Expositions: Geographic Studies in the Southern United States and Latin America. Geoscience and Man, 36. Louisiana State University, pp. 291308.Google Scholar
Daniels, J. M., Leigh, D. S., and Carson, E. C. 2019. Holocene paleohydrology and paleofloods in the Driftless Area. In Carson, E. C., Rawling, J. E., Daniels, J. M., and Attig, J. W. (eds.), The Physical Geography and Geology of the Driftless Area: The Career and Contributions of James C. Knox. Geological Society of America Special Paper 543, pp. 7592.Google Scholar
Darby, S. E., Leyland, J., Kummu, M., Reaseanen, T. A., and Lauri, H. 2013. Decoding the drivers of bank erosion on the Mekong River: the roles of the Asian monsoon, tropical storms, and snowmelt. Water Resources Research 49, 21462163. doi: 10.1002/wrcr.20205.Google Scholar
Darby, S. E., Hackney, C., Leyland, J. et al. 2016. Fluvial sediment supply to a mega-delta reduced by shifting tropical-cyclone activity. Nature 539, 276279. doi: 10.1038/nature19809.Google Scholar
Davivongs, V., Yokohari, M., and Hara, Y. 2012. Neglected canals: deterioration of indigenous irrigation system by urbanization in the West Peri-Urban area of Bangkok Metropolitan Region. Water 4, 1227. doi: 10.3390/w4010012.Google Scholar
Day, G., Dietrich, W. E., Rowland, J., and Marshall, A. 2009. The depositional web on the floodplain of the Fly River, Papua New Guinea. Journal of Geophysical Research 113, F01S02. doi: 10.1029/2006JF000622.Google Scholar
Day, J. H., et al. 2009. The impacts of pulsed reintroduction of river water on a Mississippi delta coastal basin. Journal of Coastal Research 54, 225243.Google Scholar
Day, J. H., Cable, J. E., Lane, R. R., and Kemp, G. P. 2016. Sediment deposition at the Caernarvon crevasse during the Great Mississippi Flood of 1927: implications for coastal restoration. Water 8. doi: 10.3390/w802003.Google Scholar
Day, J. W., and Giosan, L. 2008. Geomorphology: survive or subside? Nature Geoscience 1, 156157. doi: 10.1038/ngeo137.Google Scholar
Day, J. W., Hunter, R. F., Keim, R., et al. 2012. Ecological response of forested wetlands with and without Large-Scale Mississippi River input: implications for management. Ecological Engineering 46, 5767.Google Scholar
de Almeida, G. A. M., and Rodríguez, J. F. 2011. Understanding pool‐riffle dynamics through continuous morphological simulations. Water Resources Research 47, W01502. doi: 10.1029/2010WR009170.Google Scholar
de Bruin, D. 2006. Similarities and differences in the historical development of flood management in the alluvial stretches of the lower Mississippi Basin and the Rhine Basin. Irrigation and Drainage 55, 2354.Google Scholar
D’Haen, K., Dusar, B., Verstraeten, G., Degryse, P., and de Brue, H. 2013. A sediment finger-printing approach to understand the geomorphic coupling in an eastern Mediterranean mountainous river catchment. Geomorphology 197, 6475.Google Scholar
DeHaan, H., Stamper, J., and Walters, B. 2012. Mississippi River and tributaries system 2011 post-flood report. U.S. Army Corps of Engineers, Mississippi Valley Division, Vicksburg, MS, 4 plates.Google Scholar
DeLaune, R. D., Jugsujinda, A., Peterson, G. W., and Patrick, W. H. 2003. Impact of Mississippi River freshwater reintroduction on enhancing marsh accretionary processes in a Louisiana estuary. Estuarine Coastal and Shelf Science 58, 653662.Google Scholar
DEFRA. 2005. Making Space for Water: Taking Forward a New Government Strategy for Flood and Coastal Erosion Risk Management Forward in England. DEFRA, London.Google Scholar
DELP Victoria, AU. 2016. Decommissioning Dams: A Guide for Owners. Department of Environment, Land, Water and Planning, State of Victoria, Australia.Google Scholar
Denevan, W. M. 2000. Cultivated Landscapes of Native Amazonia and the Andes. Oxford University Press.Google Scholar
Dennis, J. V. Jr. 1990. Kampuchea’s ecology and resource base: natural limitations on food production. In Ablin, D. A., and Hood, M. (eds.), Revival: The Cambodian Agony. Routledge, New York, pp. 208238.Google Scholar
Department of Environment, Land, Water and Planning (DELWP). 2015. Levee Management Guidelines. The State of Victoria, Melbourne.Google Scholar
Dépret, T., Riquier, J., and Piégay, H. 2017. Evolution of abandoned channels: insights on controlling factors in a multi-pressure river system. Geomorphology 294, 99118.Google Scholar
Desloges, J. R., and Church, M. 1989. Wandering gravel bed rivers. Canadian Geographer 33, 360364.Google Scholar
Dettinger, M. 2011. Climate change, atmospheric rivers, and floods in California – a multimodel analysis of storm frequency and magnitude changes. Journal of the American Water Resources Association 47, 514523. doi: 10.1111/j.1752-1688.2011.00546.x.Google Scholar
Diamond, J. 2006. Collapse: How Societies Choose to Fail or Succeed. Penguin, New York.Google Scholar
Dieppois, B., Pohl, B., Rouault, M. New, M., Lawler, D., and Keenlyside, N. 2016. Interannual to interdecadal variability of winter and summer southern African rainfall, and their teleconnections, Journal of Geophysical Research: Atmospheres 121, 62156239. doi: 10.1002/2015JD024576.Google Scholar
Dieras, P. L., Constantine, J. A., Hales, T. C., Piégay, H., and Riquier, J. 2013. The role of oxbow lakes in the off-channel storage of bed material along the Ain River, France. Geomorphology. doi: 10.1016/j.geomorph.2012.12.024.Google Scholar
Dierauer, J., Pinter, N., and Remo, J. W. 2012. Evaluation of levee setbacks for flood-loss reduction, Middle Mississippi River, USA. Journal of Hydrology 450–451, 18. doi: 10.1016/j.jhydrol.2012.05.044.Google Scholar
Ding, L., Chen, L., Ding, C., and Tao, J. 2018. Global trends in dam removal and related research: a systematic review based on associated datasets and bibliometric analysis. Chinese Geographical Science 29, 112.Google Scholar
Dirmeyer, P. A., and IIIKinter, J. L.. 2009. The “Maya Express”: floods in the U.S. Midwest. Eos, Transactions, American Geophysical Union 90, 101102. doi: 10.1029/2009EO120001.Google Scholar
Dixon, T. H., Amelung, F., Ferretti, A., et al. 2006. Subsidence and flooding in New Orleans. Nature 441, 587588. doi: 10.1038/441587a.Google Scholar
Dokka, R. K. 2006. Modern-day tectonic subsidence in coastal Louisiana. Geology 34, 281284.Google Scholar
Domagalski, J. L., Slotton, D. G., Alpers, C. N., et al. 2004. Summary and synthesis of mercury studies in the Cache Creek Watershed, California, 2000–01. U.S. Geological Survey Water-Resources Investigations Report, 03-4335.Google Scholar
Dominquez, J. M. L. 1996. The São Francisco strand plain: a paradigm for wave-dominated deltas? Geological Society London Special Publications 117, 217231.Google Scholar
Doolittle, W. D. 2000. Cultivated Landscapes of Native North America. Oxford University Press.Google Scholar
Dorozynski, A. 1975. After the dam the depression?. Nature 255, 570. doi: 10.1038/255570a0.Google Scholar
Dos Santos, V., Stevaux, J., and Assine, M. 2017. Fluvial processes in attachment bars in the upper Paraná River, Brazil. Revista Brasileira de Geomorfologia 18, 484499. doi: 10.20502/rbg.v18i3.1135.Google Scholar
Dotterweich. 2008. The history of soil erosion and fluvial deposits in small catchments of central Europe: deciphering the long-term interaction between humans and the environment-A review. Geomorphology 101, 192208.Google Scholar
Downs, P. W., and Gregory, K. J. 2004. River Channel Management: Towards Sustainable Catchment Hydrosystems. Arnold, London.Google Scholar
Downs, P. W., Gregory, K. J., and Brookes, A. 1991. How integrated is river basin management? Environmental Management 15, 299309.Google Scholar
Doyle, M. W., Harbor, J., and Stanley, E. H. 2003a. Toward policies and decision-making for dam removal. Environmental Management 31, 453465.Google Scholar
Doyle, M. W., Stanley, E., and Harbor, J. 2003b. Channel adjustments following two dam removals in Wisconsin. Water Resources Research 39,Google Scholar
Doyle, M. W., Stanley, E., Havlick, D., et al. 2008. Aging infrastructure and ecosystem restoration. Science 319(5861), 286287.Google Scholar
Dreibrodt, S., Lubos, C., Terhorst, B., Dammc, B., and Bork, H.-R. 2009. Historical soil erosion by water in Germany: Scales and archives chronology, research perspectives. Quaternary International 222, 8095.Google Scholar
Duda, J. J., Johnson, R. C., Wieferich, D. J., Wagner, W. J., and Bellmore, J. R. 2020. USGS dam removal science database v3.0 (ver. 3.0, January 2020). U.S. Geological Survey data release.Google Scholar
Dudley, W. T. 2004. Hydraulic geometry relations for rivers in coastal and central Maine. U.S. Geological Survey, Scientific Investigation Report 2004-5042, Augusta, Maine.Google Scholar
Dugan, P. J., Barlow, C., Agostinho, A. A., et al. 2010. Fish migration, dams, and loss of ecosystem services in the Mekong basin. AMBIO 39, 344348. doi: 10.1007/s13280-010-0036-1.Google Scholar
Dunbar, J. B., IIITorrey, V. H., and Wakeley, L. D. 1999. A case history of embankment failure: Geological and geotechnical aspects of the Celotex levee failure, New Orleans, Louisiana. Final Report. US Army Engineer Research and Development Center, Technical Report GL-99-11, Vicksburg, MS.Google Scholar
Duncan, J. M., Brandon, T. L., Wright, S. G., and Vroman, N. 2008. Stability of I-walls in New Orleans during Hurricane Katrina. Journal of Geotechnical and Geoenvironmental Engineering 134, 681692.Google Scholar
Dunne, T., and Aalto, R. E. 2013. Large river floodplains. In Shroder, J. F. (ed.), Treatise on Geomorphology, vol. 9. Academic Press, San Diego, CA, pp. 645678.Google Scholar
Dunne, T., Mertes, L. A. K., Meade, R. H., Richey, J., and Forsberg, B. R. 1998. Exchanges of sediment between the floodplain and channel of the Amazon River in Brazil. Geological Society of America Bulletin 110, 450467.Google Scholar
Dusar, B., Verstraeten, G., D’Haen, K., Bakker, J., Kaptijn, E., and Waelkens, M. 2012. Sensitivity of the Eastern Mediterranean geomorphic system towards environmental change during the Late Holocene: a chronological perspective. Journal of Quaternary Science 27, 371382.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, 33383369. doi: 10.1029/2018JF004703.Google Scholar
Echevarria-Doyle, W., Biedenharn, D. S., and Little, C. D. Jr. 2020. Lake Providence to Old River geomorphology assessment. Final report. U.S. Army Corps of Engineers, Mississippi Valley Division Mississippi River Geomorphology and Potamology Program Report No. 32, Vicksburg, MS.Google Scholar
Egberts, H. 1950. De Bodemgesteldheid van de Betuwe. Stichting voor Bodemkartering Wageningen. Versl. Landbouwk Onderzoek No. 56.1 9, ’ ’S Gravenhage, NL.Google Scholar
El Bakry, M. 1994. Net radiation over the Aswan high dam lake. Theoretical and Applied Climatology 49, 129133.Google Scholar
Elliott, C. M., DeLonay, A. J., Chojnacki, K. A., Jacobson, R. B., 2020. Characterization of Pallid Sturgeon (Scaphirhynchus albus) spawning habitat in the lower Missouri River. Journal of Applied Ichthyology 36, 25–38.Google Scholar
Elliott, D. O. 1932. The improvement of the lower Mississippi River for Flood Control and Navigation, vol. 1–3. US Army Corps of Engineers Waterways Experiment Station, Vicksburg, MS.Google Scholar
Elliott, T. 1986. Deltas. In Reading, H. G. (ed.), Sedimentary Environments and Facies. Blackwell Scientific Publications, Oxford, pp. 113154.Google Scholar
El-Mahdy, M. E., Abbas, M. S., and Sobhy, H. M. 2019. Development of mass-transfer evaporation model for Lake Nasser, Egypt. Journal of Water and Climate Change 12, 223237. doi: 10.2166/wcc.2019.116.Google Scholar
El-Nashar, W. Y., and Elyamany, A. H. 2018. Managing risks of the Grand Ethiopian Renaissance Dam on Egypt. Ain Shams Engineering Journal 9, 23832388.Google Scholar
Eltahir, E. A. B. 1996. El Niño and the natural variability in the flow of the Nile River. Water Resources Research 32, 131137.Google Scholar
Eltahir, E. A. B., and Wang, G. 1999. Nilometers, E1 Niño, and climate variability. Geophysical Research Letters 26, 489492. doi: 10.1029/1999GL900013.Google Scholar
Endris, H. S., Lennard, C., Hewitson, B., Dosio, A., Nikulin, G., and Artan, G. A. 2019. Future changes in rainfall associated with ENSO, IOD and changes in the mean state over Eastern Africa. Climate Dynamics 52, 20292053. doi: 10.1007/s00382–018-4239-7.Google Scholar
Enfield, D. B., Mestas-Nunez, A. M., and Trimble, P. J. 2001. The Atlantic multidecadal oscillation and its relation to rainfall and river flows in the continental U.S. Geophysical Research Letters 28, 20772080.Google Scholar
Erban, L. E., Gorelick, S. M., and Zebker, H. A. 2014. Groundwater extraction, land subsidence, and sea-level rise in the Mekong Delta, Vietnam. Environmental Research Letters 9(8). doi: 10.1088/1748-9326/9/8/084010.Google Scholar
Erickson, C. L., and Walker, J. H. 2009. Pre-Columbian causeways and canals as landesque capital. In Snead, J., Erickson, C., and Darling, A. (eds.), Landscapes of Movement: Trails, Paths, and Roads in Anthropological Perspective. University of Pennsylvania Press, Philadelphia, pp. 232252.Google Scholar
Erkens, G. Bucx, T., Dam, R., de Lange, G., and Lambert, J. 2015. Sinking cities: an integrated approach to solutions. In Disaster Risk, edited by Global Facility for Disaster Reduction and Recovery, The Making of a Riskier Future: How Our Decisions Are Shaping Future. Case Study C 90–99. World Bank, Washington, DC.Google Scholar
Erskine, W. D. 1992. Channel response to large-scale river training works: Hunter River, Australia. River Research and Applications 7, 261278.Google Scholar
Erskine, W. D., and Webb, A. A. 2003. Desnagging to resnagging: new directions in river rehabilitation in southeastern Australia. River Research and Applications 19, 233249.Google Scholar
Escarameia, M. 1998. River and Channel Revetments: A Design Manual. Thomas Telford, London.Google Scholar
Espa, P., Batalla, R. J., Brignoli, M. L., Crosa, G., Gentili, G., and Quadroni, S. 2019. Tackling reservoir siltation by controlled sediment flushing: impact on downstream fauna and related management issues. PLoS ONE 14, e0218822. doi: 10.1371/journal.pone.0218822.Google Scholar
Esposito, C. R., Georgiou, I. Y., and Kolker, A. S. 2013. Hydrodynamic and geomorphic controls on mouth bar evolution. Geophysical Research Letters 40, 15401545. doi: 10.1002/grl.50333.Google Scholar
European Centre for River Restoration (ECRR). 2018. www.ecrr.org/RiverRestoration/Whatisriverrestoration/tabid/2614/Default.aspx (accessed October 5, 2018).Google Scholar
European Commission (EC). 2019. Fitness check of the Water Framework Directive, Groundwater Directive, Environmental Quality Standards Directive and Floods Directive. December 10, 2019.Google Scholar
European Council (EC). 2000. Directive 2000/60/EC of the European Parliament and of the Council of 23 October 2000 establishing a framework for Community action in the field of water policy.Google Scholar
European Council (EC). 2007. Directive 2007/60/EC of the European Parliament and of the Council of 23 October 2007 on the assessment and management of flood risks.Google Scholar
European Environment Agency (EEA). 2018. European Waters: Assessment of Status and Pressures 2018. European Environmental Agency. EEA Report No 7/2018. Publications Office of the European Union, Luxembourg.Google Scholar
Evans, J. E. 2015. Contaminated sediment and dam removals: problem or opportunity? Eos 96. doi: 10.1029/2015EO036385.Google Scholar
Everitt, B. 1993. Channel responses to declining flow on the Rio Grande between Ft. Quitman and Presidio, Texas. Geomorphology 6, 225242.Google Scholar
Executive Order 13690. 2015. Establishing a Federal Flood Risk Management Standard and a process for further soliciting and considering stakeholder input, January 30, 2015.Google Scholar
Fabris, M., Achilli, V., and Menin, A. 2014. Estimation of subsidence in Po Delta area (Northern Italy) by integration of GPS data, high-precision leveling and archival orthometric elevations. International Journal of Geosciences 5, 571585. doi: 10.4236/ijg.2014.56052.Google Scholar
Fan, H., Huang, H., and Zeng, T. 2006. Impacts of anthropogenic activity on the recent evolution of the Huanghe (Yellow) River Delta. Journal of Coastal Research 22, 919929.Google Scholar
Fan, H., Hea, D., and Wang, H. 2015. Environmental consequences of damming the mainstream Lancang-Mekong River: a review. Earth Science Reviews 146, 7791.Google Scholar
FAO. 2011a. AQUASTAT Country Profile – Cambodia.. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy.Google Scholar
FAO. 2011b. AQUASTAT Transboundary River Basins – Indus River Basin. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy.Google Scholar
FAO. 2011c. Irrigation in Southern and Eastern Asia in figures – AQUASTAT Survey. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy.Google Scholar
FAO. 2016a. AQUASTAT Country Profile – Egypt. Food and Agriculture Organization of the United Nations (FAO), Rome, Italy.Google Scholar
FAO. 2017. The Future of Food and Agriculture – Trends and Challenges. U.N. Food and Agricultural Organization, Rome, Italy.Google Scholar
Farrell, K. M. 1987. Sedimentology and facies architecture of overbank deposits of the Mississippi River, False River Region, Louisiana. In Etheridge, F. G., Flores, R. M., and Harvey, M. D. (eds), Recent Developments in Fluvial Sedimentology. Society of Economic Paleontologists and Mineralogists Special Publication 39, pp. 111120.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, 666685.Google Scholar
Ferguson, H. B. 1940. History of the Improvement of the Lower Mississippi River for Flood Control and Navigation, 1932–1939. Mississippi River Commission, Vicksburg, MS.Google Scholar
Ferguson, R. I. 1986. Hydraulics and hydraulic geometry. Progress in Physical Geography 10, 131.Google Scholar
Ferguson, R. I. 1987. Hydraulic and sedimentary controls on channel pattern. In Richards, K. (ed.), River Channels: Environment and Process. Basil Blackwell Ltd, Oxford, pp. 129158.Google Scholar
Ferguson, R. J., and Brierley, G. J. 1999. Levee morphology and sedimentology along the lower Tuross River, south‐eastern Australia. Sedimentology 46, 627648.Google Scholar
Fernández Garrido, P. 2018. Connecting fish, rivers, and people: Swimways around the world from local to global. Regional Conference on river habitat restoration for inland fisheries in the Danube River Basin and adjacent Black Sea áreas. Bucharest, November 13–15, 2018.Google Scholar
Ficchì, A., and Stephens, L. 2019. Climate variability alters flood timing across Africa. Geophysical Research Letters 46, 88098819. doi: 10.1029/2019GL081988.Google Scholar
Filizola, N., Latrubesse, E. M., Fraizy, P., Souza, R., Guimarães, V., and Guyot, J.-L. 2014. Was the 2009 flood the most hazardous or the largest ever recorded in the Amazon? Geomorphology 215, 99105.Google Scholar
Fischenich, J. C. 2003. Effects of riprap on riverine and riparian ecosystems. ERDC/EL TR-03-4, U.S. Army Corps of Engineer Research and Development Center, Vicksburg, MS.Google Scholar
Fisher, W. L., Brown, L. F. Jr., Scott, A. J., and McGowen, J. H. 1969. Delta Systems in the Exploration of Oil and Gas. Bureau of Economic Geology, University of Texas at Austin.Google Scholar
Fisk, H. N. 1944. Geological Investigation of the Alluvial Valley of the lower Mississippi River. Mississippi River Commission, Vicksburg, MS.Google Scholar
Fisk, H. N. 1947. Fine-Grained Alluvial Deposits and Their Effects on Mississippi River Activity. Mississippi River Commission, Vicksburg, MS.Google Scholar
Fisk, H. N. 1952. Geological Investigation of the Atchafalaya Basin and the Problem of the Mississippi River Diversion. U.S. Army Waterways Experiment Station, Mississippi River Commission, Vicksburg, MS.Google Scholar
Florsheim, J. L., and Dettinger, M. D. 2015. Promoting atmospheric-river and snowmelt-fueled biogeomorphic processes by restoring river-floodplain connectivity in California’s Central Valley. In Hudson, P. F., and Middelkoop, H. (eds.), Geomorphic Approaches to Integrated Floodplain Management of Lowland Fluvial Systems in North America and Europe. Springer-Verlag New York, New York, pp. 119141.Google Scholar
Florsheim, J. L., and Mount, J. M. 2003. Changes in lowland floodplain sedimentation processes: pre-disturbance to post-rehabilitation, Cosumnes River, CA. Geomorphology 56, 305323.Google Scholar
Florsheim, J. L., Mount, J. F., and Chin, A. 2008. Bank erosion as a desirable attribute of rivers. BioScience 58, 519529.Google Scholar
Foley, J. A., Botta, A.A., Coe, M.T., and Costa, M.H. 2002. El Niño–Southern oscillation and the climate, ecosystems and rivers of Amazonia, Global Biogeochemical Cycles 16, 1132. doi: 10.1029/2002GB001872.Google Scholar
Foley, J. A., DeFries, N., Asner, R., et al. 2005. Global consequences of land use. Science 309, 570574.Google Scholar
Foley, M. M., Warrick, J.A., Ritchie, A., et al. 2017a. Coastal habitat and biological community response to dam removal on the Elwha River. Ecological Monographs 87, 552577. doi: 10.1002/ecm.1268.Google Scholar
Foley, M. M., et al. 2017b. Dam removal: listening in. Water Resources Research 53, 52295246. doi: 10.1002/2017WR020457.Google Scholar
Folk, R. L., and Ward, W. C. 1957. Brazos River bar: a study in the significance of grain size parameters. Journal of Sedimentary Petrology 27, 326.Google Scholar
Food and Agricultural Organization of the United Nations (FAO). 1988. Nature and Management of Tropical Peat Soils. Food and Agricultural Organization of the United Nations, FAO Soils Bulletin, 59. Rome.Google Scholar
Fovet, O., Ndom, N., Crave, A., and Pannard, A. 2020. Influence of dams on river water-quality signatures at event and seasonal scales: The Sélune River (France) case study. River Research and Applications. doi: 10.1002/rra.3618.Google Scholar
Fox, C. A., Magilligan, F. J., and Snedden, 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
Franklin, P., Dunbar, M., and Whitehead, P. 2008. Flow control on lowland river macrophytes: a review. Science of the Total Environment 400, 369–78.Google Scholar
Frazier, D. E. 1967. Recent deltaic deposits of the Mississippi River: their development and chronology. Transactions Gulf Coast Association of Geological Society of America 17, 287311.Google Scholar
Friedkin, J. F. 1945. A Laboratory Study of the Meandering of Alluvial Rivers. U.S. Army Engineer Waterways Experiment Station, Vicksburg, MS.Google Scholar
Frings, R. M., Gehres, N., Promny, M., Middelkoop, H., Schuttrumpf, H., and Vollmer, S. 2014. Today’s sediment budget of the Rhine River channel, focusing on the upper Rhine Graben and Rhenish Massif. Geomorphology 204, 573587.Google Scholar
Frings, R. M, Hillebrand, G., Gehres, N., Banhold, K., Schriever, S., and Hoffmann, T. 2019. From source to mouth: basin-scale morphodynamics of the Rhine River. Earth-Science Reviews 196, 102830.Google Scholar
Fu, B., Wang, S., Liu, Y., Liu, J., Liang, W., and Miao, C. 2017. Hydrogeomorphic ecosystem responses to natural and anthropogenic changes in the Loess Plateau of China. Annual Review of Earth and Planetary Sciences 45, 223224.Google Scholar
Fu, K. D., He, D. M., and Lu, X. X. 2008. Sedimentation in the Manwan reservoir in the Upper Mekong and its downstream impacts. Quaternary International 186, 9199.Google Scholar
Fu, R., and Li, W. 2004. The influence of the land surface on the transition from dry to wet season in Amazonia. Theoretical and Applied Climatology 78, 98110.Google Scholar
Fugate, J. M. 2014. Measurements of land subsidence rates on the north-western portion of the Nile Delta using radar interferometry techniques. MSc thesis (supervisors R. H. Becker and M. Sultan), Department of Geology, University of Toledo.Google Scholar
Gagliano, S. M., and Howard, P. C. 1984. The neck cutoff oxbow lake cycle along the lower Mississippi River. In Elliot, C. M. (ed.), River Meandering: Proceedings of the Conference Rivers ’83. American Society of Civil Engineers, New Orleans, pp. 147158.Google Scholar
Galat, D. L., Fredrickson, L. H., Humburg, D. D., et al. 1998. Flooding to restore connectivity of regulated, large-river wetlands. BioScience 48, 721733.Google Scholar
Gale, E. L., and Saunders, M. A. 2013. The 2011 Thailand flood: climate causes and return periods. Weather 68, 233237. doi: 10.1002/wea.2133.Google Scholar
Galloway, D. C., Jones, D. R., and Ingebritsen, S. E. 1999. Land subsidence in the United States. U.S. Geological Survey, Circular, 1182.Google Scholar
Galloway, W. D. 1975. Process framework for describing the morphologic and stratigraphic evolution of deltaic depositional systems. In Broussard, M. L. (ed.), Deltas, Models for Exploration. Houston Geological Society, Houston, TX, pp. 8698.Google Scholar
Ganti, V., Zhongxin, C., Lamb, M. P., Nittrouer, J. A., and Parker, G. 2016. Testing morphodynamic controls on the location and frequency of river avulsions on fans versus deltas: Huanghe (Yellow River), China. Geophysical Research Letters 41, 78827890. doi: 10.1002/2014GL061918.Google Scholar
García, N. O., and Mechoso, C. R. 2005. Variability in the discharge of South American rivers and in climate. Hydrological Sciences Journal 50. doi: 10.1623/hysj.50.3.459.65030.Google Scholar
García Martínez, M. F., Gottardi, G., Marchi, M., and Tonni, L. 2020. On the reactivation of sand boils near the Po River major embankments. In Calvetti, F., Cotecchia, F., Galli, A., and Jommi, C. (eds.), Geotechnical Research for Land Protection and Development. CNRIG 2019. Lecture Notes in Civil Engineering, vol. 40. Springer, Cham, pp. 328337.Google Scholar
García-Ruiz, J. M. 2010. The effects of land uses on soil erosion in Spain: a review. Catena 81, 115.Google Scholar
Garver, J. I., and Cockburn, J. M. H. 2009. A historical perspective of ice jams on the lower Mohawk River. In Proceedings from the 2009 Mohawk Watershed Symposium, Schenectady, NY, vol. 1, pp. 25–29.Google Scholar
Gasparini, N. M., Fischer, G. C., Adams, J. M., Dawers, N. J., and Janoff, A. M. 2016. Morphological signatures of normal faulting in low-gradient alluvial rivers in south-eastern Louisiana, USA. Earth Surface Processes and Landforms 41, 642657. doi: 10.1002/esp.3852.Google Scholar
Gastaldo, R. A., Allen, G. P., and Huc, A.-Y. 1995. The tidal character of fluvial sediments of the modern Mahakam River delta, Kalimantan, Indonesia. Special Publications International Association of Sedimentologists 24, 171181.Google Scholar
Gebrehiwot, S. G., Ellison, D., Bewket, W., Seleshi, Y., Inogwabini, B.‐I., and Bishop, K. 2019. The Nile Basin waters and the West African rainforest: rethinking the boundaries. WIREs Water 6, e1317. doi: 10.1002/wat2.1317.Google Scholar
Geerling, G. W., Kater, E., van den Brink, E. K. C., Baptist, M. J., Ragas, A. M. J., and Smits, A. J. M. 2008. Nature rehabilitation by floodplain excavation: the hydraulic effect of 16 years of sedimentation and vegetation succession along the Waal River, NL. Geomorphology 99, 317328Google Scholar
Gergel, S. E., Dixon, M. D., and Turner, M. G. 2002. Consequences of human-altered floods: levees, floods, and floodplain forests along the Wisconsin River. Ecological Applications 12, 17551770.Google Scholar
Gibson, S., Osorio, A., Creech, C., et al. 2019. Two pool-to-pool spacing periods on large sand-bed rivers: mega-pools on the Madeira and Mississippi. Geomorphology 328, 196210.Google Scholar
Gilbert, G. K. 1917. Hydraulic mining debris in the Sierra Nevada. U.S. Geological Survey Professional Paper 105. Washington, DC.Google Scholar
Gimeno, L., Nieto, R., Vázquez, M., and Lavers, D. A. 2014. Atmospheric rivers: a mini-review. Frontiers in Earth Science 2. doi: 10.3389/feart.2014.00002.Google Scholar
Giosan, L., Syvitski, J., Constantinescu, S., and Day, J. 2014. Climate change: protect the world’s deltas. Nature 516, 3133.Google Scholar
Givetz, E. H. 2010. Removing erosion control projects increases bank swallow (Riparia riparia) population viability modeled along the Sacramento River, California, USA. Biological Conservation 143, 428438.Google Scholar
Gleason, C. J. 2015. Hydraulic geometry of natural rivers: a review and future directions. Progress in Physical Geography: Earth and Environment 39, 337360.Google Scholar
Glynn, M. E., and Kuszmaul, J. 2010. Prediction of piping erosion along Middle Mississippi River levees – an empirical model. Final Report, U.S. Army Corps of Engineers, Waterways Experiment Station, ERDC/GSL TR-04-12, Vicksburg, MS.Google Scholar
Gomez, B. 2006. The potential rate of bed-load transport. Proceedings of the National Academy of Sciences USA 103(46): 1717017173.Google Scholar
Gomez, B., and Marron, D. C. 1991. Neotectonic effects on sinuosity and channel migration, Belle Fourche River, Western South Dakota. Earth Surface Processes and Landforms 16, 227335. doi: 10.1002/esp.3290160304.Google Scholar
Gomez, B., Phillips, J. D., Magilligan, F. J., and James, L. A. 1997. Floodplain sedimentation and sensitivity: summer 1993 flood, Upper Mississippi River Valley. Earth Surface Processes and Landforms 22, 923936.Google Scholar
Gong, S. L., Li, C., and Lang, S. L. 2008. The microscopic characteristics of Shanghai soft clay and its effect on soil body deformation and land subsidence. Environmental Geology 6. doi: 10.1007/s00254-008-1205-4.Google Scholar
Goodbred, S. L. J., and Saito, Y. 2012. Tide-dominated deltas. In Davis, R. A. J., and Dalrymple, R. W. (eds.), Principles of Tidal Sedimentology. Springer, Dordrecht, pp. 129149.Google Scholar
Górski, K., van den Bosch, L. V., van de Wolfshaar, K. E., et al. 2012. Post-damming flow regime development in a large lowland river (Volga, Russian Federation): implications for floodplain inundation and fisheries. River Research and Applications 28, 11211134.Google Scholar
Götz, E. 2008. Improved sediment-management strategies for the sustainable development of German waterways. Sediment Dynamics in Changing Environments (Proceedings of a symposium in Christchurch, New Zealand, December 2008). International Association of Hydrological Sciences, Publ. 325, pp. 540–549.Google Scholar
Gottschalk, M. K. E. 1971–1977. Stormvloeden en Rivieroverstromingen in Nederland, vol. 3. Van Gorcum & Comp. N.V., Assen.Google Scholar
Goudie, A. S. 2013. The Human Impact on the Environment: Past, Present and Future, 7th ed. Wiley, London.Google Scholar
Gouw, M. 2007. Alluvial architecture of the Holocene Rhine-Meuse delta (The Netherlands) and the lower Mississippi Valley (USA). Netherlands Geographical Studies 364. KNAG/Faculteit Ruimtelijke Wetenschappen Universiteit Utrecht.Google Scholar
Guan, B., and Waliser, D. E. 2015. Detection of atmospheric rivers: evaluation and application of an algorithm for global studies. Journal of Geophysical Research: Atmospheres 120, 1251412535. doi: 10.1002/2015JD024257.Google Scholar
Guan, B., Molotch, N. P., Waliser, D. E., Fetzer, E. J., and Neiman, P. J. 2010. Extreme snowfall events linked to atmospheric rivers and surface air temperature via satellite measurements. Geophysical Research Letters 37, L20401. doi: 10.1029/2010GL044696.Google Scholar
Graf, W. L. 1977. The rate law in fluvial geomorphology. American Journal of Science 277, 178191.Google Scholar
Graf, W. L. 1999. Dam nation: a geographic census of large American dams and their hydrologic impacts. Water Resources Research 35, 13051311.Google Scholar
Graf, W. L. 2001. Damage control: dams and the physical integrity of America’s rivers. Annals of the Association of American Geographers 91, 127.Google Scholar
Graf, W. L. 2002. Rivers, dams, and willow flycatchers: a summary of their science and policy connections. Geomorphology 47, 169188.Google Scholar
Graf, W. L. 2003. The changing role of dams in water resources management. Universities Council on Water Resources, Water Resources Update 126, 5459.Google Scholar
Graf, W. L. 2006. Downstream hydrologic and geomorphic effects of large dams on American rivers. Geomorphology 79, 336360.Google Scholar
Graf, W. L., Wohl, E., Sinha, T., and Sabo, J. L. 2011. Sedimentation and sustainability of western American reservoirs. Water Resources Research 46. doi: 10.1029/2009WR008836.Google Scholar
Graham, O. P. 1992. Survey of land degradation in New South Wales, Australia. Environmental Management 16, 205223. doi: 10.1007/BF02393826.Google Scholar
Grall, C., Steckler, M. S., Pickering, J. L., et al. 2018. A base-level stratigraphic approach to determining Holocene subsidence of the Ganges–Meghna–Brahmaputra Delta plain. Earth and Planetary Science Letters 499, 2336.Google Scholar
Grams, P. E., and 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, 337360.Google Scholar
Gray, J. R., and Landers, M. N. 2014. Measuring suspended sediment. In Ahuja, S. (ed.), Comprehensive Water Quality and Purification, vol. 1. Elsevier, Philadelphia, PA, pp. 157204.Google Scholar
Gregory, K. J. 2006. The human role in changing river channels. Geomorphology 79, 172191.Google Scholar
Gregory, K. J. 2010. The Earth’s Land Surface: Landforms and Processes in Geomorphology. Sage, London.Google Scholar
Grenfell, M., Aalto, R., and Nicholas, A. 2012. Chute channel dynamics in large, sand-bed meandering rivers. Earth Surface Processes and Landforms 37, 315331. doi: 10.1002/esp.2257.Google Scholar
Grenfell, M. C., Nicholas, A. P., and Aalto, R. 2014. Meditative adjustment of river dynamics: the role of chute channels in tropical sand-bed meandering rivers. Sedimentary Geology 301, 93106. doi: 10.1016/j.segeo.2013.006.007.Google Scholar
Greulich, S., Franklin, S. B., Wasklewicz, T., and Grubaugh, J. 2007. Hydrogeomorphology and forest composition of sunrise towhead island in the lower Mississippi River. Southeastern Naturalist 6, 217234.Google Scholar
Grill, G., Lehner, B., Lumsdon, A. E., MacDonald, G. K., Zarfl, C., and Liermann, C. R. 2015. An index-based framework for assessing patterns and trends in river fragmentation and flow regulation by global dams at multiple scales. Environmental Research Letters 10, 1.Google Scholar
Grimardias, D., Guillard, J., and Cattanéo, F. 2017. Drawdown flushing of a hydroelectric reservoir on the Rhône River: impacts on the fish community and implications for the sediment management. Journal of Environmental Management 197, 239249. doi: 10.1016/j.jenvman.2017.03.096.Google Scholar
Guarino, J. 2013. Tribal advocacy and the art of dam removal: the lower Elwha Klallam and the Elwha Dams. American Indian Law Journal 11, 114145.Google Scholar
Guccione, M. J., van Arsdale, R. B., and Lynne, H. H. 2000. Origin and age of the Manilla high and associated Big Lake “sunklands” in the New Madrid seismic zone, northeastern Arkansas. Geological Society of America Bulletin 112, 579590.Google Scholar
Guida, R. J., Swanson, T. L., Remo, J. W. F., and Kiss, T. 2013. Strategic floodplain reconnection for the Lower Tisza River, Hungary: opportunities for flood-height reduction and floodplain-wetland reconnection. Journal of Hydrology 521, 274285.Google Scholar
Guimarães Nobre, G., Jongman, B., Aerts, J., and Ward, P. J. 2017. The role of climate variability in extreme floods in Europe. Environmental Research Letters 12, 084012. doi: 10.1088/1748-9326/aa7c22.Google Scholar
Guirguis, K., Gershunov, A., Shulgina, T., Clemesha, R. E. S., and Ralph, F. M. 2019. Atmospheric rivers impacting Northern California and their modulation by a variable climate. Climate Dynamics 52, 65696583.Google Scholar
Guntren, E. L. M., Oliver, A. J. M., and Keevin, T. M. 2016. Change in lower Mississippi River Secondary Channels: an Atlas of Bathymetric and Photographic Data. Mississippi River Geomorphology and Potamology Program, Report No. 8. St. Louis Division, U.S. Army Corps of Engineers.Google Scholar
Gurnell, A. M., Piegay, H., Swanson, F. J., and Gregory, S. V. 2002. Large wood and fluvial processes. Freshwater Biology 47, 601619.Google Scholar
Gurnell, A. M. Corenblit, D., de Jalón, D. G., et al. 2016. A conceptual model of vegetation-hydrogeomorphology interactions within river corridors. River Research and Applications 32, 142163.Google Scholar
Gupta, A. 2007. Introduction. In Gupta, A. (ed.), Large Rivers: Geomorphology and Management. John Wiley & Sons, Chichester, pp. 15.Google Scholar
Gupta, N., Kleinhans, M. G., Addink, E. A., Atkinson, P. M., and 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
Ha, D. T., Ouillon, S., and van Vinh, G. 2018. Water and suspended sediment budgets in the lower Mekong from High-Frequency Measurements (2009–2016). Water 10, 846. doi: 10.3390/w10070846.Google Scholar
Habersack, H., Jäger, E., and Hauer, C. 2013. The status of the Danube River sediment regime and morphology as a basis for future basin management. International Journal of River Basin Management 11, 153166.Google Scholar
Habersack, H., Hein, T., Stanica, A., et al. 2016. Challenges of river basin management: current status of, and prospects for, the River Danube from a river engineering perspective. Science of the Total Environment 543, 828845.Google Scholar
Hackney, C. R., Darby, S. E., Parsons, D. R., et al. 2020. River bank instability from unsustainable sand mining in the lower Mekong River. Nature Sustainability 3, 217225. doi: 10.1038/s41893–019-0455-3.Google Scholar
Hanel, M., Rakovec, O., Markonis, Y., et al. 2018. Revisiting the recent European droughts from a long-term perspective. Scientific Reports 8, 9499. doi: 10.1038/s41598–018-27464-4.Google Scholar
Hanberry, B. B., Kabrick, J. M., and He, H. S. 2015. Potential tree and soil carbon storage in a major historical floodplain forest with disrupted ecological function. Perspectives in Plant Ecology, Evolution and Systematics 17, 1723.Google Scholar
Hannan, A. 1969. Study of Mississippi River bends. Unpublished doctoral dissertation, Colorado State University, Ft. Collins, CO.Google Scholar
Hanson, S., Nicholls, R., Ranger, N., et al. 2011. A global ranking of port cities with high exposure to climate extremes. Climatic Change 104, 89111. doi: 10.1007/s10584-010-9977-4.Google Scholar
Happ, S. C., Rittenhouse, G., and Dobson, G. C. 1940. Some principles of accelerated stream and valley sedimentation. US Department of Agriculture, Technical Bulletin 695.Google Scholar
Haraguchi, M., and Lall, U. 2015. Flood risks and impacts: a case study of Thailand’s floods in 2011 and research questions for supply chain decision making. International Journal of Disaster Risk Reduction 14, 256272.Google Scholar
Harbor, D. J., Schumm, S. A., and Harvey, M D. 1994. Tectonic control of the Indus River. In Schumm, S. A., and Winkley, B. (eds), The Variability of Large Alluvial Rivers. American Society of Civil Engineers, New York, pp. 161176.Google Scholar
Hardiman, S. C., Dunstone, N. J., Scaife, A. A., et al. 2018. The asymmetric response of Yangtze river basin summer rainfall to El Niño/La Niña. Environmental Research Letters 13, 024015. doi: 10.1088/1748-9326/aaa172.Google Scholar
Harmar, O. P., Clifford, N. J., Thorne, C. R., and Biedenharn, D. S. 2005. Morphological changes of the lower Mississippi River: geomorphological response to engineering intervention. River Research and Applications 21, 11071131.Google Scholar
Harrison, L. R., Dunne, T., and 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, 12391253. doi: 10.1002/esp.3717.Google Scholar
Harrison, R. W. 1950. Flood control in the Yazoo-Mississippi Delta. Southern Economic Journal 17, 1481558.Google Scholar
Havakes, H., Koster, M., Dekking, W., Uijterlinde, R., Wensink, W., and Walker, R. 2015. Water Governance: The Dutch Water Authority Model. Dutch Water Authorities, The Hague.Google Scholar
Hayden, B. 1988. Flood hydroclimatology. In Baker, V. R., Kochel, R. C., and Patton, P. C. (eds.), Flood Geomorphology. John Wiley & Sons, New York, pp. 1326.Google Scholar
He, S., Gao, Y., Li, F., Wang, H., and He, Y. 2017. Impact of Arctic Oscillation on the East Asian climate: a review. Earth-Science Reviews 164, 4862.Google Scholar
Heath, R. C. 1983. Basic ground-water hydrology. U.S. Geological Survey Water-Supply Paper 2220.Google Scholar
Heimann, D. C. 2016. Generalized sediment budgets of the lower Missouri River, 1968–2014. U.S. Geological Survey Scientific Investigations Report 2016–5097. doi: 10.3133/sir20165097.Google Scholar
Hein, T., Funk, A., Pletterbauer, F., et al. 2018. Management challenges related to long‐term ecological impacts, complex stressor interactions, and different assessment approaches in the Danube River Basin. River Research and Applications. doi: 10.1002/rra.3243.Google Scholar
Heine, R. A., and Pinter, N. 2012. Levee effects upon flood levels: an empirical assessment. Hydrological Processes 26, 32253240. doi: 10.1002/hyp.8261.Google Scholar
Heinz Center. 2002. Dam removal: science and decision making. Final Report. The H. John Heinz III Center for Science, Economics, and the Environment.Google Scholar
Heitmuller, F. T., and Greene, L. E. 2009. Historical channel adjustment and estimates of selected hydraulic values in the lower Sabine River and lower Brazos River Basins, Texas and Louisiana: U.S. Geological Survey Scientific Investigations Report 2009–5174, 143 p.Google Scholar
Heitmuller, F. T., Hudson, P. F., and Kesel, R. H. 2017. Overbank sedimentation from the historic AD 2011 flood along the lower Mississippi River, USA. Geology 45, 107110.Google Scholar
Helmholtz Centre for Environmental Research – UFZ. 2017. New standards for better water quality in Europe: researchers present recommendations for revision of the EU Water Framework Directive. ScienceDaily, February 27, 2017. www.sciencedaily.com/releases/2017/02/170227100740.htm (accessed May 2020).Google Scholar
Hendon, H. H., Thompson, D. W. J., and Wheeler, M. C. 2007. Australian rainfall and surface temperature variations associated with the Southern Hemisphere Annular Mode. Journal of Climate 40, 24522467.Google Scholar
Hennessy, K. J., Suppiah, R., and Page, C. M. 1999. Australian rainfall changes, 1910–1995. Australian Meteorological Magazine 48, 113.Google Scholar
Henry, R. E., Sommer, T. R., and Goldman, C. R. 2010. Growth and methylmercury accumulation in juvenile Chinook salmon in the Sacramento River and its floodplain, the Yolo Bypass. Transactions of the American Fisheries Society 139, 550563.Google Scholar
Hensel, P. E., Day, J. W., and Pont, D. 1999. Wetland vertical accretion and soil elevation change in the Rhône River Delta, France: the importance of riverine flooding. Journal of Coastal Research 15, 668681.Google Scholar
Herget, J., Eckhard, B., Coch, T., Dix, E., Eggenstein, G., and Ewald, K. 2005. Engineering impact on river channels in the River Rhine catchment. Erdkunde 59, 294319.Google Scholar
Hesselink, A. W. 2002. History makes a river. Morphological development of the embanked floodplains of the Rhine and Meuse in the Netherlands in historical time. Netherlands Geographical Studies 292. KNAG/Faculteit Ruimtelijke Wetenschappen Universiteit Utrecht.Google Scholar
Hesselink, A. W., Weerts, H. J. T., and Berendsen, H. J. A. 2003. Alluvial architecture of the human-influenced river Rhine, The Netherlands. Sedimentary Geology 161, 229248.Google Scholar
Hewson, W. 1860. Principles and Practice of Embanking Lands from River-Floods as Applied to the “Levees” along the Lower Mississippi River. J. J. Reed, New York.Google Scholar
Hey, R. D., and Tovey, N. K. 1989. Processes of bank failure. In Hemphill, R. W., and Bramley, M. E. (eds.), Protection of River and Canal Banks. CIRIA/Butterworths, London, pp. 739.Google Scholar
Hickin, E. J. 1977. Hydraulic factors controlling channel migration. In Dacidson-Aenott, R., and Nicking, W. (eds.), Research in Fluvial Systems. Proceedings of the 5th Guelph Geomorph Symposium, Geobooks, Norwich, pp. 59–72.Google Scholar
Hickin, E. J. 1984. Vegetation and river channel dynamics. Canadian Geographer 28, 111126.Google Scholar
Hickin, E. J., and Nanson, G. C. 1975. The character of channel migration on the Beatton River, north-east British Columbia, Canada. Bulletin of the Geological Society of America 86, 487494.Google Scholar
Hickin, E. J., and Nanson, G. C. 1984. Lateral migration rates of river bends. Journal of Hydraulic Engineering, ASCE 110, 15571567.Google Scholar
Higgins, J. M., and Brock, W. G. 1999. Overview of reservoir release improvements at 20 TVA dams. Journal of Energy Engineering 125, 117.Google Scholar
Higgins, S. Overeem, I., Tanaka, A., and Syvitski, J. P. M. 2013. Land subsidence at aquaculture facilities in the Yellow River delta, China. Geophysical Research Letters 40, 38983902. doi: 10.1002/grl.50758.Google Scholar
Hirschboeck, K. K. 1988. Flood hydroclimatology. In Baker, V. R., Kochel, R. C., and Patton, P. C. (eds.), Flood Geomorphology. John Wiley & Sons, New York, pp. 2749.Google Scholar
Hirschboeck, K. K. 1991. Climate and floods, in National Water Summary 1988–1989, Hydrologic Events and Floods and Droughts. U.S. Geological Survey Water Supply Paper 2375, pp. 67–88.Google Scholar
Hockenos, P. 2018. A Balkan dam boom imperils Europe’s wildest rivers. Yale Environment 360. Published at the Yale School of Forestry and Environmental Studies. https://e360.yale.edu/features/a-balkan-dam-boom-imperils-europes-wildest-rivers (accessed October 14, 2019).Google Scholar
Hodgkins, G. A., Whitfield, P. H., Burn, D. H., et al. 2017. Climate-driven variability in the occurrence of major floods across North America and Europe. Journal of Hydrology 552, 704717.Google Scholar
Hoffmann, T., Erkens, G., Gerlach, R., Klostermann, J., and Lang, A. 2009. Trends and controls of Holocene floodplain sedimentation in the Rhine catchment. Catena 77, 96106.Google Scholar
Hoffmann, T., Schlummer, M., Notebaert, B., Verstraeten, G., and Korup, O. 2013. Carbon burial in soil sediments from Holocene agricultural erosion, Central Europe. Global Biogeochemical Cycles 27, 828835.Google Scholar
Hohensinner, S., Egger, G., Haidvogl, G., Jungwirth, M., Muhar, S., and Schmutz, S. 2007. Hydrological connectivity of a Danube river-floodplain system in the Austrian Machland: changes between 1812 and 1991. In, Floodplain Protection, Restoration, Management. Why and how? Editions Lavoisier SAS, Paris, pp. 53–69.Google Scholar
Holbrook, J., and Schumm, S. A. 1999. Geomorphic and sedimentary response of rivers to tectonic deformation. Tectonophysics 305, 287306.Google Scholar
Holmes, R. R. Jr., Koenig, T. A., and Karstensen, K. A. 2010. Flooding in the United States Midwest, 2008. U.S. Geological Survey Professional Paper 1775.Google Scholar
Hooke, J. M. 1979. An analysis of the processes of river bank erosion. Journal of Hydrology 42, 3962.Google Scholar
Hooke, J. M. 1986. The significance of mid-channel bars in an active meandering river. Sedimentology 33, 839850.Google Scholar
Hori, K., and Saito, Y. 2007. Classification, architecture, and evolution of Large-River Deltas. In Gupta, A. (ed.), Large Rivers: Geomorphology and Management. John Wiley & Sons, Chichester, pp. 7596.Google Scholar
Hortle, K. G., and Nam, S. 2017. Mitigation of the impacts of dams on fisheries – a primer. Mekong Development Series No. 7. Mekong River Commission, Vientiane Lao PDR.Google Scholar
Houben, P. 2008. Scale linkage and contingency effects of field-scale and hillslope-scale controls of long-term soil erosion: anthropogeomorphic sediment flux in agricultural loess watersheds of Southern Germany. Geomorphology 101, 172191.Google Scholar
Huang, H. Q., and Nanson, G. C. 1998. The influence of bank strength on channel geometry: an integrated analysis of some observations. Earth Surface Processes and Landforms 23, 865876.Google Scholar
Huang, Q. G., and Nanson, G. A. 2007. Why some alluvial rivers develop an anabranching pattern. Water Resources Research 43. doi: 10.1029/2006WR005223.Google Scholar
Huat, B. B. K., Kazemian, S., Prasad, A., and Barghchi, M. 2011. State of an art review of peat: general perspective. International Journal of the Physical Sciences 6, 19881996. doi: 10.5897/IJPS11.192.Google Scholar
Huat, B. B. K., Prasad, A., Asadi, A., and Kazemian, S. 2014. Geotechnics of Organic Soils and Peat. CRC Press.Google Scholar
Huđek, H., Žganecc, K., and Puscha, M. T. 2020. A review of hydropower dams in Southeast Europe – distribution, trends and availability of monitoring data using the example of a multinational Danube catchment subarea. Renewable and Sustainable Energy Reviews 117, 109434. doi: 10.1016/j.rser.2019.109434.Google Scholar
Hudson, P. F. 2002. Pool-riffle morphology in an actively migrating alluvial channel: the lower Mississippi River. Physical Geography 23, 154169.Google Scholar
Hudson, P. F. 2003. Event sequence and sediment exhaustion in the lower Pánuco basin, Mexico. Catena 52, 5676.Google Scholar
Hudson, P. F. 2004a. Geomorphic context of the prehistoric Huastec floodplain environments: lower Pánuco basin, Mexico. Journal of Archaeological Science 31, 653668.Google Scholar
Hudson, P. F. 2004b. The influence of the El Niño Southern Oscillation on suspended sediment load variability in a seasonally humid tropical setting: Pánuco Basin, Mexico. Geografiska Annaler: Series A, Physical Geography 85, 263275.Google Scholar
Hudson, P. F. 2017. Water engineering. In Richardson, D., Castree, N., Kobayashi, A., Liu, W., Goodchild, M., and Marston, R. (eds.), International Encyclopedia of Geography: People, Earth, Environment and Technology. American Association of Geographers, Wiley-Blackwell. DOI: 10.1002/9781118786352.wbieg0868.Google Scholar
Hudson, P. F. 2018. Towards integrated flood management along the lower Rhine and Mississippi Rivers and the international legacy of the 2005 New Orleans Hurricanes Katrina–Rita flood disaster. International Journal of River Basin Management 16, 273285.Google Scholar
Hudson, P. F. 2020. Biogeomorphic evolution of lower Mississippi islands: 1965–2015. In Boersema, M. P., Schielen, R. M. J., van Eijsbergen, E., and Rinsema, J. G. (eds.), Managing Changing Rivers: NCR Days 2020 Proceedings. Netherlands Centre for River Studies, Publication 44–2020.Google Scholar
Hudson, P. F., and Colditz, R. R. 2003. Flood delineation in a large and complex alluvial valley, lower Pánuco basin, Mexico. Journal of Hydrology 280, 229245.Google Scholar
Hudson, P. F., and Heitmuller, F. T. 2003. Local- and watershed-scale controls on the spatial variability of natural levee deposits in a large fine-grained floodplain: lower Pánuco basin, Mexico. Geomorphology 56, 255269.Google Scholar
Hudson, P. F., and Kesel, R. H. 2000. Channel migration and meander-bend curvature in the lower Mississippi River prior to major modification. Geology 28, 531534.Google Scholar
Hudson, P. F., and Kesel, R. H. 2006. Spatial and temporal adjustment of the lower Mississippi River to major human impacts. Zeitschrift für Geomorphologie, Supplementband 143, 1733.Google Scholar
Hudson, P. F., and Middelkoop, H. 2015. The Palimpsest of river-floodplain management and the role of geomorphology. In Hudson, P. F., and Middelkoop, H. (eds.), Geomorphic Approaches to Integrated Floodplain Management of Lowland Fluvial Systems in North America and Europe. Springer-Verlag, New York, pp. 337350.Google Scholar
Hudson, P. F., and Mossa, J. 1997. Suspended sediment transport effectiveness of three large impounded rivers, U.S. Gulf Coastal Plain. Environmental Geology 32, 263273.Google Scholar
Hudson, P. F., Middelkoop, H., and Stouthamer, E. 2008. Flood management along the Lower Mississippi and Rhine Rivers (The Netherlands) and the continuum of geomorphic adjustment. Geomorphology 101, 209236.Google Scholar
Hudson, P. F., Heitmuller, F. T., Leitch, M. B. 2012. Hydrologic connectivity of oxbow lakes along the lower Guadalupe River, Texas: the influence of geomorphic and climatic controls on the “Flood Pulse Concept.” Journal of Hydrology 414, 174183.Google Scholar
Hudson, P. F., Sounny-Slitine, M. A., and LaFevor, M. 2013. A new longitudinal approach to assess hydrologic connectivity: embanked floodplain inundation along the lower Mississippi River. Hydrological Processes 27, 21872196.Google Scholar
Hudson, P. F., van der Hout, E., and Verdaasdonk, M. 2019. (Re)Development of fluvial islands along the lower Mississippi River over five decades, 1965–2015. Geomorphology 331, 7891.Google Scholar
Humphreys, A. A., and Abbot, H. L. 1863. Report upon the physics and hydraulics of the Mississippi River, upon the protection of the alluvial region against overflow; and upon the deepening of the mouths. In Lippincott, J. B. (ed.), Professional Papers of the Topographical Engineers, 4. Philadelphia, PA, 456 p.Google Scholar
Hundey, E. J., and Ashmore, P. E. 2009. Length scale of braided river morphology. Water Resources Research 45, W08409. doi: 10.1029/2008WR007521.Google Scholar
Hurrell, J. W. 1995. Decadal trends in the North Atlantic Oscillation: regional temperatures and precipitation. Science 269, 676679.Google Scholar
Hupp, C. P., Demas, C. R., Kroes, D. E., Day, R. H., and Doyle, T. W. 2008. Recent sedimentation patterns within the central Atchafalaya Basin, Louisiana. Wetlands 28, 125140.Google Scholar
Ibàñez, C., Canicio, A., Day, J. W., and Curcó, A. 1997. Morphologic development, relative sea level rise and sustainable management of water and sediment in the Ebre Delta, Spain. Journal of Coastal Conservation 97, 191202. doi: 10.1007/BF02908194.Google Scholar
Ibáñez, C., Caiola, N., Rovira, A., and Real, M. 2012. Monitoring the effects of floods on submerged macrophytes in a large river. Science of the Total Environment 440, 132139.Google Scholar
IKSE. 2005. Die Elbe und ihr einzugsgebiet: Ein geographisch-hydrologischer und wasserwirtschaftlicher überblick (The Elbe and its catchment area: a geographical-hydrological and water management overview). Internationale Kommission zum Schutz der Elbe, Magdeburg. www.ikse-mkol.org/fileadmin/media/user_upload/D/06_Publikationen/07_Verschiedenes/2005_IKSE-Elbe-und-ihr-Einzugsgebiet.pdf (accessed June 5, 2020).Google Scholar
Independent Levee Investigation Team (ILIT). 2006. Investigation of the New Orleans flood protection systems in Hurricane Katrina on August 29, 2005, Final Report, July 31, 2006, www.ce.berkeley.edu/~new_orleans/ (accessed April 18, 2016).Google Scholar
Ingebritsen, S. E., Ikehara, M. E., Galloway, D. L., and Jones, D. R. 2000. Delta subsidence in California; the sinking heart of the state. US Geological Survey Fact Sheet 005–00.Google Scholar
Interagency Floodplain Management Review Committee. 1994. A blueprint for change. Sharing the challenge: floodplain management into the 21st century. Final Report to the Administration Floodplain Management Task Force, June 1994, Washington DC.Google Scholar
Interagency Levee Policy Review Committee (ILPRC). 2006. The national levee challenge: levees and the FEMA map modernization initiative. Stakeholder Committee Report on Levees. 2006. www.fema.gov/library/viewRecord.do?id=2677 (accessed February 5, 2016).Google Scholar
Interagency Performance Evaluation Task Force (IPET). 2007. Performance evaluation of the New Orleans and Southeast Louisiana Hurricane Protection System. Volume III – The Hurricane Protection System. U.S. Army Corps of Engineers. http://biotech.law.lsu.edu/katrina/ipet/.html (accessed May 15, 2016).Google Scholar
Interagency Performance Evaluation Task Force (IPET). 2009. Performance evaluation of the New Orleans and Southeast Louisiana Hurricane Protection System – Final Report, U.S. Army Corps of Engineers. http://permanent.access.gpo.gov/lps71007/ (accessed May 18, 2016)Google Scholar
International Commission for the Protection of the Danube River (ICPDR). 2015. Flood risk management plan for the Danube River Basin District. Document number 4.5, May 28, 2015, Vienna, Austria.Google Scholar
International Commission for the Protection of the Rhine (ICPR). 2013a. The Rhine and Its Catchment: An Overview. Koblenz, Germany.Google Scholar
International Commission for the Protection of the Rhine (ICPR). 2013b. Implementation of the sediment management plan. Report on the State of Implementation by End 2013. Report No. 212. Koblenz, Germany.Google Scholar
International Commission for the Protection of the Rhine (ICPR). 2015a. Strategy for the IRBD Rhine for adapting to climate change. Report No. 219. Koblenz, Germany.Google Scholar
International Commission for the Protection of the Rhine (ICPR). 2015b. Internationally coordinated flood risk management plan for the International River Basin District of the Rhine, Part A. December 2015. Koblenz, Germany.Google Scholar
International Commission for the Protection of the Rhine (ICPR). 2015c. Summary report on the Rhine Measurement Programme Biology 2012/2013 and National Assessments according to the WFD. Report No. 232. Koblenz, Germany.Google Scholar
International Commission for the Protection of the Rhine (ICPR). 2016. Assessment of flood risk reduction (Action Plan on Floods, Action Target 1) with due regard to types of measures and receptors of the Directive 2007/60/EC (FD) – synthesis report. Report No. 236. Koblenz, Germany.Google Scholar
International Commission for the Protection of the Rhine (ICPR). 2018. Salmon is progressing. www.iksr.org/en/topics/ecology/plants-and-animals/fish/salmon-is-progressing/ (accessed August 8, 2018)]Google Scholar
International Commission for the Protection of the Rhine (ICPR). 2020a. Assessment Rhine 2020. Koblenz, Germany.Google Scholar
International Commission for the Protection of the Rhine (ICPR). 2020b. Rhine 2040. The Rhine and its catchment: sustainably managed and climate-resilient. 16th Rhine Ministerial Conference, February 13, 2020, Amsterdam.Google Scholar
Ionita, M., Rimbu, N., Lohmann, G. 2011. Decadal variability of the Elbe River streamflow. International Journal of Climatology 31, 2230.Google Scholar
IWPDC. 2020. Moving on up – India now world’s 5th largest hydropower producer. International Water Power and Dam Construction, June 1, 2020.Google Scholar
Jacobson, R. B., and Oberg, K. A. 1997. Geomorphic changes on the Mississippi River flood plain at Miller City, Illinois, as a result of the flood of 1993. U.S. Geological Survey Circular 1120–J.Google Scholar
Jacobson, R. B., Blevins, D. W., and Bitner, C. J. 2009. Sediment regime constraints on river restoration – an example from the Lower Missouri River. In James, L. A., Rathburn, S. L., and Whittecar, G. R. (eds.), Management and Restoration of Fluvial Systems with Broad Historical Changes and Human Impacts. Geological Society of America Special Paper 451, Boulder, CO, pp. 122.Google Scholar
Jacobson, R. B., Linder, G., and Bitner, C. 2015. The role of floodplain restoration in mitigating flood risk, Lower Missouri River, USA. In Hudson, P. F., and Middelkoop, H. (eds.), Geomorphic Approaches to Integrated Floodplain Management of Lowland Fluvial Systems in North America and Europe. Springer-Verlag, New York, pp. 203243.Google Scholar
Jager, N. W., Challies, E., Kochskämper, E. et al. 2016. Transforming European water governance? participation and river basin management under the EU Water Framework Directive in 13 member states. Water 8, 156.Google Scholar
James, L. A. 1989. Sustained storage and transport of hydraulic mining sediment in the Bear River, California. Annals of the Association of American Geographers 79, 570592.Google Scholar
James, L. A. 2006. Bed waves at the basin scale: implications for river management and restoration. Earth Surface Processes and Landforms 31, 16921706.Google Scholar
James, L. A. 2011. Contrasting geomorphic impacts of pre-and post-Columbian land-use changes in Anglo America. Physical Geography 32, 399422.Google Scholar
James, L. A. 2013. Legacy sediment: definitions and processes of episodically produced anthropogenic sediment. Anthropocene 2, 1626.Google Scholar
James, L. A., and Lecce, S. A. 2013. Impacts of land-use and land-cover change on river systems. In Shroder, J. (editors in chief) and Wohl, E. (ed.), Treatise on Geomorphology, vol. 9. Fluvial Geomorphology. Academic Press, San Diego, CA, pp. 768793.Google Scholar
James, L. A., and Marcus, W. A. 2006. The human role in changing fluvial systems: retrospect, inventory and prospect. Geomorphology 79, 152171.Google Scholar
James, L. A., and Singer, M. B. 2008. Development of the lower Sacramento valley flood-control system: historical perspective. Natural Hazards Review 2008, 125135.Google Scholar
James, L. A., Singer, M. B., Ghoshal, S., and Megison, M. 2009. Historical channel changes in the lower Yuba and Feather Rivers, California: Long-term effects of contrasting river-management strategies. In James, L. A., Rathburn, S. L., and Whittecar, G. R. (eds.), Management and Restoration of Fluvial Systems with Broad Historical Changes and Human Impacts. Geological Society of America Special Paper 451, Boulder, CO, Chapter 5, pp. 57–81. doi: 10.1130/2009.2451(04).Google Scholar
James, L. A., Phillips, J. D., and Lecce, S. A. 2017. A centennial tribute to G.K. Gilbert’s Hydraulic Mining Débris in the Sierra Nevada. Geomorphology 294, 419.Google Scholar
Jha, A. K., Bloch, R., and Lamond, J. 2012. Cities and Flooding: A Guide to Integrated Urban Flood Risk Management for the 21st Century. World Bank, Washington, DC.Google Scholar
Jiang, B., Wang, F., and Ni, G. 2018. Heating impact of a tropical reservoir on downstream water temperature: a case study of the Jinghong Dam on the Lancang River. Water 2018, 10, 951.Google Scholar
Jiménez-Muñoz, J. C., Mattar, C., Barichivich, J., et al. 2016. Record-breaking warming and extreme drought in the Amazon rainforest during the course of El Niño 2015–2016. Scientific Reports 6, 33130. doi: 10.1038/srep33130.Google Scholar
Johnson, W. C. 1994. Woodland expansions in the Platte River, Nebraska: patterns and causes. Ecological Monographs 64, 4584. doi: 10.2307/2937055.Google Scholar
Johnson, W. C. 2002. Riparian vegetation diversity along regulated rivers: contribution of novel and relict habitats. Freshwater Biology 47, 749759.Google Scholar
Johnstone, G., and Sithirith, M. 2018. Tonle Sap: fisheries management case study. In Finlayson, C. M., Everard, M., Irvine, K. et al. (eds.), The Wetland Book. Springer, Dordrecht, pp. 10671073.Google Scholar
Jones, C., An, K., Blom, R. G., et al. 2016. Anthropogenic and geologic influences on subsidence in the vicinity of New Orleans, Louisiana. Journal of Geophysical Research: Solid Earth 121, 38673887. doi: 10.1002/2015JB012636.Google Scholar
Jones, J. A. A. 1989. Global Hydrology: Processes, Resources, and Environmental Management. Taylor & Francis, London.Google Scholar
Jones, L. S., and Schumm, S. A. 2009. Causes of avulsions: an overview. In Smith, N. D., and Rogers, J. (eds.), Fluvial Sedimentology VI. The International Association of Sedimentologists, Special publication 28. Blackwell, Oxford, pp. 171178.Google Scholar
Jonkman, S. N., Voortman, H. G., Jan Klerk, W., and van Vuren, S. 2018. Developments in the management of flood defences and hydraulic infrastructure in the Netherlands. Developments in the management of flood defences and hydraulic infrastructure in the Netherlands. Structure and Infrastructure Engineering 14, 895910. doi: 10.1080/15732479.2018.1441317.Google Scholar
Jordan, C., Tiede, J., Lojek, O., et al. 2019. Sand mining in the Mekong Delta revisited – current scales of local sediment deficits. Scientific Reports 9, 17823. doi: 10.1038/s41598–019–53804-z.Google Scholar
Jorgensen, D., and Renofalt, B. M. 2013. Damned if you do, dammed if you don’t: debates on dam removal in the Swedish media. Ecology and Society 18, 18.Google Scholar
Julien, P. Y. 1995. Erosion and Sedimentation. Cambridge University Press, Cambridge.Google Scholar
Junk, W. J., Bayley, P. B., and Sparks, R. E. 1989. The flood pulse concept in river-floodplain systems. In Dodge, D. P. (ed.), Proceedings of the International Large River Symposium (LARS), Honey Harbour, Ontario, Canada. . Special Publication of Canadian Fisheries and Aquatic Sciences, Ottawa, Canada, pp. 110127.Google Scholar
Kandus, P., and Quintana, R. D. 2018. The Paraná River Delta. In Finlayson, C. M., Everard, M., Irvine, K. et al. (eds.), The Wetland Book. Springer, Dordrecht, pp. 813821.Google Scholar
Karlsson, R., and Hansbro, S. 1981. Soil Classification and Identification, 1st ed. Swedish Academy for Building Research, Stockholm, Sweden.Google Scholar
Kasse, C., van Balen, R. T., Bohncke, S. J. P., Wallinga, , and Vreugdenhil, M. 2016. Climate and base-level controlled fluvial system change and incision during the last glacial–interglacial transition, Roer river, the Netherlands – western Germany. Netherlands Journal of Geosciences – Geologie en Mijnbouw 96, 7192.Google Scholar
Kearney, M. S., Riter, A. J. C., and Turner, E. R. 2011. Freshwater river diversions for marsh restoration in Louisiana: twenty‐six years of changing vegetative cover and marsh area. Geophysical Research Letters 38, L16405. doi: 10.1029/2011GL047847.Google Scholar
Keller, E. A., and Melhorn, W. N. 1978. Rhythmic spacing and origin of pools and riffles. Geological Society of America Bulletin 89, 723730.Google Scholar
Kemp, G. P., Willson, C. J., Rogers, D., and Binselam, A. 2014. Adapting to change in the Lowermost Mississippi River: implications for navigation, flood control and restoration of the delta ecosystem. In Perspectives on the Restoration of the Mississippi Delta. doi: 10.1007/978-94-017-8733-8_5.Google Scholar
Kerr, R. A. 2000. A North Atlantic climate pacemaker for the centuries. Science 288(5473), 19841986.Google Scholar
Kesel, R. H. 2003. Human modifications to the sediment regime of the lower Mississippi River flood plain. Geomorphology 56. doi: 10.1016/S0169-555X(03)00159-4.Google Scholar
Kesel, R. H. and McGraw, M. 2015. The role of floodplain geomorphology in policy and management decisions along the lower Mississippi River in Louisiana. In Hudson, P. F., and Middelkoop, H. (eds.), Geomorphic Approaches to Integrated Floodplain Management of Lowland Fluvial Systems in North America and Europe. Springer, Dordrecht, Chapter 13, pp. 321335.Google Scholar
Kesel, R. H., Dunne, K. C., McDonald, R. C., Allison, K. R., and Spicer, B. E. 1974. Lateral erosion and overbank deposition on the Mississippi River in Louisiana caused by 1973 flooding: Geology 2, 461464.Google Scholar
Kesel, R. H., Yodis, E. G., and McCraw, D. J. 1992. An approximation of the sediment budget of the lower Mississippi river prior to major human modification. Earth Surface Processes and Landforms 17, 711722.Google Scholar
Khanal, S., Lutz, A., Immerzeel, W. W., de Vries, H., Wanders, N., and van den Hurk, B. 2019. The impact of meteorological and hydrological memory on compound peak flows in the Rhine River basin. Atmosphere 10, 171. doi: 10.3390/atmos10040171.Google Scholar
Kidder, T. R., and Liu, H. 2017. Bridging theoretical gaps in geoarchaeology: archaeology, geoarchaeology, and history in the Yellow River valley, China. Archaeological and Anthropological Sciences 9, 15851602. doi: 10.1007/s12520-014-0184-5.Google Scholar
Kidder, T. R., and Zhuang, Y. 2015. Anthropocene archaeology of the Yellow River, China, 5000–2000 BP. The Holocene 25, 16271639.Google Scholar
Kijowska‐Strugała, M., Bucała‐Hrabia, A., and Demczuk, P. 2018. Long‐term impact of land use changes on soil erosion in an agricultural catchment (in the Western Polish Carpathians). Earth Surface Processes and Landforms 29, 18711884.Google Scholar
Kingston, D. G., McGregor, G. R., Hannah, D. M., and Lawler, D. M. 2006. River flow teleconnections across the northern North Atlantic region. Geophysical Research Letters 33, L14705. doi: 10.1029/2006GL026574.Google Scholar
Kiss, T., Amissah, G. J., and Fiala, K. 2019. Bank processes and revetment erosion of a large lowland river: case Study of the lower Tisza River, Hungary. Water 11, 1313. doi: 10.3390/w11061313.Google Scholar
Klasz, G., Reckendorfer, W., Gabriel, H., Baumgartner, C., Schmalfuss, R., and Gutknecht, D. 2014. Natural levee formation along a large and regulated river: the Danube in the national Park Donau-Auen, Austria. Geomorphology 215, 2033. doi: 10.1016/j.geomorph.2013.12.023.Google Scholar
Klein Goldewijk, K., Beusen, A., Doelman, J., and Stehfest, E. 2017. New anthropogenic land use estimates for the Holocene; HYDE 3.2. Earth Systems Science Data 9, 927953.Google Scholar
Kleinhans, M. G., and 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, 721728.Google Scholar
Kleinhans, M. G., Weerts, J. J. T., and Cohen, K. M. 2010. Avulsion in action: reconstruction and modelling sedimentation pace and upstream flood water levels following a Medieval tidal-river diversion catastrophe (Biesbosch, The Netherlands, 1421–1750 AD). Geomorphology 118, 6579.Google Scholar
Klimas, C. V., Smith, R. D., Raasch, J., and Saucier, R. T. 2005. Hydrogeomorphic classification of forested wetlands in the lower Mississippi Valley. In Fredrickson, L. H., King, S. L., and Kaminski, R. M. (eds.), Ecology and Management of Bottomland Hardwood Systems: The State of Our Understanding. University of Missouri-Columbia Gaylord Memorial Laboratory Special Publication No. 10, Puxico, Missouri, pp. 7791.Google Scholar
Klimas, C. V., Murray, E. O., Pagan, J., Langston, H. L., and Foti, T. 2011. A regional guidebook for applying the hydrogeomorphic approach to assessing functions of forested wetlands in the delta region of Arkansas, lower Mississippi River Alluvial Valley. Technical report ERDC/EL-TR-11–12, Version 2.0, U.S. Army Corps of Engineers, Research and Development Center, Vicksburg, Mississippi.Google Scholar
Knighton, A. D. 1998. Fluvial Forms and Processes – A New Perspective. Arnold, London.Google Scholar
Knighton, A. D., and Nanson, G. C. 1993. Anastomosis and the continuum of channel pattern. Earth Surface Processes and Landforms 18, 613625.Google Scholar
Knox, J. C. 1972. Valley alluviation in southwestern Wisconsin. Annals of the Association of American Geographers 62, 401410.Google Scholar
Knox, J. C. 1977. Human impacts on Wisconsin stream channels. Annals of the Association of American Geographers 67, 323342.Google Scholar
Knox, J. C. 1983. Responses of river systems to Holocene climates. In Wright, H. E. Jr., (ed.), Late Quaternary Environments of the United States, The Holocene, vol. 2. University of Minnesota Press: Minneapolis, pp. 2641.Google Scholar
Knox, J. C. 1993. Large increases in flood magnitude in response to modest changes in climate. Nature 361, 430432.Google Scholar
Knox, J. C. 2000. Sensitivity of modern and Holocene floods to climate change. Quaternary Science Reviews 19, 439457. doi: 10.1016/S0277-3791(99)00074-8.Google Scholar
Knox, J. C. 2006. Floodplain sedimentation in the Upper Mississippi Valley: natural versus human accelerated. Geomorphology 79, 286310. doi: 10.1016/j.geomorph.2006.06.031.Google Scholar
Knox, J. C., and Daniels, J. M. 2002. Watershed scale and the stratigraphic record of large floods. In House, P. K., Webb, R. H., Baker, V. R., and Levish, D. R. (eds.), Ancient Floods, Modern Hazards: Principles and Applications of Paleoflood Hydrology, vol. 5. American Geophysical Union, Washington, DC, pp. 237255.Google Scholar
Knox, R. L., and Latrubesse, E. M. 2016. A geomorphic approach to the analysis of bedload and bed morphology of the Lower Mississippi River near the Old River Control Structure. Geomorphology 268, 3547.Google Scholar
Ko, J.-Y., and Day, J. W. 2004. A review of ecological impacts of oil and gas development on coastal ecosystems in the Mississippi Delta. Ocean and Coastal Management 47, 597623.Google Scholar
Ko, J.-Y., Day, J. W., Wilkins, J. W., Heywood, J., and Lane, R. L. 2017. Challenges in collaborative governance for coastal restoration: lessons from the Caernarvon River Diversion in Louisiana. Coastal Management 45, 125142. doi: 10.1080/08920753.2017.1278145.Google Scholar
Koenig, F., Quick, I., and Vollmer, S. 2012. Defining quantitative morphological changes in large rivers for a sustainable and effective sediment management applied to the River Elbe, Germany. In Proceedings of the Tenth International Conference on Hydroscience & Engineering, Orlando, Florida, November 4–8, 2012.Google Scholar
Kolb, C. R. 1963. Sediments forming the bed and banks of the lower Mississippi River and their effect on river migration. Sedimentology 2, 227234.Google Scholar
Kolb, C. R. 1975. Geologic Control of Sand Boils along the Mississippi River Levees. Misc. Paper S 7522. U.S. Army Engineer Waterways Experiment Station, Vicksburg, MS.Google Scholar
Kolb, C. R., and Saucier, R. T. 1982. Engineering geology of New Orleans. Review of Engineering Geology 5, 7593.Google Scholar
Komori, D., Nakamura, S., Kiguchi, M., et al. 2012. Characteristics of the 2011 Chao Phraya River flood in Central Thailand. Hydrological Research Letters 6, 4146.Google Scholar
Kondolf, G. M. 1997. Hungry water: effects of dams and gravel mining on river channels. Environmental Management 21, 533551.Google Scholar
Kondolf, G. M., and Piégay, H. 2016. Tools in fluvial geomorphology: problem statement and recent practices. In Kondolf, M. G. and Piégay, H. (eds.), Tools in Fluvial Geomorphology. John Wiley & Sons, Chichester and Hoboken, NJ, Chapter 1, pp. 1–22.Google Scholar
Kondolf, G. M., Gao, Y., Annandale, G. W. et al. 2014a. Sustainable sediment management in reservoirs and regulated rivers: experiences from five continents, American Geophysical Union, Earth’s Future 2, 256280. doi: 10.1002/2013EF000184.Google Scholar
Kondolf, G. M., Rubin, Z. K., and Minear, J. T. 2014b. Dams on the Mekong: cumulative sediment starvation. Water Resources Research 50, 51585169.Google Scholar
Kong, D., Latrubesse, E. M., Miaoa, C., and Zhou, R. 2020. Morphological response of the Lower Yellow River to the operation of Xiaolangdi Dam, China. Geomorphology 350, 10.1016/j.geomorph.2019.106931Google Scholar
König, F., Quick, I., and Vollmer, S. 2012. Defining quantitative morphological changes in large rivers for a sustainable and effective sediment management applied to the River Elbe, Germany. In Proceedings Tenth International Conference of Hydroscience and Engineering, November 2012, Orlando, USA.Google Scholar
Konrad, C. P., and Dettinger, M. D. 2017. Flood runoff in relation to water vapor transport by atmospheric rivers over the western United States, 1949–2015. Geophysical Research Letters 44, 1145611462. doi: 10.1002/2017GL075399.Google Scholar
Konsoer, K. M., Rhoads, B. L., Langendoen, E. J., et al. 2016. Spatial variability in bank resistance to erosion on a large meandering, mixed bedrock-alluvial river. Geomorphology 252, 8097.Google Scholar
Koohafkan, P., Salman, M., and Casarotto, C. 2011. Investments in land and water. State of Land and Water Resources (SOLAW) Background Thematic Report No. 17. Food and Agriculture Organization of the United Nations, Rome, Italy.Google Scholar
Kooijmans, L. P. 1974. The Rhine/Meuse Delta: Four Studies on its Prehistoric Occupation and Holocene Geology. Doctoral dissertation, Institute of Archaeology, Leiden University.Google Scholar
Kornei, K. 2020. Europe’s rivers are the most obstructed on Earth. Eos 101, January 31, 2020. doi: 10.1029/2020EO139204.Google Scholar
Korotaev, V. N., Ivanov, V. V., and Sidorchuk, A. Y. 2004. Alluvial relief structure and bottom sediments of the lower Volga River. In Sediment Transfer through the Fluvial System (Proceedings of the Moscow Symposium, August 2004). International Association of Hydrological Sciences, Publication 288.Google Scholar
Korponai, K., Gyulai, I., Braun, M., Kövér, C., Papp, I., Forró, L. 2016. Reconstruction of flood events in an oxbow lake (Marótzugi-Holt-Tisza, NE Hungary) by using subfossil cladocerans remains and sediments. Advances in Oceanography and Limnology 7, 125–135. DOI: 10.4081/aiol.2016.6168Google Scholar
Kostaschuk, R., Best, J. L., Villard, P., Peakall, J., and Franklin, M. 2005. Measuring flow velocity and sediment transport with an acoustic Doppler current profiler. Geomorphology 68, 2537.Google Scholar
Koulouri, M., and Giourga, C. 2007. Land abandonment and slope gradient as key factors of soil erosion in Mediterranean terraced lands. Catena 69, 274281.Google Scholar
Krinitzsky, E. L. 1970. Radiography in the Earth Sciences and Soil Mechanics. Springer.Google Scholar
Krinitzsky, E. L., and Wire, J. C. 1964. Ground water in the alluvium of the lower Mississippi valley (upper and central areas). U.S. Army Corps of Engineers Waterways Experiment Station, TR 3–658, V 1–2.Google Scholar
Krinitzsky, E. L., Ferguson, J. S. Jr., and Smith, F. L. 1965. Geological investigation of the Yazoo Basin. U.S. Army Corps of Engineers, Waterways Experiment Station, Technical Report 3–480. Vicksburg, Mississippi.Google Scholar
Krützen, M. Beasley, I., Ackermann, C. Y., et al. 2018. Demographic collapse and low genetic diversity of the Irrawaddy dolphin population inhabiting the Mekong River. PLoS ONE 13. doi: 10.1371/journal.pone.0189200.Google Scholar
Kryjov, V. N., and Gorelits, O. V. 2019. Wintertime Arctic Oscillation and formation of river spring floods in the Barents Sea Basin. Russian Meteorology and Hydrology 44, 187195.Google Scholar
Kuhn, S., Migenda, W., Pfarr, U. 2016. The Integrated Rhine Programme: Flood Control and Restoration of Former Floodplains along the Upper Rhine, 5th ed. Ministry of the Environment, Climate Protection and the Energy Sector. Stuttgart, Baden-Württemberg, DL.Google Scholar
Kummu, M., and Varis, O. 2007. Sediment-related impacts due to upstream reservoir trapping, the Lower Mekong River. Geomorphology 85, 275293.Google Scholar
Kundzewicz, Z. W., Szwed, M., and Pińskwar, I. 2019. Climate variability and floods – a global review. Water 11, 1399. doi: 10.3390/w11071399.Google Scholar
Kunz, M. J., Wüest, A., Wehrli, B., Landert, J., and Senn, D. B. 2011. Impact of a large tropical reservoir on riverine transport of sediment, carbon, and nutrients to downstream wetlands. Water Resources Research 47, 116. doi: 10.1029/2011WR010996.Google Scholar
Kunz, M. J., Senn, D. B., Wehrli, B., Mwelwa, E. M., and Wüest, A. 2013. Optimizing turbine withdrawal from a tropical reservoir for improved water quality in downstream wetlands. Water Resources Research 49, 55705584. doi: 10.1002/wrcr.20358.Google Scholar
Labosier, C. F., and Quiring, S. M. 2013. Hydroclimatology of the Southeastern USA. Climate Research 57, 157171.Google Scholar
Laghari, A. N., Vanham, D., and Rauch, W. 2012. The Indus basin in the framework of current and future water resources management. Hydrology and Earth Systems Sciences 16, 10631083.Google Scholar
Lai, X., Yin, D., Finlayson, B. L. et al. 2017. Will river erosion below the Three Gorges Dam stop in the middle Yangtze? Journal of Hydrology 554, 2431.Google Scholar
Lal, R., and Moldenhauer, W. C. 1987. Effects of soil erosion on crop productivity. Critical Reviews in Plant Sciences 5, 303367. doi: 10.1080/07352688709382244.Google Scholar
Lambeets, K., Hendrickx, F., Vanacker, S., van Looy, K., Maelfait, J.-P., and Bonte, D. 2008. Assemblage structure and conservation value of spiders and carabid beetles from restored lowland river banks. Biodiversity Conservation 17, 31333148.Google Scholar
Lambert, A. M. 1985. The Making of the Dutch Landscape: An Historical Geography of the Netherlands, 2nd ed. Academic Press, London.Google Scholar
Lane, E. W. 1957. A study of the shape of channels formed by natural streams flowing in erodible materials. Missouri Rivers Division Sediment Series No. 9. U.S. Army Engineer Division, Missouri River, Corps of Engineers, Omaha, NE.Google Scholar
Langbein, W. B., and Leopold, L. B. 1966. River meanders – theory of minimum Variance. U.S. Geological Survey, Professional Paper 422-H.Google Scholar
Langeani, F., Casatti, L., Gameiro, H. S., and de Cerqueira Rossa-Feres, D. 2005. Riffle and pool fish communities in a large stream of southeastern Brazil. Neotropical Ichthyology 3, 305311.Google Scholar
Lara, A., Bahamondez, A., González Reyes, A., Muñoz, A. A., Cuq, E., and Ruiz Gómez, C. 2015. Reconstructing streamflow variation of the Baker River from tree-rings in Northern Patagonia since 1765. Journal of Hydrology 529, 511523.Google Scholar
Larsen, E. W. 2007. Sacramento River Ecological Flows Study. Meander migration modeling final report. Prepared for The Nature Conservancy, Chico, CA by Eric W. Larsen, Davis, CA.Google Scholar
Latrubesse, E. M. 2008. Patterns of anabranching channels: the ultimate end-member adjustment of mega rivers. Geomorphology 101, 130145.Google Scholar
Latrubesse, E. M. 2015. Large rivers, megafans and other Quaternary avulsive fluvial systems: a potential “who’s who” in the geological record. Earth Science Reviews 146, 130.Google Scholar
Latrubesse, E. M. and Franzinelli, E. 2002. The Holocene alluvial plain of the Middle Amazon River, Brazil. Geomorphology 44, 241257.Google Scholar
Latrubesse, E. M., and Restrepo, J. D. 2014. Sediment yield along the Andes: continental budget, regional variations, and comparisons with other basins from orogenic mountain belts. Geomorphology 216, 225233.Google Scholar
Latrubesse, E. M., Stevaux, J. C., and Sinha, R. 2005. Tropical rivers. Geomorphology 70, 187206.Google Scholar
Latrubesse, E. M., Amsler, M., Morais, R., and 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, 239252.Google Scholar
Latrubesse, E. M., Arima, E. Y., Dunne, T., et al. 2017. Damming the rivers of the Amazon Basin. Nature 546, 363369. doi: 10.1038/nature22333.Google Scholar
Lauro, C., Vich, A. I. J., and Moreiras, S. M. 2019. Streamflow variability and its relationship with climate indices in western rivers of Argentina. Hydrological Sciences Journal 64, 607619. doi: 10.1080/02626667.2019.1594820.Google Scholar
Lavers, D. A., and Villarini, G. 2015. The contribution of atmospheric rivers to precipitation in Europe and the United States. Journal of Hydrology 522, 382390.Google Scholar
Lavers, D. A., Allan, R. P., Wood, E. F., Villarini, G., Brayshaw, D. J., and Wade, A. J. 2011. Winter floods in Britain are connected to atmospheric rivers. Geophysical Research Letters 38, L23803. doi: 10.1029/2011GL049783.Google Scholar
Lawler, D. M., Thorne, C. R., and Hooke, J. M. 1997. Bank erosion and instability. In Thorne, C. R., Newson, M., and Hey, R. D. (eds.), Guidebook of Applied Fluvial Geomorphology for River Engineering and Management, Wiley, Chichester, pp. 137172.Google Scholar
Lawson, M. 2016. Dam removal: case studies on the fiscal, economic, social, and environmental benefits of dam removal. Headwater Economics. http://headwaterseconomics.org/economic-development/local-studies/dam-removal-case-studies (accessed May 20, 2020).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, 471486.Google Scholar
Lecce, S. A. 1997b. Spatial patterns of historical overbank sedimentation and floodplain evolution, Blue River Wisconsin. Geomorphology 18, 265277.Google Scholar
Lecce, S. A. 2000. Spatial variations in the timing of annual floods in the southeastern United States. Journal of Hydrology 235, 151169.Google Scholar
Lecce, S. A., and Pavlowsky, R. T. 2001. Use of mining-contaminated sediment tracers to investigate the timing and rates of historical floodplain sedimentation. Geomorphology 38, 85108.Google Scholar
Lechleitner, F., Breitenbach, S., Rehfeld, K., et al. 2017. Tropical rainfall over the last two millennia: evidence for a low-latitude hydrologic seesaw. Scientific Reports 7, 45809. doi: 10.1038/srep45809.Google Scholar
Lehner, B., Reidy Liermann, C., Revenga, C., et al. 2011. High-resolution mapping of the world’s reservoirs and dams for sustainable river-flow management. Frontiers in Ecology and the Environment 9, 494502. doi: 10.1890/100125.Google Scholar
Leigh, D. S. 2008. Late Quaternary climates and river channels of the Atlantic Coastal Plain, Southeastern USA. Geomorphology 101, 90108.Google Scholar
Leigh, D. S. 2018. Vertical accretion sand proxies of gaged floods along the upper Little Tennessee River, Blue Ridge Mountains, USA. Sedimentary Geology 364, 342350.Google Scholar
Lejon, A. G. C., Renöfält, B. M., and Nilsson, C. 2009. Conflicts associated with dam removal in Sweden. Ecology and Society 14(2).Google Scholar
Leli, I. T., Stevaux, J. C., and Assine, M. L. 2017. Genesis and sedimentary record of blind channel and islands of the anabranching river: an evolution model. Geomorphology 32, 3545. doi: 10.1016/j.geomorph.2017.05.001.Google Scholar
Lenhart, C. F. 2003. An assessment of NOAA community-based fish passage and dam removal projects. Coastal Management 231, 7796.Google Scholar
Leopold, L. B. 1998. Sediment Problems at Three Gorges Dam. International Rivers Network, Berkeley, CA.Google Scholar
Leopold, L. B., and Maddock, T. Jr. 1953. The hydraulic geometry of stream channels and some physiographic implications. U.S. Geological Survey, Professional Paper 252.Google Scholar
Leopold, L. B., and Wolman, M. G. 1957. River channel patterns: braided, meandering, and straight. U.S. Geological Survey, Professional Paper 282-B.Google Scholar
Leopold, L. B., and Wolman, M. G. 1960. River meanders. Geological Society of America Bulletin 71, 769793.Google Scholar
Lewin, J., and Ashworth, P. J. 2014a. The negative relief of large river floodplains. Earth Science Reviews 129, 123.Google Scholar
Lewin, J., and Ashworth, P. J. 2014b. Defining large river channel patterns: alluvial exchange and plurality. Geomorphology 215, 8398. doi: 10.1016/j.geomorph.2013.02.024.Google Scholar
Lewis, L. Y., Bohlen, C., and Wilson, S. 2008. Dams, dam removal, and river restoration: a hedonic property value analysis. Contemporary Economic Policy 26, 175186.Google Scholar
Lewin, J., Ashworth, P. J., and Strick, R. 2016. Spillage sedimentation on large river floodplains. Earth Surface Processes and Landforms 42. doi: 10.1002/esp.3996.Google Scholar
Li, T., Li, J., and Zhang, D. D. 2020. Yellow River flooding during the past two millennia from historical documents. Progress in Physical Geography: Earth and Environment 44, 661678. doi: 10.1177/0309133319899821.Google Scholar
Li, Y., Craven, J., Schweig, E. S., and Obermeier, S. F. 1996. Sand boils induced by the 1993 Mississippi River flood: could they one day be misinterpreted as earthquake-induced liquefaction? Geology 24, 171174.Google Scholar
Liermann, M., Pess, G., McHenry, M., et al. 2017. Relocation and recolonization of Coho Salmon in two tributaries to the Elwha River: implications for management and monitoring. Transactions of the American Fisheries Society 146, 946955.Google Scholar
Ligon, F. K., Dietrich, W. E., and Trush, W. J. 1995. Downstream ecological effects of dams: a geomorphic perspective. BioScience 45, 183192.Google Scholar
Lim, E.-P., Hendon, H. H., Arblaster, J.M., et al. 2016. The impact of the Southern Annular Mode on future changes in Southern Hemisphere rainfall, Geophysical Research Letters 43, 71607167. doi: 10.1002/2016GL069453.Google Scholar
Lindenschmidt, K.-E. 2018. Modelling probabilities of ice jam flooding from artificial breakup of the Athabasca River ice cover at Fort McMurray. In CGU HS Committee on River Ice Processes and the Environment, 19th Workshop on the Hydraulics of Ice Covered Rivers, Whitehorse, Yukon, Canada.Google Scholar
Lindenschmidt, K.-E., Carstensen, D., Fröhlich, W., et al. 2019. Development of an ice jam flood forecasting system for the Lower Oder River – requirements for real-time predictions of water, ice and sediment transport. Water 11, 95.Google Scholar
Liu, C., Walling, D. E., Spreafico, M., Ramasmy, J., Thulstrop, H. D., and Mishra, A. 2017. Sediment Problems and Strategies for Their Management: Experience from Several Large River Basins. United Nations Educational, Scientific and Cultural Organization (UNESCO), Paris.Google Scholar
Liu, P., Li, Q., Li, Z., Hoey, T., Liu, Y., and Wang, C. 2015. Land subsidence over oilfields in the Yellow River Delta. Remote Sensing 7, 15401564. doi: 10.3390/rs70201540.Google Scholar
LOLA. 2014. Dutch Dikes. LOLA Landscape Architects, Rotterdam.Google Scholar
Loomis, J. 2002. Quantifying recreation use values from removing dams and restoring free flowing rivers: a contingent behavior travel cost demand model for the Lower Snake River. Water Resources Research 38, 2-1.Google Scholar
Lopez, J. A., Henkel, T. K., Moshogianis, A. M., et al. 2014. Examination of deltaic processes of Mississippi River outlets – Caernarvon delta and Bohemia spillway in Southeastern Louisiana. Gulf Coast Association of Geological Societies 3, 7993.Google Scholar
Lorenz, S., Leszinski, M., and Graeber, D. 2016. Meander reconnection method determines restoration success for macroinvertebrate communities in a German lowland river: meander reconnection method. International Review of Hydrobiology 101(3–4). doi: 10.1002/iroh.201501823.Google Scholar
Lorimer, J., and Driessen, C. 2014. Wild experiments at the Oostvaardersplassen: rethinking environmentalism in the Anthropocene. Transactions of the British Geographers 39, 169181. doi: 10.1111/tran.12030.Google Scholar
Louisiana Department of Natural Resources (LDNR). 2006. Caernarvon Freshwater Diversion Project, 2005 Annual Report.Google Scholar
Lower Mississippi River Conservation Committee (LMRCC). 2015. Restoring America’s Greatest River: a habitat restoration plan for the lower Mississippi River. Vicksburg, MS. http://lmrcc.org.Google Scholar
Lott, N. 1993. The summer of 1993: flooding in the Midwest and Drought in the Southeast. National Climatic Data Center, Technical Report 93–04.Google Scholar
Lu, X., Kummu, M., and Oeurng, C. 2014. Reappraisal of sediment dynamics in the Lower Mekong River, Cambodia. Earth Surface Processes and Landforms 39, 18551865.Google Scholar
Lyell, C. 1837. Principles of Geology, Being an Inquiry How Far the Former Changes of the Earth’s Surface Are Referable to Causes Now in Operation, vol. 1. James Kay Jr and Brother: Philadelphia, PA.Google Scholar
Ma, H., Nittrouer, J. A., Naito, K., et al. 2017. The exceptional sediment load of fine-grained dispersal systems: example of the Yellow River, China. Science Advances 3, e1603114. doi: 10.1126/sciadv.1603114.Google Scholar
Ma, Y., Huang, H., Nanson, G. C., Li, Y., and Yao, W. 2012. Channel adjustments in response to the operation of large dams: the upper reach of the lower Yellow River. Geomorphology 147–148, 3548.Google Scholar
Maavara, T., Parsons, C. T., Ridenour, C., et al. 2015. Global phosphorus retention by river damming, Proceedings National Academy of Sciences 112, 1560315608.Google Scholar
Macdonald, D., Dixon, A., Newell, A., and Hallaways, A. 2012. Groundwater flooding within an urbanized flood plain. Journal of Flood Risk Management 5, 6880. doi: 10.1111/j.1753–318X.2011.01127.x.Google Scholar
Macklin, M. G., Lewin, J., and Woodward, J. C. 2012. The fluvial record of climate change. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. doi: 10.1098/rsta.2011.0608.Google Scholar
Magilligan, F. J. 1992. Sedimentology of a fine-grained aggrading floodplain. Geomorphology 4, 393408.Google Scholar
Magilligan, F. J., Phillips, J. D., Gomez, B., and James, L. A. 1998. Geomorphic and sedimentological controls on the effectiveness of an extreme flood, Journal of Geology 106, 8795.Google Scholar
Magilligan, F. J., Nislow, K. H., and Graber, B. E. 2003. A scale-independent assessment of discharge reduction and riparian disconnectivity following flow regulation by dams. Geology 31, 569572.Google Scholar
Magilligan, F. J., Haynie, H. J., and Nislow, K. H. 2008. Channel adjustments to dams in the Connecticut River Basin: implications for forested mesic watersheds. Annals of the American Association of Geographers 98, 267284.Google Scholar
Magilligan, F. J., Graber, B. E., Nislow, K. H., Chipman, J. W., Sneddon, C. S., and Fox, C. A. 2016. River restoration by dam removal: Enhancing connectivity at watershed scales. River restoration by dam removal. Elementa: Science of the Anthropocene 000108. doi: 10.12952/journal.elementa.000108.Google Scholar
Magilligan, F., Sneddon, C., and Fox, C. 2017. The social, historical, and institutional contingencies of dam removal. Environmental Management 59, 982994.Google Scholar
Major, J. J., East, A., O’Conner, J. et al. 2017. Geomorphic responses to dam removal in the United States – a two-decade perspective. In Tsutsumi, D., and Laronne, J. B. (eds.), Gravel-Bed Rivers: Processes and Disasters. John Wiley & Sons, Chichester, pp. 355383.Google Scholar
Makaske, B. 2001. Anastomosing rivers: a review of their classification, origin and sedimentary products. Earth Science Reviews 53, 149196.Google Scholar
Makaske, B., Berendsen, H. J. A., and van Ree, M. H. M. 2007. Middle Holocene avulsion-belt deposits in the Central Rhine-Meuse delta, The Netherlands. Journal of Sedimentary Research. Section A, Sedimentary Petrology and Process 77, 110123.Google Scholar
Makaske, B., Maas, G. J., van den Brink, C., and Wolfert, H. P. 2011. The influence of floodplain vegetation succession on hydraulic roughness: is ecosystem rehabilitation in Dutch embanked floodplains compatible with flood safety standards? AMBIO 40, 370376. doi: 10.1007/s13280-010-0120-6.Google Scholar
Malik, S., and Pal, S. C. 2019. Impact of groyne on channel morphology and sedimentology in an ephemeral alluvial river of Bengal Basin. Environmental Earth Sciences 78, 631. doi: 10.1007/s12665–019–8642-0.Google Scholar
Maltsev, S. A. 2009. Conservation of the sturgeon fish in lower Volga. In Carmona, R., Domezain, A., García-Gallego, M., Hernando, J. A., Rodríguez, F., and Ruiz-Rejón, M. (eds.), Biology, Conservation and Sustainable Development of Sturgeons. Fish and Fisheries Series, vol. 29. Springer, Dordrecht, pp. 265274.Google Scholar
Mann, C. C. 2005. 1491: New Revelations of the Americas before Columbus. Alfred E. Knopf, New York.Google Scholar
Mansur, C. I., Caufman, R. I., and Schultz, J. R. 1956. Investigation of Underseepage and Its Control on Lower Mississippi River Levees, 2 vols. US Army Corps of Engineers, Waterways Experiment Station Technical Memorandum, TM 3–424, Vicksburg, MS.Google Scholar
Mantua, N. J., Hare, S. R., Zhang, Y., Wallace, J. M., and Francis, R.C. 1997. A Pacific interdecadal climate oscillation with impacts on salmon production. Bulletin of the American Meteorological Society 78, 10691079.Google Scholar
Mariotti, A. 2007. How ENSO impacts precipitation in southwest central Asia. Journal of Geophysical Research Letters 34. doi: 10.1029/2007GL030078.Google Scholar
Markham, E. M. 1916. Results of experiments looking to the development of a form of subaqueous concrete revetment for the protection of river banks against scour or erosion. Professional Memoirs, Corps of Engineers, United States Army, and Engineer Department at Large 8, No. 42, pp. 721–733.Google Scholar
Mardhiah, U., Tillig, M., and Gurnell, A. 2015. Reconstructing the development of sampled sites on fluvial island surfaces of the Tagliamento River, Italy, from historical sources. Earth Surface Processes and Landforms 40, 629641. doi: 10.1002/esp.3658.Google Scholar
Marshall, G. J. 2003. Trends in the southern annular mode from observations and reanalysis. Journal of Climate 16, 41344143.Google Scholar
Marston, R. A., Mills, J. D., Wrazien, D. R., Bassett, B., and Splinter, D. K. 2005. Effects of Jackson Lake Dam on the Snake River and its floodplain, Grand Teton National Park, Wyoming, USA. Geomorphology 71, 7998.Google Scholar
Martin, S. M., Dunbar, J. B., Corcoran, M. K., and Schmitz, D. W. 2017. Geologic controls of sand boil formation at Buck Chute, Mississippi. Final report. U.S. Army Corps of Engineers, Waterways Experiment Station, Vicksburg, MS.Google Scholar
Mason, J., and Mohrig, D. 2019. Scroll bars are inner bank levees along meandering river bends. Earth Surface Processes and Landforms 44, 26492659.Google Scholar
Massey, W., Biron, P. M., and Choné, G. 2017. Impacts of river bank stabilization using riprap on fish habitat in two contrasting environments. Earth Surface Processes and Landforms 42, 635646. doi: 10.1002/esp.4010.Google Scholar
Matthai, H. F. 1967, Measurement of peak discharge at width contractions by indirect methods. U.S. Geological Survey Techniques of Water-Resources Investigations, book 3, chapter A4.Google Scholar
Maurice-Bourgoin, L., Bonnet, M.-P., Martinez, J.-M., et al. 2007. Temporal dynamics of water and sediment exchanges between the Curuaí floodplain and the Amazon River, Brazil. Journal of Hydrology 335, 140156. doi: 10.1016/j.hydrol.2006.11.023.Google Scholar
Mazi, K., Koussis, A. D., and Destouni, G. 2013. Tipping points for seawater intrusion in coastal aquifers under rising sea level. Environmental Research Letters 8, 014001.Google Scholar
Mazzotti, S., Lambert, A., van der Kooij, M., and Mainville, A. 2009. Impact of anthropogenic subsidence on relative sea-level rise in the Fraser River delta. Geology 37, 771774. doi: 10.1130/G25640A.1.Google Scholar
McBride, R. A., Taylor, M. J., and Byrnes, M. R. 2007. Coastal morphodynamics and Chenier-Plain evolution in southwestern Louisiana, USA: a geomorphic model. Geomorphology 88, 367422. doi: 10.1016/j.geomorph.2006.11.013.Google Scholar
McCabe, D. J. 2011. Rivers and streams: life in flowing water. Nature Education Knowledge 3, 19.Google Scholar
McGrath, B., Tachakitkachorn, T., and Thitakoo, D. 2013. Bangkok’s distributary waterscape urbanism: from a tributary to a distributary system. In Shannon, K., and de Meulder, B. (eds.), Water Urbanisms East, Emerging Practices and Age-old Traditions, UFO Explorations of Urbanism. Park Books, Zurich, pp. 4863.Google Scholar
McGregor, G. 2017. Hydroclimatology, modes of climatic variability and stream flow, lake and groundwater level variability: a progress report. Progress in Physical Geography 41, 496512.Google Scholar
McLeman, R. A., Dupre, J., Ford, L. B., Ford, J., Gajewski, K., and Marchildon, G. 2014. What we learned from the Dust Bowl: lessons in science, policy, and adaptation. Population and Environment 35, 417440. doi: 10.1007/s11111–013–0190-z.Google Scholar
McPhee, J. 1989. The Control of Nature. Farrar, Straus, and Giroux, New York, 272 pp.Google Scholar
Meade, R. H. 1995. Setting: Geology, hydrology, sediments, and engineering of the Mississippi River. In R. H. Meade (Ed.), Contaminants in the Mississippi River, 1987–1992. US Geological Survey Circular 1133. Reston, Virginia.Google Scholar
Meade, R. H. 2007. Transcontinental moving and storage: the Orinoco and Amazon Rivers transfer the Andes to the Atlantic, In Gupta, A. (ed.), Large Rivers: Geomorphology and Management. John Wiley & Sons, Chichester, pp. 45–63.Google Scholar
Meade, R. H., and Moody, J. 2010. Causes for the decline of suspended‐sediment discharge in the Mississippi River system, 1940–2007. Hydrological Processes 24, 3449.Google Scholar
Meadows, M. E., and Hoffman, M. T. 2010. The nature, extent and causes of land degradation in South Africa: legacy of the past, lessons for the future? Area 34, 428437.Google Scholar
Mechoso, C. R., Dias, P. S., Baethgen, W., et al. 2001. Climatology and hydrology of the Plata Basin. In Berbery, E. H., and Mechoso, R. (eds.), Document of VAMOS Scientific Study Group on the Plata Basin. World Climate Research Programme (WCRP). www2.atmos.umd.edu/~berbery/laplata/ (accessed October 3, 2019).Google Scholar
Mehta, V. M. 2017. River flow and its impacts, Ch. 5. In Natural Decadal Climate Variability: Societal Impacts. CRC Press, Taylor and Francis, pp. 113170.Google Scholar
Mertes, L. A. K. 1997. Documentation and significance of the perirheic zone on inundated floodplains. Water Resources Research 33, 17491762.Google Scholar
Mertes, L. A. K., and Dunne, T. 2007. The effects of tectonics, climatic history, and sea-level history on the form and behaviour of the modern Amazon River. In Gupta, A. (ed.), Large Rivers. Wiley, Chichester, pp. 115144.Google Scholar
Meyer, A., Combroux, I., Schmitt, L., and Trémoliéres, M. 2013. Vegetation dynamics in side-channels reconnected to the Rhine River: what are the main factors controlling communities trajectories after restoration? Hydrobiologia 714, 3547.Google Scholar
Middelkoop, H. 1997. Embanked Floodplains in the Netherlands: Geomorphological Evolution over Various Time Scales. Netherlands Geographical Studies 34. Utrecht University, Utrecht.Google Scholar
Middelkoop, H. 2000. Heavy-metal pollution of the river Rhine and Meuse floodplains in the Netherlands. Geologie en Mijnbouw / Netherlands Journal of Geosciences 79, 411428.Google Scholar
Middelkoop, H., and Asselman, N. E. M. 1998. Spatial variability of floodplain sedimentation at the event scale in the Rhine–Meuse delta, The Netherlands. Earth Surface Processes and Landforms 23, 561573.Google Scholar
Middelkoop, H., and van Haselen, C. O. G. 1999. Twice a river. Rhine and Meuse in the Netherlands. RIZA Report No. 99.033. RIZA, Arnhem.Google Scholar
Middelkoop, H., Daamen, K., Gellens, D., et al. 2001. Impact of climate change on hydrological regimes and water resources management in the Rhine Basin. Climatic Change 49, 105128.Google Scholar
Middelkoop, H., Kwadijk, J. C. J., van Deursen, W. P. A., and van Asselt, M. B. A. 2002. Scenario analyses in global change assessment for water management in the lower Rhine delta. In Beniston, M. (ed.), Climatic Change: Implications for the Hydrological Cycle and for Water Management. Advances in Global Change Research. Kluwer, Dordrecht, pp. 445463.Google Scholar
Middelkoop, H., Erkens, G., and van der Perk, M. 2010. The Rhine delta – a record of sediment trapping over time scales from millennia to decades. Journal of Soils and Sediments 10, 628639.Google Scholar
Middelkoop, H., Alabyan, A. M., Babich, D. B., and Ivanov, V. V. 2015. Post-dam channel and floodplain adjustments along the Lower Volga River, Russia. In Hudson, P. F., and Middelkoop, H. (eds.), Geomorphic Approaches to Integrated Floodplain Management of Lowland Fluvial Systems in North America and Europe. Springer-Verlag, New York, pp. 245264.Google Scholar
Millard, C., Hajek, E., and Edmonds, D. A. 2017. Evaluating controls on crevasse-splay size: implications for floodplain-basin filling. Journal of Sedimentary Research 87, 722739.Google Scholar
Milliman, J. D., and Farnsworth, K. L. 2011. River Discharge to the Coastal Ocean: A Global Synthesis. Cambridge University Press.Google Scholar
Milliman, J. D., and Meade, R. H. 1983. Worldwide delivery of river sediment to the oceans. Journal of Geology 91, 121.Google Scholar
Milliman, J. D., and Syvitski, J. 1992. Geomorphic tectonic control of sediment discharge to ocean – the importance of small mountainous rivers. The Journal of Geology 100, 525544.Google Scholar
Minderhoud, P. S. J., Middelkoop, H., Erkens, G., and Stouthamer, E. 2020. Groundwater extraction may drown mega-delta: projections of extraction-induced subsidence and elevation of the Mekong delta for the 21st century. Environmental Research Communications 2, 011005.Google Scholar
Miranda, L. E. 2005. Fish assemblages in oxbow lakes in connectivity to the Mississippi River. Transactions American Fisheries Society 134, 14801489.Google Scholar
Mississippi River Commission (MRC). Hydrographic surveys of the lower Mississippi River, Cairo, IL to Red River Landing, 84 plates, map scale 1:10,000 1948, 1963, 1975, 1988, 1999.Google Scholar
Mississippi River Commission (MRC). 1880. Report of the Mississippi River Commission, Appendix SS, H. R. Ex. Doc. 95, 46th Congress, 2nd Session.Google Scholar
Mississippi River Commission (MRC). 2008. The Mississippi River and Tributaries Project: Yazoo Backwater Area. Information Paper, November 2008, Vicksburg, Mississippi. www.mvd.usace.army.mil (accessed April 3, 2016).Google Scholar
Mississippi River Commission (MRC). 2017. Internet web page “Mississippi River and Tributaries Project.” www.mvd.usace.army.mil/mrc/mrt/index.php (accessed November 26, 2017).Google Scholar
MITECO. 2020. Sistema Nacional de Cartografia de Zonas Inundables: Inventario de Presas y Embalses (SNCZI-IPE). Spanish Ministry of Environment. https://sig.mapama.gob.es/snczi/ (accessed April 2020).Google Scholar
Mitsch, W. J., Day, J. W. Jr., Giliam, W., et al. 2001. Reducing nitrogen loading to the Gulf of Mexico from the Mississippi River Basin: strategies to counter a persistent ecological problem: BioScience 51, 373388. doi: 10.1641/00063568(2001)051[0373: RNLTTG]2.0.CO;2.Google Scholar
Moffatt, K. C., Crone, E. E., Holl, K. D., Schlorff, R. W., and Garrison, B. A. 2005. Importance of hydrologic and landscape heterogeneity for restoring bank swallow (Riparia riparia) colonies along the Sacramento River, California. Restoration Ecology 13, 391402.Google Scholar
Molle, F. 2018. Irrigation policies in Egypt since the construction of the High Aswan Dam. G-EAU Working Paper/Rapport de Recherche No.9. Montpellier, France. www.g-eau.net/ (accessed April 2020).Google Scholar
Montgomery, D. R. 2007. Dirt: The Erosion of Civilizations. University of California Press.Google Scholar
Montgomery, D. R., Collins, B. D., Buffington, J. M., and Abbe, T. B. 2003. Geomorphic effects of wood in rivers. In American Fisheries Society Symposium, American Fisheries Society, Bethesda, MD, pp. 127.Google Scholar
Morgan, J. P. 1970. Depositional processes and products in the deltaic environment. In Morgan, J. P. (ed.), Deltaic Sedimentation: Modern and Ancient. Society of Economic Paleontologists and Mineralogists, Special Publication 15, pp. 3147.Google Scholar
Morris, G. L. 2020. Classification of management alternatives to combat reservoir sedimentation. Water 12, 861. doi: 10.3390/w12030861.Google Scholar
Morris, G. L., and Fan, J. 1998. Reservoir Sedimentation Handbook: Design and Management of Dams, Reservoirs, and Watersheds for Sustainable Use. McGraw-Hill, New York.Google Scholar
Morris, M., Dyer, M., and Smith, P. 2007. Management of flood embankments: a good practice review. Department for Environment, Food and Rural Affairs, Environment Agency, Research and Development Technical Report FD2411/TR1.Google Scholar
Morton, L. W., and Olson, K. R. 2015. Sinkholes and sand boils during 2011 record flooding in Cairo, Illinois. Journal of Soil and Water Conservation 70, 49A54A.Google Scholar
Morton, R. A., and Bernier, J. C. 2010. Recent subsidence-rate reductions in the Mississippi Delta and their geological implications. Journal of Coastal Research 26, 555561.Google Scholar
Morton, R. A., Bernier, J. C., and Barras, J. A. 2006. Evidence of regional subsidence and associated interior wetland loss induced by hydrocarbon production, Gulf Coast region, USA. Environmental Geology 50, 261. doi: 10.1007/s00254-006-0207-3.Google Scholar
Morton, R. A., Bernier, J. C., Barras, J. A., and Ferina, N. F. 2016. Rapid subsidence and historical wetland loss in the Mississippi Delta Plain: likely causes and future implications. U.S. Geological Survey Open-File Report 2005–1216.Google Scholar
Mossa, J. 1996. Sediment dynamics in the lowermost Mississippi River. Engineering Geology 45, 457479.Google Scholar
Mossa, J. 2013. Historical changes of a major juncture: lower Old River, Louisiana. Physical Geography 34, 315334.Google Scholar
Mossa, J. 2015a. Geomorphic perspectives of managing, modifying, and restoring a river with prolonged flooding: Kissimmee River, Florida, USA. In Hudson, P. F., and Middelkoop, H. (eds.), Geomorphic Approaches to Integrated Floodplain Management of Lowland Fluvial Systems in North America and Europe. Springer-Verlag, New York, pp. 143169.Google Scholar
Mossa, J. 2015b. The changing geomorphology of the Atchafalaya River, Louisiana: a historical perspective. Geomorphology 252, 112127. doi: 10.1016/j.geomorph.2015.08.018.Google Scholar
Mossa, J., and Marks, S. R. 2011. Pit avulsions and planform change on a mined river floodplain: Tangipahoa River, Louisiana. Physical Geography 32, 512532.Google Scholar
Mossa, J., and McLean, M. 1997. Channel planform and land cover changes on a mined river floodplain Amite River, Louisiana, USA. Applied Geography 17, 4354.Google Scholar
Mossa, J., Hudson, P. F., Wilder, B., and Lower, J. 1993. Sediment supply from large rivers entering the Northern Gulf of Mexico. Report submitted to the US Army Waterways Experiment Station in fulfillment of Contract No DACA39-M-4918, U.S. Army Corps of Engineers, Vicksburg, MS.Google Scholar
Mossa, J., Chen, Y. H., Walls, S. P., Kondolf, G. M., and Wu, C.-Y. 2017. Anthropogenic landforms and sediments from dredging and disposing sand along the Apalachicola River and its floodplain. Geomorphology 294, 119134.Google Scholar
Mosselman, E. 2001. Morphological development of side channels. T2401, IRMA-SPONGE and Delft Cluster, DCFR project report 9, Delft, NL.Google Scholar
Mosselman, E., Shishikura, T., and Klaassen, G. J. 2000. Effect of bank stabilization on bend scour in anabranches of braided rivers. Physics and Chemistry of the Earth, Part B: Hydrology, Oceans and Atmosphere 25, 699704.Google Scholar
Mosselman, E., Kerssens, P., Van der Knaap, F., Schwanenberg, D., and Sloff, K. 2004. Sustainable river fairway maintenance and improvement: literature survey. Prepared for Rijkswaterstaat Directie Oost-Nederland. Q3757, WL Delft Hydraulics, December 2004.Google Scholar
Mount, J. F. 1995. California Rivers and Streams: The Conflict between Fluvial Processes and Land Use. University of California Press, Berkeley and Los Angleles.Google Scholar
Mulligan, M., van Soesbergen, A., and Sáenz, L. 2020. GOODD: a global dataset of more than 38,000 georeferenced dams. Scientific Data 7, 31. doi: 10.1038/s41597–020–0362-5.Google Scholar
Muñoz, A. A., González-Reyes, A., Lara, A., et al. 2016. Streamflow variability in the Chilean Temperate-Mediterranean climate transition (35°S–42°S) during the last 400 years inferred from tree-ring records. Climate Dynamics 47, 40514066. doi: 10.1007/s00382–016–3068-9.Google Scholar
Muñoz, S. E., and Dee, S. G. 2017. El Niño increases the risk of lower Mississippi River flooding. Scientific Reports 7, 1772. doi: 10.1038/s41598–017–01919-6.Google Scholar
Muñoz, S. E., Giosan, L., Therrell, M. D., et al. 2018. Climatic control of Mississippi River flood hazard amplified by river engineering. Nature 556, 9598.Google Scholar
Mussetter, R., and Trabant, S. 2005. Analysis of potential dam removal/retrofit impacts to habitat, flooding and channel stability in the Carmel Valley, California. In Watershed Management Conference 2005, July 19–22, 2005, Williamsburg, VA, pp. 1–11.Google Scholar
Mur, L. R., Skulberg, O. M., and Utkilen, H. 1999. Cyanobacteria in the environment, Ch. 2. In Chorus, I., and Bartram, J. (eds.), Toxic Cyanobacteria in Water: A Guide to Their Public Health Consequences, Monitoring and Management. World Health Organization, London. ISBN 0-419-23930-8.Google Scholar
Nagy, J., Kiss, T., Fehérváry, I., and Vaszkó, C. 2018. Changes in floodplain vegetation density and the impact of invasive amorpha fruiticosa on flood conveyance. Journal of Environmental Geography 11) 312. DOI: 10.2478/jengeo-2018-0008Google Scholar
Naik, P. K., and Jay, D. A. 2011. Distinguishing human and climate influences on the Columbia River: changes in mean flow and sediment transport. Journal of Hydrology 404, 259277.Google Scholar
Nair, S., Wen, W. K., and Ling, C. M. 2014. Bangkok flood risk management: application of foresight methodology for scenario and policy development. Journal of Futures Studies 19, 87112.Google Scholar
Nakamura, T. K., Singer, M. B., and Gabet, E. J. 2018. Remains of the 19th century: deep storage of contaminated hydraulic mining sediment along the Lower Yuba River, California. Elementa: Science of the Anthropocene 6, 70. doi: 10.1525/elementa.333.Google Scholar
Nanson, G. C. 2013. Anabranching and anastomosing rivers. In Shroder, J. F. (ed.), Treatise on Geomorphology, vol. 9. Academic Press, San Diego, CA, pp. 330345.Google Scholar
Nanson, G. C., and Croke, J. G. 1992. A genetic classification of floodplains. Geomorphology 4, 459486.Google Scholar
Nanson, G. C., and Knighton, A. D. 1996. Anabranching rivers: their cause, character and classification. Earth Surface Processes and Landforms 21, 217239.Google Scholar
National Committee on Levee Safety (NCLS). 2009. Recommendations for a National Levee Safety Program: a report to congress from the National Committee on Levee Safety.Google Scholar
National Hurricane Center. 2018. Hurricane Harvey (AL092017) 17 August – 1 September, 2017. Compiled by E. S. Blake and D. A. Zelinsky, May 9, 2018. www.nhc.noaa.gov/data/tcr/AL092017_Harvey.pdf (accessed February 2020).Google Scholar
National Levee Safety Act (NLSA). 2007. 121 Stat. 1288, Public Law 110–114-Nov. 8, 2007, 110th Congress.Google Scholar
National Oceanic Atmospheric Administration (NOAA). 2018. https://climate.nasa.gov/vital-signs/sea-level/ (accessed November 12, 2018).Google Scholar
National Oceanic Atmospheric Administration (NOAA). 2019. National Centers for Environmental Information (NCEI) U.S. Billion-Dollar Weather and Climate Disasters, Table of Events. www.ncdc.noaa.gov/billions/.Google Scholar
National Register of Dams in India (NRDI). 2016. Government of India. Archived from the original (PDF) September 20, 2016.Google Scholar
National Research Council (NRC). 1992. Restoration of Aquatic Ecosystems. National Academy Press, Washington, DC.Google Scholar
National Research Council (NRC). 2001. American Hazardscapes: The Regionalization of Hazards and Disasters. The National Academies Press, Washington, DC.Google Scholar
National Research Council (NRC). 2002. Riparian Areas: Functions and Strategies for Management. The National Academies Press, Washington, DC.Google Scholar
National Research Council (NRC). 2005. Endangered and Threatened Species of the Platte River. The National Academies Press, Washington, DC.Google Scholar
National Research Council (NRC). 2011. Sediment management alternatives and opportunities. Missouri River Planning: Recognizing and Incorporating Sediment Management.Google Scholar
National Research Council (NRC). 2013. Levees and the National Flood Insurance Program: Improving Policies and Practices. Water Science and Technology Board, Division on Earth and Life Studies, National Academies Press, Washington, DC.Google Scholar
Natural Water Retention Measures (NWRM). 2018. Reconnection of oxbow lakes and similar features. European Union Directorate General Environment. http://nwrm.eu/measure/reconnection-oxbow-lakes-and-similar-features (accessed August 2, 2018).Google Scholar
Neave, M., Rayburg, S., and Swan, A. 2009. River channel change following dam removal in an ephemeral stream. Australian Geographer 40, 235246.Google Scholar
Nègre, F. 2019. The European Union and forests. European Union Fact Sheet. www.europarl.europa.eu/factsheets/en/sheet/105/the-european-union-and-forests.Google Scholar
Nelson, S. A., and Leclair, S. F. 2006. Katrina’s unique splay deposits in a New Orleans neighborhood. GSA Today 16(9). doi: 10.1130/GSAT01609A.1.Google Scholar
New South Wales (NSW). 2014. An overview of floodplain management plans under the Water Management Act 2000. NSW Department of Industry – Water.Google Scholar
Newell, R. E., Newell, N. E., Zhu, Y., and Scott, C. 1992. Tropospheric rivers? – a pilot study. Geophysical Research Letters 19, 24012404. doi: 10.1029/92GL02916.Google Scholar
Newson, M. D., and Newson, C. L. 2000. Geomorphology, ecology and river channel habitat: mesoscale approaches to basin-scale challenges. Progress in Physical Geography 24, 195217.Google Scholar
Nienhuis, J. H. 2019. Wave-dominated river deltas. www.coastalwiki.org/wiki/Wave-dominated_river_deltas (accessed June 30, 2019).Google Scholar
Nienhuis, J. H., Ashton, A. D., and Giosan, L. 2015. What makes a delta wave-dominated?, Geology 43(6), 511514. doi: 10.1130/G36518.1.Google Scholar
Nienhuis, J. H., Hoitink, A. J. F., and Törnqvist, T. E. 2018a. Future change to tide-influenced deltas. Geophysical Research Letters 45, 34993507. doi: 10.1029/2018GL077638.Google Scholar
Nienhuis, J. H., Törnqvist, T. E., and Esposito, C. R. 2018b. Crevasse splays versus avulsions: a recipe for land building with levee breaches. Geophysical Research Letters 45, 40584067.Google Scholar
Nienhuis, J. H., Ashton, A. D., Edmonds, D. A., et al. 2020. Global-scale human impact on delta morphology has led to net land area gain. Nature 577, 514518.Google Scholar
Nienhuis, P. H. 2008. Environmental History of the Rhine-Meuse Delta: An Ecological Story on Evolving Human-Environmental Relations Coping with Climate Change and Sea-Level Rise. Springer, Dordrecht.Google Scholar
Nienhuis, P. H., and Leuven, R. S. E. W. 2001. River restoration and flood protection: controversy or synergism? Hydrobiologia 444, 8599.Google Scholar
Nillesen, A. L., and Kok, M. 2015. An integrated approach to flood risk management and spatial quality for a Netherlands’ river polder area. Mitigation and Adaptive Strategies for Global Change 20, 949966. doi: 10.1007/s11027-015-9675-7.Google Scholar
Nislow, K. H., Magilligan, F. J., Fassnacht, H., Bechtel, D., and Ruesink, A. 2002. Effects of hydrologic alteration on flood regime of natural floodplain communities in the Upper Connecticut River. Journal of the American Water Resources Association 38, 15331548.Google Scholar
Nittrouer, J. A., Allison, M. A., and Campanella, R. 2008. Bedform transport rates for the lowermost Mississippi River. Journal of Geophysical Research 113, F03004, doi: 10.1029/2007JF000795.Google Scholar
Nittrouer, C. A., Kuehl, S. A., Demaster, D. J., and Kowsmann, R. O. 1986. The deltaic nature of Amazon shelf sedimentation. Geological Society of America Bulletin 97, 444458.Google Scholar
Nixon, S. W. 2003. Replacing the Nile: are anthropogenic nutrients providing the fertility once brought to the Mediterranean by a great river? AMBIO 32, 3039.Google Scholar
Noda, K., Hamada, J., Kimura, M., and Oki, K. 2018. Debates over dam removal in Japan. Water and Environment Journal 32, 446452.Google Scholar
Notebaert, B., Verstraeten, G., Vandenberghe, D., Marinova, E., Poesen, J., and Govers, G. 2011. Changing hillslope and fluvial Holocene sediment dynamics in a Belgian loess catchment. Journal of Quaternary Science 26, 4458.Google Scholar
NSW Department of Primary Industries (DPS). 2007. Threat Abatement Plan: Removal of Large Woody Debris from NSW Rivers and Streams. Fisheries Conservation and Aquaculture. State of New South Wales, Australia.Google Scholar
Nyqvist, D., Nilsson, P. A., Alenäs, I., et al. 2017. Upstream and downstream passage of migrating adult Atlantic salmon: remedial measures improve passage performance at a hydropower dam. Ecological Engineering 102, 331343.Google Scholar
Ockerson, J. A. 1883. Review of Surveys and Gaugings of Cubitt’s Gap, Made in 1868, 1875, and 1576, 48 Cong., 1, session, H7 Docs., pt 7, 2 Sept. War, Vol. pt. 3, Appendix 0 of Appendix SS (Mississippi River Commission): 2302–2304.Google Scholar
Oczkowski, A. J., Nixon, S. W., Granger, S. L., El-Sayed, A.-F. M., and McKinney, R. A. 2009. Anthropogenic enhancement of Egypt’s Mediterranean fishery. Proceedings National Academy of Sciences 106, 13641367. doi: 10.1073/pnas.0812568106.Google Scholar
Okada, M., Iizumi, T., Sakamoto, T., et al. 2018. Varying benefits of irrigation expansion for crop production under a changing climate and competitive water use among crops. Earth’s Future 6, 12071220.Google Scholar
Olea, R. A., and Coleman, J. L. Jr. 2014. A synoptic examination of causes of land loss in southern Louisiana as related to the exploitation of subsurface geologic resources. Journal of Coastal Research 30, 10251044. doi: 10.2112/JCOASTRES-D-13-00046.1.Google Scholar
Ollero, A., Ibisate, A., Granado, D., and de Asua, R. R. 2015. Channel responses to global change and local impacts: perspectives and tools for floodplain management, Ebro River and Tributaries, NE Spain. In Hudson, P. F., and Middelkoop, H. (eds.), Geomorphic Approaches to Integrated Floodplain Management of Lowland Fluvial Systems in North America and Europe. Springer-Verlag, New York, pp. 2752.Google Scholar
Ondruch, J., Máčka, Z., Michalková, M. S., Putiška, R., Knot, M., Holík, P., Miřijovský, J., and Jenčo, M. 2018. Response of channel dynamics to recent meander neck cut-off in a lowland meandering river with artificial training history: the Morava River, Czech Republic, Hydrological Sciences Journal 63, 1236–1254. DOI: 10.1080/02626667.2018.1474218Google Scholar
Ontario Ministry of Natural Resources. 2011. Dam decommissioning and removal. Technical Bulletin. Ontario.ca/dams (accessed February 12, 2020).Google Scholar
Oppenheimer, M., Glavovic, B. C., Hinkel, J. et al. 2019: Sea level rise and implications for low-lying islands, coasts and communities. In Pörtner, H.-O., Roberts, D. C., Masson-Delmotte, V. et al. (eds.), IPCC Special Report on the Ocean and Cryosphere in a Changing Climate. Cambridge University Press, Cambridge, Chapter 4, pp. 321445.Google Scholar
Opperman, J. J., Royte, J., Banks, J., Day, L. R., and Apse, C. 2011. The Penobscot River, Maine, USA: a basin-scale approach to balancing power generation and ecosystem restoration. Ecology and Society 16, 3. doi: 10.5751/ES-04117–160307.Google Scholar
Ouchi, S. 1985. Response of alluvial rivers to slow active tectonic movement. Geological Society of American Bulletin 96, 504515.Google Scholar
Overeem, I., and Syvitski, J. P. M. 2009. Dynamics and Vulnerability of Delta Systems. LOICZ Reports and Studies No. 35. GKSS Research Center, Geesthacht.Google Scholar
Paira, A. R., and Drago, E. C. 2007. Origin, evolution, and types of floodplain water bodies. In Iriondo, M. H., Paggi, J. C., and Parma, M. J. (eds.), The Middle Paraná River: Limnology of a Subtropical Wetland. Springer-Verlag, Berlin and Heidelberg, pp. 5381.Google Scholar
Palanques, A., Guillen, J., Puig, P., and Grimalt, J. O. 2020. Effects of flushing flows on the transport of mercury-polluted particulate matter from the Flix Reservoir to the Ebro Estuary. Journal of Environmental Management 260, 110028. doi: 10.1016/j.jenvman.2019.110028.Google Scholar
Paltan, H., Waliser, D., Lim, W. H., et al. 2017. Global floods and water availability driven by atmospheric rivers. Geophysical Research Letters 44, 1038710395. doi: 10.1002/2017GL074882.Google Scholar
Panagos, P., Imeson, A., Meusburger, K., Borrelli, P., Poesen, J., and Alewell, C. 2016. Soil conservation in Europe: wish or reality? Land Degradation and Development 41, 15471551.Google Scholar
Pander, J., Mueller, M., and Geist, J. 2015. Succession of fish diversity after reconnecting a large floodplain to the upper Danube River. Ecological Engineering 75, 4150.Google Scholar
Paneque-Gálvez, J., Mas, J.-F., Guèze, M., et al. 2013. Land tenure and forest cover change. The case of southwestern Beni, Bolivian Amazon, 1986–2009. Applied Geography 43, 113126.Google Scholar
Panin, N. 1999. Global changes, sea level rise and the Danube Delta: risks and responses. In Proceedings of International Workshop on Modern and Ancient Sedimentary Environments and Processes, Moeciu, Romania, October 15–18, 1998, pp. 19–29.Google Scholar
Panin, N. 2015. The Danube Delta. Geomorphology and Holocene evolution: a synthesis. Geomorphologie: Relief, Proessus, Environment, 9 (Octobre-decembre), 247–262.Google Scholar
Park, E., and Latrubesse, E. M. 2015. Surface water types and sediment distribution patterns at the confluence of mega rivers: the Solimões-Amazon and Negro rivers junction. Water Resources Research 51, 61976213. doi: 10.1002/2014WR016757.Google Scholar
Park, E., and Latrubesse, E. M. 2017. High-resolution mapping of flood routing patterns and hydrological connectivity in the middle Amazon River floodplain. Remote Sensing of Environment 198, 321332.Google Scholar
Park, E., and Latrubesse, E. M. 2019. A geomorphological assessment of washload sediment fluxes and floodplain sediment sinks along the lower Amazon River. Geology 47, 403406.Google Scholar
Parker, C., Simon, A., and Thorne, C. 2008. The effects of variability in bank material properties on riverbank stability: Goodwin Creek, Mississippi. Geomorphology 101, 533543.Google Scholar
Patterson, L. A., Doyle, M. W., and Kuzma, S. 2018. Creating data as a service for U.S. Army Corps of Engineers Reservoirs. NI R 18–01. Duke University, Durham, NC. http://nicholasinstitute.duke.edu/publicationsGoogle Scholar
Pavelsky, T. M., and Smith, L. C. 2004. Spatial and temporal patterns in Arctic river ice breakup observed with MODIS and AVHRR time series. Remote Sensing of Environment 93, 328338.Google Scholar
Pedroli, B., de Blust, G., van Looy, K., and van Rooij, S. 2002. Setting targets in strategies for river restoration. Landscape Ecology 17, 518.Google Scholar
Pegg, M., Pierce, C., and Roy, A. 2003. Hydrological alteration along the Missouri River Basin: a time series approach. Aquatic Sciences 65, 6372.Google Scholar
Peng, C., Zhang, Y., Huang, S. et al. 2019. Sediment phosphorus release in response to flood event across different land covers in a restored wetland. Environmental Science and Pollution Research 26, 91139122. doi: 10.1007/s11356-019-04398-6.Google Scholar
Penland, S., Boyd, R., and Suter, J. R. 1988. Transgressive depositional systems of the Mississippi delta plain: a model for barrier shoreline and shelf sand development. Journal of Sedimentary Petrology 58, 932949.Google Scholar
Pepper, A. T., and RickardPepper, A. C. E. Rickard, C. 2008. Works in the river channel. In Fluvial Design Guide. U.K. Environment Agency. http://evidence.environment-agency.gov.uk/FCERM/en/FluvialDesignGuide.aspx (accessed November 17, 2017).Google Scholar
Pescaroli, G., and Nones, M. 2016. Cascading events, technology and the floods directive: future challenges. FLOODrisk 2016 – 3rd European Conference on Flood Risk Management.Google Scholar
Peteuil, C., Fruchart, F., Abadie, F., Reynaud, S., Camenen, B., and Guertault, L. 2013. Sustainable management of sediment fluxes in reservoir by environmental friendly flushing: the case study of the Genissiat dam on the upper Rhône River (France). In Conference: 12th International Conference on River Sedimentation, Kyoto, Japan.Google Scholar
Petts, G. E. 1986. Water quality characteristics of regulated rivers. Progress in Physical Geography 10, 492516.Google Scholar
Peyronnin, N. S., Caffey, R. H., Cowan, J. H. et al. 2017. Optimizing sediment diversion operations: working group recommendations for integrating complex ecological and social landscape interactions. Water 9, 368. doi: 10.3390/w9060368.Google Scholar
Phien-wej, N., Gao, P. H., and Nutalaya, P. 2006. Land subsidence in Bangkok, Thailand. Engineering Geology 82, 187201. doi: 10.1016/j.Enggeo.2005.10.004.Google Scholar
Phillips, J. D. 2003. Toledo Bend reservoir and geomorphic response in the lower Sabine River. River Research and Applications 19, 137159.Google Scholar
Phillips, J. D. 2011. Universal and local controls of avulsions in southeast Texas Rivers. Geomorphology 130, 1728.Google Scholar
Phillips, J. D. 2012. Log-jams and avulsions in the San Antonio River Delta, Texas. Earth Surface Processes and Landforms 37, 946950.Google Scholar
Phillips, J. D. 2013. Hydrological connectivity of abandoned channel water bodies on a coastal plain river. River Research and Applications 29, 149160.Google Scholar
Phillips, J. D., and Park, L. 2009. Forest blowdown impacts of Hurricane Rita on fluvial systems. Earth Surface Processes and Landforms 34, 10691081Google Scholar
Phillips, J. D., and Slattery, M. C. 2007. Downstream trends in discharge, slope, and stream power in a coastal plain river. Journal of Hydrology 334, 290303.Google Scholar
Phillips, J. D., Slattery, M. C., and Musselman, Z. A. 2004. Dam-to-delta sediment inputs and storage in the lower trinity river, Texas. Geomorphology 62, 1734.Google Scholar
Phillips, J. D., Slattery, M. C., and Musselman, Z. A. 2005. Channel adjustments of the Lower Trinity River, Texas, downstream of Livingston Dam. Earth Surface Processes and Landforms 30, 14191439.Google Scholar
Piégay, H., and Gurnell, A. M. 1997. Large woody debris and river geomorphological pattern: examples from S. E. France and S. England. Geomorphology 19, 99116.Google Scholar
Piégay, H., and Marston, R. A. 1998. Distribution of large woody debris along the outer bend of meanders in the Ain River, France. Physical Geography 19, 318340.Google Scholar
Piégay, H., Thévenet, A., and Citterio, A. 1999. Input, storage and distribution of large woody debris along a mountain river continuum. The Dróme River, France. Catena 35, 1939.Google Scholar
Piman, T., and Manish, S. 2017. Case study on sediment in the Mekong River basin: Current state and future trends. Project Report 2017–03. Stockholm Environment Institute. www.sei.org/publications/sediment-mekong-river/ (accessed April 3, 2020).Google Scholar
Pinter, N. 2005. One step forward, two steps back on U.S. floodplains. Science 308(5719), 207208.Google Scholar
Pinter, N. 2015a. Discussion of “Mississippi River streamflow measurement techniques at St. Louis, Missouri” by C. C. Watson, R. R. Holmes Jr., and D. S. Biedenharn. Journal of Hydraulic Engineering 141, 10621070. doi: 10.1061/(ASCE)HY.1943-7900.0001020.Google Scholar
Pinter, N. 2015b. Discussion of “Analysis of the Impacts of Dikes on Flood Stages in the Middle Mississippi River” by Watson, Chester C., Biedenharn, David S., and Thorne, Colin R.. Journal of Hydraulic Engineering 141. doi: 10.1061/(ASCE)HY.1943-7900.0001054.Google Scholar
Pinter, N., Ickes, B. S., Wlosinski, J. H., and van der Ploeg, R. R. 2006a. Trends in flood stages: contrasting results from the Mississippi and Rhine River systems. Journal of Hydrology 331, 554566.Google Scholar
Pinter, N., van der Ploeg, R. R., Schweigert, P., and Hoefer, G. 2006b. Flood magnification on the River Rhine. Hydrological Processes 20, 147164.Google Scholar
Pinter, N., Jemberie, A. A., Remo, J. W. F., Heine, R. A., and Ickes, B. S. 2008. Flood trends and river engineering on the Mississippi River system, Geophysical Research Letters 35, L23404. doi: 10.1029/2008GL035987.Google Scholar
Pinter, N., Jemberie, A. A., Remo, J. W. F., Heine, R. A., and Ickes, B. S. 2010. Cumulative impacts of river engineering, Mississippi and Lower Missouri rivers. River Research and Applications 26, 546571. doi: 10.1002/rra.1269.Google Scholar
Pinter, N., Huthoff, F., Dierauer, J., Remo, J. W. F., and Damptz, A. 2016. Modeling residual flood risk behind levees, Upper Mississippi River, USA. Environmental Science and Policy 58, 131. doi: 10.1016/j.envsci.2016.01.003.Google Scholar
Pitlick, J. 1997. A regional perspective of the hydrology of the 1993 Mississippi River Basin floods. Annals of the Association of American Geographers 87, 325351.Google Scholar
Pizzuto, J. E. 1987. Sediment diffusion during overbank events. Sedimentology 34, 301317.Google Scholar
Poeppl, R. E., Keestra, S. D., and Hein, T. 2015. The geomorphic legacy of small dams – an Austrian study. Anthropocene 10, 4355.Google Scholar
Poesen, J., and Govers, G. 1990. Gully erosion in the Loam Belt of Belgium: typology and control measures. In Boardman, J., Foster, I. D. L., and Dearing, J. A. (eds.), Soil Erosion on Agricultural Land. John Wiley & Sons, Chichester, pp. 515530.Google Scholar
Poesen, J., Nachtergaele, J., Verstraeten, G., and Valentin, C. 2003. Gully erosion and environmental change: importance and research needs. Catena 50(2), 91133.Google Scholar
Pohl, M. 2002. Bringing down our dams: trends in American dam removal rationales. Journal of the American Water Resources Association 38, 15111519.Google Scholar
Pokhrel, Y., Burbano, M., Roush, J., Kang, H., Sridhar, V., and Hyndman, D. W. 2018. A Review of the integrated effects of changing climate, land use, and dams on Mekong River hydrology. Water 10, 266.Google Scholar
Polyakov, I. V., and Johnson, M. A. 2000. Arctic decadal and interdecadal variability. Geophysical Research Letters 27, 40974100. doi: 10.1029/2000GL011909.Google Scholar
Pongruktham, O., and Ochs, C. 2015. The rise and fall of the lower Mississippi: effects of hydrologic connection on floodplain backwaters. Hydrobiologia 742, 169183.Google Scholar
Poole, G. C., Stanford, J. A., Frissell, C. A., and Running, S. W. 2002. Three-dimensional mapping of geomorphic controls on flood-plain hydrology and connectivity from aerial photos. Geomorphology 48, 329347.Google Scholar
Potter, P. E. 1978. Significance and origin of big rivers. Journal of Geology 86, 1333.Google Scholar
Powell, J. W. 1895. The Exploration of the Colorado River and Its Canyons. Dover Publications, New York.Google Scholar
Power, M. E., Dietrich, W. E., and Finlay, J. C. 1996. Dams and downstream aquatic biodiversity: potential food web consequences of hydrologic and geomorphic change. Environmental Management 20, 887895.Google Scholar
Powers, J. 2018. Mississippi River ship channel dredging and wetlands creation – an environmental success story. International Dredge Review November 5, 2018. www.dredgemag.com/2018/11/05/mississippi-river-ship-channel-dredging-and-wetlands-creation-an-environmental-success-story/ (accessed June 2020).Google Scholar
Prieto, M. R. 2007. ENSO signals in South America: rains and floods in the Paraná River region during colonial times. Climate Change 83, 3954.Google Scholar
Provansal, M. 2004. The Rhone delta (France). EUROSION case study. In Living with coastal erosion in Europe: Sediment and Space for Sustainability. A guide to coastal erosion management practices in Europe. www.eurosion.org/shoreline/introduction.htmlGoogle Scholar
Provencher, B., Sarakinos, H., and Meyer, T. 2008. Does small dam removal affect local property values? An empirical analysis. Contemporary Economic Policy 26, 187197.Google Scholar
Przedwojski, B. 1995. Bed topography and local scour in rivers with banks protected by groynes, Journal of Hydraulic Research 33(2), 257273. doi: 10.1080/00221689509498674.Google Scholar
Przedowjski, B., Blazejewski, R., and Pilarczyk, K. W. 1995. River Training Techniques: Fundamentals, Design and Application. A.A. Balkema, Rotterdam.Google Scholar
Querner, E. P., Jansen, P. C., van den Akker, J. J. H., and Kwakernaak, C. 2012. Analyzing water level strategies to reduce soil subsidence in Dutch peat meadows. Journal of Hydrology 446–447, 5969.Google Scholar
Quick, I., König, F., Baulig, Y., Schriever, S., and Vollmer, S. 2020. Evaluation of depth erosion as a major issue along regulated rivers using the classification tool Valmorph for the case study of the Lower Rhine. International Journal of River Basin Management 18, 191206. doi: 10.1080/15715124.2019.1672699.Google Scholar
Qureshi, A. S. 2011. Water management in the Indus Basin in Pakistan: challenges and opportunities. Mountain Research and Development 31, 252260.Google Scholar
Rader, R. B., Voelz, N. J., and Ward, J. V. 2008. Post‐flood recovery of a macroinvertebrate community in a regulated river: resilience of an anthropogenically altered ecosystem. Restoration Ecology 16, 2433.Google Scholar
Ralph, F. M., Coleman, T., Neiman, P. J., Zamora, R. J., and Dettinger, M. D. 2013. Observed impacts of duration and seasonality of atmospheric-river landfalls on soil moisture and runoff in coastal northern California. Journal of the American Meteorological Society, April 2013, 443459. doi: 10.1175/JHM-D-12–076.1.Google Scholar
Randle, T. J., and Bountry, J. 2017. Dam Removal Analysis Guidelines for Sediment. U.S. Bureau of Reclamation, Denver, CO.Google Scholar
Randle, T., Morris, G., Whelan, M., et al. 2019. Reservoir sediment management: building a legacy of sustainable water storage reservoirs. In National Reservoir Sedimentation and Sustainability Team White Paper. SEDHYD, Denver, CO, p. 57. http://www.sedhyd.org/reservoir-sedimentation/ (accessed May 26, 2020).Google Scholar
Raffles, H., and WinklerPrins, A. M. G. A. 2003. Further reflections on Amazonian environmental history: transformations of rivers and streams. Latin American Research Review 38, 165187.Google Scholar
Räsänen, T. A., and Kummu, M. 2013. Spatiotemporal influences of ENSO on precipitation and flood pulse in the Mekong River Basin. Journal of Hydrology 476, 154168.Google Scholar
Räsänen, T. A., Someth, P., Lauri, H., Koponen, J., Sarkkula, J., Kummu, M. 2017. Observed river discharge changes due to hydropower operations in the Upper Mekong Basin. Journal of Hydrology 545, 28–41.Google Scholar
Rashid, M. M. 1995. Some additional information on limnology and fisheries of Lakes Nasser (Egypt) and Nubia (Sudan). In Crul, R. C. M., and Roest, F. C. (eds.), Current Status of the Lake Nasser–Nubia Fisheries and Fish Stocks of the Four Largest African Reservoirs: Kainji, Kariba, Nasser/Nubia, and Volta. CIFA Technical Paper 30. FAO, Rome, Italy, pp. 81109.Google Scholar
Raslan, Y., and Salama, R. 2015. Development of Nile River islands between Old Aswan Dam and new Esna barrages. Water Science 29, 7792. doi: 10.1016/j.wsj.2015.03.003.Google Scholar
Raza, A., Naeem, K., Qadeer, A., et al. 2019. Water, energy and food nexus of Indus Water Treaty: water governance. Water-Energy Nexus 2, 1024.Google Scholar
Reckendorfer, W., Schmalfuss, R., Baumgartner, C., et al. 2005. The Integrated River Engineering Project for the free-flowing Danube in the Austrian Alluvial Zone National Park: contradictory goals and mutual solutions. Archives of Hydrobiology 155, 613630.Google Scholar
Reckendorfer, W., Funk, A., Gschöpf, C., Hein, T., and Schiemer, F. 2013. Aquatic ecosystem functions of an isolated floodplain and their implications for flood retention and management. Journal of Applied Ecology 50, 119128. doi: 10.1111/1365-2664.12029.Google Scholar
Redfern, S. K., Azzu, N., and Binarima, J. S. 2012. Rice in Southeast Asia: facing risks and vulnerabilities to respond to climate change. In Proceedings of the, Building resilience for adaptation to climate change in the agriculture sector. Proceedings of a Joint FAO/OECD.Google Scholar
Reisner, M. 1993 [1986]. Cadillac Desert: The American West and Its Disappearing Water. Penguin, New York.Google Scholar
Remo, J. W. F., Ickes, B. S., Ryherd, J. K., Guida, R. J., and Therrell, M. D. 2018. Assessing the impacts of dams and levees on the hydrologic record of the Middle and lower Mississippi River, USA. Geomorphology 313, 88100. doi: 10.1016/j.geomorph.2018.01.004.Google Scholar
Renwick, W. 1992. Equilibrium, disequilibrium, and nonequilibrium landforms in the landscape. Geomorphology 5, 265276.Google Scholar
Renwick, W. H., and Andereck, Z. D. 2006. Reservoir sedimentation trends in Ohio, USA: sediment delivery and response to land-use change. Sediment dynamics and the hydromorphology of fluvial systems (Proceedings of a symposium held in Dundee, UK, July 2006). International Association of Hydrological Sciences, Publication 306.Google Scholar
Report of the Secretary of War. 1855. Col. C. S. Fuller’s Survey of Red River. Map of Red River with its Bayous and Lakes in the vicinity of the Raft (1:36,000). 33d Congress, 2d session. House of Representatives. Ex. doc. no. 90; March 3, 1855, Washington, DC.Google Scholar
Reuss, M. 1998. Designing the Bayous: The Control of Water in the Atchafalaya Basin, 1800–1995. U.S. Army Corps of Engineers, Office of History, Alexandria, VA.Google Scholar
Rice, S. P., Church, M., Wooldridge, C. L., and Hickin, E. J. 2009. Morphology and evolution of bars in a wandering gravel-bed river; lower Fraser River, British Columbia, Canada. Sedimentology 56, 709736. doi: 10.1111/j.1365-3091.2008.00994.x.Google Scholar
Richards, D. R., Warren, P. H., Maltby, L., and Moggridge, H. L. 2017. Awareness of greater numbers of ecosystem services affects preferences for floodplain management. Ecosystem Services 24, 138146.Google Scholar
Richards, K. 1982. Rivers: Form and Process in Alluvial Channels. London: Methuen.Google Scholar
Richey, J. E., Nobre, C., and Deser, C. 1989. Amazon River discharge and climate variability: 1903 to 1985. Science 246(4926), 101103. doi: 10.1126/science.246.4926.101.Google Scholar
Riggsbee, J. A., Wetzel, R., and Doyle, M. W. 2012. Physical and plant community controls on nitrogen and phosphorus leaching from impounded riverine wetlands following dam removal. River Research and Applications 28, 14391450.Google Scholar
Rijkswaterstaat. 2009. Water Management in the Netherlands. Ministry of Infrastructure and the Environment, The Hague.Google Scholar
Rijkswaterstaat. 2016. Afwegingen bij plaatsing en beheer rivierhout : voor initiatiefnemers en uitvoerders bijlagen Projectteam Pilot Rivierhout (Guidelines for installation of wood in rivers). Rijkswaterstaat (RWS): RWS 12–2016.Google Scholar
Rijkswaterstaat. 2018a. Final and Partial Evaluations and Report for the Room for the River Project on the Rhine River and Zandmaas/Grensmaas: 2018. Ministerie van Infrastructuur en Waterstaat, Rijkswaterstaat, Ruimte voor de Rivier. http://publicaties.minienm.nl/documenten/eind-en-deelevaluaties-programma-s-ruimte-voor-de-rivier-en-zandmaas-grensmaas (accessed October 15, 2018)Google Scholar
Rijkswaterstaat. 2018b. Our Flood Defences: Working on Storm Safety. Rijkswaterstaat Ministry of Infrastructure and Environment, The Netherlands.Google Scholar
Rijnland. 2009. Flood Control in the Netherlands: A Strategy for Dike Reinforcement and Climate Adaptation. Hoogheemraadschap van Rijnland, Leiden.Google Scholar
Rinaldi, M., and Casagli, N. 1999. Stability of streambanks in partially saturated soils and effects of negative pore water pressures: the Sieve River. Geomorphology 26, 253277.Google Scholar
Rinaldi, M., and Simon, A. 1998. Bed-level adjustments of the Arno River, central Italy. Geomorphology 22, 5771.Google Scholar
Rincón Sanz, G., and Gortázar Rubial, J. 2016. An analysis of river fragmentation in the Spanish River Basins. Developed by Ecohidráulica, S.L. at the request of Centro Ibérico de Restauración Fluvial (CIREF) and Wetlands International European Association, EH-BIO-007–16.Google Scholar
Ritchie, A. C., Warrick, J. A., East, A. E., et al. 2018. Morphodynamic evolution following sediment release from the world’s largest dam removal. Scientific Reports 8, 13279.Google Scholar
Ritchie, H., and Roser, M. 2013. Land use. UN FAO Statistical Database Published online at OurWorldInData.org. https://ourworldindata.org/land-use (accessed May 5, 2020).Google Scholar
Riviere, G., and Drouard, M. 2015. Dynamics of the Northern Annular Mode at weekly time scales. Journal of the Atmospheric Sciences 72, 45694590.Google Scholar
Rivierenland Waterschap. 2017. A new beginning for an iconic polder: vision for 2050 for the Alblasserwaard water system (in Dutch: Een nieuw begin voor een iconische polder: Visie voor 2050 op het watersysteem in de Alblasserwaard), https://simcms.waterschaprivierenland.nl/_flysystem/media/visie-alblasserwaard-2050.pdf (accessed November, 2019).Google Scholar
Roberts, H. H. 1997. Dynamic changes of the Holocene Mississippi River delta plain: the Delta Cycle. Journal of Coastal Research 13, 605627.Google Scholar
Roberts, H. H. 1998. Delta switching: early responses to the Atchafalaya River diversion. Journal of Coastal Research 14, 882899.Google Scholar
Robinson, C. T., Uelinger, U., and Monaghan, M. T. 2003. Effects of a multi-year experimental flood regime on macroinvertebrates downstream of a reservoir. Aquatic Sciences 65, 210222.Google Scholar
Rodell, M., Famiglietti, J. S., Wiese, D. N. et al. 2018. Emerging trends in global freshwater availability. Nature 557(7707), 651659. doi: 10.1038/s41586–018–0123-1.Google Scholar
Rogers, J. D. 2008. Development of the New Orleans Flood Protection System prior to Hurricane Katrina. Journal of Geotechnical and Geoenvironmental Engineering 134, 602617. doi: 10.1061/(ASCE)1090-0241(2008)134:5(602).Google Scholar
Rogers, J. D., Boutwell, G. P., Schmitz, D. W., Karadeniz, D., Watkins, C. M., Athanasopoulos-Zekkos, A. G., and Cobos-Roa, D. 2008. Geologic conditions underlying the 2005 17th Street Canal levee failure in New Orleans. Journal of Geotechnical and Geoenvironmental Engineering 134, 583601.Google Scholar
Rojstaczer, S. A., Hamon, R. E., Deverel, S. J., and Massey, C. A. 1991. Evaluation of selected data to assess the causes of subsidence in the Sacramento-San Joaquin Delta. Open-File Report 91–193, 16. U.S. Geological Survey, California.Google Scholar
Rokaya, P., Budhathoki, S., and Lindenschmidt, K.-E. 2018. Trends in the timing and magnitude of ice-jam floods in Canada. Nature Scientific Reports 8, 5834. doi: 10.1038/s41598–018–24057z.Google Scholar
Ronchail, J., Labat, D., Callede, J., et al. 2005. Discharge variability within the Amazon basin. Regional Hydrological Impacts of Climatic Change – Hydroclimatological Variability, Proceedings of symposium S6 held during the Seventh IAHS Scientific Assembly at Foz do Iguaçu, Brazil, April 2005). IAHS Publ. 296, pp. 21–26.Google Scholar
Rose, K. A., Huang, H., Justic, D., and de Mutsert, K. 2014. Simulating fish movement responses to and potential salinity stress from large-scale river diversions. Marine and Coastal Fisheries 6, 4361, doi: 10.1080/19425120.2013.866999.Google Scholar
Rotterdam Climate Proof Initiative. 2013. Rotterdam climate change adaptation strategy. Rotterdam Office for Sustainability and Climate Change. www.rotterdamclimateinitiative.nl (accessed April 4, 2016).Google Scholar
Rovira, A., and Ibàñez, C. 2007. Sediment management options for the lower Ebro River and its delta. Journal of Soil and Sediments. doi: 10.1065/jss2007.08.244.Google Scholar
Rowland, J. C., Dietrich, W. E., Day, G., and Parker, G. 2010. Formation and maintenance of single‐thread tie channels entering floodplain lakes: observations from three diverse river systems. Journal of Geophysical Research: Earth Surface 114. doi: 10.1029/2008JF001073.Google Scholar
Rozo, M. G., Nogueira, A. C. R., and Castro, C. S. 2014. Remote sensing-based analysis of the planform changes in the Upper Amazon river over the period 1986–2006. Journal of South American Earth Sciences 51, 2844. doi: 10.1016/j.sames.2013.12.004.Google Scholar
Ru, H.-J., and Liu, X. 2013. River-lake migration of fishes in the Dongting Lake area of the Yangtze floodplain. Journal of Applied Ichthyology 29, 594601.Google Scholar
Ruban, G., Khodorevskaya, R., and Shatunovskii, M. 2019. Factors influencing the natural reproduction decline in the beluga (Huso huso, Linnaeus, 1758), Russian sturgeon (Acipenser gueldenstaedtii, Brandt and Ratzeburg, 1833), and stellate sturgeon (A. stellatus, Pallas, 1771) of the Volga–Caspian basin: a review. Journal of Applied Ichthyology 35, 387395.Google Scholar
Rutherford, J. S., Day, J. W., D’Elia, C. F., et al. 2018. Evaluating trade-offs of a large, infrequent sediment diversion for restoration of a forested wetland in the Mississippi delta. Estuarine, Coastal and Shelf Science. doi: 10.1016/j.ecss.2018.01.016.Google Scholar
Russell, R. J. 1936. Physiography of the lower Mississippi River Delta. In, Reports on the Geology of Plaquemines and St. Bernard Parishes. Louisiana Geological Survey, Bulletin 8, 3193.Google Scholar
Russell, R. J. 1939. Louisiana stream patterns. American Association of Petroleum Geologists Bulletin 23, 11991227.Google Scholar
Sacklin, H., and Ozaki, S. 1988. Environmental Assessment: Upper Dam Removal. Redwood National Park, Lost Man Creek, CA.Google Scholar
Saito, Y., Yang, Z., and Hori, K. 2001. The Huang He (Yellow River) and Chang Jiang (Yangtze River) deltas: a review of their characteristics, evolution and sediment discharge during the Holocene. Geomorphology 41, 219231.Google Scholar
Saji, N. H., Goswami, B. N., Vinayachandran, P. N., and Yamagata, T. 1999. A dipole mode in the tropical Indian Ocean. Nature 401, 360363.Google Scholar
Salmon, J. M., Friedl, M. A., Frolking, S., Wisser, D., and Douglas, E. M. 2015. Global rain-fed, irrigated, and paddy croplands: a new high resolution map derived from remote sensing, crop inventories and climate data. International Journal Applied Earth Observation Geoinformatics 38, 321334.Google Scholar
Sanchez-Arcilla, A., Jimenez, J. A., and Valdemoro, H. I. 1998. The Ebro Delta: morphodynamics and vulnerability. Journal of Coastal Research 14, 754772.Google Scholar
Sando, S. K., and Lambing, J. H. 2011. Estimated loads of suspended sediment and selected trace elements transported through the Clark Fork Basin, Montana, in selected periods before and after the breach of Milltown Dam (water years 1985–2009). U.S. Geological Survey Scientific Investigations Report 2011–5030, 64pp. at http://pubs.usgs.gov/sir/2011/5030 (accessed June 1, 2020).Google Scholar
Sando, S. K., and Vecchia, A. V. 2016. Water-quality trends and constituent-transport analysis for selected sampling sites in the Milltown Reservoir/Clark Fork River Superfund Site in the upper Clark Fork Basin, Montana, water years 1996–2015. U.S. Geological Survey Scientific Investigations Report 2016–5100.Google Scholar
Satyamurty, P., da Costa Priscila, C. P. W., Manzi, A. O., and Candido, L. A. 2013. A quick look at the 2012 record flood in the Amazon Basin. Geophysical Research Letters 40, 13961401. doi: 10.1002/grl.50245.Google Scholar
Saucier, R. T. 1963. Recent geomorphic history of the Pontchartrain Basin. Louisiana State University, Coastal Studies Series 9, 114pp.Google Scholar
Saucier, R. T. 1994. Geomorphology and Quaternary Geologic History of the Lower Mississippi Valley. U.S. Army Corps of Engineers, Waterways Experiment Station, Vicksburg, Mississippi.Google Scholar
Savelieva, N. I., Semiletov, I. P., Vasilevskaya, L. N., and Pugach, S. P. 2000. A climate shift in seasonal values of meteorological and hydrological parameters for Northeastern Asia. Progress in Oceanography 47, 279297.Google Scholar
Sayers, P., Galloway, G., Penning-Rosell, E. et al. 2015. Strategic flood management: ten “golden rules” to guide a sound approach. International Journal of River Basin Management 1, 137151. doi: 10.1080/15715124.2014.902378.Google Scholar
Scanlon, B. R., Jolly, I., Sophocleous, M., and Zhang, L. 2007. Global impacts of conversions from natural to agricultural ecosystems on water resources: quantity versus quality. Water Resources Research 43, W03437. doi: 10.1029/2006WR005486.Google Scholar
Schalk, G. K., and Jacobson, R. B. 1997. Scour, sedimentation, and sediment characteristics at six levee-break sites in Missouri from the 1993 Missouri River flood. U.S. Geological Survey Water Resources Investigations Report 97–4110.Google Scholar
Scheueklein, H. 1990. Removal of sediment deposits in reservoirs by means of flushing. Hydrology in Mountainous Regions II, Artificial Reservoirs; Water and Slopes (Proceedings of Lausanne Symposia). International Association of Hydrologic Sciences (IAHS) Publ. no. 194, 99–106.Google Scholar
Schiermeier, Q. 2018a. Europe is demolishing its dams to restore ecosystems. Nature 557, 290291.Google Scholar
Schiermeier, Q. 2018b. Dam removal restores rivers: huge European demolition projects offer hope for fragmented ecosystems. Nature 557(17 May), 290291.Google Scholar
Schmutz, S., and Moog, O. 2018. Dams: ecological impacts and management. In Schmutz, S., and Sendzimir, J. (eds.), Riverine Ecosystem Management. Aquatic Ecology Series 8 (open access). Springer, pp. 111127.Google Scholar
Schober, B., Hauer, C., and Habersack, H. 2015. A novel assessment of the role of Danube floodplains in flood hazard reduction (FEM method). Natural Hazards 75, S33S50. doi: 10.1007/s11069–013–0880-y.Google Scholar
Schramm, H. L., Cox, M. F., Tietjen, H. E., and Ezell, A. W. 2009. Nutrient dynamics in the lower Mississippi River floodplain: comparing present and historic hydrologic conditions. Wetlands 29, 476487.Google Scholar
Schulte, J. A., Najjar, R. G., and Li, M. 2016. The influence of climate modes on streamflow in the Mid-Atlantic region of the United States. Journal of Hydrology: Regional Studies 5, 8099.Google Scholar
Schumm, S. A. 1960. The shape of alluvial channels in relation to sediment type. U.S. Geological Survey Professional Paper 352-B, pp. 17–30.Google Scholar
Schumm, S. A. 1963. Sinuosity of alluvial rivers on the Great Plains. Geological Society of America Bulletin 74(9), 10891100.Google Scholar
Schumm, S. A. 1968. River adjustments to altered hydrologic regimen – Murrumbidgee River and paleochannels. U.S. Geological Survey Professional Paper 598, Australia.Google Scholar
Schumm, S. A. 1977. The Fluvial System. Wiley, New York.Google Scholar
Schumm, S. A. 1985. Patterns of alluvial rivers. Annual Review of Earth and Planetary Sciences 13, 527.Google Scholar
Schumm, S. A. 1991. To Interpret the Earth: Ten Ways to Be Wrong. Cambridge University Press.Google Scholar
Schumm, S. A. 2007. Rivers and humans – unintended consequences. In Gupta, A. (ed.), Large Rivers: Geomorphology and Management. John Wiley & Sons, Chichester, pp. 517533.Google Scholar
Schumm, S. A., and Khan, H. R. 1972. Experimental study of river patterns. Bulletin of the Geological Society of America 83, 17551770.Google Scholar
Schumm, S. A., and Spitz, W. J. 1996. Geological influences on the lower Mississippi River and its alluvial valley. Engineering Geology 45, 245261.Google Scholar
Schumm, S. A., and Winkley, B. R. 1994. The character of large alluvial rivers. In Schumm, S. A., and Winkley, B. R. (eds.), The Variability of Large Alluvial Rivers. American Society of Engineers Press, New York, pp. 19.Google Scholar
Schumm, S. A., Dumont, J. F., and Holbrook, J. M. 2000. Active Tectonics and Alluvial Rivers. Cambridge University Press, Cambridge.Google Scholar
Science 2021. Unleashing big muddy. 372 (6540), 334–337. DOI: 10.1126/science.372.6540.334Google Scholar
Scruton, P. C. 1960. Delta building and the deltaic sequence. In Shepard, F. P., Phleger, F. B., and Andel, T. H. (eds.), Recent Sediments, Northwest Gulf of Mexico: A Symposium. American Association of Petroleum Geologists, Tulsa, OK, pp. 82102.Google Scholar
Seager, R., Lis, N., Feldman, J., et al. 2018. Whither the 100th Meridian? The once and future physical and human geography of America’s Arid–Humid Divide. Part I: The story so far. Earth Interactions 22, 122. doi: 10.1175/EI-D-17-0011.1.Google Scholar
Secor, D. H., Arefjev, V., Nikolaev, A., and Sharov, A. 2000. Restoration of sturgeons: lessons from the Caspian Sea Sturgeon Ranching Programme. Fish and Fisheries 1, 215230. doi: 10.1111/j.1467–2979.2000.00021.x.Google Scholar
Seed, R. B., Bea, R. G., Athanasopoulos-Zekkos, A., et al. 2008. The New Orleans and Hurricane Katrina III: the 17th St. drainage canal. Journal of Geotechnical and Geoenvironmental Engineering 134, 740761. doi: 10.1061/(ASCE)1090–0241(2008)134:5(740).Google Scholar
Select Bipartisan Committee. 2006. A failure of initiative. Final Report of the Select Bipartisan Committee to Investigate the Preparation for and Response to Hurricane Katrina, 109th Congress, 2nd Session, pp. 109–377.Google Scholar
Shankman, D., and Smith, L. J. 2004. Stream channelization and swamp formation in the US Coastal Plain. Physical Geography 25, 2238.Google Scholar
Sharp, J. M. Jr. 1988. Alluvial aquifers along major rivers. In Back, W., Rosenshein, J. S., and Seaber, P. R (eds.), Hydrogeology, vol. O-2. Geological Society of America, The Geology of North America, Boulder, CO, pp. 273282.Google Scholar
Shaw, J. B., Mohrig, D., and Whitman, S. K. 2013. The morphology and evolution of channels on the Wax Lake Delta, Louisiana, USA. Journal of Geophysical Research: Earth Surface 118, 15621584. doi: 10.1002/jgrf.20123.Google Scholar
Shen, Z., Törnqvist, T. E., Mauz, B., Chamberlain, E. L., Nijhuis, A. G., and Sandoval, L. 2015. Episodic overbank deposition as a dominant mechanism of floodplain and delta-plain aggradation. Geology 43, 875878. doi: 10.1130/G36847.1.Google Scholar
Shi, H., and Shao, M. 2000. Soil and water loss from the Loess Plateau in China. Journal of Arid Environments 45, 920.Google Scholar
Shi, H. L., Hu, C. H., Deng, A. J., and Tian, Q. Q. 2017. Analyses on trends and reasons of runoff and sediment load of Yellow River stem. In Wieprecht, S., Haun, S., Weber, K., Noack, M., and Terheidenet, K. (eds.), River Sedimentation. Proceedings of the 13th International Symposium on River Sedimentation, Stuttgart, Germany, September 19–22, 2016.Google Scholar
Shields, F. D. Jr., and Abt, S. R. 1989. Sediment deposition in cutoff meander bends and implications for effective management. Regulated Rivers: Research and Management 4, 381396.Google Scholar
Shields, F. D. Jr., and Knight, S. S. 2013. Floodplain restoration with flood control: fish habitat value of levee borrow pits. Ecological Engineering 53, 217227.Google Scholar
Shinkle, K., and Dokka, R. K. 2004. Rates of Vertical Displacement at Benchmarks in the lower Mississippi Valley and the Northern Gulf Coast, National Oceanic and Atmospheric Administration, Technical Report 50.Google Scholar
Shull, C. A. 1922. The formation of a new island in the Mississippi River. Ecological Society of America 3, 202206.Google Scholar
Silc, T., and Sanek, W. 1977. Bulk density determination of several peats in northern Ontario using the Von Post humification scale. Canadian Journal of Soil Science 57, 75.Google Scholar
Silva, W., Klijn, F., and Dijkman, J. P. 2001. Room for the Rhine Branches in the Netherlands: What the Research Has Taught Us. RIZA, Deltares.Google Scholar
Silva, W., Dijkman, J., and Loucks, P. 2004. Flood management options for The Netherlands. International Journal of River Basin Management 2, 101112.Google Scholar
Silver, M., and Griffin, C. R. 2009. Nesting habitat Characteristics of bank swallows and belted kingfishers on the Connecticut River. Northeastern Naturalist 16, 519534.Google Scholar
Simon, A., and Collison, A. J. C. 2001. Pore-water pressure effects on the detachment of cohesive streambeds: seepage forces and matric suction. Earth Surface Processes and Landforms 26, 14211442.Google Scholar
Simon, A., and Darby, S. E. 1997. Process-form interactions in unstable sand-bed river channels: a numerical modeling approach. Geomorphology 21, 85106.Google Scholar
Simon, A., and Rinaldi, M. 2006. Disturbance, stream incision, and channel evolution: the roles of excess transport capacity and boundary materials in controlling channel response. Geomorphology 79, 361383.Google Scholar
Simon, A., Curini, A., Darby, S. E., and Langendoen, E. J. 2000. Bank and near-bank processes in an incised channel. Geomorphology 35, 193217.Google Scholar
Singer, M. B., and Aalto, R. 2009. Floodplain development in an engineered setting. Earth Surface Processes and Landforms 34, 291304.Google Scholar
Singer, M. B., Aalto, R., and James, A. 2008. Status of the lower Sacramento River flood control works in the context of its natural geomorphic setting. Natural Hazards Review 9, 104114. doi: 10.1061/(ASCE)1527-6988(2008)9:3(104).Google Scholar
Singer, M. B., Aalto, R., James, L. A., Kilham, N. E. Higson, J. L., and Ghoshal, S. 2013. Enduring legacy of a toxic fan via episodic redistribution of California gold mining debris. Proceedings National Academy of Science 46, 1843618441.Google Scholar
Slingerland, R. L., and Smith, N. D. 2004. River avulsions and their deposits. Annual Review of Earth Planetary Sciences 32, 257285.Google Scholar
Smith, D. G. 1986. Anastomosing river deposits, sedimentation rates and basin subsidence, Magdalena River, Northwestern Columbia, South America. Sedimentary Geology 46, 177196.Google Scholar
Smith, D. L., Miner, S. P., Theiling, C., Behm, R., and Nestler, J. M. 2017. Levee setbacks: an innovative, cost-effective, and sustainable solution for improved flood risk management. US Army Corps of Engineers. Report ERDC/EL SR-17–3. doi: 10.21079/11681/22736.Google Scholar
Smith, H. N. 1947. Rain follows the plow: the notion of increased rainfall for the Great Plains, 1844–1880. Huntington Library Quarterly 10, 169193.Google Scholar
Smith, L. M., and Winkley, B. R. 1996. The response of the lower Mississippi River to river engineering. Engineering Geology 45, 433455.Google Scholar
Smith, N. D., and Pérez-Arlucea, M. 1994. Fine-grained splay deposits in the avulsion belt of the lower Saskatchewan River, Canada. Journal of Sedimentary Research B64, 159168.Google Scholar
Smith, N. D., Cross, T. A., Dufficy, J. P., and Clough, S. R. 1989. The anatomy of an avulsion. Sedimentology 36, 123.Google Scholar
Smith, N. D., Morozova, G. S., Pérez-Arlucea, M., and Gibling, M. R. 2016. Dam-induced and natural channel changes in the Saskatchewan River below the E.B. Campbell Dam, Canada. Geomorphology 269, 186202.Google Scholar
Smith, V. B., and Mohrig, D. 2017. Geomorphic signature of a dammed sandy river: The lower Trinity River downstream of Livingston Dam in Texas, USA. Geomorphology 297, 122136.Google Scholar
Snedden, G. A., Cable, J. E., Swarzenski, C., and Swenson, E. 2007. Sediment discharge into a subsiding Louisiana deltaic estuary through a Mississippi River diversion. Estuarine Coastal and Shelf Science 71, 181193.Google Scholar
Somoza, L., and Rodríguez-Santalla, I. 2014. Geology and geomorphological evolution of the Ebro River Delta. In Gutiérrez, F., and Gutiérrez, M. (eds.), Landscapes and Landforms of Spain, World Geomorphological Landscapes. Springer, Dordrecht, pp. 213227.Google Scholar
Soniat, T. M., Conzelmann, C. P., Byrd, J. D. et al. 2013. Predicting the effects of proposed Mississippi River diversions on oyster habitat quality; application of an oyster habitat suitability index model. Journal of Shellfish Research 32, 629638. doi: 10.2983/035.032.0302.Google Scholar
Sparks, R. E. 1995. Need for ecosystem management of large rivers and their floodplains. BioScience 45, 168182.Google Scholar
Spreafico, M., and Lehmann, C. (eds.). 2009. Erosion, transport and deposition of sediment – case study Rhine. Contribution to the International Sediment Initiative of UNESCO/IHP, International Commission for the Hydrology of the Rhine Basin, Report no II-20 of the CHR.Google Scholar
Springborn, M., Singer, M. B., and Dunne, T. 2011. Sediment-adsorbed total mercury flux through Yolo Bypass, the primary floodway and wetland in the Sacramento Valley, California. Science of the Total Environment 412–413, 203213. doi: 10.1016/j.scitotenv.2011.10.004.Google Scholar
Stanford, J. A., and Ward, J. V. 1993. An ecosystem perspective of alluvial rivers: connectivity and the hyporheic corridor. Journal of the North American Benthological Society 12, 4860.Google Scholar
Stanley, D.-J. 1997. Mediterranean deltas: subsidence as a major control on relative sea-level rise. In CIESM Science Series n°3. Transformations and Evolution of the Mediterranean Coastline 35–62.Google Scholar
Stanley, D.-J., and Clemente, P. L. 2014. Clay distributions, grain sizes, sediment thicknesses, and compaction rates to interpret subsidence in Egypt’s Northern Nile Delta. Journal of Coastal Research 30, 88101.Google Scholar
Stanley, D.-J., and Hait, A. K. 2000. Holocene depositional patterns, neotectonics and sundarban mangroves in the Western Ganges-Brahmaputra Delta. Journal of Coastal Research 16, 2639.Google Scholar
Stanley, D.-J., and Warne, A. G. 1994. Worldwide initiation of Holocene marine deltas by deceleration of sea-level rise. Science 265(5169), 228231.Google Scholar
Stanley, D.-J., and Warne, A. G. 1998. Nile delta in its destruction phase, Journal of Coastal Research 14, 794825Google Scholar
Stanley, E. H., and Doyle, M. W. 2C3. Trading off: the ecological effects of dam removal. Frontiers in Ecology and the Environment 1, 1522.Google Scholar
Stegner, W. 1954. Beyond the Hundredth Meridian: John Wesley Powell and the Second Opening of the West. Houghton Mifflin Company, Boston, 438pp.Google Scholar
Sternberg, H. O’R. 1960. Radiocarbon Dating as Applied to a Problem of Amazonian Morphology. International Geographical Union, Congrès International de Géographie, 18th, Proceedings, vol. 2, Washington, DC, pp. 399–424.Google Scholar
Stevaux, J. C., Corradini, F. A., and Aquino, S. 2013. Connectivity processes and riparian vegetation of the upper Paraná River, Brazil. Journal of South American Earth Sciences 46, 113121. doi: 10.1016/j.jsames.2011.12.007.Google Scholar
Stouthamer, E., and Berendsen, H. J. A. 2001. Avulsion history, avulsion frequency, and intervulsion period of Holocene channel belts in the Rhine-Meuse Delta (The Netherlands). Journal of Sedimentary Research 71, 589598.Google Scholar
Stouthamer, E., and van Asselen, S. 2015. Potential of Holocene deltaic sequences for subsidence due to peat compaction. Proceedings of the International Association of Hydrological Sciences 372, 173178. doi: 10.5194/piahs-372–173-2015.Google Scholar
Stouthamer, E., Cohen, M. K., and Gouw, M. J. P. 2011. Avulsion and its implications for fluvial-deltaic architecture: Insights from the Holocene Rhine-Meuse Delta. From River to Rock Record: The Preservation of Fluvial Sediments and their Subsequent Interpretation. SEPM (Society for Sedimentary Geology), Special Publication no. 97, 215–231.Google Scholar
STOWA. 2012. Inspection manuals for flood defense systems. Technical information for the performance of inspections. STOWA, Report 14.Google Scholar
Straatsma, M. W., and Kleinhans, M. G. 2018. Flood hazard reduction from automatically applied landscaping measures in RiverScape, a Python package coupled to a two-dimensional flow model. Environmental Modelling and Software 101, 102116.Google Scholar
Straatsma, M. W., Fliervoet, J. M., Kabout, J. A. H., Baart, F., and Kleinhans, M. G. 2019. Towards multi-objective optimization of large-scale fluvial landscaping measures. Natural Hazards and Earth Systems Science 19, 11671187. doi: 10.5194/nhess-19-1167-2019.Google Scholar
Straatsma, M. W., Schipper, A., van der Perk, M., van den Brink, C., Leuven, R., and Middelkoop, H. 2009. Impact of value-driven scenarios on the geomorphology and ecology of lower Rhine floodplains under a changing climate. Landscape and Urban Planning 92, 160174.Google Scholar
Suanez, S., and Provensal, M. 1996. Morphosedimentary behaviour of the deltaic fringe in comparison to relative sea level rise on the Rhône River delta. Quaternary Science Reviews 15, 811818.Google Scholar
Summerfield, M. A. 1991. Global Geomorphology. Routledge.Google Scholar
Sun, W., Shao, Q., Liu, J., and Zhai, J. 2014. Assessing the effects of land use and topography on soil erosion on the Loess Plateau in China. Catena 121, 151163.Google Scholar
Swartz, J. M., Goudge, T. A., and Mohrig, D. C. 2020. Quantifying coastal fluvial morphodynamics over the Last 100 years on the lower Rio Grande, USA and Mexico. Journal of Geophysical Research: Earth Surface 125(6). doi: 10.1029/2019JF005443.Google Scholar
Syvitski, J. P. M., and Kettner, A. J. 2011. Sediment flux and the Anthropocene. Philosophical Transactions of the Royal Society A. doi: 10.1098/rsta.2010.0329.Google Scholar
Syvitski, J. P. M., Vörösmarty, C., Kettner, A. J., and Green, P. 2005. Impact of humans on the flux of terrestrial sediment to the global coastal ocean. Science 308, 376380.Google Scholar
Syvitski, J. P. M, Kettner, A. J., Overeem, I., et al. 2009. Sinking deltas due to human activities. Nature Geoscience 2, 681686.Google Scholar
Syvitski, J. P. M., Overeem, I., Brackenridge, R., and Hannon, M. T. 2012. Floods, floodplains, delta plains – a satellite imaging approach. Sedimentary Geology. doi: 10.1016/j.sedgeo.2012.05.014.Google Scholar
Syvitski, J. P. M., Kettner, A. J., Overeem, I., et al. 2013. Anthropocene metamorphosis of the Indus Delta and lower floodplain. Anthropocene 3, 2435.Google Scholar
Te Brake, W. H. 2002. Taming the waterwolf: hydraulic engineering and water management in the Netherlands during the Middle Ages. Technology and Culture 43, 475499.Google Scholar
Telcik, N., and Pattiaratchi, C. 2014. Influence of Northwest Cloudbands on Southwest Australian Rainfall. Journal of Climatology. doi: 10.1155/2014/671394.Google Scholar
Tellman, B., Sullivan, J.A., Kuhn, C., Kettner, A.J., Doyle, C.S., Brakenridge, G.R., Erickson, T.A., Slayback, D.A. 2021. Satellite imaging reveals increased proportion of population exposed to floods. Nature 596, 80?86. https://doi.org/10.1038/s41586-021-03695-wGoogle Scholar
Ten Brinke, W. B. T. 2005. The Dutch Rhine, a Restrained River. Veen Magazines, Diemen.Google Scholar
Ten Brinke, W. B. T., Schulze, F. H., and van der Veer, P. 2004. Sand exchange between groyne-field beaches and the navigation channel of the Dutch Rhine: the impact of navigation versus river flow. River Research and Applications 20, 889928.Google Scholar
Tena, A., and Batalla, R. J. 2013. The sediment budget of a large river regulated by dams (the lower River Ebro, NE Spain). Journal of Soils and Sediment 13, 966980.Google Scholar
Tennessee Valley Authority (TVA). A map of the profile of the Tennessee River, showing the TVA dams and facilities. 24158 Collection: Dept. of Conservation Photograph Collection Accession No.: RG 82 File Location: Box 69, File 165.Google Scholar
Terry, M. 1945. Soil erosion in Australia. The Geographical Journal 105, 121129.Google Scholar
Thompson, D. W. J., and Wallace, J. M. 1998. The Arctic Oscillation signature in the wintertime geopotential height and temperature fields. Geophysical Research Letters 25, 12971300. doi: 10.1029/98GL00950.Google Scholar
Thompson, D. W. J., and Wallace, J. M. 2000. Annular modes in the extratropical circulation. Part I: month-to-month variability. Journal of Climate 13, 10001016. doi: 10.1175/1520–0442.Google Scholar
Thompson, G., Hartman, B., Miller, B., et al. 2019. Development of borrow areas along the lower Mississippi River. In Rosati, J. D., Wang, P., and Vallee, M. (eds.), Coastal Sediments 2019 – Proceedings of the 9th International Conference. World Scientific, Hackensack, NJ and London, pp. 870879.Google Scholar
Thoms, M. C. 2003. Floodplain–river ecosystems: lateral connections and the implications of human interference. Geomorphology 56, 335350.Google Scholar
Thonon, I., Middelkoop, J., and van der Perk, M. 2007. The influence of floodplain morphology and river works on spatial patterns of overbank deposition. Netherlands Journal of Geosciences 86(1), 6375.Google Scholar
Thorne, C. R. 1989. Bank processes on the Red River between Index, Arkansas and Shreveport. Louisiana Final Report to the US Army European Research Office, London, England, under Contract No. DAJ45–88-C-0018.Google Scholar
Thorne, C. R. 1991. Bank erosion and meander migration of the Red and Mississippi Rivers, USA. Hydrology for the Water Management of Large River Basins (Proceedings of the Vienna Symposium, August 1991). International Association of Hydrological Sciences, no. 201.Google Scholar
Thorne, C. R. 1998. Channel types and morphological classification. In Thorne, C. R., Hey, R. D., and Newson, M. D. (eds.), Guidebook of Applied Fluvial Geomorphology for River Engineering and Management. Wiley, Chichester, pp. 175222.Google Scholar
Thorne, C. R. 2002. Geomorphic analysis of large alluvial rivers. Geomorphology 44, 203–219.Google Scholar
Thorne, C. R., and Abt, S. R. 1993. Analysis of riverbank instability due to toe scour and lateral erosion. Earth Surface Processes and Landforms 18, 835843.Google Scholar
Thorsteinson, L. K., Becker, P. R., and Hale, D. A. 1989. The Yukon Delta A Synthesis of Information. NOAA National Ocean Service, Ocean Assessments Division, Anchorage, AK.Google Scholar
Tian, J., Chang, J., Zhang, Z., Wang, Y., Wu, Y., and Jiang, T. 2019. Influence of Three Gorges Dam on downstream low flow. Water 11, 65. doi: 10.3390/w11010065.Google Scholar
Tiner, R. W. 2016. Wetland Indicators: A Guide to Wetland Formation, Identification, Delineation, Classification, and Mapping, 2nd ed. CRC Press, Taylor and Francis Group, Boca Raton, 630 p.Google Scholar
Tockner, K., Malard, F., and Ward, J. V. 2000. An extension of the flood pulse concept. Hydrological Processes 14, 28612883.Google Scholar
Tockner, K., Bernhardt, E. S., Koska, A., and Zarfl, C. 2016. A global view on future major water engineering projects. In Hüttl, R. F., et al. (eds.), Society-Water-Technology, Water Resources Development and Management. Springer Open, Heidelberg, pp. 4764.Google Scholar
Toole, J. K. 1980. A Confederacy of Dunces. Louisiana State University Press, Baton Rouge, LA, 405pp.Google Scholar
Toonen, W. H. J., Kleinhans, M. G., and Cohen, K. M. 2012. Sedimentary architecture of abandoned channel fills. Earth Surface Processes and Landforms 37, 459472.Google Scholar
Toonen, W. H. J., Middelkoop, H., Konijnendijk, T. Y. M., Macklin, M. G., and Cohen, K. M. 2016. The influence of hydroclimatic variability on flood frequency in the Lower Rhine. Earth Surface Processes and Landforms 41. doi: 10.1002/esp.3953.Google Scholar
Törnblom, J., Angelstam, P., Degerman, E., and Tamario, C. 2017. Prioritizing dam removal and stream restoration using critical habitat patch threshold for brown trout (Salmo trutta L.): a catchment case study from Sweden. Écoscience 34, 157166.Google Scholar
Törnqvist, T. E., and Bridge, J. S. 2003. Spatial variation of overbank aggradation rate and its influence on avulsion frequency. Sedimentology 49, 891905.Google Scholar
Törnqvist, T. E., Kidder, T. R., Autin, W. J., et al. 1996. A revised chronology for Mississippi River subdeltas. Science 273, 16931696.Google Scholar
Törnqvist, T. E., Bick, S. J., van der Borg, K., and de Jong, A. F. M. 2006. How stable is the Mississippi delta? Geology 34, 697700.Google Scholar
Törnqvist, T. E., Wallace, D. J., Storms, J. E. A., et al. 2008. Mississippi Delta subsidence primarily caused by compaction of Holocene strata. Nature Geoscience 1, 173176.Google Scholar
Torres, N., and Harrelson, D. W. 2012. The great Red River raft and its sedimentological implications. In Reconstructing Human-Landscape Interactions. Springer Briefs in Earth Systems Sciences, 1, pp. 3555.Google Scholar
Toth, L. A., Obeysekera, J. T. B., Perkins, W. A., and Loftin, M. K. 1993. Flow regulation and restoration of Florida’s Kissimmee river. Regulated Rivers: Research and Management 8, 155166.Google Scholar
Tran, D. D., van Halsema, G., Hellegers, P. J. G. J., et al. 2018. Assessing impacts of dike construction on the flood dynamics of the Mekong Delta. Hydrology and Earth Systems Science 22, 18751896. doi: 10.5194/hess-221875-2018.Google Scholar
Trenberth, K. E. 1997. The definition of El Niño. Bulletin of the American Meteorological Society 78, 27712777.Google Scholar
Trenberth, K. E., Caron, J. M., Stepaniak, D. P., and Worley, S. 2002. Evolution of El Niño–Southern Oscillation and global atmospheric surface temperatures. Journal of Geophysical Research – Atmospheres 107, AAC 5–1-AAC 5–17.Google Scholar
Trigg, M. A., Bates, P. D., Wilson, M. D., Schumann, G., and Baugh, C. 2012. Floodplain channel morphology and networks of the middle Amazon River. Water Resources Research 48. doi: 10.1029/2012WR01188.W10504.Google Scholar
Trimble, S. W. 1974. Man-Induced Soil Erosion on the Southern Piedmont, 1700–1970. Soil and Water Conservation Society, Ankeny, Iowa.Google Scholar
Trimble, S. W. 1976. Sedimentation in Coon Creek valley, Wisconsin. In Proceedings of the Third Federal Interagency Sedimentation Conference, Denver, Water Resources Council, Section 5, pp. 100–112.Google Scholar
Trimble, S. W. 2013. Historical Agriculture and Soil Erosion in the Upper Mississippi Valley Hill Country. CRC Press, Boca Raton, FL.Google Scholar
Tullos, D. D., Collins, M. J., Bellmore, J. R., et al. 2016. Synthesis of common management concerns associated with dam removal. Journal of the American Water Resources Association 52, 11791206.Google Scholar
Turnbull, W. J., Krinitzky, E. L., and Weaver, F. J. 1966. Bank erosion in soils of the lower Mississippi Valley. Journal of the Soil Mechanics and Foundations Division 92, 121136.Google Scholar
IITurner, B. L., Lambin, E. F., and Reenberg, A. 2007. Land Change Science Special Feature: the emergence of land change science for global environmental change and sustainability. Proceedings of the National Academy of Sciences 52, 2066620671.Google Scholar
Twain, M. 1883. Life on the Mississippi. James R. Osgood & Co., Boston, MA.Google Scholar
Twilley, R. R., Bentley, S. J., Chen, Q., et al. 2016. Co-evolution of wetland landscapes, flooding, and human settlement in the Mississippi River Delta Plain. Sustainability Science 11, 711. doi: 10.1007/s11625-016-0374-4.Google Scholar
Tye, R. S., and Coleman, J. M. 1989. Depositional processes and stratigraphy of fluvially dominated lacustrine deltas: Mississippi delta plain. Journal of Sedimentary Petrology 59, 973996.Google Scholar
U.K. Environment Agency. 2015. Management of the London Basin Chalk Aquifer, Status Report, Environment Agency, UK.Google Scholar
U.S. Army Corps of Engineers (USACE). 1982. Yazoo Area Pump Project: Yazoo Area and Satartia Area Backwater Levee Projects, Yazoo Backwater Area, Mississippi. U.S. Army Corps of Engineers, Vicksburg, MS.Google Scholar
U.S. Army Corps of Engineers (USACE). 1988. Final Report: Steele Bayou Gravity Control Structure, Hydraulic Model Investigation. U.S. Army Corps of Engineers, Waterways Experiment Station, Vicksburg, MS.Google Scholar
U.S. Army Corps of Engineers (USACE). 1998a. Mississippi River Mainline Levees Enlargement and Seepage Control, Cape Girardeau, Missouri to Head of Passes, Louisiana, V. I.Google Scholar
U.S. Army Corps of Engineers (USACE). 1998b. Mississippi River Mainline Levees Enlargement and Seepage Control, Cape Girardeau, Missouri to Head of Passes, Louisiana, V. II.Google Scholar
U.S. Army Corps of Engineers (USACE). 1998c. Mississippi River Mainline Levees Enlargement and Seepage Control, Cape Girardeau, Missouri to Head of Passes, Louisiana, V. III.Google Scholar
U.S. Army Corps of Engineers (USACE). 2000a. Design and construction of levees. US Army Corps of Engineers, Engineering and Design, Engineers Manual No. 1110–2-1913.Google Scholar
U.S. Army Corps of Engineers (USACE). 2000b. Yazoo Backwater Area Reformulation Report, Vols. I, II, III (17 App.), U.S. Army Corps of Engineers, Vicksburg, MS.Google Scholar
U.S. Army Corps of Engineers (USACE). 2009a. Dutch perspective appendix. U.S. Army Corps of Engineers New Orleans District, Mississippi Valley Division, Louisiana Coastal Protection and Restoration. Final Technical Report.Google Scholar
U.S. Army Corps of Engineers (USACE). 2009b. Incorporating sea-level change considerations in Civil Works Programs Engineer Circular (EC), 1165–2-211.Google Scholar
U.S. Army Corps of Engineers (USACE). 2012. Bonnet Carré Spillway, U.S. Army Corps of Engineers, New Orleans District, Louisiana.Google Scholar
U.S. Army Corps of Engineers (USACE). 2013. Modification of Caernarvon Diversion. Fact Sheet, U.S. Army Corps of Engineers, New Orleans Division. www.lca.gov/ (accessed September 29, 2018).Google Scholar
U.S. Army Corps of Engineers (USACE). 2014a. Procedures to evaluate sea level change: impacts, responses and adaptation. Engineer Technical Letter 1100–2-1.Google Scholar
U.S. Army Corps of Engineers (USACE). 2014b. Sacramento river bank protection project environmental impact statement / report. U.S. State Clearinghouse #2009012081.Google Scholar
U.S. Army Corps of Engineers (USACE). 2015. American river watershed: common features general reevaluation report. Attachment-E, Draft Erosion Protection Report. USACE Sacramento District.Google Scholar
U.S. Army Corps of Engineers – New Orleans Division (USACE). 2017. Internet Web page “Corps Needs Levee Clay.” www.mvn.usace.army.mil/Missions/HSDRRS/ Corps-Needs-Levee-Clay/Borrow-Material/ (accessed January 18, 2018).Google Scholar
U.S. Army Corps of Engineers (USACE). 2018a. A summary of risks and benefits associated with the USACE levee portfolio. U.S. Army Corps of Engineers Levee Portfolio Report. Prepared by U.S. Army Corps of Engineers Levee Safety Program.Google Scholar
U.S. Army Corps of Engineers (USACE). 2018b. Yuba river ecosystem restoration feasibility study California. Draft Interim Feasibility Report and Environmental Assessment. U.S. Army Corps of Engineers, Sacramento District.Google Scholar
U.S. Army Corps of Engineers (USACE). 2019. Lake Pontchartrain and Vicinity, Louisiana General Re-Evaluation Report with Integrated Environmental Impact Statement. U.S. Army Corps of Engineers, New Orleans.Google Scholar
U.S. Army Corps of Engineers (USACE). 2020. Beneficial use of dredged material: history 2015–2020. In Mississippi River, Baton Rouge to the Gulf of Mexico, LA Head of Passes Hopper Dredge Disposal Area. U.S. Army Corps of Engineers, New Orleans. www.mvn.usace.army.mil/About/Offices/Operations/Beneficial-Use-of-Dredged-Material/hopper-dredge-disposal-area/ (accessed June 2020).Google Scholar
U.S. Army Corps of Engineers (USACE) Sacramento District. 2020. Sacramento levee upgrades. www.spk.usace.army.mil/Missions/Civil-Works/Sacramento-Levee-Upgrades/.Google Scholar
U.S. Congress. 1928. Flood Control Act, Mississippi River and Tributaries, Appendix E, 70th Congress, Session I Ch. 596.Google Scholar
U.S. Environmental Protection Agency (U.S. EPA). 2008. Final determination of the U.S. Environmental Protection Agency’s Assistant Administrator for Water Pursuant to Section 404(C) of the Clean Water Act Concerning the Proposed Yazoo Backwater Area Pumps Project, Issaquena County, Mississippi, August 31, 2008 (with 6 appendices).Google Scholar
U.S. Environmental Protection Agency (U.S. EPA). 2001. Fact Sheet: Mercury Update: Impact on Fish Advisories. EPA-823-F-01-011. Office of Water 4305, Washington, DC.Google Scholar
U.S. Environmental Protection Agency (U.S. EPA). 2016a. Second five-year review report for the Milltown Reservoir Clark Fork River Superfund Site EPA ID MTD980717565. Milltown, Missoula, Granite, Powell, and Deer Lodge Counties, Montana. United States Environmental Protection Agency Region 8 Denver, Colorado, September 2016.Google Scholar
U.S. Environmental Protection Agency (U.S. EPA). 2016b. Frequently asked questions on removal of obsolete dams. Office of Water, EPA-840-F-16–001 December 2016.Google Scholar
U.S. Fish and Wildlife Service (U.S. FWS). 2001. U.S. Fish and Wildlife perspective on the Corps of Engineers proposed Yazoo Pumps Project. Fact Sheet, U.S. Fish and Wildlife Service, Jackson Ecological Services Field Office, Jackson, Mississippi. www.fws.gov/southeast/pubs/facts/yazooback.pdf (accessed May 19, 2016).Google Scholar
U.S. Geological Survey (USGS). 2010. Land subsidence in California. U.S. Geological Survey web site at http://ca.water.usgs.gov/groundwater/sub/ (accessed April 9, 2016).Google Scholar
U.S. Geological Survey (USGS). 2013. Trends and causes of historical wetland loss in coastal Louisiana. Fact Sheet 2013-3017. March 2013. (https://pubs.usgs.gov/fs/2013/3017/pdf/fs2013-3017.pdf)Google Scholar
U.S. Geological Survey (USGS). 2018. Endangered, discontinued and rescued stream gages. https://water.usgs.gov/networks/fundingstability/ (accessed November 11, 2018).Google Scholar
U.S. Government Accountability Office (U.S. GAO). 2005. Testimony before the subcommittee on energy and water development, Committee on Appropriations, House of Representatives, Army Corps of Engineers Lake Pontchartrain and Vicinity Hurricane Protection Project, Statement of Anu Mittal, Director Natural Resources and Environment, GAO-05–1050T.Google Scholar
U.S. Government Accountability Office (U.S. GAO). 2006a. Hurricane Katrina: strategic planning needed to guide future enhancements beyond interim levee repairs. Report to Congressional Committees, GAO-06–934.Google Scholar
U.S. Government Accountability Office (U.S. GAO). 2006b. Testimony before the Committee on Homeland Security and Governmental Affairs, U.S. Senate, Hurricane Protection: Statutory and Regulatory Framework for Levee Maintenance and Emergency Response for the Lake Pontchartrain Project, Statement for the Record by Anu K. Mittal, Director Natural Resources and Environment , GAO-06–322T.Google Scholar
U.S. Government Accountability Office (GAO). 2011. Mississippi river: actions are needed to help resolve environmental and flooding concerns about the use of river training structures. Rep. GAO-12–41. http://gao.gov/products/GAO-12-41.Google Scholar
U.S. National Inventory of Dams (NID). 2020. https://nid.sec.usace.army.mil/.Google Scholar
U.S. Society on Dams (USSD). 2015. Guidelines for Dam Decommissioning Projects. Committee on Dam Decommissioning, U. S. Society on Dams, Denver, CO.Google Scholar
Uehlinger, U., Wantzen, K. M., Leuven, R. S. E. W., and Arndt, H. 2009. The Rhine River Basin. In Tockner, K., Robinson, C. T., and Uehlinger, U. (eds.), Rivers of Europe. Elsevier – Academic Press, San Diego, CA. , pp. 147.Google Scholar
UNESCO-IHP. 2011. Sediment issues and sediment management in large river basins: interim case study synthesis report. International Sediment Initiative, Technical Documents in Hydrology, CN/2011/SC/IHP/PI/2, UNESCO International Hydrological Program, Beijing.Google Scholar
United Nations World Water Development Report (UNWWDR). 2020. Water and Climate Change. Food and Agriculture Organization of the United Nations, Rome, Italy.Google Scholar
van Asselen, S., Stouthamer, E., and van Asch, T. W. J. 2009. Effects of peat compaction on delta evolution: a review on processes, responses, measuring and modeling. Earth Science Reviews 92, 3551. doi: 10.1016/j.earscirev.2008.11.001.Google Scholar
van Asselen, S., Karsenberg, D., and Stouthamer, E. 2011. Contribution of peat compaction to relative sea-level rise within Holocene deltas. Geophysical Research Letters, L24401. doi: 10.1029/2011GL049835.Google Scholar
van Asselen, S., Cohen, K. M., and Stouthamer, E. 2017. The impact of avulsion on groundwater level and peat formation in delta floodbasins during the middle-Holocene transgression in the Rhine-Meuse delta, The Netherlands. The Holocene 27, 16941706.Google Scholar
van Baars, S. 2005. The horizontal failure mechanism of the Wilnis peat dyke. Géotechnique 55, 319323.Google Scholar
van Baars, S., and van Kempen, I. M. 2009. The causes and mechanisms of historical dike failures in the Netherlands. E-Water, Official Publication of the European Water Association (EWA), ISSN1994–8549.Google Scholar
van Balen, R. T., Houtgast, R. F., and Cloetingh, S. A. P. L. 2005. Neotectonics of The Netherlands: a review. Quaternary Science Reviews 24, 439454.Google Scholar
van Balen, R., Kasse, C., and de Moor, J. 2008. Impact of groundwater flow on meandering; example from the Geul River, The Netherlands. Earth Surface Processes and Landforms 33, 20102028.Google Scholar
van Binh, D., Kantoush, S., and Sumi, T. 2019. Changes to long-term discharge and sediment loads in the Vietnamese Mekong Delta caused by upstream dams. Geomorphology 335, March 2020, 107011.Google Scholar
van de Ven, G. 2004. Man-Made Lowlands: History of Water Management and Land Reclamation in the Netherlands, 4th ed. Uitgeverij Matrijs, Utrecht.Google Scholar
van de Ven, G. 2007. Verdeel en hbeheers! 300 jar Pannerdensch Kanaal. Veen Magazines, Diemen.Google Scholar
van den Berg, J. H. 1995. Prediction of alluvial channel pattern of perennial rivers. Geomorphology 12, 259279.Google Scholar
van den Brink, M., Meijerink, S., and Termeer, C. 2014. Climate-proof planning for flood-prone areas: assessing the adaptive capacity of planning institutions in the Netherlands. Regional Environmental Change 14, 981. doi: 10.1007/s10113–012–0401-7.Google Scholar
van Denderen, R. P., Schielen, R. M. J., Straatsma, M. W., Kleinhans, M. G., and Hulscher, S. J. M. H. 2019a. A characterization of side channel development. River Research and Applications 35, 1597–1603.Google Scholar
van Denderen, R. P., Schielen, R. M. J., Westerhof, S. G., Quartel, S., and Hulscher, S. J. M. H. 2019b. Explaining artificial side channel dynamics using data analysis and model calculations. Geomorphology 327, 93110.Google Scholar
van der Meulen, M. J., Wiersma, A. P., van der Perk, M., Middelkoop, H., and Hobo, M. 2009. Sediment management and the renewability of floodplain clay for structural ceramics. Journal of Soils and Sediments 9, 627639. doi: 10.1007/s11368–009–0115-8.Google Scholar
van der Most, M., and Hudson, P. F. 2018. The influence of floodplain geomorphology and hydrologic connectivity on alligator gar (Atractosteus spatula) habitat along the embanked floodplain of the lower Mississippi River. Geomorphology 302, 6275.Google Scholar
van der Perk, M., Sutaria, C. A. T., and Middelkoop, H. 2019. Examination of the declining trend in suspended sediment loads in the Rhine River in the period 1952–2016. In Stouthamer, E., Middelkoop, H., Kleinhans, M., van der Perk, M., and Straatsma, M. (eds.), Land of Rivers: NCR DAYS 2019 Proceedings. Netherlands Centre for River Studies Publication 43–2019.Google Scholar
van Dijk, A. I. J. M., Beck, H. E., Crosbie, R. S., et al. 2013. The Millennium Drought in southeast Australia (2001–2009): natural and human causes and implications for water resources, ecosystems, economy, and society. Water Resources Research 49, 10401057. doi: 10.1002/wrcr.20123.Google Scholar
van Dijk, W. M., Teske, R., van de Lageweg, W. I., and Kleinhans, M. G. 2013. Effects of vegetation distribution on experimental river channel dynamics. Water Resources Research 49, 75587574.Google Scholar
Van Dinter, M., and Van Zijverden, W. K. 2010. Settlement and land use on crevasse splay deposits; geoarchaeological research in the Rhine-Meuse Delta, the Netherlands. Netherlands Journal of Geosciences – Geologie en Mijnbouw 89, 2134.Google Scholar
van Heerden, I., and Roberts, H. H. 1980. The Atchafalaya Delta: Louisiana’s new prograding coast. Gulf Coast Association of Geological Societies, Transactions 30, 497506.Google Scholar
van Heezik, A. 2008) Battle over the Rivers: Two Hundred Years of River Policy in the Netherlands. Rijkswaterstaat, Haarlem, Netherlands.Google Scholar
van Loo, M. Dusar, B., Verstraeten, G., et al. 2017. Human induced soil erosion and the implications on crop yield in a small mountainous Mediterranean catchment (SW-Turkey). Catena 149, 491504.Google Scholar
van Looy, K., Honnay, O., Bossuyt, B., and Hermy, M. 2003. The effects of river embankment and forest fragmentation on the plant species richness and composition of floodplain forest in the Meuse Valley, Belgium. Belgian Journal of Botany 136, 97108.Google Scholar
van Veen, J. 1962. Dredge, Drain and Reclaim: Art of a Nation, 5th ed. Martinus Nijhoff, The Hague.Google Scholar
van Riel, W. 2011. Exploratory Study of Pluvial Flood Impacts in Dutch Urban Areas. Deltares, Delft.Google Scholar
van Vuren, S., Paarlberg, A., and Havinga, H. 2015. The aftermath of “Room for the River” and restoration works: coping with excessive maintenance dredging. Journal of Hydro-environment Research 9, 172186.Google Scholar
Vergouwe, R. 2016. The National Flood Risk Analysis for the Netherlands: Final Report. VNK Project, Rijkswaterstaat, The Netherlands.Google Scholar
Verstraeten, G., Broothaerts, N., van Loo, M., et al. 2017. Variability in fluvial geomorphic response to anthropogenic disturbance. Geomorphology 494, 2039. doi: 10.1016/j.geomorph.2017.03.027.Google Scholar
Vita-Finzi, C. 1969. The Mediterranean Valleys. Cambridge University Press, Cambridge.Google Scholar
Viale, M., Valenzuela, R., Garreaud, R. D., and Ralph, F. M. 2018. Impacts of atmospheric rivers on precipitation in Southern South America. Journal of Hydrometeorology 19, 16711687. doi: 10.1175/JHM-D-18-0006.1.Google Scholar
Vollmer, S. 2020. Geschiebezugabe – ein innovativer Ansatz zur Eindämmung der Sohlerosion. Bundesanstalt für Gewässerkunde (Attachment allowance – an innovative approach to contain erosion). Federal Institute for Hydrology. www.bafg.de/DE/01_Leistungen/02_F_E/Themen/Artikel/Geschiebezugabe.html (accessed June 5, 2020).Google Scholar
Vollmer, S., and Goelz, E. 2006. Sediment monitoring and sediment management in the Rhine River. In Sediment Dynamics and the Hydromorphology of Fluvial Systems (Proceedings of a symposium held in Dundee, UK, July 2006). IAHS Publ. 306, 231–240.Google Scholar
Vorogushyn, S., and Merz, B. 2013. Flood trends along the Rhine: the role of river training. Hydrology and Earth Systems Science 17, 38713884. doi: 10.5194/hess-17–3871-2013.Google Scholar
Vörösmarty, C. J., and Sahagian, D. 2000. Anthropogenic disturbance of the terrestrial water cycle. BioScience 50, 753765.Google Scholar
Vörösmarty, C. J., Meybeck, M., Fekete, B., Sharma, K., Green, P., and Syvitski, J. P. M. 2003. Anthropogenic sediment retention: major global impact from registered river impoundments. Global and Planetary Change 39, 169190.Google Scholar
Voulvoulis, N., Arpon, K. D., and Giakoumis, T. 2017. The EU Water Framework Directive: from great expectations to problems with implementation. Science of the Total Environment 575, 358366.Google Scholar
Wahl, K. L., Vining, K. C., and Wiche, G. J. 1993. Precipitation in the upper Mississippi River Basin, January 1 through July 31, 1993. U.S. Geological Survey Circular 1120-B.Google Scholar
Walker, H. J., and Hudson, P. F. 2003. Hydrologic and geomorphic processes in the Colville River Delta, Alaska. Geomorphology 56, 291303.Google Scholar
Walker, H. J., Arnborg, L., and Peippo, J. 1987. Riverbank erosion in the Colville Delta, Alaska. Geografiska Annaler: Series A 69, 6170.Google Scholar
Walling, D. E. 2006. Human impact on land-ocean sediment transfer by the world’s rivers. Geomorphology 79, 192216.Google Scholar
Walling, D. E. 2008a. The changing sediment loads of the world’s rivers. In Sediment Dynamics in Changing Environments (Proceedings of a symposium held in Christchurch, New Zealand, December 2008). IAHS Publication 325, 323–338.Google Scholar
Walling, D. E. 2008b. The changing sediment load of the Mekong River. AMBIO 37, 150157.Google Scholar
Walling, D. E. 2009. The sediment load of the Mekong River. In Campbell, I. (ed.), The Mekong: Biophysical Environment of an International River Basin. Aquatic Ecology. Elsevier – Academic Press, Amsterdam, pp. 113142.Google Scholar
Walter, R. C., and Merritts, D. J. 2008. Natural streams and the legacy of water-powered mills. Science 319, 299304.Google Scholar
Wang, B., and Xu, Y. J. 2015. Sediment trapping by emerged channel bars in the Lowermost Mississippi River during a major flood. Water 7, 60796096.Google Scholar
Wang, G., Wu, B., and Wang, Z. Y. 2005. Sedimentation problems and management strategies of Sanmenxia Reservoir, Yellow River, China. Water Resources Research 41, W09417. doi: 10.1029/2004WR003919.Google Scholar
Wang, H., Yang, Z., Saito, Y., Liu, J. P., and Xiaoxia, S. 2007. Stepwise decreases of the Huanghe (Yellow River) sediment load (1950–2004): impacts of climate change and human activities. Global and Planetary Change 57, 331354.Google Scholar
Wang, H., Chen, Q., La Peyre, M. K., Hu, K., and La Peyre, J. F. 2017. Predicting the impacts of Mississippi River diversions and sea-level rise on spatial patterns of eastern oyster growth rate and production. Ecological Modeling 352, 4053. doi: 10.1016/j.ecolmodel.2017.02.028.Google Scholar
Wang, H. J., Wright, T., Yu, Y., et al. 2012. InSAR reveals coastal subsidence in the Pearl River Delta, China. Geophysical Journal International 191, 11181128. doi: 10.1111/j.1365-246X.2012.05687.x.Google Scholar
Wang, J., Shen, Y., Gleason, C. J., and Wada, Y. 2013. Downstream Yangtze River levels impacted by Three Gorges Dam. Environmental Research Letters 8, 4012. doi: 10.1088/1748-9326/8/4/044012.Google Scholar
Wang, L., Ting, M., and Kushner, P. J. 2017. A robust empirical seasonal prediction of winter NAO and surface climate. Scientific Reports 7, 279. doi: 10.1038/s41598–017–00353-y.Google Scholar
Wang, Y., Zhao, W., Wang, S., Feng, X., and Liu, Y. 2019. Yellow River water rebalanced by human regulation. Scientific Reports 9, 9707. doi: 10.1038/s41598–019–46063-5.Google Scholar
Wang, Z.‐Y., and Hu, C. 2009. Strategies for managing reservoir sedimentation. International Journal of Sediment Research 24, 369384.Google Scholar
Wang, Z. Y., and Liang, Z. Y. 2000. Dynamic characteristics of the Yellow River mouth. Earth Surface Processes and Landforms 2, 765782.Google Scholar
Ward, J. V., and Stanford, J. A. 1995. The serial discontinuity concept: extending the model to floodplain rivers. Regulated Rivers: Research and Management 10, 159168.Google Scholar
Ward, P. J., Beets, W., Bouwer, L. M., Aerts, J. C. J. H., and Rensen, H. 2010. Sensitivity of river discharge to ENSO. Geophysical Research Letters 37, L12402. doi: 10.1029/2010GL043215.Google Scholar
Ward, P. J., Pauw, W. P., van Buuren, M. W., and Marfai, M. A. 2013. Governance of flood risk management in a time of climate change: the cases of Jakarta and Rotterdam. Environmental Politics 22, 518536. doi: 10.1080/09644016.2012.683155.Google Scholar
Warne, A. G., Toth, L. A., and White, W. A. 2000. Drainage-basin-scale geomorphic analysis to determine reference conditions for ecologic restoration – Kissimmee River, Florida. Geological Society of America Bulletin 112, 884899.Google Scholar
Wasklewicz, T., Greulich, S., Franklin, S., and Grubaugh, J. 2004. The 20th century hydrologic regime of the Mississippi River. Physical Geography 25, 208224.Google Scholar
Water Institute of the Gulf (WIG). 2016. Building land in Coastal Louisiana: expert recommendations for operating a successful sediment diversion that balances ecosystem and community needs. https://thewaterinstitute.org/assets/docs/reports/Peyronnin_2016-8-31.pdf (accessed October 5, 2018).Google Scholar
Watson, C. C., and Biedenharn, D. S. 2009. Specific gage analysis of stage trends on the Middle Mississippi River. U.S. Army Corps of Engineers, St. Louis, MO.Google Scholar
Watson, C. C., Biedenharn, D. S., and Thorne, C. R. 2013a. Analysis of the impacts of dikes on flood stages in the middle Mississippi River. Journal of Hydraulic Engineering 139(10), 10711078. doi: 10.1061/(ASCE)HY.1943-7900.0000786.Google Scholar
Watson, C. C., Holmes, R. R., and Biedenharn, D. S. 2013b. Mississippi River streamflow measurement techniques at St. Louis, Missouri. Journal of Hydraulic Engineering 139(10), 10621070. doi: 10.1061/(ASCE)HY.1943-7900.0000752.Google Scholar
Watson, C. C., Biedenharn, D. S., and Thorne, C. R. 2015a. Closure to “Analysis of the Impacts of Dikes on Flood Stages in the Middle Mississippi River” by Watson, Chester C., Biedenharn, David S., and Thorne, Colin R.. Journal of Hydraulic Engineering 141(8), 07015011-1. doi: 10.1061/(ASCE)HY.1943-7900.0001055.Google Scholar
Watson, C. C., Holmes, R. R., and Biedenharn, D. S. 2015b. Closure to “Mississippi River Streamflow Measurement Techniques at St. Louis, Missouri” by Watson, Chester C., Holmes, Robert R. Jr., and Biedenharn, David S.. Journal of Hydraulic Engineering 141(8), 07015008-1. doi: 10.1061/(ASCE)HY.1943-7900.0001021.Google Scholar
Watson, J. M., Coghlan, S. M. Jr., Zydlewski, J., Hayes, D. B., and Kiraly, I. A. 2018. Dam removal and fish passage improvement influence fish assemblages in the Penobscot River, Maine. Transactions of the American Fisheries Society 147, 525540.Google Scholar
Watson, K. M., Harwell, G. R., Wallace, D. S., et al. 2018. Characterization of peak streamflows and flood inundation of selected areas in southeastern Texas and southwestern Louisiana from the August and September 2017 flood resulting from Hurricane Harvey. U.S. Geological Survey Scientific Investigations Report 2018–5070.Google Scholar
World Commission on Dams (WCD). 2000. Dams and Development: A New Framework. Report of the World Commission on Dams for Decision-Making. Earthscan Publications, London.Google Scholar
Wei, W., Chang, Y., and Dai, Z. 2014. Streamflow changes of the Changjiang (Yangtze) River in the recent 60 years: impacts of the East Asian summer monsoon, ENSO, and human activities. Quaternary International 336, 98107.Google Scholar
Welder, F. A. 1955. Deltaic processes in Cubits Gap Area, Plaquemines Parish, Louisiana. PhD dissertation, Louisiana State University, Baton Rouge, LA.Google Scholar
Wells, J. T., Chinburg, S. J., and Coleman, J. M. 1983. The Atchafalaya River Delta, Report No. 4: Generic analysis of Delta Development. U.S. Corps of Army Engineers, Waterways Experiment Station, Technical Report HL-82–15.Google Scholar
Wheeler, K. G., Basheer, M., Mekonnen, Z. T., et al. 2016. Cooperative filling approaches for the Grand Ethiopian Renaissance Dam, Water International 41, 611634.Google Scholar
White, G. 1945. Human adjustment to floods. University of Chicago, Department of Geography, Research Paper, No 29.Google Scholar
Whitfield, P. H., Moore, R. D., Fleming, S. W., and Zawadzki, A. 2010. Pacific Decadal Oscillation and the hydroclimatology of western Canada – review and prospects. Canadian Water Resources Journal / Revue canadienne des ressources hydriques 35, 128.Google Scholar
Whiting, P., and Pomeranets, M. 1997. A numerical study of bank storage and its contribution to streamflow. Journal of Hydrology 202, 121136.Google Scholar
Whittaker, B. N., and Reddish, D. J. 1989. Subsidence: Occurrence, Prediction, and Control. Developments in Geotechnical Engineering, 56. Elsevier, Amsterdam.Google Scholar
Wilcock, P. R. 1998. Two‐fraction model of initial sediment motion in gravel‐bed rivers. Science 280, 410412.Google Scholar
Williams, A. P., Funk, C., Michaelsen, J., et al. 2012. Recent summer precipitation trends in the Greater Horn of Africa and the emerging role of Indian Ocean sea surface temperature. Climate Dynamics 39, 23072328.Google Scholar
Williams, G. P. 1978a. Bankfull discharge of rivers. Water Resources Research 14, 11411154.Google Scholar
Williams, G. P. 1978b. The case of the shrinking channels: The North Platte and Platte Rivers, Nebraska, US Geological Survey Circular 781.Google Scholar
Williams, G. P. 1984. Paleohydrological methods and some examples from Swedish fluvial environments, II – river meanders. Geografiska Annaler: Series A, Physical Geography 66, 89102.Google Scholar
Williams, G. P. 1986. River meanders and channel size. Journal of Hydrology 88, 147164.Google Scholar
Williams, G. P., and Wolman, M. G. 1984. Downstream effects of dams on alluvial rivers. U.S. Geological Survey, Professional Paper 1286.Google Scholar
Williams, S. J., Arsenault, M. A., Buczkowski, B. J., et al. 2009. Surficial sediment character of the Louisiana offshore continental shelf region: a GIS compilation. U.S. Geological Survey Open-File Report 2006–1195.Google Scholar
Wilson, K. V. 1979. Changes in channel characteristics, 1938–1974, of the Homochitto River and Tributaries, Mississippi. U.S. Geological Survey, Open-File Report 79–554.Google Scholar
Winemiller, K. O., Tarim, S., Shorman, D., and Cotner, J. B. 2000. Fish assemblage structure in relation to environmental variation among Brazos River oxbow lakes. Transactions of the American Fisheries Society 129, 451468.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, 128129.Google Scholar
WinklerPrins, A. M. G. A. 2002. Seasonal floodplain-upland migration along the Lower Amazon River. The Geographical Review 92, 415431.Google Scholar
Winkley, B. R. 1977. Man-made cutoffs on the lower Mississippi River, conception, construction and river response. U.S. Army Corps of Engineers, Potamology Investigations Report 300–2, Vicksburg, MS.Google Scholar
Winkley, B. R. 1994. Response of the lower Mississippi River to flood control and navigation improvements. In Schumm, S. A., and Winkley, B. R. (eds.), The Variability of Large Alluvial Rivers. American Society of Civil Engineers, New York, pp. 4574.Google Scholar
Winton, R. S., Calamita, E., and Wehrli, B. 2019. Reviews and syntheses: dams, water quality and tropical reservoir stratification. Biogeosciences 16, 16571671. doi: 10.5194/bg-16-1657.Google Scholar
Wintenberger, C., Rodrigues, S., Bréhéret, J.-G., and Villar, M. 2015. Fluvial islands: first stage of development from nonmigrating (forced) bars and woody-vegetation interactions. Geomorphology 246, 305320.Google Scholar
Wittfogel, K. 1957. Oriental Despotism: A Comparative Study of Total Power. Yale University Press, New Haven, CT.Google Scholar
Wohl, E. 2004. Disconnected Rivers: Linking Rivers to Landscapes. Yale University Press.Google Scholar
Wohl, E. 2007. Hydrology and discharge. In Gupta, A. (ed.), Large Rivers: Geomorphology and Management. John Wiley & Sons, Chichester, pp. 2944.Google Scholar
Wohl, E. 2014. A legacy of absence: wood removal in U.S. rivers. Progress in Physical Geography 38, 637663.Google Scholar
Wolman, G. M. 1959. Factors influencing the erosion of cohesive river banks. American Journal of Science 257, 204216.Google Scholar
Wolman, G. M., and Gerson, R. 1978. Relative scales of time and effectiveness of climate in watershed geomorphology. Earth Surface Processes and Landforms 3, 189208.Google Scholar
Wolman, G. M., and Giegengack, R. F. 2007. The Nile River: geology, hydrology, hydraulic society. In Gupta, A. (ed.), Large Rivers: Geomorphology and Management. John Wiley & Sons, Chichester, pp. 471490.Google Scholar
Wolman, G. M., and Leopold, L. B. 1957. River flood plains: some observations on their formation. U.S. Geological Survey Professional Paper 282-C.Google Scholar
Wolman, G. M., and Miller, G. P. 1960. Magnitude and frequency of forces in geomorphic process. Journal of Geology 68, 5474.Google Scholar
Wolman, M. G. 1967. A cycle of sedimentation and erosion in urban river channels. Geografiska Annaler 49A, 385395.Google Scholar
Wolters, H. A., Platteeuw, M., and Schoor, M. M. 2001. Guidelines for rehabilitation and management of floodplains; Ecology and Safety Combined. NCR Publication 09–2001; RIZA Report 2001.059.Google Scholar
Wolters, K. 2010. Multivariate ENSO Index (MEI). National Oceanic Atmospheric Administration.Google Scholar
Woodyer, K. D. 1968. Bankfull frequency in rivers. Journal of Hydrology 6, 114142.Google Scholar
World Bank. 2017. Implementing Nature-Based Flood Protection: Principles and Implementation Guidance. Washington, DC.Google Scholar
World Meteorological Association (WMO). 2004. Integrated Flood Management. The Associated Programme on Flood Management 1, Geneva, Switzerland.Google Scholar
Wright, L. D., and Coleman, J. M. 1972. River delta morphology, wave climate and the role of the subaqueous profile. Science 176, 282284.Google Scholar
Wright, L. D., and Coleman, J. M. 1973. Variations in morphology of major river deltas as functions of ocean waves and river discharge regimes. American Association of Petroleum Geologists Bulletin 57, 370398.Google Scholar
Wright, L. D., and Coleman, J. M. 1974. Mississippi River mouth processes: effluent dynamics and morphologic development. Journal of Geology 82, 751778.Google Scholar
Wu, C.-Y., and Mossa, J. 2019. Decadal-scale variations of thalweg morphology and riffle–pool sequences in response to flow regulation in the lowermost Mississippi River. Water 11, 1175. doi: 10.3390/w11061175.Google Scholar
Wu, Z., Milliman, J. D., Zhao, D., Cao, Z., Zhou, J., and Zhou, C. 2018. Geomorphologic changes in the lower Pearl River Delta, 1850–2015, largely due to human activity. Geomorphology 314, 4254.Google Scholar
Wüst, R. A. J., Bustin, R. M., and Lavkulich, L. M. 2003. New classification systems for tropical organic-rich deposits based on studies of the Tasek Bera Basin, Malaysia. Catena 53, 133153.Google Scholar
Xiquing, C. 1998. Changjian (Yangtze) River Delta, China. Journal of Coastal Research 14, 838858.Google Scholar
Xu, Y. J. 2006. Organic nitrogen retention in the Atchafalaya River Swamp. Hydrobiologia 560, 133143. doi: 10.1007/s10750-005-1171-8.Google Scholar
Xue, C. T. 1993. Historical changes in the Yellow River delta, China. Marine Geology 113, 321330.Google Scholar
Yang, C. T. 2006. Erosion and Sedimentation Manual. U.S. Department of the Interior Bureau of Reclamation. Technical Service Center, Sedimentation and River Hydraulics Group. Denver, CO.Google Scholar
Yang, S. L., Milliman, J. D., Li, P., and Xu, K. 2011. 50,000 dams later: erosion of the Yangtze River and its delta. Global and Planetary Change 75, 1420.Google Scholar
Yang, S. L., Xu, K. H., Milliman, J. D., Yang, H. F., and Wu, C. S. 2015. Decline of Yangtze River water and sediment discharge: impact from natural and anthropogenic changes. Scientific Reports 5, 12581. doi: 10.1038/srep12581.Google Scholar
Yia, Y., Yang, Z., and Zhang, S. 2010. Ecological influence of dam construction and river-lake connectivity on migration fish habitat in the Yangtze River basin, China. Procedia Environmental Sciences 2, 19421954.Google Scholar
Ylla Arbos, C., Schielen, R. M. J., and Blom, A. 2020. Bed level change in the Upper Rhine Delta between 1926–2018. NCR Days 2020, pp. 91–92, Nijmegen, Netherlands.Google Scholar
Yodis, E. G., and Kesel, R. H. 1992. The effects and implications of base-level changes to Mississippi River tributaries. Zeitschrift für Geomorphologie 37, 385402.Google Scholar
Yoshida, Y., Lee, H. S., Trung, B. H., et al. 2020. Impacts of mainstream hydropower dams on fisheries and agriculture in lower Mekong Basin. Sustainability 12, 2408. doi: 10.3390/su12062408.Google Scholar
Yoshiyama, R. M., Gerstung, E. R., Fisher, F. W., and Moyle, P. B. 2001. Historical and present distribution of Chinook Salmon in the Central Valley Drainage of California. Contributions to the Biology of Central Valley Salmonids. Fish Bulletin 179, Volume 1.Google Scholar
Young, W. C., Kent, D. H., and Whiteside, B. G. 1976. The influence of a deep storage reservoir on the species diversity of benthic macroinvertebrate communities on the Guadalupe River, Texas. The Texas Journal of Science 27, 213224.Google Scholar
Yuill, B. T., Lavoie, D., and Reed, D. J. 2009. Understanding subsidence processes in coastal Louisiana. Journal of Coastal Research: Special Issue 54, 2336. doi: 10.2112/SI54-012.1.Google Scholar
Yuill, B. T., Gaweesh, A., Allison, M. A., and Meselhe, E. A. 2016a. Morphodynamic evolution of a lower Mississippi River channel bar after sand mining. Earth Surface Processes and Landforms 41, 252262. doi: 10.1002/esp.3846.Google Scholar
Yuill, B. T., Khadka, A. K., Pereira, J., Allison, A., and Meselhe, E. A. 2016b. Morphodynamics of the erosional phase of crevasse-splay evolution and implications for river sediment diversion function. Geomorphology 259 1229.Google Scholar
Zarfl, C., Lumsdon, A. E., Berlekamp, J., Tydecks, L., and Tockner, K. 2015. A global boom in hydropower dam construction. Aquatic Sciences 77, 161170.Google Scholar
Zarfl, C., Berlekamp, J., He, F., et al. 2019. Future large hydropower dams impact global freshwater megafauna. Scientific Reports 9, 18531. doi: 10.1038/s41598-019-54980-8.Google Scholar
Zehetner, F., Lair, G. J., Maringer, F.-J., Gerzabek, M. J., and Hein, T. 2008. From sediment to soil: floodplain phosphorus transformations at the Danube River. Biogeochemistry 88, 117126. doi: 10.1007/s10533–008–9198-3.Google Scholar
Zevenbergen, C., Khan, S. A., van Alphen, J., van Scheltinga, C. T., and Veerbeek, W. 2018. Adaptive delta management: a comparison between the Netherlands Bangladesh Delta Program. International Journal of River Basin Management 16, 299306.Google Scholar
Zhang, J., Wen, W., Mao, H., and Shi, C. 1990. Huanghe (Yellow River) and its estuary: sediment origin, transport and deposition. Journal of Hydrology 120, 203223.Google Scholar
Zhang, W., Yuan, J., Han, J., Huang, C., and Li, M. 2016. Impact of the Three Gorges Dam on sediment deposition and erosion in the middle Yangtze River: a case study of the Shashi Reach. Hydrology Research. doi: 10.2166/nh.2016.092.Google Scholar
Zhang, W., Villarini, G., Vecchi, G. A., and Smith, J. A. 2018. Urbanization exacerbated the rainfall and flooding caused by hurricane Harvey in Houston. Nature 563(7731), 384. doi: 10.1038/s41586-018-0676-z.Google Scholar
Zhang, X., Dong, Z., Gupta, H., Wu, G., and Li, D. 2016. Impact of the Three Gorges Dam on the hydrology and ecology of the Yangtze River. Water 8, 590. doi: 10.3390/w8120590.Google Scholar
Zheng, S., Wu, B., Wang, K., Tan, G., Han, S., and Thorne, C. R. 2017. Evolution of the Yellow River delta, China: impacts of channel avulsion and progradation. International Journal of Sediment Research. doi: 10.1016/j.ijsrc.2016.10.001.Google Scholar
Zheng, S., Xu, Y. J., Cheng, H., Wang, B., Xu, W., and Wu, S. 2018. Riverbed erosion of the final 565 kilometers of the Yangtze River (Changjiang) following construction of the Three Gorges Dam. Scientific Reports 8, 11917. doi: 10.1038/s41598-018-30441.Google Scholar
Zhong, Y., and Power, G. 2015. Environmental impacts of hydroelectric projects on fish resources. Regulated Rivers: Research and Management 12, 8198.Google Scholar
Zhu, Y., and Newell, R. E. 1998. A proposed algorithm for moisture fluxes from atmospheric rivers. Monthly Weather Review 126, 725735.Google Scholar
Zinger, J. A., Rhoads, B. L., and Best, J. L. 2011. Extreme sediment pulses generated by bend cutoffs along a large meandering river. Nature Geoscience 4, 675678.Google Scholar
Ziv, G., Baran, E., Nam, S., Rodríguez-Iturbe, I., and Levin, S. A. 2012. Trading-off fish biodiversity, food security, and hydropower in the Mekong River Basin. Proceedings of the National Academy of Sciences 109, 56095614.Google Scholar
Zorn, M. R., and Waylen, P. R. 1997. Seasonal response of mean monthly streamflow to El Niño/Southern Oscillation in North Central Florida. Professional Geographer 49, 5156.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.

  • Bibliography
  • Paul F. Hudson, Universiteit Leiden
  • Book: Flooding and Management of Large Fluvial Lowlands
  • Online publication: 04 November 2021
  • Chapter DOI: https://doi.org/10.1017/9781139015738.011
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.

  • Bibliography
  • Paul F. Hudson, Universiteit Leiden
  • Book: Flooding and Management of Large Fluvial Lowlands
  • Online publication: 04 November 2021
  • Chapter DOI: https://doi.org/10.1017/9781139015738.011
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.

  • Bibliography
  • Paul F. Hudson, Universiteit Leiden
  • Book: Flooding and Management of Large Fluvial Lowlands
  • Online publication: 04 November 2021
  • Chapter DOI: https://doi.org/10.1017/9781139015738.011
Available formats
×