Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-zzh7m Total loading time: 0 Render date: 2024-04-27T11:09:56.064Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  20 October 2016

Andrew S. Goudie
Affiliation:
University of Oxford
Heather A. Viles
Affiliation:
University of Oxford
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
Publisher: Cambridge University Press
Print publication year: 2016

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

Aagaard, T. and Sorensen, P., (2012). Coastal profile response to sea level rise: a process-based approach. Earth Surface Processes and Landforms, 37, 354–62.CrossRefGoogle Scholar
Aalto, J., Venäläinen, A., Heikkinen, R. K., and Luoto, M. (2014). Potential for extreme loss in high‐latitude Earth surface processes due to climate change. Geophysical Research Letters, 41, 3914–24.CrossRefGoogle Scholar
Abidin, H. Z., Andreas, H., Gumilar, I., and Wibowo, I. R. R. (2015). On correlation between urban development, land subsidence and flooding phenomena in Jakarta. Proceedings IAHS, 370, 1520.CrossRefGoogle Scholar
Abrams, M. D. and Nowacki, G. J. (2015). Exploring the early Anthropocene burning hypothesis and climate-fire anomalies for the Eastern US. Journal of Sustainable Forestry, 34, 3048.CrossRefGoogle Scholar
Abudanh, F. and Twaissi, S. (2010). Innovation or technology immigration? The Qanat systems in the regions of Udhruh and Maʾan in Southern Jordan. Bulletin of the American Schools of Oriental Research, 360, 6787.CrossRefGoogle Scholar
Adam, J. C., Hamlet, A. F., and Lettenmaier, D. P. (2009). Implications of global climate change for snowmelt hydrology in the twenty-first century. Hydrological Processes, 23, 962–72.CrossRefGoogle Scholar
Adams, H. D., Luce, C. H., Breshears, D. D., Allen, C. D., Weiler, M., Hale, V. C., Smith, A. M. S., and Huxman, T. E. (2012). Ecohydrological consequences of drought‐and infestation‐triggered tree die‐off: insights and hypotheses. Ecohydrology, 5, 145–59.CrossRefGoogle Scholar
Adams, W. M., Perrow, M. R., and Carpenter, A. (2004). Conservatives and champions: river managers and the river restoration discourse in the United Kingdom. Environment and Planning A, 36, 1929–42.CrossRefGoogle Scholar
Aeschbach-Hertig, W. and Gleeson, T. (2012). Regional strategies for the accelerating global problem of groundwater depletion. Nature Geoscience, 5, 853–61.CrossRefGoogle Scholar
Aikens, C. M. and Lee, G. A. (2014). Postglacial inception and growth of anthropogenic landscapes in China, Korea, Japan, and the Russian Far East. Anthropocene, 4, 4656.CrossRefGoogle Scholar
Akiner, S., Cooke, R. U., and French, R. A. (1992). Salt damage to Islamic monuments in Uzbekistan. Geographical Journal, 158, 257–72.CrossRefGoogle Scholar
Albert, R. M. (2015). Anthropocene and early human behavior. The Holocene, 25, 1542–52 doi: 10.1177/0959683615588377.CrossRefGoogle Scholar
Alcántara-Ayala, I., Esteban-Chávez, O., and Parrot, J. F. (2006). Landsliding related to land-cover change: a diachronic analysis of hillslope instability distribution in the Sierra Norte, Puebla, Mexico. Catena, 65, 152–65.CrossRefGoogle Scholar
Alemayehu, T., Furi, W., and Legesse, D. (2007). Impact of water overexploitation on highland lakes of Eastern Ethiopia. Environmental Geology, 52, 147–54.CrossRefGoogle Scholar
Algeet-Abarquero, N., Marchamalo, M., Bonatti, J., Fernández-Moya, J., and Moussa, R. (2015). Implications of land use change on runoff generation at the plot scale in the humid tropics of Costa Rica. Catena, 135, 263–70.CrossRefGoogle Scholar
Al-Harthi, A. A. and Bankher, K. A. (1999). Collapsing loess-like soil in western Saudi Arabia. Journal of Arid Environments, 41, 383–9.CrossRefGoogle Scholar
Allchin, B. (1976). Palaeolithic sites in the plains of Sind and their geographical implications. Geographical Journal, 142, 471–89.CrossRefGoogle Scholar
Allen, B. (1971). Wet-field taro terraces on Mangaia, Cook Islands. The Journal of the Polynesian Society, 80, 371–8.Google Scholar
Allen, S. K., Cox, S. C., and Owens, I. F. (2011). Rock avalanches and other landslides in the central Southern Alps of New Zealand: a regional study considering possible climate change impacts. Landslides, 8, 3348.CrossRefGoogle Scholar
Allis, R. G. (2000). Review of subsidence at Wairakei field, New Zealand. Geothermics, 29, 455–78.Google Scholar
Allis, R. G., Bromley, C., and Currie, S. (2009). Update on subsidence at the Wairakei–Tauhara geothermal system, New Zealand. Geothermics, 38, 169–80.CrossRefGoogle Scholar
Ambers, R. K., Druckenbrod, D. L., and Ambers, C. P. (2006). Geomorphic response to historical agriculture at Monument Hill in the Blue Ridge foothills of central Virginia. Catena, 65, 4960.CrossRefGoogle Scholar
Anderson, A. (2009). The rat and the octopus: initial human colonization and the prehistoric introduction of domestic animals to remote Oceania. Biological Invasions, 11, 1503–19.CrossRefGoogle Scholar
Anderson, D. (1984). Depression, dust bowl, demography, and drought: the colonial state and soil conservation in East Africa during the 1930s. African Affairs, 83, 321–43.CrossRefGoogle Scholar
Anderson, J. B., Wallace, D. J., Simms, A. R., Rodriguez, A. B., and Milliken, K. T. (2014). Variable response of coastal environments of the Northwestern Gulf of Mexico to sea-level rise and climate change: implications for future change. Marine Geology, 352, 348–66.CrossRefGoogle Scholar
Andersson, A. J. and Gledhill, D. (2013). Ocean acidification and coral reefs: effects on reakdown, dissolution, and net ecosystem calcification. Annual Review of Marine Science, 5, 321–48.CrossRefGoogle ScholarPubMed
André, M. F., Vautier, F., Voldoire, O., and Roussel, E. (2014). Accelerated stone deterioration induced by forest clearance around the Angkor temples. Science of the Total Environment, 493, 98108.CrossRefGoogle ScholarPubMed
Andréassian, V. (2004). Waters and forests: from historical controversy to scientific debate. Journal of Hydrology, 291, 127.CrossRefGoogle Scholar
Angel, J. R. and Kunkel, K. E. (2010). The response of Great Lakes water levels to future climate scenarios with an emphasis on Lake Michigan-Huron. Journal of Great Lakes Research, 36, 51–8.CrossRefGoogle Scholar
Anthony, E. J. (2014). The human influence on the Mediterranean coast over the last 200 years: a brief appraisal from a geomorphological perspective. Géomorphologie: Relief, Processus, Environnement, 3, 219–26.Google Scholar
Anthony, E. J. (2015). Wave influence in the construction, shaping and destruction of river deltas: A review. Marine Geology, 361, 5378.CrossRefGoogle Scholar
Anthony, K., Maynard, J. A., Diaz‐Pulido, G., Mumby, P. J., Marshall, P. A., Cao, L., and Hoegh‐Guldberg, O. V. E. (2011). Ocean acidification and warming will lower coral reef resilience. Global Change Biology, 17, 1798–808.CrossRefGoogle ScholarPubMed
Anthony, K. R. N., Kline, D. I., Diaz-Pulido, G., Dove, S., and Hoegh-Guldberg, O. (2008). Ocean acidification causes bleaching and productivity loss in coral reef builders. Proceedings of the National Academy of Sciences, 105, 17442–6.CrossRefGoogle Scholar
April, R., Newton, R., and Coles, L. T. (1986). Chemical weathering in two Adirondack watersheds: past and present-day rates. Geological Society of America Bulletin, 97, 1232–8.2.0.CO;2>CrossRefGoogle Scholar
Arendt, A. A., Echelmeyer, K. A., Harriso, W. D., Kingle, C. S., and Valentine, V. B. (2002) Rapid wastage of Alaska glaciers and their contribution to rising sea level. Science, 297, 382–5.CrossRefGoogle ScholarPubMed
Arens, S. M., Mulder, J. P., Slings, Q. L., Geelen, L. H., and Damsma, P. (2013). Dynamic dune management, integrating objectives of nature development and coastal safety: examples from the Netherlands. Geomorphology, 199, 205–13.CrossRefGoogle Scholar
Arenson, L. U. and Jakob, M. (2015). Periglacial geohazard risks and ground temperature increases. In Engineering Geology for Society and Territory: Vol. 1, eds. Lollino, G., Manconi, A., Clague, J., Shan, W., and Chiarle, M.. Springer International Publishing, pp. 233–7.Google Scholar
Armitage, A. R., Highfield, W. E., Brody, S. D., and Louchouarn, P. (2014). The contribution of mangrove expansion to salt marsh loss on the Texas Gulf Coast. PloS one, 10(5), e0125404.CrossRefGoogle Scholar
Arnáez, J., Lana-Renault, N., Lasanta, T., Ruiz-Flaño, P., and Castroviejo, J. (2015). Effects of farming terraces on hydrological and geomorphological processes. A review. Catena, 128, 122–34.CrossRefGoogle Scholar
Arnáez, J., Larrea, V., and Ortigosa, L. (2004). Surface runoff and soil erosion on unpaved forest roads from rainfall simulation tests in northeastern Spain. Catena, 57, 114.Google Scholar
Arnaud, F., Révillon, S., Debret, M., Revel, M., Chapron, E., Jacob, J., Giguet-Covexa, C., Poulenarda, J., and Magny, M. (2012). Lake Bourget regional erosion patterns reconstruction reveals Holocene NW European Alps soil evolution and paleohydrology. Quaternary Science Reviews, 51, 8192.CrossRefGoogle Scholar
Arnell, N. W. (2002). Hydrology and Global Environmental Change. Harlow: Prentice Hall.Google Scholar
Arnell, N. W. and Gosling, S. N. (2013). The impacts of climate change on river flow regimes at the global scale. Journal of Hydrology, 486, 351–64.CrossRefGoogle Scholar
Arora, V. K. and Boer, G. J. (2001). Effects of simulated climate change on the hydrology of major river basins. Journal of Geophysical Research, 106, D4, 3335–48.CrossRefGoogle Scholar
Ashkenazy, Y., Yizhaq, H., and Tsoar, H. (2012). Sand dune mobility under climate change in the Kalahari and Australian deserts. Climatic Change, 112, 901–23.CrossRefGoogle Scholar
Ashton, A. D., Donnelly, J. P., and Evans, R. L. (2008). A discussion of the potential impacts of climate change on the shorelines of the northeastern USA. Mitigation and Adaptive Strategies for Global Change, 13, 719–43.Google Scholar
Ashton, A. D., Walkden, M. J. A., and Dickson, M. E. (2011). Equilibrium responses of cliffed coasts to changes in the rate of sea level rise. Marine Geology, 284, 217–29.CrossRefGoogle Scholar
Ateweberhan, M., Feary, D. A., Keshavmurthy, S., Chen, A., Schleyer, M. H., and Sheppard, C. R. (2013). Climate change impacts on coral reefs: Synergies with local effects, possibilities for acclimation, and management implications. Marine Pollution Bulletin, 74, 526–39.CrossRefGoogle ScholarPubMed
Atkinson, G., Assatourians, K., Cheadle, B., and Greig, W. (2015). Ground motions from three recent earthquakes in Western Alberta and Northeastern British Columbia and their implications for induced‐seismicity hazard in Eastern Regions. Seismological Research Letters, 86, 1022–31.CrossRefGoogle Scholar
Aubault, H., Webb, N. P., Strong, C. L., McTainsh, G. H., Leys, J. F., and Scanlan, J. C. (2015). Grazing impacts on the susceptibility of rangelands to wind erosion: the effects of stocking rate, stocking strategy and land condition. Aeolian Research, 17, 8999.CrossRefGoogle Scholar
Aukema, J. E., McCullough, D. G., Von Holle, B., Liebhold, A. M., Britton, K., and Frankel, S. J. (2010). Historical accumulation of nonindigenous forest pests in the continental United States. BioScience, 60, 886–97.CrossRefGoogle Scholar
Avila, F. B., Pitman, A. J., Donat, M. G., Alexander, L. V. and Abramowitz, G. (2012). Climate model simulated changes in temperature extremes due to land cover change. Journal of Geophysical Research, 117, D04108, doi: 10.1029/2011JD016382.CrossRefGoogle Scholar
Ayenew, T. and Legesse, D. (2007). The changing face of the Ethiopian rift lakes and their environs: call of the time. Lakes & Reservoirs: Research and Management, 12, 149–65.CrossRefGoogle Scholar
Azorin-Molina, C., Vicente-Serrano, S. M., McVicar, T. R., Jerez, S., Sanchez-Lorenzo, A., López-Moreno, J., Revuelto, J., Trigo, R. M., Lopez-Bustins, J. A., and Espírito-Santo, F. (2014). Homogenization and assessment of observed near-surface wind speed trends over Spain and Portugal, 1961–2011. Journal of Climate, 27, 3692–712.CrossRefGoogle Scholar
Baartman, J. E., Temme, A. J., Schoorl, J. M., Braakhekke, M. H., and Veldkamp, T. (2012). Did tillage erosion play a role in millennial scale landscape development? Earth Surface Processes and Landforms, 37, 1615–26.CrossRefGoogle Scholar
Baccari, N., Boussema, M. R., Lamachère, J. M., and Nasri, S. (2008). Efficiency of contour benches, filling-in and silting-up of a hillside reservoir in a semi-arid climate in Tunisia. C. R. Geoscience, 340, 3848.CrossRefGoogle Scholar
Badescu, V. and Cathcart, R. B. (2011). Aral Sea partial restoration.1. A Caspian importation macroproject. International Journal of Environment and Waste Water Management, 7, 161–82.Google Scholar
Badman, T. (2010). World Heritage and geomorphology. In Geomorphological Landscapes of the World, ed. Migoń, P.. Dordrecht: Springer, pp. 357–68.Google Scholar
Bailey, D. W., Tringham, R., Bass, J., Stevanović, M., Hamilton, M., Neumann, H., and Raduncheva, A. (1998). Expanding the dimensions of early agricultural tells: the Podgoritsa Archaeological Project, Bulgaria. Journal of Field Archaeology, 25, 373–96.Google Scholar
Bain, D. J., Green, M. B., Campbell, J. L., Chamblee, J. F., Chaoka, S., Fraterrigo, J. M., and Leigh, D. S. (2012). Legacy effects in material flux: structural catchment changes predate long-term studies. Bioscience, 62, 575–84.CrossRefGoogle Scholar
Baine, M. (2001). Artificial reefs: a review of their design, application, management and performance. Ocean & Coastal Management, 44, 241–59.CrossRefGoogle Scholar
Baisch, S., Vörös, R., Rothert, E., Stang, H., Jung, R., and Schellschmidt, R. (2010). A numerical model for fluid injection induced seismicity at Soultz-sous-Forêts. International Journal of Rock Mechanics and Mining Sciences, 47, 405–13.Google Scholar
Baker, A. and Simms, M. J. (1998). Active deposition of calcareous tufa in Wessex, UK, and its implications for the “late-Holocene tufa decline.” The Holocene, 8, 359365.CrossRefGoogle Scholar
Baker, A. C., Glynn, P. W., and Riegl, B. (2008). Climate change and coral reef bleaching: an ecological assessment of long-term impacts, recovery trends and future outlook. Estuarine, Coastal and Shelf Science, 8, 435–71.Google Scholar
Baker, V. R. (2014). Uniformitarianism, earth system science, and geology. Anthropocene, 5, 76–9.CrossRefGoogle Scholar
Bakker, M. M., Govers, G., van Doorn, A., Quetier, F., Chouvardas, D., and Rounsevell, M. (2008). The response of soil erosion and sediment export to land-use change in four areas of Europe: the importance of landscape pattern. Geomorphology, 98, 213–26.CrossRefGoogle Scholar
Bakoariniaina, L. N., Kusky, T., and Raharimahefa, T. (2006). Disappearing Lake Alaotra: monitoring catastrophic erosion, waterway silting, and land degradation hazards in Madagascar using Landsat imagery. Journal of African Earth Sciences, 44, 241–52.CrossRefGoogle Scholar
Balcerak, E. (2011). Thermokarst lakes expand and drain laterally as permafrost degrades. Eos, Transactions American Geophysical Union, 92(45), 408.Google Scholar
Balke, T. and Friess, D. A. (2015). Geomorphic knowledge for mangrove restoration: a pan‐tropical categorization. Earth Surface Processes and Landforms. doi: 10.1002/esp.3841.CrossRefGoogle Scholar
Balling, R. C. and Wells, S. G. (1990). Historical rainfall patterns and arroyo activity within the Zuni river drainage basin, New Mexico. Annals of the Association of American Geographers, 80, 603–17.CrossRefGoogle Scholar
Balme, J. (2013). On boats and string: the maritime colonisation of Australia. Quaternary International 285, 6875.CrossRefGoogle Scholar
Balter, M. (2010). The tangled roots of agriculture. Science, 327, 404–6.CrossRefGoogle ScholarPubMed
Balter, M. (2013). Farming’s tangled European roots. Science, 342, 181–82.CrossRefGoogle ScholarPubMed
Balthazar, V., Vanacker, V., Girma, A., Poesen, J., and Golla, S. (2013). Human impact on sediment fluxes within the Blue Nile and Atbara River basins. Geomorphology, 180, 231–41.Google Scholar
Bamber, J. L. and Aspinall, W. P. (2013). An expert judgement assessment of future sea level rise from the ice sheets. Nature Climate Change, 3, 424–7.CrossRefGoogle Scholar
Banks, V. J., Jones, P. F., Lowe, D. J., Lee, J. R., Rushton, J., and Ellis, M. A. (2012). Review of tufa deposition and palaeohydrological conditions in the White Peak, Derbyshire, UK: implications for Quaternary landscape evolution. Proceedings of the Geologists’ Association, 123, 117–29.Google Scholar
Barca, D., Comite, V., Belfiore, C. M., Bonazza, A., La Russa, M. F., Ruffolo, S. A., and Sabbioni, C. (2014). Impact of air pollution in deterioration of carbonate building materials in Italian urban environments. Applied Geochemistry, 48, 122–31.CrossRefGoogle Scholar
Barker, G. (2006). The Agricultural Revolution in Prehistory; Why Did Foragers Become Farmers? Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Barker, G. W. and Hunt, C. O. (1995). Quaternary valley floor erosion and aluviation in the Biferno Valley, Molise, Italy: the role of tectonics, climate, sea level change, and human activity. In Mediterranean Quaternary River Environments, ed. Lewin, J., Macklin, M. G. and Woodward, J. C., pp. 145–57.Google Scholar
Barnard, P. L., Owen, L. A., Sharma, M. C., and Finkel, R. C. (2001). Natural and human-induced landsliding in the Garhwal Himalaya of northern India. Geomorphology, 40, 2135.CrossRefGoogle Scholar
Barry, R. G. (1985). The cryosphere and climate change. In Detecting the Climatic Effects of Increasing Carbon Dioxide, ed. MacCracken, M. C. and Luther, F. M., Washington, DC: US Deptarment of Energy, pp. 111–48.Google Scholar
Bartley, R., Bainbridge, Z. T., Lewis, S. E., Kroon, F. J., Wilkinson, S. N., Brodie, J. E., and Silburn, D. M. (2014). Relating sediment impacts on coral reefs to watershed sources, processes and management: a review. Science of the Total Environment, 468, 1138–53.Google ScholarPubMed
Bar-Yosef, O. (1998). On the nature of transitions: the Middle to Upper Palaeolithic and the Neolithic Revolution. Cambridge Archaeological Journal, 8, 141–63.CrossRefGoogle Scholar
Basagic, H. J. and Fountain, A. G. (2011). Quantifying 20th century glacier change in the Sierra Nevada, California. Arctic, Antarctic and Alpine Research, 43, 317–30.CrossRefGoogle Scholar
Battany, M. C. and Grismer, M. E. (2000). Rainfall runoff and erosion in Napa Valley vineyards: effects of slope, cover and surface roughness. Hydrological Processes, 14, 1289–304.3.0.CO;2-R>CrossRefGoogle Scholar
Battarbee, R. W., Appleby, P. G., Odel, K., and Flower, R. J. (1985). 210Pb dating of Scottish Lake sediments, afforestation and accelerated soil erosion. Earth Surface Processes and Landforms, 10, 137–42.CrossRefGoogle Scholar
Baú, D., Gambolati, G., and Teatini, P. (2000). Residual land subsidence near abandoned gas fields raises concern over Northern Adriatic coastland. Eos, Transactions American Geophysical Union, 81(22), 245–9.CrossRefGoogle Scholar
Bauer, J. E., Cai, W. J., Raymond, P. A., Bianchi, T. S., Hopkinson, C. S., and Regnier, P. A. (2013). The changing carbon cycle of the coastal ocean. Nature, 504, 6170.CrossRefGoogle ScholarPubMed
Bautista, S., Bellot, J., and Ramón Vallejo, V. (1996). Mulching treatment for postfire soil conservation in a semi-arid ecosystem. Arid Soil Research and Rehabilitation, 10, 235–42.CrossRefGoogle Scholar
Bayliss, A., McAvoy, F., and Whittle, A. (2007). The world recreated: redating Silbury Hill in its monumental landscape. Antiquity, 81, 26.CrossRefGoogle Scholar
Bayon, G., Dennielou, B., Etoubleau, J., Ponzevera, E., Toucanne, S., and Bermell, S. (2012). Intensifying weathering and land use in iron age Central Africa. Science, 335, 1219–22.CrossRefGoogle ScholarPubMed
Beach, T., Dunning, N., Luzzadder-Beach, S., Cook, D. E., and Lohse, J. (2006). Impacts of the ancient Maya on soils and soil erosion in the central Maya Lowlands. Catena, 65, 166–78.CrossRefGoogle Scholar
Beach, T., Luzzadder-Beach, S., Cook, D., Dunning, N., Kennett, D. J., Krause, S., Terry, R., Trein, D., and Valdez, F. (2015). Ancient Maya impacts on the Earth’s surface: an early Anthropocene analog? Quaternary Science Reviews, 124, 130.CrossRefGoogle Scholar
Beaulieu, E., Goddéris, Y., Donnadieu, Y., Labat, D., and Roelandt, C. (2012). High sensitivity of the continental-weathering carbon dioxide sink to future climate change. Nature Climate Change, 2, 346–9.CrossRefGoogle Scholar
Beaumont, P. B. (2011). The edge: more on fire-making by about 1.7 million years ago at Wonderwerk Cave in South Africa. Current Anthropology, 52, 585–94.CrossRefGoogle Scholar
Beck, M. W., Brumbaugh, R. D., Airoldi, L., Carranza, A., Coen, L. D., Crawford, C., Defeo, O., Edgar, G. J., Hancock, B., Kay, M. C., and Guo, X. (2011). Oyster reefs at risk and recommendations for conservation, restoration, and management. Bioscience, 61, 107–16.CrossRefGoogle Scholar
Beckers, B., Berking, J., and Schütt, B. (2013). Ancient water harvesting methods in the drylands of the Mediterranean and Western Asia. Journal for Ancient Studies. 2, 145–64.Google Scholar
Beerten, K., Vandersmissen, N., Deforce, K., and Vandenberghe, N. (2014). Late Quaternary (15 ka to present) development of a sandy landscape in the Mol area, Campine region, north‐east Belgium. Journal of Quaternary Science, 29, 433–44.CrossRefGoogle Scholar
Bégin, C., Brooks, G., Larson, R. A., Dragićević, S., Ramos Scharrón, C. E., and Côté, I. M. (2014). Increased sediment loads over coral reefs in Saint Lucia in relation to land use change in contributing watersheds. Ocean & Coastal Management, 95, 3545.CrossRefGoogle Scholar
Beinart, W. (1984). Soil erosion, conservationism and ideas about development: a southern African exploration, 1900–1960. Journal of Southern African Studies, 11, 5283.CrossRefGoogle Scholar
Bekendam, R. F. and Pöttgens, J. J. (1995). Ground movements over the coal mines of southern Limburg, The Netherlands, and their relation to rising mine waters. Land Subsidence. IAHS Publication, 234, 312.Google Scholar
Bell, F. G., Stacey, T. R., and Genske, D. D. (2000). Mining subsidence and its effect on the environment: some differing examples. Environmental Geology, 40, 135–52.CrossRefGoogle Scholar
Bell, F. G., Donnelly, L. J., Genske, D. D., and Ojeda, J. (2005). Unusual cases of mining subsidence from Great Britain, Germany and Colombia. Environmental Geology, 47, 620–31.CrossRefGoogle Scholar
Bell, M. L. (1982). The effect of land-use and climate on valley sedimentation. In Climatic Change in Later Prehistory, ed. Harding, A. F.. Edinburgh: Edinburgh University Press, pp. 127–42.Google Scholar
Bellard, C., Leclerc, C., and Courchamp, F. (2013). Potential impact of sea level rise on French islands worldwide. Nature Conservation, 5, 7586.CrossRefGoogle Scholar
Bellin, N., van Wesemael, B., Meerkerk, A., Vanacker, V., and Barbera, G. G. (2009). Abandonment of soil and water conservation structures in Mediterranean ecosystems: a case study from south east Spain. Catena, 76, 114–21.CrossRefGoogle Scholar
Belnap, J. and Gillette, D. A. (1997). Disturbance of biological soil crusts: impacts on potential wind erodibility of sandy desert soils in southeastern Utah. Land Degradation & Development, 8, 355–62.3.0.CO;2-H>CrossRefGoogle Scholar
Belski, A. J. (1996). Viewpoint: western juniper expansion: is it a threat to arid northwest ecosystems? Journal of Range Management, 49, 53–9.Google Scholar
Benda, L., Miller, D., Bigelow, P., and Andras, K. (2003). Effects of post-wildfire erosion on channel environments, Boise River, Idaho. Forest Ecology and Management, 178, 105–19.CrossRefGoogle Scholar
Benito, G., Del Campo, P. P., Gutiérrez-Elorza, M., and Sancho, C. (1995). Natural and human-induced sinkholes in gypsum terrain and associated environmental problems in NE Spain. Environmental Geology, 25, 156–64.CrossRefGoogle Scholar
Benn, D. I., Bolch, T., Hands, K., Gulley, J., Luckman, A., Nicholson, L. I,, Quincey, D., Thompson, S., Toumi, R., and Wiseman, S. (2012). Response of debris-covered glaciers in the Mount Everest region to recent warming, and implications for outburst flood hazards. Earth-Science Reviews, 114, 156–74.CrossRefGoogle Scholar
Bennett, H. H. (1938). Soil Conservation. New York: McGraw-Hill.Google Scholar
Bennett, M. R. (2003). Ice streams as the arteries of an ice sheet: their mechanics, stability and significance. Earth-Science Reviews, 61, 309–39.CrossRefGoogle Scholar
Berg, R. D. and Carter, D. L. (1980). Furrow erosion and sediment losses on irrigated cropland. Journal of Soil and Water Conservation, 35, 267–70.Google Scholar
Berhe, A. A., Harte, J., Harden, J. W., and Torn, M. S. (2007). The significance of the erosion-induced terrestrial carbon sink. BioScience, 57, 337–46.CrossRefGoogle Scholar
Bern, C. R., Clark, M. L., Schmidt, T. S., Holloway, J. M., and McDougal, R. R. (2015). Soil disturbance as a driver of increased stream salinity in a semiarid watershed undergoing energy development. Journal of Hydrology, 524, 123–36.CrossRefGoogle Scholar
Berna, F., Goldberg, P., Horwitz, L. K., Brink, J., Holt, S., Bamford, M., and Chazan, M. (2012). Microstratigraphic evidence of in situ fire in the Acheulean strata of Wonderwerk Cave, Northern Cape Province, South Africa. Proceedings of the National Academy of Sciences, 109, E1215–20.CrossRefGoogle ScholarPubMed
Berner, R. A. (1998). The carbon cycle and CO2 over Phanerozoic time: the role of land plants. Philosophical Transactions of the Royal Society B, 353, 7582.CrossRefGoogle Scholar
Bertness, M. D. and Coverdale, T. C. (2013). An invasive species facilitates the recovery of salt marsh ecosystems on Cape Cod. Ecology, 94, 1937–43.CrossRefGoogle ScholarPubMed
Bertness, M. D. and Silliman, B. R. (2008). Consumer control of salt marshes driven by human disturbance. Conservation Biology, 22, 618–23.CrossRefGoogle ScholarPubMed
Bertran, P. (2004). Soil erosion in small catchments of the Quercy region (southwestern France) during the Holocene. The Holocene, 14, 597606.CrossRefGoogle Scholar
Beschta, R. L. (1978). Long‐term patterns of sediment production following road construction and logging in the Oregon Coast Range. Water Resources Research, 14, 1011–6.CrossRefGoogle Scholar
Bhagwat, S., Kettle, C. J., and Koh, L. P. (2014). The history of deforestation and forest fragmentation: a global perspective. In Global Forest Fragmentation, ed. Kettle, C. J. and Koh, L. P.. Wallingford: CABI, pp. 519.CrossRefGoogle Scholar
Bhark, E. W. and Small, E. E. (2003). Association between plant canopies and the spatial patterns of infiltration in shrubland and grassland of the Chihuahuan Desert, New Mexico. Ecosystems, 6, 185–96.CrossRefGoogle Scholar
Bhattachan, A., D’Odorico, P., Okin, G. S., and Dintwe, K. (2013). Potential dust emissions from the southern Kalahari’s dunelands. Journal of Geophysical Research: Earth Surface, 118, 307–14.Google Scholar
Bhattachan, A., D’Odorico, P., Dintwe, K., Okin, G. S., and Collins, S. L. (2014). Resilience and recovery potential of duneland vegetation in the southern Kalahari. Ecosphere, 5(1), art2.CrossRefGoogle Scholar
Biemans, H., Haddeland, I., Kabat, P., Ludwig, F., Hutjes, R. W. A., Heinke, J., and Gerten, D. (2011). Impact of reservoirs on river discharge and irrigation water supply during the 20th century. Water Resources Research, 47, doi: 10.1029/2009WR008929.CrossRefGoogle Scholar
Bigelow, G. F., Ferrante, S. M., Hall, S. T., Kimball, L. M., Proctor, R. E., and Remington, S. L. (2005). Researching catastrophic environmental changes on northern coastlines: a geoarchaeological case study from the Shetland Islands. Arctic Anthropology, 42, 88102.CrossRefGoogle Scholar
Binford, M. W., Brenner, M., Whitmore, T. J., Higuera-Grundy, A., Deevey, E. S., and Leyden, B. (1987). Ecosystems, palaeoecology, and human disturbance in subtropical and tropical America. Quaternary Science Review, 6, 115–28.CrossRefGoogle Scholar
Bini, C., Gemignani, S., and Zilocchi, L. (2006). Effect of different land use on soil erosion in the pre-alpine fringe (North-East Italy): Ion budget and sediment yield. Science of the Total Environment, 369, 433–40.CrossRefGoogle ScholarPubMed
Bird, E. C. F. (1979). Coastal processes. In Man and Environmental Processes, ed. Gregory, K. J. and Walling, D. E.. Folkestone: Dawson, pp. 82101.Google Scholar
Bird, E. C. F. (1993). Submerging Coasts. Chichester: Wiley.Google Scholar
Bird, E. C. F. (1996). Beach Management. Chichester: Wiley.Google Scholar
Bird, E. C. F. and Lewis, N. (2015). Causes of beach erosion. In Beach Renourishment ed. Bird, E. and Lewis, N.. Springer International Publishing. pp. 728.CrossRefGoogle Scholar
Birkeland, C. (ed.) (2015). Coral Reefs in the Anthropocene. Dordrecht: Springer.CrossRefGoogle Scholar
Birken, A. S. and Cooper, D. J. (2006). Processes of Tamarix invasion and floodplain development along the lower Green River, Utah. Ecological Applications, 16, 1103–20.CrossRefGoogle ScholarPubMed
Birkinshaw, S. J., Bathurst, J. C., and Robinson, M. (2014). 45 years of non-stationary hydrology over a forest plantation growth cycle, Coalburn catchment, Northern England. Journal of Hydrology, 519, 559–73.CrossRefGoogle Scholar
Bischoff, M., Cete, A., Fritschen, R., and Meier, T. (2010). Coal mining induced seismicity in the Ruhr area, Germany. Pure and Applied Geophysics, 167, 6375.CrossRefGoogle Scholar
Bishop, P. and Muñoz-Salinas, E. (2013). Tectonics, geomorphology and water mill location in Scotland, and the potential impacts of mill dam failure. Applied Geography, 42, 195205.CrossRefGoogle Scholar
Black, K. and Mead, S. (2009). Design of surfing reefs. Reef Journal, 1, 177–91.Google Scholar
Blais-Stevens, A., Kremer, M., Bonnaventure, P. P., Smith, S. L., Lipovsky, P., and Lewkowicz, A. G. (2015). Active layer detachment slides and retrogressive thaw slumps susceptibility mapping for current and future permafrost distribution, Yukon Alaska Highway Corridor. In Engineering Geology for Society and Territory-Volume 1 (pp. 449–53), ed. Lollino, G., Manconi, A., Clague, J., Shan, W., and Chiarle, M., Springer International Publishing, pp. 449–53.Google Scholar
Blanco, P. D., Rostagno, C. M., del Valle, H. F., Beeskow, A. M., and Wiegand, T. (2008). Grazing impacts in vegetated dune fields: predictions from spatial pattern analysis. Rangeland Ecology and Management, 61, 194203.CrossRefGoogle Scholar
Blankespoor, B., Dasgupta, S., and Laplante, B. (2014). Sea-level rise and coastal wetlands. Ambio, 43, 9961005.CrossRefGoogle ScholarPubMed
Blanton, P. and Marcus, W. A. (2013). Transportation infrastructure, river confinement, and impacts on floodplain and channel habitat, Yakima and Chehalis rivers, Washington, USA. Geomorphology, 189, 5565.CrossRefGoogle Scholar
Blavet, D., De Noni, G., Le Bissonnais, Y., Leonard, M., Maillo, L., Laurent, J. Y., and Roose, E. (2009). Effect of land use and management on the early stages of soil water erosion in French Mediterranean vineyards. Soil and Tillage Research, 106, 124–36.CrossRefGoogle Scholar
Bliss, A., Hock, R., and Radić, V. (2014). Global response of glacier runoff to twenty‐first century climate change. Journal of Geophysical Research: Earth Surface, 119, 717–30.Google Scholar
Blott, S. J., Pye, K., Van der Wal, D., and Neal, A. (2006). Long-term morphological change and its causes in the Mersey Estuary, NW England. Geomorphology, 81, 185206.CrossRefGoogle Scholar
Boardman, J. (1998). An average soil erosion rate for Europe: myth or reality. Journal of Soil and Water Conservation, 53, 4650.Google Scholar
Boardman, J. (2013). The hydrological role of “sunken lanes” with respect to sediment mobilization and delivery to watercourses with particular reference to West Sussex, southern England. Journal of Soils and Sediments, 13, 1636–44.CrossRefGoogle Scholar
Boardman, J. (2014). How old are the gullies (dongas) of the Sneeuberg uplands, Eastern Karoo, South Africa? Catena, 113, 7985.CrossRefGoogle Scholar
Böckh, A. (1973). Consequences of uncontrolled human activities in the Valencia lake basin. In The Careless Technology, ed. Farvar, M. T. and Milton, J. P.. London: Tom Stacey, pp. 301–17.Google Scholar
Boenzi, F., Caldara, M., Capolongo, D., Dellino, P., Piccarreta, M., and Simone, O. (2008). Late Pleistocene–Holocene landscape evolution in Fossa Bradanica, Basilicata (southern Italy). Geomorphology, 102, 297306.CrossRefGoogle Scholar
Boix-Fayos, C., Barberá, G. G., López-Bermúdez, F., and Castillo, V. M. (2007). Effects of check dams, reforestation and land-use changes on river channel morphology: case study of the Rogativa catchment (Murcia, Spain). Geomorphology, 91, 103–23.Google Scholar
Boix‐Fayos, C., de Vente, J., Martínez‐Mena, M., Barberá, G. G., and Castillo, V. (2008). The impact of land use change and check‐dams on catchment sediment yield. Hydrological Processes, 22, 4922–35.CrossRefGoogle Scholar
Bonazza, A., Messina, P., Sabbioni, C., Grossi, C. M., and Brimblecombe, P. (2009a). Mapping the impact of climate change on surface recession of carbonate buildings in Europe. Science of the Total Environment, 407, 2039–50.CrossRefGoogle Scholar
Bonazza, A., Sabbioni, C., Messina, P., Guaraldi, C., and De Nuntiis, P. (2009b). Climate change impact: mapping thermal stress on Carrara marble in Europe. Science of the Total Environment, 407, 4506–12.Google ScholarPubMed
Bonhomme, D., Boudouresque, C. F., Astruch, P., Bonhomme, J., Bonhomme, P., Goujard, A., and Thibaut, T. (2015). Typology of the reef formations of the Mediterranean seagrass Posidonia oceanica, and the discovery of extensive reefs in the Gulf of Hyères (Provence, Mediterranean). Scientific Report of Port-Cros National Park, 29, 4173.Google Scholar
Borrelli, P., Ballabio, C., Panagos, P., and Montanarella, L. (2014). Wind erosion susceptibility of European soils. Geoderma, 232, 471–8.Google Scholar
Bostock, H. C., Lowe, D. J., Gillespie, R., Priestley, R., Newnham, R. M., and Mooney, S. D. (2015). The advent of the Anthropocene in Australasia. Quaternary Australasia, 32(1), 7.Google Scholar
Boulanger, M. T. and Lyman, R. L. (2014). Northeastern North American Pleistocene megafauna chronologically overlapped minimally with Paleoindians. Quaternary Science Reviews, 85, 3546.Google Scholar
Bountry, J. A., Lai, Y. G., and Randle, T. J. (2013). Sediment impacts from the savage rapids dam removal, Rogue River, Oregon. Reviews in Engineering Geology, 21, 93104.CrossRefGoogle Scholar
Boussingault, J. B., 1845, Rural Economy (2nd edn). London: Baillière.Google Scholar
Bowden, W. B. (2010). Climate change in the Arctic – Permafrost, thermokarst, and why they matter to the non‐Arctic world. Geography Compass, 4, 1553–66.CrossRefGoogle Scholar
Bowman, D. M. (2014). What is the relevance of pyrogeography to the Anthropocene? The Anthropocene Review, 2, 73–6.Google Scholar
Bozec, Y. M. and Mumby, P. J. (2015). Synergistic impacts of global warming on the resilience of coral reefs. Philosophical Transactions of the Royal Society B: Biological Sciences, 370(1659), 20130267.CrossRefGoogle Scholar
Bradbury, J., Cullen, P., Dixon, G., and Pemberton, M. (1995). Monitoring and management of streambank erosion and natural revegetation on the lower Gordon River, Tasmanian Wilderness World Heritage Area, Australia. Environmental Management, 19, 259–72.CrossRefGoogle Scholar
Bradley, R. and Fraser, E. (2010). Bronze Age barrows on the heathlands of southern England: construction, forms and interpretations. Oxford Journal of Archaeology, 29, 1533.CrossRefGoogle Scholar
Bradley, R. S., Vuille, M., Diaz, H. F., and Vergara, W. (2006). Threats to water supplies in the tropical Andes. Science, 312, 1755–6.CrossRefGoogle ScholarPubMed
Bradwell, T., Sigurđsson, O., and Everest, J. (2013). Recent, very rapid retreat of a temperate glacier in SE Iceland. Boreas, 42, 959–73.CrossRefGoogle Scholar
Bragg, O. M. and Tallis, J. H. (2001). The sensitivity of peat-covered upland landscapes. Catena, 42, 345–60.CrossRefGoogle Scholar
Braje, T., Erlandson, J., Aikens, C. M., Beach, T., Fitzpatrick, S., Gonzalez, S., and Zeder, M. A. (2014). An Anthropocene without Archaeology—Should we care? The SAA Archaeological Record, 14, 26–9.CrossRefGoogle Scholar
Braje, T. J. (2015). Earth Systems, Human Agency, and the Anthropocene: Planet Earth in the Human Age. Journal of Archaeological Research, doi: 10.1007/s10814-015-9087-y.Google Scholar
Braje, T. J. and Erlandson, J. M. (2014). Looking forward, looking back: humans, anthropogenic change, and the Anthropocene. Anthropocene, 4, 116–21.Google Scholar
Brander, K. E., Owen, K. E. and Potter, K. W. (2004). Modelled impacts of development type on runoff volume and infiltration performance. Journal of the American Water Resources Association, 40, 961–9.CrossRefGoogle Scholar
Brantley, S. L., Megonigal, J. P., Scatena, F. N, Zalogh-Brunstad, Z, Barnes, R. T., Bruns, M. A., and 21 others. (2011). Twelve testable hypotheses on the geobiology of weathering. Geobiology, 9, 140–65.CrossRefGoogle ScholarPubMed
Brattebo, B. O. and Booth, D. B. (2003). Long-term stormwater quantity and quality performance of permeable pavement systems. Water Research, 37, 4369–76.CrossRefGoogle ScholarPubMed
Breckle, S. W., Yair, A., and Veste, M. (2008). General conclusions – sand dune deserts, desertification, rehabilitation and conservation. Ecological Studies 200, 441–59.CrossRefGoogle Scholar
Brenot, J., Quiquerez, A., Petit, C., and Garcia, J. P. (2008). Erosion rates and sediment budgets in vineyards at 1-m resolution based on stock unearthing (Burgundy, France). Geomorphology, 100, 345–55.CrossRefGoogle Scholar
Breslow, P. B. and Sailor, D. J. (2002). Vulnerability of wind power resources to climate change in the continental United States. Renewable Energy, 27, 585–98.CrossRefGoogle Scholar
Brestolani, F., Solari, L., Rinaldi, M., and Lollino, G. (2015). On the morphological impacts of gravel mining: The case of the Orco River. In Engineering Geology for Society and Territory-Volume 3, ed. Lollino, G., Arattano, M., Rinaldi, M., Giustolisi, O., Marechal, J. C. and Grant, G. E.. Springer International Publishing, pp. 319–22.Google Scholar
Brierley, G., Huang, H. Q., Chen, A., Aiken, S., Crozier, M., Eder, W., Goudie, A., Ma, Y., May, J. H., Migon, P., and Nanson, G. A. (2011). Naming conventions in geomorphology: contributions and controversies in the sandstone landscape of Zhangjiajie Geopark, China. Earth Surface Processes and Landforms, 36, 1981–4.CrossRefGoogle Scholar
Bright, M. (2005). 1001 Natural Wonders You Must See Before You Die. London: Cassell Illustrated.Google Scholar
Brimblecombe, P. and Camuffo, D. (2003). Long term damage to the built environment. In The Effects of Air Pollution on the Built Environment, ed. Brimblecombe, P.. London: Imperial College Press, pp. 130.CrossRefGoogle Scholar
Brookes, A. (1985). River channelization: traditional engineering methods, physical consequences, and alternative practices. Progress in Physical Geography, 9, 4473.CrossRefGoogle Scholar
Brookes, A. (1987). River channel adjustments downstream from channelization works in England and Wales. Earth Surface Processes and Landforms, 12, 337–51.CrossRefGoogle Scholar
Brookes, A. and Brierley, G. J. (1997). Geomorphic responses of lower Bega River to catchment disturbance, 1851–1926. Geomorphology, 18, 291304.CrossRefGoogle Scholar
Brooks, N. and Legrand, M. (2000). Dust variability over northern Africa and rainfall in the Sahel. In Linking Land Surface Change to Climate Change, ed. McLaren, S. J., and Kniverton, D.. Dordrecht: Kluwer, pp. 125.Google Scholar
Brooks, S. M. and Spencer, T. (2012). Shoreline retreat and sediment release in response to accelerating sea level rise: measuring and modelling cliffloine dynamics on the Suffolk coast, UK. Global and Planetary Change, 80–81, 165–79.Google Scholar
Brooks, S. S. and Lake, P. S. (2007). River restoration in Victoria, Australia: change is in the wind, and none too soon. Restoration Ecology, 15, 584–91.CrossRefGoogle Scholar
Broothaerts, N., Kissi, E., Poesen, J., Van Rompaey, A., Getahun, K., Van Ranst, E., and Diels, J. (2012). Spatial patterns, causes and consequences of landslides in the Gilgel Gibe catchment, SW Ethiopia. Catena, 97, 127–36.CrossRefGoogle Scholar
Broothaerts, N., Notebaert, B., Verstraeten, G., Kasse, C., Bohncke, S., and Vandenberghe, J. (2014). Non‐uniform and diachronous Holocene floodplain evolution: a case study from the Dijle catchment, Belgium. Journal of Quaternary Science, 29, 351–60.CrossRefGoogle Scholar
Brouns, K., Eikelboom, T., Jansen, P. C., Janssen, R., Kwakernaak, C., van den Akker, J. J., and Verhoeven, J. T. (2015). Spatial analysis of soil subsidence in peat meadow areas in Friesland in relation to land and water management, climate change, and adaptation. Environmental Management, 55, 360–72.CrossRefGoogle ScholarPubMed
Brown, A. E., Zhang, L., McMahon, T. A., Western, A. W., and Vertessy, R. A. (2005). A review of paired catchment studies for determining changes in water yield resulting from alterations in vegetation. Journal of Hydrology, 310, 2861.CrossRefGoogle Scholar
Brown, A. G. (2014). The Anthropocene: a geomorphological and sedimentary view. Strati 2013, 909–14.Google Scholar
Brown, A. G., Tooth, S., Chiverrell, R. C., Rose, J., Thomas, D. S., Wainwright, J., Bullard, J. E., Thorndycraft, V. R., Aalto, R., and Downs, P. (2013). The Anthropocene: is there a geomorphological case? Earth Surface Processes and Landforms 38, 431–4.CrossRefGoogle Scholar
Brown, A. V., Lyttle, M. M., and Brown, K. B. (1998). Impacts of gravel mining on gravel bed streams. Transactions of the American Fisheries Society, 127, 979–94.2.0.CO;2>CrossRefGoogle Scholar
Brown, D., Jorgenson, M. T., Douglas, T. A., Romanovsky, V. E., Kielland, K., Hiemstra, C., Euskirchen, E. S., and Ruess, R. W. (2015). Interactions of fire and climate exacerbate permafrost degradation in Alaskan lowland forests. Journal of Geophysical Research: Biogeosciences. 120, 1619–37, doi: 10.1002/2015JG003033.CrossRefGoogle Scholar
Brown, E. H. (1970). Man shapes the earth. Geographical Journal 136, 7485.CrossRefGoogle Scholar
Brown, S. and Nicholls, R. J. (2015). Subsidence and human influences in mega deltas: the case of the Ganges–Brahmaputra–Meghna. Science of the Total Environment, 527, 362–74.Google ScholarPubMed
Brown, S., Barton, M., and Nicholls, R. (2011). Coastal retreat and/or advance adjacent to defences in England and Wales. Journal of Coastal Conservation, 15, 659–70.CrossRefGoogle Scholar
Bruijnzeel, L. A. (1990). Hydrology of Moist Tropical Forests and Effects of Conversion: a State of Knowledge Review. Amsterdam: Free University for UNESCO International Hydrological Programme.Google Scholar
Bruins, H. J. (2012). Ancient desert agriculture in the Negev and climate-zone boundary changes during average, wet and drought years. Journal of Arid Environments, 86, 2842.CrossRefGoogle Scholar
Brunbjerg, A. K., Svenning, J. C., and Ejrnæs, R. (2014). Experimental evidence for disturbance as key to the conservation of dune grassland. Biological Conservation, 174, 101–10.CrossRefGoogle Scholar
Brunel, C. and Sabatier, F. (2009). Potential influence of sea-level rise in controlling shoreline position on the French Mediterranean Coast. Geomorphology, 107, 4757.CrossRefGoogle Scholar
Brunet, F., Potot, C., Probst, A., and Probst, J. L. (2011). Stable carbon isotope evidence for nitrogenous fertilizer impact on carbonate weathering in a small agricultural watershed. Rapid Communications in Mass Spectrometry, 25, 2682–90.CrossRefGoogle Scholar
Brungard, C. W., Boettinger, J. L., and Hipps, L. E. (2015). Wind erosion potential of lacustrine and alluvial soils before and after disturbance in the eastern Great Basin, USA: estimating threshold friction velocity using easier-to-measure soil properties. Aeolian Research, 18, 185203.CrossRefGoogle Scholar
Brunier, G., Anthony, E. J., Goichot, M., Provansal, M., and Dussouillez, P. (2014). Recent morphological changes in the Mekong and Bassac river channels, Mekong delta: the marked impact of river-bed mining and implications for delta destabilisation. Geomorphology, 224, 177–91.CrossRefGoogle Scholar
Bruno, J. F. and Selig, E. R. (2007). Regional decline of coral cover in the Indo-Pacific: timing, extent and subregional comparisons. PlosOne, 8, e711, doi: 10.1371/journal.pone.0000711.CrossRefGoogle Scholar
Brunsden, D., Coombe, K., Goudie, A. S., and Parker, A. G. (1996). The structural geomorphology of the Isle of Portland, southern England. Proceedings of the Geologists’ Association, 107, 209–30.CrossRefGoogle Scholar
Bruun, P. (1962). Sea level rise as a cause of shore erosion. American Society of Civil Engineers Proceedings: Journal of Waterways and Harbors Division, 88, 117–30.Google Scholar
Bryan, K. (1928). Historic evidence of changes in the channel of Rio Puerco, a tributary of the Rio Grande in New Mexico. Journal of Geology, 36, 265–82.CrossRefGoogle Scholar
Buendia, C., Batalla, R. J., Sabater, S., Palau, A., and Marcé, R. (2015). Runoff trends driven by climate and afforestation in a Pyrenean Basin. Land Degradation and Development, doi: 10.1002/ldr.2384.CrossRefGoogle Scholar
Buijse, A. D., Coops, H., Staras, M., Jans, L. H., Van Geest, G. J., Grift, R. E., Ibelings, B. W., Oosterberg, W., and Roozen, F. C. (2002). Restoration strategies for river floodplains along large lowland rivers in Europe. Freshwater Biology, 47, 889907.CrossRefGoogle Scholar
Bull, W. B. (1997). Discontinuous ephemeral streams. Geomorphology, 19, 227–76.CrossRefGoogle Scholar
Burrin, P. J. (1985). Holocene alluviation in southeast England and some implications for palaeohydrological studies. Earth Surface Processes and Landforms, 10, 257–71.CrossRefGoogle Scholar
Burroughs, B. A., Hayes, D. B., Klomp, K. D., Hansen, J. F., and Mistak, J. (2009). Effects of Stronach Dam removal on fluvial geomorphology in the Pine River, Michigan, United States. Geomorphology, 110, 96107.CrossRefGoogle Scholar
Burt, J. A. (2014). The environmental costs of coastal urbanization in the Arabian Gulf. City, 18, 760–70.CrossRefGoogle Scholar
Burt, J. A., Bartholomew, A., and Feary, D. A. (2012). Man-made structures as artificial reefs in the Gulf. In Coral Reefs of the Gulf, ed. Riegel, B. M. and Purkis, S. J.. Netherlands: Springer, pp. 171–86.Google Scholar
Burt, J. A., Bartholomew, A., Usseglio, P., Bauman, A., and Sale, P. F. (2009). Are artificial reefs surrogates of natural habitats for corals and fish in Dubai, United Arab Emirates? Coral Reefs, 28, 663–75.CrossRefGoogle Scholar
Burt, T. P. (1994). Long-term study of the natural environment – perceptive science or mindless monitoring? Progress in Physical Geography, 18, 475–96.Google Scholar
Butler, D. R. (2006). Human-induced changes in animal populations and distributions, and the subsequent effects on fluvial systems. Geomorphology, 79, 448–59.CrossRefGoogle Scholar
Buttrick, D. and Van Schalkwyk, A. (1998). Hazard and risk assessment for sinkhole formation on dolomite land in South Africa. Environmental Geology, 36, 170–8.CrossRefGoogle Scholar
Butzer, K. W. (1974). Accelerated soil erosion: a problem of man-land relationships. In Perspectives on Environments, ed. Manners, I. R. and Mikesell, M. W.. Washington, DC: Association of American Geographers. pp. 5777.Google Scholar
Butzer, K. W. (1976). Early Hydraulic Civilization in Egypt. Chicago: University of Chicago Press.Google Scholar
Butzer, K. W. (2005). Environmental history in the Mediterranean world: cross-disciplinary investigation of cause-and-effect for degradation and soil erosion. Journal of Archaeological Science, 32, 1773–800.CrossRefGoogle Scholar
Buynevich, I., Bitinas, A., and Pupienis, D. (2007). Reactivation of coastal dunes documented by subsurface imaging of the Great Dune Ridge, Lithuania. Journal of Coastal Research, 50, 226–30.Google Scholar
Cadol, D., Rathburn, S. L., and Cooper, D. J. (2011). Aerial photographic analysis of channel narrowing and vegetation expansion in Canyon de Chelly National Monument, Arizona, USA, 1935–2004. River Research and Applications, 27, 841–56.CrossRefGoogle Scholar
Caldararo, N. (2002). Human ecological intervention and the role of forest fires in human ecology. The Science of the Total Environment, 292, 141–65.CrossRefGoogle ScholarPubMed
Caldiera, K. and Wickett, M. G. (2003). Anthropogenic carbon and ocean pH. Nature, 425, 365.Google Scholar
Caldwell, P. V., Sun, G., McNulty, S. G., Cohen, E. C., and Moore Myers, J. A. (2012). Impacts of impervious cover, water withdrawals, and climate change on river flows in the conterminous US. Hydrology and Earth System Sciences, 16, 2839–57.CrossRefGoogle Scholar
Calmels, D., Gaillardet, J., and François, L. (2014). Sensitivity of carbonate weathering to soil CO2 production by biological activity along a temperate climate transect. Chemical Geology, 390, 7486.CrossRefGoogle Scholar
Cambi, M., Certini, G., Neri, F., and Marchi, E. (2015). The impact of heavy traffic on forest soils: a review. Forest Ecology and Management, 338, 124–38.CrossRefGoogle Scholar
Cannon, S. H., Bigio, E. R., and Mine, E. (2001a,). A process for fire-related debris flow initiation, Cerro Grande fire, New Mexico. Hydrological Processes, 15, 3011–23.CrossRefGoogle Scholar
Cannon, S. H., Kirkham, R. M., and Parise, M. (2001b). Wildfire-related debris-flow initiation processes, Storm King Mountains, Colorado. Geomorphology, 39, 171–88.CrossRefGoogle Scholar
Cao, S., Tian, T., Chen, L., Dong, X., Yu, X., and Wang, G. (2010). Damage caused to the environment by reforestation policies in arid and semi-arid areas of China. Ambio, 39, 279–83.CrossRefGoogle Scholar
Capezzuoli, E., Gandin, A., and Sandrelli, F. (2010). Calcareous tufa as indicators of climatic variability: a case study from southern Tuscany (Italy). Geological Society, London, Special Publications, 336, 263–81.CrossRefGoogle Scholar
Carbognin, L., Teatini, P., Tomasin, A., and Tosi, L. (2010). Global change and relative sea level rise at Venice: what impact in term of flooding. Climate Dynamics, 35, 1039–47.CrossRefGoogle Scholar
Carbonell, E. and 29 others (2008). The first hominim of Europe. Nature, 452, 465–9.CrossRefGoogle ScholarPubMed
Carey, J. C. and Fulweiler, R. W. (2012). Human activities directly alter watershed dissolved silica fluxes. Biogeochemistry, 111, 125–38.CrossRefGoogle Scholar
Carminati, E. and Martinelli, G. (2002). Subsidence rates in the Po Plain, northern Italy: the relative impact of natural and anthropogenic causation. Engineering Geology, 66, 241–55.CrossRefGoogle Scholar
Carmo, J. A., Reis, C. S., and Freitas, H. (2010). Working with nature by protecting sand dunes: lessons learned. Journal of Coastal Research 26, 1068–78.CrossRefGoogle Scholar
Carmona, P. and Ruiz, J. M. (2011). Historical morphogenesis of the Turia River coastal flood plain in the Mediterranean littoral of Spain. Catena, 86, 139–49.CrossRefGoogle Scholar
Carnec, C. and Fabriol, H. (1999). Monitoring and modeling land subsidence at the Cerro Prieto geothermal field, Baja California, Mexico, using SAR interferometry. Geophysical Research Letters, 26, 1211–4.CrossRefGoogle Scholar
Carr, J. R., Stokes, C., and Vieli, A. (2014). Recent retreat of major outlet glaciers on Novaya Zemlya, Russian Arctic, influenced by fjord geometry and sea-ice conditions. Journal of Glaciology, 60, 155–70.CrossRefGoogle Scholar
Carrara, P. E. and Carroll, T. R. (1979). The determination of erosion rates from exposed tree roots in the Piceance Basin, Colorado. Earth Surface Processes, 4, 407–17.CrossRefGoogle Scholar
Carrivick, J. L. and Quincey, D. J. (2014). Progressive increase in number and volume of ice-marginal lakes on the western margin of the Greenland Ice Sheet. Global and Planetary Change, 116, 156–63.CrossRefGoogle Scholar
Casalí, J., Giménez, R., De Santisteban, L., Álvarez-Mozos, J., Mena, J., and Del Valle de Lersundi, J. (2009). Determination of long-term erosion rates in vineyards of Navarre (Spain) using botanical benchmarks. Catena, 78, 12–9.CrossRefGoogle Scholar
Casana, J. (2008). Mediterranean valleys revisited: Linking soil erosion, land use and climate variability in the Northern Levant. Geomorphology, 101, 429–42.CrossRefGoogle Scholar
Castaneda, C. and Herrero, J. (2008). Assessing the degradation of saline wetlands in an arid agricultural region in Spain. Catena, 72, 205–13.CrossRefGoogle Scholar
Castillo, V. M., Mosch, W. M., Garcia, C. C., Barberá, G. G., Cano, J. A., and López-Bermúdez, F. (2007). Effectiveness and geomorphological impacts of check dams for soil erosion control in a semiarid Mediterranean catchment: El Cárcavo (Murcia, Spain). Catena, 70, 416–27.CrossRefGoogle Scholar
Castree, N. (2014a). the Anthropocene and geography I: the back story. Geography Compass, 8, 436–49.Google Scholar
Castree, N. (2014b). Geography and the Anthropocene II: current contributions. Geography Compass, 8, 450–63.Google Scholar
Castree, N. (2014c). The Anthropocene and geography III: future directions. Geography Compass, 8, 464–76.Google Scholar
Castillo, V. M., Mosch, W. M., Garcia, C. C., Barberá, G. G., Cano, J. A., and López-Bermúdez, F. (2015). The Anthropocene: a primer for geograpers. Geography, 100, 6675.Google Scholar
Cathcart, R. B. (1983). Mediterranean Basin – Sahara reclamation. Speculations in Science and Technology, 6, 150–2.Google Scholar
Catto, N. (2002). Anthropogenic pressures on coastal dunes, southwestern Newfoundland. The Canadian Geographer, 46, 1732.CrossRefGoogle Scholar
Cavallin, A., Marchetti, M., Panizza, M., and Soldati, M. (1994). The role of geomorphology in environmental impact assessment. Geomorphology, 9, 143–53.CrossRefGoogle Scholar
Cavanaugh, K. C., Kellner, J. R., Forde, A. J., Gruner, D. S., Parker, J. D., Rodriguez, W., and Feller, I. C. (2014). Poleward expansion of mangroves is a threshold response to decreased frequency of extreme cold events. Proceedings of the National Academy of Sciences, 111, 723–7.CrossRefGoogle ScholarPubMed
Cazenave, A. and Cozannet, G. L. (2014). Sea level rise and its coastal impacts. Earth’s Future, 2(2), 1534.CrossRefGoogle Scholar
Cazenave, A. and Llovel, W. (2010). Contemporary sea level rise. Annual Review of Marine Science, 2, 145–73.CrossRefGoogle ScholarPubMed
Cazenave, A., Dieng, H. B., Meyssignac, B., von Schuckmann, K., Decharme, B., and Berthier, E. (2014). The rate of sea-level rise. Nature Climate Change, 4, 358–61.CrossRefGoogle Scholar
Cerdà, A. (2007). Soil water erosion on road embankments in eastern Spain. Science of the Total Environment, 378, 151–5.CrossRefGoogle ScholarPubMed
Cerdà, A. and Doerr, S. H. (2008). The effect of ash and needle cover on surface runoff and erosion in the immediate post-fire period. Catena, 74, 256–63.CrossRefGoogle Scholar
Cerdan, O., Govers, G., Le Bissonnais, Y., Van Oost, K., Poesen, J., Saby, N., and Dostal, T. (2010). Rates and spatial variations of soil erosion in Europe: a study based on erosion plot data. Geomorphology, 122, 167–77.CrossRefGoogle Scholar
Cerney, D. L. (2010). The use of repeat photography in contemporary geomorphic studies: an evolving approach to understanding landscape change. Geography Compass, 4, 1339–57.CrossRefGoogle Scholar
Certini, G. and Scalenghe, R. (2011). Anthropogenic soils are the golden spikes for the Anthropocene. The Holocene, 21, 1269–74.CrossRefGoogle Scholar
Chai, J-C., Shen, S-L. Zhu, H. H., and Zhang, X. L. (2004). Land subsidence due to groundwater drawdown in Shanghai. Géotechnique, 54, 143–7.CrossRefGoogle Scholar
Chan, N. and Connolly, S. R. (2013). Sensitivity of coral calcification to ocean acidification: a meta‐analysis. Global Change Biology, 19, 282–90.CrossRefGoogle ScholarPubMed
Chang, J. C. and Slaymaker, O. (2002). Frequency and spatial distribution of landslides in a mountainous drainage basin: Western Foothills, Taiwan. Catena, 46, 285307.CrossRefGoogle ScholarPubMed
Chappell, A., Webb, N. P., Butler, H. J., Strong, C. L., McTainsh, G. H., Leys, J. F., and Viscarra Rossel, R. A. (2013). Soil organic carbon dust emission: an omitted global source of atmospheric CO2. Global Change Biology, 19, 3238–44.CrossRefGoogle Scholar
Charlier, R. H., Chaineux, M. C. P., and Morcos, S. (2005). Panorama of the history of coastal protection. Journal of Coastal Research, 21, 79111.CrossRefGoogle Scholar
Chaussard, E., Amelung, F., Abidin, H., and Hong, S. H. (2013). Sinking cities in Indonesia: Alos Palsar detects rapid subsidence due to groundwater and gas extraction. Remote Sensing of Environment, 128, 150–61.CrossRefGoogle Scholar
Chen, C., Pei, S., and Jiao, J. (2003). Land subsidence caused by groundwater exploitation in Suzhou City, China. Hydrogeology Journal, 11, 275–87.CrossRefGoogle Scholar
Chen, C. T., Hu, J. C., Lu, C. Y., Lee, J. C., and Chan, Y. C. (2007). Thirty-year land elevation change from subsidence to uplift following the termination of groundwater pumping and its geological implications in the Metropolitan Taipei Basin, Northern Taiwan. Engineering Geology, 95, 3047.CrossRefGoogle Scholar
Chen, J. L., Wilson, C. R., and Tapley, B. D. (2013). Contribution of ice sheet and mountain glacier melt to recent sea level rise. Nature Geoscience, 6, 549–52.CrossRefGoogle Scholar
Chen, N., Wu, Y., Wu, J., Yan, X., and Hong, H. (2014). Natural and human influences on dissolved silica export from watershed to coast in Southeast China. Journal of Geophysical Research: Biogeosciences, 119, 95109.CrossRefGoogle Scholar
Chen, N., Chen, M., Li, J., He, N., Deng, M., Tanoli, J. I., and Cai, M. (2015). Effects of human activity on erosion, sedimentation and debris flow activity– a case study of the Qionghai Lake watershed, southeastern Tibetan Plateau, China. The Holocene, 25, 973–88.CrossRefGoogle Scholar
Chen, X. Q., Cui, P., Li, Y., Yang, Z., and Qi, Y. Q. (2007). Changes in glacial lakes and glaciers of post-1986 in the Poiqu River basin, Nyalam, Xizang (Tibet). Geomorphology, 88, 298311.CrossRefGoogle Scholar
Chen, Y., Xu, C., Chen, Y., Li, W., and Liu, J. (2010). Response of glacial-lake outburst floods to climate change in the Yarkant River basin on northern slope of Karakoram Mountains, China. Quaternary International, 226, 7581.CrossRefGoogle Scholar
Chen, Z. and Saito, Y. (2011). The Megadeltas of Asia: interlinkage of land and sea, and human development. Earth Surface Processes and Landforms, 36, 1703–4.CrossRefGoogle Scholar
Cheng, F. and Granata, T. (2007). Sediment transport and channel adjustments associated with dam removal: field observations. Water Resources Research, 43, W03444, doi:10.1029/2005WR004271.CrossRefGoogle Scholar
Cheng, J. D., Lin, J. P., Lu, S. Y., Huang, L. S., and Wu, H. L. (2008). Hydrological characteristics of betel nut plantations on slopelands in central Taiwan. Hydrological Sciences Journal, 53, 1208–20.Google Scholar
Chepil, W. S. (1945). Dynamics of wind erosion. Soil Science, 60, 305–20; 397–411; 475–80.CrossRefGoogle Scholar
Chepil, W. S. and Woodruff, N. P. (1963). The physics of wind erosion and its control. Advances in Agronomy, 15, 211302.CrossRefGoogle Scholar
Chevallier, P., Pouyaud, B., Suarez, W., and Condom, T. (2011). Climate change threats to environment in the tropical Andes: glaciers and water resources. Regional Environmental Change, 11, 179–87.CrossRefGoogle Scholar
Chi, S. C. and Reilinger, R. E. (1984). Geodetic evidence for subsidence due to groundwater withdrawal in many parts of the United States of America. Journal of Hydrology, 67, 155–82.CrossRefGoogle Scholar
Chiang, S. H. and Chang, K. T. (2011). The potential impact of climate change on typhoon-triggered landslides in Taiwan, 2010–2099. Geomorphology, 133, 143–51.CrossRefGoogle Scholar
Chin, A. (2006). Urban transformation of river landscapes in a global context. Geomorphology, 79, 460–87.CrossRefGoogle ScholarPubMed
Chin, A., Florsheim, J. L., Wohl, E., and Collins, B. D. (2014). Feedbacks in human–landscape systems. Environmental Management, 53, 2841.Google Scholar
Chin, M., Diehl, T., Tan, Q., Prospero, J. M., Kahn, R. A., Remer, L. A., Yu, H., Sayer, A. M., Bian, H., Geogdzhayev, I. V., Holben, B. N., Howell, S. G., Huebert, B. J., Hsu, N. C., Kim, D., Kucsera, T. L., Levy, R. C., Mishchenko, M. I., Pan, X., Quinn, P. K., Schuster, G. L., Streets, D. G., Strode, S. A., Torres, O., and Zhao, X.-P. (2014). Multi-decadal aerosol variations from 1980 to 2009: a perspective from observations and a global model. Atmospheric Chemistry and Physics, 14, 3657–90.CrossRefGoogle Scholar
Chiverrell, R. C., Harvey, A. M., and Foster, G. C. (2007). Hillslope gullying in the Solway Firth—Morecambe Bay region, Great Britain: Responses to human impact and/or climatic deterioration? Geomorphology, 84, 317–43.CrossRefGoogle Scholar
Chu, Z. X., Zhai, S. K., Lu, X. X., Liu, J. P., Xu, J. X., and Xu, K.H. (2009). A quantitative assessment of human impacts on decrease in sediment flux from major Chinese rivers entering the western Pacific Ocean. Geophysical Research Letters, 36, L19603, doi: 10.1029/2009GL039513.Google Scholar
Church, J. A. and 13 others (2013). Sea level change. (2013). In Climate Change 2013: the Physical Science Basis, ed. Stocker, T. F., Qin, D., Plattner, G. K., Tignor, M., Allen, S. K., Boschung, J., and Midgley, P. M.. Cambridge: Cambridge University Press, pp. 1137–216.CrossRefGoogle Scholar
Church, M. J., Burt, T., Galay, V. J., and Kondolf, G. M. (2009). Rivers. In Geomorphology and Global Environmental Change, ed. Slaymaker, O., Spencer, T., and Embleton-Hamann, C.. Cambridge: Cambridge University Press, pp. 98129.CrossRefGoogle ScholarPubMed
Ciccarelli, D. (2014). Mediterranean coastal sand dune vegetation: influence of natural and anthropogenic Factors. Environmental Management, 54, 194204.CrossRefGoogle Scholar
Clarke, D. and Ayutthaya, S. S. N. (2010). Predicted effects of climate change, vegetation and tree cover on dune slack habitats at Ainsdale on the Sefton Coast, UK. Journal of Coastal Conservation, 14, 115–25.CrossRefGoogle ScholarPubMed
Clark, J. M., Bottrell, S. H., Evans, C. D., Monteith, D. T., Bartlett, R., Rose, R., and Chapman, P. J. (2010). The importance of the relationship between scale and process in understanding long-term DOC dynamics. Science of the Total Environment, 408, 2768–75.CrossRefGoogle Scholar
Clark, P. U., Church, J. A., Gregory, J. M., and Payne, A. J. (2015). Recent progress in understanding and projecting regional and global mean sea level change. Current Climate Change Reports, 1, 123.3.0.CO;2-U>CrossRefGoogle Scholar
Clark, S. and Edwards, A. J. (1999). An evaluation of artificial reef structures as tools for marine habitat rehabilitation in the Maldives. Aquatic Conservation: Marine and Freshwater Ecosystems, 9, 521.CrossRefGoogle Scholar
Clarke, G. K., Jarosch, A. H., Anslow, F. S., Radić, V., and Menounos, B. (2015). Projected deglaciation of western Canada in the twenty-first century. Nature Geoscience, doi:10.1038/NGE02407.CrossRefGoogle Scholar
Clarke, M. A. and Walsh, R. P. D. (2006). Long-term erosion and surface roughness change of rain-forest terrain following selective logging, Danum Valley, Sabah, Malaysia. Catena, 68, 109–23.CrossRefGoogle Scholar
Clarke, M. L. and Rendell, H. M. (2009). The impact of North Atlantic storminess on western European coasts: a review. Quaternary International, 195, 3141.CrossRefGoogle Scholar
Clarke, M. L. and Rendell, H. M. (2015). “This restless enemy of all fertility”: exploring paradigms of coastal dune management in Western Europe over the last 700 years. Transactions of the Institute of British Geographers, 40, 414–29.CrossRefGoogle Scholar
Clay, G. D., Dixon, S., Evans, M. G., Rowson, J. G., and Worrall, F. (2012). Carbon dioxide fluxes and DOC concentrations of eroding blanket peat gullies. Earth Surface Processes and Landforms, 37, 562–71.CrossRefGoogle Scholar
Clements, R., Sodhi, N. S., Schilthuizen, M., and Ng, P. K. (2006). Limestone karsts of Southeast Asia: imperiled arks of biodiversity. Bioscience, 56, 733–42.CrossRefGoogle Scholar
Clemmensen, L. B., Bjornsen, M., Murray, A., and Pedersen, K. (2007). Formation of aeolian dunes on Anholt, Denmark since AD 1560: a record of deforestation and increased storminess. Sedimentary Geology, 199, 171–87.CrossRefGoogle Scholar
Closson, D., LaMoreaux, P. E., Karaki, N. A. and Al-Fugha, H. (2007). Karst system developed in salt layers of the Lisan Peninsula, Dead Sea, Jordan. Environmental Geology, 52, 155–72.CrossRefGoogle Scholar
Clymans, W., Struyf, E., Govers, G., Vandevenne, F., and Conley, D. J. (2011). Anthropogenic impact on amorphous silica pools in temperate soils. Biogeosciences, 8, 2281–93.CrossRefGoogle Scholar
Coates, D. R. (1977). Landslide perspective. Reviews in Engineering Geology, 3, 328.Google Scholar
Cocco, S., Brecciaroli, G., Agnelli, A., Weindorf, D., and Corti, G. (2015). Soil genesis and evolution on calanchi (badland-like landform) of central Italy. Geomorphology, 248, 3345.CrossRefGoogle Scholar
Cohen, J. E. (2003). Human population: the next half century. Science, 302, 1172–5.CrossRefGoogle ScholarPubMed
Coleman, R. (1981). Footpath erosion in the English Lake District. Applied Geography, 1, 121–31CrossRefGoogle Scholar
Collison, A., Wade, S., Griffiths, J., and Dehn, M. (2000). Modelling the impact of predicted climate change on landslide frequency and magnitude in S. E. England. Engineering Geology, 55, 205–18.CrossRefGoogle Scholar
Comeaux, R. S., Allison, M. A., and Bianchi, T. S. (2012). Mangrove expansion in the Gulf of Mexico with climate change: implications for wetland health and resistance to rising sea levels. Estuarine, Coastal and Shelf Science, 96, 8195.CrossRefGoogle Scholar
Comiti, F., Da Canal, M., Surian, N., Mao, L., Picco, L., and Lenzi, M. A. (2011). Channel adjustments and vegetation cover dynamics in a large gravel bed river over the last 200 years. Geomorphology, 125, 147–59.CrossRefGoogle Scholar
Compton, J. S., Herbert, C. T., Hoffman, M. T., Schneider, R. R., and Dtuut, J. B. (2010). A tenfold increase in the Orange River mean Holocene mud flux: implications for soil erosion in South Africa. The Holocene, 20, 115–22.CrossRefGoogle Scholar
Conley, D. J., Likens, G. E., Buso, D. C., Saccone, L., Bailey, S. W., and Johnson, C. E. (2008). Deforestation causes increased dissolved silicate losses in the Hubbard Brook Experimental Forest. Global Change Biology, 14, 2548–54.CrossRefGoogle Scholar
Connaughton, C. A. (1935). Forest fires and accelerated erosion. Journal of Forestry, 33, 751–2.Google Scholar
Conway, V. M. (1954). Stratigraphy and pollen analysis of southern Pennine blanket peats. Journal of Ecology, 42, 117–47.CrossRefGoogle Scholar
Cook, A. J. and Vaughan, D. G. (2010). Overview of areal changes of the ice shelves on the Antarctic Peninsula over the past 50 years. The Cryosphere, 4, 7798.CrossRefGoogle Scholar
Cook, A. J., Fox, A. J., Vaughan, D. G., and Ferrigno, J. G. (2005). Retreating glacier fronts on the Antarctic Peninsula over the past half-century. Science, 308, 541–4.CrossRefGoogle ScholarPubMed
Cook, B. I., Miller, R. L., and Seager, R. (2009). Amplification of the North American “Dust Bowl” drought through human-induced land degradation. Proceedings of the National Academy of Sciences, 106, 49975001.CrossRefGoogle ScholarPubMed
Cooke, R. U. and Reeves, R. W. (1976). Arroyos and Environmental Change in the American South-west. Oxford: Clarendon Press.Google Scholar
Coones, P. and Patten, J. H. C. (1986). The landscape of England and Wales. Harmondsworth: Penguin Books.Google Scholar
Cooper, A. H. (2002). Halite karst geohazards (natural and man-made) in the United Kingdom. Environmental Geology, 42, 505–12.CrossRefGoogle Scholar
Cooper, J. A. G. and Pilkey, O. H. (2004). Sea-level rise and shoreline retreat: time to abandon the Bruun Rule. Global and Planetary Change, 43, 157–71.CrossRefGoogle Scholar
Cooper, S. D., Lake, P. S., Sabater, S., Melack, J. M., and Sabo, J. L. (2013). The effects of land use changes on streams and rivers in Mediterranean climates. Hydrobiologia, 719, 383425.CrossRefGoogle Scholar
Cordova, C. E. (2008). Floodplain degradation and settlement history in Wadi al-Wala and Wadi ash-Shallalah, Jordan. Geomorphology, 101, 443–57.CrossRefGoogle Scholar
Corella, J. P., El Amrani, A., Sigró, J., Morellón, M., Rico, E., and Valero-Garcés, B. L. (2011). Recent evolution of Lake Arreo, northern Spain: influences of land use change and climate. Journal of Paleolimnology, 46, 469–85.CrossRefGoogle Scholar
Cornell, J. D. and Miller, M. (2007). Slash and burn. Encyclopedia of Earth, 31, www.eoearth.org/view/article/156045. (Accessed October 16, 2015).Google Scholar
Corona, M. G., Vincente, A. M., and Novo, F. G. (1988). Long-term vegetation changes on the stabilized dunes of Doñana National Park (SW Spain). Vegetatio 75, 7380.CrossRefGoogle Scholar
Corti, G., Cavallo, E., Cocco, S., Biddoccu, M., Brecciaroli, G., and Agnelli, A. (2011). Evaluation of erosion intensity and some of its consequences in vineyards from two hilly environments under a Mediterranean type of climate, Italy. In Soil Erosion in Agriculture, ed. Godone, D. and Stranchi, S.. Rijeka: Intech, pp. 113–60.Google Scholar
Cosandey, C., Andréassian, V., Martin, C., Didon-Lescot, J. F., Lavabre, J., Folton, N., and Richard, D. (2005). The hydrological impact of the Mediterranean forest: a review of French research. Journal of Hydrology, 301, 235–49.CrossRefGoogle Scholar
Costa-Cabral, M., Roy, S. B., Maurer, E. P., Mills, W. B., and Chen, L. (2013). Snowpack and runoff response to climate change in Owens Valley and Mono Lake watersheds. Climatic Change, 116, 97109.CrossRefGoogle Scholar
Cosyns, E., Degezelle, T., Demeulenaere, E., and Hoffmann, M. (2001). Feeding ecology of Konik horses and donkeys in Belgian coastal dunes and its implications for nature management. Belgian Journal of Zoology, 131, Supplement 2, 111–8.Google Scholar
Cotton, J. M., Jeffery, M. L., and Sheldon, N. D. (2013). Climate controls on soil respired CO2 in the United States: implications for 21st century chemical weathering rates in temperate and arid ecosystems. Chemical Geology, 358, 3745.CrossRefGoogle Scholar
Coverdale, T. C., Herrmann, N. C., Altieri, A. H., and Bertness, M. D. (2013). Latent impacts: the role of historical human activity in coastal habitat loss. Frontiers in Ecology and the Environment, 11, 6974.CrossRefGoogle Scholar
Cowie, S. M., Knippertz, P., and Marsham, J. H. (2013). Are vegetation-related roughness changes the cause of the recent decrease in dust emission from the Sahel? Geophysical Research Letters, 40, 1868–72.CrossRefGoogle ScholarPubMed
Cox, R., Bierman, P., Jungers, M. C., and Rakotondrazafy, A. M. (2009). Erosion rates and sediment sources in Madagascar inferred from 10Be analysis of lavaka, slope, and river sediment. Journal of Geology, 117, 363–76.CrossRefGoogle Scholar
Cox, R., Zentner, D. B., Rakotondrazafy, A. F. M., and Rasoazanamparany, C. F. (2010). Shakedown in Madagascar: occurrence of lavakas (erosional gullies) associated with seismic activity. Geology, 38, 179–82.CrossRefGoogle Scholar
Cremaschi, M. (2014). When did the Anthropocene begin? A geoarchaeological approach to deciphering the consequences of human activity in pre-protohistoric times: selected cases from the Po Plain (northern Italy). Rendiconti Lincei, 25, 101–12.CrossRefGoogle Scholar
Croke, J. and Nethery, M. (2006). Modelling runoff and soil erosion in logged forests: scope and application of some existing models. Catena, 67, 3549.CrossRefGoogle Scholar
Croke, J., Hairsine, P., and Fogarty, P. (1999). Sediment transport, redistribution and storage on logged forest hillslopes in south‐eastern Australia. Hydrological Processes, 13, 2705–20.3.0.CO;2-Y>CrossRefGoogle Scholar
Croke, J., Hairsine, P., and Fogarty, P. (2001). Soil recovery from track construction and harvesting changes in surface infiltration, erosion and delivery rates with time. Forest Ecology and Management, 143, 312.CrossRefGoogle Scholar
Crooks, J. A. (2002). Characterizing ecosystem‐level consequences of biological invasions: the role of ecosystem engineers. Oikos, 97, 153–66.CrossRefGoogle Scholar
Crozier, M. J. (2010). Deciphering the effect of climate change on landslide activity: a review. Geomorphology, 124, 260–7.CrossRefGoogle Scholar
Crozier, M. J., Marx, S. L., and Grant, I. J., 1978, Impact of off-road recreational vehicles on soil and vegetation. Proceedings of the 9th New Zealand Geography Conference, Dunedin, 76–79.Google Scholar
Crutzen, P. J. (2002). Geology of mankind. Nature, 415, 23.CrossRefGoogle ScholarPubMed
Csiki, S. J. and Rhoads, B. L. (2014). Influence of four run-of-river dams on channel morphology and sediment characteristics in Illinois, USA. Geomorphology, 206, 215–29.CrossRefGoogle Scholar
Cui, Y., Wooster, J. K., Braudrick, C. A., and Orr, B. K. (2014). Lessons learned from sediment transport model predictions and long-term postremoval monitoring: Marmot Dam removal project on the Sandy River in Oregon. Journal of Hydraulic Engineering, 140, 04014044.CrossRefGoogle Scholar
Curreli, A., Wallace, H., Freeman, C., Hollingham, M., Stratford, C., Johnson, H., and Jones, L. (2013). Eco-hydrological requirements of dune slack vegetation and the implications of climate change. Science of the Total Environment, 443, 910–9.CrossRefGoogle ScholarPubMed
Cushman, G. T. (2011). Humboldtian science, creole meteorology, and the discovery of human-caused climate change in South America. Osiris, 26, 1944.CrossRefGoogle ScholarPubMed
Cuvilliez, A., Deloffre, J., Lafite, R., and Bessineton, C. (2009). Morphological responses of an estuarine intertidal mudflat to constructions since 1978 to 2005: the Seine estuary (France). Geomorphology, 104, 165–74.CrossRefGoogle Scholar
Cypser, D. A. and Davis, S. D., 1998, Induced seismicity and the potential for liability under U.S. law. Tectonophysics, 289, 239–55.CrossRefGoogle Scholar
Dabkowski, J., Brou, L., and Naton, H. G. (2015). New stratigraphic and geochemical data on the Holocene environment and climate from a tufa deposit at Direndall (Mamer Valley, Luxembourg). The Holocene, 25, 1153–64.CrossRefGoogle Scholar
Dadson, S. (2010). Geomorphology and Earth system science. Progress in Physical Geography, 34, 385–98.Google Scholar
Dai, A. (2011). Drought under global warming: a review. Interdisciplinary Reviews: Climate Change, 2, 4565.Google Scholar
Dai, S. B. and Lu, X. X. (2014). Sediment load change in the Yangtze River (Changjiang): a review. Geomorphology, 215, 6073.CrossRefGoogle Scholar
Dai, Z. and Liu, J. T. (2013). Impacts of large dams on downstream fluvial sedimentation: an example of the Three Gorges Dam (TGD) on the Changjiang (Yangtze River). Journal of Hydrology, 480, 10–8.CrossRefGoogle Scholar
Dale, T. and Carter, V. G. (1955) Topsoil and Civilization. Norman, Oklahoma: University of Oklahoma Press.Google Scholar
da Luz, R. A. and Rodrigues, C. (2015). Anthropogenic changes in urbanised hydromorphological systems in a humid tropical environment: River Pinheiros, Sao Paulo, Brazil. Zeitschrift für Geomorphologie, Supplementary Issues, 59, 109–35.Google Scholar
Daniel, D. W., Smith, L. M., and McMurry, S. T. (2015). Land use effects on sedimentation and water storage volume in playas of the rainwater basin of Nebraska. Land Use Policy, 42, 426–31.CrossRefGoogle Scholar
Daniel, T. C., McGuire, P. E., Stoffel, D., and Millfe, B. (1979). Sediment and nutrient yield from residential construction sites. Journal of Environmental Quality, 8, 304–8.CrossRefGoogle Scholar
Dapples, F., Lotter, A. F., van Leeuwen, J. F., van der Knaap, W. O., Dimitriadis, S., and Oswald, D. (2002). Paleolimnological evidence for increased landslide activity due to forest clearing and land-use since 3600 cal BP in the western Swiss Alps. Journal of Paleolimnology, 27, 239–48.CrossRefGoogle Scholar
DasGupta, R. and Shaw, R. (2013). Cumulative impacts of human interventions and climate change on mangrove ecosystems of South and Southeast Asia: an overview. Journal of Ecosystems, article 379429.Google Scholar
Davies, B. J and Glasser, N. F. (2012). Accelerating shrinkage of Patagonian glaciers from the Little Ice Age (∼AD 1870) to 2011. Journal of Glaciology, 58, 1063–84.CrossRefGoogle Scholar
Davies, R., Foulger, G., Bindley, A., and Styles, P. (2013). Induced seismicity and hydraulic fracturing for the recovery of hydrocarbons. Marine and Petroleum Geology, 45, 171–85.CrossRefGoogle Scholar
Dawson, Q., Kechavarzi, C., Leeds-Harrison, P. B., and Burton, R. G. O. (2010). Subsidence and degradation of agricultural peatlands in the Fenlands of Norfolk, U.K. Geoderma, 154, 181–7.CrossRefGoogle Scholar
Day, M. (1996). Conservation of karst in Belize. Journal of Caves and Karst Studies, 58, 139–44.Google Scholar
De Alba, S., Lindstrom, M., Schumacher, T. E., and Malo, D. D. (2004). Soil landscape evolution due to soil redistribution by tillage: a new conceptual model of soil catena evolution in agricultural landscapes. Catena, 58, 77100.CrossRefGoogle Scholar
De Angelis, H. and Skvarca, P. (2003). Glacier surge after Ice Shelf collapse. Science, 299, 1560–2.CrossRefGoogle ScholarPubMed
De Bruyn, I. A. and Bell, F. G. (2001). The occurrence of sinkholes and subsidence depressions in the Far West Rand and Gauteng Province, South Africa, and their engineering implications. Environmental & Engineering Geoscience, 7, 281–95.CrossRefGoogle Scholar
de Jong, C., Carletti, G., and Previtali, F. (2015). Assessing impacts of climate change, ski slope, snow and hydraulic engineering on slope stability in ski resorts (French and Italian Alps). In Engineering Geology for Society and Territory, 1, ed. Lollino, G., Manconi, A., Clague, J., Shan, W., and Chiarle, M.. Springer International Publishing, pp. 51–5.Google Scholar
De Meyer, A., Poesen, J., Isabirye, M., Deckers, J., and Raes, D. (2011). Soil erosion rates in tropical villages: a case study from Lake Victoria Basin, Uganda. Catena, 84, 8998.CrossRefGoogle Scholar
De Santisteban, L. M., Casalí, J., and López, J. J. (2006). Assessing soil erosion rates in cultivated areas of Navarre (Spain). Earth Surface Processes and Landforms, 31, 487506.CrossRefGoogle Scholar
De Wit, M. and Stanciewicz, J. (2006). Changes in surface water supply across Africa with predicted climate change. Science, 311, 1917–21.CrossRefGoogle ScholarPubMed
Dean, J. R., Leng, M. J., and Mackay, A. W. (2014). Is there an isotopic signature of the Anthropocene? The Anthropocene Review, 1, 276–87.CrossRefGoogle Scholar
Death, R. G., Fuller, I. C., and Macklin, M. G. (2015). Resetting the river template: the potential for climate‐related extreme floods to transform river geomorphology and ecology. Freshwater Biology, 60, 2477–96, doi: 10.1111/fwb.12639.CrossRefGoogle Scholar
DeBano, L. F. (2000). The role of fire and soil heating on water repellency in wildland environments: a review. Journal of Hydrology, 231/2, 195206.CrossRefGoogle Scholar
DeCarlo, T. M., Cohen, A. L., Barkley, H. C., Cobban, Q., Young, C., Shamberger, K. E., and Golbuu, Y. (2014). Coral macrobioerosion is accelerated by ocean acidification and nutrients. Geology, 43, 710.CrossRefGoogle Scholar
DeGeorges, A., Goreau, T. J., and Reilly, B. (2010). Land-sourced pollution with an emphasis on domestic sewage: lessons from the Caribbean and implications for coastal development on Indian Ocean and Pacific coral reefs. Sustainability, 2, 2919–49.CrossRefGoogle Scholar
Dehn, M. and Buma, J. (1999). Modelling future landslide activity based on general circulation models. Geomorphology, 30, 175–87.CrossRefGoogle Scholar
Dehn, M., Gurger, G., Buma, J., and Gasparetto, P. (2000). Impact of climate change on slope stability using expanded downscaling. Engineering Geology, 55, 193204.CrossRefGoogle Scholar
Deline, P., Gardent, M., Magnin, F,. and Ravanel, L. (2012). The morphodynamics of the Mont Blanc massif in a changing cryosphere: a comprehensive review. Geografiska Annaler, 94 A, 265–83.Google Scholar
Del Vecchio, S., Acosta, A., and Stanisci, A. (2013). The impact of Acacia saligna invasion on Italian coastal dune EC habitats. Comptes Rendus Biologies, 336, 364–9.CrossRefGoogle ScholarPubMed
DeLong, S. B., Johnson, J. P., and Whipple, K. X. (2014). Arroyo channel head evolution in a flash-flood–dominated discontinuous ephemeral stream system. Geological Society of America Bulletin, 126, 1683–701.CrossRefGoogle Scholar
DeLong, S. B., Pelletier, J. D., and Arnold, L. J. (2011). Late Holocene alluvial history of the Cuyama River, California, USA. Geological Society of America Bulletin, B30312–1.CrossRefGoogle Scholar
Delworth, T. L. and Zeng, F. (2014). Regional rainfall decline in Australia attributed to anthropogenic greenhouse gases and ozone levels. Nature Geoscience, 7, 583–7.CrossRefGoogle Scholar
Denevan, W. M. (1992). The pristine myth: the landscape of the Americas in 1492. Annals of the Association of American Geographers, 82, 369–85.CrossRefGoogle Scholar
Denevan, W. M. (2001). Cultivated Landscapes of Native Amazonia and the Andes. Oxford: Oxford University Press.CrossRefGoogle Scholar
Deo, R. C., Syktus, J. I., McAlpine, C. A., Lawrence, P. J., McGowan, H. A., and Phinn, S.R., (2009). Impact of historical land cover change on daily indices of climate extremes including droughts in eastern Australia. Geophysical Research Letters, 36, L08705, doi:10.1029/2009GL037666.CrossRefGoogle Scholar
Dickinson, W. R. (1999). Holocene sea-level record on Funafuti and potential impact of global warming on Central Pacific atolls. Quaternary Research, 51, 124–32.CrossRefGoogle Scholar
Dillehay, T. D. (2003). Tracking the first Americans. Nature, 425, 23–4.CrossRefGoogle ScholarPubMed
Dolan, R., Godfrey, P. J., and Odum, W. E. (1973). Man’s impact on the barrier islands of North Carolina. American Scientist, 61, 152–62.Google Scholar
Domínguez-Villar, D., Vázquez-Navarro, J. A., and Carrasco, R. M. (2012). Mid-Holocene erosive episodes in tufa deposits from Trabaque Canyon, central Spain, as a result of abrupt arid climate transitions. Geomorphology, 161, 1525.CrossRefGoogle Scholar
Doney, S. C. (2006). The dangers of ocean acidification. Scientific American, 294 (3), 3845.CrossRefGoogle ScholarPubMed
Dong, Z., Chen, G., He, X., Han, Z., and Wang, X. (2004). Controlling blown sand along the highway crossing the Taklimakan Desert. Journal of Arid Environments, 57, 329–44.CrossRefGoogle Scholar
Donkin, R. A. (1979). Agricultural Terracing in the Aboriginal New World. Tucson: University of Arizona Press.Google Scholar
Donnelly, L. J. (2009). A review of international cases of fault reactivation during mining subsidence and fluid abstraction. Quarterly Journal of Engineering Geology and Hydrogeology, 42, 7394.CrossRefGoogle Scholar
Donner, S. D. (2009). Coping with commitment: projected thermal stress on coral reefs under different future scenarios. Plos One, 4, e5712, 19.CrossRefGoogle ScholarPubMed
Doody, J. P. (2013). Coastal squeeze and managed realignment in southeast England, does it tell us anything about the future? Ocean and Coastal Management 79, 3441.Google Scholar
Doody, P. (ed.) (1984). Spartina anglica in Great Britain. Shrewsbury: Nature Conservancy Council.CrossRefGoogle Scholar
Doody, P. and Barnett, B. (eds). (1987). The Wash and its Environment. Peterborough: Nature Conservancy Council.Google Scholar
Doren, R. F., Richards, J. H., and Volin, J. C. (2009). A conceptual ecological model to facilitate understanding the role of invasive species in large-scale ecosystem restoration. Ecological Indicators, 9, S150–60.CrossRefGoogle Scholar
Dotterweich, M. (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.CrossRefGoogle Scholar
Dotterweich, M. (2013). The history of human-induced soil erosion: geomorphic legacies, early descriptions and research, and the development of soil conservation—a global synopsis. Geomorphology, 201, 134.CrossRefGoogle Scholar
Dotterweich, M., Ivester, A. H., Hanson, P. R., Larsen, D., and Dye, D. H. (2015). Natural and human induced prehistoric and historical soil erosion and landscape development in Southwestern Tennessee, USA. Anthropocene, 8, 624.CrossRefGoogle Scholar
Doughty, C. E., Wolf, A., and Field, C. B. (2010). Biophysical feedbacks between the Pleistocene megafauna extinction and climate: the first human‐induced global warming? Geophysical Research Letters, 37, doi: 10.1029/2010GL043985.CrossRefGoogle Scholar
Douglas, I. (1983). The Urban Environment. London: Arnold.Google Scholar
Douglas, I. (1996). The impact of land-use changes, especially logging, shifting cultivation, mining and urbanization on sediment yields in humid tropical Southeast Asia: a review with special reference to Borneo. IAHS Publications-Series of Proceedings and Reports-Intern Assoc Hydrological Sciences, 236, 463–72.Google Scholar
Douglas, I., and Lawson, N. (2001). The human dimensions of geomorphological work in Britain. Journal of Industrial Ecology, 4, 933.CrossRefGoogle Scholar
Douglas, I., and Lawson, N. (2003). Airport construction: materials use and geomorphic change. Journal of Air Transport Management, 9, 177–85.CrossRefGoogle Scholar
Douglas-Mankin, K. R., Srinivasan, R., and Arnold, J. G. (2010). Soil and Water Assessment Tool (SWAT) model: current developments and applications. Transactions of the American Society of Agricultural and Biological Engineers, 53, 1423–31.Google Scholar
Douville, H., Chauvin, F., Planton, S., Royer, J. F., Salas-Mélia, D., and Tyteca, S. (2002). Sensitivity of the hydrological cycle to increasing amounts of greenhouse gases and aerosols. Climate Dynamics, 20, 4568.Google Scholar
Downs, P. W. and Gregory, K. J. (2004). River Channel Management. Arnold: London.Google Scholar
Downs, P. W., Dusterhoff, S. R., and Sears, W. A. (2013). Reach-scale channel sensitivity to multiple human activities and natural events: lower Santa Clara River, California, USA. Geomorphology, 189, 121–34.CrossRefGoogle Scholar
Downward, S. and Skinner, K. (2005). Working rivers: the geomorphological legacy of English freshwater mills. Area, 37, 138–47.CrossRefGoogle Scholar
Doyle, M. W., Stanley, E. H., and Harbor, J. M. (2003). Channel adjustments following two dam removals in Wisconsin. Water Resources Research, 39, W1011, doi:10.1029/2002/WR001714.CrossRefGoogle Scholar
Draut, A. E. (2012). Effects of river regulation on aeolian landscapes, Colorado River, southwestern USA. Journal of Geophysical Research: Earth Surface, 117(F2), doi: 10.1029/2011JF002329.CrossRefGoogle Scholar
Draut, A. E. and Ritchie, A. C. (2013). Sedimentology of new fluvial deposits on the Elwha River, Washington, USA, formed during large‐scale dam removal. River Research and Applications, doi: 10.1002/rra.2724.CrossRefGoogle Scholar
Dreibrodt, S., Lubos, C., Terhorst, B., Damm, B., and Bork, H. R. (2010). Historical soil erosion by water in Germany: scales and archives, chronology, research perspectives. Quaternary International, 222, 8095.CrossRefGoogle Scholar
Drew, D. P. (1983). Accelerated soil erosion in a karst area: the Burren, western Ireland. Journal of Hydrology, 61, 113–24.CrossRefGoogle Scholar
Drewry, D. J. (1991). The response of the Antarctic ice sheet to climate change. In Antarctica and Global Climatic Change, ed. Harris, C. M. and Stonehouse, B.. London: Belhaven Press, pp. 90106.Google Scholar
Duarte, F., Jones, N., and Fleskens, L. (2008). Traditional olive orchards on sloping land: sustainability or abandonment? Journal of Environmental Management, 89, 8698.CrossRefGoogle Scholar
Dubois, R. N. (2013). How does a barrier shoreface respond to a sea-level rise? Journal of Coastal Research, 18, 612–28.CrossRefGoogle ScholarPubMed
Dufour, S. and Piégay, H. (2009). From the myth of a lost paradise to targeted river restoration: forget natural references and focus on human benefits. River Research and Applications, 25, 568–81.Google Scholar
Dukes, J. S. and Mooney, H. A. (2004). Disruption of ecosystem processes in western North America by invasive species. Revista Chilena de Historia Natural, 77, 411–37.CrossRefGoogle Scholar
Dunning, N. P. and Beach, T. (1994). Soil erosion, slope management, and ancient terracing in the Maya lowlands. Latin American Antiquity, 5, 5169.CrossRefGoogle Scholar
Du Toit, G. van N., Snyman, H. A., and Malan, P. J. (2009). Physical impact of grazing by sheep on soil parameters in the Nama Karoo subshrub/grass rangeland of South Africa. Journal of Arid Environments, 73, 804–10.CrossRefGoogle Scholar
Dupin, B., De Rouw, A., Phantahvong, K. B., and Valentin, C. (2009). Assessment of tillage erosion rates on steep slopes in northern Laos. Soil and Tillage Research, 103, 119–26.CrossRefGoogle Scholar
Durá-Gómez, I. and Talwani, P. (2010). Reservoir-induced seismicity associated with the Itoiz Reservoir, Spain: a case study. Geophysical Journal International, 181, 343–56.CrossRefGoogle Scholar
Durán Zuazo, V. H., Francia Martínez, J. R., and Martínez Raya, A. (2004). Impact of vegetative cover on runoff and soil erosion at hillslope scale in Lanjaron, Spain. The Environmentalist, 24, 3948.CrossRefGoogle Scholar
Dusar, B., Verstraeten, G., Notebaert, B., and Bakker, J. (2011). Holocene environmental change and its impact on sediment dynamics in the Eastern Mediterranean. Earth-Science Reviews, 108, 137–57.CrossRefGoogle Scholar
East, A. E., Pess, G. R., Bountry, J. A., Magirl, C. S., Ritchie, A. C., Logan, J. B., and Shafroth, P. B. (2014). Large-scale dam removal on the Elwha River, Washington, USA: river channel and floodplain geomorphic change. Geomorphology, 228, 765–86.Google Scholar
Edeso, J. M., Merino, A., Gonzalez, M. J., and Marauri, P. (1999). Soil erosion under different harvesting managements in steep forestlands from northern Spain. Land Degradation & Development, 10, 7988.3.0.CO;2-4>CrossRefGoogle Scholar
Edgeworth, M., deB Richter, D., Waters, C., Haff, P., Neal, C., and Price, S. J. (2015). Diachronous beginnings of the Anthropocene: the lower bounding surface of anthropogenic deposits. The Anthropocene Review, 2, 3358.CrossRefGoogle Scholar
Edinger, E. N. and Risk, M. J. (2013). Effect of land-based pollution on Central Java coral reefs. Journal of Coastal Development, 3, 593613.Google Scholar
Edwards, K. J. and Whittington, G. (2001). Lake sediments, erosion and landscape change during the Holocene in Britain and Ireland. Catena, 42, 143–73.CrossRefGoogle Scholar
Ehrenfeld, J. G. (2010). Ecosystem consequences of biological invasions. Annual Review of Ecology, Evolution, and Systematics, 41, 5980.CrossRefGoogle Scholar
Eitner, V. (1996). Geomorphological response to the East Frisian barrier islands to sea level rise: an investigation of past and future evolution. Geomorphology, 15, 5765.CrossRefGoogle Scholar
El Banna, M. M. and Frihy, O. E. (2009). Human-induced changes in the geomorphology of the northeastern coast of the Nile delta, Egypt. Geomorphology, 107, 72–8.CrossRefGoogle Scholar
Eldridge, D. J. (1998). Trampling of microphytic crusts on calcareous soils, and its impact on erosion under rain-impacted flow. Catena, 33, 221–39.Google Scholar
Eldridge, D. J. and Robson, A. D. (1997). Bladeploughing and exclosure influence soil properties in a semi-arid Australian woodland. Journal of Range Management, 50, 191–8.CrossRefGoogle Scholar
Elguindi, N. and Giorgi, F. (2006). Projected changes in the Caspian Sea level for the 21st century based on the latest AOGCM simulations. Geophysical Research Letters, 33, L08706, doi: 10.1029/2006GL025943.CrossRefGoogle Scholar
El Hariri, M., Abercrombie, R. E., Rowe, C. A., and Do Nascimento, A. F. (2010). The role of fluids in triggering earthquakes: observations from reservoir induced seismicity in Brazil. Geophysical Journal International, 181, 1566–74.CrossRefGoogle Scholar
Elliot, W. J., Hall, D. E., and Graves, S. R. (1999). Predicting sedimentation from forest roads. Journal of Forestry, 97, 23–9.Google Scholar
Ellis, E. C., Fuller, D. Q., Kaplan, J. O., and Lutters, W. G. (2013b). Dating the Anthropocene: towards an empirical global history of human transformation of the terrestrial biosphere. Elementa: Science of the Anthropocene, doi: 10.12952/journal.elementa.000018.CrossRefGoogle ScholarPubMed
Ellis, E. C., Kaplan, J. O., Fuller, D. Q., Vavrus, S., Goldewijk, K. K., and Verburg, P. H. (2013a). Used planet: a global history. Proceedings of the National Academy of Sciences, 110, 7978–85.CrossRefGoogle Scholar
Ellison, J. C. (2015). Vulnerability assessment of mangroves to climate change and sea-level rise impacts. Wetlands Ecology and Management, 23, 115–37.CrossRefGoogle Scholar
Ellsworth, W. L. (2013). Injection-induced earthquakes. Science, 341, doi: 10.1126/science.1225942.CrossRefGoogle ScholarPubMed
Elschot, K., Bouma, T. J., Temmerman, S., and Bakker, J. P. (2013). Effects of long-term grazing on sediment deposition and salt-marsh accretion rates. Estuarine, Coastal and Shelf Science, 133, 109–15.CrossRefGoogle Scholar
Elsner, J. B., Kossin, J. P., and Jagger, T. H. (2008). The increasing intensity of the strongest tropical cyclones. Nature, 455, 92–5.CrossRefGoogle ScholarPubMed
Emadodin, I., Reiss, S., and Bork, H. R. (2011). Colluviation and soil formation as geoindicators to study long-term environmental changes. Environmental Earth Sciences, 62, 1695–706.CrossRefGoogle Scholar
Emery, F. V. (1962). Moated settlements in England. Geography, 47, 378–88.Google Scholar
Engelstaedter, S., Tegen, I., and Washington, R. (2006). North African dust emissions and transport. Earth-Science Reviews, 79, 73100.CrossRefGoogle Scholar
Engstrom, D. R., Almendinger, J. E., and Wolin, J. A. (2009). Historical changes in sediment and phosphorus loading to the upper Mississippi River: mass-balance reconstructions from the sediments of Lake Pepin. Journal of Paleolimnology, 41, 563–88.CrossRefGoogle 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, 084010.CrossRefGoogle Scholar
Ericson, J. P., Vorosmarty, C. J., Dingham, L., Ward, L. G., and Meybeck, M. (2006). Effective sea-level rise and deltas: causes of change and human dimension implications. Global and Planetary Change, 50, 6382.CrossRefGoogle Scholar
Eriksson, M. G., Olley, J. M., and Payton, R. W. (2000). Soil erosion history in central Tanzania based on OSL dating of colluvial and alluvial hillslope deposits. Geomorphology, 36, 107128.CrossRefGoogle Scholar
Erlandson, J. M. (2013). Shell middens and other anthropogenic soils as global stratigraphic signatures of the Anthropocene. Anthropocene, 4, 2432.CrossRefGoogle Scholar
Erlandson, J. M. and Braje, T. J. (2014). Archaeology and the Anthropocene. Anthropocene, 4, 17.CrossRefGoogle Scholar
Erlandson, J. M., Rick, T. C., and Peterson, C. (2005). A geoarchaeological chronology of Holocene dune building on San Miguel Island, California. The Holocene, 15, 1227–35.Google Scholar
Erskine, W. D., Saynor, M. J., Chalmers, A., and Riley, S. J. (2012). Water, wind, wood, and trees: interactions, spatial variations, temporal dynamics, and their potential role in river rehabilitation. Geographical Research, 50, 6074.CrossRefGoogle Scholar
Escalente, S. A. and Pimentel, A .S. (2008). Coastal dune stabilization using geotextile tubes at Las Colorados. Geosynthetics, 26, 1624.Google Scholar
Esteves, L. S. (2014). Managed Realignment: a Viable Long-Term Coastal Management Strategy? Dordrecht: Springer.CrossRefGoogle Scholar
Etienne, D., Ruffaldi, P., Goepp, S., Ritz, F., Georges-Leroy, M., Pollier, B., and Dambrine, E. (2011). The origin of closed depressions in Northeastern France: a new assessment. Geomorphology, 126, 121–31.CrossRefGoogle Scholar
Evans, D. J., Ewertowski, M., Jamieson, S. S., and Orton, C. (2015). Surficial geology and geomorphology of the Kumtor Gold Mine, Kyrgyzstan: human impacts on mountain glacier landsystems. Journal of Maps, doi: 10.1080/17445647.2015.1071720.CrossRefGoogle Scholar
Evans, G. (2008). Man’s impact on the coastline. Journal of Iberian Geology, 34, 167–90.Google Scholar
Evans, K. G., Saynor, M. J., Willgoose, G. R., and Riley, S. J. (2000). Post‐mining landform evolution modelling: 1. Derivation of sediment transport model and rainfall–runoff model parameters. Earth Surface Processes and Landforms, 25, 743–63.3.0.CO;2-0>CrossRefGoogle Scholar
Evans, M. and Lindsay, J. (2010). High resolution quantification of gully erosion in upland peatlands at the landscape scale. Earth Surface Processes and Landforms, 35, 876–86.CrossRefGoogle Scholar
Evans, M. and Warburton, J. (2007). Geomorphology of Upland Peat. Oxford: Blackwell.CrossRefGoogle Scholar
Evans, M., Warburton, J., and Yang, J. (2006). Eroding blanket peat catchments: global and local implications of upland organic sediment budgets. Geomorphology 79, 4557.CrossRefGoogle Scholar
Evans, R. (1997). Soil erosion in the UK initiated by grazing animals: a need for a national survey. Applied Geography, 17, 127–41.CrossRefGoogle Scholar
Evans, R. (1998). The erosional impacts of grazing animals. Progress in Physical Geography, 22, 251–68.Google Scholar
Evans, R. (2005). Curtailing grazing-induced erosion in a small catchment and its environs, the Peak District, Central England. Applied Geography, 25, 8195.CrossRefGoogle Scholar
Evenari, M., Shanan, L., and Tadmor, N. (1982). The Negev: The Challenge of a Desert. Cambridge, Mass.: Harvard University Press.CrossRefGoogle Scholar
Fabi, G., Spagnolo, A., Bellan-Santini, D., Charbonnel, E., Çiçek, B. A., García, J. J. G., and Santos, M. N. D. (2011). Overview on artificial reefs in Europe. Brazilian Journal of Oceanography, 59(SPE1), 155–66.CrossRefGoogle Scholar
Fabricius, K. E. (2005). Effects of terrestrial runoff on the ecology of corals and coral reefs: review and synthesis. Marine Pollution Bulletin, 50, 125–46.CrossRefGoogle ScholarPubMed
Faith, J. T. (2014). Late Pleistocene and Holocene mammal extinctions on continental Africa. Earth-Science Reviews, 128, 105–21.CrossRefGoogle Scholar
Fan, B., Guo, L., Li, N., Chen, J., Lin, H., Zhang, X. and Ma, L. (2014). Earlier vegetation green-up has reduced spring dust storms. Scientific Reports, 4, 6749, doi: 10.1038/srep06749.CrossRefGoogle ScholarPubMed
Farley, K. A., Jobbágy, E. G., and Jackson, R. B. (2005). Effects of afforestation on water yield: a global synthesis with implications for policy. Global Change Biology, 11, 1565–76.CrossRefGoogle Scholar
Faulkner, H. (1995). Gully erosion associated with the expansion of unterraced almond cultivation in the coastal Sierra de Lujar, S. Spain. Land Degradation & Development, 6, 179200.CrossRefGoogle Scholar
Faulkner, H., Ruiz, J., Zukowskyj, P., and Downward, S. (2003). Erosion risk associated with rapid and extensive agricultural clearances on dispersive materials in southeast Spain. Environmental Science & Policy, 6, 115–27.CrossRefGoogle Scholar
Favis-Mortlock, D. and Boardman, J. (1995). Non-linear responses of soil erosion to climate change: modelling study on the UK South Downs. Catena, 25, 365–87.CrossRefGoogle Scholar
Favis-Mortlock, D. T. and Guerra, A. J. T. (1999). The implications of general circulation model estimates of rainfall for future erosion: a case study from Brazil. Catena, 37, 329–54.CrossRefGoogle Scholar
Fedick, S. L. (1994). Ancient Maya agricultural terracing in the upper Belize River area. Ancient Mesoamerica, 5, 107–27.CrossRefGoogle Scholar
Feeser, I. and O’Connell, M. (2009). Fresh insights into long‐term changes in flora, vegetation, land use and soil erosion in the karstic environment of the Burren, western Ireland. Journal of Ecology, 97, 1083–100.CrossRefGoogle Scholar
Fei, S., Phillips, J., and Shouse, M. (2014). Biogeomorphic impacts of invasive species. Annual Review of Ecology, Evolution, and Systematics, 45, 6987.CrossRefGoogle Scholar
Feld, C. K., Birk, S., Bradley, D. C., Hering, D., Kail, J., Marzin, A., and Friberg, N. (2011). From natural to degraded rivers and back again: a test of restoration ecology theory and practice. Advances in Ecological Research, 44, 119209.CrossRefGoogle Scholar
Fencl, J. S., Mather, M. E., Costigan, K. H., and Daniels, M. D. (2014). How big of an effect do small dams have? Using geomorphological footprints to quantify spatial impact of low-head dams and identify patterns of across-dam variation. PloS one, 10(11), e0141210.CrossRefGoogle Scholar
Feng, Q. Y., Liu, G. J., Meng, L., Fu, E. J., Zhang, H. R., and Zhang, K. F. (2008). Land subsidence induced by groundwater extraction and building damage level assessment—a case study of Datun, China. Journal of China University of Mining and Technology, 18, 556–60.CrossRefGoogle Scholar
Feola, S., Carranza, M. L., Schaminée, J. H. J., Janssen, J. A. M., and Acosta, A. T. R. (2011). EU habitats of interest: an insight into Atlantic and Mediterranean beach and foredunes. Biodiversity and Conservation, 20, 1457–68.CrossRefGoogle Scholar
Ferguson, G. and Gleeson, T. (2012). Vulnerability of coastal aquifers to groundwater use and climate change. Nature Climate Change, 2, 342–5.CrossRefGoogle Scholar
Fernandez, D. P., Neff, J. C., and Reynolds, R. L. (2008). Biogeochemical and ecological impacts of livestock grazing in semi-arid southeastern Utah, USA. Journal of Arid Environments, 72, 777–91.CrossRefGoogle Scholar
Fernández-Moya, J., Alvarado, A., Forsythe, W., Ramírez, L., Algeet-Abarquero, N., and Marchamalo-Sacristán, M. (2014). Soil erosion under teak (Tectona grandis Lf) plantations: general patterns, assumptions and controversies. Catena, 123, 236–42.CrossRefGoogle Scholar
Ferris, T. M. C., Lowther, K. A., and Smith, B. J. (1993). Changes in footpath degradation 1983–1992: a study of the Brandy Pad, Mourne Mountains. Irish Geography, 26, 133–40.CrossRefGoogle Scholar
Fidelibus, M. D., Gutiérrez, F., and Spilotro, G. (2011). Human-induced hydrogeological changes and sinkholes in the coastal gypsum karst of Lesina Marina area (Foggia Province, Italy). Engineering Geology, 118, 119.CrossRefGoogle Scholar
Field, M. E., Ogston, A. S., and Storlazzi, C. D. (2011). Rising sea level may cause decline of fringing coral reefs. Eos, 92, 273–80.CrossRefGoogle Scholar
Filin, S., Avni, Y., Baruch, A., Morik, S., Arav, R., and Marco, S. (2014). Characterization of land degradation along the receding Dead Sea coastal zone using airborne laser scanning. Geomorphology, 206, 403–20.CrossRefGoogle Scholar
Filip, F. and Giosan, L. (2014). Evolution of Chilia lobes of the Danube delta: reorganization of deltaic processes under cultural pressures. Anthropocene, 5, 6570.CrossRefGoogle Scholar
Fillenham, L. F. (1963). Holme Fen Post. Geographical Journal, 129, 502–3.CrossRefGoogle Scholar
Fisher, C. T., Pollard, H. P., Israde-Alcántara, I., Garduño-Monroy, V. H., and Banerjee, S. K. (2003). A reexamination of human-induced environmental change within the Lake Patzcuaro Basin, Michoacan, Mexico. Proceedings of the National Academy of Sciences, 100, 4957–62.CrossRefGoogle ScholarPubMed
Fitzhugh, T. W. and Vogel, R. M. (2011). The impact of dams on flood flows in the United States. River Research and Applications, 27, 1192–215.CrossRefGoogle Scholar
Fitzner, B., Heinrichs, K., and La Bouchardiere, D. (2002). Limestone weathering on historical monuments in Cairo, Egypt. Special Publication of the Geological Society of London, 205, 217–40.CrossRefGoogle Scholar
Flannigan, M., Cantin, A. S., de Groot, W. J., Wotton, M., Newbery, A., and Gowman, L. M. (2013). Global wildland fire season severity in the 21st century. Forest Ecology and Management, 294, 5461.CrossRefGoogle Scholar
Fleitmann, D., Dunbar, R. B., McCulloch, M., Mudelsee, M., Vuille, M., McClanahan, T. R., and Eggins, S. (2007). East African soil erosion recorded in a 300 year old coral colony from Kenya. Geophysical Research Letters, 34, doi: 10.1029/2006GL028525.CrossRefGoogle Scholar
Flenley, J. R. (1979). The Equatorial Rain Forest: A Geological History. London: Butterworth.Google Scholar
Fleskens, L. and Stroosnijder, L. (2007). Is soil erosion in olive groves as bad as often claimed? Geoderma, 141, 260–71.CrossRefGoogle Scholar
Fletcher, S., Bateman, P., and Emery, A. (2011). The governance of the Boscombe artificial surf reef, UK. Land Use Policy, 28, 395401.CrossRefGoogle Scholar
Flor-Blanco, G., Pando, L., Morales, J. A., and Flor, G. (2015). Evolution of beach–dune fields systems following the construction of jetties in estuarine mouths (Cantabrian coast, NW Spain). Environmental Earth Sciences, 73, 1317–30.CrossRefGoogle Scholar
Foley, R. A. and Lahr, M. M. (2015). Lithic landscapes: early human impact from stone tool production on the central Saharan environment. PloS one, 10(3), e0116482.CrossRefGoogle ScholarPubMed
Foley, S. F., Gronenborn, D., Andreae, M. O., Kadereit, J. W., Esper, J., Scholz, D., and Crutzen, P. J. (2013). The Palaeoanthropocene – The beginnings of anthropogenic environmental change. Anthropocene, 3, 83–8.CrossRefGoogle Scholar
Foltz, R. B., Copeland, N. S., and Elliot, W. J. (2009). Reopening abandoned forest roads in northern Idaho, USA: quantification of runoff, sediment concentration, infiltration, and interrill erosion parameters. Journal of Environmental Management, 90, 2542–50.CrossRefGoogle ScholarPubMed
Ford, D. C. and Williams, P. W. (1989). Karst Geomorphology and Hydrology. London: Unwin Hyman.CrossRefGoogle Scholar
Ford, J. R., Price, S. J., Cooper, A. H., and Waters, C. N. (2014). An assessment of lithostratigraphy for anthropogenic deposits. Geological Society, London, Special Publications, 395, 5589.CrossRefGoogle Scholar
Fornasiero, A., Putti, M.., Teatini, P., Ferraris, S., Rizzetto, F., and Tosi, L. (2003). Monitoring of hydrological parameters related to peat oxidation in a subsiding coastal basin south of Venice, Italy. International Association of Hydrological Sciences, Publication, 278, 458–62.Google Scholar
Foster, G. C., Chiverrell, R. C., Thomas, G. S. P., Marshall, P., and Hamilton, D. (2009). Fluvial development and the sediment regime of the lower Calder, Ribble catchment, northwest England. Catena, 77, 8195.CrossRefGoogle Scholar
Foster, I. D. L., Dearing, J. A., and Appleby, R. G. (1986). Historical trends in catchment sediment yields: a case study in reconstruction from lake-sediment records in Warwickshire, UK. Hydrological Science Journal, 31, 427–43.CrossRefGoogle Scholar
Foster, I. D. L., Collins, A. L., Naden, P. S., Sear, D. A., Jones, J. I., and Zhang, Y. (2011). The potential for paleolimnology to determine historic sediment delivery to rivers. Journal of Paleolimnology, 45, 287306.CrossRefGoogle Scholar
Foulds, S. A. and Macklin, M. G. (2006). Holocene land-use change and its impact on river basin dynamics in Great Britain and Ireland. Progress in Physical Geography, 30, 589604.CrossRefGoogle Scholar
Fox, D. M., Martin, N., Carrega, P., Andrieu, J., Adnès, C., Emsellem, K., and Fox, E. A. (2015). Increases in fire risk due to warmer summer temperatures and wildland urban interface changes do not necessarily lead to more fires. Applied Geography, 56, 112.CrossRefGoogle Scholar
Fox, H. L. (1976). The urbanizing river: a case study in the Maryland piedmont. In Geomorphology and Engineering, ed. Coates, D. R.. Stroudsburg: Dowden, Hutchinson and Ross, pp. 245–71.Google Scholar
Francia Martínez, J. R., Durán Zuazo, V. H., and Martínez Raya, A. (2006). Environmental impact from mountainous olive orchards under different soil-management systems (SE Spain). Science of the Total Environment, 358, 4660.CrossRefGoogle ScholarPubMed
Frankl, A., Nyssen, J., De Dapper, M., Haile, M., Billi, P., Munro, R. N., and Poesen, J. (2011). Linking long-term gully and river channel dynamics to environmental change using repeat photography (Northern Ethiopia). Geomorphology, 129, 238–51.CrossRefGoogle Scholar
Frans, C., Istanbulluoglu, E., Mishra, V., Munoz‐Arriola, F., and Lettenmaier, D. P. (2013). Are climatic or land cover changes the dominant cause of runoff trends in the Upper Mississippi River Basin? Geophysical Research Letters, 40, 1104–10.CrossRefGoogle Scholar
Frazier, T. G., Wood, N., Yarnal, B., and Bauer, D. H. (2010). Influence of potential sea level rise on societal vulnerability to hurricane storm-surge hazards, Sarasota County, Florida. Applied Geography, 30, 490505.CrossRefGoogle Scholar
French, C. A. and Whitelaw, T. M. (1999). Soil erosion, agricultural terracing and site formation processes at Markiani, Amorgos, Greece: the micromorphological perspective. Geoarchaeology, 14, 151–89.3.0.CO;2-R>CrossRefGoogle Scholar
French, C., Periman, R., Cummings, L. S., Hall, S., Goodman-Elgar, M., and Boreham, J. (2009). Holocene alluvial sequences, cumulic soils and fire signatures in the Middle Rio Puerco Basin at Guadelupe Ruin, New Mexico. Geoarchaeology, 24, 638–76.CrossRefGoogle Scholar
French, H. M. (1976). The Periglacial Environment. London: Longman.Google Scholar
French, J. R., Spencer, T., and Reed, D. J. (eds). (1995). Geomorphic responses to sea level rise: existing evidence and future impacts. Earth Surface Processes and Landforms, 20, 16.CrossRefGoogle Scholar
Friedman, J. M., Auble, G. T., Shafroth, P. B., Scott, M. L., Merigliano, M. F., Freehling, M. D., and Griffin, E. R. (2005). Dominance of non-native riparian trees in western USA. Biological Invasions, 7, 747–51.CrossRefGoogle Scholar
Friedman, J. M., Vincent, K. R., Griffin, E. R., Scott, M. L., Shafroth, P. B., and Auble, G. T. (2014). Processes of arroyo filling in northern New Mexico, USA. Geological Society of America Bulletin, 127, 621–40.Google Scholar
Friess, D. A., Möller, I., Spencer, T., Smith, G. M., Thomson, A. G., and Hill, R. A. (2014). Coastal saltmarsh managed realignment drives rapid breach inlet and external creek evolution, Freiston Shore (UK). Geomorphology, 208, 2233.CrossRefGoogle Scholar
Fuchs, M., Lang, A., and Wagner, G. A. (2004). The history of Holocene soil erosion in the Phlious Basin, NE Peloponnese, Greece, based on optical dating. The Holocene, 14, 334–45.CrossRefGoogle Scholar
Fuller, I. C., Macklin, M. G., and Richardson, J. M. (2015). The geography of the Anthropocene in New Zealand: differential river catchment response to human impact. Geographical Research, doi: 10.1111/1745-5871.12121.CrossRefGoogle Scholar
Gabet, E. J. (2014). Fire increases dust production from chaparral soils. Geomorphology 217, 182–92.CrossRefGoogle Scholar
Gale, S. J. and Hoare, P. G. (2012). The stratigraphic status of the Anthropocene. The Holocene, 22, 1491–4.CrossRefGoogle Scholar
Gallardo, A. H., Marui, A., Takeda, S., and Okuda, F. (2009). Groundwater supply under land subsidence constrains in the Nobi Plain. Geosciences Journal, 13, 151–9.CrossRefGoogle Scholar
Gallego-Sala, A. V. and Prentice, I. C. (2012). Blanket peat biome endangered by climate change. Nature Climate Change, 3, 152–5.CrossRefGoogle Scholar
Galloway, D. L. and Burbey, T. J. (2011). Review: regional land subsidence accompanying groundwater extraction. Hydrogeology Journal, 19, 1459–86.Google Scholar
Gambolati, G., Ricceri, G., Bertoni, W., Brighenti, G., and Vuillermin, E. (1991). Mathematical simulation of the subsidence of Ravenna. Water Resources Research, 27, 2899–918.CrossRefGoogle Scholar
Gambolati, G., Putti, M., Teatini, P., Camporese, M., Ferraris, S., Stori, G. G., and Tosi, L. (2005). Peat land oxidation enhances subsidence in the Venice watershed. Eos, Transactions American Geophysical Union, 86(23), 217–20.CrossRefGoogle Scholar
Gambolati, G., Putti, M., Teatini, P., and Stori, G. G. (2006). Subsidence due to peat oxidation and impact on drainage infrastructures in a farmland catchment south of the Venice Lagoon. Environmental Geology, 49, 814–20.CrossRefGoogle Scholar
Gao, J. H., Xu, X., Jia, J., Kettner, A. J., Xing, F., Wang, Y. P., and Gao, S. (2015). A numerical investigation of freshwater and sediment discharge variations of Poyang Lake catchment, China over the last 1000 years. The Holocene, 25, 1470–82.CrossRefGoogle Scholar
Garbutt, R. A., Reading, C. J., Wolters, M., Gray, A. J., and Rothery, P. (2006). Monitoring the development of intertidal habitats on former agricultural land after the managed realignment of coastal defences at Tollesbury, Essex, UK. Marine Pollution Bulletin, 53, 155–64.CrossRefGoogle ScholarPubMed
García-Moreno, I. and Mateos, R. M. (2011). Sinkholes related to discontinuous pumping: susceptibility mapping based on geophysical studies. The case of Crestatx (Majorca, Spain). Environmental Earth Sciences, 64, 523–37.CrossRefGoogle Scholar
García-Ruiz, J., Lasanta, T., and Alberto, F. (1997). Soil erosion by piping in irrigated fields. Geomorphology, 20, 269–78.CrossRefGoogle Scholar
García-Ruiz, J. M. (2010). The effects of land uses on soil erosion in Spain: a review. Catena, 81, 111.Google Scholar
García-Ruiz, J. M. and Valero-Garcés, B. L. (1998). Historical geomorphic processes and human activities in the Central Spanish Pyrenees. Mountain Research and Development, 10, 267–79.CrossRefGoogle Scholar
García-Ruiz, J. M., Beguería, S., Alatorre, L. C., and Puigdefábregas, J. (2010). Land cover changes and shallow landsliding in the flysch sector of the Spanish Pyrenees. Geomorphology, 124, 250–9.CrossRefGoogle Scholar
García-Ruiz, J. M., López-Moreno, J. I., Vicente-Serrano, S. M., Lasanta–Martínez, T., and Beguería, S. (2011). Mediterranean water resources in a global change scenario. Earth-Science Reviews, 105, 121–39.CrossRefGoogle Scholar
Gardner, R. (2009). Trees as technology: planting shelterbelts on the Great Plains. History and Technology, 25, 325–41.CrossRefGoogle Scholar
Garland, G. G., Hudson, C., and Blackshaw, J. (1985). An approach to the study of path erosion in the Natal Drakensberg, a mountain wilderness area. Environmental Conservation, 12, 337–42.CrossRefGoogle Scholar
Gautney, J. R. and Holliday, T. W. (2015). New estimations of habitable land area and human population size at the last glacial maximum. Journal of Archaeological Science, 58, 103–12.CrossRefGoogle Scholar
Gedan, K. B., Altieri, A. H., and Bertness, M. D. (2011). Uncertain future of New England salt marshes. Marine Ecology Progress Series, 434, 229–37.CrossRefGoogle ScholarPubMed
Gedan, K. B., Silliman, B. R., and Bertness, M. D. (2009). Centuries of human-driven change in salt marsh ecosystems. Annual Review of Marine Science, 1, 117–41.CrossRefGoogle Scholar
Gedney, N., Cox, P. M., Betts, R. A., Boucher, O. Huntingford, C., and Stott, P. A. (2006). Detection of a direct carbon dioxide effect in continental river runoff records. Nature, 439, 835–8.CrossRefGoogle ScholarPubMed
Geertsema, M., Clague, J. J., Schwab, J. W., and Evans, S. G. (2006). An overview of recent large catastrophic landslides in northern British Columbia, Canada. Engineering Geology, 83, 120–43.CrossRefGoogle Scholar
Gelfenbaum, G., Stevens, A. W., Miller, I., Warrick, J. A., Ogston, A. S., and Eidam, E. (2015). Large-scale dam removal on the Elwha River, Washington, USA: coastal geomorphic change. Geomorphology, 246, 649–68.CrossRefGoogle Scholar
Gellatly, A. F., Whalley, W. B., and Gordon, J. E. (1986). Footpath deterioration in the Lyngen Peninsula, north Norway. Mountain Research and Development, 6, 167–76.CrossRefGoogle Scholar
Gellis, A. C., Pavich, M. J., Ellwein, A. L., Aby, S., Clark, I., Wieczorek, M. E., and Viger, R. (2012). Erosion, storage, and transport of sediment in two subbasins of the Rio Puerco, New Mexico. Geological Society of America Bulletin, 124, 817–41.CrossRefGoogle Scholar
Genevois, R. and Ghirotti, M. (2005). The 1963 Vaiont landslide. Giornale di Geologia Applicata, 1, 4152.Google Scholar
Gerasimov, I. P. (1979). Anthropogene and its major problem. Boreas, 8, 2330.CrossRefGoogle Scholar
Germer, S., Neill, C., Krusche, A. V., and Elsenbeer, H. (2010). Influence of land-use on near-surface hydrological processes: undisturbed forest to pasture. Journal of Hydrology, 380, 473–80.CrossRefGoogle Scholar
Ghoneim, E., Mashaly, J., Gamble, D., Halls, J., and AbuBakr, M. (2015). Nile Delta exhibited a spatial reversal in the rates of shoreline retreat on the Rosetta promontory comparing pre-and post-beach protection. Geomorphology, 228, 114.CrossRefGoogle Scholar
Gifford, G. F. and Hawkins, R. H. (1978). Hydrologic impact of grazing on infiltration: a critical review. Water Resources Research, 14, 305–13.CrossRefGoogle Scholar
Giguet-Covex, C., Pansu, J., Arnaud, F., Rey, P. J., Griggo, C., Gielly, L., and Taberlet, P. (2014). Long livestock farming history and human landscape shaping revealed by lake sediment DNA. Nature communications, 5, doi:10.1038/ncomms4211.CrossRefGoogle ScholarPubMed
Gilbert, G. K. (1917). Hydraulic-mining débris in the Sierra Nevada. United States Geological Survey Professional Paper, 105.Google Scholar
Gilbertson, D. D., Schwenninger, , Kemp, R. A., and Rhodes, E. J. (1999). Sand-drift and soil formation along an exposed North Atlantic coastline: 14,000 years of diverse geomorphological, climatic and human impacts. Journal of Archaeological Science, 26, 439–69.CrossRefGoogle Scholar
Gill, R. A. (2007). Influence of 90 years of protection from grazing on plant and soil processes in the subalpine of the Wasatch Plateau, USA. Rangeland Ecology & Management, 60, 8898.CrossRefGoogle Scholar
Gill, T. E. (1996). Eolian sediments generated by anthropogenic disturbance of playas: human impacts on the geomorphic system and geomorphic impacts on the human system. Geomorphology, 17, 207–28.CrossRefGoogle Scholar
Gillies, J. A. (2013). Fundamentals of aeolian sediment transport: dust emissions and transport near surface. In Treatise on Geomorphology, ed. Shroder, J. (Editor in Chief), Lancaster, N., Sherman, D. J., and Baas, A. C. W.. Academic Press, San Diego, CA, 11, Aeolian Geomorphology, pp. 4363.CrossRefGoogle Scholar
Gillies, J. A., Etyemezian, V., Kuhns, H., Nikolic, D., and Gillette, D. A. (2005). Effect of vehicle characteristics on unpaved road dust emissions. Atmospheric Environment, 39, 2341–7.CrossRefGoogle Scholar
Gillies, J. A., Kuhns, H., Engelbrecht, J. P., Uppapalli, S., Etyemezian, V., and Nikolich, G. (2007). Particulate emissions from US Department of Defense artillery backblast testing. Journal of the Air & Waste Management Association, 57, 551–60.Google ScholarPubMed
Gilvear, D. J., Waters, T. M., and Milner, A. M. (1995). Image analysis of aerial photography to quantify changes in channel morphology and instream habitat following placer mining in interior Alaska. Freshwater Biology, 34, 389–98.CrossRefGoogle Scholar
Ginoux, P., Prospero, J. M., Gill, T. E., Hsu, N. C., and Zhao, M. (2012). Global-scale attribution of anthropogenic and natural dust sources and their emission rates based on MODIS deep blue aerosol products. Reviews of Geophysics, 50, RG3005, doi: 10.1029/2012RG000388.CrossRefGoogle Scholar
Giosan, L., Syvitski, J., Constantinescu, S., and Day, J. (2014). Climate change: protect the world’s deltas. Nature, 516, 31–3.CrossRefGoogle ScholarPubMed
Giraldéz, J. V., Ayuso, J. L., Garcia, A., Lopez, J. G., and Roldán, J. (1988). Water harvesting strategies in the semiarid climate of southeastern Spain. Agricultural Water Management, 14, 253–63.CrossRefGoogle Scholar
Gislason, S. R., Oelkers, E. H., Eiriksdottir, E. S., Kardjilov, M. I., Gisladottir, G., Sigfusson, B., and Oskarsson, N. (2009). Direct evidence of the feedback between climate and weathering. Earth and Planetary Science Letters, 277, 213–22.CrossRefGoogle Scholar
Glacken, C. (1967). Traces on the Rhodian Shore: Nature and Culture in Western Thought from Ancient Times to the End of the Eighteenth Century. Berkeley: University of California Press.CrossRefGoogle Scholar
Glade, T. (2003). Landslide occurrence as a response to land use change: a review of evidence from New Zealand. Catena, 51, 297314.CrossRefGoogle Scholar
Glikson, A. (2013). Fire and human evolution: the deep-time blueprints of the Anthropocene. Anthropocene, 3, 8992.CrossRefGoogle Scholar
Glowacka, E., Gonzalez, J., and Nava, F. A. (2000). Subsidence in Cerro Prieto geothermal field, Baja California, Mexico. In Proceedings World Geothermal Congress, Kyushu, Japan: pp. 591–6.Google Scholar
Goddéris, Y. and Brantley, S. L. (2013). Earthcasting the future Critical Zone. Elementa: Science of the Anthropocene, 1(1), 000019.Google Scholar
Godoy, J. M., Padovani, C. R., Guimarães, J. R., Pereira, J. C., Vieira, L. M., Carvalho, Z. L., and Galdino, S. (2002). Evaluation of the siltation of River Taquari, Pantanal, Brazil, through 210Pb geochronology of floodplain lake sediments. Journal of the Brazilian Chemical Society, 13, 71–7.CrossRefGoogle Scholar
Godoy, M. D. P. and de Lacerda, L. D. (2013). River-island morphological response to basin land-use change within the Jaguaribe River Estuary, NE Brazil. Journal of Coastal Research, 30, 399410.Google Scholar
Goebel, T., Waters, M. R. and O’Rourke, D. H. (2008). The late Pleistocene dispersal of modern humans in the Americas. Science, 319, 1497–502.CrossRefGoogle ScholarPubMed
Goelles, T., Bøggild, C. E., and Greve, R. (2015). Ice sheet mass loss caused by dust and black carbon accumulation. The Cryosphere Discussions, 9, 2563–96.Google Scholar
Goldewijk, K. K. (2001). Estimating global land use change over the past 300 years: the HYDE database, Global Biogeochemical Cycles, 15, 417–33.CrossRefGoogle Scholar
Golomb, B. and Eder, H. M. (1964). Landforms made by man. Landscape, 14, 47.Google Scholar
Golosov, V. N., Sosin, P. M., Belyaev, V. R., Wolfgramm, B., and Khodzhaev, S. I. (2015). Effect of irrigation-induced erosion on the degradation of soils in river valleys of the alpine Pamir. Eurasian Soil Science, 48, 325–35.CrossRefGoogle Scholar
Gómez, J. A., Giráldez, J. V., and Fereres, E. (2005). Water erosion in olive orchards in Andalusia (Southern Spain): a review. Geophysical Research Abstracts, 7, 08406.Google Scholar
Gómez, J. A., Battany, M., Renschler, C. S., and Fereres, E. (2003). Evaluating the impact of soil management on soil loss in olive orchards. Soil Use and Management, 19, 127–34.CrossRefGoogle Scholar
Gómez, J. A., Romero, P., Giráldez, J. V., and Fereres, E. (2004). Experimental assessment of runoff and soil erosion in an olive grove on a Vertic soil in southern Spain as affected by soil management. Soil Use and Management, 20, 426–31.CrossRefGoogle Scholar
Gómez, J. A., Guzmán, M. G., Giráldez, J. V., and Fereres, E. (2009). The influence of cover crops and tillage on water and sediment yield, and on nutrient, and organic matter losses in an olive orchard on a sandy loam soil. Soil and Tillage Research, 106, 137–44.CrossRefGoogle Scholar
Gómez, J. A., Llewellyn, C., Basch, G., Sutton, P. B., Dyson, J. S., and Jones, C. A. (2011). The effects of cover crops and conventional tillage on soil and runoff loss in vineyards and olive groves in several Mediterranean countries. Soil Use and Management, 27, 502–14.CrossRefGoogle Scholar
Gómez-Pina, G., Munoz-Pérez, J. J., Ramírez, J. L., and Ley, C. (2002). Sand dune management problems and techniques, Spain. Journal of Coastal Research S1 36, 325–32.Google Scholar
Gonzalez, M. A. (2001). Recent formation of arroyos in the Little Missouri Badlands of southwestern Dakota. Geomorphology, 38, 6384.CrossRefGoogle Scholar
González-Amuchastegui, M. J. and Serrano, E. (2015). Tufa buildups, landscape evolution and human impact during the Holocene in the Upper Ebro Basin. Quaternary International, 364, 5464.CrossRefGoogle Scholar
Goode, J. R., Luce, C. H., and Buffington, J. M. (2012). Enhanced sediment delivery in a changing climate in semi-arid mountain basins: implications for water resource management and aquatic habitat in the northern Rocky Mountains. Geomorphology, 139–140, 115.CrossRefGoogle Scholar
Goossens, D., Buck, B., and McLaurin, B. (2012). Contributions to atmospheric dust production of natural and anthropogenic emissions in a recreational area designated for off-road vehicular activity (Nellis Dunes, Nevada, USA). Journal of Arid Environments, 78, 8099.CrossRefGoogle Scholar
Gornitz, V., Couch, S., and Hartig, E. K. (2002). Impacts of sea level rise in the New York City metropolitan area. Global and Planetary Change, 32, 6188.CrossRefGoogle Scholar
Gornitz, V., Rosenzweig, C., and Hillel, D. (1997). Effects of anthropogenic intervention in the land hydrologic cycle on global sea level rise. Global and Planetary Change, 14, 147–61.CrossRefGoogle Scholar
Gottschalk, L. C. (1945). Effects of soil erosion on navigation in Upper Chesapeake Bay. Geographical Review, 35, 219–38.CrossRefGoogle Scholar
Goudie, A. S. (1973). Duricrusts of Tropical and Subtropical Landscapes. Oxford: Clarendon Press.Google Scholar
Goudie, A.S. (1977). Sodium sulphate weathering and the disintegration of Mohenjo-Daro, Pakistan. Earth Surface Processes, 2, 7586.CrossRefGoogle Scholar
Goudie, A.S. (1978). Dust storms and their geomorphological implications. Journal of Arid Environments, 1, 291310.CrossRefGoogle Scholar
Goudie, A. S. (1983). Dust storms in space and time. Progress in Physical Geography, 7, 502–30.Google Scholar
Goudie, A. S. (1996). Geomorphological ‘hotspots’ and global warming. Interdisciplinary Science Reviews, 21, 253–9.CrossRefGoogle Scholar
Goudie, A.S. (2013a). Arid and Semi-arid Geomorphology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Goudie, A.S. (2013b). The Human Impact on the Natural Environment (7th edn). Oxford: Wiley-Blackwell.Google Scholar
Goudie, A.S. (2014). Desert dust and human health disorders. Environment International, 63, 101–13.CrossRefGoogle ScholarPubMed
Goudie, A. and Seely, M. (2011). World Heritage desert landscapes: potential priorities for the recognition of desert landscapes and geomorphological sites on the World Heritage List. IUCN World Heritage Studies, 9.CrossRefGoogle Scholar
Goudie, A. S. and Middleton, N. J. (1992). The changing frequency of dust storms through time. Climate Change, 20, 197225.Google Scholar
Goudie, A. S. and Middleton, N.J. (2006) Desert Dust in the Global System. Berlin and Heidelberg: Springer.Google Scholar
Goudie, A. S. and Viles, H. A. (1997). Salt Weathering Hazards. Chichester: Wiley.Google Scholar
Goudie, A. S. and Viles, H. A. (2012). Weathering and the global carbon cycle: geomorphological perspectives. Earth-Science Reviews, 113, 5971.CrossRefGoogle Scholar
Goudie, A. S. and Viles, H. A. (2015). Landscapes and Landforms of Namibia. Springer: Dordrecht.CrossRefGoogle Scholar
Goudie, A. S., Viles, H. A., and Pentecost, A. (1993). The late-Holocene tufa decline in Europe. The Holocene, 3, 181–6.CrossRefGoogle Scholar
Gourou, P. (1961). The Tropical World (3rd edn). London: Longman.Google Scholar
Goutal, N., Keller, T., Défossez, P., and Ranger, J. (2013). Soil compaction due to heavy forest traffic: measurements and simulations using an analytical soil compaction model. Annals of Forest Science, 70, 545–56.CrossRefGoogle Scholar
Gowlett, J. A. J., Harris, J. W. K., Walton, D., and Wood, B. A. (1981). Early archaeological sites, hominid remains and traces of fire from Chesowanja, Kenya. Nature, 284, 125–9.Google Scholar
Graber, E. R., Fine, P., and Levy, G. J. (2006). Soil stabilization in semiarid and arid land agriculture. Journal of Materials in Civil Engineering, 18, 190205.CrossRefGoogle Scholar
Gradziński, M., Hercman, H., Jaśkiewicz, M., and Szczurek, S. (2013). Holocene tufa in the Slovak Karst: facies, sedimentary environments and depositional history. Geological Quarterly, 57, 769–88.CrossRefGoogle Scholar
Graf, J. B., Webb, R. H., and Hereford, R. (1991). Relation of sediment load and flood-plain formation to climatic variability, Paria River drainage basin, Utah and Arizona. Bulletin of the Geological Society of America, 103, 1405–15.2.3.CO;2>CrossRefGoogle Scholar
Graf, W. L. (1977). Network characteristics in suburbanizing streams. Water Resources Research, 13, 459–63.CrossRefGoogle Scholar
Graf, W. L. (1979). Mining and channel response. Annals of the Association of American Geographers, 69, 262–75.CrossRefGoogle Scholar
Grainger, S. and Conway, D. (2014). Climate change and international river boundaries: fixed points in shifting sands. Wiley Interdisciplinary Reviews: Climate Change, 5, 835–48.Google Scholar
Grant, G. E. and Lewis, S. L. (2015). The remains of the dam: What have we learned from 15 years of US dam removals? Engineering Geology for Society and Territory, 3, 31–5.CrossRefGoogle Scholar
Grantham, T. E., Figueroa, R., and Prat, N. (2013). Water management in mediterranean river basins: a comparison of management frameworks, physical impacts, and ecological responses. Hydrobiologia, 719, 451–82.CrossRefGoogle Scholar
Grattan, J. P., Gilbertson, D. D., and Hunt, C. O. (2007). The local and global dimensions of metalliferous pollution derived from a reconstruction of an eight thousand year record of copper smelting and mining at a desert-mountain frontier in southern Jordan. Journal of Archaeological Science, 34, 83110.CrossRefGoogle Scholar
Gray, M. (2013). Geodiversity: Valuing and Conserving Abiotic Nature (2nd edn). Chichester: John Wiley & Sons.Google Scholar
Grayson, R., Holden, J., and Rose, R. (2010). Long-term change in storm hydrographs in response to peatland vegetation change. Journal of Hydrology, 389, 336–43.CrossRefGoogle Scholar
Greenfield, H. J. (2010). The secondary products revolution: the past, the present and the future. World Archaeology, 42, 2954.CrossRefGoogle Scholar
Greenwood, P. and Kuhn, N. J. (2014). Does the invasive plant, Impatiens glandulifera, promote soil erosion along the riparian zone? An investigation on a small watercourse in northwest Switzerland. Journal of Soils and Sediments, 14, 637–50.CrossRefGoogle Scholar
Gregory, K. J. (1997). Fluvial Geomorphology of Great Britain. London: Chapman and Hall.CrossRefGoogle Scholar
Gregory, J. M., White, N. J., Church, J. A., Bierkens, M. F. P., Box, J. E., Van den Broeke, M. R., and Van de Wal, R. S. W. (2013). Twentieth-Century global-mean sea level rise: is the whole greater than the sum of the parts? Journal of Climate, 26, 4476–99.CrossRefGoogle Scholar
Griffith, M. B., Norton, S. B., Alexander, L. C., Pollard, A. L., and LeDuc, S. D. (2012). The effects of mountaintop mines and valley fills on the physicochemical quality of stream ecosystems in the central Appalachians: a review. Science of the Total Environment, 417, 112.CrossRefGoogle ScholarPubMed
Grimm, N. B., Chacon, A., Dahm, C. N., Hostetler, S. W., Lind, O. T., Starkweather, P. L., and Wertsbaugh, W. W. (1997). Sensitivity of aquatic ecosystems to climatic and anthropogenic changes: the Basin and Range, American south-west and Mexico. In Freshwater Ecosystems and Climate Change in North America: a Regional Assessment, ed. Cushing, C. E.. Chichester: Wiley, pp. 205–23.Google Scholar
Grinsted, A., Moore, J. C., and Jevrejeva, S. (2013). Projected Atlantic hurricane surge threat from rising temperatures. Proceedings of the National Academy of Sciences, 110, 5369–73.CrossRefGoogle ScholarPubMed
Grossi, C. M., Brimblecombe, P., and Harris, I. (2007). Predicting long term freeze–thaw risks on Europe built heritage and archaeological sites in a changing climate. Science of the Total Environment, 377, 273–81.CrossRefGoogle Scholar
Grossi, C. M., Bonazza, A., Brimblecombe, P., Harris, I., and Sabbioni, C. (2008). Predicting twenty-first century recession of architectural limestone in European cities. Environmental Geology, 56, 455–61.CrossRefGoogle Scholar
Grossi, C. M., Brimblecombe, P., Menéndez, B., Benavente, D., Harris, I., and Déqué, M. (2011). Climatology of salt transitions and implications for stone weathering. Science of the Total Environment, 409, 2577–85.CrossRefGoogle ScholarPubMed
Grove, A. T. and Rackham, O. (2001). The Nature of Mediterranean Europe: an Ecological History. New Haven and London: Yale University Press.Google Scholar
Grove, A. T. and Sutton, J. E. G. (1989). Agricultural terracing south of the Sahara. Azania: Journal of the British Institute in Eastern Africa, 24, 113–22.CrossRefGoogle Scholar
Grove, R. H. (1996). Green Imperialism: Colonial Expansion, Tropical Island Edens and the Origins of Environmentalism, 1600–1860. Cambridge: Cambridge University Press.Google Scholar
Grove, R. H. and Damodaran, V. (2006). Imperialism, intellectual networks and environmental change. Economic and Social Weekly, October 14th, 4345–54.Google Scholar
Grünthal, G. (2014). Induced seismicity related to geothermal projects versus natural tectonic earthquakes and other types of induced seismic events in central Europe. Geothermics. 52, 2235.CrossRefGoogle Scholar
Guha, S. K. (ed.) (2000). Induced Earthquakes. Dordrecht: Kluwer.CrossRefGoogle Scholar
Gumilar, I., Abidin, H. Z., Hutasoit, L. M., Hakim, D. M., Sidiq, T. P., and Andreas, H. (2015). Land subsidence in Bandung Basin and its possible caused factors. Procedia Earth and Planetary Science, 12, 4762.CrossRefGoogle Scholar
Gunnell, Y. and Krishnamurthy, A. (2003). Past and present status of runoff harvesting systems in dryland peninsular India: a critical review. Ambio, 32, 320–4.CrossRefGoogle ScholarPubMed
Guns, M. and Vanacker, V. (2014). Shifts in landslide frequency–area distribution after forest conversion in the tropical Andes. Anthropocene, 6, 7585.CrossRefGoogle Scholar
Guo, M., Wu, W., Zhou, X., Chen, Y., and Li, J. (2015). Investigation of the dramatic changes in lake level of the Bosten Lake in northwestern China. Theoretical and Applied Climatology, 119, 341–51.CrossRefGoogle Scholar
Gupta, H. K. (2002). A review of recent studies of triggered earthquakes by artificial water reservoirs with special emphasis on earthquakes in Koyna, India. Earth-Science Reviews, 58, 279310.CrossRefGoogle Scholar
Gupta, H., Kao, S. J., and Dai, M. (2012). The role of mega dams in reducing sediment fluxes: a case study of large Asian rivers. Journal of Hydrology, 464, 447–58.Google Scholar
Gutiérrez, F., Parise, M., De Waele, J., and Jourde, H. (2014). A review on natural and human-induced geohazards and impacts in karst. Earth-Science Reviews, 138, 6188.CrossRefGoogle Scholar
Habersack, H. and Piégay, H. (2007). River restoration in the Alps and their surroundings: past experience and future challenges. Developments in Earth Surface Processes, 11, 703–35.CrossRefGoogle Scholar
Habersack, H., Haspel, D., and Kondolf, M. (2014). Large rivers in the Anthropocene: insights and tools for understanding climatic, land use, and reservoir influences. Water Resources Research, 50, 3641–6.CrossRefGoogle Scholar
Haddock, K. (1998). Giant Earthmovers: an Illustrated History. Osceola, WI: MotorBooks International.Google Scholar
Haeberli, W. and Burn, C. R. (2002). Natural hazards in forests: glacier and permafrost effects as related to climate change. In Environmental Change and Geomorphic Effects in Forests, ed. Sidle, R. C.. Wallingford: CABI, pp. 167202.CrossRefGoogle Scholar
Haeberli, W., Clague, J. J., Huggel, C., and Kääb, A. (2008). Hazards from lakes in high-mountain glacier and permafrost regions: climate change effects and process interactions. Avances de la Geomorphología en España, 2010, 439–46.Google Scholar
Haff, P. (2014). Humans and technology in the Anthropocene: six rules. The Anthropocene Review, 1, 126–36.CrossRefGoogle Scholar
Haff, P. K. (2010). Hillslopes, rivers, plows and trucks: mass transport on Earth’s surface by natural and technological processes. Earth Surface Processes and Landforms, 35, 1157–66.CrossRefGoogle Scholar
Haig, J., Nott, J., and Reichart, G. J. (2014). Australian tropical cyclone activity lower than at any time over the past 550–1,500 years. Nature, 505, 667–71.CrossRefGoogle Scholar
Hall, B. A. (1994). Formation processes of large earthen residential mounds in La Mixtequilla, Veracruz, Mexico. Latin American Antiquity, 5, 3150.CrossRefGoogle Scholar
Hall, C., Hamilton, A., Hoff, W. D., Viles, H. A., and Eklund, J. A. (2011). Moisture dynamics in walls: response to micro-environment and climate change. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Science, 467, 194211.CrossRefGoogle Scholar
Hall, R. (2015). Population and the future. Geography, 100, 3644.CrossRefGoogle Scholar
Hallegatte, S., Green, C., Nicholls, R. J., and Corfee-Morlot, J. (2013). Future flood losses in major coastal cities. Nature Climate Change, 3, 802–6.CrossRefGoogle Scholar
Hamilton, C. (2015). Getting the Anthropocene so wrong. The Anthropocene Review, 2, 102–7.CrossRefGoogle Scholar
Hamilton, C., Bonneuil, C., and Gemenne, F. (eds.) (2015). The Anthropocene and the Global Environmental Crisis. London and New York: Routledge.CrossRefGoogle Scholar
Hamza, M. A. and Anderson, W. K. (2005). Soil compaction in cropping systems: a review of the nature, causes and possible solutions. Soil and Tillage Research, 82, 121–45.CrossRefGoogle Scholar
Han, Z., Wang, T., Dong, Z., Hu, Y., and Yao, Z. (2007). Chemical stabilization of mobile dunefields along a highway in the Taklimakan Desert of China. Journal of Arid Environments, 68, 260–70.CrossRefGoogle Scholar
Hanley, M. E., Hoggart, S. P. G., Simmonds, D. J., Bichot, A., Colangelo, M. A., Bozzeda, F., and Thompson, R. C. (2014). Shifting sands? Coastal protection by sand banks, beaches and dunes. Coastal Engineering, 87, 136–46.CrossRefGoogle Scholar
Hanson, P. R., Joeckle, R. M., Young, A. R., and Horn, J. (2009). Late Holocene dune activity in the eastern Platte River Valley, Nebraska. Geomorphology, 103, 555–61.CrossRefGoogle Scholar
Hapke, C. J., Kratzmann, M. G., and Himmelstoss, E. A. (2013). Geomorphic and human influence on large-scale coastal change. Geomorphology, 199, 160–70.CrossRefGoogle Scholar
Happ, S. C., Rittenhouse, G., and Dobson, G. C. (1940). Some Principles of Accelerated Stream and Valley Sedimentation (No. 695). US Department of Agriculture.Google Scholar
Harbor, J. (1999). Engineering geomorphology at the cutting edge of land disturbance: erosion and sediment control on construction sites. Geomorphology, 31, 247–63.CrossRefGoogle Scholar
Harden, C. P. (2006). Human impacts on headwater fluvial systems in the northern and central Andes. Geomorphology, 79, 249–63.CrossRefGoogle Scholar
Harley, G. L., Polk, J. S., North, L. A., and Reeder, P. P. (2011). Application of a cave inventory system to stimulate development of management strategies: the case of west-central Florida, USA. Journal of Environmental Management, 92, 2547–57.CrossRefGoogle ScholarPubMed
Harmand, S., Lewis, J. E., Feibel, C. S., Lepre, C. J., Prat, S., Lenoble, A., and Roche, H. (2015). 3.3-million-year-old stone tools from Lomekwi 3, West Turkana, Kenya. Nature, 521, 310–5.CrossRefGoogle ScholarPubMed
Harnischmacher, S. (2010). Quantification of mining subsidence in the Ruhr District (Germany). Géomorphologie, 3, 261–74.Google Scholar
Harnischmacher, S. and Zepp, H. (2014). Mining and its impact on the earth surface in the Ruhr District (Germany). Zeitschrift für Geomorphologie, 58, Supplement, 322.CrossRefGoogle Scholar
Harris, N. and Evans, J. E. (2014). Channel evolution of sandy reservoir sediments following low-head dam removal, Ottawa River, Northwestern Ohio, USA. Open Journal of Modern Hydrology, 2014, doi:10.4236/ojmh.2014.42004.Google Scholar
Harrison, S., Glasser, N., Winchester, V., Haresign, E., Warren, C., and Jansson, K. (2006). A glacial lake outburst flood associated with recent mountain glacier retreat, Patagonian Andes. The Holocene, 16, 611–20.CrossRefGoogle Scholar
Hartemink, A. (2006). Soil erosion: perennial crop plantations. In Encycolpedia of Soil Science, ed. Lal, R.. Taylor and Francis: New York, pp. 1613–7.Google Scholar
Hartmann, H. C. (1990). Climate change impacts on Laurentian Great Lakes levels. Climatic Change, 17, 4967.CrossRefGoogle Scholar
Harvey, A. M. and Renwick, W. H. (1987). Holocene alluvial fan and terrace formation in the Bowland Fells, Northwest England. Earth Surface Processes and Landforms, 12, 249–57.CrossRefGoogle Scholar
Harvey, A. M., Oldfield, F., Baron, A. F., and Pearson, G. W., (1981). Dating of post-glacial landforms in the central Howgills. Earth Surface Processes and Landforms, 6, 401–12.CrossRefGoogle Scholar
Harvey, G. L., Moorhouse, T. P., Clifford, N. J., Henshaw, A. J., Johnson, M. F., Macdonald, D. W., and Rice, S. (2011). Evaluating the role of invasive aquatic species as drivers of fine sediment-related river management problems: the case of the signal crayfish (Pacifastacus leniusculus). Progress in Physical Geography, 35, 517–33.Google Scholar
Harvey, J. E. and Pederson, J. L. (2011). Reconciling arroyo cycle and paleoflood approaches to late Holocene alluvial records in dryland streams. Quaternary Science Reviews, 30, 855–66.CrossRefGoogle Scholar
Harvey, J. E., Pederson, J. L., and Rittenour, T. M. (2011). Exploring relations between arroyo cycles and canyon paleoflood records in Buckskin Wash, Utah: reconciling scientific paradigms. Geological Society of America Bulletin, 123, 2266–76.CrossRefGoogle Scholar
Hawley, R. J., MacMannis, K. R., and Wooten, M. S. (2013). Bed coarsening, riffle shortening, and channel enlargement in urbanizing watersheds, northern Kentucky, USA. Geomorphology, 201, 111–26.CrossRefGoogle Scholar
Haycraft, W. R. (2002). Yellow Steel: the Story of the Earth Moving Equipment Industry. Urbana and Chicago: University of Illinois Press.Google Scholar
Hayhoe, K., VanDorn, J., Croley, T. II, Schlegal, N., and Wuebbles, D. (2010). Regional climate change projections for Chicago and the US Great Lakes. Journal of Great Lakes Research, 36, 721.CrossRefGoogle Scholar
Hayhoe, S. J., Neill, C., Porder, S., McHorney, R., LeFebvre, P., Coe, M. T., and Krusche, A. V. (2011). Conversion to soy on the Amazonian agricultural frontier increases streamflow without affecting stormflow dynamics. Global Change Biology, 17, 1821–33.CrossRefGoogle Scholar
Haynes, G. (2012). Elephants (and extinct relatives) as earth-movers and ecosystem engineers. Geomorphology, 157, 99107.CrossRefGoogle Scholar
He, F., Vavrus, S. J., Kutzbach, J. E., Ruddiman, W. F., Kaplan, J. O., and Krumhardt, K. M. (2014). Simulating global and local surface temperature changes due to Holocene anthropogenic land cover change. Geophysical Research Letters, 41, 623–31.Google Scholar
He, K., Bin, W., and Dunyun, Z. (2004). Mechanism and mechanical model of karst collapse in an over-pumping area. Environmental Geology, 46, 1102–7.CrossRefGoogle Scholar
He, K., Liu, C. and Wang, S. (2003). Karst collapse related to over-pumping and a criterion for its stability. Environmental Geology, 43, 720–4.CrossRefGoogle Scholar
Healy, T. (1996). Sea level rise and impacts on nearshore sedimentation. Geologische Rundschau, 85, 546–53.CrossRefGoogle Scholar
Heberger, M., Cooley, H., Herrera, P., Gleick, P. H. and Moore, E. (2011).Potential impacts of increasing coastal flooding in California due to sea-level rise. Climatic Change, 109 (supplement), S229–49.CrossRefGoogle Scholar
Heckmann, M. (2014). Farmers, smelters and caravans: Two thousand years of land use and soil erosion in North Pare, NE Tanzania. Catena, 113, 187201.CrossRefGoogle Scholar
Heine, K., Niller, H. P., Nuber, T., and Scheibe, R. (2005). Slope and valley sediments as evidence of deforestation and land-use in prehistporic and historic Eastern Bavaria. Zeitschrift für Geomorphologie Supplementband, 139, 147–71.Google Scholar
Heitkamp, F., Sylvester, S. P., Kessler, M., Sylvester, M. D., and Jungkunst, H. F. (2014). Inaccessible Andean sites reveal human-induced weathering in grazed soils. Progress in Physical Geography, 38, 576601.CrossRefGoogle Scholar
Hernández, A. J., Lacasta, C., and Pastor, J. (2005). Effects of different management practices on soil conservation and soil water in a rainfed olive orchard. Agricultural Water Management, 77, 232–48.CrossRefGoogle Scholar
Hernández-Calvento, L., Jackson, D. W. T., Medina, R., Hernández-Cordero, A. I., Cruz, N., and Requejo, S. (2014). Downwind effects on an arid dunefield from an evolving urbanised area. Aeolian Research, 15, 301–9.CrossRefGoogle Scholar
Hesp, P. A. (2001). The Manawatu Dunefield: environmental change and human impacts. New Zealand Geographer, 57, 3340.CrossRefGoogle Scholar
Hesp, P. A., Schmutz, P., Martinez, M. L., Driskell, L., Orgera, R., Renken, K., Revelo, N. A. R., and Orocio, O. A. J. (2010). The effect on coastal vegetation of trampling on a parabolic dune. Aeolian Research, 2, 105–11.CrossRefGoogle Scholar
Hewitt, K. (2005).The Karakoram Anomaly? Glacier expansion and the “elevation effect,” Karakoram Himalaya. Mountain Research and Development, 25, 332–40.CrossRefGoogle Scholar
Hewitt, K. and Liu, J. (2010). Ice-dammed lakes and outburst floods, Karakoram Himalaya: historical perspectives on emerging threats. Physical Geography, 31, 528–51.CrossRefGoogle Scholar
Higgitt, D. L. and Lee, E. M. (eds.) (2001). Geomorphological Processes and Landscape Change: Britain in the Last 1000 Years, Chichester: Wiley, pp. 269–87.CrossRefGoogle Scholar
Hilfinger IV, M. F., Mullins, H. T., Burnett, A., and Kirby, M. E. (2001). A 2500 year sediment record from Fayetteville Green Lake, New York: evidence for anthropogenic impacts and historic isotope shift. Journal of Paleolimnology, 26, 293305.CrossRefGoogle Scholar
Hilmes, M. M. and Wohl, E. E. (1995). Changes in channel morphology associated with placer mining. Physical Geography, 16, 223–42.CrossRefGoogle Scholar
Hilton, M. J. (2006). The loss of New Zealand’s active dunes and the spread of marram grass (Ammophila arenaria). New Zealand Geographer, 62, 105–20.CrossRefGoogle Scholar
Hinkel, J., Nicholls, R. J., Tol, R. S., Wang, Z. B., Hamilton, J. M., Boot, G., and Klein, R. J. (2013). A global analysis of erosion of sandy beaches and sea-level rise: an application of DIVA. Global and Planetary Change, 111, 150–8.CrossRefGoogle Scholar
Hirabayashi, Y., Mahendran, R., Koirala, S., Konoshima, L., Yamazaki, D., Watanabe, S., and Kanae, S. (2013). Global flood risk under climate change. Nature Climate Change, 3, 816–21.CrossRefGoogle Scholar
Hiscock, K. M., Lister, D. H., Boar, R. R., and Green, F. M. L. (2001). An integrated assessment of long-term changes in the hydrology of three lowland rivers in eastern England. Journal of Environmental Management, 61, 195214.CrossRefGoogle ScholarPubMed
Hodgen, M. T. (1939). Domesday water mills. Antiquity, 13, 261–79.CrossRefGoogle Scholar
Hoegh-Guldberg, O. (1999). Climate change, coral bleaching and the future of the world’s coral reefs. Marine and Freshwater Research, 50, 839–66.Google Scholar
Hoegh-Guldberg, O. (2001). Sizing the impact: coral reef ecosystems as early casualties of climate change. In Fingerprints of Climate Change, ed. Walther, G. R., Burga, C. A., and Edwards, P. J.. New York: Kluwer/Plenum, pp 203–28.Google Scholar
Hoegh-Guldberg, O. (2011). Coral reef ecosystems and anthropogenic climate change. Regional Environmental Change, 11 (suppl.1), S215–27.CrossRefGoogle Scholar
Hoegh-Guldberg, O. (2014). Coral reefs in the Anthropocene: persistence or the end of the line? Geological Society, London, Special Publications, 395, 167–83.CrossRefGoogle Scholar
Hoeksema, R. J. (2007). Three stages in the history of land reclamation in the Netherlands. Irrigation and Drainage, 56(S1), S113–26.CrossRefGoogle 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.CrossRefGoogle Scholar
Hoffmann, T., Thorndycraft, V. R., Brown, A. G., Coulthard, T. J., Damnati, B., Kale, V. S., and Walling, D. E. (2010). Human impact on fluvial regimes and sediment flux during the Holocene: review and future research agenda. Global and Planetary Change, 72, 8798.CrossRefGoogle Scholar
Hoffmann, T., Mudd, S. M., Van Oost, K., Verstraeten, G., Erkens, G., Lang, A., and Aalto, R. (2013). Short communication: humans and the missing C-sink: erosion and burial of soil carbon through time. Earth Surface Dynamics, 1, 4552.CrossRefGoogle Scholar
Hollis, G. E. (1975). The effects of urbanization on floods of different recurrence interval. Water Resources Research, 11, 431–5.CrossRefGoogle Scholar
Holm, K., Bovis, M., and Jacob, M. (2004). The landslide response of alpine basins to post-Little Ice Age glacial thinning and retreat in southwestern British Columbia. Geomorphology, 57, 201–16.CrossRefGoogle Scholar
Holmes, S. C. A. (1980). Geology of the Country around Faversham. London: IGS/HMSO.Google Scholar
Hoogland, T., Van Den Akker, J. J. H., and Brus, D. J. (2012). Modeling the subsidence of peat soils in the Dutch coastal area. Geoderma, 171, 92–7.Google Scholar
Hooijer, A., Page, S., Jauhiainen, J., Lee, W. A., Lu, X. X., Idris, A., and Anshari, G. (2012). Subsidence and carbon loss in drained tropical peatlands. Biogeosciences, 9, 1053–71.Google Scholar
Hooke, J. M. and Mant, J. (2002). Floodwater use and management strategies in valleys of southeast Spain. Land Degradation & Development, 13, 165–75.CrossRefGoogle Scholar
Hooke, R. L. (1994). On the efficacy of humans as geomorphic agents. GSA Today, 4, 217, 224–5.CrossRefGoogle Scholar
Hooke, R. L. (1999). Spatial distribution of human geomorphic activity in the United States: comparison with rivers. Earth Surface Processes and Landforms, 24, 687–92.Google Scholar
Hooke, R. L. (2000). On the history of humans as geomorphic agents. Geology, 28, 843–6.3.0.CO;2-#>CrossRefGoogle Scholar
Hooke, R. L., Martín-Duque, J. F., and Pedraza, J. (2012). Land transformation by humans: a review. GSA Today, 22(12), 410.2.0.CO;2>CrossRefGoogle Scholar
Hoomehr, S., Schwartz, J. S., and Yoder, D. C. (2015). Potential changes in rainfall erosivity under GCM climate change scenarios for the southern Appalachian region, USA. Catena, doi:10.1016/j.catena.2015.01.012.CrossRefGoogle Scholar
Hope, G. (1999). Vegetation and fire response to late Holocene human occupation in island and mainland north west Tasmania. Quaternary International, 59, 4760.CrossRefGoogle Scholar
Hopkins, K. G., Morse, N. B., Bain, D. J., Bettez, N. D., Grimm, N. B., Morse, J. L., and Palta, M. M. (2015). Type and timing of stream flow changes in urbanizing watersheds in the Eastern US. Elementa: Science of the Anthropocene, 3(1), 000056.CrossRefGoogle Scholar
Hopkins, R. L., Altier, B. M., Haselman, D., Merry, A. D., and White, J. J. (2013). Exploring the legacy effects of surface coal mining on stream chemistry. Hydrobiologia, 713, 8795.Google Scholar
Horn, J. D., Joeckel, R. M., and Fielding, C. R. (2012). Progressive abandonment and planform changes of the central Platte River in Nebraska, central USA, over historical timeframes. Geomorphology, 139, 372–83.CrossRefGoogle Scholar
Hornbach, M. J., DeShon, H. R., Ellsworth, W. L., Stump, B. W., Hayward, C., Frohlich, C., and Luetgert, J. H. (2015). Causal factors for seismicity near Azle, Texas. Nature Communications, 6, doi:10.1038/ncomms7728.Google ScholarPubMed
Horrocks, M., Nichol, S. L., D-Costa, D., Augustinius, P., Jacobi, T., Shane, P. A., and Middleton, A. (2007). A late quaternary record of natural change and human impact from Rangihoua Bay, Bay of Islands, northern New Zealand. Journal of Coastal Research, 23, 592604.CrossRefGoogle Scholar
Hoshino, S., Esteban, M., Mikami, T., Takagi, H., and Shibayama, T. (2015). Estimation of increase in storm surge damage due to climate change and sea level rise in the Greater Tokyo area. Natural Hazards, doi: 10.1007/s11069–015–1983–4.Google Scholar
Houston, S. L., Houston, W. N., Zapata, C. E., and Lawrence, C. (2001). Geotechnical engineering practice for collapsible soils. Geotechnical and Geological Engineering, 19, 333–55.CrossRefGoogle Scholar
Howard, J. L. (2014). Proposal to add anthrostratigraphic and technostratigraphic units to the stratigraphic code for classification of anthropogenic Holocene deposits. The Holocene, 24, 1856–61.CrossRefGoogle Scholar
Hoyos, C.D., Agudelo, P. A., Webster, P. J., and Curry, J.A. (2006). Deconvolution of the factors contributing to the increase in global hurricane intensity. Science, 312, 94–7.CrossRefGoogle Scholar
Hu, A. and Deser, C. (2013). Uncertainty in future regional sea level rise due to internal climate variability. Geophysical Research Letters, 40, 2768–72.CrossRefGoogle Scholar
Hu, D., Clift, P. D., Böning, P., Hannigan, R., Hillier, S., Blusztajn, J., and Fuller, D. Q. (2013). Holocene evolution in weathering and erosion patterns in the Pearl River delta. Geochemistry, Geophysics, Geosystems, 14, 2349–68.CrossRefGoogle Scholar
Hu, R. (2006). Urban land subsidence in China. IAEG2006 Paper 786, Geological Society of London.Google Scholar
Hu, R. L., Yue, Z. Q., Wang, L. U., and Wang, S. J. (2004). Review on current status and challenging issues of land subsidence in China. Engineering Geology, 76, 6577.CrossRefGoogle Scholar
Huang, L., Liu, J., Shao, Q., and Liu, R. (2011). Changing inland lakes responding to climate warming in Northeastern Tibetan Plateau. Climatic Change, 109, 479502.CrossRefGoogle Scholar
Huang, R. and Chan, L. (2004). Human-induced landslides in China: mechanism study and its implications on slope management. Chinese Journal of Rock Mechanics and Engineering, 23, 2766–77.Google Scholar
Hübner, R., Herbert, R. J. H. and Astin, K. B. (2010). Cadmium release caused by the die-back of the saltmarsh cord grass Spartina anglica in Poole Harbour (UK). Estuarine, Coastal and Shelf Science, 87, 553–60.CrossRefGoogle Scholar
Huckleberry, G. and Duff, A. I. (2008). Alluvial cycles, climate and Pueblan settlement shifts near Zuni Salt Lake, New Mexico, USA. Geoarchaeology, 23, 107–30.CrossRefGoogle Scholar
Hudson, P. and Middelkoop, H. (eds.) (2015). Geomorphic Approaches to Integrated Floodplain Management of Lowland Fluvial Systems in North America and Europe. New York: Springer.CrossRefGoogle 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, 209–36.CrossRefGoogle Scholar
Hudson-Edwards, K., Macklin, M., and Taylor, M. (1997). Historic metal mining inputs to Tees river sediment. Science of the Total Environment, 194, 437–45.Google Scholar
Huggel, C., Allen, S., Deline, P., Fischer, L., Noetzli, J., and Ravanel, L. (2012). Ice thawing, mountains falling—are alpine rock slope failures increasing? Geology Today, 28, 98104.CrossRefGoogle Scholar
Huggel, C., Clague, J. J., and Korup, O. (2012). Is climate change responsible for changing landslide activity in high mountains? Earth Surface Processes and Landforms, 37, 7791.CrossRefGoogle Scholar
Hughes, R. G. and Paramor, O. A. L. (2004). On the loss of saltmarshes in south-east England and methods for their restoration. Journal of Applied Ecology, 41, 440–8.CrossRefGoogle Scholar
Hughes, R. J., Sullivan, M. E., and Yok, D. (1991). Human-induced erosion in a highlands catchment in Papua New Guinea: the prehistoric and contemporary records. Zeitschrift für Geomorphologie, supplementband, 83, 227–39.Google Scholar
Hughes, T. P., Bellwood, D. R., Baird, A. H., Brodie, J., Bruno, J. F., and Pandolfi, J. M. (2011). Shifting base-lines, declining coral cover, and the erosion of reef resilience: comment on Sweatman et al. (2011). Coral Reefs, 30, 653–60.CrossRefGoogle Scholar
Hughes, T. P., Linares, C., Dakos, V., van de Leemput, I. A., and van Nes, E. H. (2013). Living dangerously on borrowed time during slow, unrecognized regime shifts. Trends in Ecology & Evolution, 28, 149–55.CrossRefGoogle ScholarPubMed
Hultine, K. R. and Bush, S. E. (2011). Ecohydrological consequences of non-native riparian vegetation in the southwestern United States: a review from an ecophysiological perspective. Water Resources Research, 47, W07542, doi:10.1029/2010WR010317.CrossRefGoogle Scholar
Humboldt, A. and Bonpland, A. (1815). Personal Narrative of Travels to the Equinoctial Regions of the New Continent, during the Years 1799–1804. London: M. Carey.Google Scholar
Huntington, E. (1914). The Climatic Factor as Illustrated in Arid America. Carnegie Institution of Washington publication, 192.Google Scholar
Hupp, C. R. and Osterkamp, W. R. (1996). Riparian vegetation and fluvial geomorphic processes. Geomorphology, 14, 277–95.CrossRefGoogle Scholar
Hupp, C. R., Noe, G. B., Schenk, E. R., and Benthem, A. J. (2013). Recent and historic sediment dynamics along Difficult Run, a suburban Virginia Piedmont stream. Geomorphology, 180, 156–69.Google Scholar
Hupy, J. P. and Koehler, T. (2012). Modern warfare as a significant form of zoogeomorphic disturbance upon the landscape. Geomorphology, 157, 169–82.Google Scholar
Hupy, J. P. and Schaetzl, R. J. (2006). Introducing “bombturbation,” a singular type of soil disturbance and mixing. Soil Science, 171, 823–36.CrossRefGoogle Scholar
Husen, S., Kissling, E., and von Deschwanden, A. (2012). Induced seismicity during the construction of the Gotthard Base Tunnel, Switzerland: hypocenter locations and source dimensions. Journal of Seismology, 16, 195213.CrossRefGoogle Scholar
Hussain, I., Abu-Rizaiza, O. S., Habib, M. A., and Ashfaq, M. (2008). Revitalizing a traditional dryland water supply system: the karezes in Afghanistan, Iran, Pakistan and the Kingdom of Saudi Arabia. Water International, 33, 333–49.CrossRefGoogle Scholar
Hüttl, R. F. (1998). Ecology of post strip-mining landscapes in Lusatia, Germany. Environmental Science & Policy, 1, 129–35.CrossRefGoogle Scholar
Huxman, T. E., Wilcox, B. P., Breshears, D. D., Scott, R. L., Snyder, K. A., Small, E. E., and Jackson, R. B. (2005). Ecohydrological implications of woody plant encroachment. Ecology, 86, 308–19.CrossRefGoogle Scholar
Hyde, K. D., Wilcox, A. C., Jencso, K., and Woods, S. (2014). Effects of vegetation disturbance by fire on channel initiation thresholds. Geomorphology, 214, 8496.CrossRefGoogle Scholar
Ibáñez, C., Day, J. W., and Reyes, E. (2014). The response of deltas to sea-level rise: natural mechanisms and management options to adapt to high-end scenarios. Ecological Engineering, 65, 122–30.CrossRefGoogle Scholar
Ilyés, Z. (2010) Military activities: warfare and defence. In Anthropogenic Geomorphology: a Guide to Man-Made Landforms, ed. Szabó, J., Dávid, L., and Loczy, D.. Dordrecht: Springer, pp. 217–31.Google Scholar
Imaizumi, F., Sidle, R. C., and Kamei, R. (2008). Effects of forest harvesting on the occurrence of landslides and debris flows in steep terrain of central Japan. Earth Surface Processes and Landforms, 33, 827–40.CrossRefGoogle Scholar
Immer Zeel, W. W., van Beek, L. P. H., and Bierkens, M. F. P. (2010). Climate change will affect the Asian water towers. Science, 328, 1382–5.Google Scholar
Improta, L., Valoroso, L., Piccinini, D., and Chiarabba, C. (2015). A detailed analysis of wastewater‐induced seismicity in the Val d’Agri oil field (Italy). Geophysical Research Letters, 42, 2682–90.CrossRefGoogle Scholar
Inbar, M. and Llerena, C. A. (2000). Erosion processes in high mountain agricultural terraces in Peru. Mountain Research and Development, 20, 72–9.CrossRefGoogle Scholar
Inbar, M., Tamir, M. I., and Wittenberg, L. (1998). Runoff and erosion processes after a forest fire in Mount Carmel, a Mediterranean area. Geomorphology, 24, 1733.CrossRefGoogle Scholar
Indoitu, R., Kozhoridze, G., Batyrbaeva, M., Vitkovskaya, I., Orlovsky, N., Blumberg, D., and Orlovsky, L. (2015). Dust emission and environmental changes in the dried bottom of the Aral Sea. Aeolian Research, 17, 101–15.CrossRefGoogle Scholar
Inkpen, R., Viles, H., Moses, C., and Baily, B. (2012a). Modelling the impact of changing atmospheric pollution levels on limestone erosion rates in central London, 1980–2010. Atmospheric Environment, 61, 476–81.CrossRefGoogle Scholar
Inkpen, R. J., Viles, H. A., Moses, C., Baily, B., Collier, P., Trudgill, S. T., and Cooke, R. U. (2012b). Thirty years of erosion and declining atmospheric pollution at St Paul’s Cathedral, London. Atmospheric Environment, 62, 521–9.CrossRefGoogle Scholar
Innes, J. L. (1983). Lichenometric dating of debris‐flow deposits in the Scottish Highlands. Earth Surface Processes and Landforms, 8, 579–88.CrossRefGoogle Scholar
Innes, J., Blackford, J., and Simmons, I. (2010). Woodland disturbance and possible land-use regimes during the Late Mesolithic in the English uplands: pollen, charcoal and non-pollen palynomorph evidence from Bluewath Beck, North York Moors, UK. Vegetation History and Archaeobotany, 19, 439–52.CrossRefGoogle Scholar
IPCC (Intergovernmental Panel on Climate Change) (2007). Climate Change: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, ed. Solomon, S. et al., Cambridge and New York: Cambridge University Press.Google Scholar
IPCC (Intergovernmental Panel on Climate Change) (2013). Climate Change 2013: the Physical Science Basis: Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, ed. Stocker, T. F., Qin, D., Plattner, G.-K., Tignor, M., Allen, S. K., Boschung, J., Nauels, A., Xia, Y., Bex, V., and Midgley, P. M.. Cambride and New York: Cambridge University Press.Google Scholar
Irabien, M. J., García-Artola, A., Cearreta, A., and Leorri, E. (2015). Chemostratigraphic and lithostratigraphic signatures of the Anthropocene in estuarine areas from the eastern Cantabrian coast (N. Spain). Quaternary International, 364, 196205.CrossRefGoogle Scholar
Ireland, A. W., Clifford, M. J., and Booth, R. K. (2014). Widespread dust deposition on North American peatlands coincident with European land-clearance. Vegetation History and Archaeobotany, 23, 693700.CrossRefGoogle Scholar
Irish, J. L. and 8 others (2010). Potential implications of global warming and barrier island degradation on future hurricane inundation, property damages, and population impacted. Ocean and Coastal Management, 53, 645–57.CrossRefGoogle Scholar
Isebrands, J. G. and Richardson, J. (eds.) 2014. Poplars and Willows: Trees for Society and the Environment. Wallingford, U.K.: CABI.CrossRefGoogle Scholar
Isermann, M., Diekmann, M., and Heemann, S. (2007). Effects of the expansion by Hippophaë rhamnoides on plant species richness in coastal dunes. Applied Vegetation Science, 10, 3342.CrossRefGoogle Scholar
Jabaloy-Sánchez, A., Lobo, F. J., Azor, A., Bárcenas, P., Fernández-Salas, L. M., del Río, V. D., and Pérez-Peña, J. V. (2010). Human-driven coastline changes in the Adra River deltaic system, southeast Spain. Geomorphology, 119, 922.CrossRefGoogle Scholar
Jacks, G. V. and Whyte, R. O. (1939). The Rape of the Earth: a World Survey of Soil Erosion. London: Faber and Faber.Google Scholar
Jaeger, K. L. (2015). Reach-scale geomorphic differences between headwater streams draining mountaintop mined and unmined catchments. Geomorphology, 236, 2533.CrossRefGoogle Scholar
Jaeger, K. L. and Wohl, E. (2011). Channel response in a semiarid stream to removal of tamarisk and Russian olive. Water Resources Research, 47(2), W02536, doi: 10.1029/2009WR008741.CrossRefGoogle Scholar
Jaffe, B. E., Smith, R. E., and Foxgrover, A. C. (2007). Anthropogenic influence on sedimentation and intertidal mudflat change in San Pablo Bay, California: 1856–1983. Estuarine, Coastal and Shelf Science, 73, 175–87.CrossRefGoogle Scholar
Jalowska, A. M., Rodriguez, A. B., and McKee, B. A. (2015). Responses of the Roanoke Bayhead Delta to variations in sea level rise and sediment uupply during the Holocene and Anthropocene. Anthropocene, 9, 4155.CrossRefGoogle Scholar
James, L. A. (1989). Sustained storage and transport of hydraulic gold mining sediment in the Bear River, California. Annals of the Association of American Geographers, 79, 570–92.CrossRefGoogle Scholar
James, L. A. (1991). Incision and morphologic evolution of an alluvial channel recovering from hydraulic mining sediment. Geological Society of America Bulletin, 103, 723–36.2.3.CO;2>CrossRefGoogle Scholar
James, L. A. (2004). Tailings fans and valley‐spur cutoffs created by hydraulic mining. Earth Surface Processes and Landforms, 29, 869–82.CrossRefGoogle Scholar
James, L. A. (2010). Secular sediment waves, channel bed waves, and legacy sediment. Geography Compass, 4, 576–98.Google Scholar
James, L. A. (2013). Legacy sediment: definitions and processes of episodically produced anthropogenic sediment. Anthropocene, 2, 1626.CrossRefGoogle Scholar
James, L. A. and Marcus, W. A. (2006). The human role in changing fluvial systems: retrospect, inventory and prospect. Geomorphology, 79, 152–71.CrossRefGoogle Scholar
Jeffries, M. J. (2012). Ponds and the importance of their history: an audit of pond numbers, turnover and the relationship between the origins of ponds and their contemporary plant communities in south-east Northumberland, UK. Hydrobiologia, 689, 1121.CrossRefGoogle Scholar
Jennings, J. N. (1952). The Origin of the Broads. London: Royal Geographical Society.Google Scholar
Jensen, A. (2002). Artificial reefs of Europe: perspective and future. ICES Journal of Marine Science: Journal du Conseil, 59(suppl), S3-S13.CrossRefGoogle Scholar
Jensen, K., Trepel, M., Merritt, D., and Rosenthal, G. (2006). Restoration ecology of river valleys. Basic and Applied Ecology, 7, 383–7.CrossRefGoogle Scholar
Jeppesen, E., Brucet, S., Naselli-Flores, L., Papastergiadou, E., Stefanidis, K., Nõges, T., and Beklioğlu, M. (2015). Ecological impacts of global warming and water abstraction on lakes and reservoirs due to changes in water level and related changes in salinity. Hydrobiologia, 750, 201–27.CrossRefGoogle Scholar
Jewell, M., Houser, C., and Trimble, S. (2014). Initiation and evolution of blowouts within Padre Island National Seashore, Texas. Ocean & Coastal Management, 95, 156–64.CrossRefGoogle Scholar
Jiang, Y. Luo, Y., Zhao, Z., and Tao, S. (2009). Changes in wind speed over China during 1956–2004. Theoretical and Applied Climatology, doi: 10.1007/s00704-009-0152-7.CrossRefGoogle Scholar
Jiang, Z., Lian, Y., and Qin, X. (2014). Rocky desertification in Southwest China: impacts, causes, and restoration. Earth-Science Reviews, 132, 112.CrossRefGoogle Scholar
Jilbert, T., Reichard, G.-J., Aeschlimann, B., Günther, D., Boer, W., and de Lange, G. (2010). Climate-controlled multidecadal variability in North African dust transport to the Mediterranean. Geology, 38, 1922.CrossRefGoogle Scholar
Jin, H., Li, S., Cheng, G., Shaoling, W., and Li, X. (2000). Permafrost and climatic change in China. Global and Planetary Change, 26, 387404.CrossRefGoogle Scholar
Jin, R., Li, X., Che, T., Wu, L., and Mool, P. (2005). Glacier area changes in the Pumqu river basin, Tibetan Plateau, between the 1970s and 2001. Journal of Glaciology, 51, 607–10.CrossRefGoogle Scholar
Johnson, A. I. (ed.) (1991). Land Subsidence. International Association of Hydrological Sciences Publication 200.Google Scholar
Johnson, K. S. (1997). Evaporite karst in the United States. Carbonates and Evaporites, 12, 214.CrossRefGoogle Scholar
Johnson, K. S. (2005). Subsidence hazards due to evaporite dissolution in the United States. Environmental Geology, 48, 395409.CrossRefGoogle Scholar
Jomelli, V., Brunstein, D., Déqué, M., Vrac, M., and Grancher, D. (2009). Impacts of future climatic change (2070–2099) on the potential occurrence of debris flows: a case study in the Massif des Ecrins (French Alps). Climatic Change, 97, 171–91.CrossRefGoogle Scholar
Jonah, F. E., Adjei-Boateng, D., Agbo, N. W., Mensah, E. A., and Edziyie, R. E. (2015). Assessment of sand and stone mining along the coastline of Cape Coast, Ghana. Annals of GIS, doi:10.1080/19475683.2015.1007894.CrossRefGoogle Scholar
Jones, B. M., Arp, C. D., Jorgenson, M. T., Hinkel, K. M., Schmutz, J. A., and Flint, P. L. (2009). Increase in the rate and uniformity of coastline erosion in Arctic Alaska. Geophysical Research Letters, 36, doi: 10.1029/2008GL036205.CrossRefGoogle Scholar
Jones, B. M., Grosse, G., Arp, C. D., Jones, M. C., Walter Anthony, K. M., and Romanovsky, V. E. (2011). Modern thermokarst lake dynamics in the continuous permafrost zone, northern Seward Peninsula, Alaska. Journal of Geophysical Research: Biogeosciences 116(G2), doi: 10.1029/2011JG001666.CrossRefGoogle Scholar
Jones, B. M., Stoker, J. M., Gibbs, A. E., Grosse, G., Romanovsky, V. E., Douglas, T. A., and Richmond, B. M. (2013). Quantifying landscape change in an arctic coastal lowland using repeat airborne LiDAR. Environmental Research Letters, 8, 045025.CrossRefGoogle Scholar
Jones, C. G., Lawton, J. H., and Shachak, M. (1997). Positive and negative effects of organisms as physical ecosystem engineers. Ecology, 78, 1946–57.CrossRefGoogle Scholar
Jones, J. A., Swanson, F. J., Wemple, B. C., and Snyder, K. U. (2000). Effects of roads on hydrology, geomorphology, and disturbance patches in stream networks. Conservation Biology, 14, 7685.CrossRefGoogle Scholar
Jones, L. S., Rosenburg, M., Figueroa, M. D. M., McKee, K., Haravitch, B., and Hunter, J. (2010). Holocene valley-floor deposition and incision in a small drainage basin in western Colorado, USA. Quaternary Research, 74, 199206.CrossRefGoogle Scholar
Jones, M. D., Djamali, M., Holmes, J., Weeks, L., Leng, M. J., Lashkari, A., and Metcalfe, S. E. (2015). Human impact on the hydroenvironment of Lake Parishan, SW Iran, through the late Holocene. The Holocene, 25, 1651–61.CrossRefGoogle Scholar
Jones, R., Benson-Evans, K., and Chambers, F. M. (1985). Human influence upon sedimentation in Llangorse Lake, Wales. Earth Surface Processes and Landforms, 10, 227–35.CrossRefGoogle Scholar
Jordán, A. and Martínez-Zavala, L. (2008). Soil loss and runoff rates on unpaved forest roads in southern Spain after simulated rainfall. Forest Ecology and Management, 255, 913–9.CrossRefGoogle Scholar
Jordan, H., Hamilton, K., Lawley, R., and Price, S. J. (2016). Anthropogenic contribution to the geological and geomorphological record: a case study from Great Yarmouth, Norfolk, UK. Geomorphology, 253, 534–46, doi:10.1016/j.geomorph.2014.07.008.CrossRefGoogle Scholar
Jordán-López, A., Martínez-Zavala, L., and Bellinfante, N. (2009). Impact of different parts of unpaved forest roads on runoff and sediment yield in a Mediterranean area. Science of the Total Environment, 407, 937–44.CrossRefGoogle Scholar
Jorgenson, M. T., Shur, Y. L., and Pullman, E. R. (2006). Abrupt increase in permafrost degradation in Arctic Alaska. Geophysical Research Letters, 33, L02503, doi: 10.1029/2005GL024960.CrossRefGoogle Scholar
Joughlin, I. and Alley, R. B. (2011). Stability of the West Antarctic ice sheet in a warming world. Nature Geoscience, 4, 506–13.Google Scholar
Jouvet, G., Huss, M., Funk, M., and Blatter, H. (2011). Modelling the retreat of Grosser Aletschgletscher, Switzerland, in a changing climate. Journal of Glaciology, 57, 1033–45.CrossRefGoogle Scholar
Joyce, E. B. (2010). Australia’s geoheritage: history of study, a new inventory of geosites and applications to geotourism and geoparks. Geoheritage, 2(1–2), 3956.CrossRefGoogle Scholar
Julian, J. P., Wilgruber, N. A., de Beurs, K. M., Mayer, P. M., and Jawarneh, R. N. (2015). Long-term impacts of land cover changes on stream channel loss. Science of the Total Environment, 537, 399410.CrossRefGoogle ScholarPubMed
Julian, M. and Anthony, E. (1996). Aspects of landslide activity in the Mercantour Massif and the French Riviera, southeastern France. Geomorphology, 15, 275–89.CrossRefGoogle Scholar
Kabakci, H., Chevalier, P. M., and Papendick, R. I. (1993). Impact of tillage and residue management on dryland spring wheat development. Soil and Tillage Research, 26, 127–37.CrossRefGoogle Scholar
Kadiri, M., Ahmadian, R., Bockelmann-Evans, B., Rauen, W., and Falconer, R. (2012). A review of the potential water quality impacts of tidal renewable energy systems. Renewable and Sustainable Energy Reviews, 16, 329–41.CrossRefGoogle Scholar
Kadomura, H. (1983). Some aspects of large-scale land transformation due to urbanization and agricultural evelopment in recent Japan. Advances in Space Research, 2, 169–78.CrossRefGoogle Scholar
Kairis, O., Karavitis, C., Salvati, L., Kounalaki, A., and Kosmas, K. (2015). Exploring the impact of overgrazing on soil erosion and land degradation in a dry Mediterranean agro-forest landscape (Crete, Greece). Arid Land Research and Management, 29, 360–74.CrossRefGoogle Scholar
Kamh, G. M. E., Kallash, A., and Azzam, R. (2008). Factors controlling building susceptibility to earthquakes: 14-year recordings of Islamic archaeological sites in Old Cairo, Egypt: a case study. Environmental Geology, Special Issue, 56, 269–79, doi: 10.10007/s00254–007–1162–3.CrossRefGoogle Scholar
Kanevskiy, M., Shur, Y., Strauss, J., Jorgenson, T., Fortier, D., Stephani, E., and Vasiliev, A. (2016). Patterns and rates of riverbank erosion involving ice-rich permafrost (yedoma) in northern Alaska. Geomorphology, 253, 370–84, doi: 10.1016/j.geomorph.2015.10.023.CrossRefGoogle Scholar
Kang, S., Xu, Y., You, Q., Flügel, W. A., Pepin, N., and Yao, T. (2010). Review of climate and cryospheric change in the Tibetan Plateau. Environmental Research Letters, 5(1), 015101.CrossRefGoogle Scholar
Karl, T. R., Melillo, J. M., and Peterson, T. C. (2009). Global Climate Change Impacts in the United States. Cambridge: Cambridge University Press.Google Scholar
Kasai, M., Brierley, G. J., Page, M. J., Marutani, T., and Trustrum, N. A. (2005). Impacts of land use change on patterns of sediment flux in Weraamaia catchment, New Zealand. Catena, 64, 2760.CrossRefGoogle Scholar
Kaser, G. (1999). A review of the modern fluctuations of tropical glaciers. Global and Planetary Change, 22, 93103.CrossRefGoogle Scholar
Kaspari, S., McKenzie Skiles, S., Delaney, I., Dixon, D., and Painter, T. H. (2015). Accelerated glacier melt on Snow Dome, Mount Olympus, Washington, USA, due to deposition of black carbon and mineral dust from wildfire. Journal of Geophysical Research: Atmospheres, 120, 2793–807.Google Scholar
Kassam, A., Derpsch, R., and Friedrich, T. (2014). Global achievements in soil and water conservation: the case of conservation agriculture. International Soil and Water Conservation Research, 2, 513.CrossRefGoogle Scholar
Kates, R. W., Turner, B. L. I., and Clark, W. C. (1990). The great transformation. In The Earth as Transformed by Human Action, ed. Turner, B. L., Clark, W. C., Kates, R. W., Richards, J. F., Matthews, J. T., and Meyer, W. B.. Cambridge: Cambridge University Press, pp. 117.Google Scholar
Kaushal, S. S., Likens, G. E., Utz, R. M., Pace, M. L., Grese, M., and Yepsen, M. (2013). Increased river alkalinization in the Eastern US. Environmental Science & Technology, 47, 10302–11.Google Scholar
Kearns, T. J., Wang, G., Bao, Y., Jiang, J., and Lee, D. (2015). Current land subsidence and groundwater level changes in the Houston Metropolitan Area (2005–2012). Journal of Surveying Engineering, 05015002.CrossRefGoogle Scholar
Keatings, K., Tassie, G. J., Flower, R. J., Hassan, F. A., Hamdan, M. A. T., Hughes, M., and Arrowsmith, C. (2007). An examination of groundwater within the Hawara Pyramid, Egypt. Geoarchaeology, 22, 533–54.CrossRefGoogle Scholar
Keenan, T. F., Hollinger, D. Y., Bohrer, G., Dragoni, D., Munger, J. W., Schmid, H. P., and Richardson, A. D. (2013). Increase in forest water-use efficiency as atmospheric carbon dioxide concentrations rise. Nature, 499, 324–7.CrossRefGoogle ScholarPubMed
Kellerer‐Pirklbauer, A. Lieb, G. K., Avian, M., and Carrivick, J. (2012). Climate change and rock fall events in high mountain areas: numerous and extensive rock falls in 2007 at Mittlerer Burgstall, Central Austria. Geografiska Annaler: Series A, Physical Geography, 94, 5978.CrossRefGoogle Scholar
Kench, P. S., Smithers, S. G., McLean, R. F., and Nichol, S. L. (2009). Holocene reef growth in the Maldives: evidence of a mid-Holocene sea-level highstand in the central Indian Ocean. Geology, 37, 455–8.CrossRefGoogle Scholar
Kennett, D. J. and Beach, T. P. (2013). Archeological and environmental lessons for the Anthropocene from the Classic Maya collapse. Anthropocene, 4, 88100.CrossRefGoogle Scholar
Kennish, M. J. (2001). Coastal salt marsh systems in the U.S.: a review of anthropogenic impacts. Journal of Coastal Research, 17, 731–48.Google Scholar
Keranen, K. M., Savage, H. M., Abers, G. A., and Cochran, E. S. (2013). Potentially induced earthquakes in Oklahoma, USA: links between wastewater injection and the 2011 Mw 5.7 earthquake sequence. Geology, 41, 699702.CrossRefGoogle Scholar
Khanal, S., Anex, R. P., Anderson, C. J., Herzmann, D. E., and Jha, M. K. (2013). Implications of biofuel policy‐driven land cover change for rainfall erosivity and soil erosion in the United States. Global Change Biology: Bioenergy, 5, 713–22.Google Scholar
Khazendar, A., Borstad, C. P., Scheuchl, B., Rignot, E., and Seroussi, H. (2015). The evolving instability of the remnant Larsen B Ice Shelf and its tributary glaciers. Earth and Planetary Science Letters, 419, 199210.CrossRefGoogle Scholar
Khromova, T., Nosenko, G., Kutuzov, S., Muraviev, A., and Chernova, L. (2014). Glacier area changes in Northern Eurasia. Environmental Research Letters, 9, 015003.CrossRefGoogle Scholar
Khvorostyanov, D. V., Ciais, P., Krinner, G., and Zimov, S. A. (2008). Vulnerability of east Siberia’s frozen carbon stores to future warming. Geophysical Research Letters, 35, L10703, doi: 10.1029/2008GL033639.CrossRefGoogle Scholar
Kibler, K., Tullos, D., and Kondolf, M. (2011). Evolving expectations of dam removal outcomes: downstream geomorphic effects following removal of a small, gravel-filled dam. Journal of the American Water Resources Association, 47, 408–22.CrossRefGoogle Scholar
Kiernan, K. (2015). Nature, severity and persistence of geomorphological damage caused by armed conflict. Land Degradation and Development, 26, 380–96.CrossRefGoogle Scholar
Killick, D. (2015). Invention and innovation in African iron-smelting technologies. Cambridge Archaeological Journal, 25, 307–19.CrossRefGoogle Scholar
Kim, K. D., Lee, S., Oh, H. J., Choi, J. K., and Won, J. S. (2006). Assessment of ground subsidence hazard near an abandoned underground coal mine using GIS. Environmental Geology, 50, 1183–91.CrossRefGoogle Scholar
Kim, W. Y. (2013). Induced seismicity associated with fluid injection into a deep well in Youngstown, Ohio. Journal of Geophysical Research: Solid Earth, 118, 3506–18.Google Scholar
Kirby, M. X. (2004). Fishing down the coast: Historical expansion and collapse of oyster fisheries along continental margins. Proceedings of the National Academy of Sciences of the United States of America, 101, 13096–9.Google ScholarPubMed
Kirk, S. and Herbert, A. W. (2002). Assessing the impact of groundwater abstractions on river flows. Geological Society, London, Special Publications, 193, 211–33.CrossRefGoogle Scholar
Kirkbride, M. P. and Warren, C. R. (1999). Tasman Glacier, New Zealand: 20th century thinning and predicted calving retreat. Global and Planetary Change, 22, 1128.CrossRefGoogle Scholar
Kirkels, F. M. S. A., Cammeraat, L. H., and Kuhn, N. J. (2014). The fate of soil organic carbon upon erosion, transport and deposition in agricultural landscapes—A review of different concepts. Geomorphology, 226, 94105.CrossRefGoogle Scholar
Kirpotin, S. N., Polishchuk, Y., and Bryksina, N. (2009). Abrupt changes of thermokarst lakes in Western Siberia: impacts of climatic warming on permafrost melting. International Journal of Environmental Studies, 66, 423–31.CrossRefGoogle Scholar
Kirwan, M. L. and Megonigal, J. P. (2013). Tidal wetland stability in the face of human impacts and sea-level rise. Nature, 504, 5360.CrossRefGoogle ScholarPubMed
Kirwan, M. L., Murray, A. B., Donnelly, J. P., and Corbett, D. R. (2011). Rapid wetland expansion during European settlement and its implication for marsh survival under modern sediment delivery rates. Geology, 39, 507–10.CrossRefGoogle Scholar
Kiss, T., Sipos, G., Mauz, B., and Mezösi, G. (2012). Holocene aeolian sand mobilization, vegetation history and human impact on the stabilized sand dune area of the southern Nyírség, Hungary. Quaternary Research, 78, 492501.CrossRefGoogle Scholar
Knight, J. and Burningham, H. (2011). Sand dune morphodynamics and prehistoric human occupation in NW Ireland. Geological Society of America Special Papers, 476, 8192.CrossRefGoogle Scholar
Knight, J. and Harrison, S. (2014a). Limitations of uniformitarianism in the Anthropocene. Anthropocene, 5, 71–5.Google Scholar
Knight, J. and Harrison, S. (2014b). Mountain glacial and paraglacial environments under global climate change: lessons from the past, future directions and policy implications. Geografiska Annaler: Series A, Physical Geography, 96, 245–64.Google Scholar
Knight, M., Thomas, D. S. G., and Wiggs, G. F. S. (2004). Challenges of calculating dunefield mobility over the 21st century. Geomorphology, 59, 197213.CrossRefGoogle Scholar
Knighton, A. D. (1989). River adjustment to changes in sediment load: the effects of tin mining on the Ringarooma River, Tasmania, 1875–1984. Earth Surface Processes and Landforms, 14, 333–59.CrossRefGoogle Scholar
Knippertz, P. and Stuut, J. -B. W. (eds.) (2014). Mineral Dust. Dordrecht: Springer.CrossRefGoogle Scholar
Knox, J. C. (1977). Human impacts on Wisconsin stream channels. Annals of the Association of American Geographers, 67, 323–42.CrossRefGoogle Scholar
Knox, J. C. (1987). Historical valley floor sedimentation in the Upper Mississippi Valley. Annals of the Association of American Geographers, 77, 224–44.CrossRefGoogle Scholar
Knutson, T. R., Sirutis, J. J., Garner, S. T., Vecchi, G. A., and Held, I. M. (2008). Simulated reduction in Atlantic hurricane frequency under twenty-first-century warming conditions. Nature Geoscience, 1, 359–64.Google Scholar
Koluvek, P. K., Tanji, K. K., and Trout, T. J. (1993). Overview of soil erosion from irrigation. Journal of Irrigation and Drainage Engineering, 119, 929–46.CrossRefGoogle Scholar
Komar, P. D. and Allan, J. C. (2008). Increasing hurricane-generated wave heights along the U.S. East Coast and their climate controls. Journal of Coastal Research, 24, 479–88.Google Scholar
Komori, J. (2008). Recent expansions of glacial lakes in the Bhutan Himalayas. Quaternary International, 184, 177–86.CrossRefGoogle Scholar
Kondolf, G. M. (1997). Profile: hungry water: effects of dams and gravel mining on river channels. Environmental Management, 21, 533–51.CrossRefGoogle ScholarPubMed
Kondolf, G. M. (2013). Sustainable sediment management in reservoirs and regulated rivers: experiences from five continents, Earth’s Future, 2, 256–80 doi: 10.1002/2013EF000184.Google Scholar
Kondolf, G. M. and Podolak, K. (2014). Space and time scales in human-landscape systems. Environmental Management, 53, 7687.CrossRefGoogle ScholarPubMed
Kondolf, G. M., Anderson, S., Lave, R., Pagano, L., Merenlender, A., and Bernhardt, E. S. (2007). Two decades of river restoration in California: what can we learn? Restoration Ecology, 15, 516–23.CrossRefGoogle Scholar
Korpak, J. (2007). The influence of river training on mountain channel changes (Polish Carpathian Mountains). Geomorphology, 92, 166–81.CrossRefGoogle Scholar
Kortekaas, S., Bagdanaviciute, I., Gyssels, P., Huerta, J. M. A., and Héquette, A. (2010). Assessment of the effects of marine aggregate extraction on the coastline: an example from the German Baltic Sea Coast. Journal of Coastal Research, S1, 51, 205–14.Google Scholar
Kossoff, D., Dubbin, W. E., Alfredsson, M., Edwards, S. J., Macklin, M. G., and Hudson Edwards, K. A. (2014). Mine tailings dams: characteristics, failure, environmental impacts, and remediation. Applied Geochemistry, 51, 229–45.CrossRefGoogle Scholar
Koulouri, M. and Giourga, C. (2007). Land abandonment and slope gradient as key factors of soil erosion in Mediterranean terraced lands. Catena, 69, 274–81.CrossRefGoogle Scholar
Kozlov, M. V. and Zvereva, E. L. (2007). Industrial barrens: extreme habitats created by non-ferrous metallurgy. Review of Environmental Science and Bio-Technology, 6, 231–59.Google Scholar
Kraaijpoel, D. and Dost, B. (2013). Implications of salt-related propagation and mode conversion effects on the analysis of induced seismicity. Journal of Seismology, 17, 95107.CrossRefGoogle Scholar
Krabill, W., Frederick, E., Manizade, S., Martin, C., Sonntag, J., Swift, R., Thomas, R., Wright, W., and Yungel, J. (1999). Rapid thinning of parts of the southern Greenland Ice Sheet. Science, 283, 1522–4.CrossRefGoogle ScholarPubMed
Krauss, K. W., Lovelock, C. E., McKee, K. L., López-Hoffman, L., Ewe, S. M., and Sousa, W. P. (2008). Environmental drivers in mangrove establishment and early development: a review. Aquatic Botany, 89, 105–27.CrossRefGoogle Scholar
Krauss, K. W., McKee, K. L., Lovelock, C. E., Cahoon, D. R., Saintilan, N., Reef, R., and Chen, L. (2014). How mangrove forests adjust to rising sea level. New Phytologist, 202, 1934.CrossRefGoogle ScholarPubMed
Kravtsova, V. I. and Tarasenko, T. V. (2010). Space monitoring of Aral Sea degradation. Water Resources, 37, 285–96.CrossRefGoogle Scholar
Kroeker, K. J., Kordas, R. L., Crim, R., Hendriks, I. E., Ramajo, L., Singh, G. S., and Gattuso, J. P. (2013). Impacts of ocean acidification on marine organisms: quantifying sensitivities and interaction with warming. Global Change Biology, 19, 1884–96.CrossRefGoogle ScholarPubMed
Kroes, D. E., and Hupp, C. R. (2010). The effect of channelization on floodplain sediment deposition and subsidence along the Pocomoke River, Maryland. Journal of the American Water Resources Association, 46, 686–99.CrossRefGoogle Scholar
Kueny, J. A. and Day, M. J. (2002). Designation of protected karstlands in Central America: a regional assessment. Journal of Cave and Karst Studies, 64, 165–74.Google Scholar
Kuenzer, C. and Stracher, G. B. (2012). Geomorphology of coal seam fires. Geomorphology, 138, 209–22.CrossRefGoogle Scholar
Kuenzer, C., Zhang, J., Tetzlaff, A., Van Dijk, P., Voigt, S., Mehl, H., and Wagner, W. (2007). Uncontrolled coal fires and their environmental impacts: investigating two arid mining regions in north-central China. Applied Geography, 27, 4262.CrossRefGoogle Scholar
Kull, C. A. (2000). Deforestation, erosion, and fire: degradation myths in the environmental history of Madagascar. Environment and History, 6, 423–50.CrossRefGoogle Scholar
Kumar, A. (2009). Reclaimed islands and new offshore townships in the Arabian Gulf: potential natural hazards. Current Science, 96, 480–5.Google Scholar
Kumar, M., Goossens, E., and Goossens, R. (1993). Assessment of sand dune change detection in Rajasthan (Thar) Desert, India. International Journal of Remote Sensing, 14, 1689–703.CrossRefGoogle Scholar
Kumar, S. V. and Bhagavanulu, D. V. S. (2008). Effect of deforestation on landslides in Nilgiris district—a case study. Journal of the Indian Society of Remote Sensing, 36, 105–8.Google Scholar
Kummu, M., Penny, D., Sarkkula, J., and Koponen, J. (2008). Sediment: curse or blessing for Tonle Sap Lake? Ambio, 37, 158–63.CrossRefGoogle ScholarPubMed
Küster, M., Fülling, A., Kaiser, K., and Ulrich, J. (2014). Aeolian sands and buried soils in the Mecklenburg Lake District, NE Germany: Holocene land-use history and pedo-geomorphic response. Geomorphology 211, 6476.CrossRefGoogle Scholar
Kwong, Y. T. J. and Tau, T. Y. (1994). Northward migration of permafrost along the Mackenzie Highway and climatic warming. Climatic Change, 26, 399419.CrossRefGoogle Scholar
Labadz, J. C., Burt, T. P., and Potter, A. W. L. (1991). Sediment yield and delivery in the blanket peat moorlands of the southern Pennines. Earth Surface Processes and Landforms, 16, 255–71.CrossRefGoogle Scholar
Labat, D., Goddéris, Y., Probst, J. L., and Guyot, J. L. (2004). Evidence for global runoff increase related to climate warming. Advances in Water Resources, 27, 631–42.CrossRefGoogle Scholar
Labrière, N., Locatelli, B., Laumonier, Y., Freycon, V., and Bernoux, M. (2015). Soil erosion in the humid tropics: a systematic quantitative review. Agriculture, Ecosystems & Environment, 203, 127–39.CrossRefGoogle Scholar
Lach, J. and Wyżga, B. (2002). Channel incision and flow increase of the upper Wisłoka River, southern Poland, subsequent to the reafforestation of its catchment. Earth Surface Processes and Landforms, 27, 445–62.CrossRefGoogle Scholar
Lahaye, C., Guérin, G., Boëda, E., Fontugne, M., Hatté, C., Frouin, M., and Da Costa, A. (2015). New insights into a late-Pleistocene human occupation in America: the Vale da Pedra Furada complete chronological study. Quaternary Geochronology 30, 445–51, doi: 10.1016/j.quageo.2015.03.009.CrossRefGoogle Scholar
Lai, S., Loke, L. H., Hilton, M. J., Bouma, T. J., and Todd, P. A. (2015). The effects of urbanisation on coastal habitats and the potential for ecological engineering: a Singapore case study. Ocean & Coastal Management, 103, 7885.CrossRefGoogle Scholar
Laity, J. E. (2003). Aeolian destabilization along the Mojave River, Mojave Desert, California: linkages among fluvial, groundwater, and aeolian systems. Physical Geography, 24, 196221.CrossRefGoogle Scholar
Lambert, J. H., Jennings, J. N., Smith, C. T., Green, C., and Hutchinson, J. N. (1970). The Making of the Broads: a Reconsideration of their Origin in the Light of New Evidence. London: Royal Geographical Society.Google Scholar
Lammers, J. M., van Soelen, E. E., Donders, T. H., Wagner-Cremer, F., Damsté, J. S., and Reichart, G. J. (2013). Natural environmental changes versus human impact in a Florida estuary (Rookery Bay, USA). Estuaries and Coasts, 36, 149–57.CrossRefGoogle Scholar
Lamoureux, S. F. and Lafrenière, M. J. (2009). Fluvial impact of extensive active layer detachments, Cape Bounty, Melville Island, Canada. Arctic, Antarctic, and Alpine Research, 41, 5968.CrossRefGoogle Scholar
Lane, P. (2009). Environmental narratives and the history of soil erosion in Kondoa District, Tanzania: an archaeological perspective. International Journal of African Historical Studies, 42, 457–83.Google Scholar
Lane, P. N. and Sheridan, G. J. (2002). Impact of an unsealed forest road stream crossing: water quality and sediment sources. Hydrological Processes, 16, 2599–612.CrossRefGoogle Scholar
Lang, A. (2003). Phases of soil erosion-derived colluviation in the loess hills of South Germany. Catena, 51, 209–21.CrossRefGoogle Scholar
Langedal, M. (1997). The influence of a large anthropogenic sediment source on the fluvial geomorphology of the Knabeåna-Kvina rivers, Norway. Geomorphology, 19, 117–32.CrossRefGoogle Scholar
Langdon, J. (1991). Water–mills and windmills in the west midlands, 1086–1500. The Economic History Review, 44, 424–44.CrossRefGoogle Scholar
Lantuit, H. and Pollard, W. H. (2008). Fifty years of coastal erosion and retrogressive thaw slump activity on Herschel Island, southern Beaufort Sea, Yukon Territory, Canada. Geomorphology, 95, 84102.CrossRefGoogle Scholar
Larsen, I. J., MacDonald, L. H., Brown, E., Rough, D., Welsh, M. J., Pietraszek, J. H., and Schaffrath, K. (2009). Causes of post-fire runoff and erosion: water repellency, cover, or soil sealing? Soil Science Society of America Journal, 73, 1393–407.CrossRefGoogle Scholar
Laruelle, G. G., Roubeix, V., Sferratore, A., Brodherr, B., Ciuffa, D., Conley, D. J., and Van Cappellen, P. (2009). Anthropogenic perturbations of the silicon cycle at the global scale: Key role of the land‐ocean transition. Global Biogeochemical Cycles, 23, GB4031, doi: 10.1029/2008GB003267.CrossRefGoogle Scholar
Lasanta, T., García-Ruez, J. M., Pérez-Rontomé, C., and Sancho-Marcén, (2000). Runoff and sediment yield in a semi-ard environment: the effect of land management after farmland abandonment. Catena, 38, 265–78.CrossRefGoogle Scholar
Laumets, L., Kalm, V., Poska, A., Kele, S., Lasberg, K., and Amon, L. (2014). Palaeoclimate inferred from δ18O and palaeobotanical indicators in freshwater tufa of Lake Äntu Sinijärv, Estonia. Journal of Paleolimnology, 51, 99111.CrossRefGoogle Scholar
Lavee, H., Poesen, J., and Yair, A. (1997). Evidence of high efficiency water-harvesting by ancient farmers in the Negev Desert, Israel. Journal of Arid Environments, 35, 341–8.CrossRefGoogle Scholar
Lavee, H., Kutiel, P., Segev, M., and Benyamini, Y. (1995). Effect of surface roughness on runoff and erosion in a Mediterranean ecosystem: the role of fire. Geomorphology, 11, 227–34.CrossRefGoogle Scholar
Lawson, D. E. (1986). Response of permafrost terrain to disturbance: a synthesis of observations from northern Alaska, USA. Arctic and Alpine Research, 18, 117.CrossRefGoogle Scholar
Le Maitre, D.C., Versfield, D. B., and Chapman, R.A. (2000). The impact of invading alien plants on surface water resources in South Africa: a preliminary assessment. Water SA, 26, 397408.Google Scholar
Lecce, S. A. and Pavlowsky, R. T. (2014). Floodplain storage of sediment contaminated by mercury and copper from historic gold mining at Gold Hill, North Carolina, USA. Geomorphology, 206, 122–32.CrossRefGoogle Scholar
Lecce, S., Pavlowsky, R., and Schlomer, G. (2008). Mercury contamination of active channel sediment and floodplain deposits from historic gold mining at Gold Hill, North Carolina, USA. Environmental Geology, 55, 113–21.CrossRefGoogle Scholar
Leclercq, N., Gattuso, J.-P., and Jaubert, J. (2000). CO2 partial pressure controls the calcification rate of a coral community. Global Change Biology, 6, 329–34.CrossRefGoogle Scholar
Lee, J. A. and Gill, T. E. (2015). Multiple causes of wind erosion in the Dust Bowl. Aeolian Research, 19, 1536.CrossRefGoogle Scholar
Lei, X., Ma, S., Chen, W., Pang, C., Zeng, J., and Jiang, B. (2013). A detailed view of the injection‐induced seismicity in a natural gas reservoir in Zigong, southwestern Sichuan Basin, China. Journal of Geophysical Research: Solid Earth, 118, 4296–311.Google Scholar
Lei, Y., Yang, K., Wang, B., Sheng, Y., Bird, B. W., Zhang, G., and Tian, L. (2014). Response of inland lake dynamics over the Tibetan Plateau to climate change. Climatic Change, 125, 281–90.CrossRefGoogle Scholar
Lenzi, M. A. (2002). Stream bed stabilization using boulder check dams that mimic step-pool morphology features in Northern Italy. Geomorphology, 45, 243–60.CrossRefGoogle Scholar
Leopold, L. B. (1951). Rainfall frequency: an aspect of climatic variation. Transactions of the American Geophysics Union, 32, 347–57.Google Scholar
Lespez, L. (2003). Geomorphic responses to long-term land use changes in Eastern Macedonia (Greece). Catena, 51, 181208.CrossRefGoogle Scholar
Lesschen, J. P., Cammeraat, L. H., and Nieman, T. (2008). Erosion and terrace failure due to agricultural land abandonment in a semi‐arid environment. Earth Surface Processes and Landforms, 33, 1574–84.CrossRefGoogle Scholar
Lesschen, J. P., Kok, K., Verburg, P. H., and Cammeraat, L. H. (2007). Identification of vulnerable areas for gully erosion under different scenarios of land abandonment in Southeast Spain. Catena, 71, 110–21.CrossRefGoogle Scholar
Lestrelin, G., Vigiak, O., Pelletreau, A., Keohavong, B., and Valentin, C. (2012, May). Challenging established narratives on soil erosion and shifting cultivation in Laos. Natural Resources Forum, 36, 6375.CrossRefGoogle Scholar
Letey, J. (2001). Cases and consequences of fire-induced soil water repellency. Hydrological Processes, 15, 2867–75.CrossRefGoogle Scholar
Levermann, A., Clark, P. U., Marzeion, B., Milne, G. A., Pollard, D., Radic, V., and Robinson, A. (2013). The multimillennial sea-level commitment of global warming. Proceedings of the National Academy of Sciences, 110, 13745–50.CrossRefGoogle ScholarPubMed
Lewis, K. C., Zyvoloski, G. A., Travis, B., Wilson, C., and Rowland, J. (2012). Drainage subsidence associated with Arctic permafrost degradation. Journal of Geophysical Research: Earth Surface 117(F4), F04019, doi: 10.1029/2011JF002284.CrossRefGoogle Scholar
Lewis, R. W., Makurat, A., and Pao, W. K. (2003). Fully coupled modeling of seabed subsidence and reservoir compaction of North Sea oil fields. Hydrogeology Journal, 11, 142–61.CrossRefGoogle Scholar
Lewis, S. L. and Maslin, M. A. (2015). Defining the Anthropocene. Nature, 519, 171–80.CrossRefGoogle ScholarPubMed
Leyva, J. C., Martínez, J. F., and Roa, M. G. (2007). Analysis of the adoption of soil conservation practices in olive groves: the case of mountainous areas in southern Spain. Spanish Journal of Agricultural Research, 5, 249–58.Google Scholar
Li, C., Tang, X., and Ma, T. (2006). Land subsidence caused by groundwater exploitation in the Hangzhou-Jiaxing-Huzhou Plain, China. Hydrogeology Journal, 14, 1652–65.CrossRefGoogle Scholar
Li, D. D., Lerman, A., and Mackenzie, F. T. (2011). Human perturbations on the global biogeochemical cycles of coupled Si–C and responses of terrestrial processes and the coastal ocean. Applied Geochemistry, 26, S289–91.CrossRefGoogle Scholar
Li, J., Okin, G. S., Tatarko, J., Webb, N. P. and Herrick, J. E. (2014). Consistency of wind erosion assessments across land use and land cover types: a critical analysis. Aeolian Research, 15, 253–60.CrossRefGoogle Scholar
Li, X. and 9 others (2008). Cryospheric change in China. Global and Planetary Change, 62, 210–8.CrossRefGoogle Scholar
Li, Y., Cui, J., Zhangh, T., Okuro, T., and Drake, S. (2009). Effectiveness of sand-fixing measures on desert land restoration in Kerqin Sandy Land, northern China. Ecological Engineering, 35, 118–27.CrossRefGoogle Scholar
Liébault, F. and Piégay, H. (2002). Causes of 20th century channel narrowing in mountain and piedmont rivers of southeastern France. Earth Surface Processes and Landforms, 27, 425–44.CrossRefGoogle Scholar
Lightfoot, D. R. (1996a). Moroccan khettara: traditional irrigation and progressive desiccation. Geoforum, 27, 261–73.CrossRefGoogle Scholar
Lightfoot, D. R. (1996b). Syrian qanat Romani: history, ecology, abandonment. Journal of Arid Environments, 33, 321–6.CrossRefGoogle Scholar
Lightfoot, D. R. (2000). The origin and diffusion of qanats in Arabia: new evidence from the northern and southern peninsula. Geographical Journal, 166, 215–26.CrossRefGoogle Scholar
Lightfoot, K. G. and Cuthrell, R. Q. (2015). Anthropogenic burning and the Anthropocene in late Holocene California. The Holocene, 25, 1581–7.CrossRefGoogle Scholar
Limondin‐Lozouet, N., Preece, R. C., and Antoine, P. (2013). The Holocene tufa at Daours (Somme Valley, northern France): Malacological succession and palaeohydrological implications. Boreas, 42, 650–63.CrossRefGoogle Scholar
Lin, C., Yang, K., Qin, J., and Fu, R. (2013). Observed coherent trends of surface and upper-air wind speed over China since 1960. Journal of Climate, 26, 2891–903.CrossRefGoogle Scholar
Lin, N., Emanuel, K., Oppenheimer, M., and Vanmarcke, E. (2012). Physically based assessment of hurricane surge threat under climate change. Nature Climate Change, 2, 462–7.CrossRefGoogle Scholar
Lin, Y. and Wei, X. (2008). The impact of large-scale forest harvesting on hydrology in the Willow watershed of Central British Columbia. Journal of Hydrology, 359, 141–49.CrossRefGoogle Scholar
Lindberg, T. T., Bernhardt, E. S., Bier, R., Helton, A. M., Merola, R. B., Vengosh, A., and Di Giulio, R. T. (2011). Cumulative impacts of mountaintop mining on an Appalachian watershed. Proceedings of the National Academy of Sciences, 108, 20929–34.CrossRefGoogle Scholar
Ling, F. and Zhang, T. (2003). Impact of the timing and duration of seasonal snow cover on the active layer and permafrost in the Alaskan Arctic. Permafrost and Periglacial Processes, 14, 141–50.CrossRefGoogle Scholar
List, J. H., Sallenger, A. H., Hansen, M. E., and Jaffe, B. E. (1997). Accelerated sea level rise and rapid coastal erosion: testing a causal relationship for the Louisiana barrier islands. Marine Geology, 140, 437–65.CrossRefGoogle Scholar
Liu, C. H., Pan, Y. W., Liao, J. J., Huang, C. T., and Ouyang, S. (2004). Characterization of land subsidence in the Choshui River alluvial fan, Taiwan. Environmental Geology, 45, 1154–66.CrossRefGoogle Scholar
Liu, C. W., Lin, W. S., Shang, C., and Liu, S. H. (2001). The effect of clay dehydration on land subsidence in the Yun-Lin coastal area, Taiwan. Environmental Geology, 40, 518–27.CrossRefGoogle Scholar
Liu, S. and Wang, T. (2014). Aeolian processes and landscape change under human disturbances on the Sonid grassland of inner Mongolian Plateau, northern China. Environmental Earth Sciences, 71, 2399–407.CrossRefGoogle Scholar
Liu, S., Xu, L., and Talwani, P. (2011). Reservoir-induced seismicity in the Danjiangkou Reservoir: a quantitative analysis. Geophysical Journal International, 185, 514–28.CrossRefGoogle Scholar
Liu, T., Kinouchi, T., and Ledezma, F. (2013). Characterization of recent glacier decline in the Cordillera Real by LANDSAT, ALOS, and ASTER data. Remote Sensing of Environment, 137, 158–72.CrossRefGoogle Scholar
Liu, W., Chen, W., and Peng, C. (2014). Assessing the effectiveness of green infrastructures on urban flooding reduction: a community scale study. Ecological Modelling, 291, 614.CrossRefGoogle Scholar
Lixin, Y., Fang, Z., He, X., Shijie, C., Wei, W., and Qiang, Y. (2011). Land subsidence in Tianjin, China. Environmental Earth Sciences, 62, 1151–61.CrossRefGoogle Scholar
Llenos, A. L. and Michael, A. J. (2013). Modeling earthquake rate changes in Oklahoma and Arkansas: possible signatures of induced seismicity. Bulletin of the Seismological Society of America, 103, 2850–61.CrossRefGoogle Scholar
Lóczy, D. and Sütö, L. (2011). Human agency and geomorphology. In The SAGE Handbook of Geomorphology, ed. Gregory, K. J. and Goudie, A. S.. Sage: London, pp. 260–78.Google Scholar
Logsdon, S. D. (2013). Depth dependence of chisel plow tillage erosion. Soil and Tillage Research, 128, 119–24.CrossRefGoogle Scholar
Lokier, S. W. (2013). Coastal sabkha preservation in the Arabian Gulf. Geoheritage, 5, 1122.CrossRefGoogle Scholar
Londoño, A. C. (2008). Pattern and rate of erosion inferred from Inca agricultural terraces in arid southern Peru. Geomorphology, 99, 1325.CrossRefGoogle Scholar
Lopez-Bermudez, F., Romero-Díaz, A., Martínez-Fernandez, J., and Martínez-Fernandez, J. (1998). Vegetation and soil erosion under a semi-arid Mediterranean climate: a case study from Murcia (Spain). Geomorphology, 24, 51–8.CrossRefGoogle Scholar
López-Moreno, J. I., Beguería, S., and García-Ruiz, J. M. (2006). Trends in high flows in the central Spanish Pyrenees: response to climatic factors or to land-use change? Hydrological Sciences Journal, 51, 1039–50.CrossRefGoogle Scholar
López-Moreno, J. I., Fontaneda, S., Bazo, J., Revuelto, J., Azorin-Molina, C., Valero-Garcés, B., and Alejo-Cochachín, J. (2014). Recent glacier retreat and climate trends in Cordillera Huaytapallana, Peru. Global and Planetary Change, 112, 111.CrossRefGoogle Scholar
Lorimer, J. (2012). Multinatural geographies for the Anthropocene. Progress in Human Geography, 36, 593612.CrossRefGoogle Scholar
Loughran, R. J., Campbell, B. L., Elliott, G. L., and Shelly, D. J. (1990). Determination of the rate of sheet erosion on grazing land using caesium-137. Applied Geography, 10, 125–33.CrossRefGoogle Scholar
Lowdermilk, W. C. (1934). Acceleration of erosion above geologic norms. Eos, Transactions American Geophysical Union, 15, 505–9.Google Scholar
Lowdermilk, W. C. (1935). Man-made deserts. Pacific Affairs, 8, 409–19.CrossRefGoogle Scholar
Lowe, A. (1986). Bronze Age burial mounds on Bahrain. Iraq, 48, 7384.CrossRefGoogle Scholar
Lowenthal, D. (2016). Origins of Anthropocene awareness. The Anthropocene Review, 3, 5263, doi: 10.1177/2053019615609953CrossRefGoogle Scholar
Lu, H. and 12 others (2009). Earliest domestication of common millet (Panicum miliaceum) in East Asia extended to 10,000 years ago. Proceedings of the National Academy of Sciences 106, 7367–72.CrossRefGoogle Scholar
Lu, Z., Streets, D. G., Zhang, Q., Wang, S., Carmichael, G. R., Cheng, Y. F., Wei, C., Chin, M., Diehl, T., and Tan, Q. (2010). Sulfur dioxide emissions in China and sulphur trends in East Asia since 2000. Atmospheric Chemistry and Physics, 10, 6311–31.Google Scholar
Lubke, R. A. (2013). Restoration of dune ecosystems following mining in Madagascar and Namibia: contrasting restoration approaches adopted in regions of high and low population density. In Restoration of Coastal Dunes, ed. Martínez, M. L., Gallego-Fernández, J. B., and Hesp, P. A.. Berlin and Heidelberg: Springer, pp 199215.CrossRefGoogle Scholar
Lubos, C. C. M., Dreibrodt, S., Nelle, O., Klamm, M., Friederich, S., Meller, H., and Bork, H. R. (2011). A multi-layered prehistoric settlement structure (tell?) at Niederröblingen, Germany and its implications. Journal of Archaeological Science, 38, 1101–10.CrossRefGoogle Scholar
Luce, C. H. (2002). Hydrological processes and pathways affected by forest roads: what do we still need to learn? Hydrological Processes, 16, 2901–4.CrossRefGoogle Scholar
Luce, C. H. and Black, T. A. (1999). Sediment production from forest roads in western Oregon. Water Resources Research, 35, 2561–70.CrossRefGoogle Scholar
Lucha, P., Cardona, F., Gutiérrez, F., and Guerrero, J. (2008). Natural and human-induced dissolution and subsidence processes in the salt outcrop of the Cardona Diapir (NE Spain). Environmental Geology, 53, 1023–35.CrossRefGoogle Scholar
Luo, H. R., Smith, L. M., Allen, B. L., and Haukos, D. A. (1997). Effects of sedimentation on playa wetland volume. Ecological Applications, 7, 247–52.CrossRefGoogle Scholar
Luzón, M. A., Pérez, A., Borrego, A. G., Mayayo, M. J., and Soria, A. R. (2011). Interrelated continental sedimentary environments in the central Iberian Range (Spain): facies characterization and main palaeoenvironmental changes during the Holocene. Sedimentary Geology, 239, 87103.CrossRefGoogle Scholar
Lwenya, C. and Yongo, E. (2010). Human aspects of siltation of Lake Baringo: causes, impacts and interventions. Aquatic Ecosystem Health & Management, 13, 437–41.CrossRefGoogle Scholar
Lyell, C. (1835). Principles of Geology (4th edn). London: MurrayGoogle Scholar
Lyell, C. (1875). Principles of Geology (12th edn). London: MurrayGoogle Scholar
Lynas, M. (2011). The God species. London: Fourth Estate.Google Scholar
Lyons, R., Tooth, S., and Duller, G. A. (2013). Chronology and controls of donga (gully) formation in the upper Blood River catchment, KwaZulu-Natal, South Africa: Evidence for a climatic driver of erosion. The Holocene, 23, 1875–87.CrossRefGoogle Scholar
Ma, R., Duan, H., Hu, C., Feng, X., Li, A., Ju, W., and Yang, G. (2010). A half‐century of changes in China’s lakes: Global warming or human influence? Geophysical Research Letters, 37, L24106, DOI: 10.1029/2010GL045514.CrossRefGoogle Scholar
Macdonald, K. C. (1997). More forgotten tells of Mali: an archaeologist’s journey from here to Timbuktu. Archaeology International, 1, 40–2.CrossRefGoogle Scholar
Macklin, M. G. and Lewin, J. (1986). Terraced fills of Pleistocene and Holocene age in the Rheidol Valley, Wales. Journal of Quaternary Science, 1, 2134.CrossRefGoogle Scholar
Macklin, M. G. and Woodward, J. C. (2009). River systems and environmental change. In The Physical Geography of the Mediterranean, ed. Woodward, J. C., Oxford: Oxford University Press. pp. 319–52.CrossRefGoogle Scholar
Macklin, M. G., Jones, A. F. and Lewin, J. (2010). River response to rapid Holocene environmental change: evidence and explanation in British catchments. Quaternary Science Reviews, 29, 1555–76.CrossRefGoogle Scholar
Macklin, M. G., Lewin, J., and Jones, A. F. (2013). River entrenchment and terrace formation in the UK Holocene. Quaternary Science Reviews, 76, 194206.CrossRefGoogle Scholar
Macklin, M. G., Lewin, J. and Jones, A. F. (2014). Anthropogenic alluvium: an evidence-based meta-analysis for the UK Holocene. Anthropocene, 6, 2638.CrossRefGoogle Scholar
Macklin, M. G., Passmore, D. G., Stevenson, A. C., Colwey, A. C., Edwards, D. N., and O’Brien, C. F. (1991). Holocene alluviation and land-use change on Callaly Moor, Northumberland, England. Journal of Quaternary Science, 6, 225–32.CrossRefGoogle Scholar
Maghsoudi, M., Simpson, I. A., Kourampas, N., and Nashli, H. F. (2014). Archaeological sediments from settlement mounds of the Sagzabad Cluster, central Iran: human-induced deposition on an arid alluvial plain. Quaternary International, 324, 6783.CrossRefGoogle Scholar
Maguregui, M., Sarmiento, A., Martinez-Arkarazo, I., Angulo, M., Castro, K., Arana, G., and Madariaga, J. M. (2008). Analytical diagnosis methodology to evaluate nitrate impact on historical building materials. Analytical and Bioanalytical Chemistry, 391, 13611370.Google Scholar
Mahowald, N., Albani, S., Kok, J. F., Engelstaeder, S., Scanza, R., Ward, D. S., and Flanner, M. G. (2014). The size distribution of desert dust aerosols and its impact on the Earth system. Aeolian Research, 15, 5371.CrossRefGoogle Scholar
Mahowald, N. M. and 19 others. (2010). Observed 20th century desert dust variability: impact on climate and biogeochemistry. Atmospheric Chemistry and Physics, 10, 10875–93.CrossRefGoogle ScholarPubMed
Maina, J., de Moel, H., Zinke, J., Madin, J., McClanahan, T., and Vermaat, J. E. (2013). Human deforestation outweighs future climate change impacts of sedimentation on coral reefs. Nature Communications, 4, article number: 1986, doi:, doi: 10.1038/ncomms2986.CrossRefGoogle Scholar
Majer, E. L., Baria, R., Start, M., Oates, S., Bommer, J., Smith, B., and Asanuma, H. (2007). Induced seismicity associated with Enhanced Geothermal Systems. Geothermics, 36, 185222.CrossRefGoogle Scholar
Majewski, M. (2014). Human impact on Subatlantic slopewash processes and landform development at Lake Jasień (northern Poland). Quaternary International, 324, 5666.Google Scholar
Major, J. and 10 others. (2008). Initial fluvial response to the removal of Oregon’s Marmot Dam. Eos, 27, 241–3.CrossRefGoogle Scholar
Maley, J., Giresse, P., Doumenge, C., and Favier, C. (2012). Comment on “Intensifying weathering and land use in Iron Age Central Africa.” Science, 337, 1040.CrossRefGoogle Scholar
Malm, W. C., Schichtel, B. A., Ames, R. B., and Gebhart, K. A. (2002). A 10-year spatial and temporal trend of sulfate across the United States. Journal of Geophysical Research, 107 (D22), article no. 4627.CrossRefGoogle Scholar
Mancini, F., Stecchi, F., Zanni, M., and Gabbianelli, G. (2009). Monitoring ground subsidence induced by salt mining in the city of Tuzla (Bosnia and Herzegovina). Environmental Geology, 58, 381–9.CrossRefGoogle Scholar
Mann, D. H. and Meltzer, D. J. (2007). Millennial-scale dynamics of valley fills over the past 12000 14C yr in northeastern New Mexico. Bulletin of the Geological Society of America, 119, 1433–48.CrossRefGoogle Scholar
Mann, K. C., Peck, J. A., and Peck, M. C. (2013). Assessing dam pool sediment for understanding past, present and future watershed dynamics: an example from the Cuyahoga River, Ohio. Anthropocene, 2, 7688.CrossRefGoogle Scholar
Mao, D., Wang, Z., Li, L., Song, K., and Jia, M. (2014). Quantitative assessment of human-induced impacts on marshes in Northeast China from 2000 to 2011. Ecological Engineering, 68, 97104.CrossRefGoogle Scholar
Marden, M., Arnold, G. Seymour, A., and Hambling, R. (2012). History and distribution of steepland gullies in response to land use change, East Coast Region, North Island, New Zealand. Geomorphology, 153–154, 8190.CrossRefGoogle Scholar
Marfai, M. A. and King, L. (2007). Monitoring land subsidence in Semarang, Indonesia. Environmental Geology, 53, 651–9.Google Scholar
Mark, A. F. and McSweeney, G. D. (1990). Patterns of impoverishment in natural communities: case studies in forest ecosystems – New Zealand. In The Earth in Transition: Patterns and Processes of Biotic Impoverishment, ed. Woodwell, G. M.. Cambridge: Cambridge University Press, pp. 151–76.CrossRefGoogle Scholar
Marker, M. E. (1967). The Dee estuary: its progressive silting and salt marsh development. Transactions of the Institute of British Geographers, 41, 6571.CrossRefGoogle Scholar
Marren, P. M., Grove, J. R., Webb, J. A., and Stewardson, M. J. (2014). The potential for dams to impact lowland meandering river floodplain geomorphology. The Scientific World Journal, 309673, http://dx.doi.org/10.1155/2014/309673 (accessed, 12th November, 2015).CrossRefGoogle Scholar
Mars, J. C. and Houseknecht, D. W. (2007). Quantitative remote sensing study indicates doubling of coastal erosion rate in past 50 yr along a segment of the Arctic coast of Alaska. Geology, 35, 583–6.CrossRefGoogle Scholar
Marsh, B. and Kealhofer, L. (2014). Scales of impact: Settlement history and landscape change in the Gordion Region, central Anatolia. The Holocene, 24, 689701.Google Scholar
Marsh, G. P. (1864). Man and Nature. New York: Scribner.Google Scholar
Marshall, E., Weinberg, M., Wunder, S., and Kaphengst, T. (2011). Environmental dimensions of bioenergy development. EuroChoices, 10, 43–8.CrossRefGoogle Scholar
Marshall, E. J. P., Wade, P. M., and Clare, P. (1978). Land drainage channels in England and Wales. Geographical Journal, 144, 254–63.CrossRefGoogle Scholar
Marston, R. A. and Dolan, L. S. (1999). Effectiveness of sediment control structures relative to spatial patterns of upland soil loss in an arid watershed, Wyoming. Geomorphology, 31, 313–23.CrossRefGoogle Scholar
Marston, R. A., Bravard, J. P., and Green, T. (2003). Impacts of reforestation and gravel mining on the Malnant River, Haute-Savoie, French Alps. Geomorphology, 55, 6574.CrossRefGoogle Scholar
Marszałek, M., Alexandrowicz, Z., and Rzepa, G. (2014). Composition of weathering crusts on sandstones from natural outcrops and architectonic elements in an urban environment. Environmental Science and Pollution Research, 21, 14023–36.Google Scholar
Martin, M. A., Levermann, A., and Winkelmann, R. (2015). Comparing ice discharge through West Antarctic Gateways: Weddell vs. Amundsen Sea warming. The Cryosphere Discussions, 9, 1705–33.CrossRefGoogle Scholar
Martin, Y. E., Johnson, E. A., Gallaway, J. M., and Chaikina, O. (2011). Negligible soil erosion in a burned mountain watershed, Canadian Rockies: field and modelling investigations considering the role of duff. Earth Surface Processes and Landforms, 36, 2097–113.CrossRefGoogle Scholar
Martínez, M. L., Gallego-Fernández, J. B., and Hesp, P. A. (2013a). Restoration of Coastal Dunes. Berlin and Heidelberg: Springer.CrossRefGoogle Scholar
Martínez, M. L., Hesp, P. A., and Gallego-Fernández, J. B. (2013b). Coastal dune restoration: trends and perspectives. In Restoration of Coastal Dunes, ed. Martínez, M. L., Gallego-Fernández, J. B., and Hesp, P. A.. Berlin and Heidelberg: Springer, pp. 323–39.CrossRefGoogle Scholar
Martínez-Casasnovas, J. A. and Ramos, M. C. (2006). The cost of soil erosion in vineyard fields in the Penedès–Anoia Region (NE Spain). Catena, 68, 194–9.CrossRefGoogle Scholar
Martínez‐Casasnovas, J. A. and Ramos, M.C. (2009). Soil alteration due to erosion, ploughing and levelling of vineyards in north east Spain. Soil Use and Management, 25, 183–92.CrossRefGoogle Scholar
Martínez‐Casasnovas, J. A., Ramos, M. C., and García‐Hernández, D. (2009). Effects of land‐use changes in vegetation cover and sidewall erosion in a gully head of the Penedès region (northeast Spain). Earth Surface Processes and Landforms, 34, 1927–37.CrossRefGoogle Scholar
Martínez-Casasnovas, J. A., Ramos, M. C., and Ribes-Dasi, M. (2005). On-site effects of concentrated flow erosion in vineyard fields: some economic implications. Catena, 60, 129–46.CrossRefGoogle Scholar
Martinez Raya, A., Duran Zuazo, V. H. and Francia Martinez, J. R. (2006). Soil erosion and runoff response to plant-cover strips on semiarid slopes (SE Spain). Land Degradation and Development, 17, 111.CrossRefGoogle Scholar
Martín-Vide, J. P., Ferrer-Boix, C., and Ollero, A. (2010). Incision due to gravel mining: modeling a case study from the Gállego River, Spain. Geomorphology, 117, 261–71.CrossRefGoogle Scholar
Marx, S. K., McGowan, H. A., Kamber, B. S., Knight, J. M., Denholm, J., and Zawadzki, A. (2014). Unprecedented wind erosion and perturbation of surface geochemistry marks the Anthropocene in Australia. Journal of Geophysical Research: Earth Surface 119, 4561.CrossRefGoogle ScholarPubMed
Maselli, V. and Trincardi, F. (2013). Man made deltas. Scientific Reports, 3, doi:10.1038/srep01926.CrossRefGoogle Scholar
Massey, S. W. (1999). The effects of ozone and NOx on the deterioration of calcareous stone. Science of the Total Environment, 227, 109–21.CrossRefGoogle Scholar
Massonnet, D., Holzer, T., and Vadon, H. (1997). Land subsidence caused by the East Mesa geothermal field, California, observed using SAR interferometry. Geophysical Research Letters, 24, 901–4.CrossRefGoogle Scholar
Matless, D. (2014). Nature of Landscape: Cultural Geography of the Norfolk Broads. Wiley Blackwell: Chichester and Oxford.CrossRefGoogle Scholar
Mattheus, C. R., Rodriguez, A. B., McKee, B. A., and Curring, C. A. (2010). Impact of land-use change and hard structures on the evolution of fringing marsh shorelines. Estuarine, Coastal and Shelf Science, 88, 365–76.CrossRefGoogle Scholar
Mattsson, T., Kortelainen, P., Räike, A., Lepistö, A., and Thomas, D. N. (2015). Spatial and temporal variability of organic C and N concentrations and export from 30 boreal rivers induced by land use and climate. Science of the Total Environment, 508, 145–54.CrossRefGoogle ScholarPubMed
May, V. (2015). Coastal cliff conservation and management: the Dorset and East Devon Coast World Heritage Site. Journal of Coastal Conservation, 19, 821–9, doi:10.1007/s11852-014-0358-4.CrossRefGoogle Scholar
May, V. J. and Hansom, J. D. (2003). Coastal Geomorphoogy of Great Britain. Peterborough: Joint Nature Conservation Committee.Google Scholar
Mazzoldi, A., Rinaldi, A. P., Borgia, A., and Rutqvist, J. (2012). Induced seismicity within geological carbon sequestration projects: maximum earthquake magnitude and leakage potential from undetected faults. International Journal of Greenhouse Gas Control, 10, 434–42.CrossRefGoogle Scholar
McCabe, S., Smith, B. J., McAlister, J. J., Gomez-Heras, M., McAllister, D., Warke, P. A., and Basheer, P. A. M. (2013). Changing climate, changing process: implications for salt transportation and weathering within building sandstones in the UK. Environmental Earth Sciences, 69, 1225–35.CrossRefGoogle Scholar
McCarney-Castle, K., Voulgaris, G., and Kettner, A. J. (2010). Analysis of fluvial suspended sediment load contribution through Anthropocene history to the South Atlantic Bight coastal zone, U.S.A. Journal of Geology, 118, 399416.CrossRefGoogle Scholar
McConnell, J. R., Aristarain, A. J., Banta, J. R., Edwards, P. R., and Simões, J. C. (2007). 20th-century doubling in dust archived in an Antarctic Peninsula ice core parallels climate change and desertification in South America. Proceedings National Academy of Sciences, 104, 5743–8.CrossRefGoogle Scholar
McCorriston, J. and Oches, E. (2001). Two early Holocene check dams from Southern Arabia. Antiquity, 75, 675–6.CrossRefGoogle Scholar
McCulloch, M., Fallon, S., Wyndham, T., Hendy, E., Lynch, J., and Barnes, D. (2003). Coral record of increased sediment flux to the inner Great Barrier Reef since European settlement. Nature, 421, 727–30.CrossRefGoogle Scholar
McCulloch, M., Falter, J., Trotter, J., and Montagna, P. (2012). Coral resilience to ocean acidification and global warming through pH up-regulation. Nature Climate Change, 2, 623–7.CrossRefGoogle Scholar
McDonald, A., Lane, S. N., Haycock, N. E., and Chalk, E. A. (2004). Rivers of dreams: on the gulf between theoretical and practical aspects of an upland river restoration. Transactions of the Institute of British Geographers, 29, 257–81.CrossRefGoogle Scholar
McFadden, L., Spencer, T., and Nicholls, R. J. (2007). Broad-scale modelling of coastal wetlands: what is required? Hydrobiologia, 577, 515.CrossRefGoogle Scholar
McFadden, L. D. and McAuliffe, J. R. (1997). Lithologically influenced geomorphic responses to Holocene climatic changes in the Southern Colorado Plateau, Arizona: a soil-geomorphic and ecologic perspective. Geomorphology, 19, 303–32.CrossRefGoogle Scholar
McGarr, A. (2014). Maximum magnitude earthquakes induced by fluid injection. Journal of Geophysical Research: Solid Earth, 119, 1008–19.Google Scholar
McGarr, A., Bekins, B., Burkardt, N., Dewey, J., Earle, P., Ellsworth, W., and Sheehan, A. (2015). Coping with earthquakes induced by fluid injection. Science, 347, 830–1.CrossRefGoogle ScholarPubMed
McGlone, M. S. and Wilmshurst, J. M. (1999). Dating initial Maori environmental impact in New Zealand. Quaternary International, 59, 516.CrossRefGoogle Scholar
McMillan, B. R., Pfeiffer, K. A., and Kaufman, D. W. (2011). Vegetation responses to an animal-generated disturbance (bison wallows) in tallgrass prairie. The American Midland Naturalist, 165, 6073.CrossRefGoogle Scholar
McNeill, J. R. (2000). Something New Under the Sun: an Environmental History of the Twentieth-Century World. New York: W.W. Norton & Company.Google Scholar
McNeill, J. R. (2003). Resource exploitation and over-exploitation: a look at the 20th century. In Exploitation and Overexploitation in Societies Past and Present, ed. Benzing, T. S. and Herrmann, B.. Münster: LIT Verlag, pp. 5160.Google Scholar
McWethy, D. B., Whitlock, C., Wilmshurst, J. M., McGlone, M. S., and Li, X. (2009). Rapid deforestation of South Island, New Zealand, by early Polynesian fires. The Holocene, 19, 883–97.CrossRefGoogle Scholar
McWethy, D. B., Wilmshurst, J. M., Whitlock, C., Wood, J. R., and McGlone, M. S. (2014) A high-resolution chronology of rapid forest transitions following Polynesian arrival in New Zealand. PLoS ONE 9(11), e111328. doi:10.1371/journal.pone.0111328.CrossRefGoogle ScholarPubMed
Meade, R. H. (1991). Reservoirs and earthquakes. Engineering Geology, 30, 245–62.CrossRefGoogle Scholar
Meade, R. H. and Moody, J. A. (2010). Causes for the decline of suspended-sediment discharge in the Mississippi River system, 1940–2007. Hydrological Processes, 24, 3549.CrossRefGoogle Scholar
Meade, R. H. and Parker, R. S. (1985). Sediment in rivers in the United States. United States Geological Survey Water Supply Paper, 2276, 4960.Google Scholar
Meade, R. H. and Trimble, S. W. (1974). Changes in sediment loads in rivers of the Atlantic drainage of the United States since 1900. Publication of the International Association of Hydrological Science, 113, 99104.Google Scholar
Meadows, P. S. and Meadows, A. (1999). The Indus River: Biodiversity, Resources, Humankind. Karachi: Oxford University Press.Google Scholar
Megahan, W. F., Wilson, M. and Monsen, S. B. (2001). Sediment production from granitic cutslopes on forest roads in Idaho, USA. Earth Surface Processes and Landforms, 26, 153–63.3.0.CO;2-0>CrossRefGoogle Scholar
Meijer, A. D., Heitman, J. L., White, J. G., and Austin, R. E. (2013). Measuring erosion in long-term tillage plots using ground-based lidar. Soil and Tillage Research, 126, 110.CrossRefGoogle Scholar
Mekonnen, M. M., Hoekstra, A. Y., and Becht, R. (2012). Mitigating the water footprint of export cut flowers from the Lake Naivasha Basin, Kenya. Water Resources Management, 26, 3725–42CrossRefGoogle Scholar
Mendonça, A., Fortes, C. J., Capitão, R., Neves, M. G., Antunes do Carmo, J. S., and Moura, T. (2011). Hydrodynamics around an artificial surfing reef at Leirosa, Portugal. Journal of Waterway, Port, Coastal, and Ocean Engineering, 138, 226–35.Google Scholar
Mendoza‐González, G., Martínez, M. L., Rojas‐Soto, O. R., Vázquez, G., and Gallego‐Fernández, J. B. (2013). Ecological niche modeling of coastal dune plants and future potential distribution in response to climate change and sea level rise. Global Change Biology, 19, 2524–35.CrossRefGoogle ScholarPubMed
Menze, B. H. and Ur, J. A. (2012). Mapping patterns of long-term settlement in Northern Mesopotamia at a large scale. Proceedings of the National Academy of Sciences, 109, E778–87.CrossRefGoogle ScholarPubMed
Metzidakis, I., Martinez-Vilela, A., Castro Nieto, G., and Basso, B. (2008). Intensive olive orchards on sloping land: good water and pest management are essential. Journal of Environmental Management, 89, 120–8.CrossRefGoogle ScholarPubMed
Meulen, F. and Salman, A. H. P. M. (1996). Management of Mediterranean coastal dunes. Ocean and Coastal Management, 30, 177–95.Google Scholar
Meulen, F. van der and Jungerius, P. D. (1989). Landscape development in Dutch coastal dunes: the breakdown and restoration of geomorphological and geohydrological processes. Proceedings of the Royal Society of Edinburgh. Section B. Biological Sciences, 96, 219–29.Google Scholar
Meunier, J. D., Braun, J. J., Riotte, J., Kumar, C., and Sekhar, M. (2011). Importance of weathering and human perturbations on the riverine transport of Si. Applied Geochemistry, 26, S360–2.CrossRefGoogle Scholar
Meusberger, K. and Alewell, C. (2008). Impacts of anthropogenic and environmental factors on the occurrence of shallow landslides in an alpine catchment (Urseren Valley, Switzerland). Natural Hazards and Earth System Sciences, 8, 509–20.Google Scholar
Meybeck, M. (2003). Global analysis of river systems: from Earth system controls to Anthropocene syndromes. Philosophical Transactions of the Royal Society B: Biological Sciences, 358, 1935–55.CrossRefGoogle ScholarPubMed
Meybeck, M. and Vörösmarty, C. (2005). Fluvial filtering of land-to-ocean fluxes: from natural Holocene variations to Anthropocene. Comptes Rendus Geoscience, 337, 107–23.CrossRefGoogle Scholar
Meyer, A. and Martınez-Casasnovas, J. A. (1999). Prediction of existing gully erosion in vineyard parcels of the NE Spain: a logistic modelling approach. Soil and Tillage Research, 50, 319–31.CrossRefGoogle Scholar
Meyfroidt, P. and Lambin, E. (2009). Geographic and historical patterns of reforestation. Bulletin des Séances Academie Royale des Sciences d’outre-mer, 55, 477502.Google Scholar
Micheli, F., Halpern, B. S., Walbridge, S., Ciriaco, S., Ferretti, F., Fraschetti, S., and Rosenberg, A. A. (2013). Cumulative human impacts on mediterranean and black sea marine ecosystems: assessing current pressures and opportunities. PloS one, 8(12), e79889.CrossRefGoogle ScholarPubMed
Micheletti, N., Lane, S. N., and Chandler, J. H. (2015). Application of archival aerial photogrammetry to quantify climate forcing of Alpine landscapes. The Photogrammetric Record, doi: 10.1111/phor.12099.CrossRefGoogle Scholar
Middleton, N. J. (1990). Wind erosion and dust storm prevention. In Desert Reclamation, ed. Goudie, A. S.. Wiley, Chichester: Wiley, pp. 87108.Google Scholar
Mietton, M., Cordier, S., Frechen, M., Dubar, M., Beiner, M., and Andrianaivoarivony, R. (2014). New insights into the age and formation of the Ankarokaroka lavaka and its associated sandy cover (NW Madagascar, Ankarafantsika natural reserve). Earth Surface Processes and Landforms, 39, 1467–77.CrossRefGoogle Scholar
Migoń, P. (Ed.). (2010). Geomorphological Landscapes of the World. Springer.CrossRefGoogle Scholar
Migoń, P. (2014). The significance of landforms–the contribution of geomorphology to the World Heritage Programme of UNESCO. Earth Surface Processes and Landforms, 39, 836843.CrossRefGoogle Scholar
Miller, A. J. (2011). Identifying landslide activity as a function of economic development: a case study of increased landslide frequency surrounding Dominical, Costa Rica. Environment, Development and Sustainability, 13, 901–21.CrossRefGoogle Scholar
Miller, A. J. and Zégre, N. P. (2014). Mountaintop removal mining and catchment hydrology. Water, 6, 472–99.CrossRefGoogle Scholar
Miller, J. D., Kim, H., Kjeldsen, T. R., Packman, J., Grebby, S., and Dearden, R. (2014). Assessing the impact of urbanization on storm runoff in a peri-urban catchment using historical change in impervious cover. Journal of Hydrology, 515, 5970.CrossRefGoogle Scholar
Miller, L. and Douglas, B. C. (2004). Mass and volume contribution to twentieth-century global sea level rise. Nature, 428, 406–9.CrossRefGoogle ScholarPubMed
Miller, R. L. and Tegen, I. (1998). Climate response to soil dust aerosols. Journal of Climate, 11, 3247–67.2.0.CO;2>CrossRefGoogle Scholar
Miller Rosen, A. (1997). The geoarchaeology of Holocene environments and land use at Kazane Höyük, SE Turkey. Geoarchaeology, 12, 395416.3.0.CO;2-W>CrossRefGoogle Scholar
Milliman, J. D., Qin, Y. S., Ren, M. E., and Saita, Yoshiki (1987). Man’s influence on erosion and transport of sediment by Asian rivers: the Yellow River (Huanghe) example. Journal of Geology, 95, 751–62.CrossRefGoogle Scholar
Millspaugh, S. H. and Whitlock, C. (1995). A 750-year fire history based on lake sediment records in central Yellowstone National Park, USA. The Holocene, 5, 283–92.CrossRefGoogle Scholar
Min, S.-K., Zhang, X., Zwiers, F. W., and Hegerl, G. C. (2011). Human contribution to more-intense precipitation extremes. Nature, 470, 378–81.CrossRefGoogle ScholarPubMed
Miyasaka, T., Okuro, T., Miyamori, E., Zhao, X., and Takeuchi, K. (2014). Effects of different restoration measures and sand dune topography on short-and long-term vegetation restoration in northeast China. Journal of Arid Environments 111, 16.CrossRefGoogle Scholar
Moffet, C. A., Pierson, F. B., Robichaud, P. R., Spaeth, K. E., and Hardegree, S. P. (2007). Modeling soil erosion on steep sagebrush rangeland before and after prescribed fire. Catena, 71, 218–28.CrossRefGoogle Scholar
Mokhtari, D., Karami, F., and Bayati Khatibi, M. (2011). An investigation on Parsian alluvial fan’s tufa in northwest of Iran and its implications for the Holocene tufa decline. Geography and Environmental Planning, 22, 122.Google Scholar
Mölg, N. (2014). Hasty retreat of glaciers in northern Patagonia from 1985–2011. Journal of Glaciology, 60, 1033–43.Google Scholar
Molg, T., Georges, C., and Kaser, G. (2003). The contribution of increased incoming shortwave radiation to the retreat of the Rwenzori Glaciers, East Africa, during the 20th Century. International Journal of Climatology, 23, 291303.CrossRefGoogle Scholar
Moncel, M.-H. (2010). Oldest human expansions in Eurasia: favouring and limiting factors. Quaternary International, 223–224, 19.CrossRefGoogle Scholar
Monteith, D. T., Stoddard, J. L., Evans, C. D., de Wit, H. A., Forsius, M., Høgåsen, T., and Vesely, J. (2007). Dissolved organic carbon trends resulting from changes in atmospheric deposition chemistry. Nature, 450, 537–40.CrossRefGoogle ScholarPubMed
Montgomery, D. R. (1997). What’s best on the banks? Nature, 388, 328–9.CrossRefGoogle Scholar
Montgomery, D. R. (2007). Dirt: The Erosion of Civilizations. Berkeley: University of California Press.CrossRefGoogle Scholar
Moody, J. A. and Martin, D. A. (2001). Initial hydrologic and geomorphic response following a wildfire in the Colorado Front Range. Earth Surface Processes and Landforms, 26, 1049–70.CrossRefGoogle Scholar
Moody, J. A., Shakesby, R. A., Robichaud, P. R., Cannon, S. H., and Martin, D. A. (2013). Current research issues related to post-wildfire runoff and erosion processes. Earth-Science Reviews, 122, 1037.CrossRefGoogle Scholar
Moore, P. D. (1987). Man and mire: a long and wet relationship. Transactions of the Botanical Society of Edinburgh, 45, 7795.CrossRefGoogle Scholar
Moore, P. D. (2002). The future of cool temperate bogs. Environmental Conservation, 29, 320.CrossRefGoogle Scholar
Moosdorf, N., Renforth, P., and Hartmann, J. (2014). Carbon dioxide efficiency of terrestrial enhanced weathering. Environmental Science & Technology, 48, 4809–16.CrossRefGoogle ScholarPubMed
Mora, J. W. and Burdick, D. M. (2013). The impact of man-made earthen barriers on the physical structure of New England tidal marshes (USA). Wetlands Ecology and Management, 21, 387–98.CrossRefGoogle Scholar
Morales, C. (ed.) (1979). Saharan Dust, Chichester: John Wiley.Google Scholar
Morgan, J. A. and 14 others (2004). Water relations in grassland and desert ecosystems exposed to elevated atmospheric CO2. Oecologia, 140, 1125.CrossRefGoogle Scholar
Morgan, R. P. C. (2005). Soil Erosion and Conservation (3rd edn). Oxford: Blackwell Publishing.Google Scholar
Moritz, M. A., Parisien, M. A., Batllori, E., Krawchuk, M. A., Van Dorn, J., Ganz, D. J., and Hayhoe, K. (2012). Climate change and disruptions to global fire activity. Ecosphere, 3(6), art49.CrossRefGoogle Scholar
Morris, S. E. and Moses, T. A. (1987). Forest fire and the natural soil erosion regime in the Colorado Front Range. Annals of the Association of American Geographers, 77, 245–54.CrossRefGoogle Scholar
Mortimore, M. (1989). Adapting to Drought: Farmers, Famines, and Desertification in West Africa. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Morton, R. A., Bernier, J. C., and Barras, J. A. (2006). Evidence of regional subsidence and associated wetland loss induced by hydrocarbon production, Gulf Coast region, USA. Environmental Geology, 50, 261–74.CrossRefGoogle Scholar
Morton, R. A., Bernier, J. C., Barras, J. A., and Ferina, N. F. (2005). Historical subsidence and wetland loss in the Mississippi Delta Plain. Gulf Coast Association of Geological Societies, Transactions, 55, 555–71.Google Scholar
Mortsch, L., Hengeveld, H., Lister, M., Wenger, L., Lofgren, B., Quinn, F., and Slivitzky, M. (2000). Climate change impacts on the hydrology of the Great Lakes-St. Lawrence system. Canadian Water Resources Journal, 25, 153–79.CrossRefGoogle Scholar
Morzaria-Luna, H. N., Turk-Boyer, P., Rosemartin, A., and Camacho-Ibar, V. F. (2014). Vulnerability to climate change of hypersaline salt marshes in the Northern Gulf of California. Ocean & Coastal Management, 93, 3750.CrossRefGoogle 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.CrossRefGoogle Scholar
Motyka, R. J., O’Neal, S., Connor, C. L., and Echelmeyer, K. A. (2002). Twentieth century thinning of Mendenhall Glacier, Alaska, and its relationship to climate, lake calving and glacier run-off. Global and Planetary Change, 35, 93112.CrossRefGoogle Scholar
Moulin, C. and Chiapello, I. (2006). Impact of human‐induced desertification on the intensification of Sahel dust emission and export over the last decades. Geophysical Research Letters, 33, doi: 10.1029/2006GL025923.CrossRefGoogle Scholar
Mousavi, M. E., Irish, J. L., Frey, A. E., Olivera, F., and Edge, B. L. (2011). Global warming and hurricanes: the potential impact of hurricane intensification and sea level rise on coastal flooding. Climatic Change, 104, 575–97.CrossRefGoogle Scholar
Mugagga, F., Kakembo, V., and Buyinza, M. (2012). Land use changes on the slopes of Mount Elgon and the implications for the occurrence of landslides. Catena, 90, 3946.CrossRefGoogle Scholar
Muhly, J. D. (1997). Artifacts of the Neolithic, Bronze and Iron Ages. In The Oxford Encyclopaedia of Archaeology in the Near East, ed. Myers, E. M.. New York: Oxford University Press, vol. 4, pp. 515.Google Scholar
Mulitza, S. and 10 others. (2010). Increase in African dust flux at the onset of commercial agriculture in the Sahel region. Nature, 466, 226–8.CrossRefGoogle ScholarPubMed
Mulla, D. J. and Sekely, A. C. (2009). Historical trends affecting accumulation of sediment and phosphorus in Lake Pepin, upper Mississippi River, USA. Journal of Paleolimnology, 41, 589602.CrossRefGoogle Scholar
Mullan, D. (2013). Soil erosion under the impacts of future climate change: assessing the statistical significance of future changes and the potential on-site and off-site problems. Catena, 109, 234–46.CrossRefGoogle Scholar
Müller‐Nedebock, D. and Chaplot, V. (2015). Soil carbon losses by sheet erosion: a potentially critical contribution to the global carbon cycle. Earth Surface Processes and Landforms, 40, 1803–13.CrossRefGoogle Scholar
Munson, S., Belnap, J., and Okin, G. S. (2011). Responses of wind erosion to climate-induced vegetation changes on the Colorado Plateau. Proceedings of the National Academy of Sciences, 108, 3854–9.CrossRefGoogle ScholarPubMed
Murray, N. J., Clemens, R. S., Phinn, S. R., Possingham, H. P., and Fuller, R. A. (2014). Tracking the rapid loss of tidal wetlands in the Yellow Sea. Frontiers in Ecology and the Environment, 12, 267–72.CrossRefGoogle Scholar
Murray-Rust, D. H. (1972). Soil erosion and reservoir sedimentation in a grazing area west of Arusha, northern Tanzania. Geografiska Annaler, 54A, 325–43.Google Scholar
Murria, J. (1991). Subsidence due to oil production in Western Venezuela: engineering problems and solutions. IAHS Publication, 200, 129–39.Google Scholar
Muturi, G. M., Mohren, G. M. J., and Kimani, J. N. (2009). Prediction of Prosopis species invasion in Kenya using geographical information system techniques. African Journal of Ecology, 48, 628–36.Google Scholar
Myer, L. R. and Daley, T. M. (2011). Elements of a best practices approach to induced seismicity in geologic storage. Energy Procedia, 4, 3707–13.CrossRefGoogle Scholar
Nadeu, E., Gobin, A., Fiener, P., Wesemael, B., and Oost, K. (2015). Modelling the impact of agricultural management on soil carbon stocks at the regional scale: the role of lateral fluxes. Global Change Biology, doi: 10.1111/gcb.12889.CrossRefGoogle Scholar
Nagel, N. B. (2001). Compaction and subsidence issues within the petroleum industry: from Wilmington to Ekofisk and beyond. Physics and Chemistry of the Earth, Part A: Solid Earth and Geodesy, 26, 314.CrossRefGoogle 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, 259–77.CrossRefGoogle Scholar
Nakamura, F., Sudo, T., Kameyama, S., and Jitsu, M. (1997). Influences of channelization on discharge of suspended sediment and wetland vegetation in Kushiro Marsh, northern Japan. Geomorphology, 18, 279–89.CrossRefGoogle Scholar
Nakamura, K., Tockner, K., and Amano, K. (2006). River and wetland restoration: lessons from Japan. BioScience, 56, 419–29.CrossRefGoogle Scholar
Nakano, T. and Matsuda, I. (1976). A note on land subsidence in Japan. Geographical Reports of Tokyo Metropolitan University, 11, 147–62.Google Scholar
Natural England. (2009). Assessing impacts of wind farm development on blanket peatland in England. Project Report and Guidance, Final Report, http://publications.naturalengland.org.uk/publication/43010 (Accessed June 4, 2015).Google Scholar
Navarro-Pons, M., Muñoz-Perez, J. J., Román-Sierra, J., Tejedor, B., Rodriguez, I., and Gomez-Pina, G. (2007). Morphological evolution in the migrating dune of Valdevaqueros (SW Spain) during an eleven-year period. In International Conference on Management and Restoration of Coastal Dunes, Santander, Spain, pp. 80–5.Google Scholar
Navratil, O., Breil, P., Schmitt, L., Grosprêtre, L., and Albert, M. B. (2013). Hydrogeomorphic adjustments of stream channels disturbed by urban runoff (Yzeron River basin, France). Journal of Hydrology, 485, 2436.CrossRefGoogle Scholar
Nawaz, M. F., Bourrie, G., and Trolard, F. (2013). Soil compaction impact and modelling. A review. Agronomy for Sustainable Development, 33, 291309.CrossRefGoogle Scholar
Nearing, M. A. (2001). Potential changes in rainfall erosivity in the US with climate change during the 21st century. Journal of Soil and Water Conservation, 56, 229–32.Google Scholar
Nearing, W. and 11 others (2005). Modeling response of soil erosion and runoff to changes in precipitation and cover. Catena, 61, 131–54.CrossRefGoogle Scholar
Neave, M., Rayburg, S., and Swan, A. (2009). River channel change following dam removal in an ephemeral stream. Australian Geographer, 40, 235–46.CrossRefGoogle Scholar
Neff, J. C. and 9 others. (2008). Increasing eolian dust deposition in the western United States linked to human activity. Nature Geoscience, 1, 189–95.CrossRefGoogle Scholar
Nelson, F. E. and Anisimov, O. A. (1993). Permafrost zonation in Russia under anthropogenic climate change. Permafrost and Periglacial Processes, 4, 137–48.CrossRefGoogle Scholar
Neris, J., Tejedor, M., Fuentes, J., and Jiménez, C. (2013). Infiltration, runoff and soil loss in Andisols affected by forest fire (Canary Islands, Spain). Hydrological Processes, 27, 2814–24.CrossRefGoogle Scholar
Nesje, A., Lie, O., and Dahl, S. O. (2000). Is the North Atlantic Oscillation reflected in glacier mass balance records?’ Journal of Quaternary Science, 15, 587601.3.0.CO;2-2>CrossRefGoogle Scholar
Nesje, A., Bakke, J., Dahl, S. O., Lie, O., and Matthews, J. A. (2008). Norwegian mountain glaciers in the past, present and future. Global and Planetary Change, 60, 1027.CrossRefGoogle Scholar
New, M., Todd, M., Hulme, M., and Jones, P. (2001). Precipitation measurements and trends in the twentieth century. International Journal of Climatology, 21, 1899–922.CrossRefGoogle Scholar
Newton, J. G. (1976). Induced and natural sinkholes in Alabama: continuing problem along highway corridors. In Subsidence over Mines and Caverns, ed. Zwanig, F. R.. Washington, DC: National Academy of Sciences, pp. 916.Google Scholar
Nichol, S. L., Augustinus, P. L., Gregory, M. R., Creese, R., and Horrocks, M. (2000). Geomorphic and sedimentary evidence of human impact on the New Zealand landscape. Physical Geography, 21, 109–32.CrossRefGoogle Scholar
Nicholls, R. J., Hoozemans, F. M. J., and Marchand, M. (1999). Increasing flood risk and wetland losses due to global sea level rise: regional and global analyses. Global Environmental Change, 9, S69–87.CrossRefGoogle Scholar
Nicholls, R. J., Marinova, N., Lowe, J. A., Brown, S., Vellinga, P., de Gusmão, D., Hinkel, J., and Tol, R. S. J. (2011). Sea-level rise and its possible impacts given a “beyond 4oC world” in the twenty-first century. Philosophical Transactions of the Royal Society, A, 369, 161–81.Google Scholar
Nichols, R. A. and Ketcheson, G. L. (2013). A two‐decade watershed approach to stream restoration log jam design and stream recovery monitoring: Finney Creek, Washington. Journal of the American Water Resources Association, 49, 1367–84.CrossRefGoogle Scholar
Nick, F. M., Vieli, A., Andersen, M. L., Joughin, I., Payne, A., Edwards, T. L., and van de Wal, R. S. (2013). Future sea-level rise from Greenland/’s main outlet glaciers in a warming climate. Nature, 497, 235–8.CrossRefGoogle Scholar
Nicod, J. (1986). Facteurs physico-chimiques de l’accumulation des formations travertineuses. Mediterranée, 10, 161–4.Google Scholar
Nicol, A., Gerstenberger, M., Bromley, C., Carne, R., Chardot, L., Ellis, S., and Viskovic, P. (2013). Induced seismicity; observations, risks and mitigation measures at CO2 Storage Sites. Energy Procedia, 37, 4749–56.CrossRefGoogle Scholar
Nicolay, A., Raab, A., Raab, T., Rösler, H., Bönisch, E., and Murray, A. S. (2014). Evidence of (pre-) historic to modern landscape and land use history near Jänschwalde (Brandenburg, Germany). Zeitschrift für Geomorphologie 58, Suppl. 2, 731.CrossRefGoogle Scholar
Nieuwenhuis, H. S. and Schokking, F. (1997). Land subsidence in drained peat areas of the Province of Friesland, The Netherlands. Quarterly Journal of Engineering Geology and Hydrogeology, 30, 3748.CrossRefGoogle Scholar
Nijssen, B., O’Donnell, G. M., Hamlet, A. F. and Lettenmaier, D. P. (2001). Hydrologic sensitivity of global rivers to climate change. Climatic Change, 50, 143–75.CrossRefGoogle Scholar
Nilsson, C., Reidy, C. A., Dynesius, M., and Revenga, C. (2005). Fragmentation and flow regulation of the world’s large river systems. Science, 308, 405–8.CrossRefGoogle ScholarPubMed
Nir, D. (1983). Man, a Geomorphological Agent: an Introduction to Anthropic Geomorphology. Jerusalem: Keter.Google Scholar
Nitto, D. D., Neukermans, G., Koedam, N., Defever, H., Pattyn, F., Kairo, J. G., and Dahdouh-Guebas, F. (2014). Mangroves facing climate change: landward migration potential in response to projected scenarios of sea level rise. Biogeosciences, 11, 857–71.CrossRefGoogle Scholar
Nolte, S., Esselink, P., Bakker, J. P., and Smit, C. (2014). Effects of livestock species and stocking density on accretion rates in grazed salt marshes. Estuarine, Coastal and Shelf Science, 185, 41–7.Google Scholar
Nordstrom, K. F. (1994). Beaches and dunes of human-altered coasts. Progress in Physical Geography, 18, 497516.CrossRefGoogle Scholar
Nordstrom, K. F. (2000). Beaches and Dunes of Developed Coasts. Cambridge University Press: Cambridge.CrossRefGoogle Scholar
Nordstrom, K. F. (2014). Living with shore protection structures: a review. Estuarine, Coastal and Shelf Science, 150, 1123.CrossRefGoogle Scholar
Nordstrom, K. F., and Hotta, S. (2004). Wind erosion from cropland in the USA: a review of problems, solutions and prospects. Geoderma, 121, 157–67.CrossRefGoogle Scholar
Nordstrom, K. F., Jackson, N. L., Freestone, A. L., Korotky, K. H., and Puleo, J. A. (2012). Effects of beach raking and sand fences on dune dimensions and morphology. Geomorphology, 179, 106–15.CrossRefGoogle ScholarPubMed
Nordstrom, K. F., Lampe, R., and Vandemark, L. M. (2000). Reestablishing naturally functioning dunes on developed coasts. Environmental Management 25, 3751.CrossRefGoogle Scholar
Noss, R. F. (2011). Between the devil and the deep blue sea: Florida’s unenviable position with respect to sea level rise. Climatic Change, 107, 116.CrossRefGoogle Scholar
Notebaert, B. and Verstraeten, G. (2010). Sensitivity of West and Central European river systems to environmental changes during the Holocene: a review. Earth-Science Reviews, 103, 163–82.CrossRefGoogle Scholar
Novara, A., Gristina, L., Saladino, S. S., Santoro, A., and Cerda, A. (2011). Soil erosion assessment on tillage and alternative soil managements in a Sicilian vineyard. Soil and Tillage Research, 117, 140–7.CrossRefGoogle Scholar
Nyssen, J. and Vermeersch, D. (2010). Slope aspect affects geomorphic dynamics of coal mining spoil heaps in Belgium. Geomorphology, 123, 109–21.CrossRefGoogle Scholar
Nyssen, J., Debever, M., Poesen, J., and Deckers, J. (2014). Lynchets in eastern Belgium—a geomorphic feature resulting from non-mechanised crop farming. Catena, 121, 164–75.CrossRefGoogle Scholar
Nyssen, J., Haile, M., Moeyersons, J., Poesen, J., and Deckers, J. (2000). Soil and water conservation in Tigray (Northern Ethiopia): the traditional dagat technique and its integration with introduced techniques. Land Degradation & Development, 11, 199208.3.0.CO;2-Y>CrossRefGoogle Scholar
Nyssen, J., Poesen, J., Moeyersons, J., Luyten, E., Veyret‐Picot, M., Deckers, J., and Govers, G. (2002). Impact of road building on gully erosion risk: a case study from the northern Ethiopian highlands. Earth Surface Processes and Landforms, 27, 1267–83.CrossRefGoogle Scholar
Nyssen, J., Veyret‐Picot, M., Poesen, J., Moeyersons, J., Haile, M., Deckers, J., and Govers, G. (2004). The effectiveness of loose rock check dams for gully control in Tigray, northern Ethiopia. Soil Use and Management, 20, 5564.CrossRefGoogle Scholar
Olthof, I., Fraser, R. H., and Schmitt, C. (2015). Landsat-based mapping of thermokarst lake dynamics on the Tuktoyaktuk Coastal Plain, Northwest Territories, Canada since 1985. Remote Sensing of Environment, 168, 194204.CrossRefGoogle Scholar
O’Neal, M. R., Nearing, M. A., Vining, R. C., Southworth, J., and Pfeifer, R. A. (2005). Climate change impacts on soil erosion in Midwest United States with changes in crop management. Catena, 61, 165–84.Google Scholar
O’Sullivan, P. E., Coard, M. A., and Pickering, D. A. (1982). The use of laminated lake sediments in the estimation and calibration of erosion rates. Publication of the International Association of Hydrological Science, 137, 385–96.Google Scholar
O’Hara, S. L., Street-Perrott, F. A., and Burt, T. P. (1993). Accelerated soil erosion around a Mexican highland lake caused by prehispanic agriculture. Nature, 362, 4851.CrossRefGoogle Scholar
Oliver, F. W. (1946). Dust-storms in Egypt as noted in Maryut: a supplement. Geographical Journal, 108, 221–6.CrossRefGoogle Scholar
Oliveira, P. T. S., Nearing, M. A., and Wendland, E. (2015). Orders of magnitude increase in soil erosion associated with land use change from native to cultivated vegetation in a Brazilian savannah environment. Earth Surface Processes and Landforms, 40, 1524–33.CrossRefGoogle Scholar
Olivier, S., Blaser, C., Brütsch, S., Frolova, N., Gäggeler, H. W., Henderson, K. A., Palmer, A. S., Papina, T., and Schwikowski, M. (2006). Temporal variations of mineral dust, biogenic tracers, and anthropogenic species during the past two centuries from Belukha ice core, Siberian Altai. Journal of Geophysical Research: Atmospheres 111, D5, doi: 10.1029/2005JD005830.CrossRefGoogle Scholar
Olley, J. M. and Wasson, R. J. (2003). Changes in the flux of sediment in the Upper Murrumbidgee catchment, Southeastern Australia, since European settlement. Hydrological Processes, 17, 3307–20.CrossRefGoogle Scholar
Olson, K. R., Gennadiyev, A. N., Zhidkin, A. P., Markelov, M. V., Golosov, V. N., and Lang, J. M. (2013). Use of magnetic tracer and radio-cesium methods to determine past cropland soil erosion amounts and rates. Catena, 104, 103–10.CrossRefGoogle Scholar
Oppenheimer, M. (1998). Global warming and the stability of the West Antarctic ice sheet. Nature, 393, 325–32.CrossRefGoogle Scholar
Ore, G. and Bruins, H. J. (2012). Design features of ancient agricultural terrace walls in the Negev Desert: human‐made geodiversity. Land Degradation & Development, 23, 409–18.CrossRefGoogle Scholar
Orme, A. R. (2002). Human imprints on the primeval landscape. In The Physical Geography of North America, ed. Orme, A. R.. New York: Oxford University Press, pp. 459–81.Google Scholar
Ormerod, S. J. (2004). A golden age of river restoration science? Aquatic Conservation: Marine and Freshwater Ecosystems, 14, 543–9.CrossRefGoogle Scholar
Orr, J. C, Pantoja, S., and Pörtner, H.-O. (2005a). Introduction to special section: the ocean in a high-CO2 world. Journal of Geophysical Research, 110 (C), doi: 10 1029/2005 JC 003086.CrossRefGoogle Scholar
Orr, J. C., and 26 others. (2005b). Anthropogenic ocean acidification over the twenty-first century and its impact on organisms. Nature, 437, 681686.CrossRefGoogle ScholarPubMed
Orts, W. J., Roa-Espinosa, A., Sojka, R. E., Glenn, G. M., Imam, S. H., Erlacher, K., and Pedersen, J. S. (2007). Use of synthetic polymers and biopolymers for soil stabilization in agricultural, construction and military applications. Journal of Materials in Civil Engineering, 19, 5866.CrossRefGoogle Scholar
Orwig, D. A. (2002). Ecosystem to regional impacts of introduced pests and pathogens: historical context, questions and issues. Journal of Biogeography, 29, 1471–4.CrossRefGoogle Scholar
Osterkamp, T. E. and Romanovsky, V. E. (1999). Evidence for warming and thawing of discontinuous permafrost in Alaska. Permafrost and Periglacial Processes, 10, 1737.3.0.CO;2-4>CrossRefGoogle Scholar
Osterkamp, T. E., Jorgenson, M. T., Schuur, E. A. G., Shur, Y. L., Kanevskiy, M. Z., Vogel, J. G., and Tumskoy, V. E. (2009). Physical and ecological changes associated with warming permafrost and thermokarst in interior Alaska. Permafrost and Periglacial Processes, 20, 235–6.CrossRefGoogle Scholar
Out, W. A. and Verhoeven, K. (2014). Late Mesolithic and Early Neolithic human impact at Dutch wetland sites: the case study of Hardinxveld-Giessendam De Bruin. Vegetation History and Archaeobotany, 23, 4156.CrossRefGoogle Scholar
Overeem, I., Anderson, R. S., Wobus, C. W., Clow, G. D., Urban, F. E., and Matell, N. (2011). Sea ice loss enhances wave action at the Arctic coast. Geophysical Research Letters, 38, doi: 10.1029/2011GL048681.CrossRefGoogle Scholar
Owen, L. A., Kamp, U., Khattak, G. A., Harp, E. L., Keefer, D. K., and Bauer, M. A. (2008). Landslides triggered by the 8 October 2005 Kashmir earthquake. Geomorphology, 94, 19.CrossRefGoogle Scholar
Ozer, P. (2003). Fifty years of African mineral dust production. Bulletin Scientifique Academie Royale Sciences d’Outre-Mer, 49, 371–96.Google Scholar
Özkan, H., Willcox, G., Graner, A., Salamini, F., and Kilian, B. (2011). Geographic distribution and domestication of wild emmer wheat (Triticum dicoccoides). Genetic Resources and Crop Evolution, 58, 1153.CrossRefGoogle Scholar
Padonou, E. A., Assogbadjo, A. E., Bachmann, Y., and Sinsin, B. (2013). How far bowalization affects phytodiversity, life forms and plant morphology in Sub‐humid tropic in West Africa. African Journal of Ecology, 51, 255–62.CrossRefGoogle Scholar
Padonou, E. A., Fandohan, B., Bachmann, Y., and Sinsin, B. (2014). How farmers perceive and cope with ‘bowalization’: a case study from West Africa. Land Use Policy, 36, 461–7.CrossRefGoogle Scholar
Page, M. J. and Trustrum, N. A. (1997). A late Holocene lake sediment record of the erosion response to land use change in a steepland catchment, New Zealand. Zeitschrift für Geomorphologie, 41, 369–92.CrossRefGoogle Scholar
Painter, T. H., Flanner, M. G., Kaser, G., Marzeion, B., VanCuren, R. A., and Abdalati, W. (2013). End of the Little Ice Age in the Alps forced by industrial black carbon. Proceedings of the National Academy of Sciences, 110, 15216–21.CrossRefGoogle ScholarPubMed
Palmer, M., Allan, J. D., Meyer, J., and Bernhardt, E. S. (2007). River restoration in the twenty‐first century: data and experiential knowledge to inform future efforts. Restoration Ecology, 15, 472–81.CrossRefGoogle Scholar
Palmer, M. A., Bernhardt, E. S., Allan, J. D., Lake, P. S., Alexander, G., Brooks, S., and Sudduth, E. (2005). Standards for ecologically successful river restoration. Journal of Applied Ecology, 42, 208–17.CrossRefGoogle Scholar
Palumbi, S. R., Barshis, D. J., Traylor-Knowles, N. and Bay, R. A. (2014). Mechanisms of reef coral resistance to future climate change. Science, 344, 895–8.CrossRefGoogle ScholarPubMed
Pandey, D. N. (2000). Sacred water and sanctified vegetation: tanks and trees in India. In Conference of the International Association for the Study of Common Property (IASCP), in the Panel “Constituting the Riparian Commons,” Bloomington, Indiana, USA, 31, 21 p.Google Scholar
Pando, L., Pulgar, J. A., and Gutiérrez-Claverol, M. (2013). A case of man-induced ground subsidence and building settlement related to karstified gypsum (Oviedo, NW Spain). Environmental Earth Sciences, 68, 507–19.CrossRefGoogle Scholar
Pandolfi, J. M., and 10 others. (2005). Are U.S.coral reefs on the slippery slope to slime? Science, 307, 1725–6.CrossRefGoogle ScholarPubMed
Pandolfi, J. M., Connolly, S. R., Marshall, D. J., and Cohen, A. L. (2011). Projecting coral reef futures under global warming and ocean acidification. Science, 333, 418–22.CrossRefGoogle ScholarPubMed
Pardini, G., Gispert, M., and Dunjó, G. (2004). Relative influence of wildfire on soil properties and erosion processes in different Mediterranean environments in NE Spain. Science of the Total Environment, 328, 237–46.CrossRefGoogle ScholarPubMed
Parfitt, S. A., and 15 others (2010). Early Pleistocene human occupance at the edge of the boreal zone in northwest Europe. Nature, 466, 229–33.CrossRefGoogle Scholar
Parizek, B. R. and Alley, R. B. (2004). Implications of increased Greenland surface melt under global-warming scenarios: ice-sheet simulations. Quaternary Science Reviews, 23, 1013–27.CrossRefGoogle Scholar
Park, R. A., Armentano, T. V., and Cloonan, C. L. (1986). Predicting the effects of sea level rise on coastal wetlands. In Effects of Changes in Stratospheric Ozone and Global Climate, Vol. 4, Sea Level Rise, ed. Titus, J. G.. Washington, DC: UNEP/USEPA, pp. 129–52.Google Scholar
Parkinson, R. W., Harlem, P. W., and Meeder, J. F. (2014). Managing the Anthropocene marine transgression to the year 2100 and beyond in the State of Florida, USA. Climatic Change, 128, 8598.CrossRefGoogle Scholar
Paroissien, J. B., Lagacherie, P., and Le Bissonnais, Y. (2010). A regional-scale study of multi-decennial erosion of vineyard fields using vine-stock unearthing–burying measurements. Catena, 82, 159–68.CrossRefGoogle Scholar
Parry, L. E., Holden, J., and Chapman, P. J. (2014). Restoration of blanket peatlands. Journal of Environmental Management, 133, 193205.CrossRefGoogle ScholarPubMed
Parshall, T. and Foster, D. R. (2002). Fire on the New England landscape: regional and temporal variation, cultural and environmental controls. Journal of Biogeography, 29, 1305–17.CrossRefGoogle Scholar
Parson, E. A., Carter, L., Anderson, P., Wang, B., and Weller, G. (2001). Potential consequences of climate variability and change for Alaska. In Climate Change Impacts on the United States: the Potential Consequences of Climate Variability and Change, ed. National Assessment Synthesis Team. Cambridge: Cambridge University Press, pp. 283312.Google Scholar
Parsons, A. J., Abrahams, A. D., and Wainwright, J. (1996). Responses of interrill runoff and erosion rates to vegetation change in southern Arizona. Geomorphology, 14, 311–7.CrossRefGoogle Scholar
Pasternack, G. B., Brush, G. S., and Hilgartner, W. B. (2001). Impact of historic land-use change on sediment delivery to a Chesapeake Bay subestuarine delta. Earth Surface Processes and Landforms, 26, 409–27.CrossRefGoogle Scholar
Pausas, J. G., Llovet, J., Rodrigo, A., and Vallejo, R. (2009). Are wildfires a disaster in the Mediterranean basin? – A review. International Journal of Wildland Fire, 17, 713–23.Google Scholar
Pearson, A. J., Snyder, N. P., and Collins, M. J. (2011). Rates and processes of channel response to dam removal with a sand-filled impoundment. Water Resources Research 47, W08504, doi:10.1029/2010WR009733.CrossRefGoogle Scholar
Pechony, O. and Shindell, D. T. (2010). Driving forces of global wildfires over the past millennium and the forthcoming century. Proceedings of the National Academy of Sciences, 107, 19167–70.CrossRefGoogle ScholarPubMed
Peck, J. A. and Kasper, N. R. (2013). Multiyear assessment of the sedimentological impacts of the removal of the Munroe Falls Dam on the middle Cuyahoga River, Ohio. Reviews in Engineering Geology, 21, 8192.Google Scholar
Pederson, G. T., Gray, S. T., Woodhouse, C. A., Betancourt, J. L., Fagre, D. B., Littell, J. S., and Graumlich, L. J. (2011). The unusual nature of recent snowpack declines in the North American Cordillera. Science, 333, 332–5.CrossRefGoogle ScholarPubMed
Pelejero, C., Calvo, E., and Hoegh-Guldberg, O. (2010). Paleo-perspectives on ocean acidification. Trends in Ecology and Evolution, 25, 332–44.CrossRefGoogle ScholarPubMed
Pellatt, M. G. and Gedalof, Z. E. (2014). Environmental change in Garry oak (Quercus garryana) ecosystems: the evolution of an eco-cultural landscape. Biodiversity and Conservation, 23, 2053–67, doi: 10.1007/s10531-014-0703-9.CrossRefGoogle Scholar
Pelletier, J. D., Brad Murray, A., Pierce, J. L., Bierman, P. R., Breshears, D. D., Crosby, B. T., and Yager, E. M. (2015). Forecasting the response of Earth’s surface to future climatic and land‐use changes: A review of methods and research needs. Earth’s Future, 3, 220–51, doi: 10.1002/2014EF000290.CrossRefGoogle Scholar
Pellicciotti, F., Carenzo, M., Bordoy, R., and Stoffel, M. (2014). Changes in glaciers in the Swiss Alps and impact on basin hydrology: current state of the art and future research. Science of the Total Environment, 493, 1152–70.CrossRefGoogle ScholarPubMed
Peng, H., Ma, W., Mu, Y. H., Jin, L., and Yuan, K. (2015). Degradation characteristics of permafrost under the effect of climate warming and engineering disturbance along the Qinghai–Tibet Highway. Natural Hazards, 75, 2589–605.CrossRefGoogle Scholar
Pennington, W. (1981). Records of a lake’s life in time: the sediments. Hydrobiologia, 79, 197219.CrossRefGoogle Scholar
Pereira, H. C. (1973). Land Use and Water Resources in Temperate and Tropical Climates. Cambridge: Cambridge University Press.Google Scholar
Perevolotsky, A. and Seligman, N. A. G. (1998). Role of grazing in Mediterranean rangeland ecosystems. Bioscience, 48, 1007–17.CrossRefGoogle Scholar
Perkol-Finkel, S., Shashar, N., and Benayahu, Y. (2006). Can artificial reefs mimic natural reef communities? The roles of structural features and age. Marine Environmental Research, 61, 121–35.CrossRefGoogle ScholarPubMed
Perroy, R. L., Bookhagen, B., Chadwick, O. A., and Howarth, J. T. (2012). Holocene and anthropocene landscape change: arroyo formation on Santa Cruz Island, California. Annals of the Association of American Geographers, 102, 1229–50.CrossRefGoogle Scholar
Perry, C. T., Murphy, G. N., Kench, P. S., Smithers, S. G., Edinger, E. N., Steneck, R. S., and Mumby, P. J. (2013). Caribbean-wide decline in carbonate production threatens coral reef growth. Nature Communications, 4, 1402.CrossRefGoogle ScholarPubMed
Perski, Z., Hanssen, R., Wojcik, A., and Wojciechowski, T. (2009). InSAR analyses of terrain deformation near the Wieliczka Salt Mine, Poland. Engineering Geology, 106, 5867.CrossRefGoogle Scholar
Peters, F. (1999). Bronze Age barrows: factors influencing their survival and destruction. Oxford Journal of Archaeology, 18, 255–64.CrossRefGoogle Scholar
Pethick, J. (1993). Shoreline adjustments and coastal management: physical and biological processes under accelerated sea level rise. Geographical Journal, 159, 162–8.CrossRefGoogle Scholar
Pethick, J. and Orford, J. D. (2013). Rapid rise in effective sea-level in southwest Bangladesh: its causes and contemporary rates. Global and Planetary Change, 111, 237–45.CrossRefGoogle Scholar
Petit, F., Poinsart, D., and Bravard, J. P. (1996). Channel incision, gravel mining and bedload transport in the Rhône river upstream of Lyon, France (“Canal de Miribel”). Catena, 26, 209–26.CrossRefGoogle Scholar
Petley, D. N., Hearn, G. J., Hart, A., Rosser, N., Dunning, S. A., Oven, K., and Mitchell, W. A. (2007). Trends in landslide occurrence in Nepal. Natural Hazards, 43, 2344.CrossRefGoogle Scholar
Péwé, T. L. (ed.) (1981). Desert dust: origin, characteristics and effect on man. Geological Society of America Special Paper 186.Google Scholar
Phien-wej, N., Giao, P. H., and Nutalaya, P. (2006). Land subsidence in Bangkok, Thailand. Engineering Geology, 82, 187201.CrossRefGoogle Scholar
Phillips, J. D. (1997). Humans as geological agents and the question of scale. American Journal of Science, 297, 98115.CrossRefGoogle Scholar
Pianalto, F. S. and Yool, S. R. (2013). Monitoring fugitive dust emission sources arising from construction: a remote-sensing approach. GIScience & Remote Sensing, 50, 251–70.CrossRefGoogle Scholar
Piccarreta, M., Caldara, M., Capolongo, D., and Boenzi, F. (2011). Holocene geomorphic activity related to climatic change and human impact in Basilicata, Southern Italy. Geomorphology, 128, 137–47.CrossRefGoogle Scholar
Pierce, J. L., Meyer, G. A., and Jull, A. T. (2004). Fire-induced erosion and millennial-scale climate change in northern ponderosa pine forests. Nature, 432, 8790.CrossRefGoogle ScholarPubMed
Pierson, F. B., Robichaud, P. R., and Spaeth, K. E. (2001). Spatial and temporal effects of wildfire on the hydrology of a steep rangeland watershed. Hydrological Processes, 15, 2905–16.CrossRefGoogle Scholar
Pierson, F. B., Moffet, C. A., Williams, C. J., Hardegree, S. P., and Clark, P. E. (2009). Prescribed‐fire effects on rill and interrill runoff and erosion in a mountainous sagebrush landscape. Earth Surface Processes and Landforms, 34, 193203.CrossRefGoogle Scholar
Pierson, F. B., Robichaud, P. R., Moffet, C. A., Spaeth, K. E., Hardegree, S. P., Clark, P. E., and Williams, C. J. (2008). Fire effects on rangeland hydrology and erosion in a steep sagebrush‐dominated landscape. Hydrological Processes, 22, 2916–29.CrossRefGoogle Scholar
Pierson, F. B., Williams, C. J., Hardegree, S. P., Weltz, M. A., Stone, J. J., and Clark, P. E. (2011). Fire, plant invasions, and erosion events on western rangelands. Rangeland Ecology & Management, 64, 439–49.CrossRefGoogle Scholar
Piggott, S. (1935). A note on the relative chronology of the English long barrows. Proceedings of the Prehistoric Society (New Series), 1, 115–26.Google Scholar
Pimentel, D. (1976). Land degradation: effects on food and energy resources. Science, 194, 149–55.CrossRefGoogle ScholarPubMed
Pimentel, D., Harvey, C., Resosuddarmo, P., Sinclair, K., Kurz, D., McNair, M., Crist, S., Shpritz, L., Fitton, L., Saffouri, R., and Blair, R. (1995). Environmental and economic costs of soil erosion and conservation benefits. Science, 267, 1117–22.CrossRefGoogle ScholarPubMed
Piperno, D. R., McMichael, C., and Bush, M. B. (2015). Amazonia and the Anthropocene: what was the spatial extent and intensity of human landscape modification in the Amazon Basin at the end of prehistory? The Holocene, 25, 1588–97.CrossRefGoogle Scholar
Pirazzoli, P. A. (1996). Sea-level Changes—The Last 20,000 Years. Chichester: Wiley.Google Scholar
Pizzuto, J. and O’Neal, M. (2009). Increased mid-twentieth century riverbank erosion rates related to the demise of mill dams, South River, Virginia. Geology, 37, 1922.CrossRefGoogle Scholar
Plug, L. J., Walls, C., and Scott, B. M. (2008). Tundra lake changes from 1978 to 2001 on the Tuktoyaktuk Peninsula, western Canadian Arctic. Geophysical Research Letters, 35, doi: 10.1029/2007GL032303.CrossRefGoogle Scholar
Poeppl, R. E., Keesstra, S. D., and Hein, T. (2015). The geomorphic legacy of small dams—An Austrian study. Anthropocene, 10, 4355.CrossRefGoogle Scholar
Poesen, J. W., Torri, D., and Bunte, K. (1994). Effects of rock fragments on soil erosion by water and different spatial scales: a review. Catena, 23, 141–66.CrossRefGoogle Scholar
Poesen, J. W. E., Nachtergaele, J., Verstraeten, G., and Valentin, C. (2003). Gully erosion and environmental change: importance and research needs. Catena, 50, 91133.CrossRefGoogle Scholar
Poirier, C., Chaumillon, E., and Arnaud, F. (2011). Siltation of river-influenced coastal environments: respective impact of late Holocene land use and high-frequency climate changes. Marine Geology, 290, 5162.CrossRefGoogle Scholar
Polidoro, B. A. and 20 others. (2010). The loss of species: mangrove extinction risk and geographic areas of global concern. PLoS One, 5, e10095.CrossRefGoogle ScholarPubMed
Pollard, E. and Miller, A. (1968). Wind erosion of the East Anglian Fens. Weather, 23, 414–7.CrossRefGoogle Scholar
Pollen-Bankhead, N., Simon, A., Jaeger, K., and Wohl, E. (2009). Destabilization of streambanks by removal of invasive species in Canyon de Chelly National Monument, Arizona. Geomorphology, 103, 363–74.CrossRefGoogle Scholar
Pomeroy, J., Fang, X., and Ellis, C. (2012). Sensitivity of snowmelt hydrology in Marmot Creek, Alberta, to forest cover disturbance. Hydrological Processes, 26, 1891–904.CrossRefGoogle Scholar
Pongratz, J., Reick, C., Raddatz, T., and Claussen, M. (2008). A reconstruction of global agricultural areas and land cover for the last millennium. Global Biogeochemical Cycles, 22(3), GB 3018, doi: 10.1029/2007GB003153.CrossRefGoogle Scholar
Pope, G. A. and Rubenstein, R. (1999). Anthroweathering: theoretical framework and case study for human‐impacted weathering. Geoarchaeology, 14, 247–64.3.0.CO;2-6>CrossRefGoogle Scholar
Potemkina, T. G. and Potemkin, V. L. (2014). Sediment load of the main rivers of Lake Baikal in a changing environment (east Siberia, Russia). Quaternary International, 380–381, 342–9.Google Scholar
Pranzini, E., Wetzel, L., and Williams, A. T. (2015). Aspects of coastal erosion and protection in Europe. Journal of Coastal Conservation, 19, 445–59, doi: 10.1007/s11852-015-0399-3CrossRefGoogle Scholar
Pratt, W. E. and Johnson, D. W. (1926). Local subsidence of the Goose Creek oil field. Journal of Geology, 34, 577–90.CrossRefGoogle Scholar
Price, K. (2011). Effects of watershed topography, soils, land use, and climate on baseflow hydrology in humid regions: A review. Progress in Physical Geography, 35, 465–92.Google Scholar
Price, S. J., Ford, J. R., Cooper, A. H., and Neal, C. (2011). Humans as major geological and geomorphological agents in the Anthropocene: the significance of artificial ground in Great Britain. Philosophical Transactions of the Royal Society, A, 369, 1056–84.Google ScholarPubMed
Prince, H. C. (1962). Pits and ponds in Norfolk. Erdkunde, 16, 1031.CrossRefGoogle Scholar
Prince, H.C. (1964). The origin of pits and depressions in Norfolk. Geography, 49, 1532.Google Scholar
Pringle, A. W. (1996). History, geomorphological problems and effects of dredging in Cleveland Bay, Queensland. Australian Geographical Studies, 34, 5880.CrossRefGoogle Scholar
Pritchard, H. D. and Vaughan, D. G. (2007). Widespread acceleration of tidewater glaciers on the Antarctic Peninsula. Journal of Geophysical Research, 112: F03S29, doi:10.1029/2006JF000597.CrossRefGoogle Scholar
Prosser, C. D., Burek, C. V., Evans, D. H., Gordon, J. E., Kirkbride, V. B., Rennie, A. F., and Walmsley, C. A. (2010). Conserving geodiversity sites in a changing climate: management challenges and responses. Geoheritage, 2, 123–36.CrossRefGoogle Scholar
Prosser, I. P. and Williams, L. (1998). The effect of wildfire on runoff and erosion in native Eucalyptus forest. Hydrological Processes, 12, 251–65.3.0.CO;2-4>CrossRefGoogle Scholar
Provoost, S., Jones, M. L. M. and Edmondson, . (2011). Changes in landscape and vegetation of coastal dunes in northwest Europe: a review. Journal of Coastal Conservation, 15, 207–26.CrossRefGoogle Scholar
Purvis, K. G., Gramling, J. M., and Murren, C. J. (2015). Assessment of beach cccess paths on dune vegetation: diversity, abundance, and cover. Journal of Coastal Research, 31, 1222–8.Google Scholar
Pye, K. and Blott, S. J. (2014). The geomorphology of UK estuaries: the role of geological controls, antecedent conditions and human activities. Estuarine, Coastal and Shelf Science, 150B, 196214.CrossRefGoogle Scholar
Pye, K., Blott, S. J., and Howe, M. A. (2014). Coastal dune stabilization in Wales and requirements for rejuvenation. Journal of Coastal Conservation, 18, 2754.CrossRefGoogle Scholar
Pye, K. and Neal, A. (1994). Coastal dune erosion at Formby Point, north Merseyside, England: causes and mechanisms. Marine Geology, 119, 3956.CrossRefGoogle Scholar
Pye, K. and Tsoar, H. (1990). Aeolian Sand and Sand Dunes. London: Unwin Hyman.CrossRefGoogle Scholar
Quataert, E., Storlazzi, C., Rooijen, A., Cheriton, O., and Dongeren, A. (2015). The influence of coral reefs and climate change on wave‐driven flooding of tropical coastlines. Geophysical Research Letters, 42, 6407–15. doi: 10.1002/2015GL064861.CrossRefGoogle Scholar
Qiu, J. (2012). Evidence mounts for dam-quake link. Science, 336, 291.CrossRefGoogle ScholarPubMed
Qiu, X. and Fenton, C. (2015). Factors controlling the occurrence of reservoir-induced seismicity. Engineering Geology for Society and Territory, 6, 567–70.CrossRefGoogle Scholar
Quinton, J. N., Govers, G., Van Oost, K., and Bardgett, R. D. (2010). The impact of agricultural soil erosion on biogeochemical cycling. Nature Geoscience, 3, 311–4.CrossRefGoogle Scholar
Quiquerez, A., Brenot, J., Garcia, J. P., and Petit, C. (2008). Soil degradation caused by a high-intensity rainfall event: implications for medium-term soil sustainability in Burgundian vineyards. Catena, 73, 8997.CrossRefGoogle Scholar
Quisthoudt, K., Adams, J., Rajkaran, A., Dahdouh-Guebas, F., Koedam, N., and Randin, C. F. (2013). Disentangling the effects of global climate and regional land-use change on the current and future distribution of mangroves in South Africa. Biodiversity and Conservation, 22, 1369–90.CrossRefGoogle Scholar
Raabe, E. A. and Stumpf, R. P. (2015). Expansion of tidal marsh in response to sea-level rise: Gulf Coast of Florida, USA. Estuaries and Coasts, doi: 10.1007/s12237-015-9974-y.CrossRefGoogle Scholar
Raclot, D., Le Bissonnais, Y., Louchart, X., Andrieux, P., Moussa, R., and Voltz, M. (2009). Soil tillage and scale effects on erosion from fields to catchment in a Mediterranean vineyard area. Agriculture, Ecosystems and Environment, 134, 201–10.CrossRefGoogle Scholar
Raczky, P. (2015). Settlements in South-east Europe. In The Oxford Handbook of Neolithic Europe, ed. Fowler, C., Harding, J., Hofmann, D.. Oxford: Oxford University Press. pp. 235–53.Google Scholar
Radic, V. and Hock, R. (2011). Regionally differentiated contribution of mountain glaciers and ice caps to future sea-level rise. Nature Geoscience, 4, 91–4.CrossRefGoogle Scholar
Radić, V., Bliss, A., Beedlow, A. C., Hock, R., Miles, E., and Cogley, J. G. (2014). Regional and global projections of twenty-first century glacier mass changes in response to climate scenarios from global climate models. Climate Dynamics, 42, 3758.CrossRefGoogle Scholar
Radivojević, M., Rehren, T., Pernicka, E., Šljivar, D., Brauns, M., and Borić, D. (2010). On the origins of extractive metallurgy: new evidence from Europe. Journal of Archaeological Science, 37, 2775–87.CrossRefGoogle Scholar
Radley, J. (1962). Peat erosion on the high moors of Derbyshire and west Yorkshire. East Midlands Geographer, 3, 4050.Google Scholar
Rădoane, M., Obreja, F., Cristea, I., and Mihailă, D. (2013). Changes in the channel-bed level of the eastern Carpathian rivers: climatic vs. human control over the last 50 years. Geomorphology, 193, 91111.CrossRefGoogle Scholar
Raharimahefa, T. and Kusky, T. M. (2010). Environmental monitoring of Bombetoka Bay and the Betsiboka Estuary, Madagascar, using multi-temporal satellite data. Journal of Earth Science, 21, 210–26.CrossRefGoogle Scholar
Rahm, D. (2011). Regulating hydraulic fracturing in shale gas plays: the case of Texas. Energy policy, 39, 2974–81.CrossRefGoogle Scholar
Rahman, M. (1981). Ecology of Karez irrigation: a case of Pakistan. GeoJournal, 5, 715.CrossRefGoogle Scholar
Rahmstorf, S. (2007). A semi-empirical approach to projecting future sea-level rise. Science, 315, 368–70.CrossRefGoogle ScholarPubMed
Räike, A., Kortelainen, P., Mattsson, T., and Thomas, D. N. (2012). 36 year trends in dissolved organic carbon export from Finnish rivers to the Baltic Sea. Science of the Total Environment, 435, 188201.CrossRefGoogle Scholar
Raji, B. A., Utovbisere, E. O. and Momodu, A. B. (2004). Impact of sand dune stabilization structures on soil and yield of millet in the semi-arid region of NW Nigeria. Environmental Monitoring and Assessment, 99, 181–96.CrossRefGoogle ScholarPubMed
Ramankutty, N., Heller, E., and Rhemtulla, J. (2010). Prevailing myths about agricultural abandonment and forest regrowth in the United States. Annals of the Association of American Geographers, 100, 502–12.CrossRefGoogle Scholar
Ramos, M. C. and Martínez-Casasnovas, J. A. (2007). Soil loss and soil water content affected by land levelling in Penedès vineyards, NE Spain. Catena, 71, 210–7.CrossRefGoogle Scholar
Ramos‐Scharrón, C. E. and MacDonald, L. H. (2005). Measurement and prediction of sediment production from unpaved roads, St John, US Virgin Islands. Earth Surface Processes and Landforms, 30, 1283–304.CrossRefGoogle Scholar
Ran, L., Lu, X. X., and Xin, Z. (2014). Erosion-induced massive organic carbon burial and carbon emission in the Yellow River basin, China. Biogeosciences, 11, 945–59.CrossRefGoogle Scholar
Ran, L., Lu, X. X., Xin, Z., and Yang, X. (2013). Cumulative sediment trapping by reservoirs in large river basins: a case study of the Yellow River basin. Global and Planetary Change, 100, 308–19.CrossRefGoogle Scholar
Ranwell, D. S. (1964). Spartina salt marshes in southern England, II: rate and seasonal pattern of sediment accretion. Journal of Ecology, 52, 7994.CrossRefGoogle Scholar
Ranwell, D. S. and Boar, R. (1986). Coast Dune Management Guide. Abbots Ripton: Institute of Terrestrial Ecology.Google Scholar
Rao, K. N., Subraelu, P., Naga Kumar, K., Demudu, G., Hema Malini, B., and Rajawat, A. S. (2010). Impacts of sediment retention by dams on delta shoreline recession: evidences from the Krishna and Godavari deltas, India. Earth Surface Processes and Landforms, 35, 817–27.CrossRefGoogle Scholar
Rapp, A., Murray-Rust, D. H., Christiansson, C., and Berry, L. (1972). Soil erosion and sedimentation in four catchments near Dodoma, Tanzania. Geografiska Annaler, 54A, 255318.CrossRefGoogle Scholar
Rasmussen, P. and Bradshaw, E. G. (2005). Mid-to late-Holocene land-use change and lake development at Dallund S0, Denmark: study aims, natural and cultural setting, chronology and soil erosion history. The Holocene, 15, 1105–15.Google Scholar
Ravanel, L. and Deline, P. (2011). Climate influence on rockfalls in high-Alpine steep rockwalls: the north side of the Aiguilles de Chamonix (Mont Blanc massif) since the end of the ‘Little Ice Age’. The Holocene, 21, 357–65.CrossRefGoogle Scholar
Ravanel, L. and Deline, P. (2015). Rockfall hazard in the Mont Blanc Massif increased by the current atmospheric warming. In Engineering Geology for Society and Territory-Volume 1, ed. Lollino, G., Manconi, A., Clague, J., Shan, W., and Chiarle, M.. Switzerland: Springer International Publishing, pp. 425–8.Google Scholar
Ravi, S., D’Odorico, P., Wang, J., White, C. S., Okin, G. S., Macko, S. A., and Collins, S. L. (2009). Post-fire resource redistribution in desert grasslands: a possible negative feedback on land degradation. Ecosystems, 12, 434–44.CrossRefGoogle Scholar
Ravi, S., D'Odorico, P., Breshears, D. D., Field, J. P., Goudie, A. S., Huxman, T. E., Li, J., Okin, G. S., Swap, R. J., Thomas, A. D., and Van Pelt, S. (2011). Aeolian processes and the biosphere. Review of Geophysics, 49, RG3001, doi: 10.1029/2010RG000328.CrossRefGoogle Scholar
Raymond, P. A. and Oh, N. H. (2009). Long term changes of chemical weathering products in rivers heavily impacted from acid mine drainage: Insights on the impact of coal mining on regional and global carbon and sulfur budgets. Earth and Planetary Science Letters, 284, 50–6.CrossRefGoogle Scholar
Reed, A. J., Mann, M. E., Emanuel, K. A., Lin, N., Horton, B. P., Kemp, A. C., and Donnelly, J. P. (2015). Increased threat of tropical cyclones and coastal flooding to New York City during the anthropogenic era. Proceedings of the National Academy of Sciences, 201513127.CrossRefGoogle Scholar
Reed, D. J. (1990).The impact of sea level rise on coastal salt marshes. Progress in Physical Geography, 14, 465–81.Google Scholar
Reed, D. J. (1995). The response of coastal marshes to sea level rise: survival or submergence? Earth Surface Processes and Landforms, 20, 3948.CrossRefGoogle Scholar
Reed, D. J. (2002). Sea-level rise and coastal marsh sustainability: geological and ecological factors in the Mississippi delta plain. Geomorphology, 48, 233–43.CrossRefGoogle Scholar
Rees, H. G. and Collins, D. N. (2006). Regional differences in response of flow in glacier-fed Himalayan rivers to climatic warming. Hydrological Processes, 20, 2157–69.CrossRefGoogle Scholar
Reid, D. and Church, M. (2015). Geomorphic and ecological consequences of riprap placement in river systems. Journal of the American Water Resources Association, 51, 1043–59.CrossRefGoogle Scholar
Reid, L. M. and Dunne, T. (1984). Sediment production from forest road surfaces. Water Resources Research, 20, 1753–61.CrossRefGoogle Scholar
Remini, B., Achour, B., and Albergel, J. (2011). Timimoun’s foggara (Algeria): an heritage in danger. Arabian Journal of Geosciences, 4, 495506.CrossRefGoogle Scholar
Remondo, J., Soto, J., González-Díez, A., Díaz de Terán, J. R., and Cendrero, A. (2005). Human impact on geomorphic processes and hazards in mountain areas in northern Spain. Geomorphology, 66, 6984.CrossRefGoogle Scholar
Rempel, L. L. and Church, M. (2009). Physical and ecological response to disturbance by gravel mining in a large alluvial river. Canadian Journal of Fisheries and Aquatic Sciences, 66, 5271.CrossRefGoogle Scholar
Renaud, F. G., Syvitski, J. P., Sebesvari, Z., Werners, S. E., Kremer, H., Kuenzer, C., and Friedrich, J. (2013). Tipping from the Holocene to the Anthropocene: how threatened are major world deltas? Current Opinion in Environmental Sustainability, 5, 644–54.CrossRefGoogle Scholar
Rendle, E. J. and Esteves, L. (2011). Developing protocols for assessing the performance of artificial surfing reefs – a new breed of coastal engineering. In Littoral 2010 – Adapting to Global Change at the Coast: Leadership, Innovation, and Investment (p. 12008). EDP Sciences.Google Scholar
Rendle, E. J. and Rodwell, L. D. (2014). Artificial surf reefs: A preliminary assessment of the potential to enhance a coastal economy. Marine Policy, 45, 349–58.CrossRefGoogle Scholar
Restrepo, A. and Juan, D. (2012). Assessing the effect of sea-level change and human activities on a major delta on the Pacific coast of northern South America: the Patía River. Geomorphology, 151, 207–23.Google Scholar
Reusser, L., Bierman, P., and Rood, D. (2015). Quantifying human impacts on rates of erosion and sediment transport at a landscape scale. Geology, 43, 171–4.CrossRefGoogle Scholar
Revell, D. L., Battalio, R., Spear, B., Ruggiero, P., and Vandever, J. (2011). A methodology for predicting future coastal hazards due to sea-level rise on the California Coast. Climatic Change, 109, (supplement), S251–76.CrossRefGoogle Scholar
Reynard, E. (2007). Geomorphosites and geodiversity: a new domain of research. Géomorphologie, 3, 181188.Google Scholar
Rhind, P. and Jones, R. (2009). A framework for the management of sand dune systems in Wales. Journal of Coastal Conservation, 13, 1523.CrossRefGoogle Scholar
Ricaurte, L. F., Boesch, S., Jokela, J., and Tockner, K. (2012). The distribution and environmental state of vegetated islands within human‐impacted European rivers. Freshwater Biology, 57, 2539–49.CrossRefGoogle Scholar
Ricci, M., Bertini, A., Capezzuoli, E., Horvatinčić, N., Andrews, J. E., Fauquette, S., and Fedi, M. (2014). Palynological investigation of a Late Quaternary calcareous tufa and travertine deposit: a case study of Bagnoli in the Valdelsa Basin (Tuscany, central Italy). Review of Palaeobotany and Palynology, 218, 184–97.Google Scholar
Richards, J. F. (1991). Land transformation. In The Earth as Transformed by Human Action, ed. Turner, B. L., Clark, W. C., Kates, R. W., Richards, J. F., Matthews, J. T., and Meyer, W. B.. Cambridge: Cambridge University Press, pp. 163–78.Google Scholar
Richardson, D. M., Holmes, P. M., Esler, K. J., Galatowitsch, S. M., Stromberg, J. C., Kirkman, S. P., and Hobbs, R. J. (2007). Riparian vegetation: degradation, alien plant invasions, and restoration prospects. Diversity and Distributions, 13, 126–39.CrossRefGoogle Scholar
Richardson, J. A. (1976). Pit heap into pasture. In Reclamation, ed. Lenihan, J. and Fletcher, W. W.. Glasgow: Blackie, pp. 6093.Google Scholar
Richardson, J. M., Fuller, I. C., Holt, K. A., Litchfield, N. J., and Macklin, M. G. (2014). Rapid post-settlement floodplain accumulation in Northland, New Zealand. Catena, 113, 292305.CrossRefGoogle Scholar
Richardson, S. J. and Smith, J. (1977). Peat wastage in the East Anglian Fens. Journal of Soil Science, 28, 485–9.CrossRefGoogle Scholar
Rick, T. C., Sillett, T. S., Ghalambor, C. K., Hofman, C. A., Ralls, K., Anderson, R. S., and Morrison, S. A. (2014). Ecological Change on California’s Channel Islands from the Pleistocene to the Anthropocene. BioScience, doi: 10.1093/biosci/biu094.CrossRefGoogle Scholar
Rickson, R. J. (2006). Controlling sediment at source: an evaluation of erosion control geotextiles. Earth Surface Processes and Landforms, 31, 550–60.CrossRefGoogle Scholar
Ridley, D. A., Heald, C. L., and Prospero, J. M. (2014). What controls the recent changes in African mineral dust aerosol across the Atlantic? Atmospheric Chemistry and Physics Discussions, 14, 3583–627.Google Scholar
Ries, J. B., Andres, K., Wirtz, S., Tumbrink, J., Wilms, T., Peter, K. D., Burczyk, M., Butzen, V., and Seeger, M. (2014). Sheep and goat erosion–experimental geomorphology a as an approach for the quantification of underestimated processes. Zeitschrift für Geomorphologie, Supplement, 58, 2345.CrossRefGoogle Scholar
Rignot, E. and Kanagaratnam, P. (2006). Changes in the velocity structure of the Greenland Ice Sheet. Science, 311, 986–90.CrossRefGoogle ScholarPubMed
Rignot, E. and Thomas, R. H. (2002). Mass balance of polar ice sheets. Science, 297, 1502–6.CrossRefGoogle ScholarPubMed
Rignot, E., Velicogna, I., van der Broejke, M. R. Monaghan, A., and Lenaerts, J. (2011). Acceleration of the contribution of the Greenland and Antarctica ice sheets to sea level rise. Geophysical Research Letters, 38, L05503, doi: 10.1029/2011GL046583.CrossRefGoogle Scholar
Riksen, M., Spaan, W., Arrué, J. L., and López, M. V. (2003). What to do about wind erosion. In Wind Erosion on Agricultural Land in Europe, ed. Warren, A.. Luxembourg: European Commission, pp. 3952.Google Scholar
Rinaldi, M., Wyżga, B., and Surian, N. (2005). Sediment mining in alluvial channels: physical effects and management perspectives. River Research and Applications, 21, 805–28.CrossRefGoogle Scholar
Riordan, B., Verbyla, D., and McGuire, A. D. (2006). Shrinking ponds in subarctic Alaska based on 1950–2002 remotely sensed images. Journal of Geophysical Research: Biogeosciences, 111(G4), doi: 10.1029/2005JG000150.CrossRefGoogle Scholar
Rippon, S. (2000). The Transformation of Coastal Wetlands: Exploitation and Management of Marshland Landscapes in North West Europe during the Roman and Medieval Periods. Oxford: Oxford University Press.Google Scholar
Rippon, S. (2006). The Somerset Wetlands: an Ever Changing Environment. Taunton: Somerset Archaeological & Natural History Society, pp. 4756.Google Scholar
Risk, M. J. (2014). Assessing the effects of sediments and nutrients on coral reefs. Current Opinion in Environmental Sustainability, 7, 108–17.CrossRefGoogle Scholar
Ritchie, W. and Gimingham, C. H. (1989). Restoration of coastal dunes breached by pipeline landfalls in north-east Scotland. Proceedings of the Royal Society of Edinburgh. Section B. Biological Sciences, 96, 231–45.CrossRefGoogle Scholar
Rivas, V., Rix, K., Frances, E., Cendrero, A., and Brunsden, D. (1997). Geomorphological indicators for environmental impact assessment: consumable and non-consumable geomorphological resources. Geomorphology, 18, 169–82.CrossRefGoogle Scholar
Rivas, V., Cendrero, A., Hurtado, M., Cabral, M., Giménez, J., Forte, L., and Becker, A. (2006). Geomorphic consequences of urban development and mining activities; an analysis of study areas in Spain and Argentina. Geomorphology, 73, 185206.CrossRefGoogle Scholar
Roberts, B. W., Thornton, C. P., and Pigott, V. C. (2009). Development of metallurgy in Eurasia. Antiquity, 83, 1012–22.CrossRefGoogle Scholar
Roberts, R. G. and 10 others. (2001). New ages for the least Australian megafauna: continent-wide extinction about 46,000 years ago. Science, 292, 1888–92.CrossRefGoogle Scholar
Robichaud, P. R. (2000). Fire effects on infiltration rates after prescribed fire in Northern Rocky Mountain forests, USA. Journal of Hydrology, 231, 220–9.Google Scholar
Robinson, D. N. (1968). Soil erosion by wind in Lincolnshire, March, 1968. East Midland Geogapher, 4, 351–62.Google Scholar
Robinson, M. (1990). Impact of Improved Land Drainage on River Flows. Institute of Hydrology, Wallingford, Report 113.Google Scholar
Robinson, M. A. and Lambrick, G. H. (1984). Holocene alluviation and hydrology in the Upper Thames Basin. Nature, 308, 809–14.CrossRefGoogle Scholar
Robroek, B. J., Smart, R. P., and Holden, J. (2010). Sensitivity of blanket peat vegetation and hydrochemistry to local disturbances. Science of the Total Environment, 408, 5028–34.CrossRefGoogle ScholarPubMed
Rockström, J. (2015). In The Observer, November 15, 2015, p. 35.Google Scholar
Rockström, J. and 28 others. (2009). Planetary boundaries: exploring the safe operating space for humanity. Ecology and Society, 14, http://www.ecologyandsociety.org/vol14/iss2/art32/.CrossRefGoogle Scholar
Rodgers, M., O’Connor, M., Robinson, M., Muller, M., Poole, R., and Xiao, L. (2011). Suspended solid yield from forest harvesting on upland blanket peat. Hydrological Processes, 25, 207–16.CrossRefGoogle Scholar
Rodriguez, A. B., Fodrie, F. J., Ridge, J. T., Lindquist, N. L., Theuerkauf, E. J., Coleman, S. E., and Kenworthy, M. D. (2014). Oyster reefs can outpace sea-level rise. Nature Climate Change, 4, 493–7.CrossRefGoogle Scholar
Rodway-Dyer, S. J. and Walling, D. E. (2010). The use of 137Cs to establish longer-term soil erosion rates on footpaths in the UK. Journal of Environmental Management, 91, 1952–62.CrossRefGoogle ScholarPubMed
Rogge, W. F., Medeiros, P. M., and Simoneit, B. R. (2006). Organic marker compounds for surface soil and fugitive dust from open lot dairies and cattle feedlots. Atmospheric Environment 40, 2749.CrossRefGoogle Scholar
Rojstaczer, S. and Deverel, S. J. (1995). Land subsidence in drained histosols and highly organic mineral soils of California. Soil Science Society of America Journal, 59, 1162–7.CrossRefGoogle Scholar
Romero-Diaz, A., Belmonte-Serrato, F., and Ruiz-Sinoga, J. D. (2010). The geomorphic impact of afforestations on soil erosion in southeast Spain. Land Degradation and Development, 21, 188–95.CrossRefGoogle Scholar
Romero Díaz, A., Marin Sanleandro, P., Sanchez Soriano, A., Belmonte Serrato, F., and Faulkner, H. (2007). The causes of piping in a set of abandoned agricultural terraces in southeast Spain. Catena, 69, 282–93.CrossRefGoogle Scholar
Romps, D. M., Seeley, J. T., Vollaro, D., and Molinari, J. (2014). Projected increase in lightning strikes in the United States due to global warming. Science, 346, 851–4.CrossRefGoogle ScholarPubMed
Roosevelt, A. C. (2014). The Amazon and the Anthropocene: 13,000 years of human influence in a tropical rainforest. Anthropocene, 4, 6987.CrossRefGoogle Scholar
Roosevelt, C. H. (2006). Tumulus survey and museum research in Lydia, western Turkey: Determining Lydian-and Persian-period settlement patterns. Journal of Field Archaeology, 31, 6176.CrossRefGoogle Scholar
Rose, N. L. (2015). Spheroidal carbonaceous fly-ash particles provide a globally synchronous stratigraphic marker for the Anthropocene. Environmental Science & Technology, 49, 4155–62.CrossRefGoogle Scholar
Rose, N. L., Morley, D., Appleby, P. G., Battarbee, R. W., Alliksaar, T., Guilizzoni, P., and Punning, J. M. (2011). Sediment accumulation rates in European lakes since AD 1850: trends, reference conditions and exceedence. Journal of Paleolimnology, 45, 447–68.CrossRefGoogle ScholarPubMed
Rosen, A. M., Lee, J., Li, M., Wright, J., Wright, H. T., and Fang, H. (2015). The Anthropocene and the landscape of Confucius: a historical ecology of landscape changes in northern and eastern China during the middle to late-Holocene. The Holocene, 25, 1640–50.CrossRefGoogle Scholar
Rosenzweig, C. and Hillel, D. (1993). The dust bowl of the 1930s: analogy of greenhouse effect in the Great Plains? Journal of Environmental Quality, 22, 922.CrossRefGoogle Scholar
Rosepiler, M. J. and Reilinger, R. (1977). Land subsidence due to water withdrawal in the vicinity of Pecos, Texas. Engineering Geology, 11, 295304.CrossRefGoogle Scholar
Roskin, J., Katra, I., and Blumberg, D. G. (2013). Late Holocene dune mobilizations in the northwestern Negev dunefield, Israel: a response to combined anthropogenic activity and short-term intensified windiness. Quaternary International, 303, 1023.CrossRefGoogle Scholar
Rott, H., Müller, F., Nagler, T., and Floricioiu, D. (2010). The imbalance of glaciers after disintegration of Larsen B ice shelf, Antarctic Peninsula. The Cryosphere Discussions, 4, 1607–33.Google Scholar
Rovira, A., Batalla, R. J., and Sala, M. (2005). Response of a river sediment budget after historical gravel mining (The lower Tordera, NE Spain). River Research and Applications, 21, 829–47.CrossRefGoogle Scholar
Rowland, J. C., Jones, C. E., Altmann, G., Bryan, R., Crosby, B. T., Hinzman, L. D., and Geernaert, G. L. (2010). Arctic landscapes in transition: responses to thawing permafrost. Eos, Transactions American Geophysical Union, 91(26), 229–30.CrossRefGoogle Scholar
Rowntree, K. (1991). An assessment of the potential impact of alien invasive vegetation on the geomorphology of river channels in South Africa. Southern African Journal of Aquatic Science, 17, 2843.CrossRefGoogle Scholar
Rowntree, K., Duma, M., Kakembo, V., and Thornes, J. (2004). Debunking the myth of overgrazing and soil erosion. Land Degradation & Development, 15, 203–14.CrossRefGoogle Scholar
Royal Society (2005). Ocean Acidification Due to Increased Atmospheric Carbon Dioxide. London: Royal Society.Google Scholar
Rubinstein, J. L. and Mahani, A. B. (2015). Myths and facts on wastewater injection, hydraulic fracturing, enhanced oil recovery, and induced seismicity. Seismological Research Letters, 86, 1060–7.CrossRefGoogle Scholar
Ruddiman, W., Vavrus, S., Kutzbach, J., and He, F. (2014). Does pre-industrial warming double the anthropogenic total? The Anthropocene Review, 1, 147–53.CrossRefGoogle Scholar
Ruddiman, W. F. (2003). The anthropogenic greenhouse era began thousands of years ago. Climatic Change, 61, 261–93.CrossRefGoogle Scholar
Ruddiman, W. F. (2013). The Anthropocene. Annual Review of Earth and Planetary Science 41, 4568.Google Scholar
Ruddiman, W. F. (2014). Earth Transformed. New York: W.H. Freeman and Company.CrossRefGoogle Scholar
Ruddiman, W. F., Ellis, E. C., Kaplan, J. O., and Fuller, D. Q. (2015). Defining the epoch we live in. Science, 348, 38–9.CrossRefGoogle ScholarPubMed
Ruecker, G., Schad, P., Alcubilla, M. M., and Ferrer, C. (1998). Natural regeneration of degraded soils and site changes on abandoned agricultural terraces in Mediterranean Spain. Land Degradation & Development, 9, 179–88.3.0.CO;2-R>CrossRefGoogle Scholar
Ruffing, C. M., Daniels, M. D., and Dwire, K. A. (2015). Disturbance legacies of historic tie-drives persistently alter geomorphology and large wood characteristics in headwater streams, southeast Wyoming. Geomorphology, 231, 114.CrossRefGoogle Scholar
Rugel, K., Jackson, C. R., Romeis, J. J., Golladay, S. W., Hicks, D. W., and Dowd, J. F. (2012). Effects of irrigation withdrawals on streamflows in a karst environment: lower Flint River Basin, Georgia, USA. Hydrological Processes, 26, 523–34.CrossRefGoogle Scholar
Ruiz-Colmenero, M., Bienes, R., and Marques, M. J. (2011). Soil and water conservation dilemmas associated with the use of green cover in steep vineyards. Soil and Tillage Research, 117, 211–23.CrossRefGoogle Scholar
Ruiz-Colmenero, M., Bienes, R., Eldridge, D. J., and Marques, M. J. (2013). Vegetation cover reduces erosion and enhances soil organic carbon in a vineyard in the central Spain. Catena, 104, 153–60.CrossRefGoogle Scholar
Rull, V. (2013). A futurist perspective on the Anthropocene. The Holocene, 23, 1198–201.CrossRefGoogle Scholar
Rumschlag, J. H. and Peck, J. A. (2007). Short-term sediment and morphologic response of the Middle Cuyahoga River to the removal of the Munroe Falls Dam, Summit County, Ohio. Journal of Great Lakes Research, 33, 142–53.CrossRefGoogle Scholar
Rutherford, I. (2000). Some human impacts on Australian stream channel morphology. In River Management: the Australian Experience, ed. Brizga, S. and Linlayson, B. L.. Chichester: Wiley, pp. 1147.Google Scholar
Rutqvist, J., Rinaldi, A. P., Cappa, F., and Moridis, G. J. (2013). Modeling of fault reactivation and induced seismicity during hydraulic fracturing of shale-gas reservoirs. Journal of Petroleum Science and Engineering, 107, 3144.CrossRefGoogle Scholar
Rutz, K. (2012). Artificial Islands versus natural reefs: the environmental cost of development in Dubai. International Journal of Islamic Architecture, 1, 243–67.CrossRefGoogle Scholar
Ruz, M.-H., Anthony, E. A. and Faucon, L. (2005). Coastal dune evolution on a shoreline subject to strong human pressure: the Dunkirk area, northern France. Dunes and Estuaries, 19, 441–9.Google Scholar
Ryken, N., Vanmaercke, M., Wanyama, J., Isabirye, M., Vanonckelen, S., Deckers, J., and Poesen, J. (2015). Impact of papyrus wetland encroachment on spatial and temporal variabilities of stream flow and sediment export from wet tropical catchments. Science of the Total Environment, 511, 756–66.CrossRefGoogle ScholarPubMed
Saez, J. L., Corona, C., Stoffel, M., and Berger, F. (2013). Climate change increases frequency of shallow spring landslides in the French Alps. Geology, 41, 619–22.CrossRefGoogle Scholar
Saiko, T. A. and Zonn, I. S. (2000). Irrigation expansion and dynamics of desertification in the Circum-Aral region of Central Asia. Applied Geography, 20, 349–67.CrossRefGoogle Scholar
Saintilan, N., Wilson, N. C., Rogers, K., Rajkaran, A. and Krauss, K. W. (2014). Mangrove expansion and salt marsh decline at mangrove poleward limits. Global Change Biology, 20, 147–57.CrossRefGoogle ScholarPubMed
Sakals, M. E., Innes, J. L., Wilford, D. J., Sidle, R. C., and Grant, G. E. (2006). The role of forests in reducing hydrogeomorphic hazards. Forest Snow Landscape Research, 80, 1122.Google Scholar
Sale, P. F. (2013). The futures of coral reefs. In The Balance of Nature and Human Impact, ed. Rhode, K.. Cambridge: Cambridge University Press, pp. 325–34.Google Scholar
Sallenger, A. H. Jr., Doran, K. S., and Howd, P. A. (2012). Hotspot of accelerated sea-level rise on the Atlantic coast of North America. Nature Climate Change, 2, 884–8.CrossRefGoogle Scholar
Salman, A. B., Howari, F. M., El-Sankary, M. M., Wali, A. M., and Saleh, M. M. (2010). Environmental impact and natural hazards on Kharga Oasis monumental sites, Western Desert of Egypt. Journal of African Earth Sciences, 58, 341–53.CrossRefGoogle Scholar
Salmoral, G., Willaarts, B. A., Troch, P. A., and Garrido, A. (2015). Drivers influencing streamflow changes in the Upper Turia basin, Spain. Science of the Total Environment, 503, 258268.CrossRefGoogle ScholarPubMed
Sandor, J. A., Gersper, P. L., and Hawley, J. W. (1990). Prehistoric agricultural terraces and soils in the Mimbres area, New Mexico. World Archaeology, 22, 7086.CrossRefGoogle Scholar
Sannel, A. B. K. and Kuhry, P. (2011). Warming‐induced destabilization of peat plateau/thermokarst lake complexes. Journal of Geophysical Research: Biogeosciences 116(G3), doi: 10.1029/2010JG001635.CrossRefGoogle Scholar
Santer, B. D., Wigley, T. M. L., Gleckler, P. J., Bonfils, C., Wehner, M. F., AchutaRao, K., Barnett, T. P., Boyle, J. S., Brüggemann, W., Fiorino, M., and Gillett, N. (2006). Forced and unforced ocean temperature changes in Atlantic and Pacific tropical cyclogenesis regions. Proceedings of the National Academy of Sciences, 103, 13905–10.CrossRefGoogle ScholarPubMed
Sarkar, A. (2013), Tractor Production and Sales in India, 1989–2009. Review of Agrarian Studies, 3, 1, available at www.ras.org.in/tractor_production_and_sales_in_india_1989_2009 (Accessed 7th June, 2015).Google Scholar
Sattler, K., Keiler, M., Zischg, A., and Schrott, L. (2011). On the connection between debris flow activity and permafrost degradation: a case study from the Schnalstal, South Tyrolean Alps, Italy. Permafrost and Periglacial Processes, 22, 254–65.CrossRefGoogle Scholar
Sauer, C. O. (1938). Destructive exploitation in modern colonial expansion. Proceedings International Geographical Congress, Amsterdam, Vol. III, section IIIC, 494–9.Google Scholar
Saunders, M. A. and Lee, A. S. (2008). Large contribution of sea surface warming to recent increase in Atlantic hurricane activity. Nature, 451, 557–60.CrossRefGoogle ScholarPubMed
Saussure, H. B. de (1796). Voyages dans les Alpes. Paris: Fauche.Google Scholar
Saye, S. E. and Pye, K. (2007). Implications of sea level rise for coastal dune habitat conservation in Wales, UK. Journal of Coastal Conservation 11, 3152.CrossRefGoogle Scholar
Sazhin, A. N. (1988). Regional aspects of dust storms in steppe regions of the east European and West Siberian plains. Soviet Geography, 29, 935–46.CrossRefGoogle Scholar
Schaefer, K., Lantuit, H., Romanovsky, V. E., Schuur, E. A., and Witt, R. (2014). The impact of the permafrost carbon feedback on global climate. Environmental Research Letters, 9, 085003.CrossRefGoogle Scholar
Schelker, J., Kuglerova, L., Eklöf, K., Bishop, K., and Laudon, H. (2013). Hydrological effects of clear-cutting in a boreal forest–Snowpack dynamics, snowmelt and streamflow responses. Journal of Hydrology, 484, 105–14.CrossRefGoogle Scholar
Scherler, D., Bookhagen, B., and Strecker, M.R. (2011). Spatially variable response of Himalayan glaciers to climate change affected by debris cover. Nature Geoscience, 4, 156–9.CrossRefGoogle Scholar
Schiavon, N., Chiavari, G., and Fabbri, D. (2004). Soiling of limestone in an urban environment characterized by heavy vehicular exhaust emissions. Environmental Geology, 46, 448–55.CrossRefGoogle Scholar
Schiefer, E., Petticrew, E. L., Immell, R., Hassan, M. A., and Sonderegger, D. L. (2013). Land use and climate change impacts on lake sedimentation rates in western Canada. Anthropocene, 3, 6171.CrossRefGoogle Scholar
Schilling, K.E., Chan, K.-S., Liu, H., and Zhang, Y.-K. (2010). Quantifying the effect of land use land cover change on increasing discharge in the Upper Mississippi River. Journal of Hydrology, 387, 343–5.CrossRefGoogle Scholar
Schmitt, A., Dotterweich, M., Schmidtchen, G., and Bork, H. R. (2003). Vineyards, hopgardens and recent afforestation: effects of late Holocene land use change on soil erosion in northern Bavaria, Germany. Catena, 51, 241–54.CrossRefGoogle Scholar
Schmittbuhl, J., Lengliné, O., Cornet, F., Cuenot, N., and Genter, A. (2014). Induced seismicity in EGS reservoir: the creep route. Geothermal Energy, 2, 113.CrossRefGoogle Scholar
Schoellhamer, D. H., Wright, S. A., and Drexler, J. Z. (2013). Adjustment of the San Francisco estuary and watershed to decreasing sediment supply in the 20th century. Marine Geology, 345, 6371.CrossRefGoogle Scholar
Schothorst, C. J. (1977). Subsidence of low moor peat soils in the western Netherlands. Geoderma, 17, 265–91.CrossRefGoogle Scholar
Schottler, S. P., Ulrich, J., Belmont, P., Moore, R., Lauer, J., Engstrom, D. R., and Almendinger, J. E. (2014). Twentieth century agricultural drainage creates more erosive rivers. Hydrological Processes, 28, 1951–61.CrossRefGoogle Scholar
Schumm, S. A., Harvey, M. D., and Watson, C. C. (1984). Incised Channels: Morphology, Dynamics and Control. Littleton, Colorado: Water Resources Publications.Google Scholar
Schuur, E. A. G., McGuire, A. D., Schädel, C., Grosse, G., Harden, J. W., Hayes, D. J., and Vonk, J. E. (2015). Climate change and the permafrost carbon feedback. Nature, 520, 171–9.CrossRefGoogle ScholarPubMed
Schuster, R. L. (1979). Reservoir-induced landslides. Bulletin of the International Association of Engineering Geology, 20, 815.CrossRefGoogle Scholar
Schwarz, H. E., Emel, J., Dickens, W. J., Rogers, P., and Thompson, J. (1991).Water quality and flows. In The Earth as Transformed by Human Action, ed. Turner, B. L., Clark, W. C., Kates, R. W., Richards, J. F., Matthews, J. T., and Meyer, W. B.. Cambridge: Cambridge University Press, pp. 253–70.Google Scholar
Scyphers, S. B., Powers, S. P., Heck, K. L. Jr, and Byron, D. (2011). Oyster reefs as natural breakwaters mitigate shoreline loss and facilitate fisheries. PloS one, 6(8), e22396.CrossRefGoogle ScholarPubMed
Sear, D., Newson, M., Hill, C., Old, J., and Branson, J. (2009). A method for applying fluvial geomorphology in support of catchment‐scale river restoration planning. Aquatic Conservation: Marine and Freshwater Ecosystems, 19, 506–19.CrossRefGoogle Scholar
Sefelnasr, A. and Sherif, M. (2014). Impacts of seawater rise on seawater intrusion in the Nile delta aquifer, Egypt. Groundwater, 52, 264–76.CrossRefGoogle ScholarPubMed
Segura-Beltrán, F. and Sanchis-Ibor, C. (2013). Assessment of channel changes in a Mediterranean ephemeral stream since the early twentieth century. The Rambla de Cervera, eastern Spain. Geomorphology, 201, 199214.CrossRefGoogle Scholar
Seifan, N. (2009). Long-term effects of anthropogenic activities on semi-arid sand dunes. Journal of Arid Environments, 73, 332–7.CrossRefGoogle Scholar
Senter, J. (2003). Live dunes and ghost forests: Stability and change in the history of North Carolina’s maritime forests. North Carolina Historical Review, 80, 334–71.Google Scholar
Seyoum, W. M., Milewski, A. M., and Durham, M. C. (2015). Understanding the relative impacts of natural processes and human activities on the hydrology of the Central Rift Valley lakes, East Africa. Hydrological Processes, 29, 4312–24, doi: 10.1002/hyp.10490CrossRefGoogle Scholar
Shakesby, R. A. (2011). Post-wildfire soil erosion in the Mediterranean: review and future research directions. Earth-Science Reviews, 105, 71100.CrossRefGoogle Scholar
Shakesby, R. A., Doerr, S. H., and Walsh, R. P. D. (2000). The erosional impact of soil hydrophobicity: current problems and future research directions. Journal of Hydrology, 231/2, 178–91.Google Scholar
Shakesby, R. A., Wallbrink, P. J., Doerr, S. H., English, P. M., Chafre, C. J., Humphreys, G. S., Blake, W. H. and Tomkins, K. M. (2007). Distinctiveness of wildfire effects on soil erosion in south-east Australian eucalyptus forests assessed in a global context. Forest Ecology and Management, 238, 347–64.CrossRefGoogle Scholar
Shaler, N. S. (1912). Man and the Earth. New York: Duffield.Google Scholar
Shams, A. (2014). A rediscovered-new “Qanat”system in the high mountains of Sinai Peninsula, with Levantine reflections. Journal of Arid Environments, 110, 6974.CrossRefGoogle Scholar
Shan, W., Hu, Z., Guo, Y., Zhang, C., Wang, C., Jiang, H.,and Xiao, J. (2015). The impact of climate change on landslides in southeastern of high-latitude permafrost regions of China. Frontiers in Earth Science, 3, doi: 10.3389/feart.2015.00007.CrossRefGoogle Scholar
Shankman, D. and Smith, L. J. (2004). Stream channelization and swamp formation in the U.S. coastal plain. Physical Geography, 25, 2238.CrossRefGoogle Scholar
Shao, Y., Wyrwoll, K. H., Chappell, A., Huang, J., Lin, Z., McTainsh, G. H., and Yoon, S. (2011). Dust cycle: an emerging core theme in Earth system science. Aeolian Research, 2, 181204.CrossRefGoogle Scholar
Sheffield, A. T., Healy, T. R., and McGlone, M. S. (1995). Infilling rates of a steepland catchment estuary, Whangamata, New Zealand. Journal of Coastal Research, 11, 1294–308.Google Scholar
Sheng, J. and Wilson, J. P. (2009). Watershed urbanization and changing flood behaviour across the Los Angeles metropolitan region. Natural Hazards, 48, 4157.CrossRefGoogle Scholar
Shepard, C. C., Agostini, V. N., Gilmer, B., Allem, T., Stone, J., Brooks, W., and Beck, M. W. (2012). Assessing future risk: quantifying the effects of sea level rise on storm surge risk for the southern shores of Long Island, New York. Natural Hazards, 60, 727–45.CrossRefGoogle Scholar
Sheppard, C. R. C. (2003). Predicted recurrences of mass coral morality in the Indian Ocean. Nature, 425, 294–7.CrossRefGoogle Scholar
Sheppard, C. R. C., Davy, S. K., and Pilling, G. M. (2009). The Biology of Coral Reefs. Oxford: Oxford University Press.CrossRefGoogle Scholar
Sheridan, G. J. and Noske, P. J. (2007). A quantitative study of sediment delivery and stream pollution from different forest road types. Hydrological Processes, 21, 387–98.CrossRefGoogle Scholar
Sheridan, G. J., Noske, P. J., Whipp, R. K., and Wijesinghe, N. (2006). The effect of truck traffic and road water content on sediment delivery from unpaved forest roads. Hydrological Processes, 20, 1683–99.CrossRefGoogle Scholar
Sherlock, R. L. (1922). Man as a Geological Agent. London: Witherby.Google Scholar
Sherratt, A. (1983). The secondary exploitation of animals in the Old World. World Archaeology, 15, 90104.CrossRefGoogle Scholar
Shi, W., Wang, M., and Guo, W. (2014). Long‐term hydrological changes of the Aral Sea observed by satellites. Journal of Geophysical Research: Oceans, 119, 3313–26.Google Scholar
Shiklomanov, I. A. (1985). Large scale water transfers. In Facets of hydrology II, ed. Rodda, J. C.. Chichester: Wiley, pp. 345–87.Google Scholar
Shlemon, R. J. (1995). Groundwater rise and hydrocollapse: technical and political implications of Special Geologic Report Zones’ in Riverside County, California, USA. International Association of Hydrological Sciences, Publication, 234, 481–6.Google Scholar
Shomurodov, H. F., Rakhimova, T. T., Saribaeva, S. U., Rakhimova, N. K., Esov, R. A., and Adilov, B. A. (2013). Perspective plant species for stabilization of sand dunes on the exposed Aral Sea bed. Journal of Earth Science and Engineering, 3, 655–62.Google Scholar
Shriner, D. S. and Street, R. B. (1998). North America. In The Regional Impacts of Climate Change, ed. Watson, R. T., Zinyowera, M.C. and Moss, R.H.. Cambridge: Cambridge University Press, pp. 273–8.Google Scholar
Shuttleworth, E. L., Evans, M. G., Hutchinson, S. M., and Rothwell, J. J. (2015). Peatland restoration: controls on sediment production and reductions in carbon and pollutant export. Earth Surface Processes and Landforms, 40, 459–72.CrossRefGoogle Scholar
Siakeu, J., Oguchi, T., Aoki, T., Esaki, Y., and Jarvie, H. P. (2004). Change in riverine suspended sediment concentration in central Japan in response to late 20th century human activities. Catena, 55, 231–54.CrossRefGoogle Scholar
Sidle, R. C. and Dhakal, A. S. (2002). Potential effect of environmental change on landslide hazards in forest environments. In Environmental Change and Geomorphic Hazards in Forests, ed. Sidle, R. C.. Wallingford: CABI, pp. 123–65.CrossRefGoogle Scholar
Sidle, R. C. and Burt, T. P.. (2009). Temperate forests and rangelands. In Geomorphology and Global Environmental Change, ed. Slaymaker, O., Spencer, T., and Embleton-Hamman, C.. Cambridge: Cambridge University Press, pp. 321–43.Google Scholar
Sidle, R. C., Furuichi, T., and Kono, Y. (2011). Unprecedented rates of landslide and surface erosion along a newly constructed road in Yunnan, China. Natural Hazards, 57, 313–26.CrossRefGoogle Scholar
Sidle, R. C., Ghestem, M., and Stokes, A. (2014). Epic landslide erosion from mountain roads in Yunnan, China–challenges for sustainable development. Natural Hazards and Earth System Science, 14, 3093–104.CrossRefGoogle Scholar
Sidle, R. C., Sasaki, S., Otsuki, M., Noguchi, S., and Rahim Nik, A. (2004). Sediment pathways in a tropical forest: effects of logging roads and skid trails. Hydrological Processes, 18, 703–20.CrossRefGoogle Scholar
Sidle, R. C., Kamil, I., Sharma, A., and Yamashita, S. (2000). Stream response to subsidence from underground coal mining in central Utah. Environmental Geology, 39, 279–91.CrossRefGoogle Scholar
Siebert, S., Burke, J., Faures, J. M., Frenken, K., Hoogeven, J., Döll, P., and Portman, F. T. (2010). Groundwater use for irrigation – a global inventory. Hydrology and Earth System Sciences Discussions, 7, 39774021.Google Scholar
Silva, R., Martínez, M. L., Hesp, P. A., Catalan, P., Osorio, A. F., Martell, R., and Govaere, G. (2014). Present and future challenges of coastal erosion in Latin America. Journal of Coastal Research, 71(sp1), 116.CrossRefGoogle Scholar
Silveira, L. and Alonso, J. (2009). Runoff modifications due to the conversion of natural grasslands to forests in a large basin in Uruguay. Hydrological Processes, 23, 320–9.CrossRefGoogle Scholar
Silverberg, R. (2013). Mound Builders. Athens, Ohio: Ohio University Press.Google Scholar
Simas, T., Nunes, J. P., and Ferreira, J. G. (2001). Effects of global climate change on coastal salt marshes. Ecological Modelling, 139, 115.CrossRefGoogle Scholar
Simpson, J. M., Darrow, M. M., Huang, S. L., Daanen, R. P., and Hubbard, T. D. (2015). Investigating movement and characteristics of a frozen debris lobe, South-central Brooks Range, Alaska. Environmental & Engineering Geoscience, 1078–7275, doi: 10.2113/EEG-1728.CrossRefGoogle Scholar
Sinclair, W. C. (1982). Sinkhole Development Resulting from Ground-Water Withdrawal in the Tampa Area, Florida. US Geological Survey, Water Resources Division.Google Scholar
Siriwardena, L., Finlayson, B. L. and McMahon, T. A. (2006). The impact of land use change on catchment hydrology in large catchments: the Comet River, Central Queensland, Australia. Journal of Hydrology, 326, 199214.CrossRefGoogle Scholar
Six, D., Reynaud, L., and Letreguilly, A. (2001). Bilans de masse des glaciers alpines et scandinaves, leurs relations avec l’oscillation due climat de l’Atlantique nord. Comptes Rendus Academie des Sciences, Sciences de la Terre et des Planètes, 333, 693–8.Google Scholar
Skeat, A. J., East, T. J., and Corbett, L. K. (2012). Impact of feral water buffalo. In Landscape and Vegetation Ecology of the Kakadu Region, Northern Australia, ed. Finlayson, C. M. and von Oetzen, I.. Dordrecht: Springer, pp. 155–77.Google Scholar
Slater, L. J., Singer, M. B. and Kirchner, J. W. (2015). Hydrologic versus geomorphic drivers of trends in flood hazard. Geophysical Research Letters, 42, 370–6.CrossRefGoogle Scholar
Slaymaker, O., Spencer, T., and Embleton-Hamann, C. (eds.) (2009). Geomorphology and Global Environmental Change. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Sloss, C. R., Jones, B. G., Brooke, B. P., Heijnis, H. and Murray-Wallace, C. V. (2011). Contrasting sedimentation rates in Lake Illawarra and St George’s Basin, two large barrier estuaries on the southeast coast of Australia. Journal of Paleolimnology, 46, 561–77.CrossRefGoogle Scholar
Smil, V. (2011). Harvesting the biosphere: the human impact. Population and Development Review, 37, 613–36.CrossRefGoogle ScholarPubMed
Smil, V. (2015). It’s too soon to call this the Anthropocene. Spectrum, IEEE, 52, 28.Google Scholar
Smith, B. D. and Zeder, M. A. (2013). The onset of the Anthropocene. Anthropocene, 4, 813.CrossRefGoogle Scholar
Smith, B. J., McCabe, S., McAllister, D., Adamson, C., Viles, H. A., and Curran, J. M. (2011). A commentary on climate change, stone decay dynamics and the ‘greening’of natural stone buildings: new perspectives on ‘deep wetting’. Environmental Earth Sciences, 63, 1691–700.CrossRefGoogle Scholar
Smith, D. M., Zalasiewicz, J. A., Williams, M., Wilkinson, I. P., Redding, M., and Begg, C. (2010). Holocene drainage systems of the English Fenland: roddons and their environmental significance. Proceedings of the Geologists’ Association, 121, 256–69.CrossRefGoogle Scholar
Smith, H. G. and Dragovich, D. (2008). Post-fire hillslope erosion response in a sub-alpine environment, south-eastern Australia. Catena, 73, 274–85.CrossRefGoogle Scholar
Smith, H. G., Sheridan, G. J., Lane, P. N., and Bren, L. J. (2011). Wildfire and salvage harvesting effects on runoff generation and sediment exports from radiata pine and eucalypt forest catchments, south-eastern Australia. Forest Ecology and Management, 261, 570–81.CrossRefGoogle Scholar
Smith, L. M., Haukos, D. A., McMurry, S. T., LaGrange, T., and Willis, D. (2011). Ecosystem services provided by playas in the High Plains: potential influences of USDA conservation programs. Ecological Applications, 21(sp1), S82–92.CrossRefGoogle Scholar
Sobota, I. and Nowak, M. (2014). Changes in the dynamics and thermal regime of the permafrost and active layer of the high Arctic coastal area in North‐West Spitsbergen, Svalbard. Geografiska Annaler: Series A, Physical Geography, 96, 227–40.Google Scholar
Sofia, G., Prosdocimi, M., Dalla Fontana, G., and Tarolli, P. (2014). Modification of artificial drainage networks during the past half-century: Evidence and effects in a reclamation area in the Veneto floodplain (Italy). Anthropocene, 6, 4682.CrossRefGoogle Scholar
Sokolov, N. A. (1884). Dunes, Their Formation, Development and Internal Structure. St.Petersburg: St. Petersburg University (in Russian).Google Scholar
Sokolov, N. A. (1894) Die Dünen, Bildung, Entwicklung und Ihrer Bau, Berlin: Springer.Google Scholar
Solé-Benet, A., Lázaro, R., Domingo, F., Cantón, Y., and Puigdefábregas, J. (2010). Why most agricultural terraces in steep slopes in semiarid SE Spain remain well preserved since their abandonment 50 years go? Pirineos, 165, 215–35.CrossRefGoogle Scholar
Sorg, A., Kääb, A., Roesch, A., Bigler, C., and Stoffel, M. (2015). Contrasting responses of Central Asian rock glaciers to global warming. Scientific Reports, 5, doi:10.1038/srep08228.CrossRefGoogle ScholarPubMed
Soulsby, C., Birkel, C., and Tetzlaff, D. (2014). Assessing urbanization impacts on catchment transit times. Geophysical Research Letters, 41, 442–8.CrossRefGoogle Scholar
Sousa, A., García-Barrón, L., García-Murillo, P., Vetter, M., and Morales, J. (2015). The use of changes in small coastal Atlantic brooks in southwestern Europe as indicators of anthropogenic and climatic impacts over the last 400 years. Journal of Paleolimnology, 53, 7388.CrossRefGoogle Scholar
Souter, D. W. and Linden, O. (2000). The health and future of coral reef systems. Ocean and Coastal Management, 43, 657–88.CrossRefGoogle Scholar
Spalding, M. D. and Brown, B. E. (2015). Warm-water coral reefs and climate change. Science, 350, 769–71.CrossRefGoogle ScholarPubMed
Spate, O. H. K. and Learmonth, A. T. A. (1967). India and Pakistan. London: Methuen.Google Scholar
Sperna Weiland, F. C., van Beek, L. P. H., Kwadijk, K. C. J., and Blerkens, M. F. P. (2011). Global patterns of change in discharge regimes for 2100. Hydrology and Earth System Sciences Discussions, 8, 10973–1014.Google Scholar
Spigel, K. M. and Robichaud, P. R. (2007). First‐year post‐fire erosion rates in Bitterroot National Forest, Montana. Hydrological Processes, 21, 9981005.CrossRefGoogle Scholar
Stabile, T. A., Giocoli, A., Perrone, A., Piscitelli, S., and Lapenna, V. (2014). Fluid injection induced seismicity reveals a NE dipping fault in the southeastern sector of the High Agri Valley (southern Italy). Geophysical Research Letters, 41, 5847–54.CrossRefGoogle Scholar
Stabile, T. A., Giocoli, A., Lapenna, V., Perrone, A., Piscitelli, S., and Telesca, L. (2014). Evidence of low‐magnitude continued reservoir‐induced seismicity associated with the Pertusillo Artificial Lake (Southern Italy). Bulletin of the Seismological Society of America, 104, 1820–8, doi: 10.1785/0120130333.CrossRefGoogle Scholar
Stallins, J. A. (2006). Geomorphology and ecology: unifying themes for complex systems in biogeomorphology. Geomorphology, 77, 207–16.CrossRefGoogle Scholar
Stancheva, M., Ratas, U., Orviku, K., Palazov, A., Rivis, R., Kont, A., and Stanchev, H. (2011). Sand dune destruction due to increased human impacts along the Bulgarian Black Sea and Estonian Baltic Sea coasts. Journal of Coastal Research, S1, 64, 324–8.Google Scholar
Stanley, D. J. (1996). Nile delta: extreme case of sediment entrapment on a delta plain and consequent coastal land loss. Marine Geology, 129, 189–95.CrossRefGoogle Scholar
Steadman, D. W., Stafford, T. W., Donahue, D. J., and Jull, A. J. T. (1991). Chronology of Holocene vertebrate extinction in the Galápagos Islands. Quaternary Research, 36, 126–33.CrossRefGoogle Scholar
Steffen, W. (2010). Observed trends in Earth System behaviour. Interdisciplinary Reviews, Climate Change, 1, 428–49.Google Scholar
Steffen, W. and 10 others. (2004). Global change and the Earth System. Berlin: SpringerGoogle Scholar
Steffen, W., Crutzen, P. J., and McNeill, J. R. (2007). The Anthropocene: are humans now overwhelming the great forces of nature? Ambio, 36, 614–21.CrossRefGoogle ScholarPubMed
Steffen, W., Grinevald, J., Crutzen, P., and McNeill, J. (2011). The Anthropocene: conceptual and historical perspectives. Philosophical Transactions of the Royal Society, 369A, 842–67.Google Scholar
Steffen, W., Broadgate, W., Deutsch, L., Gaffney, O., and Ludwig, C. (2015). The trajectory of the Anthropocene: the Great Acceleration. The Anthropocene Review, 2, 8198.CrossRefGoogle Scholar
Stephens, J. C. (1956). Subsidence of organic soils in the Florida Everglades. Soil Science Society of America Journal, 20, 7780.CrossRefGoogle Scholar
Stephens, J. C., Allen, L. H., and Chen, E. (1984). Organic soil subsidence. Reviews in Engineering Geology, 6, 107–22.CrossRefGoogle Scholar
Sterk, G. (2003). Causes, consequences and control of wind erosion in Sahelian Africa: a review. Land Degradation and Development, 14, 95108.CrossRefGoogle Scholar
Sterling, S. M., Ducharne, A., and Polcher, J. (2013). The impact of global land-cover change on the terrestrial water cycle. Nature Climate Change, 3, 385–90.CrossRefGoogle Scholar
Stevenson, A. C., Jones, V. J., and Battarbee, R. W. (1990).The cause of peat erosion: a palaeolimnological approach. New Phytologist, 114, 727–35.CrossRefGoogle Scholar
Stewart, I. T. (2009). Changes in snowpack and snowmelt runoff for key mountain regions. Hydrological Processes, 23, 7894.CrossRefGoogle Scholar
Stinchcomb, G. E., Stewart, R.M., Messner, T.C., Nordt, L. C., Driese, S. G., and Allen, P. M., (2013). Using event stratigraphy to map the Anthropocene – an example from the historic coal mining region in eastern Pennsylvania, USA. Anthropocene, 2, 4250.CrossRefGoogle Scholar
Stive, M. J., de Schipper, M. A., Luijendijk, A. P., Aarninkhof, S. G., van Gelder-Maas, C., van Thiel de Vries, J. S., and Ranasinghe, R. (2013). A new alternative to saving our beaches from sea-level rise: The sand engine. Journal of Coastal Research, 29, 1001–8.Google Scholar
Stoddart, D. R. (1971). Coral reefs and islands and catastrophic storms. In Applied Coastal Geomorphology, ed. Steers, J. A.. London: Macmillan, pp. 154–97.Google Scholar
Stoffel, M. and Beniston, M. (2006). On the incidence of debris flows from the early Little Ice Age to a future greenhouse climate: a case study from the Swiss Alps. Geophysical Research Letters, 33, L16404, doi: 10.1029/2006GL026805.CrossRefGoogle Scholar
Stoffel, M. and Huggel, C. (2012). Effects of climate change on mass movements in mountain environments. Progress in Physical Geography, 36, 421–39.Google Scholar
Stoffel, M., Tiranti, D., and Huggel, C. (2014). Climate change impacts on mass movements—case studies from the European Alps. Science of the Total Environment, 493, 1255–66.CrossRefGoogle ScholarPubMed
Stokes, A., Atger, C., Bengough, A. G., Fourcaud, T., and Sidle, R. C. (2009). Desirable plant root traits for protecting natural and engineered slopes against landslides. Plant and Soil, 324, 130.CrossRefGoogle Scholar
Stokes, C. R., Popovnin, V., Aleynikov, A., Gurney, S. D., and Shahgedanova, M. (2007). Recent glacier retreat in the Caucasus Mountains, Russia, and associated increase in supraglacial debris cover and supra-/proglacial lake development. Annals of Glaciology, 46, 195203.CrossRefGoogle Scholar
Stokes, S., Goudie, A. S, Colls, A., and Al-Farraj, A. (2003). Optical dating as a tool for studying dune reactivation, accretion rates and desertification over decadal, centennial and millennial time-scales. In Desertification in the Third Millennium, ed. Alsharhan, A. S., Wood, W. W., Goudie, A. S., Fowler, A., and Abdellatif, E. M.. Balkema: Lisse, pp. 5360.CrossRefGoogle Scholar
Stovern, M., Betterton, E. A., Sáez, A. E., Villar, O. I. F., Rine, K. P., Russell, M. R., and King, M. (2014). Modeling the emission, transport and deposition of contaminated dust from a mine tailing site. Reviews on Environmental Health, 29, 91–4.CrossRefGoogle ScholarPubMed
Straneo, F. and Heimbach, P. (2013). North Atlantic warming and the retreat of Greenland’s outlet glaciers. Nature, 504, 3643.CrossRefGoogle ScholarPubMed
Stringer, C. (2003). Out of Africa. Nature, 423, 692–9.Google Scholar
Stromberg, J. C., Lite, S. J., Marler, R., Paradzick, C., Shafroth, P. B., Shorrock, D., and White, M. S. (2007). Altered stream‐flow regimes and invasive plant species: the Tamarix case. Global Ecology and Biogeography, 16, 381–93.CrossRefGoogle Scholar
Strong, C. L., Bullard, J. E., Dubois, C., McTainsh, G. H., and Baddock, M. C. (2010). Impact of wildfire on interdune ecology and sediments: an example from the Simpson Desert, Australia. Journal of Arid Environments, 74, 1577–81.CrossRefGoogle Scholar
Strong, D. R. and Ayres, D. A. (2009). Spartina introduction and consequences in salt marshes. In Human Impacts on Salt Marshes: a Global Perspective, ed. Silliman, B. R., Grosholz, E. D., and Bertness, M. D.. Vancouver: University of British Columbia Press, pp. 322.CrossRefGoogle Scholar
Struyf, E., Smis, A., Van Damme, S., Garnier, J., Govers, G., Van Wesemael, B., and Meire, P. (2010). Historical land use change has lowered terrestrial silica mobilization. Nature Communications, 1, 129.CrossRefGoogle ScholarPubMed
Su, Z. A., Zhang, J. H., Qin, F. C., and Nie, X. J. (2012). Landform change due to soil redistribution by intense tillage based on high-resolution DEMs. Geomorphology, 175, 190–8.Google Scholar
Suchodoletz, H. von, Oberhänsli, H., Faust, D., Fuchs, M., Blanchet, C., Goldhammer, T., and Zöller, L. (2010). The evolution of Saharan dust input on Lanzarote (Canary Islands) – influenced by human activity during the early Holocene? The Holocene, 20, 169–79.Google Scholar
Suckale, J. (2009). Induced seismicity in hydrocarbon fields. Advances in Geophysics, 51, 55106.CrossRefGoogle Scholar
Summa-Nelson, M. C. and Rittenour, T. M. (2012). Application of OSL dating to middle to late Holocene arroyo sediments in Kanab Creek, southern Utah, USA. Quaternary Geochronology, 10, 167–74.CrossRefGoogle Scholar
Sun, G. E., McNulty, S. G., Moore, J., Bunch, C., and Ni, J. (2002). Potential impacts of climate change on rainfall erosivity and water availability in China in the next 100 years. Proceedings of the 12th International Soil Conservation Conference, Beijing, 244–50.Google Scholar
Surell, A. (1841). Étude sur les torrents des Hautes-Alpes par Alexandre Surell. Paris: Carilian-Goeury et Dalmont.Google Scholar
Surian, N. (1999). Channel changes due to river regulation: the case of the Piave River, Italy. Earth Surface Processes and Landforms, 24, 1135–51.3.0.CO;2-F>CrossRefGoogle Scholar
Surian, N. and Rinaldi, M. (2003). Morphological response to river engineering and management in alluvial channels in Italy. Geomorphology, 50, 307–26.CrossRefGoogle Scholar
Sušnik, J., Vamvakeridou-Lyroudia, L. S., Baumert, N., Kloos, J., Renaud, F. G., La Jeunesse, I., and Zografos, C. (2015). Interdisciplinary assessment of sea-level rise and climate change impacts on the lower Nile Delta, Egypt. Science of the Total Environment, 503, 279–88.Google ScholarPubMed
Sweatman, H., Delean, S., and Syms, C. (2011). Assessing loss of coral cover on Australia’s Great Barrier Reef over two decades, with implications for longer term trends. Coral Reefs, 30, 521–31.CrossRefGoogle Scholar
Sylla, M. B., Gaye, A. T., Jenkins, G. S., Pal, J. S., and Giorgi, F. (2010). Consistency of projected drought over the Sahel with changes in the monsoon circulation and extremes in a regional climate model projections. Journal of Geophysical Research, 115, D16108, doi: 10.1029/2009JD012983.CrossRefGoogle Scholar
Syvitski, J. P. M. and Kettner, A. (2011). Sediment flux and the Anthropocene. Philosophical Transactions of the Royal Society, A, 369, 957–75.Google ScholarPubMed
Syvitski, J. P. M. and Milliman, J. D. (2007). Geology, geograpohy and humans battle for dominance over the delivery of fluvial sediment to the coastal ocean. Journal of Geology, 115, 119.CrossRefGoogle Scholar
Syvitski, J. P. M., and Saito, Y. (2007). Morphodynamics of deltas under the influence of humans. Global and Planetary Change, 57, 261–82.CrossRefGoogle Scholar
Syvitski, J. P. M., Vörösmarty, C. J., Kettner, A. J., and Green, P. (2005). Impact of humans on the flux of terrestrial sediment to the global coastal ocean. Science, 308, 376–80.CrossRefGoogle Scholar
Syvitski, J. P. M. and 10 others. (2009). Sinking deltas due to human activities. Nature Geosciences, 2, 681–6.CrossRefGoogle Scholar
Syvitski, J. P., Kettner, A. J., Overeem, I., Giosan, L., Brakenridge, G. R., Hannon, M., and Bilham, R. (2013). Anthropocene metamorphosis of the Indus Delta and lower floodplain. Anthropocene, 3, 2435.CrossRefGoogle Scholar
Szabó, J., Dávid, L., and Lóczy, D. (eds.). (2010) Anthropogenic Geomorphology: a Guide to Man-made Landforms. Heidelberg: Springer.CrossRefGoogle Scholar
Ta, W., Dong, Z., and Sanzhi, C. (2006). Effect of the 1950s large-scale migration for land reclamation on spring dust storms in Northwest China. Atmospheric Environment, 40, 5815–23.CrossRefGoogle Scholar
Taborda, R. and Ribeiro, M. A. (2015). A simple model to estimate the impact of sea-level rise on platform beaches. Geomorphology, 234, 204–10.CrossRefGoogle Scholar
Taheri, K., Gutiérrez, F., Mohseni, H., Raeisi, E., and Taheri, M. (2015). Sinkhole susceptibility mapping using the analytical hierarchy process (AHP) and magnitude–frequency relationships: a case study in Hamadan province, Iran. Geomorphology, 234, 6479.CrossRefGoogle Scholar
Talbot, T. and Lapointe, M. (2002). Modes of response of a gravel bed river to meander straightening: the case of the Sainte‐Marguerite River, Saguenay Region, Quebec, Canada.Water Resources Research, 38, 91.CrossRefGoogle Scholar
Tan, M. and Li, X. (2015). Does the Green Great Wall effectively decrease dust storm intensity in China? A study based on NOAA NDVI and weather station data. Land Use Policy, 43, 42–7.CrossRefGoogle Scholar
Tang, Q. and Lettenmaier, D. P. (2012). 21st century runoff sensitivities of major global river basins. Geophysical Research Letters, 39, L06403, doi: 10.1029/2011GL050834.CrossRefGoogle Scholar
Tang, Y., Zhong, S., Luo, L., Bian, X., Heilman, W. E., and Winkler, J. (2015). The potential impact of regional climate change on fire weather in the United States. Annals of the Association of American Geographers, 105, 121.CrossRefGoogle Scholar
Tang, Z., Gu, Y., Drahota, J., LaGrange, T., Bishop, A., and Kuzila, M. S. (2015). Using fly ash as a marker to quantify culturally‐accelerated sediment accumulation in playa wetlands. JAWRA Journal of the American Water Resources Association, 51, 1643–55, doi: 10.1111/1752-1688.12347.CrossRefGoogle Scholar
Taniguchi, K. T. and Biggs, T. W. (2015). Regional impacts of urbanization on stream channel geometry: a case study in semiarid southern California. Geomorphology, 248, 228–36.CrossRefGoogle Scholar
Tarolli, P., Preti, F., and Romano, N. (2014). Terraced landscapes: from an old best practice to a potential hazard for soil degradation due to land abandonment. Anthropocene, 6, 1025.CrossRefGoogle Scholar
Tarolli, P., Sofia, G., Calligaro, S., Prosdocimi, M., Preti, F., and Dalla Fontana, G. (2015). Vineyards in terraced landscapes: new opportunities from LiDAR data. Land Degradation & Development, 26, 92102.CrossRefGoogle Scholar
Tarriño, A., Elorrieta, I., García-Rojas, M., Orue, I., and Sánchez, A. (2014). Neolithic Flint Mines of Treviño (Basque-Cantabrian Basin, Western Pyrenees, Spain). Journal of Lithic Studies, 1, 129–47.CrossRefGoogle Scholar
Taylor, M. P. and Little, J. A. (2013). Environmental impact of a major copper mine spill on a river and floodplain system. Anthropocene, 3, 3650.CrossRefGoogle Scholar
Teatini, P., Ferronato, M., Gambolati, G., and Gonella, M. (2006). Groundwater pumping and land subsidence in the Emilia‐Romagna coastland, Italy: modeling the past occurrence and the future trend. Water Resources Research, 42(1), W01406, doi: 10.1029/2005WR004242.CrossRefGoogle Scholar
Tegen, I. and Fung, I., (1995) Contribution to the atmospheric mineral aerosol load from land surface modification. Journal of Geophysical Research, 100(D9), 18707–26.Google Scholar
Tegen, I., Werner, M., Harrison, S. P., and Kohfeld, K. E. (2004) Relative importance of climate and land use in determining present and future global soil dust emissions. Geophysical Science Reviews, article L05105.CrossRefGoogle Scholar
Temmerman, S., Meire, P., Bouma, T. J., Herman, P. M., Ysebaert, T., and De Vriend, H. J. (2013). Ecosystem-based coastal defence in the face of global change. Nature, 504, 7983.CrossRefGoogle ScholarPubMed
Tempany, H. A., Roddan, G. M., and Lord, L. (1944). Soil erosion and soil conservation in the colonial empire. Empire Forestry Journal, 23, 142–59.Google Scholar
Teneva, L., Karnauskas, M., Logan, C. A., Bianucci, L., Currie, J. C., and Kleypas, J. A. (2012). Predicting coral bleaching hotspots: the role of regional variability in thermal stress and potential adaptation rates. Coral Reefs, 31, 112.CrossRefGoogle Scholar
Tennant, C. and Menounos, B. (2013). Glacier change of the Columbia Icefield, Canadian Rocky Mountains, 1919–2009. Journal of Glaciology, 59, 671–86.CrossRefGoogle Scholar
Teo, E. A. and Marren, P. M. (2015). Interaction of ENSO-driven flood variability and anthropogenic changes in driving channel evolution: Corryong/Nariel Creek, Australia. Australian Geographer, 46, 339–62.CrossRefGoogle Scholar
Ter-Stepanian, G. (1988). Beginning of the Technogene. Bulletin of the International Association of Engineering Geology, 38, 133–42.CrossRefGoogle Scholar
Ternan, J. L., Williams, A. G., Elmes, A., and Fitzjohn, C. (1996). The effectiveness of bench-terracing and afforestation for erosion control on Rana sediments in central Spain. Land Degradation and Development, 7, 337–51.3.0.CO;2-G>CrossRefGoogle Scholar
Thomas, D. S. G. and Twyman, C. (2004). Good or bad rangeland? Hybrid knowledge, science, and local understandings of vegetation dynamics in the Kalahari. Land Degradation & Development, 15, 215–31.CrossRefGoogle Scholar
Thomas, D. S. G., Knight, M., and Wiggs, G. F. S. (2005). Remobilization of southern African desert dune systems by twenty-first century global warming. Nature, 435, 1218–21.CrossRefGoogle ScholarPubMed
Thomas, J. (2006). On the origins and development of cursus monuments in Britain. Proceedings of the Prehistoric Society, 72, 229–41.CrossRefGoogle Scholar
Thomas, W. F. (ed.) (1956). Man’s Role in Changing the Face of the Earth. Chicago: University of Chicago Press.Google Scholar
Thompson, J. R. (1970). Soil erosion in the Detroit metropolitan area. Journal of Soil and Water Conservation, 25, 810.CrossRefGoogle Scholar
Thompson, L. M. C. and Shlacher, T. A. (2008). Physical damage to coastal dunes and ecological impacts caused by vehicle tracks associated with beach camping on sandy shores: a case study from Fraser Island, Australia. Journal of Coastal Conservation, 12, 6782.Google Scholar
Thompson, L. G. (2000). Ice core evidence for climate change in the Tropics: implications for our future. Quaternary Science Reviews, 19, 1935.CrossRefGoogle Scholar
Thorndycraft, V. R., Pirrie, D., and Brown, A. G. (2004). Alluvial records of medieval and prehistoric tin mining on Dartmoor, southwest England. Geoarchaeology, 19, 219–36.CrossRefGoogle Scholar
Thornes, J. B. (2007). Modelling soil erosion by grazing: recent developments and new approaches. Geographical Research, 45, 1326.CrossRefGoogle Scholar
Tickner, D. P., Angold, P. G., Gurnell, A. M., and Mountford, J. O. (2001). Riparian plant invasions: hydrogeomorphological control and ecological impacts. Progress in Physical Geography, 25, 2252.CrossRefGoogle Scholar
Tielidze, L. G., Lomidze, N., and Asanidze, L. (2015). Glaciers retreat and climate change effect during the last one century in the Mestiachala River Basin, Caucasus Mountains, Georgia. Earth Sciences, 4, 7279.CrossRefGoogle Scholar
Tiffen, M., Mortimore, M., and Gichuki, F. (1994 ). More People, Less Erosion. Environmental Recovery in Kenya. Chichester: Wiley.Google Scholar
Tihansky, A. B. (1999). Sinkholes, west-central Florida. Land Subsidence in the United States. United States Geological Survey Circular, 1182, 121–40.Google Scholar
Tolksdorf, J. F. and Kaiser, K. (2012). Holocene aeolian dynamics in the European sand-belt as indicated by geochronological data. Boreas, 41, 408–21.CrossRefGoogle Scholar
Tolksdorf, J. F., Klasen, N., and Hilgers, A. (2013). The existence of open areas during the Mesolithic: evidence from aeolian sediments in the Elbe–Jeetzel area, northern Germany. Journal of Archaeological Science, 40, 2813–23.CrossRefGoogle Scholar
Tomás, R., Herrera, G., Delgado, J., Lopez-Sanchez, J. M., Mallorquí, J. J., and Mulas, J. (2010). A ground subsidence study based on DInSAR data: calibration of soil parameters and subsidence prediction in Murcia City (Spain). Engineering Geology, 111, 1930.CrossRefGoogle Scholar
Tomás, R., Márquez, Y., Lopez-Sanchez, J. M., Delgado, J., Blanco, P., Mallorquí, J. J., and Mulas, J. (2005). Mapping ground subsidence induced by aquifer overexploitation using advanced Differential SAR Interferometry: Vega Media of the Segura River (SE Spain) case study. Remote Sensing of Environment, 98, 269–83.CrossRefGoogle Scholar
Tonjes, D. J. (2013). Impacts from ditching salt marshes in the mid-Atlantic and northeastern United States. Environmental Reviews, 21, 116–26.CrossRefGoogle Scholar
Torri, D., Santi, E., Marignani, M., Rossi, M., Borselli, L., and Maccherini, S. (2013). The recurring cycles of “biancana” badlands: Erosion, vegetation and human impact. Catena, 106, 2230.CrossRefGoogle Scholar
Tóth, C. (2004). Functional changes of the tumuli at the different stages of history. Anthropogenic Aspects of Landscape Transformations, 3, 93102.Google Scholar
Tourian, M. J., Elmi, O., Chen, Q., Devaraju, B., Roohi, S., and Sneeuw, N. (2015). A spaceborne multisensor approach to monitor the desiccation of Lake Urmia in Iran. Remote Sensing of Environment, 156, 349–60.CrossRefGoogle Scholar
Touysinhthiphonexay, K. C. and Gardner, T. W. (1984). Threshold response of small streams to surface coal mining, bituminous coal fields, Central Pennsylvania. Earth Surface Processes and Landforms, 9, 4358.CrossRefGoogle Scholar
Townsend‐Small, A., Pataki, D. E., Liu, H., Li, Z., Wu, Q., and Thomas, B. (2013). Increasing summer river discharge in southern California, USA, linked to urbanization. Geophysical Research Letters, 40, 4643–7.CrossRefGoogle Scholar
Tréguer, P. J. and De La Rocha, C. L. (2013). The world ocean silica cycle. Annual Review of Marine Science, 5, 477501.CrossRefGoogle ScholarPubMed
Trenhaile, A. S. (2014). Climate change and its impact on rock coasts. Geological Society, London, Memoirs, 40, 717.CrossRefGoogle Scholar
Trimble, S. W. (1974). Man-induced Soil Erosion on the Southern Piedmont. Ankeny, IA: Soil Conservation Society of America.Google Scholar
Trimble, S.W. (1988). The impact of organisms on overall erosion rates within catchments in temperate regions. In Biogeomorphology, ed. Viles, H. A.. Oxford: Basil Blackwell, pp. 83142.Google Scholar
Trimble, S.W. (1997). Stream channel erosion and change resulting from riparian forests. Geology, 25, 467–9.2.3.CO;2>CrossRefGoogle Scholar
Trimble, S.W. (2003). Historical hydrographic and hydrologic changes in the San Diego Creek watershed, Newport Bay, California. Journal of Historical Geography, 29, 422–44.CrossRefGoogle Scholar
Trimble, S.W. (2004). Effects of riparian vegetation on stream channel stability and sediment budgets. Water Science and Application, 8, 153–69.CrossRefGoogle Scholar
Trimble, S.W. (2008a). Man-induced soil erosion on the southern Piedmont (2nd edn). Ankeny, IA: Soil Conservation Society of America.Google Scholar
Trimble, S.W. (2008b). The use of historical data and artifacts in geomorphology. Progress in Physical Geography, 32, 329.CrossRefGoogle Scholar
Trimble, S.W. (2013). Historical Agriculture and Soil Erosion in the Upper Mississippi Valley Hill Country. Boca Raton: CRC Press.Google Scholar
Trimble, S. W. and Cooke, R. U. (1991). Historical sources for geomorphological research in the United States. Professional Geographer, 43, 212–28.CrossRefGoogle Scholar
Trimble, S. W. and Crosson, S. (2000). US soil erosion rates – myth and reality. Science, 289, 248–50.CrossRefGoogle Scholar
Trimble, S. W. and Mendel, A. C. (1995). The cow as a geomorphic agent: a critical review. Geomorphology, 13, 233–53.CrossRefGoogle Scholar
Triplett, L. D., Engstrom, D. R., Conley, D. J., and Schellhaass, S. M. (2008). Silica fluxes and trapping in two contrasting natural impoundments of the upper Mississippi River. Biogeochemistry, 87, 217–30.CrossRefGoogle Scholar
Trnka, M., Kersebaum, K. C., Eitzinger, J., Hayes, M., Hlavinka, P., Svoboda, M., and Žalud, Z. (2013). Consequences of climate change for the soil climate in Central Europe and the central plains of the United States. Climatic Change, 120, 405–18.CrossRefGoogle Scholar
Tropeano, D. (1984). Rate of soil erosion processes on vineyards in central Piedmont (NW Italy). Earth Surface Processes and Landforms, 9, 253–66.CrossRefGoogle Scholar
Trout, T. and Neibling, W. (1993). Erosion and sedimentation processes on irrigated fields. Journal of Irrigation and Drainage Engineering, 119, 947–63.CrossRefGoogle Scholar
Trzhtsinsky, Y. B. (2002). Human-induced activation of gypsum karst in the southern Priangaria (East Siberia, Russia). Carbonates and Evaporites, 17, 154–58.CrossRefGoogle Scholar
Tsai, J. S., Venne, L. S., McMurry, S. T., and Smith, L. M. (2007). Influences of land use and wetland characteristics on water loss rates and hydroperiods of playas in the Southern High Plains, USA. Wetlands, 27, 683–92.CrossRefGoogle Scholar
Tsoar, H. and Blumberg, D. G. (2002). Formation of parabolic dunes from barchans and transverse dunes along Israel’s Mediterranean coast. Earth Surface Processes and Landforms, 27, 1147–61.CrossRefGoogle Scholar
Tsuboki, K., Yoshioka, M., Shinoda, T., Kato, M., Kanada, S., and Kitoh, A. (2014). Future increase of super‐typhoon intensity associated with climate change. Geophysical Research Letters, 42, 646–52.Google Scholar
Turetsky, M. R., Benscoter, B., Page, S., Rein, G., van der Werf, G. R., and Watts, A. (2015). Global vulnerability of peatlands to fire and carbon loss. Nature Geoscience, 8, 11–4.CrossRefGoogle Scholar
Turnbull, L., Wainwright, J., and Brazier, R. E. (2010). Changes in hydrology and erosion over a transition from grassland to shrubland. Hydrological Processes, 24, 393414.CrossRefGoogle Scholar
Turner, B. L., Kasperson, R. E., Meyer, W. B., Dow, K. M., Golding, D., Kasperson, J. X., Mitchell, R. C. and Ratick, S. J., 1990, Two types of global environmental change: definitional and spatial-scale issues in their human dimensions. Global Environmental Change, 1, 1422.CrossRefGoogle Scholar
Turvey, S. T. (ed.) (2009). Holocene Extinctions. Oxford: Oxford University Press.CrossRefGoogle Scholar
Tuyet, D. (2001). Characteristics of karst ecosystems of Vietnam and their vulnerability to human impact. Acta Geologica Sinica, 75, 325–9.CrossRefGoogle Scholar
Tweel, A. W. and Turner, R. E. (2012). Watershed land use and river eigineering drive wetland formation and loss in the Mississippi birdfoot delta. Limnology and Oceanography, 57, 1828.CrossRefGoogle Scholar
Tylmann, W. (2005). Lithological and geochemical record of anthropogenic changes in recent sediments of a small and shallow lake (Lake Pusty Staw, northern Poland). Journal of Paleolimnology, 33, 313–25.CrossRefGoogle Scholar
Urich, P. B., Day, M. J., and Lynagh, F. (2001). Policy and practice in karst landscape protection: Bohol, the Philippines. Geographical Journal, 167, 305–23.CrossRefGoogle Scholar
Unger, P. W., Stewart, B. A., Parr, J. F., and Singh, R. P. (1991). Crop residue management and tillage methods for conserving soil and water in semi-arid regions. Soil and Tillage Research, 20, 219–40.CrossRefGoogle Scholar
Urban, M. A. and Rhoads, B. L. (2003). Catastrophic human-induced change in stream-channel planform and geometry in an agricultural watershed, Illinois, USA. Annals of the Association of American Geographers, 93, 783–96.CrossRefGoogle Scholar
Vacca, A., Loddo, S., Ollesch, G., Puddu, R., Serra, G., Tomasi, D., and Aru, A. (2000). Measurement of runoff and soil erosion in three areas under different land use in Sardinia (Italy). Catena, 40, 6992.CrossRefGoogle Scholar
Valentin, C., Rajot, J.-L., and Mitja, D. (2004). Response of soil crusting, runoff and erosion to fallowing in the sub-humid and semi-arid regions of West Africa. Agriculture, Ecosystems and Environment, 104, 287302.CrossRefGoogle Scholar
Valese, E., Conedera, M., Held, A. C., and Ascoli, D. (2014). Fire, humans and landscape in the European Alpine region during the Holocene. Anthropocene, 6, 6374.CrossRefGoogle Scholar
van Andel, T. H., Zangger, E., and Demitrack, A. (1990). Land use and soil erosion in prehistoric and historical Greece. Journal of Field Archaeology, 17, 379–96.Google Scholar
van Beynen, P. E. (Ed.). (2011). Karst management. Dordrecht: Springer Science & Business Media.CrossRefGoogle Scholar
Van Dam, P. J. (2001). Sinking peat bogs: environmental change Holland, 1350–1550. Environmental History, 6, 3245.CrossRefGoogle Scholar
Van der Broeke, M. and 8 others. (2009). Partitioning recent Greenland mass loss. Science, 326, 984–6.Google Scholar
Van der Noort, R. (2013). Climate Change Archaeology. Building Resilence from Research in the World’s Coastal Wetlands. Oxford: Oxford University Press.CrossRefGoogle Scholar
Van der Pluijm, B. (2014). Hello Anthropocene, goodbye Holocene. Earth’s Future, 2, 566–8.CrossRefGoogle Scholar
Van der Post, K. D., Oldfield, F., Haworth, E. Y., Crooks, P. R. J., and Appleby, P. G. (1997). A record of accelerated erosion in the recent sediments of Blelham Tarn in the English Lake District. Journal of Paleolimnology, 18, 103–20.CrossRefGoogle Scholar
Van der Wal, D., Pye, K., and Neal, A. (2002). Long-term morphological change in the Ribble Estuary, northwest England. Marine Geology, 189, 249–66.CrossRefGoogle Scholar
Van der Ween, C. J. (2002). Polar ice sheets and global sea-level: how well can we predict the future? Global and Planetary Change, 32, 165–94.Google Scholar
Van Donk, S. J., Huang, X., Skidmore, E. L., Anderson, A. B., Gebhart, D. L., Prehoda, V. E., and Kellogg, E. M. (2003). Wind erosion from military training lands in the Mojave Desert, California, USA. Journal of Arid Environments, 54, 687703.CrossRefGoogle Scholar
Van Manh, N., Dung, N. V., Hung, N. N., Kummu, M., Merz, B., and Apel, H. (2015). Future sediment dynamics in the Mekong Delta floodplains: impacts of hydropower development, climate change and sea level rise. Global and Planetary Change, 127, 2233.CrossRefGoogle Scholar
Van Oost, K., Govers, G., de Alba, S., and Quine, T. A. (2006). Tillage erosion: a review of controlling factors and implications for soil quality. Progress in Physical Geography, 30, 443–66.Google Scholar
Van Rompaey, A. J., Govers, G., and Puttemans, C. (2002). Modelling land use changes and their impact on soil erosion and sediment supply to rivers. Earth Surface Processes and Landforms, 27, 481–94.CrossRefGoogle Scholar
Van Vliet, M. T., Franssen, W. H., Yearsley, J. R., Ludwig, F., Haddeland, I., Lettenmaier, D. P., and Kabat, P. (2013). Global river discharge and water temperature under climate change. Global Environmental Change, 23, 450–64.CrossRefGoogle Scholar
van Wesenbeeck, B. K., Mulder, J. P., Marchand, M., Reed, D. J., de Vries, M. B., de Vriend, H. J., and Herman, P. M. (2014). Damming deltas: a practice of the past? Towards nature-based flood defenses. Estuarine, Coastal and Shelf Science, 140, 16.CrossRefGoogle Scholar
Vannière, B., Bossuet, G., Walter-Simonnet, A. V., Gauthier, E., Barral, P., Petit, C., and Daubigney, A. (2003). Land use change, soil erosion and alluvial dynamic in the lower Doubs Valley over the 1st millenium AD (Neublans, Jura, France). Journal of Archaeological Science, 30, 1283–99.CrossRefGoogle Scholar
Vanwalleghem, T., Amate, J. I., de Molina, M. G., Fernández, D. S., and Gómez, J. A. (2011). Quantifying the effect of historical soil management on soil erosion rates in Mediterranean olive orchards. Agriculture, Ecosystems & Environment, 142, 341–51.CrossRefGoogle Scholar
Vanwalleghem, T., Laguna, A., Giráldez, J. V., and Jiménez-Hornero, F. J. (2010). Applying a simple methodology to assess historical soil erosion in olive orchards. Geomorphology, 114, 294302.CrossRefGoogle Scholar
Vařilová, Z., Přikryl, R., and Zvelebil, J. (2015). Factors and processes in deterioration of a sandstone rock form (Pravčická brána Arch, Bohemian Switzerland NP, Czech Republic). Zeitschrift für Geomorphologie, Supplementary Issues, 59, 81101.CrossRefGoogle Scholar
Vaudour, J. (1986). Travertins holocènes et pression anthropique. Mediterranée, 10, 168–73.Google Scholar
Vaughan, D. G. and Doake, C. S. M. (1996). Recent atmospheric warming and retreat of ice shelves on the Antarctic Peninsula. Nature, 379, 328–31.CrossRefGoogle Scholar
Vaughan, D. G. and Spouge, J. R. (2002). Risk estimation of collapse of the west Antarctic ice sheet. Climate Change, 52, 6591.CrossRefGoogle Scholar
Vautard, R., Cattiaux, J., Yiou, P., Thépaut, J.-N., and Ciais, P. (2010). Northern Hemisphere atmospheric stilling partly attributed to an increase in surface roughness. Nature Geoscience, 3, 756–61.CrossRefGoogle Scholar
Veni, G. (1999). A geomorphological strategy for conducting environmental impact assessments in karst areas. Geomorphology, 31,151–80.CrossRefGoogle Scholar
Venteris, E. R. (1999). Rapid tide water glacier retreat: a comparison between Columbia Glacier, Alaska and Patagonian calving glaciers. Global and Planetary Change, 22, 131–8.CrossRefGoogle Scholar
Verheijen, F. G., Jones, R. J., Rickson, R. J., and Smith, C. J. (2009). Tolerable versus actual soil erosion rates in Europe. Earth-Science Reviews, 94, 2338.CrossRefGoogle Scholar
Vermaire, J. C., Pisaric, M. F., Thienpont, J. R., Courtney Mustaphi, C. J., Kokelj, S. V., and Smol, J. P. (2013). Arctic climate warming and sea ice declines lead to increased storm surge activity. Geophysical Research Letters, 40, 1386–90.CrossRefGoogle Scholar
Vermeer, M. and Rahmstorf, S. (2009). Global sea level linked to global temperature. Proceedings of the National Academy of Sciences, 106, 21527–32.CrossRefGoogle ScholarPubMed
Verstraeten, G., Poesen, J., de Vente, J., and Koninckx, X. (2003). Sediment yield variability in Spain: a quantitative and semiqualitative analysis using reservoir sedimentation rates. Geomorphology, 50, 327–48.CrossRefGoogle Scholar
Vice, R. B., Guy, H. P., and Ferguson, G. E. (1969). Sediment movement in an area of suburban highway construction, Scott Run Basin, Fairfax, County, Virginia, 1961–64. United States Geological Survey Water Supply Paper, 1591-E.Google Scholar
Vieira, D. A. and Dabney, S. M. (2011). Modeling edge effects of tillage erosion. Soil and Tillage Research, 111, 197207.CrossRefGoogle Scholar
Viles, H. A. (2002). Implications of future climate change for stone deterioration. In Natural Stone, Weathering Phenomena, Conservation Strategies and Case Studies, ed. Siegesmund, S., Weiss, T., and Vollbrecht, J. A.. Geological Society of London Special Publication 205, pp. 407–18.Google Scholar
Viles, H. A. (2003). Conceptual modelling of the impacts of climate change on karst geomorphology in the UK and Ireland. Journal for Nature Conservation, 11, 5966.CrossRefGoogle Scholar
Viles, H. A. and Cutler, N. A. (2012). Global environmental change and the biology of heritage structures. Global Change Biology, 18, 2406–18.CrossRefGoogle Scholar
Viles, H. A. and Spencer, T. (1995). Coastal Problems. Geomorphology, Ecology and Society at the Coast. London: .ArnoldGoogle Scholar
Vilímek, V., Zapata, M. L., Klimeš, J., Patzelt, Z., and Santillán, N. (2005). Influence of glacial retreat on natural hazards of the Palcacocha Lake area, Peru. Landslides, 2, 107–15.CrossRefGoogle Scholar
Villmoare, B., Kimbel, W. H., Seyoum, C., Campisano, C. J., DiMaggio, E. N., Rowan, J., and Reed, K. E. (2015). Early Homo at 2.8 Ma from Ledi-Geraru, Afar, Ethiopia. Science, 347, 1352–5.CrossRefGoogle ScholarPubMed
Vincent, K. R., Friedman, J. M., and Griffin, E. R. (2009). Erosional consequence of saltcedar control. Environmental Management, 44, 218–27.CrossRefGoogle ScholarPubMed
Visconti, G. (2014). Anthropocene: another academic invention? Rendiconti Lincei, 25, 381–92.CrossRefGoogle Scholar
Vita-Finzi, C. (1969). The Mediterranean Valleys. Cambridge: Cambridge University Press.Google Scholar
Voarintsoa, N. R. G., Cox, R., Razanatseheno, M. O. M., and Rakotondrazafy, A. F. M. (2012). Relation between bedrock geology, topography and lavaka distribution in Madagascar. South African Journal of Geology, 115, 225–50.CrossRefGoogle Scholar
Völkel, J., Leopold, M., Dötterl, S., Schneider, A., Hürkamp, K., and Hilgers, A. (2011). Origin and age of the Lower Bavarian sand dune landscape around Abensberg and Siegenburg. Zeitschrift für Geomorphologie, 55, 515–36.CrossRefGoogle Scholar
Vörösmarty, C. J., Meybeck, M., Fekete, B., Sharma, K., Green, P., and Syvitski, J. P. (2003). Anthropogenic sediment retention: major global impact from registered river impoundments. Global and Planetary Change, 39, 169–90.CrossRefGoogle Scholar
Wada, Y., van Beek, L. P. H., van Kempen, C. M., Reckman, J. W. T. M., Vasak, S., and Bierkens, M. F. P. (2010). Global depletion of groundwater resources. Geophysical Research Letters, 37, L20402, doi:10.1029/2010GL044571.CrossRefGoogle Scholar
Wakindiki, I. I. C. and Ben-Hur, M. (2002). Indigenous soil and water conservation techniques: effects on runoff, erosion, and crop yields under semi-arid conditions. Australian Journal of Soil Research, 40, 367–79.Google Scholar
Walker, H. J. (1988). Artificial Structures and Shorelines. Dordrecht: Kluwer.CrossRefGoogle Scholar
Walker, H. J., Coleman, J. M., Roberts, H. H., and Tye, R. S. (1987). Wetland loss in Louisiana. Geografiska Annaler, 69A, 189200.CrossRefGoogle Scholar
Walker, M., Gibbard, P., and Lowe, J. (2015). Comment on “When did the Anthropocene begin? A mid-twentieth century boundary is stratigraphically optimal” by Zalasiewicz, Jan et al. (2015), Quaternary International, 383, 196203.Google Scholar
Walling, D. E. (2006). Human impact on land–ocean sediment transfer by the world's rivers. Geomorphology, 79, 192216.CrossRefGoogle Scholar
Walling, D. E. and Gregory, K. J. (1970). The measurement of the effects of building construction on drainage basin dynamics. Journal of Hydrology, 11, 129–44.CrossRefGoogle Scholar
Wallwork, K. L. (1974). Derelict Land. Newton Abbott: David and Charles.Google Scholar
Walsh, R. P. and Blake, W. H. (2009). Tropical rainforests. In Geomorphology and Global Environmental Change, ed. Slaymaker, O., Spencer, T., and Embleton-Hamman, C.. Cambridge: Cambridge University Press, pp. 214–47.Google Scholar
Walter, R. C. and Merritts, D. J. (2008). Natural streams and the legacy of water-powered mills. Science, 2319, 299304.CrossRefGoogle Scholar
Wan, S., Toucanne, S., Clift, P. D., Zhao, D., Bayon, G., Yu, Z., and Li, T. (2015). Human impact overwhelms long-term climate control of weathering and erosion in southwest China. Geology, 43, 439–42.CrossRefGoogle Scholar
Wang, C., Dong, S., Evan, A. T., Foltz, G. R., and Lee, S. K. (2012). Multidecadal covariability of North Atlantic sea surface temperature, African dust, Sahel rainfall, and Atlantic hurricanes. Journal of Climate, 25, 5404–15.CrossRefGoogle Scholar
Wang, F., Zhang, Y., Huo, Z., Peng, X., Araiba, K., and Wang, G. (2008). Movement of the Shuping landslide in the first four years after the initial impoundment of the Three Gorges Dam Reservoir, China. Landslides, 5, 321–9.Google Scholar
Wang, H., Saito, Y., Zhang, Y., Bi, N., Sun, X., and Yang, Z. (2011). Recent changes of sediment flux to the western Pacific Ocean from major rivers in East and Southeast Asia. Earth-Science Reviews, 108, 80100.CrossRefGoogle Scholar
Wang, P., Li, Z., Luo, S., Bai, J., Huai, B., Wang, F., and Wang, L. (2015). Five decades of changes in the glaciers on the Friendship Peak in the Altai Mountains, China: changes in area and ice surface elevation. Cold Regions Science and Technology, 116, 2431.CrossRefGoogle Scholar
Wang, T., Belle, I., and Hassler, U. (2015). Modelling of Singapore’s topographic transformation based on DEMs. Geomorphology, 231, 367–75.CrossRefGoogle Scholar
Wang, W., Liu, H., Li, Y., and Su, J. (2014). Development and management of land reclamation in China. Ocean & Coastal Management, 102, 415–28.CrossRefGoogle Scholar
Wang, X., Eerdun, H., Zhou, Z., and Liu, X. (2007). Significance of variations in the wind energy environment over the past 50 years with respect to dune activity and desertification in arid and semiarid northern China. Geomorphology, 86, 252–66.CrossRefGoogle Scholar
Wang, X., Siegert, F., Zhou, A. G., and Franke, J. (2013). Glacier and glacial lake changes and their relationship in the context of climate change, Central Tibetan Plateau 1972–2010. Global and Planetary Change, 111, 246–57.CrossRefGoogle Scholar
Wang, X., Thompson, D. K., Marshall, G. A., Tymstra, C., Carr, R., and Flannigan, M. D. (2015). Increasing frequency of extreme fire weather in Canada with climate change. Climatic Change, 130, 573–86.CrossRefGoogle Scholar
Wang, X. M., Zhang, C. X., Hasi, E., and Dong, Z. B. (2010). Has the Three Norths Forest Shelterbelt Program solved the desertification and dust storm problems in arid and semiarid China? Journal of Arid Environments, 74, 1322.CrossRefGoogle Scholar
Warburton, J. and Evans, M. (2011). Geomorphic, sedimentary, and potential palaeoenvironmental significance of peat blocks in alluvial river systems. Geomorphology, 130, 101–14.CrossRefGoogle Scholar
Warner, R. C. and Budd, W. F. (1990). Modelling the long-term response of the Antarctic Ice Sheet to global warming. Annals of Glaciology, 27, 161–68.Google Scholar
Warren, A. (2013). Dunes. Chichester: Wiley.CrossRefGoogle Scholar
Warrick, J. A., Madej, M. A., Goñi, M. A., and Wheatcroft, R. A. (2013). Trends in the suspended-sediment yields of coastal rivers of northern California, 1955–2010. Journal of Hydrology, 489, 108–23.CrossRefGoogle Scholar
Washington, R. and 9 others. (2009). Dust as a tipping element: the Bodélé Depression, Chad. Proceedings of the National Academy of Sciences, 106, 20564–71.CrossRefGoogle ScholarPubMed
Wasson, R. J. (2012). Geomorphic histories for river and catchment management. Philosophical Transactions of the Royal Society, A, 370, 2240–63.Google ScholarPubMed
Watanabe, T., Lamsal, D., and Ives, J. D. (2009). Evaluating the growth characteristics of a glacial lake and its degree of danger of outburst flooding: Imja Glacier, Khumbu Himal, Nepal. Norsk Geografisk Tidsskrift, 63, 255–67.CrossRefGoogle Scholar
Waters, M. R. and Haynes, C. V. (2001). Late Quaternary arroyo formation and climate change in the American southwest. Geology, 29, 399402.2.0.CO;2>CrossRefGoogle Scholar
Watkins, T. (2010). New light on Neolithic revolution in south-west Asia. Antiquity, 84, 621–34.CrossRefGoogle Scholar
Watson, A. (1990). The control of blowing sand and mobile desert dunes. In Techniques for Desert Reclamation, ed. Goudie, A. S.. Chichester: Wiley, pp. 3585.Google Scholar
Waycott, M., Duarte, C. M., Carruthers, T. J., Orth, R. J., Dennison, W. C., Olyarnik, S., and Williams, S. L. (2009). Accelerating loss of seagrasses across the globe threatens coastal ecosystems. Proceedings of the National Academy of Sciences, 106, 12377–81.CrossRefGoogle ScholarPubMed
Webb, E. L., Friess, D. A., Krauss, K. W., Cahoon, D. R., Guntenspergen, G. R., and Phelps, J. (2013). A global standard for monitoring coastal wetland vulnerability to accelerated sea-level rise. Nature Climate Change, 3, 458–65.CrossRefGoogle Scholar
Webb, N. P., Chappell, A., Strong, C. L., Marx, S. K., and McTainsh, G. H. (2012). The significance of carbon‐enriched dust for global carbon accounting. Global Change Biology, 18, 3275–8.CrossRefGoogle Scholar
Webb, R. H., Boyer, D. E., and Turner, R. M. (eds.) (2010). Repeat Photography: Methods and Applications in the Natural Sciences. Washington, DC: Island Press.Google Scholar
Weber, K. A. and Perry, R. G. (2006). Groundwater abstraction impacts on spring flow and base flow in the Hillsborough River Basin, Florida, USA. Hydrogeology Journal, 14, 1252–64.CrossRefGoogle Scholar
Webster, P. J., Holland, G. J., Curry, J. A., and Chang, H.-R. (2005). Changes in tropical cyclone number, duration, and intensity in a warming environment. Science, 309, 1844–6.CrossRefGoogle Scholar
Wehrli, M., Mitchell, E. A., van der Knaap, W. O., Ammann, B., and Tinner, W. (2010). Effects of climatic change and bog development on Holocene tufa formation in the Lorze Valley (central Switzerland). The Holocene, 20, 325–36.CrossRefGoogle Scholar
Weinhold, B. (2012). Energy development linked with earthquakes. Environmental Health Perspectives, 120, a388.CrossRefGoogle ScholarPubMed
Weisrock, A. (1986). Variations climatiques et periodes de sedimentation carbonatée a l’Holocene - l’age des depôts. Mediterranée, 10, 165–7.Google Scholar
Wells, J. T. (1995). Effects of sea level rise on coastal sedimentation and erosion. In Climate Change Impact on Coastal Habitation, ed. Eisma, D.. Boca Raton: Lewis, pp. 111–36.Google Scholar
Wells, N. A. and Andriamihaja, B. (1993). The Initiation and growth of gullies in Madagascar-are humans to blame. Geomorphology, 8, 146.CrossRefGoogle Scholar
Wemple, B. C. and Jones, J. A. (2003). Runoff production on forest roads in a steep, mountain catchment. Water Resources Research, 39, doi: 10.1029/2002WR001744.CrossRefGoogle Scholar
Wemple, B. C., Swanson, F. J., and Jones, J. A. (2001). Forest roads and geomorphic process interactions, Cascade Range, Oregon. Earth Surface Processes and Landforms, 26, 191204.3.0.CO;2-U>CrossRefGoogle Scholar
Wen, F. and Chen, X. (2006). Evaluation of the impact of groundwater irrigation on streamflow in Nebraska. Journal of Hydrology, 327, 603–17.CrossRefGoogle Scholar
Westerling, A. L., Turner, M. G., Smithwick, E. A., Romme, W. H., and Ryan, M. G. (2011). Continued warming could transform Greater Yellowstone fire regimes by mid-21st century. Proceedings of the National Academy of Sciences, 108, 13165–70.CrossRefGoogle ScholarPubMed
Westing, A. and Pfeiffer, E. W. (1972). The cratering of Indochina. Scientific American, 226, 5, 21–9.CrossRefGoogle Scholar
Wharton, G. and Gilvear, D. J. (2007). River restoration in the UK: meeting the dual needs of the European Union Water Framework Directive and flood defence? International Journal of River Basin Management, 5, 143–54.CrossRefGoogle Scholar
Wheatcroft, R. A., Goñi, M. A., Richardson, K. N., and Borgeld, J. C. (2013). Natural and human impacts on centennial sediment accumulation patterns on the Umpqua River margin, Oregon. Marine Geology, 339, 4456.CrossRefGoogle Scholar
Wheaton, E. E. (1990). Frequency and severity of drought and dust storms. Canadian Journal of Agricultural Economics, 38, 695700.CrossRefGoogle Scholar
White, A. F. and Blum, A. E. (1995). Climatic effects on chemical weathering in watersheds: application of mass balance approaches. In Solute Modelling in Catchment Systems, ed. Trudgill, S.. Chichester: John Wiley and Sons, pp. 101–31.Google Scholar
Whitmore, T. M. and Turner, B. L. (2001). Cultivated Landscapes of Middle America on the Eve of Conquest. Oxford: Oxford University Press.Google Scholar
Whitaker, J. R. (1940). World view of destruction and conservation of natural resources. Annals of the Association of American Geographers, 30, 143–62.CrossRefGoogle Scholar
Whittington, G. (1962). The distribution of strip lynchets. Transactions and Papers of the Institute of British Geographers, 31, 115–30.Google Scholar
Wienhold, M. L. (2013). Prehistoric land use and hydrology: a multi-scalar spatial analysis in central Arizona. Journal of Archaeological Science, 40, 850–9.CrossRefGoogle Scholar
Wilby, R. L., Dalgleish, H. Y., and Foster, I. D. L. (1997). The impact of weather patterns on historic and contemporary catchment sediment yields. Earth Surface Processes and Landforms, 22, 353–63.3.0.CO;2-G>CrossRefGoogle Scholar
Wildman, L. A. S. and Macbroom, J. G. (2005). The evolution of gravel bed channels after dam removal: case study of the Anaconda and Union City dam removals. Geomorphology, 71, 245–62.CrossRefGoogle Scholar
Wilkinson, B. H. and McElroy, B. J. (2007). The impact of humans on continental erosion and sedimentation. Bulletin of the Geological Society of America, 119, 140–56.CrossRefGoogle Scholar
Wilkinson, J. C., (1977) Water and Tribal Settlement in South-East Arabia: a Study of the Aflāj of Oman. Oxford: Clarendon PressGoogle Scholar
Wilkinson, T. J. and Rayne, L. (2010). Hydraulic landscapes and imperial power in the Near East. Water History, 2, 115–44.CrossRefGoogle Scholar
Wilkinson, T. J., French, C., Ur, J. A., and Semple, M. (2010). The geoarchaeology of route systems in northern Syria. Geoarchaeology, 25, 745–71.CrossRefGoogle Scholar
Williams, M. (2003). Deforesting the Earth: From Prehistory to Global Crisis. Chicago: The University of Chicago Press.Google Scholar
Williams, M., Zalasiewicz, J., Davies, N., Mazzini, I., Goiran, J. P., and Kane, S. (2014). Humans as the third evolutionary stage of biosphere engineering of rivers. Anthropocene, 7, 5763.CrossRefGoogle Scholar
Williams, P., Biggs, J., Crowe, A., Murphy, J., Nicolet, P., Waetherby, A., and Dunbar, M. (2010). Ponds report from 2007. Countryside Survey Technical Report No. 7/07.Google Scholar
Williams, P. W. (ed.) (1993). Karst terrains: environmental changes and human impact. Catena supplement, 25.Google Scholar
Williams, P. W. (2008). World heritage caves and karst. Gland, Switzerland: IUCN.Google Scholar
Williams, S. J. (2013). Sea-level rise implications for coastal regions. Journal of Coastal Research, 63(sp1), 184–96.CrossRefGoogle Scholar
Williams, T. J., Quinton, W. L., and Baltzer, J. L. (2013). Linear disturbances on discontinuous permafrost: implications for thaw-induced changes to land cover and drainage patterns. Environmental Research Letters, 8(2), 025006.CrossRefGoogle Scholar
Willis, C. M. and Griggs, G. B. (2003). Reductions in fluvial sediment discharge by coastal dams in California and implications for beach sustainability. Journal of Geology, 111, 167–82.CrossRefGoogle Scholar
Willis, K. J., Gillson, L., and Brncic, T. M. (2004). How “virgin” is virgin rainforest? Science, 304, 402–3.CrossRefGoogle Scholar
Wilson, C. J. (1999). Effects of logging and fire on runoff and erosion on highly erodible granitic soils in Tasmania. Water Resources Research, 35, 3531–46.CrossRefGoogle Scholar
Wilson, L., Wilson, J., Holden, J., Johnstone, I., Armstrong, A., and Morris, M. (2011). Ditch blocking, water chemistry and organic carbon flux: evidence that blanket bog restoration reduces erosion and fluvial carbon loss. Science of the Total Environment, 409, 2010–8.CrossRefGoogle ScholarPubMed
Winkler, E. M. (1970). The importance of air pollution in the corrosion of stone and metals. Engineering Geology, 4, 327–34.CrossRefGoogle Scholar
Winter, M. G., Dixon, N., Wasowski, J., and Dijkstra, T. A. (2010). Introduction to land-use and climate change impacts on landslides. Quarterly Journal of Engineering Geology and Hydrogeology, 43, 367–70.CrossRefGoogle Scholar
Winter, R. C. de, Sterl, A., de Vries, J. W., Weber, S. L., and Ruessink, G. (2012). The effect of climate change on extreme waves in front of the Dutch coast. Ocean Dynamics, 62, 1139–52.CrossRefGoogle Scholar
Wishart, D. and Warburton, J. (2001). An assessment of blanket mire degradation and peatland gully development in the Cheviot Hills, Northumberland. Scottish Geographical Magazine, 117, 185206.CrossRefGoogle Scholar
Wisser, D., Frolking, S., Hagen, S., and Bierkens, M. F. (2013). Beyond peak reservoir storage? A global estimate of declining water storage capacity in large reservoirs. Water Resources Research, 49, 5732–9.CrossRefGoogle Scholar
Woeikof, A. (1901). De l’influence de 1’homme sur la terre. Annales de Géographie, 10, 97114, 193–215.CrossRefGoogle Scholar
Wohl, E. (2013). Wilderness is dead: whither critical zone studies and geomorphology in the Anthropocene. Anthropocene, 2, 415.CrossRefGoogle Scholar
Wohl, E. (2015). Legacy effects on sediments in river corridors. Earth-Science Reviews, 147, 3053.CrossRefGoogle Scholar
Wohl, E. and Merritts, D. J. (2007). What is a natural river? Geography Compass, 1, 871900.CrossRefGoogle Scholar
Wohl, E., Angermeier, P. L., Bledsoe, B., Kondolf, G. M., MacDonnell, L., Merritt, D. M., and Tarboton, D. (2005). River restoration. Water Resources Research, 41, W10301, doi: 10.1029/2005WR003985.CrossRefGoogle Scholar
Wolanski, E., Moore, K., Spagnol, S., D’adamo, N., and Pattiaratchi, C. (2001). Rapid, human-induced siltation of the macro-tidal Ord River Estuary, Western Australia. Estuarine, Coastal and Shelf Science, 53, 717–32.CrossRefGoogle Scholar
Wolfe, S. A., Hugenholtz, C. H., Evans, C. P., Huntley, D. J., and Ollerhead, J. (2007). Potential aboriginal-occupation-induced dune activity, Elbow Sand Hills, northern Great Plains, Canada. Great Plains Research, 17, 173–92.Google Scholar
Wolff, W. J. (1992). The end of a tradition: 1000 years of embankment and reclamation of wetlands in the Netherlands. Ambio, 21, 287–91.Google Scholar
Wolman, M. G. (1967). A cycle of sedimentation and erosion in urban river channels. Geografiska Annaler, 49A, 385–95.Google Scholar
Wolman, M. G. and Schick, A. P. (1967). Effects of construction on fluvial sediment, urban and suburban areas of Maryland. Water Resources Research, 3, 451–64.CrossRefGoogle Scholar
Wondzell, S. M. and King, J. G. (2003). Postfire erosional processes in the Pacific Northwest and Rocky Mountain regions. Forest Ecology and Management, 178, 7587.CrossRefGoogle Scholar
Wong, S., Dessler, A. E., Mahowald, N., Colarco, P. R., and da Silva, A. (2008). Long-term variability in Saharan dust transport and its link to North Atlantic sea surface temperature. Geophysical Research Letters, 35, L07812, doi: 10.1029/2007GL032297.CrossRefGoogle Scholar
Woo, M.-K., Lewkowicz, A. G., and Rouse, W. R. (1992). Response of the Canadian permafrost environment to climate change. Physical Geography, 13, 287317.CrossRefGoogle Scholar
Wood, C. (2009). World Heritage Volcanoes: Thematic Study, Global Review of Volcanic World Heritage Properties: Present Situation, Future Prospects and Management Requirements. Gland: IUCN.Google Scholar
Woodroffe, C. D. (1990). The impact of sea-level rise on mangrove shorelines. Progress in Physical Geography, 14, 483520.CrossRefGoogle Scholar
Woodroffe, C. D. (2008). Reef-island topography and the vulnerability of atolls to sea-level rise. Global and Planetary Change, 62, 7796.CrossRefGoogle Scholar
Woodruff, J. D., Irish, J. L., and Camargo, S. J. (2013). Coastal flooding by tropical cyclones and sea-level rise. Nature, 504, 4452.CrossRefGoogle ScholarPubMed
Woodruff, J. D., Martini, A. P., Elzidani, E. Z., Naughton, T. J., Kekacs, D. J., and MacDonald, D. G. (2013). Off-river waterbodies on tidal rivers: Human impact on rates of infilling and the accumulation of pollutants. Geomorphology, 184, 3850.CrossRefGoogle Scholar
Woodward, C., Shulmeister, J., Larsen, J., Jacobsen, G. E., and Zawadzki, A. (2014). The hydrological legacy of deforestation on global wetlands. Science, 346, 844–7.CrossRefGoogle ScholarPubMed
Woolsey, S., Capelli, F., Gonser, T. O. M., Hoehn, E., Hostmann, M., Junker, B., and Peter, A. (2007). A strategy to assess river restoration success. Freshwater Biology, 52, 752–69.CrossRefGoogle Scholar
World Glacier Monitoring Service. (2008). Global glacier changes: facts and figures, www.grid.unep.ch/glaciers/ (Accessed October 29, 2011).Google Scholar
Wosten, J. H. M., Ismail, A. B., andVan Wijk, A. L. M. (1997). Peat subsidence and its practical implications: a case study in Malaysia. Geoderma, 78, 2536.CrossRefGoogle Scholar
Woth, K., Weisse, R., and von Storch, H. (2006). Climate change and North Sea storm surge extremes: an ensemble study of storm surge extremes expected in a changed climate projected by four different regional climate models. Ocean Dynamics, 56, 315.CrossRefGoogle Scholar
Wright, S. A. and Schoellhamer, D. H. (2004). Trends in the sediment yield of the Sacramento River, California, 1957–2001. San Francisco Estuary and Watershed Science (on line serial), 2(2), Article 2.CrossRefGoogle Scholar
Wu, C. S., Yang, S. L., and Lei, Y. P. (2012). Quantifying the anthropogenic and climatic impacts on water discharge and sediment load in the Pearl River (Zhujiang), China (1954–2009). Journal of Hydrology, 452, 190204.CrossRefGoogle Scholar
Wu, J., Shi, X., Xue, Y., Zhang, Y., Wei, Z., and Yu, J. (2008). The development and control of the land subsidence in the Yangtze Delta, China. Environmental Geology, 55, 1725–35.CrossRefGoogle Scholar
Wu, Q., Hou, Y., Yun, H., and Liu, Y. (2015). Changes in active-layer thickness and near-surface permafrost between 2002 and 2012 in alpine ecosystems, Qinghai–Xizang (Tibet) Plateau, China. Global and Planetary Change, 124, 149–55.CrossRefGoogle Scholar
Wu, Q. Jiewu, P., Shanzhong, Q., Yiping, L., Congcong, H., Tingxiang, L., and Limei, H. (2009). Impacts of coal mining subsidence on the surface landscape in Longkou city, Shandong Province of China. Environmental Earth Sciences, 59, 783–91.Google Scholar
Wu, R. S., Sue, W. R., Chien, C. B., Chen, C. H., Chang, J. S., and Lin, K. M. (2001). A simulation model for investigating the effects of rice paddy fields on the runoff system. Mathematical and Computer Modelling, 33, 649–58.CrossRefGoogle Scholar
Wu, Y. and Zhu, L. (2008). The response of lake-glacier variations to climate change in Nam Co Catchment, central Tibetan Plateau, during 1970–2000. Journal of Geographical Sciences, 18, 177–89.CrossRefGoogle Scholar
Wüst, R. A. J. and Schlüchter, C. (2000). The origin of soluble salts in rocks of the Thebes Mountains, Egypt: the damage potential to ancient Egyptian wall art. Journal of Archaeological Science, 27, 1161–72.CrossRefGoogle Scholar
Wyzga, B. (1996). Changes in the magnitude and transformation of flood waves subsequent to the channelization of the Raba River, Polish Carpathians. Earth Surface Processes and Landforms, 21, 749–63.3.0.CO;2-5>CrossRefGoogle Scholar
Xu, K. and Milliman, J. D. (2009). Seasonal variations of sediment discharge from the Yangtze River before and after impoundment of the Three Gorges Dam. Geomorphology, 104, 276–83.CrossRefGoogle Scholar
Xu, K., Milliman, J. D., Yang, Z., and Xu, H. (2007). Climatic and anthropogenic impacts on water and sediment discharges from the Yangtze River (Changjiang), 1950–2005. In Large Rivers: Geomorphology and Management, ed. Gupta, A.. Chichester: Wiley, pp. 609–26.Google Scholar
Xu, Y. S., Shen, S. L., Cai, Z. Y., and Zhou, G. Y. (2008). The state of land subsidence and prediction approaches due to groundwater withdrawal in China. Natural Hazards, 45, 123–35.CrossRefGoogle Scholar
Xue, Y. Q., Zhang, Y., Ye, S. J., Wu, J. C., and Li, Q. F. (2005). Land subsidence in China. Environmental Geology, 48, 713–20.CrossRefGoogle Scholar
Yamano, H., Sugihara, K., and Nomura, K. (2011). Rapid poleward range expansion of tropical reef corals in response to rising sea surface temperatures. Geophysical Research Letters, 38, L04601, doi:10.1029/2010GL046474.CrossRefGoogle Scholar
Yang, D., Kanae, S., Oki, T., Koike, T., and Musiake, K. (2003). Global potential soil erosion with reference to land use and climate changes. Hydrological Processes, 17, 2913–28.CrossRefGoogle Scholar
Yang, Q., Wang, K., Zhang, C., Yue, Y., Tian, R., and Fan, F. (2011). Spatio-temporal evolution of rocky desertification and its driving forces in karst areas of Northwestern Guangxi, China. Environmental Earth Science, 64, 383–93.CrossRefGoogle Scholar
Yang, X. and Lu, X. (2014). Drastic change in China’s lakes and reservoirs over the past decades. Scientific Reports, 4, article number: 6041, doi: 10.1038/srep06041.Google ScholarPubMed
Yang, X., Ding, Z., Fan, X., Zhou, Z., and Ma, N. (2007). Processes and mechanisms of desertification in northern China during the last 30 years, with a special reference to the Hunshandake Sandy Land, eastern Inner Mongolia. Catena, 71, 212.CrossRefGoogle Scholar
Yang, X., Shen, S., Yang, F., He, Q., Ali, M., Huo, W., and Liu, X. (2015). Spatial and temporal variations of blowing dust events in the Taklimakan Desert. Theoretical and Applied Climatology, doi: 10.1007/s00704-015-1537-4.CrossRefGoogle Scholar
Yang, Z., Wang, T., Leung, R., Hibbard, K., Janetos, T., Kraucunas, I., and Wilbanks, T. (2014). A modeling study of coastal inundation induced by storm surge, sea-level rise, and subsidence in the Gulf of Mexico. Natural Hazards, 71, 1771–94.CrossRefGoogle Scholar
Yao, T., Pu, J., Lu, A., Wang, Y., and Yu, W. (2007). Recent glacial retreat and its impact on hydrological processes on the Tibetan Plateau, China, and surrounding regions. Arctic, Antarctic, and Alpine Research, 39, 642–50.CrossRefGoogle Scholar
Yarushina, V. M. and Bercovici, D. (2013). Mineral carbon sequestration and induced seismicity. Geophysical Research Letters, 40, 814–8.CrossRefGoogle Scholar
Yechieli, Y., Abelson, M., Bein, A., Crouvi, O., and Shtivelman, V. (2006). Sinkhole “swarms” along the Dead Sea coast: reflection of disturbance of lake and adjacent groundwater systems. Bulletin of the Geological Society of America, 118, 1075–87.CrossRefGoogle Scholar
Yeck, W. L., Block, L. V., Wood, C. K., and King, V. M. (2015). Maximum magnitude estimations of induced earthquakes at Paradox Valley, Colorado, from cumulative injection volume and geometry of seismicity clusters. Geophysical Journal International, 200, 322–36.CrossRefGoogle Scholar
Yesner, D. R. (2001). Human dispersal into interior Alaska: antecedent conditions, mode of colonization, and adaptations. Quaternary Science Reviews, 20, 315–27.CrossRefGoogle Scholar
Yin, Y., Zhang, K., and Li, X. (2006). Urbanization and land subsidence in China. IAEG2006 Paper 31, Geological Society of London.Google Scholar
Yonggui, Y., Xuefa, S., Houjie, W., Chengkun, Y., Shenliang, C., Yanguang, L., and Shuqing, Q. (2013). Effects of dams on water and sediment delivery to the sea by the Huanghe (Yellow River): the special role of water-sediment modulation. Anthropocene, 3, 7282.CrossRefGoogle Scholar
Yorke, T. H. and Herb, W. J. (1978). Effects of urbanization on streamflow and sediment transport in the Rock Creek and Anacostia River Basins, Montgomery County, Maryland, 1962–74. U.S. Geological Survey Professional Paper, 1003.CrossRefGoogle Scholar
Yoshikawa, K. and Hinzman, L. D. (2003). Shrinking thermokarst ponds and groundwater dynamics in discontinuous permafrost near Council, Alaska. Permafrost and Periglacial Processes, 14, 151–60.CrossRefGoogle Scholar
Young, J. E. (1992). Mining the Earth. Worldwatch Institute Paper, 109, 153.Google Scholar
Yu, K., D’Odorico, P., Bhattachan, A., Okin, G. S., and Evan, A. T., (2015), Dust-rainfall feedback in West African Sahel, Geophysical Research Letters, 42, 7563–71.CrossRefGoogle Scholar
Yuill, B., Lavoie, D., and Reed, D. (2009). Understanding subsidence processes in coastal Louisiana. Journal of Coastal Research, SI, 54, 2336.CrossRefGoogle Scholar
Zaimes, G. N. and Schultz, R. C. (2015). Riparian land-use impacts on bank erosion and deposition of an incised stream in north-central Iowa, USA. Catena, 125, 6173.CrossRefGoogle Scholar
Zalasiewicz, J., Kryza, R., and Williams, M. (2014). The mineral signature of the Anthropocene in its deep-time context. Geological Society, London, Special Publications, 395, 109–17.Google Scholar
Zalasiewicz, J., Waters, C. N., and Williams, M. (2014). Human bioturbation, and the subterranean landscape of the Anthropocene. Anthropocene, 6, 39.Google Scholar
Zalasiewicz, J., Williams, M., Haywood, A., and Ellis, M. (2011a). The Anthropocene: a new epoch of geological time? Philosophical Transactions of the Royal Society, 369A: 835–41.CrossRefGoogle Scholar
Zalasiewicz, J., Williams, M., Fortey, R., et al. (2011b). Stratigraphy of the Anthropocene. Philosophical Transactions of the Royal Society, 369A, 1036–55.CrossRefGoogle Scholar
Zalasiewicz, J., Waters, C. N., Williams, M., Barnosky, A. D., Cearreta, A., Crutzen, P., and Oreskes, N. (2015). When did the Anthropocene begin? A mid-twentieth century boundary level is stratigraphically optimal. Quaternary International, 383, 196203.CrossRefGoogle Scholar
Zanello, F., Teatini, P., Putti, M., and Gambolati, G. (2011). Long term peatland subsidence: Experimental study and modeling scenarios in the Venice coastland. Journal of Geophysical Research: Earth Surface, 116(F4), F04002, doi: 10.1029/2011JF002010.CrossRefGoogle Scholar
Zang, A., Oye, V., Jousset, P., Deichmann, N., Gritto, R., McGarr, A., and Bruhn, D. (2014). Analysis of induced seismicity in geothermal reservoirs – An overview. Geothermics, 52, 621.CrossRefGoogle Scholar
Zarfl, C., Lumsdon, A. E., Berlekamp, J., Tydecks, L., and Tockner, K. (2015). A global boom in hydropower dam construction. Aquatic Sciences, 77, 161–70.CrossRefGoogle Scholar
Zarnetske, P. L., Hacker, S. D., Seabloom, E. W., Ruggiero, P., Killian, J. R., Maddux, T. B., and Cox, D. (2012). Biophysical feedback mediates effects of invasive grasses on coastal dune shape. Ecology, 93, 1439–50.CrossRefGoogle ScholarPubMed
Zavada, M. S., Wang, Y., Rambolamanana, G., Raveloson, A., and Razanatsoa, H. (2009). The significance of human induced and natural erosion features (lavakas) on the central highlands of Madagascar. Madagascar Conservation & Development, 4, 120–7.CrossRefGoogle Scholar
Zawiejska, J. and Wyżga, B. (2010). Twentieth-century channel change on the Dunajec River, southern Poland: patterns, causes and controls. Geomorphology, 117, 234–46.CrossRefGoogle Scholar
Zeeberg, J. and Forman, S. L. (2001). Changes in glacier extent of north Novaya Zemlya in the twentieth century. The Holocene, 11, 161–75.CrossRefGoogle Scholar
Zektser, S., Loáiciga, H. A., and Wolf, J. T. (2005). Environmental impacts of groundwater overdraft: selected case studies in the southwestern United States. Environmental Geology, 47, 396404.CrossRefGoogle Scholar
Zemp, M., Frey, H., Gärtner-Roer, I., Nussbaumer, S., Hoelzle, M., Paul, F., and Vincent, C. (2015). Historically unprecedented global glacier decline in the early 21st century. Journal of Glaciology, 61, 745–62, doi: 10.3189/2015JoG15J017.CrossRefGoogle Scholar
Zeng, N. and Yoon, J. (2009). Expansion of the world’s deserts due to vegetation-albedo feedback under global warming. Geophysical Research Letters, 36, L17401, doi: 10.1029/2009GL039699.CrossRefGoogle Scholar
Zennaro, P., Kehrwald, N., Marlon, J., Ruddiman, W., Brücher, T., Agostinelli, C., and Barbante, C. (2015). Europe on fire three thousand years ago: Arson or climate?. Geophysical Research Letters. DOI: 10.1002/2015GL064259.CrossRefGoogle Scholar
Zhang, L., Wu, B., Yin, K., Li, X., Kia, K., and Zhu, L. (2014). Impacts of human activities on the evolution of estuarine wetland in the Yangtze Delta from 2000 to 2010. Environmental Earth Sciences, 73, 435–47, doi: 10.1007/s12665-014-3565-2.CrossRefGoogle Scholar
Zhang, T.-H., Zhao, H.-L., Li, S.-G., Li, F.-R., Shirato, Y., Ohkuro, T., and Taniyama, I. (2004). A comparison of different measures for stabilizing moving sand dunes in the Horqin Sandy Land of Inner Mongolia, China. Journal of Arid Environments, 58, 203–14.CrossRefGoogle Scholar
Zhang, Y., Person, M., Rupp, J., Ellett, K., Celia, M. A., Gable, C. W., and Elliot, T. (2013). Hydrogeologic controls on induced seismicity in crystalline basement rocks due to fluid injection into basal reservoirs. Groundwater, 51, 525–38.CrossRefGoogle ScholarPubMed
Zhang, Y., Xue, Y. Q., Wu, J. C., Shi, X. Q., and Yu, J. (2010). Excessive groundwater withdrawal and resultant land subsidence in the Su-Xi-Chang area, China. Environmental Earth Sciences, 61, 1135–43.CrossRefGoogle Scholar
Zhang, Y., Xue, Y. Q., Wu, J. C., Yu, J., Wei, Z. X., and Li, Q. F. (2008). Land subsidence and earth fissures due to groundwater withdrawal in the Southern Yangtse Delta, China. Environmental Geology, 55, 751–62.CrossRefGoogle Scholar
Zhang, Z. and Wu, Q. (2012). Thermal hazards zonation and permafrost change over the Qinghai–Tibet Plateau. Natural Hazards, 61, 403–23.CrossRefGoogle Scholar
Zhao, S., Fang, J., Miao, S., Gu, B., Tao, S., Peng, C., and Tang, Z. (2005). The 7-decade degradation of a large freshwater lake in Central Yangtze River, China. Environmental Science & Technology, 39, 431–36.CrossRefGoogle ScholarPubMed
Zhou, Z. C., Gan, Z. T., Shangguan, Z. P., and Dong, Z. B. (2010). Effects of grazing on soil physical properties and soil erodibility in semiarid grassland of the Northern Loess Plateau (China). Catena, 82, 8791.CrossRefGoogle Scholar
Zhu, C., Wang, B., and Qian, W. (2008). Why do dust storms decrease in northern China concurrently with the recent global warming? Geophysical Research Letters, 35, L18702, doi: 10.1029/2008GL034886.CrossRefGoogle Scholar
Zhu, D., Tian, L., Wang, J., Wang, Y., and Cui, J. (2014). Rapid glacier retreat in the Naimona’Nyi region, western Himalayas, between 2003 and 2013. Journal of Applied Remote Sensing, 8, 083508–083508.CrossRefGoogle Scholar
Zhu, J. and Olsen, C. R. (2014). Sedimentation and Organic Carbon Burial in the Yangtze River and Hudson River Estuaries: implications for the Global Carbon Budget. Aquatic Geochemistry, 20, 325–42.CrossRefGoogle Scholar
Zhu, R. X., Potts, R., Pan, Y. X., Yao, H. T., , L. Q., Zhao, X., Gao, X., Chen, L. W., Gao, F., and Deng, C. L. (2008). Early evidence of the genus Homo in East Asia. Journal of Human Evolution, 55, 1075–85.CrossRefGoogle ScholarPubMed
Zhuang, Y. and Kidder, T. R. (2014). Archaeology of the Anthropocene in the Yellow River region, China, 8000–2000 cal. BP. The Holocene, 24, 1602–23.CrossRefGoogle Scholar
Ziegler, A. D. and Giambelluca, T. W. (1997). Importance of rural roads as source areas for runoff in mountainous areas of northern Thailand. Journal of Hydrology, 196, 204–29.CrossRefGoogle Scholar
Ziegler, A. D., Bruun, T. B., Guardiola-Claramonte, M., Giambelluca, T. W., Lawrence, D., and Lam, N. T. (2009). Environmental consequences of the demise in swidden cultivation in montane mainland Southeast Asia: hydrology and geomorphology. Human Ecology, 37, 361–73.CrossRefGoogle Scholar
Zimmermann, A., Francke, T., and Elsenbeer, H. (2012). Forests and erosion: insights from a study of suspended-sediment dynamics in an overland flow-prone rainforest catchment. Journal of Hydrology, 428, 170–81.Google Scholar
Zizumbo-Villarreal, D. and Colunga-GarcíaMarín, P. (2010). Origin of agriculture and plant domestication in West Mesoamerica. Genetic Resources and Crop Evolution, 57, 813–25.CrossRefGoogle Scholar
Zobeck, T. M., Baddock, M. C., and Van Pelt, R. S. (2013). Anthropogenic environments. In: Treatise on Geomorphology, ed. Shroder, J., (Editor in Chief), Lancaster, N., Sherman, D. J., Baas, A. C. W.. San Diego, CA: Academic Press, vol. 11, Aeolian Geomorphology, pp. 395413.CrossRefGoogle Scholar
Zöbisch, M. A. (1993). Erosion susceptibility and soil loss on grazing lands in some semiarid and subhumid locations of eastern Kenya. Journal of Soil and Water Conservation, 48, 445–8.Google Scholar
Zolitschka, B., Behre, K. E., and Schneider, J. (2003). Human and climatic impact on the environment as derived from colluvial, fluvial and lacustrine archives—examples from the Bronze Age to the Migration period, Germany. Quaternary Science Reviews, 22, 81100.CrossRefGoogle Scholar
Zuazo, V. H. D. and Pleguezuelo, C. R. R. (2008). Soil-erosion and runoff prevention by plant covers: a review. Agronomy for Sustainable Development, 28, 6586.CrossRefGoogle Scholar
Zu Ermgassen, P. S., Spalding, M. D., Blake, B., Coen, L. D., Dumbauld, B., Geiger, S., and Brumbaugh, R. (2012). Historical ecology with real numbers: past and present extent and biomass of an imperilled estuarine habitat. Proceedings of the Royal Society B: Biological Sciences, 279, 3393–400.Google ScholarPubMed
Zwally, H. J., Abdalafi, W., Herring, T., Larson, K., Saba, J., and Steffen, K. (2002). Surface melt-induced acceleration of Greenland ice-sheet flow. Science, 297, 218–21.CrossRefGoogle ScholarPubMed

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.

  • References
  • Andrew S. Goudie, University of Oxford, Heather A. Viles, University of Oxford
  • Book: Geomorphology in the Anthropocene
  • Online publication: 20 October 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781316498910.012
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.

  • References
  • Andrew S. Goudie, University of Oxford, Heather A. Viles, University of Oxford
  • Book: Geomorphology in the Anthropocene
  • Online publication: 20 October 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781316498910.012
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.

  • References
  • Andrew S. Goudie, University of Oxford, Heather A. Viles, University of Oxford
  • Book: Geomorphology in the Anthropocene
  • Online publication: 20 October 2016
  • Chapter DOI: https://doi.org/10.1017/CBO9781316498910.012
Available formats
×