Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-25wd4 Total loading time: 0 Render date: 2024-04-25T13:26:40.276Z Has data issue: false hasContentIssue false

5 - Landscape Evolution Models and Ecohydrologic Processes

from Part III - Coupling Hillslope Geomorphology, Soils, Hydrology, and Ecosystems

Published online by Cambridge University Press:  27 October 2016

Erkan Istanbulluoglu
Affiliation:
University of Washington
Edward A. Johnson
Affiliation:
University of Calgary
Yvonne E. Martin
Affiliation:
University of Calgary
Get access

Summary

Introduction

Landscape can be broadly defined as a land area that comprises a physical (abiotic) topography component with an organized structure of hillslopes, valleys, and channels, and a biotic component including microorganisms, plants, animals, and humans that inhabit and influence this topographic platform. These landscape components are inherently coupled across a range of space and time scales. From the perspective of thermodynamic principles, landscapes can be viewed as open dissipative systems, with abiotic and biotic components exchanging mass, energy, and momentum within the basin template, co-organizing into dynamic and interacting structures (Fisher et al., 2007; Muneepeerakul et al., 2008; Rodríguez-Iturbe et al., 2009; del Jesus et al., 2012). Components of the system undertake distinctly different roles. Terrestrial ecosystems dictate the local land surface energy, water, and nutrient balance, and influence geomorphic processes (GP). The flows of water, sediments, nutrients, and organisms across the landscape are regulated by the network topology of a river basin, and in turn, they act as boundary conditions for the evolution of the river basin itself and its characteristic ecosystem (Murray et al., 2008; del Jesus et al., 2012). Complex interactions between abiotic and biotic landscape components have long been qualitatively recognized among scientists, however, such interactions remain largely unexplored quantitatively, as they lie in the interface among different disciplines such as biology, ecology, hydrology, and geology. An interdisciplinary research area, commonly referred to as “ecogeomorphology,” has emerged in the last couple of decades to study the coupled physical and biological processes of landscapes, and identify the emergent properties of the coupled system (Fisher et al., 2007; Murray et al., 2008; Reinhardt et al., 2010; Saco and Rodríguez, 2013).

Ecologists and hydrologists have researched to understand how the biota responds to the physical environment. Particular examples include the observed ecosystem differences in relation to landscape morphology at catchment scale (Hack and Goodlett, 1960; Swanson et al., 1988; Florinsky and Kuryakova, 1996; Ivanov et al., 2008b; Svoray and Karnieli, 2010; Jost et al., 2012; Flores-Cervantes et al., 2014), and the organization of ecosystems across regional climate and elevation gradients (Coblentz and Riitters, 2004; Caylor et al., 2005; Hillyer and Silman, 2010; Malhi et al., 2010).

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

Abrahams, A. D. (1984), Channel networks: a geomorphological perspective, Water Resources Research, 20(2), 161–88.Google Scholar
Abrahams, A. D., Parsons, A. J. and Wainwright, J. (1995). Effects of vegetation change in interrill runoff and erosion, Walnut Gulch, southern Arizona, Geomorphology, 13(1–4), 37–48.Google Scholar
Anderson, R. S., Anderson, S. P. and Tucker, G. E. (2013). Rock damage and regolith transport by frost: an example of climate modulation of the geomorphology of the critical zone. Earth Surface Processes and Landforms, 38, 299–316.Google Scholar
Ardizzone, F., Cardinali, M., Galli, M., Guzzetti, F. and Reichenbach, P. (2007). Identification and mapping of recent rainfall-induced landslides using elevation data collected by airborne Lidar. Natural Hazards and Earth System Sciences, 7, 637–50.Google Scholar
Arnone, E., Noto, L., Lepore, C. and Bras, R. (2011). Physically-based and distributed approach to analyze rainfall-triggered landslides at watershed scale. Geomorphology, 133, 121–31.Google Scholar
Baas, A. C. W. (2007). Complex systems in eolian geomorphology. Geomorphology, 91, 311–31.Google Scholar
Baas, A. C. W. and Nield, J. M. (2007). Modelling vegetated dune landscapes. Geophysical Research Letters 34, L06405.Google Scholar
Band, L. E., Hwang, T., Hales, T. C., Vose, J. and Ford, C. (2012). Ecosystem processes at the watershed scale: mapping and modeling ecohydrological controls of landslides. Geomorphology, 137, 159–67.Google Scholar
Beaumont, C., Fullsack, P. and Hamilton, J. (1992). Erosional control of active compressional orogens. In Thrust Tectonics, ed. McClay, K. R.. New York: Chapman and Hall, 1–18.
Benda, L. and Dunne, T. (1997). Stochastic forcing of sediment supply to channel networks from landsliding and debris flow. Water Resources Research, 33, 2849–63.Google Scholar
Bertoldi, G., Rigon, R. and Over, T. M. (2006). Impact of watershed geomorphic characteristics on the energy and water budgets. Journal of Hydrometeorology, 7, 389–403.Google Scholar
Black, T. A. and Montgomery, D. R. (1991). Sediment transport by burrowing mammals, Marin County, California. Earth Surface Processes and Landforms, 16, 163–72.Google Scholar
Bogaart, P. W., Tucker, G. E. and de Vries, J. J. (2003). Channel network morphology and sediment dynamics under alternating periglacial and temperate regimes: a numerical simulation study. Geomorphology, 54, 257–77.Google Scholar
Brandon, F. A., and Shown, L. M. (1990). Contrasts of Vegetation, Soils Microclimates, and Geomorphic Processes between North and South Facing Slopes on Green Mountain near Denver, Colorado. Water Resources Investigations Report 89–4094. Denver: US Department of the Interior, US Geological Survey.
Braun, J., Heimsath, A. M. and Chappell, J. (2001). Sediment transport mechanisms on soil-mantled hillslopes. Geology, 29, 683–6.Google Scholar
Brown, D. E. 1994. Biotic Communities: Southwestern United States and Northwestern Mexico. Salt Lake City: University of Utah Press.
Burnett, B. N., Meyer, G. A. and McFadden, L. D. (2008). Aspect-related microclimatic influences on slope forms and processes, northeastern Arizona. Journal of Geophysical Research-Earth Surface, 113, F03002.Google Scholar
Cantlon, J. E. (1953). Vegetation and microclimates on north and south slopes of Cushetunk Mountain, New Jersey. Ecological Monographs, 23, 241–70.Google Scholar
Carson, M. A. and Kirkby, M. J. (1972). Hillslope Form and Process. Cambridge UK: Cambridge University Press, 475.
Caylor, K. K., Manfreda, S. and Rodriguez-Iturbe, I. (2005). On the coupled geomorphological and ecohydrological organization of river basins. Advances in Water Resources, 28, 69–86.Google Scholar
Caylor, K. K., Scanlon, T. M. and Rodriguez-Iturbe, I. (2004). Feasible optimality of vegetation patterns in river basins. Geophysical Research Letters, 31, L13502.Google Scholar
Choi, M. and Jacobs, J. M. (2007). Soil moisture variability of root zone profiles within SMEX02 remote sensing footprints. Advances in Water Resources, 30, 883–96.Google Scholar
Coblentz, D. and Keating, P. L. (2008). Topographic controls on the distribution of tree islands in the high Andes of south-western Ecuador. Journal of Biogeography, 35, 2026–2038.Google Scholar
Coblentz, D. D. and Riitters, K. H. (2004). Topographic controls on the regional-scale biodiversity of the south-western USA. Journal of Biogeography, 31, 1125–38.Google Scholar
Collins, D. B. G. and Bras, R. L. (2008). Climatic control of sediment yield in dry lands following climate and land cover change. Water Resources Research, 44, W10405.Google Scholar
Collins, D. B. G. and Bras, R. L. (2010). Climatic and ecological controls of equilibrium drainage density, relief, and channel concavity in dry lands. Water Resources Research, 46, W04508.Google Scholar
Collins, D. B. G., Bras, R. L. and Tucker, G. E. (2004). Modeling the effects of vegetation-erosion coupling on landscape evolution. Journal of Geophysical Research-Earth Surface, 109, F03004.Google Scholar
Corenblit, D., Baas, A. C. W., Bornette, G. et al. (2011). Feedbacks between geomorphology and biota controlling Earth surface processes and landforms: a review of foundation concepts and current understandings. Earth-Science Reviews 106 (3–4), 307–31.Google Scholar
Cottle, H. J. (1932). Vegetation on north and south slopes of mountains in southwestern Texas. Ecology, 13, 121–34.Google Scholar
Coulthard, T. J. (2001). Landscape evolution models: a software review. Hydrological Processes 15, 165, doi:10.1002/hyp.426.Google Scholar
Cristea, N. C., Kampf, S. K. and Burges, S. J. (2013). Linear models for estimating annual and growing season reference evapotranspiration using averages of weather variables. International Journal of Climatology, 33, 376–87.Google Scholar
del Jesus, M., Foti, R., Rinaldo, A. and Rodriguez-Iturbe, I. (2012). Maximum entropy production, carbon assimilation, and the spatial organization of vegetation in river basins. Proceedings of the National Academy of Sciences of the United States of America, 109, 20837–41.Google Scholar
Densmore, A. L., Ellis, M. A. and Anderson, R. S. (1998). Landsliding and the evolution of normal-fault-bounded mountains. Journal of Geophysical Research-Solid Earth, 103, 15203–19.Google Scholar
Desta, F., Colbert, J. J., Rentch, J. S. and Gottschalk, K. W. (2004). Aspect induced differences in vegetation, soil, and microclimatic characteristics of an Appalachian watershed. Castanea, 69, 92–108.Google Scholar
DiBiase, R. A., Whipple, K. X., Heimsath, A. M. and Ouimet, W. B. (2010). Landscape form and millennial erosion rates in the San Gabriel Mountains, CA. Earth and Planetary Science Letters, 289, 134–44.Google Scholar
Dick-Peddie, W. A. 1999. New Mexico Vegetation: Past, Present, and Future. Albuquerque, NM: University of New Mexico Press.
Dietrich, W. E., Bellugi, D. G., Sklar, L. S. et al. (2003). Geomorphic transport laws for predicting landscape form and dynamics. In Prediction in Geomorphology, ed. Wilcock, P. R. and Iverson, R. M.. Washington, DC: American Geophysical Union, 103–32.
Dietrich, W. E. and Perron, J. T. (2006). The search for a topographic signature of life. Nature, 439, 411–8.Google Scholar
Dingman, S. L. 2002. Physical Hydrology,. Upper Saddle River, NJ: Prentice Hall.
Dunne, T. and Black, R. D. (1970). Partial area contributions to storm runoff in a small New England watershed. Water Resources Research, 6, 1296–1311.Google Scholar
Dunne, T., Moore, T. and Taylor, C. (1975). Recognition and prediction of runoff-producing zones in humid regions. Hydrological Sciences Sciences–Bulletin, 20, 305–27.Google Scholar
Eagleson, P. S. 2002. Ecohydrology: Darwinian Expression of Vegetation Form and Function, Cambridge: Cambridge University Press.
Ellis, M. A., Densmore, A. L. and Anderson, R. S. (1999). Development of mountainous topography in the Basin Ranges, USA. Basin Research, 11, 21–41.Google Scholar
Famiglietti, J. S., Ryu, D. R., Berg, A. A., Rodell, M. and Jackson, T. J. (2008). Field observations of soil moisture variability across scales. Water Resources Research, 44, W01423.Google Scholar
Fatichi, S., Ivanov, V. Y. and Caporali, E. (2012a). A mechanistic ecohydrological model to investigate complex interactions in cold and warm water-controlled environments: 1. Theoretical framework and plot-scale analysis. Journal of Advances in Modeling Earth Systems, 4, M05002.Google Scholar
Fatichi, S., Ivanov, V. Y. and Caporali, E. (2012b). A mechanistic ecohydrological model to investigate complex interactions in cold and warm water-controlled environments: 2. Spatiotemporal analyses. Journal of Advances in Modeling Earth Systems, 4, M05003.Google Scholar
Fekedulegn, D., Hicks, R. R. and Colbert, J. J. (2003). Influence of topographic aspect, precipitation and drought on radial growth of four major tree species in an Appalachian watershed. Forest Ecology and Management, 177, 409–25.Google Scholar
Flores Cervantes, J. H., Istanbulluoglu, E., Vivoni, E. R., Holifield Collins, C. D. and Bras, R. L. (2014). A geomorphic perspective on terrain-modulated organization of vegetation productivity: analysis in two semiarid grassland ecosystems in Southwestern United States. Ecohydrology, 7, 242–57.Google Scholar
Fisher, S. G., Heffernan, J. B., Sponseller, R. A. and Welter, J. R. (2007). Functional ecomorphology: feedbacks between form and function in fluvial landscape ecosystems. Geomorphology, 89, 84–96.Google Scholar
Florinsky, I. V. and Kuryakova, G. A. (1996). Influence of topography on some vegetation cover properties. Catena, 27, 123–41.Google Scholar
Forzieri, G., Castelli, F. and Vivoni, E. R. (2011). Vegetation dynamics within the North American monsoon region. Journal of Climate, 24, 1763–83.Google Scholar
Foster, G. (1982). Modeling the erosion process. In Hydrologic Modeling of Small Watersheds, ed. Haan, C. T., Johson, H. P. and Brakensiek, D. L.. St. Joseph, MI: American Society of Agricultural Engineers, 295–380.
Franz, T. E., Caylor, K. K., Nordbotten, J. M., Rodriguez-Iturbe, I. and Celia, M. A. (2010). An ecohydrological approach to predicting regional woody species distribution patterns in dryland ecosystems. Advances in Water Resources, 33, 215–30.Google Scholar
Gabet, E. J. (2003). Sediment transport by dry ravel. Journal of Geophysical Research-Solid Earth, 108, doi:10.1029/2001JB001686.Google Scholar
Gabet, E. J., and Dunne, T. (2003b). A stochastic sediment delivery model for a steep Mediterranean landscape. Water Resources Research, 39(9), 1237, doi:10.1029/2003WR002341.Google Scholar
Gabet, E. J. and Mudd, S. M. (2010). Bedrock erosion by root fracture and tree throw: a coupled biogeomorphic model to explore the humped soil production function and the persistence of hillslope soils, Journal of Geophysical Research, 115, F04005, doi:10.1029/2009JF001526.Google Scholar
Gabet, E. J., Reichman, O. J. and Seabloom, E. W. (2003). The effects of bioturbation on soil processes and sediment transport. Annual Review of Earth and Planetary Sciences, 31, 249–73.Google Scholar
Gabet, E. J., Perron, J. T. and Johnson, D. L. (2014). Biotic origin for Mima mounds supported by numerical modeling. Geomorphology, 206, 58–66, doi:10.1016/j.geomorph.2013.09.018.Google Scholar
Garde, R. J. and Raju, K. R. (1985). Mechanics of Sediment Transportation and Alluvial Stream Problems,. New Delhi: Wiley Eastern Ltd.
Gasparini, N. M., Tucker, G. E. and Bras, R. L. (1999). Downstream fining through selective particle sorting in an equilibrium drainage network. Geology, 27, 1079–82.Google Scholar
Gasparini, N. M., Tucker, G. E. and Bras, R. L. (2004). Network-scale dynamics of grain-size sorting: implications for downstream fining, stream-profile concavity, and drainage basin morphology. Earth Surface Processes and Landforms, 29, 401–21.Google Scholar
Gasparini, N. M., Whipple, K. X. and Bras, R. L. (2007). Predictions of steady state and transient landscape morphology using sediment-flux-dependent river incision models. Journal of Geophysical Research-Earth Surface, 112, F03S09.Google Scholar
Gebremichael, M., Vivoni, E. R., Watts, C. J. and Rodriguez, J. C. (2007). Submesoscale spatiotemporal variability of North American monsoon rainfall over complex terrain. Journal of Climate, 20, 1751–73.Google Scholar
Gochis, D. J., Watts, C. J., Garatuza-Payàn, J. and Rodrìguez, J. C. (2007). Spatial and temporal patterns of precipitation intensity as observed by the NAME event rain gauge network from 2002 to 2004. Journal of Climate, 20, 1734–50.Google Scholar
Govers, G. 1992. Evaluation of transporting capacity formulae for overland flow conditions. In Overland Flow, Hydraulics and Erosion Mechanics, ed. Parsons, A. J. and Abrahams, A. D.. London: UCL Press, 243–273.
Graf, W. L. (1979). The development of montane arroyos and gullies. Earth Surface Processes and Landforms, 4, 1–14.Google Scholar
Gregory, K. J. (1976), Drainage networks and climate. In Geomorphology and Climate, ed. Derbyshire, E.. London: John Wiley & Sons, 289–315.
Gutiérrez-Jurado, H. A. and Vivoni, E. R. (2013). Ecogeomorphic expressions of an aspect-controlled semiarid basin: I. Topographic analyses with high-resolution data sets. Ecohydrology , 6, 8–23.Google Scholar
Gutiérrez-Jurado, H. A., Vivoni, E. R., Istanbulluoglu, E. and Bras, R. L. (2007). Ecohydrological response to a geomorphically significant flood event in a semiarid catchment with contrasting ecosystems. Geophysical Research Letters, 34, L24S25.Google Scholar
Gutiérrez-Jurado, H. A., Vivoni, E. R., Cikoski, C. et al. (2013). On the observed ecohydrologic dynamics of a semiarid basin with aspect-delimited ecosystems. Water Resources Research, 49, 8263–84.Google Scholar
Hack, J. T. and Goodlett, J. C. (1960). Geomorphology and Forest Ecology of a Mountain Region in the Central Appalachians. Washington, DC: US Government Printing Office.
Hancock, G. R., Willgoose, G. R. and Evans, K. G. (2002). Testing of the SIBERIA landscape evolution model using the Tin Camp Creek, Northern Territory, Australia, field catchment. Earth Surface Processes and Landforms, 27, 125–43.Google Scholar
Hanks, T. C. (2000). The age of scarplike landforms from diffusion-equation analysis. In Quaternary Geochronology: Methods and Applications, ed. Noller, J. S., Sowers, J. M. and Lettis, W. R.. Washington DC: American Geophysical Union, 313–38.
Heimsath, A. M., Dietrich, W. E., Nishiizumi, K. and Finkel, R. C. (1999). Cosmogenic nuclides, topography, and the spatial variation of soil depth. Geomorphology, 27, 151–72.Google Scholar
Hillyer, R. and Silman, M. R. (2010). Changes in species interactions across a 2.5 km elevation gradient: effects on plant migration in response to climate change. Global Change Biology, 16, 3205–14 Google Scholar
Holland, P. G. and Steyn, D. G. (1975). Vegetational responses to latitudinal variations in slope angle and aspect. Journal of Biogeography, 2, 179–83.Google Scholar
Howard, A. D. (1980). Thresholds in river regimes. In Thresholds in Geomorphology, ed. Coates, D. R. and Vitek, J. D.. London: Allen and Unwin, 227–58.
Howard, A. D. (1994). A detachment-limited model of drainage basin evolution. Water Resources Research, 30, 2261–85.Google Scholar
Huang, X. and Niemann, J. D. (2006). Modeling the potential impacts of groundwater hydrology on long-term landscape evolution. Earth Surface Processes and Landforms, 31(14), 1802–23.Google Scholar
Hughes, M. W., Almond, P. C. and Roering, J. J. (2009). Increased sediment transport via bioturbation at the last glacial-interglacial transition. Geology, 37, 919–22.Google Scholar
Hurst, M. D., Mudd, S. M., Yoo, K., Attal, M. and Walcott, R. (2013). Influence of lithology on hillslope morphology and response to tectonic forcing in the northern Sierra Nevada of California. Journal of Geophysical Research-Earth Surface, 118, 832–51.Google Scholar
Huxman, T. E., Smith, M. D., Fay, P. A. et al. (2004). Convergence across biomes to a common rain-use efficiency. Nature, 429, 651–4.Google Scholar
Hwang, T., Band, L. and Hales, T. C. (2009). Ecosystem processes at the watershed scale: extending optimality theory from plot to catchment. Water Resources Research, 45, W11425.Google Scholar
Istanbulluoglu, E. and Bras, R. L. (2005). Vegetation-modulated landscape evolution: effects of vegetation on landscape processes, drainage density, and topography. Journal of Geophysical Research-Earth Surface, 110, F02012.Google Scholar
Istanbulluoglu, E. and Bras, R. L. (2006). On the dynamics of soil moisture, vegetation, and erosion: implications of climate variability and change. Water Resources Research, 42, W06418.Google Scholar
Istanbulluoglu, E., Tarboton, D. G., Pack, R. T. and Luce, C. H. (2003). A sediment transport model for incision of gullies on steep topography. Water Resources Research, 39, 1103.Google Scholar
Istanbulluoglu, E., Tarboton, D. G., Pack, R. T. and Luce, C. H. (2004). Modeling of the interactions between forest vegetation, disturbances, and sediment yields. Journal of Geophysical Research-Earth Surface, 109, F01009.Google Scholar
Istanbulluoglu, E., Yetemen, O., Vivoni, E. R., Gutiérrez-Jurado, H. A. and Bras, R. L. (2008). Eco-geomorphic implications of hillslope aspect: inferences from analysis of landscape morphology in central New Mexico. Geophysical Research Letters, 35, L14403.Google Scholar
Ivanov, V. Y., Bras, R. L. and Vivoni, E. R. (2008a). Vegetation-hydrology dynamics in complex terrain of semiarid areas: 1. A mechanistic approach to modeling dynamic feedbacks. Water Resources Research, 44, W03429.Google Scholar
Ivanov, V. Y., Bras, R. L. and Vivoni, E. R. (2008b). Vegetation-hydrology dynamics in complex terrain of semiarid areas: 2. Energy-water controls of vegetation spatiotemporal dynamics and topographic niches of favorability. Water Resources Research, 44, W03430.Google Scholar
Ivanov, V. Y., Fatichi, S., Jenerette, G. D. et al. (2010). Hysteresis of soil moisture spatial heterogeneity and the “homogenizing” effect of vegetation. Water Resources Research, 46, W09521.Google Scholar
Jencso, K. G. and McGlynn, B. L. (2011). Hierarchical controls on runoff generation: topographically driven hydrologic connectivity, geology, and vegetation. Water Resources Research, 47, W11527.Google Scholar
Jencso, K. G., McGlynn, B. L., Gooseff, M. N. et al. (2009). Hydrologic connectivity between landscapes and streams: transferring reach-and plot-scale understanding to the catchment scale. Water Resources Research, 45, W04428.Google Scholar
Jost, G., Schume, H., Hager, H., Markart, G. and Kohl, B. (2012). A hillslope scale comparison of tree species influence on soil moisture dynamics and runoff processes during intense rainfall. Journal of Hydrology 420–421, 112–24 Google Scholar
Laflen, J. M., Elliot, W. J., Simanton, J. R., Holzhey, C. S. and Kohl, K. D. (1991). WEPP: soil erodibility experiments for rangeland and cropland soils. Journal of Soil and Water Conservation, 46, 39–44.Google Scholar
Laio, F., Porporato, A., Ridolfi, L. and Rodriguez-Iturbe, I. (2001). Plants in water-controlled ecosystems: active role in hydrologic processes and response to water stress II. Probabilistic soil moisture dynamics. Advances in Water Resources, 24, 707–23.Google Scholar
Laland, K. N., Odling-Smee, F. J. and Feldman, M. W. (1999). Evolutionary consequences of niche construction and their implications for ecology. Proceedings of the National Academy of Sciences of the United States of America 96, 10242–7.Google Scholar
Lancaster, S. T., Hayes, S. K. and Grant, G. E. (2003). Effects of wood on debris flow runout in small mountain watersheds. Water Resources Research, 39, doi:10.1029/2001WR001227.Google Scholar
Langbein, W. B. and Schumm, S. (1958). Yield of sediment in relation to mean annual precipitation. Eos, Transactions American Geophysical Union, 39, 1076–84.Google Scholar
Lanni, C., Borga, M., Rigon, R. and Tarolli, P. (2012). Modelling shallow landslide susceptibility by means of a subsurface flow path connectivity index and estimates of soil depth spatial distribution. Hydrology and Earth System Sciences, 16, 3959–71.Google Scholar
Lauenroth, W. K. and Bradford, J. B. (2009), Ecohydrology of dry regions of the United States: precipitation pulses and intraseasonal drought, Ecohydrology, 2, 173–81, doi:10.1002/Eco.53.Google Scholar
Lepore, C., Arnone, E., Noto, L. V., Sivandra, G. and Bras, R. L. (2013). Physically based modeling of rainfall-triggered landslides: a case study in the Luquillo forest, Puerto Rico. Hydrology and Earth System Sciences, 17, 3371–87.Google Scholar
Luoto, M. (2007). New insights into factors controlling drainage density in subarctic landscapes. Arctic, Antarctic, and Alpine Research, 39(1), 117–26.Google Scholar
Mahmood, T. H. and Vivoni, E. R. (2011). A climate-induced threshold in hydrologic response in a semiarid ponderosa pine hillslope. Water Resources Research, 47, W09529.Google Scholar
Malhi, Y., Silman, M., Salin, N. et al. (2010). Introduction: elevation gradients in the tropics: laboratories for ecosystem ecology and global change research. Global Change Biology 16, 3171–5, doi:10.1111/j.1365–2486.2010.02323.x.Google Scholar
Martin, Y. (2000). Modelling hillslope evolution: linear and nonlinear transport relations. Geomorphology, 34, 1–21.Google Scholar
Mascaro, G., Vivoni, E. R. and Deidda, R. (2011). Soil moisture downscaling across climate regions and its emergent properties. Journal of Geophysical Research-Atmospheres, 116, D22114.Google Scholar
McGuire, L. A., Pelletier, J. D. and Roering, J. J. (2014). Development of topographic asymmetry: insights from dated cinder cones in the western United States. Journal of Geophysical Research: Earth Surface, 2014JF003081.
Melton, M. (1957). An Analysis of the Relations among Elements of Climate, Surface Properties and Geomorphology. Technical Report (Columbia University, Department of Geology) No. 11. New York: Department of Geology, Columbia University.
Minder, J. R., Mote, P. W. and Lundquist, J. D. (2010). Surface temperature lapse rates over complex terrain: lessons from the Cascade Mountains. Journal of Geophysical Research-Atmospheres, 115, D14122.Google Scholar
Moglen, G. E. and Bras, R. L. (1995). The importance of spatially heterogeneous erosivity and the cumulative area distribution within a basin evolution model. Geomorphology, 12, 173–85.Google Scholar
Montaldo, N., Corona, R. and Albertson, J. D. (2013). On the separate effects of soil and land cover on Mediterranean ecohydrology: two contrasting case studies in Sardinia, Italy. Water Resources Research, 49, 1123–36.Google Scholar
Moglen, G. E., Eltahir, E. A. B. and Bras, R. L. (1998), On the sensitivity of drainage density to climate change, Water Resources Research, 34, 855–62, doi:10.1029/97WR02709.Google Scholar
Montaldo, N., Rondena, R., Albertson, J. D. and Mancini, M. (2005). Parsimonious modeling of vegetation dynamics for ecohydrologic studies of water-limited ecosystems. Water Resources Research, 41, W10416.Google Scholar
Montgomery, D. R. (1994). Road surface drainage, channel initiation, and slope instability. Water Resources Research, 30, 1925–32.Google Scholar
Montgomery, D. R. and Dietrich, W. E. (1994). A physically-based model for the topographic control on shallow landsliding, Water Resources Research, 30, 1153–71.Google Scholar
Montgomery, D. R., Dietrich, W. E. and Heffner, J. T. (2002). Piezometric response in shallow bedrock at CB1: implications for runoff generation and landsliding. Water Resources Research, 38, doi:10.1029/2002WR001429.Google Scholar
Montgomery, D. R., Schmidt, K. M., Greenberg, H. M. and Dietrich, W. E. (2000). Forest clearing and regional landsliding. Geology, 28, 311–4.Google Scholar
Muneepeerakul, R., Rinaldo, A. and Rodriguez-Iturbe, I. (2008). Patterns of vegetation biodiversity: the roles of dispersal directionality and network structure, Journal of Theoretical Biology, 252, 221–9, doi:10.1016/j.jtbi.2008.02.001.Google Scholar
Murray, A. B., Knaapen, M. A. F., Tal, M. and Kirwan, M. L. (2008). Biomorphodynamics: physical-biological feedbacks that shape landscapes. Water Resources Research, 44, W11301, doi:10.1029/2007WR006410.Google Scholar
Nash, D. (1980). Forms of bluffs degraded for different lengths of time in Emmet County, Michigan, USA. Earth Surface Processes and Landforms, 5, 331–45.Google Scholar
Nearing, M. A., Simanton, J. R., Norton, L. D., Bulygin, S. J. and Stone, J. (1999). Soil erosion by surface water flow on a stony, semiarid hillslope. Earth Surface Processes and Landforms, 24, 677–86.Google Scholar
Nepf, H. M. (1999). Drag, turbulence, and diffusion in flow through emergent vegetation. Water Resources Research, 35, 479–89.Google Scholar
Nepf, H. M. and Vivoni, E. R. (2000). Flow structure in depth-limited, vegetated flow. Journal of Geophysical Research-Oceans, 105, 28547–57.Google Scholar
Nyman, P., Sheridan, G. J. and Lane, P. N. J. (2013). Hydro-geomorphic response models for burned areas and their applications in land management. Progress in Physical Geography, 37, 787–812, doi:10.1177/0309133313508802.Google Scholar
Odling-Smee, F. J., Laland, K. N. and Feldman, M. W. (2003). Niche Construction: A Neglected Process in Evolution. Princeton, NJ: Princeton University Press.
Pelletier, J. D., Barron-Gafford, G. A., Breshears, D. D. et al. (2013). Coevolution of nonlinear trends in vegetation, soils, and topography with elevation and slope aspect: a case study in the sky islands of southern Arizona. Journal of Geophysical Research-Earth Surface, 118, 741–58.Google Scholar
Perron, J. T., Kirchner, J. W. and Dietrich, W. E. (2009). Formation of evenly spaced ridges and valleys. Nature, 460, 502–5.Google Scholar
Poulos, M. J., Pierce, J. L., Flores, A. N. and Benner, S. G. (2012). Hillslope asymmetry maps reveal widespread, multi-scale organization. Geophysical Research Letters, 39, L06406, doi:10.1029/2012GL051283.Google Scholar
Prosser, I. P., Dietrich, W. E. and Stevenson, J. (1995). Flow resistance and sediment transport by concentrated overland flow in a grassland valley. Geomorphology, 13, 71–86.Google Scholar
Rasmussen, C. and Tabor, N. (2007). Application of a quantitative pedogenic energy model to predict soil development across a range of environmental gradients. Soil Science Society of America Journal, 71(5), 1719–29.Google Scholar
Rasmussen, C., Troch, P. A., Chorover, J. et al. (2011). An open system framework for integrating critical zone structure and function. Biogeochemistry, 102, 15–29.Google Scholar
Reinhardt, L., Jerolmack, D., Cardinale, B. J., Vanacker, V. and Wright, J. (2010). Dynamic interactions of life and its landscape: feedbacks at the interface of geomorphology and ecology. Earth Surface Processes and Landforms, 35, 78–101.Google Scholar
Riitters, K. H., Wickham, J. D., Vogelmann, J. E. and Jones, K. B. (2000). National land-cover pattern data. Ecology, 81, 604.Google Scholar
Riveros-Iregui, D. A. and McGlynn, B. L. (2009). Landscape structure control on soil CO2 efflux variability in complex terrain: scaling from point observations to watershed scale fluxes. Journal of Geophysical Research-Biogeosciences, 114, G02010.Google Scholar
Robichaud, P. R., Wagenbrenner, J. W. and Brown, R. E. (2010). Rill erosion in natural and disturbed forests: 1. Measurements. Water Resources Research, 46, W10506.Google Scholar
Rodríguez-Iturbe, I. and Porporato, A. (2004). Ecohydrology of Water-controlled Ecosystems. Cambridge: Cambridge University Press.
Rodríguez-Iturbe, I., Muneepeerakul, R., Bertuzzo, E., Levin, S. A. and Rinaldo, A. (2009). River networks as ecological corridors: a complex systems perspective for integrating hydrologic, geomorphologic, and ecologic dynamics. Water Resources Research, 45, W01413, doi:10.1029/2008WR007124.Google Scholar
Roering, J. J. (2008). How well can hillslope evolution models “explain” topography? Simulating soil transport and production with high-resolution topographic data. Geological Society of America Bulletin, 120, 1248–62.Google Scholar
Roering, J. J., Almond, P., Tonkin, P. and McKean, J. (2002). Soil transport driven by biological processes over millennial time scales. Geology, 30, 1115–8.Google Scholar
Roering, J. J., Almond, P., Tonkin, P. and McKean, J. (2004). Constraining climatic controls on hillslope dynamics using a coupled model for the transport of soil and tracers: application to loess-mantled hillslopes, South Island, New Zealand. Journal of Geophysical Research-Earth Surface, 109, F01010.Google Scholar
Roering, J. J. and Gerber, M. (2005). Fire and the evolution of steep, soil-mantled landscapes. Geology, 33, 349–52.Google Scholar
Roering, J. J., Kirchner, J. W. and Dietrich, W. E. (1999). Evidence for nonlinear, diffusive sediment transport on hillslopes and implications for landscape morphology. Water Resources Research, 35, 853–70.Google Scholar
Roering, J. J., Marshall, J. A., Booth, A., Mort, M. and Jin, Q. (2010). Evidence for biotic controls on topography and soil production. Earth and Planetary Science Letters, 298, 183–90.Google Scholar
Roering, J. J., Schmidt, K. M., Stock, J. D., Dietrich, W. E. and Montgomery, D. R. (2003). Shallow landsliding, root reinforcement, and the spatial distribution of trees in the Oregon Coast Range. Canadian Geotechnical Journal, 40, 237–53.Google Scholar
Saco, P. M. and Rodríguez, J. F. ( 2013 ). Modeling ecogeomorphic systems. In Treatise on Geomorphology, Volume 2, ed. Shroder, J. F.. San Diego, CA: Academic Press, 201–20.
Saco, P. M., Willgoose, G. R. and Hancock, G. R. (2007). Eco-geomorphology of banded vegetation patterns in arid and semi-arid regions. Hydrology and Earth System Sciences, 11, 1717–30.Google Scholar
Sakals, M. E., and Sidle, R. C. (2004). A spatial and temporal model of root strength in forest soils. Canadian Journal of Forest Research, 34, 950–8.Google Scholar
Schmidt, K. M., Roering, J. J., Stock, J. D. et al. (2001). The variability of root cohesion as an influence on shallow landslide susceptibility in the Oregon Coast Range. Canadian Geotechnical Journal, 38, 995–1024.Google Scholar
Selby, M. J. (1993). Hillslope Materials and Processes,. Oxford: Oxford University Press.
Sólyom, P. B. and Tucker, G. E. (2004). Effect of limited storm duration on landscape evolution, drainage basin geometry, and hydrograph shapes. Journal of Geophysical Research-Earth Surface, 109, F03012.Google Scholar
Stock, J. and Dietrich, W. E. (2003). Valley incision by debris flows: evidence of a topographic signature. Water Resources Research, 39, doi:10.1029/2001WR001057.Google Scholar
Stock, J. D. and Montgomery, D. R. (1999). Geologic constraints on bedrock river incision using the stream power law. Journal of Geophysical Research-Solid Earth, 104, 4983–93.Google Scholar
Summerfield, M. A. and Hulton, N. J. (1994). Natural controls of fluvial denudation rates in major world drainage basins. Journal of Geophysical Research-Solid Earth, 99(B7), 13871–83, doi:10.1029/94jb00715.Google Scholar
Svoray, T. and Karnieli, A. (2010). Rainfall, topography and primary production relationships in a semiarid ecosystem. Ecohydrology, 4, 56–66.Google Scholar
Swanson, F. J., Kratz, T. K., Caine, N. and Woodmansee, R. G. (1988). Landform effects on ecosystem patterns and processes. BioScience 38(2), 92–8.Google Scholar
Tague, C., Heyn, K. and Christensen, L. (2009). Topographic controls on spatial patterns of conifer transpiration and net primary productivity under climate warming in mountain ecosystems. Ecohydrology, 2, 541–54.Google Scholar
Tang, Z. Y. and Fang, J. Y. (2006). Temperature variation along the northern and southern slopes of Mt. Taibai, China. Agricultural and Forest Meteorology, 139, 200–7.Google Scholar
Tilman, D. (1994). Competition and biodiversity in spatially structured habitats. Ecology 75, 2–16.Google Scholar
Thompson, R., Whitlock, C., Bartlein, P., Harrison, S. and Spaulding, W. (1993). Climatic changes in the western United States since 18,000 yr BP. In Global Climates since the Last Glacial Maximum, ed. Wright, H. E. et al. St. Paul, MN: University of Minnesota Press, 468–513.
Thompson, S. E., Harman, C. J., Troch, P. A., Brooks, P. D. and Sivapalan, M. (2011). Spatial scale dependence of ecohydrologically mediated water balance partitioning: a synthesis framework for catchment ecohydrology. Water Resources Research, 47, W00J03.Google Scholar
Tromp-van Meerveld, H. J. and McDonnell, J. J. (2006). Threshold relations in subsurface stormflow: 1. A 147-storm analysis of the Panola hillslope. Water Resources Research, 42, W02410.Google Scholar
Tucker, G. E. and Bras, R. L. (1998). Hillslope processes, drainage density, and landscape morphology. Water Resources Research, 34, 2751–64.Google Scholar
Tucker, G. E. and Bras, R. L. (1999). Dynamics of vegetation and runoff erosion. In A 3D Computer Simulation Model of Drainage Basin and Floodplain Evolution: Theory and Applications. Technical Report to US Army Corps of Engineers Construction Engineering Research Lab. Cambridge, MA: Department of Civil and Environmental Engineering.
Tucker, G. E. and Bras, R. L. (2000). A stochastic approach to modeling the role of rainfall variability in drainage basin evolution. Water Resources Research, 36, 1953–64.Google Scholar
Tucker, G. E. and Whipple, K. X. (2002). Topographic outcomes predicted by stream erosion models: sensitivity analysis and intermodel comparison. Journal of Geophysical Research-Solid Earth, 107, doi:10.1029/2001JB000162 2179.Google Scholar
Tucker, G. E. and Hancock, G. R. (2010). Modelling landscape evolution. Earth Surface Processes and Landforms, 35, 28–50.Google Scholar
Tucker, G. E., Lancaster, S. T., Gasparini, N. M. and Bras, R. L. (2001). The Channel-Hillslope Integrated Landscape Development model (CHILD). In Landscape Erosion and Evolution Modeling, ed. Harmon, R. S. and Doe, W. W.. New York: Kluwer Academic/Plenum Publishers, 349–88.
Vanwalleghem, T., Stockmann, U., Minasny, B. and McBratney, A. B. (2013). A quantitative model for integrating landscape evolution and soil formation. Journal of Geophysical Research-Earth Surface, 118, 331–47, doi:10.1029/2011JF002296.Google Scholar
Viles, H.A. (1988). Biogeomorphology. Oxford: Blackwell.
Viles, H. A., Naylor, L. A., Carter, N. E. A. and Chaput, D. (2008). Biogeomorphological disturbance regimes: progress in linking ecological and geomorphological systems. Earth Surface Processes and Landforms 33, 1419–35.Google Scholar
Vivoni, E. R., Tai, K. and Gochis, D. J. (2009). Effects of initial soil moisture on rainfall generation and subsequent hydrologic response during the North American Monsoon. Journal of Hydrometeorology, 10, 644–64.Google Scholar
Walling, D. E. and Webb, B. W. (1983). Patterns of sediment yield. In Background to Palaeohydrology, ed. Gregory, K. J.. Chichester, UK: John Wiley, 69–100.
Wagenbrenner, J. W., Robichaud, P. R. and Elliot, W. J. (2010). Rill erosion in natural and disturbed forests: 2. Modeling approaches. Water Resources Research, 46, W10507.Google Scholar
Warren, R. J. (2010). An experimental test of well-described vegetation patterns across slope aspects using woodland herb transplants and manipulated abiotic drivers. New Phytologist, 185, 1038–49.Google Scholar
West, N., Kirby, E., Bierman, P. and Clarke, B. A. (2014). Aspect-dependent variations in regolith creep revealed by meteoric 10Be. Geology, 42, 507–10.Google Scholar
Western, A. W., Grayson, R. B., Bloschl, G., Willgoose, G. R. and McMahon, T. A. (1999). Observed spatial organization of soil moisture and its relation to terrain indices. Water Resources Research, 35, 797–810.Google Scholar
Western, A. W., Zhou, S. L., Grayson, R. B. et al. (2004). Spatial correlation of soil moisture in small catchments and its relationship to dominant spatial hydrological processes. Journal of Hydrology, 286, 113–34.Google Scholar
Wilson, L. (1973). Variations in mean annual sediment yield as a function of mean annual precipitation, American Journal of Science, 273, 335–49.Google Scholar
Whipple, K. X. (2004). Bedrock rivers and the geomorphology of active orogens. Annual Review of Earth and Planetary Sciences, 32, 151–85.Google Scholar
Whiteman, C. D. (2000). Mountain Meteorology: Fundamentals and Applications. New York: Oxford University Press, p. 355
Whittaker, R. H. (1977). Evolution of species diversity in land communities. Evolutionary Biology, 10, 1–67.Google Scholar
Willett, S. D., McCoy, S. W., Perron, J. T., Goren, L. and Chen, C. Y. (2014). Dynamic reorganization of river basins. Science, 343, doi:10.1126/science.1248765.Google Scholar
Willgoose, G. (1994). A physical explanation for an observed area-slope-elevation relationship for catchments with declining relief. Water Resources Research, 30, 151–9.Google Scholar
Willgoose, G., Bras, R. L. and Rodriguez-Iturbe, I. (1991). A coupled channel network growth and hillslope evolution model. 1. Theory. Water Resources Research, 27, 1671–84.Google Scholar
Williams, C. A. and Albertson, J. D. (2005). Contrasting short- and long-timescale effects of vegetation dynamics on water and carbon fluxes in water-limited ecosystems. Water Resources Research, 41, W06005.Google Scholar
Yetemen, O., Istanbulluoglu, E. and Vivoni, E. R. (2010). The implications of geology, soils, and vegetation on landscape morphology: inferences from semi-arid basins with complex vegetation patterns in Central New Mexico, USA. Geomorphology, 116, 246–63.Google Scholar
Yetemen, O., Istanbulluoglu, E., Flores-Cervantes, J. H., Vivoni, E. R. and Bras, R. L. (2015). Ecohydrologic role of solar radiation on landscape evolution. Water Resources Research, 51, doi:10.1002/2014WR016169.Google Scholar
Yoo, K., Amundson, R., Heimsath, A. M. and Dietrich, W. E. (2005). Process-based model linking pocket gopher (Thomomys bottae) activity to sediment transport and soil thickness. Geology, 33, 917–20.Google Scholar
Zhou, X., Istanbulluoglu, E. and Vivoni, E. R. (2013). Modeling the ecohydrological role of aspect-controlled radiation on tree-grass-shrub coexistence in a semiarid climate. Water Resources Research, 49, 2872–95.Google Scholar
Walling, D. E. and Kleo, A. H. A. (1979). Sediment yields of rivers in areas of low precipitation: a global review. In: The Hydrology of Areas of Low Precipitation. Proceedings of the Canberra Symposium. International Association of Hydrological Sciences Publication, 128, 479–93.

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

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

Available formats
×

Save book to Google Drive

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

Available formats
×