TODO: Change the header's content in Project Settings : HTML (CHM) Export : Setup HTML Template and Layout

References

Aadland, L. P. (1993). "Stream Habitat Types: Their Fish Assemblages and Relationship to Flow " North American Journal of Fisheries Management 13: 790-806.
Abbe, T. B. and D. R. Montgomery (1996). "Large woody debris jams, channel hydraulics and habitat formation in large rivers." Regulated Rivers: Research & Management 12: 1-21.
Abbe, T. B. and D. R. Montgomery (2003). "Patterns and processes of wood debris accumulation in the Queets river basin, Washington." Geomorphology 51: 81-107.
Field surveys in the 724-km2 Queets river basin on the west slope of the Olympic Mountains in NW Washington reveal
basin-wide patterns of distinctive wood debris (WD) accumulations that arise from different mechanisms of WD recruitment,
hydraulic geometry, and physical characteristics of WD. Individual pieces of WD in an accumulation or jam can be
separated into key, racked, and loose members. Ten types of WD accumulations are identified based on the mode of
recruitment and the orientation of key, racked, and loose debris relative to the channel axis. Although some types of WD
accumulation have few geomorphic effects, others form stable in-stream structures that influence alluvial morphology at
both subreach- and reach-length scales ranging from less than 1 to greater than 10 channel widths. In the Queets river,
stable accumulations of WD directly influence channel anabranching, planform geometry, flood plain topography, and
establishment of long-term riparian refugia for old-growth forest development. The classification of wood debris
accumulations in the Queets river basin is based on physical observations that offer a template potentially applicable to
other forested mountain regions.
 
Abbott, L. L. (1987). The effect of fire on subsequent growth of surviving trees in an old growth redwood forest in Redwood National Park, California. Arcata, California, Humboldt State University.
Abbott, M. B., J.C. Bathurst, J.A. Cunge, P.E. O'Connell, and J. Rasmussen (1986). "An introduction to the european hydrological system - systeme hydologique europeen, "SHE", 2: structure of a physically based, distributed modelling system." Journal of Hydrology 87: 61-77.
Abrahams, A. D., G Li, C Krishnan, and JF Atkinson. "A sediment transport equation for interrill overland flow on rough surfaces."
A model for predicting the sediment transport capacity of turbulent interrill flow on rough surfaces is developed from 1295 flume experiments with flow depths ranging from 3.4 to 43.4 mm, flow velocities from 0.09 to 0.65 m s super(-1), Reynolds numbers from 5000 to 26949, Froude numbers from 0.23 to 2.93, bed slopes from 2.7 degree to 10 degree , sediment diameters from 0.098 to 1.16 mm, volumetric sediment concentrations from 0.002 to 0.304, roughness concentrations from 0 to 0.57, roughness diameters from 1.0 to 91.3 mm, rainfall intensities from 0 to 159 mm h super(-1), flow densities from 1002 to 1501 kg m super(-3), and flow kinematic viscosities from 0.913 to 2.556 x 10 super(-6) m super(2) s super(-1). Stones, cylinders and miniature ornamental trees are used as roughness elements. Given the diverse shapes, sizes and concentrations of these elements, the transport model is likely to apply to a wide range of ground surface morphologies. Using dimensional analysis, a total-load transport equation is developed for open-channel flows, and this equation is shown to apply to interrill flows both with and without rainfall. The equation indicates that the dimensionless sediment transport rate phi is a function of, and therefore can be predicted by, the dimensionless shear stress theta , its critical value theta sub(c), the resistance coefficient u/u*, the inertial settling velocity of the sediment w sub(i), the roughness concentration C sub(r), and the roughness diameter D sub(r). Testing reveals that the model gives good unbiased predictions of phi in flows with sediment concentrations less than 0.20. Flows with higher concentrations appear to be hyperconcentrated and to have sediment transport rates higher than those predicted by the model.
 
Abrahams, A. D. (1972). "Drainage densities and sediment yields in eastern Australia." Australian Geographical Studies 10: 19-41.
Abrahams, A. D., and  J.F. Atkinson (1993). "Relation between grain velocity and sediment concentration in overland flow." Water Resources Research 29(9): 3021-3028.
Abt, S. R., et al. (1992). "Ability of Streambed Vegetation to Entrap Fine Sediments." Interdiscipinary Approaches in Hydrology and Hydrogeology: 249-259.
Abt, S. R., et al. (1993). Sediment Entrapment in Vegetated Streambeds. Preserving Our Environment - The Race Is On, Indianapolis, Indiana, International Erosion Control Association.
Achord, S., et al. (2007). "Migration timing, growth, and estimated parr-to-smolt survival rates of wild Snake River spring-summer chinook salmon from the Salmon River Basin, Idaho, to the Lower Snake River." Transactions of the American Fisheries Society 136: 142-154.
Survival, growth, and juvenile migration timing are key life history traits for at-risk salmon
populations. To estimate these traits in threatened wild Snake River spring–summer Chinook salmon
Oncorhynchus tshawytscha, we tagged fish as parr in 3–17 natal streams per year from 1991 to 2003. We
injected passive integrated transponder tags into parr collected from streams within the Salmon River basin in
Idaho. Each spring, after the previous summer’s tagging, fish were detected as smolts in the juvenile fish
bypass systems of lower Snake River dams. Estimated parr-to-smolt survival to Lower Granite Dam
(excluding migration year 1992) ranged from 3% to 48% for individual populations and from 8% to 25%
(yearly average ¼ 16%) for all streams combined. From 1998 to 2004, estimated parr-to-smolt survival
declined from 25% to 8%, in part because of parr density-dependent effects. Overall annual average growth
rates from tagging to detection at Little Goose Dam ranged from 39.7 to 43.3 mm during 2001–2004, and
significant differences in growth were observed among sites and years. Growth of individuals was positively
related to elapsed time between tagging and recapture and negatively related to fork length at tagging. Annual
migration timing distributions for fish populations from the different streams varied highly within and
between years. Timing of the 10th to 90th percentile passing Lower Granite Dam ranged from 20 to 45 d for
the combined wild populations (average ¼ 38 d). Median passage date was negatively related to autumn
temperature, spring temperature, and March river flow, and was positively related to elevation of the tagging
site. Baseline data generated by this project provide a foundation for understanding the biocomplexity of these
populations, which is critical to effective recovery efforts for these threatened wild fish stocks.
 
Acker, S. A., et al. (2003). "Composition, complexity and tree mortality in riparian forests in the central Western Cascades of Oregon." Forest Ecology and Management 173: 293-308.
Ackroyd, P. (1984). "En Masse Debris Transport in a Mountain Stream." Earth Surface Process and Landforms 9: 307-320.
Acornley, R. M. and D. A. Sear (1999). "Sediment transport and siltation of brown trout (Salmo trutta L.) spawning gravels in chalk streams." Hydrological Processes 13: 447--458.
Adair, M. J., et al. (1989). "Geology and Seismicity of the Skagit Nuclear Power Plant Site." Engineering Geology in Washington I: 607-624.
Adams, J. (1980). "Contemporary uplift and erosion of the Southern Alps, New Zealand." Geological Society of America Bulletin 91: 1-114.
Adams, J. (1984). "Active Deformation of the Pacific Northwest Continental Margin." Tectonics 3(4): 449-472.
Adams, J. (1985). Large-scale tectonic geomorphology of the Southern Alps, New Zealand. Tectonic Geomorphology. M. Morisawa and J. T. Hack, Allen and Unwin.
Adams, J. N. (1979). Variations in Gravel Bed Compostion of Small Streams in the Oregon Coast Range, Oregon State University: 160.
Adams, J. N. and R. L. Beschta (1980). "Gravel Bed Composition in Oregon Coastal Streams " Canadian Journal of Fisheries and Aquatic Sciences 37: 1514-1521.
Adams, M. B., L.H. Loughry, and L.L. Plaugher. (2004). Experimental forests and ranges of the USDA Forest Service., General Technical Report GTR-NE-321. USDA Forest Service, North-East Region, Parsons, WV: 178 p.
Adams, T. O., D.D. Hook, and M.A. Floyd (1995). "Effectiveness monitoring of silvicultural best management practices in South Carolina." SJAF 19(4): 170-176.
Agee, J. K. (1990). Fire History Along an Elevational Gradient in the Siskiyou Mountains, Oregon (DRAFT). Seattle, Washington, University of Washington: 24.
Agee, J. K. (1993). Fire Ecology of Pacific Northwest Forests. Washington, D.C., Island Press.
Agee, J. K. (1994). An analysis of catastrophic forest disturbance on the Olympic Peninsula (DRAFT), prepared for Rayonier, Inc.: 2-29.
Agee, J. K. (?). Fire history of Douglas-fir forests of the Pacific Northwest.
Agee, J. K., et al. (1990). "Forest fire history of Desolation Peak, Washington." Canadian Journal of Forest Resources 20: 350-356.
Agee, J. K. and R. Flewelling (1983). A fire cycle model based on climate for the Olympic Mountains, Washington. Seventh Conference on Fire and Forest Meteorology, Boston, MA, American Meteorological Society.
Agee, J. K. and R. Flewelling (1983). A fire cycle model based on climate for the Olympic Mountains, Washington. Proceedings of Seventh Conference on Fire and Forest Meteorology. Boston, American Meteorology Society. 7: 32-37.
Agee, J. K. and M. H. Huff (1987). "Fuel succession in a western hemlock/Douglas-fir forest." Canadian Journal of Forest Resources 17: 697-704.
Agee, J. K. and D. R. Johnson (1988). Ecosystem management for parks and wilderness: workshop synthesis. Seattle, Washington, Institute of Forest Resources: 2-39.
Agency, U. S. E. P. (1991). Guidance for water quality-based decisions: The TMDL Process, United States Environmental Protection Agency.
Agnese, C., et al. (1988). "Estimation of the time scale of the geomorphologic instantaneous unit hydrograph from effective streamflow velocity." Water Resources Research 24(7): 969-978.
Agrawal, A., et al. (2005). Predicting the potential for historical coho, chinook and steelhead habitat in northern California, National Oceanic and Atmospheric Administration: 25.
Ahl, R. S., S. Woods,  and M. DiLuzio (2005). "Integrating landscape and hydrologic simulation models to assess the influence of fire-suppression on water yield from forested Rocky Mountain watersheds." Eos Trans. AGU,  86(52) Fall Meet. Suppl., Abstract #H41D-0436. San Francisco, California.
Ahl, R. S., S.W. Woods, and H.R. Zuuring (2008). "Hydrologic calibration and validation of SWAT in a snow-dominated Rocky Mountain watershed, Montana, U.S.A." Journal of the American Water Resources Association 44(6): 1411-1430.
Ahnert, F. (1994). "Equilibrium, scale and inheritance in geomorphology." Geomorphology 11: 125-140.
Ai, N. S. and T. D. Miao (1987). "A model of progressive slope failure under the effect of the neotectonic stress field." Catena Supplement 10: 21-29.
Aitt, R. B. and R. M. Thorson (?). The Cordilleran Ice Sheet in Washingon, Idaho, and Montana. ?
Ajward MH, a. I. M. (2000). "A spatially varied unit hydrograph model." Journal of Environmental Hydrology 8(7).
Alberti, M., et al. (2007). "The impact of urban patterns on aquatic ecosystems: An empirical analysis in Puget lowland sub-basins." Landscape and Urban Planning 80: 345-361.
Landscape change associated with urbanization poses major challenges to aquatic ecosystems. Extensive studies have shown that the composition
of land cover within a watershed can account for much of the variability in water quality and stream ecological conditions. While several studies
have addressed the relationship between watershed urbanization and biotic integrity in streams, few have directly addressed the question of how
urban patterns influence ecological conditions. These studies typically correlate changes in ecological conditions with simple aggregated measures
of urbanization (e.g., human population density or percent impervious surface).We develop an empirical study of the impact of urban development
patterns on stream ecological conditions in 42 sub-basins in the Puget Sound lowland region on a gradient of urbanization. We hypothesize that
ecological conditions in urbanizing landscapes are influenced through biophysical changes by four urban pattern variables: land use intensity, land
cover composition, landscape configuration, and connectivity of the impervious area. Using community measures of benthic macroinvertebrates
as indicators of in-stream biotic integrity we examined the relationships between urban development patterns and ecological conditions in these
basins. Significant statistical relationships were found between landscape patterns—both amount and configuration of impervious area and forest
land—and biotic integrity of streams suggesting that patterns of urban development matter to aquatic ecosystems.
 
 
Alexander, C. a. C. V. (2003). "Modern sedimentary processes in the Santa Monica, California continental margin: sediment accumulation, mixing and budget." Marine Environmental Research 56(1-2): 177-204.
Alexander, G. R. and E. A. Hansen (1986). "Sand Bed Load in a Brook Trout Stream." North American Journal of Fisheries Management 6: 9-23.
Alila, Y., and M. Schnorbus (2005). "Peak flow responses to forest harvesting and roads  in the Maritime Regions of the Pacific Northwest: a preferential hillslope runoff perspective." Eos Trans. AGU,  86(52) Fall Meet. Suppl., Abstract #H24A-04.
Allan, J. D. (1995). Stream ecology: structure and function of running waters. London, Chapman & Hall.
Allan, J. D. (2004). "landscapes and riverscapes: the influence of land use on stream ecosystems." Annu. Rev. Ecol. Evol. Syst. 35: 257-284.
Local habitat and biological diversity of streams and rivers are strongly
influenced by landform and land use within the surrounding valley at multiple scales.
However, empirical associations between land use and stream response only varyingly
succeed in implicating pathways of influence. This is the case for a number of reasons,
including (a) covariation of anthropogenic and natural gradients in the landscape;
(b) the existence of multiple, scale-dependent mechanisms; (c) nonlinear responses;
and (d) the difficulties of separating present-day from historical influences. Further
research is needed that examines responses to land use under different management
strategies and that employs response variables that have greater diagnostic value than
many of the aggregated measures in current use.
 
 
Allan, J. D., et al. (1997). "The influence of catchment land use on stream integrity across multiple spatial scales." Freshwater Biology 37: 149-161.
1. Despite wide recognition of the need for catchment-scale management to ensure the
integrity of river ecosystems, the science and policy basis for joint management of land
and water remains poorly understood. An interdisciplinary case study of a river basin
in south-eastern Michigan is presented.
2. The River Raisin drains an area of 2776 km2, of which some 70% is agricultural land.
The upper basin consists of till and outwash, and both topography and land use/cover
are diverse. The lower basin consists of fine textured lake deposits, is of low relief, and
land use is primarily agricultural.
3. The River Raisin basin historically was a region of oak-savannah and wetlands. It
was deforested, drained and converted to farmland during the mid-nineteenth century.
Human population reached a plateau at about 1880, and then underwent a second
period of growth after 1950, mainly in small urban areas. More recently, the amount of
agricultural land has declined and forested land has increased, in accord with a general
decline in farming activity.
4. It could be suggested that the influence of land use on stream integrity is scaledependent.
Instream habitat structure and organic matter inputs are determined
primarily by local conditions such as vegetative cover at a site, whereas nutrient supply,
sediment delivery, hydrology and channel characteristics are influenced by regional
conditions, including landscape features and land use/cover at some distance upstream
and lateral to stream sites.
5. Sediment concentrations measured during low flows were higher in areas of greater
agriculture. In a comparison of two subcatchments, sediment yields were up to ten
times greater in the more agricultural location, in response to similar storm events. A
distributed parameter model linked to a geographical information system predicted that
an increase in forested land cover would result in dramatic declines in runoff and
sediment and nutrient yields.
6. Habitat quality and biotic integrity varied widely among individual stream sites in
accord with patterns in land use/cover. Extent of agricultural land at the subcatchment
scale was the best single predictor of local stream conditions. Local riparian vegetation
was uncorrelated with overall land use and was a weak secondary predictor of habitat
quality and biotic integrity.
7. Investigation of the regulatory agencies involved in land and water management in
the basin revealed a complex web of overlapping political jurisdictions. Most land-use
decision-making occurs at the local level of township, city or village. Unfortunately, local decision-making bodies typically lack the information and jurisdictional authority
to influence up- and downstream events.
 
 
Allen, J. D. (1995). Stream Ecology: Structure and Function of Running Waters. Norwell, MA, Academic Publishers.
Allen, T. F. H. and T. W. Hoekstra (1987). Problems of scaling in restoration ecology: a practical application. Restoration Ecology. W. R. Jordan. Cambridge, MA, Cambridge University Press: 289-299.
Allen, T. F. H. and T. B. Starr (1982). Hierarchy: Perspectives for Ecological Complexity. Chicago, University of Chicago Press.
Allmendinger, R. W., et al. (1983). "Paleogeography and Andean Structural Geometry, Northwest Argentina." Tectonics 2(1): 1-16.
Amaranthus, M. P., et al. (1985). "Logging and Forest Roads Related to Increased Debris Slides in Southwestern Oregon." Journal of Forestry 83(4): 229-233.
Ambers, R. K. R. (2000). "Using the Sediment Record in a Western Oregon Flood-control Reservoir to Assess the Influence of Storm History and Logging on Sediment Yield (DRAFT)." submitted to the Journal of Hydrology: 1-12.
Amoros, C., et al. (1987). "A method for applied ecological studies of fluvial hydrosystems." Regulated Rivers 1: 17-36.
Anbalagan, R. (1992). "Landslide hazard evaluation and zonation mapping in mountainous terrain." Engineering Geology 32: 269-277.
Anders, A. M., et al. (2008). "Influence of precipitation phase on the form of mountain ranges." Geology 36(6): 479-482.
Anderson, D. M., and L.H. MacDonald (1998). "Modelling road surface sediment production using a vector GIS." Earth Surface Processes and Landforms 23: 95-107.
Anderson, E. W. and N. J. Cox (1978). "A comparison of different instruments for measuring soil creep." Catena 5: 81-93.
Anderson, J. H. and T. P. Quinn (2007). "Movements of adult coho salmon (Oncorhynchus kisutch) during colonizatino of newly accessible habitat." Canadian Journal of Fisheries and Aquatic Science 64: 1143-1154.
Anderson, J. K., et al. (2005). "Patterns in stream longitudinal profiles and implications for hyporheic exchange flow at the H.J. Andrews Experimental Forest, Oregon, USA." Hydrological Processes 19: 2931-2949.
There is a need to identify measurable characteristics of stream channel morphology that vary predictably throughout
stream networks and that influence patterns of hyporheic exchange flow in mountain streams. In this paper we
characterize stream longitudinal profiles according to channel unit spacing and the concavity of the water surface
profile. We demonstrate that: (1) the spacing between zones of upwelling and downwelling in the beds of mountain
streams is closely related to channel unit spacing; (2) the magnitude of the vertical hydraulic gradients (VHGs)
driving hyporheic exchange flow increase with increasing water surface concavity, measured at specific points along
the longitudinal profile; (3) channel unit spacing and water surface concavity are useful metrics for predicting how
patterns in hyporheic exchange vary amongst headwater and mid-order streams. We use regression models to describe
changes in channel unit spacing and concavity in longitudinal profiles for 12 randomly selected stream reaches spanning
62 km2 in the H.J. Andrews Experimental Forest in Oregon. Channel unit spacing increased significantly, whereas
average water surface concavity (AWSC) decreased significantly with increasing basin area. Piezometer transects
installed longitudinally in a subset of stream reaches were used to measure VHG in the hyporheic zone, and to
determine the location of upwelling and downwelling zones. Predictions for median pool length and median distance
between steps in piezometer reaches bracketed the median distance separating zones of upwelling in the stream
bed. VHG in individual piezometers increased with increasing water surface concavity at individual points in the
longitudinal profile along piezometer transects. Absolute values of VHG, averaged throughout piezometer transects,
increased with increasing AWSC, indicating increased potential for hyporheic exchange flow. These findings suggest
that average hyporheic flow path lengths increase—and the potential for hyporheic exchange flow in stream reaches
decreases—along the continuum from headwater to mid-order mountain streams.
 
 
Anderson, P. L. and M. M. Meerschaert (1998). "Modeling river flows with heavy tails." Water Resources Research 34(9): 2271-2280.
Anderson, R., J., et al. (2004). "Width of streams and rivers in response to vegetation, bank material, and other factors." Journal of the American Water Resources Association 40(5): 1159-1172.
An extensive group of datasets was analyzed to examine
factors affecting widths of streams and rivers. Results indicate
that vegetative controls on channel size are scale dependent. In
channels with watersheds greater than 10 to 100 km2, widths are
narrower in channels with thick woody bank vegetation than in
grass lined or nonforested banks. The converse is true in smaller
streams apparently due to interactions between woody debris,
shading, understory vegetation, rooting characteristics, and channel
size. A tree based statistical method (regression tree) is introduced
and tested as a tool for identifying thresholds of response
and interpreting interactions between variables. The implications
of scale dependent controls on channel width are discussed in the
context of stable channel design methods and development of
regional hydraulic geometry curves.
 
 
Anderson, S. A. and N. Sitar (1995). "Analysis of rainfall-induced debris flows." Journal of Geotechnical Engineering 121(7): 544-.
Rainfall-induced debris flows are flow failures in residual and colluvial soils initiated by a reduction of confining stress as a result of pore-water pressure rise during or following periods of intense rainfall. To establish appropriate stability analysis procedures, the behavior of soil from a debris flow source area is investigated. Constant-shear-drained triaxial tests, tests that mimnic the field stress path, and anisotropically consolidated undrained triaxial tests performed at the in-situ stress level indicite that undrained loading is a prerequisite for flow failure. Mechanisms of stress transfer are proposed to explain how the initially drained deformation that occurs along the field stress path can lead to undrained mobilization. Because a debris flow involves both drained initiation and undrained mobilization, a complete stability analysis includes an effective stress analysis with constant-shear-drained parameters to determine the potential for failure initiation, and an undrained strength analysis using undrained residual strength to determine the potential for debris flow mobilization.
 
Anderson, S. P., et al. (1997). "Subsurface flow paths in a steep, unchanneled catchment." Water Resources Research 33(12): 2637-2653.
Andrea Tribe (1992). "Automated recognition of valley lines and drainage networks from grid digital elevation models: a review and a new method." Journal of Hydrology 139: 263-293.
Andrews, E. D. (1979). Scour and Fill In a Stream Channel, East Fork River, Western Wyoming US Geological Survey: 49.
Andrews, E. D. (1984). "Bed-material entrainment and hydraulic geometry of gravel-bed rivers in Colorado." Geological Society of America Bulletin 95: 371-378.
Andrews, H. J. and R. W. Cowlin (1934). Forest resources of the Douglas-fir region, USDA: 169.
Andrle, R. (1994). "Flow structure and development of circular meander pools " Geomorphology 9: 261-270.
Andrus, C. and H. A. Froehlich (1987). Riparian forest development after logging or fire in the Oregon coast range: wildlife habitat and timber value. Symposium on Streamside Management: Riparian Wildlife and Forestry Interactions, Washington, Seattle.
Appt, J. (2002). Discriminating between landslide sites and potentially unstable terrain using topographic indices. Forest Engineering. Corvallis, OR, Oregon State University. Master of Science: 116.
A landslide inventory, statistical analyses and a Geographic Information System
(GIS) are used to analyze landslide sites and potentially unstable terrain in the
Oregon Coast Range. The objectives are to evaluate the efficacy of locating
landslide sites with topographic variables and discriminate the difference between
sites where landslides have and have not occurred. The population of known
landslides are characterized as up-slope, non-road related, and associated with 1996
storm events. Topographic variables are derived from a Digital Elevation Model
(DEM) for index construction forming six groups; i) slopes, ii) contributing areas,
iii) ratios of slope and contributing area, iv) curvature v) infinite slope models, and
vi) functions of slope and contributing area based on statistical models. Index
groups employ different algorithms. Index performance is measured with landslide
and aerial densities. Cumulative landslide occurrence is plotted against cumulative
area on a continuous domain of the index to locate a maximum landslide density on
equal size areas. Indices are used to generate model definitions of potentially
unstable terrain based on similarity to the landslide population. Aerial densities of
potentially unstable terrain based on index definitions are determined but no
common metric is achieved. Statistical analyses on spatially stratified data suggest
a significant (α < 0.05) difference between landslides sites and adjoined terrain.
The minimum resolution at which a significant difference is achieved based on
spatial stratification is a three cell radius surrounding the slide population.
Curvature and area discriminate better than simple slope and topographic ratios.
The relative performance is mostly a function of DEM error and resolution, and
spatial correlation. Hydrologic geomorphic models perform about as well as the
topographic ratios, and much less than the simple area index. There is no statistical
evidence to suggest that the hydrologic geomorphic models accurately describe a
threshold in the Mapleton slide population. The lack of significance is likely due to
limitations on the available parameter sets. Logistic regression produced an index
with the highest discrimination performance due to a maximum likelihood
algorithm. Regression models have a physical basis in and parallel the behavior of
linked hydrologic geomorphic and slope stability models. The measured
differences in performance are a useful assessment of the DEM – index
combination.
 
Arave, N. (2008). Locating channel heads with digital elevation models, geology, soils, and spectral data in a geographic information system. Department of Geology, Idaho state University. Master of Science: 74.
This thesis presents a two part investigation into methods for modeling channel
initiation in a geographic information system. For this study channel heads mapped on
USGS 7.5” Topographic quadrangles were used as the “known” dataset. The region of
study was the southern extents of the Beaverhead and Centennial Ranges in Idaho. The
first approach was to evaluate terrain characteristics at channel heads on a point by point
basis. The results of this study indicate the majority of the channel heads in the study
region are on southern facing slopes. The second approach was a paired watershed
assessment of the Beaver Creek and Medicine Lodge Creek basins. The assessment
compared land cover to mean flow accumulation at channel heads. The results from this
study suggested that as percent brush cover increases relative to forest cover the mean
flow accumulation at channel heads will increase
 
Araya, T. and S. Higashi (?). Debris movement in torrential rivers of volcanic areas (DRAFT), Hokkaido University: 22.
Ardizzone, F., et al. (2002). "Impact of mapping errors on the reliability of landslide hazard maps." Natural Hazards and Earth System Sciences 2: 3-14.
Identification and mapping of landslide deposits are an intrinsically difficult and subjective operation that requires a great effort to minimise the inherent uncertainty. For the Staffora Basin, which extends for almost 300 km2 in the northern Apennines, three landslide inventory maps were independently produced by three groups of geomorphologists. In comparing each map with the others, large positional discrepancies arise (in the range of 55–65%). When all three maps are overlain, the locational mismatch of landslide deposit polygons increases to over 80%. To assess the impact of these errors on predictive models of landslide hazard, for the study area discriminant models were built up from the same set of geological-geomorphological factors as predictors, and the occurrence of landslide deposits within each terrain-unit, derived from each inventory map, as dependent variable. The comparison of these models demonstrates that statistical modelling greatly minimises the impact of input data errors which remain, however, a major limitation on the reliability of landslide hazard map
 
Arhonditsis, G., et al. (2002). "Quantitative assessment of agricultural runoff and soil erosion using mathematical modeling: Applications in the Mediterranean region." Environmental Management 30(3): 434-453.
Armantrout, N. B. (1990). Index areas as population indicators. 1990 Northeast Pacific Chinook and Coho Salmon Workshop.
Armantrout, N. B. (1991). "Restructuring Stream for Anadromous Salmonids." American Fisheries Society Symposium 10: 136-149.
Arnold, J. F. (1989). Hierarchiacal Land Systems Stratifications: A Holistic Perspective for Ecologically Integrated Informations Systems. Geoprocessing and Surveying Conference, Boise, Idaho.
Arnold, J. G., R. Srinivasan, R.S. Muttiah, and J.R. Williams (1998). "Large area hydrologic modeling and assessment part I: model development." Journal of the American Water Resources Association 34(1): 73-89.
ASCE (1986). Landslide Dams: Processes, Risk, and Mitigation. Landslide Dams: Processes, Risk, and Mitigation, Seattle, Washington, American Society of Civil Engineers.
ASCE Task Committee on Glossary of Hydraulic Modeling Terms of the Committee on Research of the Hydraulics Division (1982). "Modeling Hydraulic Phenomena: A Glossary of Terms." Journal of the Hydraulics Division 108: 845-1982.
ASCE Task Committee on Sediment Transport and Aquatic Habitats, S. C. (1992). "Sediment and aquatic habitat in river systems." Journal of Hydraulic Engineering 118(5): 669-687.
Asch, T. W. J. V. and U. V. Steijn (1991). "Temporal patterns of mass movements in the French Alps." Catena 18: 515-527.
Ashworth, P. J. and R. I. Ferguson (1989). "Size-Selective Entrainment of Bed Load in Gravel Bed Streams." Water Resources Research 25(4): 627-634.
Atwater, B. F. (1987). "Evidence for Great Holocene Earthquakes Along the Outer Coast of Washington State." Science 236: 942-944.
Atwater, B. F. and A. L. Moore (1992). "A Tsunami About 1,000 Years Ago in Pudget Sound, Washington." Science 258: 1614-1616.
Atwater, B. F., et al. (1991). "Radiocarbon test of earthquake magnitude at the Cascadia subduction zone." Nature 353(12): 156-158.
Austin, S. A. (1999). Streamflow response to forest management: a meta-analysis using published data and flow duration curves. Department of Earth Resources. Fort Collins, CO, Colorado State University: 265.
Axelsson, P. (1999). "Processing of laser scanner data - algorithms and applications." ISPRS Journal of Photogrammetry and Remote Sensing 54: 138-147.
Babovic, V., and M. Keijzer (2002). "Rainfall runoff modelling based on genetic programming." Nordic Hydrology 33(5): 331-346.
Backiel, A., and R.W. Gorte (1992). Clearcutting in the National Forests. C. R. Service, Congressional Research Service, Library of Congress, Washington DC.: 18p.
Badger, T. (1993). Structurally controlled landslides northwestern Olympic Peninsula, USA. Geotechnical Engineering of Hard Soils - Soft Rocks. A. e. al: 1051-1056.
Badger, T. (1993). Structurally-controlled landslides northwestern Olympic Peninsula, USA. Geotechnical Engineering of Hard Soils - Soft Rocks. A. e. al.: 1051.
Badura, G. J., et al. (1974). Siuslaw National Forest Soil resource survey. Corvallis, Oregon, USDA Forest Service, Siuslaw National Forest: 139.
Baffaut, C., M.A.Nearing, J.C. Ascough, and B. Liu (1997). "The WEPP watershed model: II Sensitivity analysis and discretization on small watersheds." Transactions of the ASAE 40(4): 935-943.
Bagnold, R. (1977). "Bedload transport by natural rivers." Water Resources Research 13(2): 303-312.
Bagnold, R. (1980). An empirical correlation of bedload transport rates in flumes and natural rivers. Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences.
Bagnold, R. (1986). Transport of solids by natural water flow: Evidence of a worldwide correlation. Proceedings of the Royal Society of London, Series A, Mathematical and physical sciences.
Bagnold, R. A. (1954). "Experiments on a gravity free dispersion of large solid spheres in a Newtonian fluid under shear." Proceedings of the Royal Society of London 225A: 49-63.
Bagnold, R. A. (1966). An approach to the sediment transport problem from general physics, United States Geological Survey.
Bagnold, R. A. (1973). "The nature of saltation and of 'bed-load' transport in water." Proceedings of the Royal Society of London, Serial A 332(473-504).
Bagnold, R. A. (1980). "An empirical correlation of bedload transport rates in flumes and natural rivers." Proceedings of the Royal Society of London, Serial A(372): 453-473.
Bahuguna, D., et al. (2010). "Windthrow and recruitment of large woody debris in riparian stands." Forest Ecology and Management 259: 2048-2055.
To document the impacts of windthrow in riparian leave strips and identify the components needed for
small stream large woody debris (LWD) recruitment modeling, we monitored nine small streams at a
temperate rainforest site in coastal British Columbia. This study was a component of a larger integrated
study of forest management impacts on small streams. A series of small clearcuts were harvested in 1998
in a 70-year-old second growth stand that had regenerated naturally following logging and wildfire. Three
cutblocks each were assigned to 10m and 30m buffer width treatments and three areas were assigned
as unharvested controls. Seven years after the 1998 logging, all logs greater than 10cm diameter that
spanned at least part of stream channel width were measured. A total of 179 logs were recorded. Postharvest
windthrow was higher in the 10m buffer treatment, while competition-related standing tree
mortality was higher in the controls. The major windthrow events had occurred in the first and second
years after logging of adjacent stands. There was no significant difference in the number of spanning
and in-stream logs in the 10 m, 30m buffer and control treatments. More than 90% of the LWD was in
the 10–30cm diameter classes. The majority of logs were oriented perpendicular to the stream channel.
At the time of measurement, the majority of these trees were still suspended above the stream channel,
indicating that the recruitment of logs into the stream channel is a long-term process. Time to recruitment
into the channel is dependent on log and valley geometry, log size, species, and log condition prior to
toppling. Log height above stream was negatively correlated with log decay class and valley width. Log
length was negatively correlated with state of decay, and many windthrown logs were in an advanced
state of decay before they entered the stream.
 
Bailey, R. G., P.E. Avers,  T. King, and W.H. McNab (1994). Ecoregions and subregions of the United States (map). 1:7,500,000. With supplementary table of map unit descriptions, compiled and edited by W. H. McNab and R. G. Bailey. USDA Forest Service, Washington, DC.
Bajracharya, D. (1983). "Deforestation in the food/fuel context historical and political perspectives from Nepal." Mountain Research and Development 3(3): 227-240.
Bak, P. (1994). Self-Organized Criticality: A Holistic View of Nature. Complexity: Metaphors, Models, and Reality. G. Cowan, D. Pines and D. Maltzer. Upton, New York, Addison-Wesley. XIX: 477-496.
Baker, V. R. and C. R. Twidale (1991). "The reenchantment of geomorphology." Geomorphology 4: 73-100.
Baker, W. L. (1989). "Effect of scale and spatial heterogeneity on fire-interval distributions." Canadian Journal of Forest Research 19: 700-706.
The distribution of forest-fire intervals has been characterized by fitting statistical distributions, such as that of Weibull. the papameters of fitted distributions can then be used to compare fire regimes. Fire-interval distributions for the 187-year presettlement fire history record in the Boundary Waters Canoe Area, Minnesota, were analyzed using reconstructed "fire-year" maps. Distributions were determined for sampling units at five spatial scales, from about 25,000 to 400,000 ha. Fire-interval distributions varied from positively to negatively skewed, but for most units the Weibull distribution fit significantly. The distributions varied spatially, and cluster analysis suggested that three fire regions, each containing a relatively homogeneous fire regime, could be identified. The sources of theis spatial variation are unknown. There was less variation between scales within a fire region thatn between fire regions. This contrasts with a previous finding, using the same fire-history data, that scale substantially affects observed landscape age-class distribution. This disparity arises because landscape age-class distributions may fluctuate even if fire-interval distributions do not fluctuate. Reconstruction of fire-interval distributions requires historical data; landscape age-class distributions at an instant in time are insufficient.
 
Balasscio, C. C., D.J. Palmeri, and H. Gao (1998). "Use of a genetic algorithm and multi-objective programming for callibration of a hydrologic model." Transactions of the ASAE 41(3): 615-619.
Barnes, H. H. (1967). Roughness characteristics of natural channels, U.S. Geological Survey: 1849.
Barnowsky, C. W. (1981). "A record of late Quaternary vegetation from Davis Lake, southern Puget Lowland, Washington." Quaternary Research 16: 221-239.
Barrett, J. C. and W. J. Conroy (2001). Evaluation of timber harvest impacts on water quality. American Water Resources Association Technical Publication Series TPS, vol.01-1. J. J. Warwick: 205-210.
The Pacific Lumber Company (PALCO) in 1999 signed a Habitat Conservation Plan (HCP) that commits it to conduct watershed analysis studies, monitoring, and research to determine how its timber harvesting practices impact water quality, and how best to reduce those impacts. The watershed analysis assesses impacts to water quality by evaluating sediment inputs to streams and the ability of riparian forest stands to protect beneficial uses of water. Studies to date have documented that surface erosion from roads is the primary impact to water quality in low relief portions of the ownership, whereas mass wasting processes are the predominant impact in high relief areas. Research efforts have focused on calibration and use of the USFS Water Erosion Potential Project (WEPP) model to estimate sediment volumes generated by timber harvest operations. The work includes evaluation of the effectiveness of hillslope management strategies and riparian buffers in preventing sediment delivery to streams.
 
Barrett, J. C. and W. J. Conroy (2002). Use of watershed scale sediment budgets for timberlands mitigation development. American Water Resources Association Technical Publication Series TPS, vol.02-1. J. R. Lesnik. Bethesda, AWRA - American Water Resources Association: 459-463.
The Pacific Lumber Company (PALCO) owns 225,000 acres of industrial timberlands in coastal Northern California watersheds that drain to tidal estuaries, or salt-water bays used by anadromous salmonids (e.g., coho salmon, Oncorhynchus kisutch). PALCO's Habitat Conservation Plan (HCP) commits the Company to conduct research and monitoring studies, including watershed analysis, to determine how management activities affect sediment delivery rates to streams. Much of this work has focused on development of watershed scale sediment budgets that quantify inputs from harvest, roads, site preparation, and other silvicultural activities. These sediment budgets, because they specifically identify both the mechanisms of sediment delivery, and their relative magnitude, can be used to develop cost-effective sediment mitigation strategies. Examples from the Freshwater Creek and Van Duzen watersheds are presented.
 
Barro, S. C., and S. G. Conrad (1991). "Fire effects on California chaparral systems: an overview." Environment International 17: 135-149.
Barry, J., JM Buffington, and JG King (2004). "A general power equation for predicting bed load transport rates in gravel-bed rivers." Water Resources Research 40(w10401): 22p.
Barry, J. J., et al. (2004). "A general power equation for predicting bed load transport rates in gravel bed rivers." Water Resources Research 40(doi:10.1029/2004WR003190): 22.
Barry, R. G. (1990). "Changes in mountain climate and glacio-hydrological responses." Mountain Research and Development 10(2): 161-170.
Bartle, H. (1999). "Analyzing the Effectiveness of Common Road Construction and Deactivation Techniques." Streamline: Water Restoration Technical Bulletin 4(2): 9.
Bartlein, P. J., S.W. Hostetler, S.L. Shafer, J.O. Holman, and A.M. Solomon (2003). The seasonal cycle of wildfire and climate in the western United States. The 5th Symposium on Fire and Meteorology, American Meteorological Society, Orlando, FL. Paper p3.9.
Bartlein, P. J., S.W. Hostetler, S.L. Shafer, J.O. Holman, and A.M. Solomon (2008). "Temporal and spatial structure in a daily wildfire-start data set from the Western United States (1986-96)." International Journal of Wildland Fire 17: 8-17.
Bartley, R. and I. Rutherfurd (2002). Techniques to quantify the variability of thalveg profiles. The Structure, Function and Management Implications of Fluvial Sedimentary Systems. F. J. Dyer, M. C. Thoms and J. M. Olley, International Association of Hydrological Sciences. 276: 35-44.
This paper presents a range of techniques for quantifying the variability of thalweg profiles. Previous research has shown that the thalweg is an important morphological and ecological feature in stream systems, however, the tools to measure the morphological diversity of thalweg profiles are limited (at the reach scale using a horizontal interval of ~2 m). The methods presented are evaluated against synthetic data to determine which techniques would be most appropriate for quantifying thalweg data collected in the field. The results suggest that three techniques are the most appropriate for quantifying thalweg diversity (the wiggliness factor w, fractal dimension D, and standard deviation of depths).
 
Bartz, K. K., et al. (2006). "Translating restoration scenarios into habitat conditions: an initial step in evaluating recovery strategies for Chinook salmon (Oncorhynchus tshawytscha)." Canadian Journal of Fisheries and Aquatic Science 63: 1578-1595.
One of the challenges associated with recovering imperiled species, such as Chinook salmon (Oncorhynchus
tshawytscha), is identifying a set of actions that will ensure species’ persistence. Here we evaluate the effects of alternative
land use scenarios on habitat conditions potentially important to Chinook salmon. We first summarize the alternative
scenarios as target levels for certain land use characteristics. We then use the target levels to estimate changes in
current habitat conditions. The scenarios we explore indicate considerable potential to improve both the quality and
quantity of salmon habitat through protection and restoration. Results from this analysis constitute the habitat inputs to
a population model linking changes in habitat to salmon population status. By transparently documenting the approach
we use to translate land use actions into changes in salmon habitat conditions, we provide decision makers with a clear
basis for choosing strategies to recover salmon.
 
Basnyat, P., et al. The use of remote sensing and GIS in watershed level analyses of non-point source pollution problems. Forest Ecology and Management 128 (1/2), 65-73 pp.;   34 ref.; 2000. L. G. Arvanitis.
Basin characteristics such as land use/land cover, slope, and soil attributes affect water quality by regulating sediment and chemical concentration. Among these characteristics, land use/land cover can be manipulated to gain improvements in water quality. These land use/land cover types can serve as nutrient detention media or as nutrient transformers as dissolved or suspended nutrients move towards the stream. This study examines a methodology to determine nitrate pollution 'contributing zones' within a given basin based on basin characteristics. The study area was a portion of the Fish River drainage basin (watershed) in Alabama, USA, which feeds into Weeks Bay, and is eventually connected directly to the Gulf of Mexico. The analysis involved classifying land use/land cover types, and delineating basins and 'contributing zones' using geographic information system (GIS) and remote sensing (RS, Landsat thematic mapper data, SPOT panchromatic imagery, aerial photography) analysis tools. A 'land use/land cover-nutrient-linkage-model' was developed which suggests that forests act as a sink, and that as the proportion of forest inside a contributing zone increases (or the proportion of agricultural land decreases), nitrate levels downstream will decrease. In the model, the residential/urban/built-up areas have been identified as strong contributors of nitrate. Other contributors were orchards, and row crops and other agricultural activities.
 
Batalla, R. J., et al. (1999). "Field observations on hyperconcentrated flows in mountain torrents." Earth Surface Processes and Landforms 24: 247-253.
Bates, D. (1988). Water Temperature: A Case of Cumulative Effects?, Gifford Pinchot National Forest, Wind River Ranger District: 10.
filed
 
Bates, D., et al. (1998). "North Santiam River Turbidity Study, 1996/1997." Watershed Management Council Networker: 1-19.
Bathurst, J. C. (1978). "Flow resistance of large-scale roughness." Proceedings of the American Society of Civil Engineers, Journal of the Hydraulics Division 104(HY12): 1587-1603.
Bathurst, J. C. (1982). Flow resistance in boulder-bed streams. Gravel-bed Rivers. R. D. Hey, J. C. Bathurst and C. R. Thorne. Chichester, John Wiley & Sons 443-465.
Bathurst, J. C. (1993). Flow resistance through the channel network. Channel Network Hydrology. K. Beven and M. J. Kirkby, John Wiley & Lons Ltd.: 69-98.
Bathurst, J. C. (2002). "At-a-site variation and minimum flow resistance for mountain rivers " Journal of Hydrology 269: 11-26.
Bathurst, J. C., et al. (1997). "Debris flow run-out and landslide sediment delivery model tests." Journal of Hydraulic Engineering 123: 410-419.
Bathurst, J. C., et al. (1983). Bedforms and flow resistance in steep gravel-bed channels. Mechanics of Sediment Transport. B. M. Sumer and A. Muller. Rotterdam, the Netherlands, Balkema: 215-221.
Battin, J., et al. (2007). "Projected impacts of climate change on salmon habitat restoration." Proceedings of the Natonal Academy of Sciences of the USA 104(16): 6720-6725.
Throughout the world, efforts are under way to restore watersheds,
but restoration planning rarely accounts for future climate
change. Using a series of linked models of climate, land cover,
hydrology, and salmon population dynamics, we investigated the
impacts of climate change on the effectiveness of proposed habitat
restoration efforts designed to recover depleted Chinook salmon
populations in a Pacific Northwest river basin. Model results
indicate a large negative impact of climate change on freshwater
salmon habitat. Habitat restoration and protection can help to
mitigate these effects and may allow populations to increase in the
face of climate change. The habitat deterioration associated with
climate change will, however, make salmon recovery targets much
more difficult to attain. Because the negative impacts of climate
change in this basin are projected to be most pronounced in
relatively pristine, high-elevation streams where little restoration
is possible, climate change and habitat restoration together are
likely to cause a spatial shift in salmon abundance. River basins that
span the current snow line appear especially vulnerable to climate
change, and salmon recovery plans that enhance lower-elevation
habitats are likely to be more successful over the next 50 years than
those that target the higher-elevation basins likely to experience
the greatest snow–rain transition.
 
 
Baum, R. L., et al. (2002). Regional Landslide-Hazard Assessment - An Example from Seattle, Washington. IX Great Lakes Geotechnical/Geoenvironmental Conference, Dayton, Ohio.
Baxter, C., V., et al. (1999). "Geomorphology, logging roads, and the distribution of Bull Trout spawning in a forested river basin: implications for management and conservation." Transactions of the American Fisheries Society 128: 854-867.
The Swan Basin in Montana is considered a stronghold of regional significance for
the bull trout Salvelinus confluentus, a native char whose populations are fragmented and declining
throughout its range. We used correlation analysis to examine spatial and temporal variation of
bull trout redd count data (1982–1995) relative to geomorphic and land-use factors among nine
principal spawning tributaries of the Swan River. Bull trout redd numbers were positively correlated
with the extent of alluvial valley segments bounded by knickpoints and negatively correlated with
the density of logging roads in spawning tributary catchments. The density of logging roads in
spawning tributary catchments was not significantly correlated with geomorphic factors. Temporal
trends among the principal spawning streams were variable. In four of the nine principal spawning
streams, redd numbers increased significantly during the survey period, and in the remaining
streams, redd numbers showed no significant change. Changes in redd numbers with time were
negatively correlated with catchment road density and positively correlated with the extent of
bounded alluvial valley segments. The significance of bounded alluvial valley segments to bull
trout spawning habitat may be related to groundwater–surface water exchange occurring within
these segments. Our results emphasize the importance of valley geomorphology to bull trout, and
our results suggest that prior land use may have adversely affected bull trout populations in the
Swan Basin. Protection of critical spawning tributary catchments from additional road building
and associated land-use disturbance will likely be necessary for the maintenance of viable bull
trout populations in the Swan Basin. Geomorphic context and land-use status of spawning tributaries
are important considerations for future monitoring and management of this species.
 
 
Baxter, C., V. and F. R. Hauer (2000). "Geomorphology, hyporheic exchange, and selection of spawning habitat by bull trout (Salvelinus confluentus)." Canadian Journal of Fisheries and Aquatic Science 57: 1470-1481.
The distribution and abundance of bull trout (Salvelinus confluentus) spawning were affected by geomorphology
and hyporheic groundwater – stream water exchange across multiple spatial scales in streams of the Swan River
basin, northwestern Montana. Among spawning tributary streams, the abundance of bull trout redds increased with increased
area of alluvial valley segments that were longitudinally confined by geomorphic knickpoints. Among all valley
segment types, bull trout redds were primarily found in these bounded alluvial valley segments, which possessed complex
patterns of hyporheic exchange and extensive upwelling zones. Bull trout used stream reaches for spawning that
were strongly influenced by upwelling. However, within these selected reaches, bull trout redds were primarily located
in transitional bedforms that possessed strong localized downwelling and high intragravel flow rates. The changing relationship
of spawning habitat selection, in which bull trout selected upwelling zones at one spatial scale and
downwelling zones at another spatial scale, emphasizes the importance of considering multiple spatial scales within a
hierarchical geomorphic context when considering the ecology of this species or plans for bull trout conservation and
restoration.
 
 
Baxter, C. V. (2001). Fish movement and assemblage dynamics in a Pacific Northwest Riverscape. Corvallis, OR, Oregon State University.
Baxter, C. V. and F. R. Hauer (2000). "Geomorphology, hyporheic exchange, and selection of spawning habitat by bull trout (Salvelinus confluentus)." Can. J. Fish. Aquat. Sci 57: 1470-1481.
Bayley, P. B. and H. W. Li (1999). Riverine Fishes. The River Handbook Vol. 1. P. Calow and G. E. Petts. Oxford, Blackwell Scientific Publications: 251-281.
Beamer, E. M., et al. (2005). Linking watershed conditions to egg-to-fry survival of Skagit chinook salmon. Skagit Chinook Recovery Plan. S. R. S. Cooperative.
Beamer, E. M. and G. Pess (2005). Effects of peak flows on Chinook (Oncorhynchus tshawytscha) spawning success in two Puget Sound river basins. Stillaguamish Watershed Chinook Salmon Recovery Plan. S. I. R. Committee, Snohomish County Department of Public Works: 4.
Bear, J. (1972). Dynamics of Fluids in Porous Media. New York, American Elsevier Publishing Co.
Beatty, D. A. and N. Tatsuaki (1984). "Modeling of River Channel Changes "? 110(2): 262-265.
Beaty, C. B. (1974). "Debris flows, alluvial fans, and a revitalized catastrophism." Z. Geomorph. N.F. 21: 39-51.
Beaty, K. G. (1994). "Sediment transport in a small stream following 2 successive forest-fires." Canadian Journal of Fisheries and Aquatic Sciences 51(12): 2723-2733.
The transport of stream bedload sediment was monitored continuously in a small stream from 1975 to 1982 following forest fires in 1974 and 1980. The stream is located in the east subcatchment (170 ha) of Lake 239 in the Experimental Lakes Area, northwestern Ontario. Precipitation, stream discharge, bedload transport, and concentration of suspended materials were measured quantitatively and organic debris was observed and collected. Bedload transport increased 20-fold following the first fire and threefold after the second. Particle sizes tended to increase during the period of study. Bedload data suggest a recovery period of 5-6 yr following the first fire and a shorter one following the second. A mass budget of material load transported in a single year following recovery indicated a dominance of dissolved load (87%), followed by suspended load (10%), and bedload (3%).
 
Becker, A., and P. Braun (1999). "Disaggregation, aggregation, and spatial scaling in hydrological modelling." 217: 239-252.
Beechie, T. (1988). Stream Classification Literature Search, University of Washington: 5.
Beechie, T., et al. (1995). Restoration of habitat-forming processes in Pacific Northwest watersheds: a locally adaptable approach to aquatic habitat restoration. Proceedings of Society for Ecological Restoration, Seattle, Washington.
Beechie, T., et al. (1994). "Estimating Coho Salmon Rearing Habitat and Smolt Production Losses in a Large River Basin, and Implications for Habitat Restoration." North American Journal of Fisheries Management 14: 797-811.
Beechie, T., et al. (1994). "Estimating coho salmon rearing habitat and smolt production losses in a large river basin, and implications for habitat restoration." North American Journal of Fisheries Management 14: 797-811.
Beechie, t., et al. (2008). Hierarchical physical controls on salmonid spawning location and timing. Salmonid spawning habitat in rivers: physical controls, biological responses, and approaches to remediation. D. A. Sear and P. DeVries. Bethesda, Maryland, American Fisheries Society. Symposium 65: 83-101.
Beechie, T. J. (1998). Rates and pathways of recovery for sediment supply and woody debris recruitment in northwestern Washington streams, and implications for salmonid habitat restoration, University of Washington: 91.
Beechie, T. J. (2001). "Empirical predictors of annual bed load travel distance, and implications for salmonid habitat resotration and protection." Earth Surface Processes and Landforms 26: 1025-1034.
Beechie, T. J. (2001). "Empirical predictors of annual bed load travel distance, and implications for salmonid habitat restoration and protection." Earth Surface Processes and Landforms 26: 1025-1034.
Beechie, T. J. and S. Bolton (1999). "an approach to restoring salmonid habitat-forming processes in Pacific Northwest watersheds." Fisheries 24(4): 6-15.
We present an approach to diagnosing salmonid habitat degradation and restoring habitat-forming
processes that is focused on causes of habitat degradation rather than on effects of degradation.
The approach is based on the understanding that salmonid stocks are adapted to local freshwater conditions
and that their environments are naturally temporally dynamic. In this context, we define a
goal of restoring the natural rates and magnitudes of habitat-forming processes, and we allow for locally
defined restoration priorities. The goal requires that historical reconstruction focus on diagnosing
disruptions to processes rather than conditions. Historical reconstruction defines the suite of restoration
tasks, which then may be prioritized based on local biological objectives. We illustrate the use of this
approach for two habitat-forming processes: sediment supply and stream shading. We also briefly
contrast this approach to several others that may be used as components of a restoration strategy.
 
Beechie, T. J., et al. (2001). "Holocene and Recent Geomorphic Proceses, Land Use, and Salmonid Habitat in two North Puget Sound River Basins." Geomorphic Processes and Riverine Habitat Water Science and Application 4: 37-54.
Beechie, T. J., et al. (2006). "Incorporation parameter uncertainty into evaluation of spawning habitat limitations on Chinook salmon (Oncorhynchus tshawytscha) populations." Canadian Journal of Fisheries and Aquatic Science 63: 1242-1250.
Incorporating parameter uncertainty into a Monte Carlo procedure for estimating spawning habitat capacity
helped determine that spawning habitat availability is unlikely to limit recovery of six populations of Chinook salmon
(Oncorhynchus tshawytscha) in Puget Sound. Spawner capacity estimates spanned up to four orders of magnitude, yet
there was virtually no overlap of distributions of capacity estimates with distributions of current spawner abundance
(<0.2% overlap), except for the Suiattle River population (51% overlap). Empirical distributions of input parameters
contained several important sources of uncertainty, insuring reasonably wide distributions of capacity estimates. The
most defensible ranges of input parameters tended to produce conservative capacity estimates, indicating that increased
model accuracy would only strengthen our conclusion that spawning habitat is not a constraint on these populations.
There are insufficient data with which to develop parameter distributions that better represent historical capacity, which
would certainly be higher than our estimates. Our results suggest that factors other than spawning capacity limit population
size and that recovery efforts for Skagit River Chinook salmon need not focus on spawning habitat restoration
 
Beechie, T. J., et al. (2005). "A classification of habitat types in a large river and their use by juvenile salmonids." Transactions of the American Fisheries Society 134: 717-729.
We describe six habitat types for large rivers (.100 m bank-full width), including
pools, riffles, and glides in midchannel and bank edges, bar edges, and backwaters along channel
margins. Midchannel units were deeper and faster than edge units on average. Among edge habitat
types, backwater units had the lowest velocities and contained complex cover consisting mainly
of wood accumulations and aquatic plants. Banks and bars had similar velocity distributions, but
banks had more complex cover such as rootwads and debris jams. Because sampling of juvenile
salmonids was ineffective in the midchannel units (electrofishing capture efficiency was low, and
the units were too deep and fast to snorkel), we focused our sampling efforts on juvenile salmonid
use of edge habitats during winter, spring, and late summer. Densities of juvenile Chinook salmon
Oncorhynchus tshawytscha and coho salmon O. kisutch were highest in bank and backwater units
in winter, whereas age-0 and age-1 or older steelhead densities were highest in bank units in
winter. In summer, only coho salmon densities were significantly different among edge unit types,
densities being highest in banks and backwaters. Microhabitat selection (velocity, depth, and cover
type) by juvenile salmonids mirrored that in small streams, most fish occupying areas with a
velocity less than 15 cm/s and wood cover. Among ocean-type salmon, Chinook and chum salmon
fry were captured in large numbers in all edge units and exhibited only slightly higher densities
in low-velocity areas (,15 cm/s).
 
Beechie, T. J., et al. (2006). "Channel pattern and river-floodplain dynamics in forested mountain river systems." Geomorphology 78: 124-141.
Channel pattern effectively stratifies the dynamics of rivers and floodplains in forested mountain river systems of the Pacific
Northwest, USA. Straight channels are least dynamic, with relatively slow floodplain turnover and floodplains dominated by old
surfaces. Braided channels are most dynamic, with floodplain turnover as low as 25 years and predominantly young floodplain
surfaces. Island-braided and meandering channels have intermediate dynamics, with moderately frequent disturbances (erosion of
floodplain patches) maintaining a mix of old and young surfaces. Return intervals for the erosion of floodplains increase in the
order of braided, island-braided, meandering, and straight (8, 33, 60, and 89 years, respectively). A threshold for the lateral
migration of a channel occurs at a bankfull width of 15–20 m. The most likely mechanism underlying this threshold is that larger
channels are deep enough to erode below the rooting zone of bank vegetation. Above this threshold, channels not confined between
valley walls exhibit channel patterns distinguishable by slope and discharge, and slope–discharge domains can be used to predict
channel patterns. Meandering and braided patterns are most consistently identified by the model, and prediction errors are largely
associated with indistinct transitions among channel patterns that are adjacent in plots of slope against discharge. Locations of
straight channels are difficult to identify accurately with the current model. The predicted spatial distribution of channel patterns
reflects a downstream decline in channel slope, which is likely correlated with a declining ratio of bed load to suspended load.
Ecological theory suggests that biological diversity should be highest where the intermediate disturbance regime of island-braided
channels sustains high diversity of habitat and successional states through time.
 
Beechie, T. J., et al. (2000). "Modeling recovery rates and pathways for woody debris recruitment in northwestern Washington streams." North American Journal of Fisheries Management 20: 436-452.
We modeled large woody debris (LWD) recruitment and pool formation in northwestern Washington
streams after simulated stand-clearing disturbance using two computer models:
Forest Vegetation Simulator for stand development and Riparian-in-a-Box for LWD recruitment,
depletion, and pool formation. We evaluated differences in LWD recruitment and pool formation
among different combinations of channel size, successional pathway, and stand management scenario.
The models predict that time to first recruitment of pool-forming LWD is about 50% shorter
for red alder Alnus rubra than for Douglas-fir Pseudotsuga menziesii at all channel widths. Total
LWD abundance increases faster in red alder stands than in Douglas-fir stands but declines rapidly
after 70 years as the stand dies and pieces decompose. Initial recovery is slower for Douglas-fir
stands, but LWD recruitment is sustained longer. Total LWD abundance increases faster with
decreasing channel size, and pool abundance increases faster with decreasing channel width and
increasing channel slope. The models predict that thinning of the riparian forest does not increase
recruitment of pool-forming LWD where the trees are already large enough to form pools in the
adjacent channel and that thinning reduces the availability of adequately sized wood. Thinning
increases LWD recruitment where trees are too small to form pools and, because of reduced
competition, trees more rapidly attain pool-forming size. On channels less than 20 m wide, thinning
of red alder and underplanting shade-tolerant conifers will reduce near-term alder recruitment and
increase long-term conifer recruitment. However, the same treatment on channels more than 20
m wide may increase both near-term and long-term recruitment. Compared with the natural fire
regime, timber harvest rotations of 40–80 years during the past century have reduced the percentage
of riparian stands that can provide LWD of pool-forming size to streams, especially in channels
at least 20 m wide.
 
Beechie, T. J., et al. (2000). "Modeling Recovery Rates and Pathways for Woody Debris Recruitment in Northwestern Washington Streams." North American Journal of Fisheries Management 20: 436-452.
Beechie, T. J. and T. Sibley, H. (1997). "Relationships Between Channel Characteristics, Woody Debris, and Fish Habitat in Northwestern Washington Streams." Transactions of the American Fisheries Society 126: 217-229.
Beechie, T. J. and T. H. Sibley (1997). "Relationships between channel characteristics, woody debris, and fish habitat in Northwestern Washington streams." Transactions of the American Fisheries Society 126: 217-229.
Beechie, T. J., et al., Eds. (2003). Ecosystem recovery planning for listed salmon: an integrated assessment approach for salmon habitat. NOAA Tech. Memo. NMFS-NWFSC-58, U.S. Dept. Commerce.
Beeson, C. E. and P. F. Doyle (1995). "Comparison of bank erosion at vegetated and non-vegetated channel bends." Journal of the American Water Resources Association 31(6): 983-990.
Beeson, P. C., S.N.Martens, and D.D. Breshears (2001). "Simulating overland flow following wildfire: mapping vulnerability to landscape disturbance." Hydrological Processes 15: 2917-2930.
Beget, J. E. (1984). "Tephrochronology of Late Wisconsin Deglaciation and Holocene Glacier Fluctuations Near Glacier Peak, North Cascade Range, Washington." Quaternary Research 21: 304-316.
Begin, Z. B. (1982). Application of "diffusion" degradation to some aspects of drainage net development. Badland Geomorphology and Piping. R. Bryan and A. Yain, Geobooks. chapter 9: 169-179.
Begin, Z. B. (1988). "Application of a diffusion-erosion model to alluvial channels which degrade due to base-level lowering." Earth Surface Processes and Landforms 13: 487-500.
Begin, Z. e. B. (1987). "ERFUS - A FORTRAN Program for Calculating the Response of Alluvial Channels to Baselevel Lowering." Computers and Geosciences 13(4): 389-398.
Bell, R. and T. Glade (2004). "Quantitative risk analysis for landslides - Examples from Bíldudalur, NW-Iceland." Natural Hazards and Earth System Sciences 4: 117-131.
Although various methods to carry out quantitative
landslide risk analyses are available, applications are
still rare and mostly dependent on the occurrence of disasters.
In Iceland, two catastrophic snow avalanches killed 34
people in 1995. As a consequence the Ministry of the Environment
issued a new regulation on hazard zoning due to
snow avalanches and landslides in 2000, which aims to prevent
people living or working within the areas most at risk
until 2010. The regulation requires to carry out landslide and
snow avalanche risk analyses, however, a method to calculate
landslide risk adopted to Icelandic conditions is still missing.
Therefore, the ultimate goal of this study is to develop such
a method for landslides, focussing on debris flows and rock
falls and to test it in B´ýldudalur, NW-Iceland.
Risk analysis, beside risk evaluation and risk management,
is part of the holistic concept of risk assessment. Within this
study, risk analysis is considered only, focussing on the risks
to life. To calculate landslide risk, the spatial and temporal
probability of occurrence of potential damaging events, as
well as the distribution of the elements at risk in space and
time, considering also changing vulnerabilities, must be determined.
Within this study, a new raster-based approach is developed.
Thus, all existent vector data are transferred into raster
data using a resolution of 1m×1m. The specific attribute
data are attributed to the grid cells, resulting in specific raster
data layers for each input parameter. The calculation of the
landslide risk follows a function of the input parameters hazard,
damage potential of the elements at risk, vulnerability,
probability of the spatial impact, probability of the temporal
impact and probability of the seasonal occurrence. Finally,
results are upscaled to a resolution of 20m×20m and are
presented as individual risk to life and object risk to life for
each process. Within the quantitative landslide risk analysis
the associated uncertainties are estimated qualitatively.
In the study area the highest risks throughout all of the
analyses (individual risk to life and object risk to life) are
caused by debris flows, followed by rock falls, showing that
risk heavily varies depending on the process considered. The
resultant maps show areas, in which the individual risk to
life exceeds the acceptable risk (defined in the aforementioned
regulation), so that for these locations risk reduction
measures should be developed and implemented. It can be
concluded that the newly developed method works satisfactory
and is applicable to further catchments in Iceland, and
potentially to further countries with different environmental
settings.
 
Bellamy, K., et al. (1992). River morphology, sediments and fish habitats Erosion and Sediment Transport Monitoring Programmes in River Basins, IAHS.
Benavides-Solorio, J., and L.H. MacDonald (2001). "Post-fire runoff and erosion from simulated rainfall on small plots, Colorado Front Range." Hydrological Processes 15(15): 2931-2952.
Wildfires in the Colorado Front Range can trigger dramatic increases in runoff and erosion. A better understanding of the causes of these increases is needed to predict the effects of future wildfires, estimate runoff and erosion risks front prescribed fires. and design effective post-fire rehabilitation treatments. The objective of this project was to determine whether runoff and sediment yields were significantly related to the site variables of burn severity, percent cover, soil water repellency. soil moisture, time since burning, and slope. To eliminate the variability due to natural rainfall events, we applied an artificial storm of approximately 80 mm h(-1) on 26 1 m(2) plots in the summer and fall of 2000, The plots were distributed among a June 2000 wildfire, a November 1999 prescribed fire, and a July 1994 wildfire.For 23 of the 26 plots the ratio of runoff to rainfall exceeded 50%. Nearly all sites exhibited strong natural or fire-induced water repellency, so the runoff ratios were only 15-30% larger for the high-severity plots in the two more recent fires than for the unburned or low-severity plots, The two high-severity plots in the 1994 wildfire had very low runoff ratios., and this probably was due to the high soil moisture conditions at the time of the simulated rainfall and the resulting reduction in the natural water repellency. Sediment yields from the high-severity sites in the two more recent fires were 10-26 times greater than the unburned and low-severity plots. The plots burned at high severity in 1994 yielded only slightly more sediment than the unburned plots. Percent ground cover explained 81% of the variability in sediment yields, and the sediment yields from the plots in the 1994 wildfire are consistent with the observed recovery in percent ground cover. Copyright (C) 2001 John Wiley & Sons, Ltd.
 
Benavides-Solorio, J. (2003). Post-fire runoff and erosion at the plot and hillslope scale, Colorado Front Range.  Ph.D. dissertation. Fort Collins, Colorado, Colorado State University: 218pp.
Benavides-Solorio, J., and L.H. MacDonald (2005). "Measurement and prediction of post-fire erosion at the hillslope scale, Colorado Front Range." International Journal of Wildland Fire 14(1): 1-18.
Benda, L.
Benda, L. (1980). Greenwater Debris Flow Lab #5.
Benda, L. (1988). Debris flows in the Oregon Coast Range. Department of Geological Sciences. Seattle, University of Washington: 140.
Benda, L. (1988). Debris flows in the Tyee Sandstone Formation of the Oregon Coast Range. Dept. of Geological Sciences. Seattle, WA, University of Washington. Master's: 125.
Benda, L. (1990). "The influence of debris flows on channels and valley floors in the Oregon Coast Range, U.S.A." Earth Surface Processes and Landforms 15: 457-466.
Benda, L. (1990). Reconstructing precipitation regimes during the last 10,000 years for the purpose of modelling hillslope erosion in the Pacific Northwest. Seattle, WA, University of Washington.
Benda, L. (1994). Stochastic Geomorphology in a Humid Mountain Landscape. Geological Sciences. Seattle, WA, University of Washington: 356.
Benda, L., and T. Dunne (1997). "Stochastic forcing of sediment supply to channel networks from landsliding and debris flow." Water Resources Research 33(12): 2849-2863.
Benda, L., and D. Miller (2001). "Beyond arm waving: thinking critically at large scales." Watershed Management Council Networker 10(1): 1-16.
Benda, L. (2008). Confluence environments at the scale of river networks. River Confluences, Tributaries and the Fluvial Network. A. R. S. Rice, B. Rhoads. West Sussex, England, John Wiley & Sons, Ltd.: Chapter 13: 271-297.
Benda, L. (2010). "TEST."
Benda, L. (TEST). TEST.
Benda, L., et al. (2004). "Tributary effects in river networks: role of basin scale, basin shape, network geometry, and disturbance regimes." Water Resources Research 40: 1-15.
Benda, L., et al. (1992). "Morphology and Evolution of Salmonid Habitats in a Recently Deglaciated River Basin, Washington State, USA." Canadian Journal of Fisheries and Aquatic Sciences 49(6): 1246-1256.
Benda, L. and T. Dunne (1987). Sediment routing by debris flow. Erosion and Sediment Transport in Pacific Rim Steeplands. Christchurch, IAHS. Publ. no 165: 213-223.
Benda, L. and T. Dunne (1997). "Stochastic forcing of sediment supply to channel networks from landsliding and debris flow." Water Resources Research 33(12): 2849-2863.
Benda, L., et al. (2005). "Geomorphology of Steepland Headwaters: The Transition from Hillslopes to Channels." Journal of American Water Resources Association 41(4): 835-851.
Benda, L., et al. (1992). "The geomorphic structure and evolution of fish habitat in a recently deglaciated river valley, Washington, USA." Canadian Journal Fisheries and Aquatic Sciences 49(6): 1246-1256.
Benda, L. and D. Miller (2002). Landscape dynamics and forest management. General Technical Report RMR@-GTR-101-CD, Rocky Mountain Research Station.
Benda, L., et al. (2003). "Effects of post-wildfire erosion on channel environments, Boise River, Idaho." Journal of Forest Ecology and Management 178: 105-119.
Benda, L., et al. (2004). "Network dynamics hypothesis: spatial and temporal organization of physical heterogeneity in rivers." BioScience 54(5): 413-427.
Benda, L., et al. (accepted). "Network disturbance hypothesis: spatial and temporal organization of physical heterogeneity in rivers." BioScience.
Benda, L., et al. (2009). "Future of applied watershed science at regional scales." EOS, Transaction American Geophysical Union 90(18 (May 5, 2009)): 156-157.
Benda, L., et al. (in press). Wood budgeting: quantitative theory, field practice, and modeling. Wood in World Rivers, American Fisheries Society. Chapter XX.
Benda, L., et al. (2003). Wood Recruitment Processes and Wood Budgeting. The Ecology and Management of Wood in World Rivers. S. V. Gregory, K. Boyer and A. M. Gurnell, American Fisheries Society: 49 - 73.
Benda, L., et al. (2007). "NetMap: A new tool in support of watershed science and resource management." Forest Science 52(2): 206-219.
Benda, L., et al. (2007). "NetMap: A new tool in support of watershed science and resource management." Forest Science 53(2): 206-219.
An integrated suite of numerical models and analysis tools (NetMap) is created for three purposes: (1) Develop regional scale terrain databases in support of watershed science and resource management, (2) Automate numerous kinds of watershed analyses keying on environmental variability for diversifying resource management options, and (3) Improve tools and skills for interpreting watershed-level controls on aquatic systems, including natural disturbance. Hillslope attributes, such as erosion potential, sediment supply, road density, forest age, and fire risk are aggregated down to the channel habitat scale (20-200 m) allowing unique overlap analyses, and they are accumulated downstream in networks revealing patterns across multiple scales. Watershed attributes are aggregated up to subbasin scales (∼10,000 ha), allowing comparative analyses across large watersheds and landscapes. Approximately 25 automated tools address erosion risk, habitat indices, channel classification, habitat core areas, habitat diversity, and sediment and wood supply, among others. Search functions target overlaps between specific hillslope and channel conditions and between roads and landslide or debris flow potential. To facilitate its use, NetMap contains hyperlinked users' manuals and reference materials, including a library of 50 watershed parameters. NetMap provides decision support for forestry, restoration, monitoring, conservation, and regulation
 
Benda, L., et al. (1998). Dynamic Landscape Systems. River Ecology and Management: Lessons from the Pacific Coastal Ecoregion. R. J. Naiman and R. E. Bilby, Springer-Verlag. Chapter 11: 261-288.
Benda, L., et al. (2003). Wood recruitment processes and wood budgeting. the Ecology and Management of Wood in World Rivers. S. V. Gregory, K. L. Boyer and A. M. Gurnell. Bethesda, Maryland, American Fisheries Society. Symposium 37: 49-73.
Benda, L., et al. (in press). "Improving Interdisciplinary Collaborations: Comparative Analysis."
Benda, L., et al. (2002). "How to Avoid Train Wrecks When Using Science in Environmental Problem Solving." BioScience 52(12): 1127-1136.
Benda, L. and G. Reeves (1999). Stochastic and Deterministic Origins of Aquatic Habitats. 1999 Meeting, San Francisco, CA, American Geophysical Union.
Benda, L. and J. Sias (1998). Landscape controls on wood abundance in streams. Seattle, WA, Earth Systems Institute: 60.
Benda, L. and J. Sias (2003). "A quantitative framework for evaluating the mass balance of wood in streams." Journal of Forest Ecology and Management 172: 1-16.
Benda, L., et al. (2003). "Debris flows as agents of morphological heterogeneity at low-order confluences, Olympic Mountains, Washington." Geological Society of America Bulletin 115(9): 1110-1121.
Benda, L., et al. (1997). Slope Instability and Forest Land Managers. Seattle, Washington, Earth Systems Institute.
Benda, L., et al. (1997). Slope Instability and Forest Land Managers: A Primer and Field Guide, Earth Systems Institute.
Benda, L. E. (1985). Behavior and effect of debris flows on streams in the Oregon Coast Range. Delineation of Landslide, Flash Flood, and Debris Flow hazards in Utah. D. S. Bowles: 153-162.
Benda, L. E. (1988). Debris flows in the Oregon Coast Range. Department of Geological Sciences. Seattle, Washington, University of Washington: 127.
Benda, L. E. (1990). "The influence of debris flows on channels and valley floors in the Oregon Coast Range, U.S.A." Earth Surface Processes and Landforms 15: 457-466.
Benda, L. E. (1991). "Predicting downstream impacts from slope failures."
Benda, L. E., and T. Dunne (1997). "Stochastic forcing of sediment routing and storage in channel networks." Water Resources Research 33(12): 2865-2880.
Benda, L. E. (1999). Science vs. Reality: The Scale Crisis in the Watershed Sciences. Annual Wildland Hydrology Conference, Bozeman, Montana.
Benda, L. E. (?). The Role of Science in Resolving Environmental Conflicts: Recent Failures Point to the Next Step - Ecological Risk Assessment. Seattle, Washington, Earth Systems Institute.
Benda, L. E. (?). Watershed Analysis: A Sound Idea Limited by Poor Formulations and the Opportunity of NMFS to Correct the Problem. Seattle, Washington, Earth Systems Institute: 4.
Benda, L. E. (?). Watershed Analysis: How Good is the Science? Seatle, Washington.
Benda, L. E., et al. (2004). "Confluence effects in rivers: Interactions of basin scale, network geometry, and disturbance regimes." Water Resources Research 40: W05402.
Benda, L. E., et al. (1992). "Morphology and evolution of salmonid habitats in a recently deglaciated river basin, Washington State, U.S.A." Canadian Journal of Fisheries and Aquatic Sciences 49(6): 1246-1256.
Benda, L. E., et al. (1992). "Morphology and evolution of salmonid habitats in a recently deglaciated river basin, Washington state, USA." Canadian Journal of Fisheries and Aquatic Science 49: 1246-1256.
Benda, L. E., et al. (2002). "Recruitment of wood to streams in old-growth and second-growth redwood forests, northern California, U.S.A." Canadian Journal of Fisheries and Aquatic Sciences 32: 1460-1477.
Benda, L. E. and T. W. Cundy (1990). "Predicting deposition of debris flows in mountain channels." Canadian Geotechnical Journal 27: 409-417.
Benda, L. E. and T. Dunne (1987). "Sediment routing by debris flow." International Association of Hydrological Sciences 165: 213-223.
Benda, L. E. and T. Dunne (1997). "Stochastic forcing of sediment routing and storage in channel networks." Water Resources Research 33(12): 2865-2880.
Benda, L. E. and T. Dunne (1997). "Stochastic forcing of sediment supply to channel networks from landsliding and debris flow." Water Resources Research 33(12): 2849-2863.
Benda, L. E. and T. Dunne (1997). "Stochiastic forcing of sediment routing and storage in channel networks." Water Resources Research 33(12): 2865-2880.
Benda, L. E., et al. (1987). Influences of forest management on channel environments in the Oregon Coast Range. Oregon Riparian Symposium.
Benda, L. E. and D. J. Miller (2001). "Beyond Arm Waving: Thinking Critically at Large Scales." Watershed Management Council Networker 10(1): 1, 4-16.
Benda, L. E., et al. (2011). "Creating a catchment scale perspective for river restoration." Hydrology and Earth System Science 15.
Benda, L. E., et al. (2003). "Effects of post-wildfire erosion on channel environments, Boise River, Idaho." Forest Ecology and Management 178: 105-119.
Benda, L. E., et al. (2003). "Effects of post-wildlife erosion on channel environments Boise River, Idaho." Forest Ecology and Management 178: 105-119.
Benda, L. E., et al. (1999). "Fires and geomorphology in mountain drainage basins; theoretical predictions and field observations." Abstracts with Programs - Geological Society of America 31(7): 313.
The ubiquitous occurrence of charcoal in lake core sediments, landslide sites, debris flow deposits, alluvial terraces, and spatial patterns of forest vegetation suggest how fires are linked to geomorphology in the Pacific Northwest region of North America. The empirical record, however compelling, is presently too fragmentary to deduce general principles on the role of fires in geomorphic regimes. A stochastic simulation model of erosion and sedimentation, built upon slope stability and sediment transport theories, and on empirical studies of more complicated processes including fire behavior and debris flows, was developed for the Oregon Coast Range and for the southern Cascades of Washington. In these terrains, fires are predicted to result in a punctuated supply of sediment and wood by landslides and debris flows, a pattern circumscribed by the size and frequency of fires, the number of erosion source areas, soil production rates, and the geometry of the low-order channel network. In higher-order channel networks, fires can influence the spectra of sediment flux and storage depending on basin scale, particle attrition rates, sediment transport rates, and on availability of sediment storage reservoirs on the valley floor. Periodic erosion and sedimentation can influence the spatial distribution and volume of boulders and alluvium in channels, patterns that are documented in field studies. Fire-induced erosion can also lead to the genesis of fans and terraces. Variations in topography, lithology, and in fire regimes across mountainous regions of North America can be expected to result in different spatial and temporal patterns of punctuated erosion and sedimentation. Fire-related disturbance regimes can be studied through a combination of field studies and theory development.
 
Benda, L. E., et al. (1998). Dynamic Landscape Systems. River Ecology and Management. R. J. Naiman and R. E. Bilby. New York, Springer-Verlag: 261-288.
Benda, L. E., et al. (1998). Analysis of Natural Disturbance Murray Pacific Tree Farm: Consequences for Environmental Assessment and Forest Management. Seattle, Washington, Earth Systems Institute: 1-50.
Benda, L. E. and L. R. Miller (1991). Geomorphical Watershed Analysis: A Conceptual Framework and Review of Techniques. Seattle, Washington, Department of Geological Sciences: Mountain Drainage Basin Geomorphology Group: 54.
Benda, L. E., et al. (2004). "The network dynamics hypothesis: how channel networks structure riverine habitats." BioScience 54(5): 413-427.
Hierarchical and branching river networks interact with dynamic watershed disturbances, such as fires, storms, and floods, to impose a spatial and
temporal organization on the nonuniform distribution of riverine habitats, with consequences for biological diversity and productivity. Abrupt
changes in water and sediment flux occur at channel confluences in river networks and trigger changes in channel and floodplain morphology. This
observation, when taken in the context of a river network as a population of channels and their confluences, allows the development of testable predictions
about how basin size, basin shape, drainage density, and network geometry interact to regulate the spatial distribution of physical diversity
in channel and riparian attributes throughout a river basin. The spatial structure of river networks also regulates how stochastic watershed disturbances
influence the morphology and ages of fluvial features found at confluences.
 
Benda, L. E., et al. (2002). "How to avoid train wrecks when using science in environmental problem solving." BioScience 52(12): 1127-1136.
Benda, L. E., et al. (2002). "How to Avoid Train Wrecks When Using Science in Environmental Problem Solving." BioScience 52(12): 1127-1136.
Benda, L. E. and J. C. Sias (1998). Landscape Controls on Wood Abundance in Streams. Seattle, WA, Earth Systems Institute: 59.
Benda, L. E. and J. C. Sias (2003). "A quantitative framework for evaluating the mass balance of in-stream organic debris." Forest Ecology and Management 172: 1-6.
Benda, L. E., et al. (1988). Report of the I. D. Team: Investigation of the Hazel Landslide of the North Fork of the Stillaguamish River: 1-12.
Benda, L. E., et al. (2003). "Debris flows as agents of morphological heterogeneity at low-order confluences, Olympic Mountains, Washington." Geological Society of America Bulletin 115(9): 1110-1121.
Benda, L. E. and W. Zhang (1990). Accounting for the stochiastic occurrence of landslides when predicting sediment yields. Research Needs and Applications to Reduce Erosion and Sedimentation in Tropical Steeplands, Fiji.
Benke, A. C. (2001). "Importance of flood regime to invertebrate habitat in an unregulated river-floodplain ecosystem " Journal of the North American Benthological Society 20(2): 225-240.
Benke, A. C., et al. (2000). "Flood pulse dynamics of an unregulated river floodplain in the southeastern U.S. Coastal Plain." Ecology 81(10): 2730-2741.
Benner, P. A., et al. (1988). From the Forest to the Sea: A Story of Fallen Trees. Portland, Oregon, U.S. Forest Service: 154.
Benner, P. A. and J. R. Sedell (1987). Chronic Reduction of Large Woody Debris on Beaches at Oregon River Mouths. Wetland and Riparian Ecosystems of the American West: 8th Annual Meeting of the Society of Wetland Scientists, Seattle, Washington.
Bennet, J. P. (1974). "Concepts of mathematical modelling of sediment yield." Water Resources Research 10: 485-493.
Bennett, D. A., M.P.Armstrong, and G.A.Wade (1999). "Exploring the solution space of semi-structured geographical problems using genetic algorithms." Transactions in GIS 3(1): 51-71.
Bent, G. C. and P. A. Steeves (2006). "A revised logistic regression equation and an automated procedure for mapping the probability of a stream flowing perennially in Massachusetts: U.S. Geological Survey Scientific Investigations Report 2006-5031." 107.
A Revised Logistic Regression Equation and an Automated Procedure for Mapping the Probability of a Stream Flowing Perennially in Massachusetts
By Gardner C. Bent and Peter A. Steeves
A revised logistic regression equation and an automated procedure were developed for mapping the probability of a stream flowing perennially in Massachusetts. The equation provides city and town conservation commissions and the Massachusetts Department of Environmental Protection a method for assessing whether streams are intermittent or perennial at a specific site in Massachusetts by estimating the probability of a stream flowing perennially at that site. This information could assist the environmental agencies who administer the Commonwealth of Massachusetts Rivers Protection Act of 1996, which establishes a 200-foot-wide protected riverfront area extending from the mean annual high-water line along each side of a perennial stream, with exceptions for some urban areas. The equation was developed by relating the observed intermittent or perennial status of a stream site to selected basin characteristics of naturally flowing streams (defined as having no regulation by dams, surface-water withdrawals, ground-water withdrawals, diversion, wastewater discharge, and so forth) in Massachusetts. This revised equation differs from the equation developed in a previous U.S. Geological Survey study in that it is solely based on visual observations of the intermittent or perennial status of stream sites across Massachusetts and on the evaluation of several additional basin and land-use characteristics as potential explanatory variables in the logistic regression analysis. The revised equation estimated more accurately the intermittent or perennial status of the observed stream sites than the equation from the previous study.
Stream sites used in the analysis were identified as intermittent or perennial based on visual observation during low-flow periods from late July through early September 2001. The database of intermittent and perennial streams included a total of 351 naturally flowing (no regulation) sites, of which 85 were observed to be intermittent and 266 perennial. Stream sites included in the database had drainage areas that ranged from 0.04 to 10.96 square miles. Of the 66 stream sites with drainage areas greater than 2.00 square miles, 2 sites were intermittent and 64 sites were perennial. Thus, stream sites with drainage areas greater than 2.00 square miles were assumed to flow perennially, and the database used to develop the logistic regression equation included only those stream sites with drainage areas less than 2.00 square miles. The database for the equation included 285 stream sites that had drainage areas less than 2.00 square miles, of which 83 sites were intermittent and 202 sites were perennial.
Results of the logistic regression analysis indicate that the probability of a stream flowing perennially at a specific site in Massachusetts can be estimated as a function of four explanatory variables: (1) drainage area (natural logarithm), (2) areal percentage of sand and gravel deposits, (3) areal percentage of forest land, and (4) region of the state (eastern region or western region). Although the equation provides an objective means of determining the probability of a stream flowing perennially at a specific site, the reliability of the equation is constrained by the data used in its development. The equation is not recommended for (1) losing stream reaches or (2) streams whose ground-water contributing areas do not coincide with their surface-water drainage areas, such as many streams draining the Southeast Coastal Region-the southern part of the South Coastal Basin, the eastern part of the Buzzards Bay Basin, and the entire area of the Cape Cod and the Islands Basins. If the equation were used on a regulated stream site, the estimated intermittent or perennial status would reflect the natural flow conditions for that site.
An automated mapping procedure was developed to determine the intermittent or perennial status of stream sites along reaches throughout a basin. The procedure delineates the drainage area boundaries, determines values for the four explanatory variables, and solves the equation for estimating the probability of a stream flowing perennially at two locations on a headwater (first-order) stream reach-one near its confluence or end point and one near its headwaters or start point. The automated procedure then determines the intermittent or perennial status of the reach on the basis of the calculated probability values and a probability cutpoint (a stream is considered to flow perennially at a cutpoint of 0.56 or greater for this study) for the two locations or continues to loop upstream or downstream between locations less than and greater than the cutpoint of 0.56 to determine the transition point from an
intermittent to a perennial stream. If the first-order stream reach is determined to be intermittent, the procedure moves to the next downstream reach and repeats the same process. The automated procedure then moves to the next first-order stream and repeats the process until the entire basin is mapped.
A map of the intermittent and perennial stream reaches in the Shawsheen River Basin is provided on a CD-ROM that accompanies this report. The CD-ROM also contains ArcReader 9.0, a freeware product, that allows a user to zoom in and out, set a scale, pan, turn on and off map layers (such as a USGS topographic map), and print a map of the stream site with a scale bar. Maps of the intermittent and perennial stream reaches in Massachusetts will provide city and town conservation commissions and the Massachusetts Department of Environmental Protection with an additional method for assessing the intermittent or perennial status of stream sites.
 
Berelson, W. L., et al. (2004). "Mapping hydrologic units for the national Watershed Boundary Dataset." Journal of the American Water Resources Association 40(5): 1231-1246.
ABSTRACT: In 2002, Wyoming became the first state to complete
development of a statewide 1:24,000-scale Watershed Boundary
Dataset (WBD) under the new Federal Standards for Delineation
of Hydrologic Unit Boundaries. The product was developed through
the coordinated efforts of numerous state, federal, and local entities
both within Wyoming and in neighboring states. Development of a
comprehensive, standardized hydrologic unit boundary dataset in a
“headwaters” state such as Wyoming poses a number of unique
challenges. This paper details the WBD’s development in Wyoming,
highlighting technical methodology development and interagency
coordination strategies. Evolution of the WBD standard is
reviewed, addressing inconsistencies between definitions for hydrologic
units and “true” watershed delineations. While automated
methods are improving, manual and semi-automated techniques
continue to serve as valuable approaches to hydrologic unit boundary
delineation given the quality of digital terrain models and the
multijurisdictional nature of watershed based management. This
case study provides insight on future development and maintenance
of the WBD within and across other states and regions of the
country and on opportunities for linking the WBD to related water
resource geospatial data products like the National Hydrography
Dataset.
 
Berg, D. R. (1995). "Riparian silvicultural system design and assessment in the Pacific Northwest Cascade Mountains, USA." Ecological Applications 5(1): 87-96.
Berg, N., et al. (1998). "Function and dynamics of woody debris in stream reaches in the central Sierra Nevada, California." Canadian Journal of Fisheries and Aquatic Sciences 55: 1807-1820.
Berger, A. M. and R. E. Gresswell (2009). "Factors influencing coastal cutthroat trout (Oncorhynchus clarkii clarkii) seasonal survival rates: a spatially continuous approach within stream networks." Canadian Journal of Fisheries and Aquatic Science 66: 613-632.
Mark–recapture methods were used to examine watershed-scale survival of coastal cutthroat trout (Oncorhynchus
clarkii clarkii) from two headwater stream networks. A total of 1725 individuals (‡100 mm, fork length) were individually
marked and monitored seasonally over a 3-year period. Differences in survival were compared among spatial
(stream segment, subwatershed, and watershed) and temporal (season and year) analytical scales, and the effects of abiotic
(discharge, temperature, and cover) and biotic (length, growth, condition, density, movement, and relative fish abundance)
factors were evaluated. Seasonal survival was consistently lowest and least variable (years combined) during autumn
(16 September – 15 December), and evidence suggested that survival was negatively associated with periods of low stream
discharge. In addition, relatively low (–) and high (+) water temperatures, fish length (–), and boulder cover (+) were
weakly associated with survival. Seasonal abiotic conditions affected the adult cutthroat trout population in these watersheds,
and low-discharge periods (e.g., autumn) were annual survival bottlenecks. Results emphasize the importance of
watershed-scale processes to the understanding of population-level survival.
 
Berghe, E. P. v. d. and M. R. Gross (1984). "Female Size and Nest Depth in Coho Salmon (Oncorhynchus kisutch) " Canadian Journal of Fisheries and Aquatic Sciences 41: 204-206.
Berghe, E. P. v. d. and M. R. Gross (1989). "Natural Selection Resulting From Female Breeding Competition in a Pacific Salmon (Coho: Oncorhynchus Kistuch) " Evolution 43(1): 125-140.
Bergstedt, L. C. and E. P. Bergersen (1997). "Health and movements of fish in response to sediment sluicing in the Wind River, Wyoming." Canadian Journal of Fisheries and Aquatic Sciences 54: 312-319.
Bergstrom, F. W. and S. A. Schumm (1981). Episodic behaviour in badlands. Erosion and Sediment Transport in Pacific Rim Steeplands, I.A.H.S.: 478-493.
Berris, S. N. and D. R. Harr (1987). "Comparative Snow Accumulation and Melt During Rainfall in Forested and Clear-cut Plots in the Western Cascades of Oregon." Water Resources Research 23(1): 135-142.
Berry, J. D. and J. R. Maxwell (1981). Land systems inventory for the Pacific Northwest Region and the Siuslaw National Forest. Corvallis, OR, Siuslaw National Forest: 22.
Bertolo, P. and F. Wieczorek (2005). "Calibration of numerical models for small debris flows in Yosemite Valley, California, USA." Natural Hazards and Earth System Sciences 5: 993-1001.
This study compares documented debris flow
runout distances with numerical simulations in the Yosemite
Valley of California, USA, where about 15% of historical
events of slope instability can be classified as debris flows
and debris slides (Wieczorek and Snyder, 2004).
To model debris flows in the Yosemite Valley, we selected
six streams with evidence of historical debris flows; three of
the debris flow deposits have single channels, and the other
three split their pattern in the fan area into two or more channels.
From field observations all of the debris flows involved
coarse material, with only very small clay content.
We applied the one dimensional DAN (Dynamic ANalysis)
model (Hungr, 1995) and the two-dimensional FLO-
2D model (O’Brien et al., 1993) to predict and compare the
runout distance and the velocity of the debris flows observed
in the study area. As a first step, we calibrated the parameters
for the two softwares through the back analysis of three
debris- flows channels using a trial-and-error procedure starting
with values suggested in the literature. In the second step
we applied the selected values to the other channels, in order
to evaluate their predictive capabilities.
After parameter calibration using three debris flows we obtained
results similar to field observationsWe also obtained a
good agreement between the two models for velocities. Both
models are strongly influenced by topography: we used the
30m cell size DTM available for the study area, that is probably
not accurate enough for a highly detailed analysis, but it
can be sufficient for a first screening.
 
Beschta, B. (1997). "Riparian shade and stream temperature." Rangelands 19(2): 25-28.
Beschta, R. L. (1978). "Long-term Patterns of Sediment Production Following Road Construction and Logging in the Oregon Coast Range." Water Resources Research 14(6): 1011-1016.
Beschta, R. L. (1978). "Long-term patterns of sediment production following road construction and logging in the Oregon Coastal Range." Water Resources Research 14(6): 1011-1016.
Suspended sediment production after road construction, logging and slash disposal was significantly increased (p=0.95) on two watersheds in Oregon's Coast Range. A 25% patch cut watershed showed increases during 3-8 posttreatment years. these increases were caused primarily by mass soil erosion from roads. Monthly sediment concentrations before the occurrence of the annual peak  flow were increased more than those following the annual peak. Surface erosion from a severe slash burnwas the primary cause of increased sediment yields for 5 post treatment years on a watershed that was 82% cleacut. Monthly sediment concentrations were generally increased throughout the winter runoff period on this watershed.  The flushing of suspended sediment in Oregon Coast Range watersheds is apparent from seasonal changes of suspended sediment rating curves.
 
Beschta, R. L. (1981). Patterns of sediment and organic-matter transport in Oregon Coast Range streams. Symposium on Erosion and Sediment Transport in Pacific Rim Steeplands, IAHS. 132: 179-188.
Beschta, R. L. (1983). The effects of large organic debris upon channel morphology: a flume study. Symposium on Erosion and Sedimentation, Simons, Li and Associates, Fort Collins, CO.
Beschta, R. L. (1984). River Channel Response to Accelerated Mass Soil Erosion. Symposium on the Effects of Forest Land Use on Erosion and Slope Stability, Honolulu, Hawaii.
Beschta, R. L. (1984). River channel response to acclerated mass soil erosion. Proceedings, Symposium on Effects of Forest Land Use on Erosion and Slope Stability. Honolulu, East-West Center: 155-164.
Beschta, R. L. (1984). River channel responses to accelerated mass soil erosion. Symposium on Effects of Forest Land Use on Erosion and Slope Stability. Honolulu, Hawaii.
Long-term channel adjustments to accelerated hillslope erosion were evaluated from aerial photographs for two river systems (i.e., the Middle fork Willamette River, USA, and the Kowai River, NZ) which drain steep mountainous terrain. Hillslope erosion rates in both basins have been altered by landuse practices and the occurrence of an extremely large flow event (recurrence interval > 100 years). Results indicate that active channel width is a sensitive indicator of channel aggradation, channel adjustments are most pronounced at or immediately downstream of sediment inputs, and channel changes caused by sediment inputs are relatively persistent. Sediment, and not peak flow per se, caused the most pronounced changes in channel characteristics.
 
Beschta, R. L. (1998). "Forest Hydrology in the Pacific Northwest: Additional Research Notes." Journal of the American Water Resources Association 34(4): 729-741.
Beschta, R. L. and W. L. Jackson (1979). "The Intrusion of Fine Sediments into a Stable Gravel Bed." Journal of the Fisheries Research Board of Canada 36(2): 204-210.
Beschta, R. L. and A. R. Orme (1990). Upper Stiltner Creek, Lewis County, Washington. D. D. Zender.
Beschta, R. L. and W. S. Platts (1986). "Morphological Features of Small Streams: Signifigance and Function." Water Resources Bulletin 22(3): 369-379.
Beschta, R. L. and W. J. Ripple (2008). "Wolves, trophic cascades, and rivers in the Olympic National Park, USA." Ecohydrology 1: 118-130.
Gray wolves (Canis lupus) were extirpated in the early 1900s from the Olympic Peninsula of northwestern Washington.
Thus, we studied potential cascading effects of wolf removal by undertaking a retrospective study of Roosevelt elk (Cervus
elaphus) populations, riparian forests, and river channel morphology. For three riparian sites within the western portion of
Olympic National Park, the age structure of black cottonwood and bigleaf maple indicated a pattern of significantly decreased
recruitment (growth of seedlings/sprouts into tall saplings and trees) associated with intensive elk browsing in the decades
following the loss of wolves. At a riparian site outside the park, which represented a refugium from elk browsing, cottonwood
recruitment has been ongoing during the 20th century, indicating that climate and flow regimes, in the absence of intensive
herbivory, have not limited the establishment and growth of this deciduous woody species. Using 1994 orthophotos, we also
measured channel dimensions and planform morphology of 8-km-long river reaches at each vegetation sampling site and an
additional reach outside the park. Channels inside the park versus those outside the park had greater percent braiding (37 vs
2%) and larger ratios of active channel width/wetted width (3Ð0 vs 1Ð5 m/m). Results for western Olympic National Park were
consistent with a truncated trophic cascade hypothesis whereby ungulate browsing following the extirpation of wolves caused
significant long-term impacts to riparian plant communities which, in turn, allowed increased riverbank erosion and channel
widening to occur.
 
Besctha, J. L., et al. (1987). Stream temperature and aquatic habitat. Streamside Management: Forestry and Fishery Interactions, University of Washington, Seattle, WA, Institute of Forest Resour4ces.
Best, J. L. (1986). "The morphology of river channel confluences." Progress in Physical Geography 10: 157-174.
Best, J. L. (1986). "Sediment transport and bed morphology at river channel confluences." Sedimentology 35(481-498).
Best, J. L. (1988). "Sediment transport and bed morphology at river channel confluences." Sedimentology 35: 481-498.
Bettinger, P. and M. Lennette (2004). LAndscape Management Policy Simulator (LAMPS) Version 1.1, User Guide. Corvallis, OR, Oregon State University.
Bettinger, P., et al. (2005). "A hierarchical spatial framework for forest landscape planning." Ecological Modeling 182: 25-48.
A hierarchical spatial framework for large-scale, long-term forest landscape planning is presented along with example policy
analyses for a 560,000 ha area of the Oregon Coast Range. The modeling framework suggests utilizing the detail provided by
satellite imagery to track forest vegetation condition and for representation of fine-scale features, such as riparian areas. Spatial
data are then aggregated up to management units, where forest management decisions are simulated. Management units may also
be aggregated into harvest blocks to closer emulate management behavior. Land allocations, subdivisions of landowner groups,
can be used to represent different levels of management. A management unit may contain multiple land allocations, such as
riparian management emphases that vary based on distance from the stream system. The management emphasis required by each
land allocation is retained in the simulation of policies. When applied within a large-scale forest landscape planning context, the
implications of policies that suggest clearcut size restrictions, minimum harvest ages, or the development of interior habitat areas
can be assessed. Simulations indicated that the minimum harvest age constraint has a stronger influence on even-flow harvest
levels than do maximum clearcut size or interior habitat area constraints. Even-flow timber harvest level objectives, however,
also have an effect on the results: time periods beyond the constraining time period show a build-up of timber inventory, which
suggests a possible relaxation or modification of the objective in order to achieve average harvest ages that are closer to the
minimum harvest age.
 
Beuselinck, L., et al. (2002). "The influence of rainfall on sediment transport by overland flow over areas of net deposition." Journal of Hydrology 257(1-4): 145-163.
Beuselinck, L., et al. (1999). "Evaluation of the simple settling theory for predicting sediment deposition by overland flow." Earth Surface Processes and Landforms 24(11): 993-1007.
Beven, K. (1981). The effect of ordering on the geomorphic effectiveness of hydrologic events. Erosion and Sediment Transport in Pacific Rim Steeplands. Christchurch, I.A.H.S. Publication Number 132.
Beven, K. (1987). "Towards the Use of Catchment Geomorphology in Flood Frequency Predictions." Earth Surface Processes and Landforms 12: 69-82.
Beven, K. J. (1993). "Prophecy, reality and uncertainty in distributed hydrological modeling." Advances in Water Resources 16: 41-51.
Beven, K. J. (2001). "How far can we go in distributed hydrological modelling." Hydrology and Earth System Sciences 5(1): 1-12.
Beven, K. J. (2002). "Towards a coherent philosophy for modelling the environment." Proceedings of the Royal Society, London, A 458: 1-20.
Beven, K. J. (2002). "Towards an alternative blueprint for a physically based digitally simulated hydrologic response modelling system." Hydrological Processes 16: 189-206.
Beven, K. J. (2005). "A manifesto for the equifinality thesis." Journal of Hydrology xx: 1-19.
Beven, K. J., et al. (1988). "On Hydrological Heterogeneity - Catchment Morphology and Catchment Response." Journal of Hydrology 100: 353-375.
Beverage, J. P. and J. K. Culbertson (1964). "Hyperconcentrations of Suspended Sediment." Journal of the Hydraulics Division 6: 117-128.
Bevers, M., et al. (1996). "Spatially optimizing forest management schedules to meet stormflow constraints." Water Resources Bulletin 32(5): 1007-1015.
A spatial optimization model is developed and used to limit cumulative effects resulting from storm events by strategically arranging and scheduling forest treatments to meet peak storm-flow constraints. A mixed integer forest management scheduling formulation is used to select the location and timing of forest treatments. The approach includes simulated spatial routing of storm-flows imbedded as hydrologic process constraints (in a nested schedule) within longer-term forest management planning periods in the mathematical programming model. Although difficult to solve, the model shows promise for further research.
 
Bevin, K. (1981). "Kinematic subsurface stormflow." Water Resources Research 17(5): 1419-1424.
Bevin, K. (1982). "On subsurface stormflow: predictions with simple kinematic theory for saturated and unsaturated flows." Water Resources Research 18(6): 1627-1633.
Bevington, P. R. (1969). Data Reduction and Error analysis for the Physical Sciences. New York, McGraw-Hill.
Bigelow, P., et al. (2007). "On debris flows, river networks, and the spatial structure of channel morphology." Forest Science 52(2): 220-238.
Bigelow, P., et al. (2001). Bolinas Lagoon Watershed Study Input Sediment Budget San Francisco, California, Prepared by Tetra Tech, Inc. for US Army Corps of Engineers: 77.
Bigelow, P. E., et al. (2007). "On debris flows, river networks, and the spatial structure of channel morphology." Forest Science 53(2): 220-238.
We evaluated the morphological effects of debris flows from headwater streams in larger, fish-bearing channels of the central Oregon Coast Range, including their influence on fans, wood recruitment, and channel morphology. Continuous channel surveys (6.4 km) were conducted in third- through fifth-order streams (drainage area <10 km2 and slope <7%) where debris fan effects at confluences were most evident. This basin size contains the majority of channels (67%) in the central Coast Range with gradients that are used by coho salmon (Oncorhynchus kisutch Walbaum). The close spacing between headwater tributaries susceptible to debris flows (118 m average) resulted in long continuous sections of fish-bearing streams that were bordered by debris fans (103 m average) and debris fans impinging on 54% of the total channel length surveyed. Debris flows also supplied the majority of wood (58% of pieces) to the surveyed fish-bearing channels. The highest values of large wood, boulders, and channel gradients were associated with debris fans at confluences with headwater tributaries, while deeper sediment deposits were often associated with fans but also extended up and downstream from fans. The spacing and network pattern of debris flow-prone headwater tributaries influenced the spatial structure of channel morphology and aquatic habitats leading to a high degree of physical heterogeneity and patchiness in channel environments. Our study contributes to a growing emphasis on the role of tributary confluences in structuring channel morphology and aquatic habitats in mountain drainage basins and argues for including a confluence component to stream classification and habitat typing schemes
 
Bilby, R. E. (1981). "Role of organic debris dams in regulating the export of dissolved and particulate matter from a watershed." Ecology 62: 1234-1243.
Bilby, R. E. (1984). "Removal of Woody Debris May Affect Stream Channel Stability." Journal of Forestry 82(10): 609-612.
Bilby, R. E. (1985). "Contributions of Road Surface Sediment to a Western Washington Stream." Forest Science 31(4): 827-838.
Bilby, R. E., K. Sullivan, and S.H. Duncan (1989). "The generation and fate of road-surface sediment in forested watersheds in southwestern Washington." Forest Science 35(2): 453-468.
Bilby, R. E. and P. A. Bisson (1998). Function and Distribution of Large Woody Debris. River Ecology and management. Lessons from the Pacific Coastal Ecoregion. R. J. Naiman and R. E. Bilby. New York, Springer: 324-346.
Bilby, R. E. and P. A. Bisson (2004). Function and Distribution of Large Woody Debris. River Ecology and Management. R. J. Naiman and R. E. Bilby. New York, Springer: 324-346.
Bilby, R. E., et al. (1996). "Incorporation of nitrogen and carbon from spawning coho salmon into the trophic system of small streams: evidence from stable isotopes." Fish. Aquat. Sci. 53: 164-173.
Bilby, R. E., et al. (1999). "Effects of immersion in water on deterioration of wood from five species of trees used for habitat enhancement projects." North American Journal of Fisheries Management 19: 687-695.
Logs of standard dimensions from five species of trees were submerged in a stream
to evaluate changes in strength and decomposition over a period of 5 years. Changes in structural
properties occurred only for wood near the outer surface of the logs. Nearly all bark was removed
from the logs within 12 months. Diameter loss for the five species ranged from 10.6 mm (western
hemlock Tsuga heterophylla) to 21.8 mm (bigleaf maple Acer macrophyllum) after 5 years. De-creases
in the density of surface wood for the five species ranged from 23% (red alder Alnus rubra)
to 31% (western hemlock). Modulus of rupture, modulus of elasticity, and wood density did not
change for wood more than 12 mm from the log surface for any of the species. Bigleaf maple
exhibited the highest resistance to rupture, and western redcedar Thuja plicata exhibited the lowest.
Western redcedar was also the most easily flexed. Microbial activity on the surface of the logs
was highest at the start of the experiment and decreased rapidly with time of immersion. The two
hardwood species (bigleaf maple and red alder) generally had higher levels of microbial activity
than the conifer species (Douglas fir Pseudotsuga menzesii, western hemlock, western redcedar)
from 12 months through 60 months of immersion. Differences in the rate of decomposition between
conifer and hardwood logs were much less than in terrestrial environments. Our results suggest
that hardwood logs can be used in stream enhancement projects where the wood will be submerged.
 
Bilby, R. E., et al. (?). "Effects of Stream Submersion on Deterioration of Wood from Five Tree Species: Implications for Habitat Enhancement Projects (DRAFT)."
Bilby, R. E. and G. E. Likens (1980). "Importance of organic debris jams in the structure and function of stream ecosystems." Ecology 61: 1107-1113.
Bilby, R. E. and L. A. Mollot (2008). "Effect of changing land use patterns on the distribution of coho salmon (Oncorhynchus kisutch) in the Puget Sound region." Canadian Journal of Fisheries and Aquatic Science 65: 2138-2148.
Population increase in the Pacific Northwest of North America over the last century has led to the removal of
forests for various purposes. Evidence of salmon response to these alterations in land use is rare owing to a scarcity of
fish population data and a high degree of interannual variation in abundance. We examined the relationship between the
spatial distribution of spawning coho salmon (Oncorhynchus kisutch) and changes in land use from 1984 through 2001 at
84 sites in four rivers draining into northern Puget Sound. Changes in land use over this period were determined from
LandSat imagery, county zoning designations, and aerial photographs. Substantial reduction in forest cover occurred in
many of the index watersheds during this time. The proportion of salmon using sites subjected to increased urban land use
over the study period declined about 75%. Increases were observed at forested sites and those with increased rural residential
use. Maintaining salmon populations in rapidly developing areas may require the identification and protection of sites
that support large salmon populations and steering development to areas supporting few fish.
 
 
Bilby, R. E., et al. (2003). Sources of variability in aquatic ecosystems: factors controlling biotic production and diversity. Strategies for Restoring River Ecosystems: Sources of Variability and Uncertainty in Natural and Managed Systems. R. C. Wissmar and P. A. Bisson. Bethesda, Maryland, American Fisheries Society: 129-146.
Bilby, R. E. and J. W. Ward (1989). "Changes in characteristics and function of woody debris with increasing size of streams in Western Washington." Transactions of the American Fisheries Society 118: 368-378.
Bingner, R. L. (1996). "Runoff simulated from Goodwin Creek watershed using SWAT." Transactions of the ASAE 39(1): 85-90.
Binkley, D., and L.H. MacDonald (1994). Forests as nonpoint sources of pollution and effectiveness of best management practices. NCASI Technical Bulletin No. 672. NY.: 57p.
Birkeland, P. W. (1990). "Soil-geomorphic research  - a selective overview." Geomorphology 3: 207-224.
Birkeland, P. W., et al. (1991). Soils as a tool for applied quaternary geology, Utah Department of Natural Resources: 63.
Bisaillon, J.-F., et al. (2007). Effect of winter harshness on Atlantic salmon (Salmo salar L.) egg to fry (0+) and fry to parr (1+) over-winter mortality. 14th Workshop on the Hydraulics of Ice Covered Rivers. Quebec City, CGU HS Committee on River Ice Processes and the Environment: 7.
Although winter mortality has long been regarded as a major factor affecting salmonid
production in cold-region rivers, only few studies have explored the impact of winter
harshness on salmonid mortality. In this study, we conducted an analysis of juvenile
Atlantic salmon egg to fry (0+) and fry to parr (1+) over-winter mortality in relation to
various hydro-climatic variables. We analysed the egg and fish abundances data
collected between 1983 and 1992 on the Trinité River (Québec, Canada) by the Société
de la Faune et des Parcs du Québec (FAPAQ, Québec). The results indicate that egg to
fry over-winter mortality was high (min 83%, max 93%) but inversely related to winter
coldness as defined by the cumulated freezing degree-days between October 1st and
January 31st (r2 = 0,81, p < 0,001) and positively related to winter low discharge as
defined by the ratio of November to February mean discharges (r2 = 0,80, p < 0,001).
Fry to parr overwinter mortality ranged from 42% to 73% between years and was
inversely related to winter coldness (cumulated freezing-degree days in January and
February) (r2 = 0,91, p < 0,0001) and positively related to conditions conducive to frazil
and anchor ice formation in early winter (number of days in November with mean air
temperature below –10oC) (r2 = 0,77, p < 0,01). These results are discussed in relation
with other similar studies conducted on the Indian River, New Foundland (Chadwick,
1982) and Catamaran Brook, New Brunswick (Cunjak and Therrien, 1998).
 
Bischoff, J. L., et al. (1997). "Climatic Oscillations 10,000-155,000 yr B.P. at Owen's Lake, California Reflected in Glacial Rock Flour Abundance and Lake Salinity in Core OL-92." Quaternary Research 48: 313-325.
Bishop, A. W. (1955). "The use of the slip circle in the stability analysis of slopes." Geotechnique 5: 7-17.
Bishop, K. M. (1999). "Determination of translational landslide slip surface depth using balanced cross sections." Environmental & Engineering Geoscience 5(2): 147-156.
Bisson, P. A., J.M. Buffington, and D.R. Montgomery (2006). Valley segments, stream reaches, and channel units. Methods in Stream Ecology, Elsevier.
Bisson, P. A., et al. (1987). Large Woody Debris in Forested Streams in the Pacific Northwest: Past, Present, and Future. Streamside Management: Forestry and Fishery Interactions. E. O. Salo and T. W. Cundy. Seattle, Washington: 143-190.
Bisson, P. A., et al. (1987). Large woody debris in forested streams in the Pacific Northwest: past, present, future. Streamside Management: Forestry and Fisheries Interactions. University of Washington, College of Forest Resources, Contribution 57. E. O. Salo and T. W. Cundy. Seattle, University of Washington: 143-190.
Bisson, P. A., et al. (2009). "Freshwater ecosystems and resilience of Pacific Salmon: habitat management based on natural variability." Ecology and Society 14(1): 45.
In spite of numerous habitat restoration programs in fresh waters with an aggregate annual
funding of millions of dollars, many populations of Pacific salmon remain significantly imperiled. Habitat
restoration strategies that address limited environmental attributes and partial salmon life-history
requirements or approaches that attempt to force aquatic habitat to conform to idealized but ecologically
unsustainable conditions may partly explain this lack of response. Natural watershed processes generate
highly variable environmental conditions and population responses, i.e., multiple life histories, that are
often not considered in restoration. Examples from several locations underscore the importance of natural
variability to the resilience of Pacific salmon. The implication is that habitat restoration efforts will be more
likely to foster salmon resilience if they consider processes that generate and maintain natural variability
in fresh water. We identify three specific criteria for management based on natural variability: the capacity
of aquatic habitat to recover from disturbance, a range of habitats distributed across stream networks through
time sufficient to fulfill the requirements of diverse salmon life histories, and ecological connectivity. In
light of these considerations, we discuss current threats to habitat resilience and describe how regulatory
and restoration approaches can be modified to better incorporate natural variability.
 
Bisson, P. A. and D. R. Montgomery, Eds. (1996). Valley segments, stream reaches, and channel units. Methods in Stream Ecology. San Diego, CA, Academic Press.
Bisson, P. A. and D. R. Montgomery (1996). Valley segments, stream reaches, and channel units. Methods in Stream Ecology. F. R. Hauer and G. A. Lamberti, Academic Press, N.Y.: 23-52.
Bisson, P. A., et al. (1982). A system of naming habitat types in small streams, with examples of habitat utilization by salmonids during low streamflow. Aquisition and utilization of aquatic habitat inventory information. N. B. Armantrout. Portland, OR, Western Division, American Fisheries Socity: 62-73.
Bisson, P. A., et al. (1982). A system of naming habitat types in small streams, with examples of habitat utilization by salmonids during low streamflow. Aquisition and Utilization of Aquatic Habitat Inventory Information. N. B. Armantrout. Portland, OR, Western Division, American Fisheries Socity: 62-73.
Bisson, P. A., et al. (1988). "Summer Production of Coho Salmon Stocked in Mount St. Helens Streams 3-6 Years after the 1980 Eruption " Transactions of the American Fisheries Society 117: 322-335.
Bisson, P. A., et al. (2003). "Fire and aquatic ecosystems of the western USA: current knowledge and key questions." Forest Ecology and Management 178: 213-229.
Understanding of the effects of wildland fire and fire management on aquatic and riparian ecosystems is an evolving field, with
many questions still to be resolved. Limitations of current knowledge, and the certainty that fire management will continue,
underscore the need to summarize available information. Integrating fire and fuels management with aquatic ecosystem
conservation begins with recognizing that terrestrial and aquatic ecosystems are linked and dynamic, and that fire can play a
critical role in maintaining aquatic ecological diversity. To protect aquatic ecosystems we argue that it will be important to: (1)
accommodate fire-related and other ecological processes that maintain aquatic habitats and biodiversity, and not simply control
fires or fuels; (2) prioritize projects according to risks and opportunities for fire control and the protection of aquatic ecosystems;
and (3) develop new consistency in the management and regulatory process. Ultimately, all natural resource management is
uncertain; the role of science is to apply experimental design and hypothesis testing to management applications that affect fire
and aquatic ecosystems. Policy-makers and the public will benefit from an expanded appreciation of fire ecology that enables
them to implement watershed management projects as experiments with hypothesized outcomes, adequate controls, and
replication.
 
 
Bisson, P. A. a. c.-a. (1999). Roads analysis:informing decisions about managing the National Forest Transportation System, United States Department of Agriculture
Forest Service: 1-222.
Bisson Peter, A. a. s. (1999). Roads Analysis: Informing Decisions About Managing the National Forest Transportaition System, US Forest Service: 1-222.
Bjonness, I.-M. (1983). "External Economic Dependency and Changing Human Adjustment to Marginal Environment in the High Himalaya, Nepal." Mountain Research and Development 3(3): 263-272.
Bjornn, T. C., et al. (1991). "Relation of Cover Alterations to the Summer Standing Crop of Young Salmonids in Small Southeast Alaska Streams." Transactions of the American Fisheries Society 120: 562-570.
Bjornn, T. C. and D. W. Reiser (1991). Habitat requirements of salmonids in streams. Influences of Forest and Rangeland Management on Salmonid Fishes and Their Habitats. W. R. Meehan. Bethesda, Maryland, American Fisheries Society.
Blackwelder, E. (1928). "Mudflow as a geologic agent in semiarid mountains." Bulletin of the Geological Society of America 39: 465-484.
Blair, J. B., et al. (1999). "The Laser Vegetation Imaging Sensor: a medium-altitude, digitisation-only, airborne laser altimeter for mapping vegetation and topography." EPRS Journal of Photogrammetry and Remote Sensing 54: 115-122.
Blanchet, D. (1983). Evaluation of Recent Channel Changes on the Scott River. Anchorage, Alaska, US Forest Service: 18.
Blasing, T. J. and H. C. Fritts (1976). "Reconstructing Past Climatic Anomalies in the North Pacific and Western North America From Tree-ring Data." Quaternary Research 6: 563-579.
Bledsoe, B. P., M.C. Brown, and D. A. Raff (2007). "Geotools: a toolkit for fluvial systems analysis." Journal of the American Water Resources Association 43(3): 757-772.
Bledsoe, B. P. and C. C. Watson (2001). "Effects of urbanization on channel instability." Journal of the American Water Resources Association 37(2): 255-270.
Channel instability and aquatic ecosystem degradation have been linked to watershed imperviousness in humid regions of the U.S. In an effort to provide a more process-based linkage between observed thresholds of aquatic ecosystem degradation and urbanization, standard single event approaches (U.S. Geological Survey Flood Regression Equations and rational) and continuous hydrologic models (HSPF and CASC2D) were used to examine potential changes in flow regime associated with varying levels of watershed imperviousness. The predicted changes in flow parameters were then interpreted in concert with risk-based models of channel form and instability. Although low levels of imperviousness (10 to 20 percent) clearly have the potential to destabilize streams, changes in discharge, and thus stream power, associated with increased impervious area are highly variable and dependent upon watershed-specific conditions. In addition to the storage characteristics of the pre-development watershed, the magnitude of change is sensitive to the connectivity and conveyance of impervious areas as well as the specific characteristics of the receiving channels. Different stream types are likely to exhibit varying degrees and types of instability, depending on entrenchment, relative erodibility of bed and banks, riparian condition, mode of sediment transport (bedload versus suspended load), and proximity to geomorphic thresholds. Nonetheless, simple risk-based analyses of the potential impacts of land use change on aquatic ecosystems have the potential to redirect and improve the effectiveness of watershed management strategies by facilitating the identification of channels that may be most sensitive to changes in stream power.
 
Bledsoe, B. P. and C. C. Watson (2001). "Logistic analysis of channel pattern thresholds: meandering, braiding, and incising." Geomorphology 38(3-4): 281-300.
A large and geographically diverse data set consisting of meandering, braiding, incising, and post-incision equilibrium streams was used in conjunction with logistic regression analysis to develop a probabilistic approach to predicting thresholds of channel pattern and instability. An energy-based index was developed for estimating the risk of channel instability associated with specific stream power relative to sedimentary characteristics. The strong significance of the 74 statistical models examined suggests that logistic regression analysis is an appropriate and effective technique for associating basic hydraulic data with Various channel forms. The probabilistic diagrams resulting from these analyses depict a more realistic assessment of the uncertainty associated with previously identified thresholds of channel form and instability and provide a means of gauging channel sensitivity to changes in controlling variables. (C) 2001 Elsevier Science B.V. All rights reserved.
 
Bledsoe, B. P., et al. (2002). "Quantification of incised channel evolution and equilibrium." Journal of the American Water Resources Association 38(3): 861-870.
Incised channels are caused by an imbalance between sediment transport capacity and sediment supply that alters channel morphology through bed and bank erosion, Consistent sequential changes in incised channel morphology may be quantified and used to develop relationships describing quasi-equilibrium conditions in these channels. We analyzed the hydraulic characteristics of streams in the Yazoo River Basin, Mississippi in various stages of incised channel evolution. The hydraulic characteristics of incising channels were observed to follow the sequence predicted by previous conceptual models of incised channel response. Multiple regression models of stable slopes in quasi-equilibrium channels that have completed a full evolutionary sequence were developed, These models compare favorably with analytical solutions based on the extremal hypothesis of minimum stream power and empirical relationships from other regions. Appropriate application of these empirical relationships may be useful in preliminary design of stream rehabilitation strategies.
 
Bleeck, J. A., et al. (1993). "The 'Geology' of Papermaking." Washington Geology 21(3): 3-27.
Blöschl, G. and M. Sivapalan (1997). "Process controls on regional flood frequency: Coefficient of variation and basin scale." Water Resources Research 33(12): 2967-2980.
Board, W. F. P. (1997). Board Manual: Standard Methodology for Conducting Watershed Analysis. Olympia, Washington Department of Natural Resources.
Bodhaine, G. L. and D. M. Thomas (1964). Magnitude and frequency of floods in the U.S. Part 12. Pacific slope basins in Washington and Upper Columbia River basin, US Geological Survey: 1687.
Boer, D. H. (1992). "Hierarchies and spatial scale in process geomorphology: a review." Geomorphology 4: 303-318.
Boer, D. H. d. (4). "Hierarchies and spatial scale in process geomorphology: a review." Geomorphology 4: 303-318.
Boer, D. H. d. (1992). "Hierarchies and spatial scale in process geomorphology: a review." Geomorphology 4: 303-318.
Bogardi, J. J., et al. (1988). "Practical Generation of Synthetic Rainfall Event Time Series in a Semi-Arid Climatic Zone." Journal of Hydrology 103: 357-373.
Bogena, H. R., and B. Diekkruger (2002). "Modelling solute and sediment transport at different temporal and spatial scales." Earth Surface Processes and Landforms 27: 1475-1489.
A concept for continuous modeling of solute and sediment transport at the catchment scale is presented. the simulation is based on the application of the OPUS system model system, which is designed for long-term simulations of agricultural processes. As OPUS is a hillslope model, the catchment is discretized into numerous slopes. Modifications and enhancements of the OPUS model were made in order to simulate interflow induced by macropores. The simulation results are compared with measurements at different spatial and temporal scales of a meso-scale catchment used for drinking water supply. Five small subcatchments of 22 to 29 ha are used to validate the model at the local scale. The modified and validated model is then applied to the Wahnbach catchment with an area of 54km2. Simulations at the local and catchment scales are presented and compared with measurements. To model the long term behaviour of solute and sediment transport with a deterministic model, a complete set of climate and land use data is necessary. In this case study a method was developed to overcome the lack of data by using a weather generator. the long term simulations are validated by determining the mass of sediments trapped in the drinking water reservoir. The results confirm that the model concpet is applicable to a larger range of scales from single events to decades and from single slopes to meso-scale catchments. By using these methods it is conceivable to forecast simulations of the future solute and sediment tranport discharge using land use scenarios.
 
Bolstad, P. V., and W.T.Swank (1997). "Cumulative impacts of landuse on water quality in a southern Appalachian watershed." Journal of the American Water Resources Association 33(3): 519-533.
Bond, N. (2004). "Spatial variation in fine seediment transport in small upland streams: the effect of flow regulation and catchment geology." River Research and Applications 20: 705-717.
Bonney, C. T. G. (1902). "Moraines and Mud-streams in the Alps." Geological Magazine 9(1): 8-17.
Bonniwell, E. C., et al. (1999). "Determining the times and distances of particle transit in a mountain streams using fallout radionuclides." Geomorphology 27: 75-92.
Booth, D. B., et al. (1991). Sediment transport along the South Fork and Mainstem of the Snoqualmie River. Seattle, Washington, King County Surface Water Management Division: 56.
Booth, D. B. and B. Hallet (1993). "Channel networks carved by subglacial water: observations and reconstruction in the eastern Puget Lowland of Washington." Geological Society of America Bulletin 105: 671-683.
Booth, D. B., et al. (2003). The geology of Puget Lowland Rivers. Restoration of Puget Sound Rivers. D. R. Montgomery, S. Bolton, D. B. Booth and L. Wall. Seattle, WA, University of Washington Press: 14-45.
Booth, D. E. (1991). "Estimating Prelogging Old-Growth in the Pacific Northwest." Journal of Forestry 89(10): 25-29.
Borah, D. K. (2002). Watershed-scale nonpoint source pollution models: mathematical bases. ASAE Annual Meeting / CIGR XVth World Congress Paper Number 022091, Chicago, Illinois, ASAE.
Borcard, D. "Partial r2, contribution and fraction[a]." UNPUBLISHED WEB NOTES.
Borden, R. K. and K. G. Troost (2001). Late Pleistocene Stratigraphy in the South-Central Puget Lowland, Pierce County, Washington, Washington State Department of Natural Resources: 1-33.
Borga, M., et al. (2002). "Assessment of shallow landsliding by using a physically based model of hillslope stability." Hydrological Processes 16(14): 2833-2851.
A model for the simulation of shallow landsliding triggered by heavy rainstorms is analysed and discussed. The model is applied in two mountainous catchments in the Dolomites (Eastern Italian Alps): the Cordon catchment (5 km2) and the Vauz catchment (1·9 km2), where field surveys provided a description of hydraulic and geotechnical properties of soils and an inventory of landslide scars is available. The stability mapping procedure, which is similar to that proposed by Montgomery and Dietrich (1994 Water Resources Research 30: 1153), combines steady-state hydrologic concepts with the infinite slope stability model. The model provides an estimate of the spatial distribution of the critical rainfall, which is the minimum steady-state rainfall predicted to cause instability. The comparison of the landslides observed in the study basins with model predictions shows that the distribution of critical rainfall obtained from the model provides a surrogate for failure initiation probability as a function of topographic location
 
Borga, M., et al. (2002). "Analysis of topographic and climatic control on rainfall-triggered shallow landsliding using a quasi-dynamic wetness index." Journal of Hydrology 268: 56-71.
Bormann, B. T., et al. (1995). "Rapid soil development after windthrow disturbance in pristine forests." Journal of Ecology 83: 747-757.
Bormann, F. H. and G. E. Likens (1979). "Catastrophic Disturbance and the Steady State in Northern Hardwood Forests." American Scientist 67: 660-669.
Bormann, F. H. and G. E. Likens (1979). Pattern and Processes in a Forested Ecosystem. New York, Springer-Verlag.
Bosch, D. D., R. G. Williams, S.P. Inamdar, J. M. Sheridan, R. R. Lowrance, and M.T. Teeng. (1998). Erosion and sediment transport through riparian forest buffers.
Bosch, J. M., and J.D. Hewlett (1982). "A review of catchment experiments to determine the effect of vegetation changes on water yield and evapotranspiration." Journal of Hydrology 55: 3-23.
Bosscher, P. J., et al. (1988). Evaluation of Risks of Slope Instability Along a Coastal Reach. 5th International Symposium on Landslides, Lausanne, Switzerland.
Botkin, D., et al. (1994). Status and Future of Salmon of Western Oregon and Northern California: Findings and Options (DRAFT). Santa Barbara, California, The Center for the Study of the Environment: 74-113.
Botkin, D. B. (1990). Discordant Harmonies: A New Ecology for the Twenty-First Century. New York, Oxford University Press.
Boughton, C. J., et al. (?). Stream and Ground-Water Monitoring Program, Lake Tahoe Basin, Nevada and California US Geological Survey: 1-6.
Bovis, Michael J. and M. Jakob (1999). "The role of debris supply conditions in predicting debris flow activity." Earth Surface Processes and Landforms 24: 1039-1054.
Bowen, Z. H. and R. G. Waltermire (2002). "Evaluation of Light Detection and Ranging (Lidar) for Measuring River Corridor Topography." Journal of American Water Resources Association 38(1): 33-41.
Bowles, J. E. (1986). Engineering properties of soils and their measurement. New York, McGraw-Hill.
Boychuk, D., et al. (1997). "Modelling the effect of spatial scale and correlated fire disturbances on forest age distribution." Ecological Modelling 95: 145-164.
Bozek, M. A. and M. K. Young (1994). "Fish mortality resulting from delayed effects of fire in the greater Yellowstone ecosystem." Great Basin Naturalist 54(1): 91-95.
Bracken, L., and J. Croke (2007). "The concept of hydrological connectivity and its contribution to understanding runoff-dominated geomorphic systems." Hydrologic Processes 21: 1749-1763.
Bracken, L. J., J. Croke, and M. Kirkby (2004). Connectivity in geomorphology: definition and evaluation. Geophysical Research Abstracts, European Geosciences Union, Volume 6, 04039.
Bradford, M. J. and J. R. Irvine (2000). "Land use, fishing, climate change, and the decline of Thompson River, British Columbia, coho salmon." Canadian Journal of Fisheries and Aquatic Science 57: 13-16.
We investigated a recent, major decline in the abundance of a large aggregate of coho salmon
(Oncorhynchus kisutch) spawning in the Thompson River, British Columbia, watershed. We found that the decline
could be attributed to a declining trend in productivity likely related to changing ocean conditions, overfishing, and
freshwater habitat alteration. Among individual watersheds, rates of decline in adult coho salmon abundance were correlated
with agricultural land use, road density, and a qualitative measure of stream habitat status but not with the proportion
of land recently logged. The recovery of these populations will require the prudent regulation of fishing, the
restoration of salmon producing watersheds, and an improvement in ocean conditions.
 
Bradley, J. B. and D. W. Reiser (?). Sedimentation and Fisheries Evaluation, Stikine area, Tongass National Forest. Carlsbad, CA, WEST Consultants, Inc.: 64.
Bradley, J. B. and P. J. Whiting (?). Draft-stream characterization and a stream classification for small streams. Seattle, Washington, submitted by WEST Consultats, submitted to the Washington Department of Natural Resources: 1-38.
Brady, L. M., F. Gray, C.A. Wissler, and D.P. Guertin (2001). Spatial Variability of sediment erosion processes using GIS analysis within watersheds in a historically mined region, Patagonia Mountains, Arizona. Open-File Report 01-267. Tucson, USGS, University of Arizona: 30.
A hillslope-scale erosion prediction model (USLE) and a spatially derived sediment delivery model (SEDMOD) are applied within a raster GIS to estimate erosion, sediment yield, and sediment deposition for five adjacent sub-basins impacted by historical mining in the Patagonia Mountains of southern Arizona. Geospatial landscape data (elevation, soil type, vegetation, mine locations, and stream networks) were divided into 30m2 cell grids, allowing for consistent high-resolution analysis within each watershed. The automation of paper soils maps is described. The model results identify non-point sources and sinks of trace-metal bearing sediment.
 
Bragg, D. C. (2000). "Simulating catastrophic and individualistic large woody debris recruitment for a small riparian system." Ecology 81(5): 1383-1394.
Surprisingly little research has been done to partition the contribution of
catastrophic disturbance from that of small-scale individualistic mortality events on riparian
large woody debris (LWD) recruitment. This study compared the impact of both processes
on recruitment through simulation of several catastrophic disturbances (a spruce beetle
outbreak, a moderately intense fire, and a clearcut) and undisturbed (individualistic mortality
only) old growth for a small headwater stream in the Intermountain West of the United
States. All scenarios progressed through a two-stage process, with the Forest Vegetation
Simulator growth and yield model controlling forest dynamics and a postprocessor (CWD,
version 1.2) predicting riparian LWD recruitment. Projections indicate that individualisticonly
conditions delivered 2.5 m3 LWD·100 m reach21·10-yr cycle21; while the spruce beetle-,
fire-, and clearcut-affected stands averaged 2.9, 3.2, and 1.5 m3 LWD·100 m reach21·cycle21,
respectively. Stands impacted by natural catastrophic disturbance significantly (P , 0.05)
increased cumulative (300 yr) LWD recruitment over the individualistic-only scenario,
whereas clear-cutting significantly decreased total delivery. In-stream LWD loads, relatively
stable in undisturbed riparian zones, fluctuated sharply under catastrophic disturbance. Peak
channel loads associated with natural perturbation occurred ;30 yr after the event while
debris volumes under clear-cutting immediately declined. The postevent recruitment and
in-stream LWD stocks of all disturbance scenarios eventually fell below undisturbed conditions,
requiring decades to recover historical volumes. Catastrophic disturbances induced
such steep oscillations in riparian LWD load that the systems experiencing frequent largescale
perturbations never achieved a long-term steady state, as some have postulated. Because
of the inflation in cumulative LWD delivery, it may prove advantageous to encourage
(or imitate) some catastrophic disturbance in forests along streams noticeably depauperate
of LWD.
 
Bragg, D. C., et al. (2000). Modeling Large Woody Debris Recruitment for Small Streams of the Central Rocky Mountains. Fort Collins, CO, U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station: 36.
Bragg, D. C., et al. (2000). Modeling large woody debris recruitment for small streams of the central Rocky Mountains. Fort Collins, CO, U. S. Department of Agriculture, Rocky Mountain Research Station.
Brake, D. J., et al. (1999). Sediment delivery below roads in the Oregon Coast Range. Boise ID, Rocky Mountain Research Station: 18.
Bramblett, R. G., et al. (2002). "Seasonal use of small tributary and main-stem habitats by juvenile steelhead, coho salmon, and dolly varden in a southeastern Alaska drainage basin." Transactions of the American Fisheries Society 131: 498-506.
The movement of juvenile salmonids between small tributaries and main-stem habitats
in southeast Alaska watersheds is poorly understood. We observed movements of steelhead Oncorhynchus
mykiss, coho salmon O. kisutch, and Dolly Varden Salvelinus malma between mainstem
and tributary habitats at weirs located on tributaries in the Staney Creek watershed in southeast
Alaska. We used seasonal relative abundance (catch per unit effort) in eight main-stem reaches
and eight tributaries to corroborate observed movement in the two streams with weirs.We observed
juvenile steelhead and coho salmon moving through the weirs into tributaries during the fall as
flows increased and temperatures decreased. The relative abundance of steelhead was greater in
main-stem sites than in tributaries during the summer, whereas during spring and fall relative
abundance in the tributaries was similar to that in the main stem. Juvenile coho salmon were
abundant in tributaries during all seasons. The relative abundance of Dolly Varden was greater in
the tributaries than in the main-stem during all seasons. These results underscore the significance
of links between main-stem habitats and small tributaries for production of juvenile salmonids.
 
Brandinoni, F. and M. A. Hassan (2006). "Glacial erosion, evolution of river long profiles, and the organization of process domains in mountain drainage basins in coastal British Columbia." Journal of Geophysical Research 3.
Brardinoni, F. (2001). Identification of Natural and Logging-Related Landslides in the Capilano River Basin (Coastal British Columbia): A Comparison Between Remotely Sensed Survey and Field Survey. Department of Geography. Vancouver, B.C., The University of British Columbia: 127.
Brardinoni, F. and M. Church (2004). "Representing the landslide magnitude-frequency relation: Capilano River Basin, British Columbia." Earth Surface Processes and Landforms 29: 115-124.
Brardinoni, F. and M. A. Hassan (2006). "Glacial erosion, evolution of river long profiles, and the organization of process domains in mountain drainage basins of coastal British Columbia." Journal of  Geophysical Research 111(F01013): 12.
In glaciated British Columbia, Canada, Quaternary climate changes are responsible
for profound spatial reorganization of Earth surface processes. These changes have left a
landscape characterized by topographic anisotropy associated with a hierarchy of
glacial troughs. The evolution of glaciated landscapes is examined by analyzing the
structure of geomorphic process domains and channel long profiles. To identify process
domains we use channel surveys and GIS analysis to construct slope-area transects of the
channel network. This analysis reveals generalized process-form disequilibrium with a
mismatch between topographic signatures and currently active geomorphic process
domains. At the landscape scale of ‘‘source’’ colluvial channels (contributing area
<1 km2), the glacial/paraglacial signature commonly overrides that produced by
contemporary debris flows. Along the axis of former ice flows, relict glacial cirques
introduce a ‘‘hanging’’ fluvial domain at contributing areas as small as 8   10 2 km2 and
produce complex channel long profiles similar to those observed for rivers responding
to tectonic forcing. Slope-area relations typical of unglaciated equilibrium environments
do not apply here. The concept of process domains appears to hold, however, some major
glacially forced modifications in the alluvial-colluvial transition are observed and the
definition of a depositional colluvial subdomain is proposed. Comparison between
field- and GIS-measured slopes reveals that GIS-associated error is not uniform between
process domains, and that GIS-based plots do not successfully discriminate field-based
process domains. The combination of glacial and post-glacial fingerprints and the
effects of ongoing Earth surface processes generate a complex landscape whose glacial
signatures may persist until the onset of the next ice age.
 
Brardinoni, F., et al. (2002). "Complex mass wasting response of drainage basins to forest management in coastal British Columbia." Geomorphology 49: 109-124.
Brardinoni, F., et al. (2002). "Complex mass wasting response of drainage basins to forest management in coastal British Columbia." Geomorphology 49: 109-124.
The impacts of logging activities on mass wasting were examined in five watersheds in the coastal mountains of British
Columbia. Historical aerial photos were used to document mass wasting events, and their occurrence was related to logging
activities in the study basins. Logged and forested areas were compared in terms of mass wasting magnitude and frequency,
with reference to site characteristics. The recovery time of the landscape after logging was assessed. Bedrock type and basin
physiography had no identifiable effect on mass wasting frequency and magnitude. Mass wasting failure was primarily
controlled by slope gradient. Basin vulnerability increased, following clearcutting relative to forested areas, in that mass wasting
was initiated on gentler slopes. The volume of sediment produced from logged slopes is of the same order as that from forested
areas, which are steeper by as much as 10j. In both logged and forested areas, the size distribution of mass wasting events
follows an exponential distribution. However, the variability in mass wasting size in forested areas is much higher than that
obtained for logged areas. The recovery time after forest harvesting is over 20 years, which confirms published estimates based
on vegetation reestablishment. Continuous disturbance of the basin, however, may extend the recovery time for the whole basin
well beyond 20 years.
 
Brardinoni, F., et al. (2003). "Landslide inventory in a rugged forested watershed: a comparison between air-photo and field survey data." Geomorphology 54: 179-196.
Landslide inventories are routinely compiled by means of aerial photo interpretation (API). when examining photo pairs, the forest canopy (notably in old-growth forest) hides a population of "not visible" landslides. In the present study, we attempt to estimate how important is the contribution of landslides not detectable from aerial photographs to the global mass of sediment production from mass failures on forested terrain of the Capilano basin, coastal British Columbia. API was coupled with intensive fieldwork for identification and measurement of all landslides. A 30-year framework was adopted. We show that "not visible" landslides can represent up to 85% of the total number of failures and account for 30% of the volume of debris mobilised. Such percentages display high sub-basin variability with rates of sediment production varying by one order of magnitude betwen tow sub-basins of the study area. This is explained qualitatively by GIS-based analysis of slope frequency distributions, drainage density, and spatial distribution of surficial materials. Such observations find further support in the definitions of transport-limited and supply-limited basins. As a practical consideration to land managers, we envisage that supplementary fieldwork for landslide identification is mandatory in transport-limited systems only. Fieldwork has deomonstrated that gully-related failures have a greater importance than one could expect from API.
 
Brardinoni, F., et al. (2003). "Landslide inventory in a rugged forested watershed: a comparison between remotely sensed and field survey data." Greomorphology 54: 179-196.
Braun, P., T. Molnar, and H-B Kleeberg (1997). "The problem of scaling in grid-related hydroogical process modelling." Hydrologic Processes 11: 1219-1230.
This artical shows some important scaling aspects that exist in hydrological process grid modelling with GIS, with respect to selecting the optimal grid width. Hydrologically relevant space parameter fields should often be regarded as self-similar fractals for which certain invariance characteristics are applied. These characteristics are especially important if there is an extensive change in the model grid width. The self-similarity generates simple scaling laws, which can be used for rescaling procedures.
 
Bravard, J.-P., et al. (1986). "Impact of civil engineering works on the successions of communities in a fluvial system." Oikos 47: 92-111.
Bravo-Espinosa, M., W.R. Osterkamp, and V.L.Lopes (2003). "bedload transport in alluvial channels." Journal of Hydraulic Engineering 129(10): 783-795.
Brayshaw, D. and M. A. Hassan (2009). "Debris flow initiation and sediment recharge in gullies." Geomorphology.
Landslides that enter gullied low-order drainages can either initiate debris flow or stop,
depositing sediment in the channel. This process is one of the most common ways that
debris flows initiate, but little attention to date has been paid to evaluating the factors that
affect whether or not the initial landslide will become a debris flow or deposit sediment
in the channel. Statistically significant parameters that determine whether slope failures
become debris flows or act to recharge in-channel sediment are channel gradient, angle of
entry of failure into the channel, initial failure volume, and the amount of in-channel
stored sediment. Steeper channels, low angles of entry, lower volumes of in-channel
sediment, and larger initial failures were more likely to result in debris flows. This study
found that as the volume of in-channel stored sediment increased, the volume of initial
failure required to initiate a debris flow also increased. This result calls into question the
simple supply-limited model of cyclical debris recharge and debris flow in low-order
gullied drainages and suggests a negative feedback mechanism between debris
accumulation and debris flow susceptibility
 
Breiman, L., J.H. Friedman, R.A. Ohlshen, and C.J. Stone (1998). Classification and Regression Trees., Chapman and Hall, Washington, D.C., 358 pp.
Breiman, L., et al. (1984). Classification and regression trees. Monterey, CA, Wadsworth & Brooks / Cole Advanced Books & Software.
Bren, L. J. (1997). "Effects of slope vegetation removal on the diurnal variations of a small mountain stream." Water Resources Research 33(2): 321-331.
The effect on removal of lower, mid, and upper slope vegetation on the diurnal
variation in streamflow from a 46-ha catchment was observed. The diurnal variation in
streamflow of the small stream was measurable during the late-spring-to-late-autumn
period. The amplitude in streamflow variation reached a maximum in early summer and
declined during autumn. Observation of diurnal variations during the periods of higher
flow in winter and spring showed that they may occur but were masked by much larger
variations associated with storm runoff. Simulation of the characteristics of the flow
measurement system showed that diurnal variations can only be studied using V-notch
weirs and float recorders during periods of low flow. No effect of the clearing of slope
vegetation on the phase of the outflow could be found. However, there was evidence of a
significanint creasien amplitudep, robablyd uet o increasegdr oundwateoru tflowfr omt he
slopes.I t was concludedt hat the diurnalv ariationi s due to transpirationb y the riparian
and near-riparianv egetationo nly, and that the lower to mid slopev egetationp laysl ittle
role in thisv ariationS. imulationssu ggestetdh at increaseadm plitudeis associatewdi th
increased flow rates, and that the amplitude is not directly affected by water use of
vegetationo n the catchments lopes.I t was concludedt hat the amplitudeo f the variationi s
insensitiveto changesin slopeh ydrologya nd cannotb e usedt o providei nsighti nto deep
slope processe
 
Brenden, T. O., et al. (2007). "Comparison between model-predicted and field-measured stream habitat features for evaluating fish assemblage-habatit relationships." Transactions of the American Fisheries Society 136: 580-592.
The use of model-predicted, local-scale habitat
data as inputs in analyses intended to evaluate multiscale fish
assemblage–habitat relationships in streams has become
increasingly common as the scale at which such studies are
conducted has increased. We used fish assemblage and habitat
data from 208 wadeable streams in Wisconsin and Michigan
to determine whether model-predicted habitat data would
yield results similar to those of field-measured data in
multiscale analyses of fish assemblage–habitat relationships.
Predictions of local habitat features from landscape variables
were generated by means of generalized additive modeling
with likelihood-based boosting. Relationships between fish
assemblage measures and landscape and local habitat features
were studied via partial constrained multivariate ordination
analyses. The total variation explained in the fish assemblage
data sets was similar for model-predicted local habitat features
and field-measured data, as was the proportion of variation
explained that was due independently to local and regional
(i.e., landscape) habitat features. We observed dissimilar
results in the magnitude of ordination scores for local habitat
features and the directional relationships between local habitat
ordination scores and individual species and assemblage
metric scores. Our findings indicate that model-predicted,
local-scale habitat data can be useful for evaluating the relative
strengths of local and regional habitat features in structuring
fish assemblages, but caution may be necessary when
evaluating species–habitat or assemblage metric–habitat
relationships.
 
Brenden, T. O., et al. (2007). "A spatially constrained clustering program for river valley segment delineation from GIS digital river networks." Envrionmental Modelling and Software 23(5): 638-649.
River valley segments are adjacent sections of streams and rivers that are relatively homogeneous in hydrology, limnology, channel morphology, riparian dynamics, and biological communities. River valley segments have been advocated as appropriate spatial units for assessing, monitoring, and managing rivers and streams for several reasons; however, methods for delineating these spatial units have been tedious to implement or have lacked objectivity, which arguably has limited their use as river and stream management units by natural resource agencies. We describe a spatially constrained clustering program that we developed for delineating river valley segments from geographic information system digital river network databases that is flexible, easy-to-use, and improves objectivity in the river valley segment delineation process. This program, which we refer to as the valley segment affinity search technique (VAST), includes a variety of options for determining spatial adjacency in stream reaches, as well as several data transformation methods, types of resemblance coefficients, and cluster linkage methods. The usefulness of VAST is demonstrated by using it to delineate river valley segments for river network databases for Michigan and Wisconsin, USA, and by comparing river valley segments delineated by VAST to an expert-opinion delineation previously completed for a Michigan river network database.
 
Brenning, A. (2005). "Spatial prediction models for landslide hazards: review, comparison and evaluation." Natural Hazards and Earth System Sciences 5: 853-862.
The predictive power of logistic regression, support
vector machines and bootstrap-aggregated classification
trees (bagging, double-bagging) is compared using misclassification
error rates on independent test data sets. Based on
a resampling approach that takes into account spatial autocorrelation,
error rates for predicting “present” and “future”
landslides are estimated within and outside the training area.
In a case study from the Ecuadorian Andes, logistic regression
with stepwise backward variable selection yields lowest
error rates and demonstrates the best generalization capabilities.
The evaluation outside the training area reveals that
tree-based methods tend to overfit the data.
 
Breshears, D. D., J.J. Whicker, M.P. Johansen, and J.E. Pinder lll (2003). "Wind and water erosion and transport in semi-arid shrubland, grassland and forest ecosystems: quantifying dominance of horizontal wind-driven transport." Earth Surface Processes and Landforms 28: 1189-1209.
Breuer, L., J.A. Huisman, H.G. Frede. (2006). "Monte Carlo estimation of uncertainty in the simulated hydrological response to land use change." Environmental Model Assessment 11: 2009-2218.
Bridge, J. S. and S. J. Bennett (1992). "A Model for the Entrainment and Transport of Sediment Grains of Mixed Sizes, Shapes, and Densities." Water  Resources Research 28(2): 337-363.
Brierley, G. J. and E. J. Hickin (1985). "The Downstream Gradation of Particle Sizes in the Squamish River, British Columbia." Earth Surface Processes and Landforms 10: 597-606.
Brinkmann, R., et al. (1993). Soil loss and the archeological record in the Plumas National Forest, California. Abstracts, Annual Meeting - Association of American Geographers, vol.89: 24-25.
Bristow, C. C., et al. (1993). "Morphology and facies models of channel confluences." Spec. Publs Int. Ass. Sedimentology 17: 91-100.
Bristow, C. S., et al. (1993). Morphology and facies models of channel confluences. Alluvial Sedimentation. M. Marzo and C. Puigdefabregas, International Association of Sedimentologists. Special Publication 17: 91-100.
British Columbia Ministry of Forests (1998). Statistical Methods for Adaptive Management Studies, B.C. Ministry of Forests.
Brocard, G. Y. and P. A. van der Beek (2006). Influence of incision rate, rock strength, and bedload supply on becrock river gradients and valley-flat widths: field-based evidence and calibrations from western alpine rivers (southeast France). Tectonics, Climate, and Landscape Evolution. S. D. Willett, N. Hovius, M. T. Brandon and D. Fisher, Geological Society of america: 101-126.
Several process-based models of river incision have been proposed in recent years
that attempt to describe fluvial landform development. Although some field tests have
been performed, more data are required to test the ability of these models to predict
the observed evolution of fluvial landforms. We have investigated several tens of rivers
located in the French western Alps that flow across folded sedimentary rocks with
strongly contrasting rock strengths. These rivers record significant variations in some
of the parameters controlling river incision, notably bedrock lithology, stream power,
incision rate, and sediment flux, potentially allowing discrimination between existing
models. Variations in incision rates are driven by variations in the amount of disequilibrium
introduced in the river profiles during the last glaciation. We use diagnostic
indices to investigate transport- and detachment-limited conditions, which include
the channel morphology, the occurrence of lithogenic knickpoints, the continuity of
alluvial and bedrock reaches, and the slope-area scaling of the river long profile. We
observe transitions from detachment-limited to transport-limited conditions with
increasing discharge/drainage area and decreasing incision rate. Bedrock strength
influences the location of the transition predictably. The formation of transport-limited
rivers coincides with the development of a valley flat wider than the active channel,
which accommodates variations in bedrock strength, stream power, and incision rate
along the transport-limited reaches. We propose and calibrate a model for the development
of valley flats along transport-limited rivers and explore some properties of
landscape development in mountain ranges controlled by transport-limited rivers.
 
Brooks, K. N. (1997). "Integrated Watershed Management Research." Journal of Soil and Water Conservation 29(3): 196-219.
Brown, A. E., Lu Zhang, T.A. McMahon, A. W. Western, and R. A. Vertessy (2005). "A review of paired catchment studies for determining changes in water yield resulting from alterations in vegetation." Journal of Hydrology 310: 28-61.
Brown, A. V., Y. Aguila, K. B. Brown and W. P. Fowler (1997). "Responses of benthic macroinvertebrates in small intermittent streams to silvicultural practices." Hydrobiologia 347(1-3): 119-125.
We examined macroinvertebrate communities in small(0.1-1.0 m2) pools of intermittent streams (alwayscontainingsome water but without perennial flow) with small watersheds(2-6 ha) subjected to five types of forest harvest to assesspotential impacts of the different harvest methods. Bufferstrips10 m wide were left on each side of the streams. Each harvesttreatment was coupled with a similar unharvested referencestand.An incomplete block design included three 0.05 m2 vacuumsamples from each treatment paired with three from theadjacentreferences. There was a high degree of similarity amongreferencesfor parameters other than taxonomic composition (e.g.macroinvertebrate density, number of species, Shannondiversity,functional groups, etc.). Statistically significantdifferenceswere found between references and treatments and among harvestmethods but the responses varied among response variables(density,Shannon-Weiner diversity, species composition), differentspeciesassemblages (all invertebrates, chironomids,Ephemeroptera-Plecoptera-Trichoptera [EPT], isopods), andfunctional group categories (shredders, collector-gatherers).Wecollected 56 taxa, 7–16 per site, with low communitysimilarity(mean Jaccardlsquosthinsp=thinsp0.18, mean Bray-Curtis percentdissimilaritythinsp=thinsp81). The most severe harvest treatmentsresultedin the highest diversities of total invertebrates in thesesmallspring pool communities.
 
Brown, E. H., et al. (1986). Geologic map of the northwest Cascades, Washington. Bellingham, Washington.
Brown, E. R. t. e. (1985). Management of wildlife and fish habitat in forests of western Oregon and Washington. Portland, Oregon, USDA Forest Service, Pacific Northwest Region: 2 Volumes.
Brown, G. W. and J. T. Krygier (1971). "Clear-Cut Logging and Sediment Production in the Oregon Coast Range." Water  Resources Research 7(5): 1189-1198.
Brown, L., et al. (1988). "Erosion of the Eastern United States observed with 10Be." Earth Surface Processes and Landforms 13: 441-457.
Brown, L. R., et al. (1994). "Historical Decline and Current Status of Coho Salmon in California " North American Journal of Fisheries Management 14(2): 237-260.
Brown, T. C., D. Binkley (1994). Effect of management on water quality in North American forests. GTR-RM-248. F. C. USDA Forest Service, CO.
Brown, T. C., J.C. Bergstrom, and J.B. Loomis (2007). "Defining, valuing, and providing ecosystem services ." Natural Resources Journal 47(2): 331-376.
Brown, T. C., M.T. Hobbins, and J. A. Ramirez (2008). "Spatial distribution of water supply in the coterminous United States." Journal of the American Water Resources Association 44(5): 1-14.
Brown, T. J., B.L. Hall, and A. L. Westerling (2004). "The impact of twenty-first century climate change on wildland fire danager in the western United States: an applications perspective." Climatic Change 62: 365-388.
Brownlie, W. R. and B. D. Taylor (1981). "Sediment management for southern California mountains, coastal plains and shoreline; Part C, Coastal sediment delivery by major rivers in southern California." EQL Report 17-C: 334.
Bruhn, R. L. (1987). "Continental Tectonics: Selected Topics." Reviews of Geophysics 25(6): 1293-1304.
Brummer, C. J., et al. (2006). "Influence of vertical channel change associeated with wood accumulations on delineating channel migration zones, Washington, USA." Geomorphology.
We combine hydraulic modeling and field investigations of logjams to evaluate linkages between wood-mediated fluctuations
in channel-bed-and water-surface elevations and the potential for lateral channel migration in forest rivers of Washington state. In
the eleven unconfined rivers we investigated, logjams were associated with reduced channel gradient and bank height. Detailed
river gauging and hydraulic modeling document significant increases in the water-surface elevation upstream of channel-spanning
wood accumulations. Logjams initiated lateral channel migration by increasing bed-or water-surface elevations above adjacent
banks. Because the potential for a channel to avulse and migrate across its floodplain increases with the size and volume of
instream wood, the area of the valley bottom potentially occupied by a channel over a specified timeframe —the channel migration
zone (CMZ) —is dependent on the state of riparian forests. The return of riparian forests afforded by current land management
practices will increase the volume and caliber of wood entering Washington rivers to a degree unprecedented since widespread
clearing of wood from forests and rivers nearly 150 years ago. A greater supply of wood from maturing riparian forests will
increase the frequency and spatial extent of channel migration relative to observations from wood-poor channels in the period of
post-European settlement. We propose conceptual guidelines for the delineation of the CMZs that include allowances for vertical
fluctuations in channel elevation caused by accumulations of large woody debris.
 
Brummer, C. J. and D. R. Montgomery (2003). "Downstream coarsening in headwater channels." Water Resources Research 39(10): 1294.
Field data from four mountain drainage basins in western Washington document
systematic downstream coarsening of median bed surface grain size (D50 ) and a subsequent
shift to downstream fining at a drainage area of about 10 km
2
. Analyses of network-wide
patterns of unit stream power derived from both channel surveys and digital elevation
models reveal maximum unit stream power that in all four study areas roughly corresponds
with both the grain size maxima and an inflection in the drainage area-slope relation
thought to represent the transition from debris flow-dominated channels to fluvially
dominated channels. Our results support the hypothesis that basin-wide trends in D50 are
hydraulically controlled by systematic variations in unit stream power in addition to lag
deposits forced by mass-wasting processes. The similar relations found in our four study
areas suggest that the tendency for downstream coarsening may be ubiquitous in headwater
reaches of mountain drainage basins where debris flow processes set the channel
gradient.
 
Brummer, C. J. and D. R. Montgomery (2006). "Influence of coarse lag formation on the mechanics of sediment pulse dispersion in a mountain stream, Squire Creek, North Cascades, Washington, United States." Water Resources Research 42(7).
Mountain channels closely coupled to landslide-prone hillslopes often exhibit bed
surface grain sizes coarser than transportable by annual high flows. Coarse particles within
poorly sorted sediment delivered to channels by mass-wasting processes may not be
readily transported as bed load and can consequently form lag deposits that influence the
morphology, hydraulic roughness, and sediment storage within mountain channel
networks. A tracer study and comparison of supply and bed grain size distributions from a
valley-spanning landslide in the North Cascades of Washington state were used to
derive relations between shear stress and the probability of particle entrainment and
erosion from the sediment pulse. Rapid bed surface armoring formed a relatively immobile
lag deposit within 2 years. Covering of 20% of the bed by lag boulders with <5%
probability of entrainment was sufficient to retard vertical incision and force considerable
channel widening during a flood with an 8- to 152-year recurrence discharge on locally
gauged streams. Our results imply that numerical models of sediment pulse evolution
that do not explicitly incorporate the influence of lag formation may substantially
overestimate long-term dispersion rates. The grain size distribution and lithology of a
sediment input relative to the flow competence of the receiving channel are important
factors influencing the rates and mechanisms of sediment pulse dispersion and the sediment
capacitance provided by coarse-grained sediment pulses in mountain drainage basins.
 
Brummer CJ, a. D. M. (2003). "Downstream coarsening in headwater channels." Water Resources Research 39(10): 1294-1308.
Brunengo, M. J., et al. (1994). Earth, Water, Trees, and Fish: Geomorphology and Land-Use Problems in Forested Mountains of the Pacific Northwest. 1994 Geological Society of America Annual Meeting, Seattle, Washington.
Bruns, D. A., et al. (1984). "Tributaries as modifiers of the river continuum concept: analysis by polar ordination and regression models." Archiv für Hydrobiologie 99: 208-220.
Bruns, D. A., et al. (1984). "Tributaries as modifiers of the river continuum concept: analysis by polar ordination and regression models." Archeological Hydrobiology 99(2): 208-220.
Brunsden, D. (1999). "Some geomorphological considerations for the future development of landslide models." Geomorphology 30: 13-24.
Brunsden, D. and D. K. C. Jones (1972). "The morphology of degraded landslide slopes in South West Dorset." Q. Jl. Eng.Geol. 5: 205-222.
Brunsden, D. and J. B. Thornes (1979). "Landscape sensitivity and change." Transactions, Institute of British Geographers NS 4: 485-515.
Brunton, D. A. and R. B. Bryan (2000). "Rill network development and sediment budgets." Earth Surface Processes and Landforms 25(7): 783-800.
Threshold conditions for rill incision are well known, but few studies have examined the effect of network geometry on water and sediment fluxes within an evolving rill system. This paper reports the first in a series of studies designed to identify the influence of soil properties on rill network and confluence geometry, and on water and sediment fluxes in rill systems. The object of this study was to prepare detailed rill network water and sediment budgets identifying important sources and sinks. Simulated rainfall experiments were carried out on a Canadian silt loam soil in a 7.1 m x 2.4 m flume on a 5 degrees slope. Rill networks of varying complexity developed, which were ultimately constrained by flume boundaries. Sediment and water fluxes and hydraulic conditions were measured within networks and at a terminal weir.Networks evolved by initial knickpoint incision in the lower flume, as flow shear velocities reached critical levels of 4.5-5 cm s(-1) followed by headward migration and tributary development. Microtopography determined tributary location, but the timing and intensity of development were controlled by the incision and migration rate of the main channel, which changed local thalwegs and base levels, raising shear Velocities on side slopes above critical levels. Sediment discharge at the weir broadly reflected rill incision intensity and transport-limited conditions, but as active incision moved headward, the linkage became attenuated and identification of discrete erosional incidents (e.g. local bank collapse) in the signal of the weir record became very difficult.Detailed water and sediment budgets showed much more complex patterns of localized incision and deposition within networks, strongly influenced by local changes in thalweg and in water discharge due to seepage or return flow, and by the effect of confluences on hydraulic conditions. Results indicate the value of detailed sediment budgets in interpreting weir water and sediment flux records, and the necessity of linking such measurements for erosion plots and hillslope segments to rill network characterization. Experimental results are consistent with a simple model of rill system evolution based on channel incision, headward migration, and the critical shear velocities for rill initiation. Copyright (C) 2000 John Wiley & Sons, Ltd.
 
Brutsaert, W. (1994). "The unit response of graoundwater outflow from a hillslope." Water Resources Research 30(10): 2759-2763.
Brutsaert, W. (1994). "The unit response of groundwater outflow from a hillslope." Water Resources Research 30(10): 2759-2763.
Bryan, R. B. (1979). "The Influence of Slope Angle on Soil Entrainment by Sheetwash and Rainsplash." Earth Surface Processes 4: 43-58.
Bryan, R. B. (2000). "Soil erodibility and processes of water erosion on hillslope."
The importance of the inherent resistance of soil to erosional processes, or soil erodibility, is generally recognized in hillslope and fluvial geomorphology, but the full implications of the dynamic soil properties that affect erodibility are seldom considered. In Canada, a wide spectrum of soils and erosional processes has stimulated much research related to soil erodibility. This paper aims to place this work in an international framework of research on water erosion processes, and to identify critical emerging research questions. It focuses particularly on experimental research on rill and interrill erosion using simulated rainfall and recently developed techniques that provide data at appropriate temporal and spatial scales, essential for event-based soil erosion prediction. Results show that many components of erosional response, such as partitioning between rill and interrill or surface and subsurface processes, threshold hydraulic conditions for rill incision, rill network configuration and hillslope sediment delivery, are strongly affected by spatially variable and temporally dynamic soil properties. This agrees with other recent studies, but contrasts markedly with long-held concepts of soil credibility as an essentially constant property for any soil type. Properties that determine erodibility, such as soil aggregation and shear strength, are strongly affected by climatic factors such as rainfall distribution and frost action, and show systematic seasonal variation. They can also change significantly over much shorter time scales with subtle variations in soil water conditions, organic composition, microbiological activity, age-hardening and the structural effect of applied stresses. Property changes between and during rainstorms can dramatically affect the incidence and intensity of rill and interrill erosion and, therefore, both short and long-term hillslope erosional response. Similar property changes, linked to climatic conditions, may also significantly influence the stability and resilience of plant species and vegetation systems. Full understanding of such changes is essential if current event-based soil erosion models such as WEPP and EUROSEM are to attain their full potential predictive precision. The complexity of the interacting processes involved may, however, ultimately make stochastic modelling more effective than physically based modelling in predicting hillslope response to erodibility dynamics.
 
Bryan, R. B. and D. L. Rockwell (1998). "Water table control on rill initiation and implications for erosional response." Geomorphology 23(2-4): 151-169.
On sloping sandy agricultural soil sites near Toronto, Canada, summer storms adequate to generate runoff occur frequently, but rill development occurs mainly in spring when snowmelt or rainfall-induced runoff occurs above frozen subsoil. This suggests that on low and moderate slopes on these soils rill initiation is controlled primarily by hydraulic impedance close to the surface, rather than critical hydraulic conditions in runoff. Laboratory flume experiments were carried out on 10 m slopes at 1.5 degrees, 5 degrees and 9 degrees with loamy sand/clay composite soil sample to test this hypothesis. Runoff with hydraulic conditions adequate for rill initiation occurred rapidly in most tests, but on 1.5 degrees and 5 degrees slopes little knickpoint scour or sediment transport occurred before water table development. This coincided with reduced surface soil strength, knickpoint scour and marked increase in sediment discharge, particularly on 5 degrees slopes where increase was 20- to 30-fold. Further increase in sediment discharge occurred when water tables reached the surface. On 9 degrees slopes runoff occurred more quickly, with higher hydraulic values. Significant rill incision and sediment discharge occurred well before water table development, and ultimately reached much higher values than on lower slopes. Results show that soil erodibility can change dramatically over short time periods during storms due to soil moisture conditions, and that the presence of a hydraulic impedance close to the surface which causes a perched water table to develop can strongly influence rill incision and sediment transport. The influence is unlikely to be marked on soils which are very erodible regardless of moisture conditions, or on extremely resistant soils. It will also be Limited on very gentle or steep sites, but can be a significant factor in rill development on intermediate slopes. (C) 1998 Elsevier Science B.V. All rights reserved.
 
Bryant, M. D., et al. (1990). Evaluation of a Stream Channel-Type System for Southeast Alaska. Portland, OR, U. S. Department of Agriculture, Forest Service: 20.
Bryant, M. D., et al. (2004). "Salmonids on the fringe: abundance, species composition, and habitat use of salmonids in high-gradient headwater streams, southeast Alaska." Transactions of the American Fisheries Society 133: 1529-1538.
We evaluated the species distribution,
abundance, and habitat relationships of salmonids in
small first- to second-order headwater streams in south-east
Alaska. Streams were separated into three zones
based on gradient and sampled during the spring, sum-mer,
and fall. Dolly Varden Salvelinus malma were found
in all streams where fish were present. They were the
dominant species in moderate- (mean gradient 5 5.5%)
and high-gradient (mean gradient 5 12.9%) zones. Coho
salmon Oncorhynchus kisutch fry and parr were the dom-inant
species in the low-gradient zone (mean gradient
5 3.1%) but were present in higher-gradient zones.
Small numbers of steelhead O. mykiss parr were present
in all three zones in the spring and fall. Few were cap-tured
during the summer. Coastal cutthroat trout O.
clarkii were found primarily in one stream and in all
three zones. The density of all species decreased as gra-dient
increased. Anadromous Dolly Varden in spawning
condition were observed in the fall up to the highest
accessible locations in four streams. Salmonids use high-gradient
reaches when pools are present and accessible.
Headwater tributaries comprise a large proportion of
most southeast Alaska watersheds, and the combined
contribution from all of these tributaries to the fish com-munity
may be large. The results from this study un-derscore
the importance of maintaining access for fish
throughout watersheds and into small high-gradient
streams.
 
Buchanan, P. (1988). Debris avalanche and debris torrent initiation, Whatcom County, Washington, U. S. A. Geological Sciences. Vancouver, British Columbia, University of British Columbia: 235.
Buchanan, P. and K. W. Savigny (1990). "Factors controlling debirs avalanche initiation." Can. Geotech. J. 27: 659-675.
Buffington, J. M. (1996). "An alternative method for determining subsurface grain size distributions of gravel-bedded rivers." EOS, Transactions, AGU 77(46): 250.
Buffington, J. M., et al. (2002). "Controls on the size and occurrence of pools in coarse-grained forest rivers." River Research and Applications 18: 507-531.
Controls on pool formation are examined in gravel- and cobble-bed rivers in forest mountain drainage basins of northern
California, southern Oregon, and southeastern Alaska. We demonstrate that the majority of pools at our study sites are
formed by flow obstructions and that pool geometry and frequency largely depend on obstruction characteristics (size,
type, and frequency). However, the effectiveness of obstructions to induce scour also depends on channel characteristics,
such as channel gradient, width:depth ratio, relative submergence (ratio of flow depth to grain size), and the calibre
and rate of bed material supply. Moreover, different reach-scale channel types impose different characteristic physical
processes and boundary conditions that further control the occurrence of pools within a catchment. Our findings indicate
that effective management of pools and associated aquatic habitat requires consideration of a variety of factors, each
of which may be more or less important depending on channel type and location within a catchment. Consequently,
strategies for managing pools that are based solely on single-factor, regional target values (e.g. a certain number of
wood pieces or pools per stream length) are likely to be ineffective because they do not account for the variety of local
and catchment controls on pool scour and, therefore, may be of limited value for proactive management of complex
ecosystems.
 
Buffington, J. M. and D. R. Montgomery (1997). "A systematic analysis of eight decades of incipient motion studies, with special reference to gravel-bedded rivers." Water Resources Research 33(8): 1993-2029.
Buffington, J. M. and D. R. Montgomery (1999). "Effects of hydraulic roughness on surface textures of gravel-bed rivers." Water Resources Research 35(11): 3507-3521.
Buffington, J. M. and D. R. Montgomery (1999). "Effects of sediment supply on surface textures of gravel-bed rivers." Water Resources Research 35(11): 3523-3530.
Buffington, J. M. and D. R. Montgomery (2001). "Reply." Water Resources Research 37(5): 1529-1533.
Buffington, J. M., et al. (2004). "Basin scale availability of salmonid spawning gravel as influenced by channel type and hydraulic roughness in mountain catchments." Can. Geotech. J. 61(11): 2085-2096.
Buffington, J. M., et al. (2004). "Basin-scale availability of salmonid spawning gravel as influenced by channel type and hydraulic roughness in mountain catchments." Canadian Journal of Fisheries and Aquatic Science 61: 2085-2096.
A general framework is presented for examining the effects of channel type and associated hydraulic roughness
on salmonid spawning-gravel availability in mountain catchments. Digital elevation models are coupled with grain-size predictions
to provide basin-scale assessments of the potential extent and spatial pattern of spawning gravels. To demonstrate
both the model and the significance of hydraulic roughness, we present a scenario for optimizing the spatial extent of spawning
gravels as a function of channel type in Pacific Northwest catchments. Predictions indicate that hydraulic roughness
could control more than 65% of the potential available spawning habitat at our study sites. Results further indicate that bar
roughness can be important for maintaining spawning gravels in lower mainstem reaches, while wood roughness may be required
for spawning-gravel maintenance in steeper, upper mainstem channels. Our analysis indicates that wood loss and consequent
textural coarsening could deplete up to one third of the potentially usable spawning area at our study sites.
 
 
Buffington, J. M., et al. (2003). Fluvial processes in Puget Sound Rivers and the Pacific Northwest. Restoration of Puget Sound Rivers. D. R. Montgomery, S. Bolton, D. B. Booth and L. Wall. Seattle, Washington, University of Washington Press: 46-78.
Bull, W. B. (1962). "Relation of textural (CM) patterns to depositional environment of alluvial-fan deposits." Journal of Sedimentary Petrology 32(2): 211-216.
Bull, W. B. (1963). "Alluvial-fan Deposits in Western Fresno County California." Journal of Geology 71(2): 243-251.
Bull, W. B. (1964). Geomorphology of segmented alluvial fans in western Fresno County, California, US Geological Survey.
Bull, W. B. (1977). "The alluvial fan environment." Progress in Physical Geography 1: 222-270.
Bull, W. B. (1977). "The alluvial-fan environment." Progress in Physical Geography 1: 222-270.
Bull, W. B. (1979). "Threshold of critical power in streams." Geological Society of America Bulletin 90: 453-464.
Bull, W. B. (1984). Correlation of flights of global marine terraces. Tectonic Geormorphology. M. Morisawa and J. T. Hack, Allen and Unwin: 129-152.
Bull, W. B. (1999). Quaternary Climatic Changes and the Ice Ages. Surface Processes and Landforms. D. J. Eaterbrook. Upper Saddle River, New Jersey, Prentice Hall: 365-399.
Bunte, K., and L.H. MacDonald (1995). Detecting change in sediment loads: where and how is it possible? Effects of Scale on Interpretation and Management of Sediment and Water Quality (Proceedings of a Boulder Symposium), IAHS Publ. No. 226.
Bunte, K., and L.H. MacDonald (1999). Scale considerations and the detectability of sedimentary cumulative watershed effects. Technical Bulletin No. 776., National Council for Air and Stream Improvement,  Research Triangle Park, NC. 327 pp.: 327.
Bunte, K., and L.H. MacDonald (2002). Predicting the mean annual travel distance of streambed particles: a meta-analysis of published data. EOS Transactions, Abstract #H21G-05, Fall Meeting, San Francisco, CA., American Geophysical Union.
Abstract H21G-05. The cumulative effect of sediment inputs on channel morphology and downstream aquatic resources depends on the rate of downstream particle transport. The use of physically-based models to predict the downstream migration of bedload waves is limited by the need for detailed channel and discharge data. The goal of our study is to determine whether simpler procedures can provide first-order estimates of sediment transport distances and hence the likely magnitude and timing of downstream effects. The specific objectives were to: (1) collect and review existing data on bedload and bedform travel distances; and (2) determine whether mean annual travel distance could be related to easily-measured variables such as stream type, stream gradient, stream width, basin area, size class of the particle(s) being observed, or bed material D50. Our literature review yielded 55 references with 84 observations of travel distance from individual tracer studies or the downstream migration of bedforms and bedload waves. In many cases we had to extrapolate the mean annual travel distance from shorter-term data. Hence the large variability in annual travel distances may partly due to the uncertainty in these extrapolations as well as differences in methodology and study design. Initial results indicate that mean annual travel distance decreases with stream gradient (R2=0.29), increases with basin area (R2=0.29), and decreases with increasing bed material D50 (R2=0.26). In contrast to Beechie (2001), stream width was a relatively poor predictor of mean annual travel distance. Strong correlations between the independent variables meant that multiple regression could explain only 39% of the variability in mean annual transport distances. Efforts to stratify the data showed that mean annual travel distances tended to be larger for streams with finer bed material, braided streams, and migrating bedforms. Coarse-bedded pool-riffle streams, step-pool streams, and ephemeral streams tended to have shorter mean annual travel distances than predicted by the global regressions. These patterns in bedload transport distances suggest that site- or stream-specific correction factors could be developed to improve the predictions derived from the complete data set. Additional analyses are being conducted to better understand the causes of the observed variability and determine the usefulness of this approach for predicting cumulative watershed effects.
 
Bunte, K. and L. MacDonald (1999). Scale considerations and the detectability of sedimentary cumulative watershed effects, National Council for Air and Stream Improvement.
Bunte, K. and L. H. MacDonald (1999). Scale considerations and the detectability of sedimentary cumulative watershed effects, NCASI Technical Bulletin 776. Research Triangle Park, North Carolina: 327.
Burbank, D. W., et al. (1996). "Bedrock incision, rock uplift and threshold hillslopes in the northwestern Himalayas." Nature 379(8): 505-510.
Bureau of Land Management (2000). Science Strategy. Denver, Colorado, Bureau of Land Management: 19.
Bureau of Land Management, U.S. Department of Interior (2008). Final Environmental Impact Statement for the Revision of the Resource Management Plans of the Western Oregon Bureau of Land Mangement, Vol. 3, Appendix J. Oregon and Washington State Office. Portland, OR. Volume III,.
Bureau, U. S. W. (1960). Generalized estimates of probable maximum precipitation for the United States west of the 105th meridian for areas to 400 square miles and duration to 24 hours: 66.
Burgan, R. E. and R. C. Rothermel (1984). BEHAVE: Fire behavior prediction and fuel modeling subsystem - fuel subsystem, USDA Forest Service.
Bürgmann, R., et al. (2000). "Synthetic aperture radar interferometry to measure Earth's surface topography and its deformation." Annual Review Earth and Planetary Sciences 28: 169-209.
Synthetic aperture radar interferometry (InSAR) from Earth-orbiting
spacecraft provides a new tool to map global topography and deformation of the
Earth’s surface. Radar images taken from slightly different viewing directions allow
the construction of digital elevation models of meter-scale accuracy. These data sets
aid in the analysis and interpretation of tectonic and volcanic landscapes. If the Earth’s
surface deformed between two radar image acquisitions, a map of the surface displacement
with tens-of-meters resolution and subcentimeter accuracy can be constructed.
This review gives a basic overview of InSAR for Earth scientists and presents
a selection of geologic applications that demonstrate the unique capabilities of InSAR
for mapping the topography and deformation of the Earth.
 
 
Burke, K. J., et al. (in press). Holocene terraces, sand dunes, and debris fans along the Colorado River in Grand Canyon. Grand Canyon Geology. S. S. Beus and M. Morales. London, Oxford University Press.
Burkhardt, H. J. (1994). Maximizing Forest Productivity.
Burnett, K., et al. (2003). A first step toward broad scale identification of freshwater protected areas for pacific salmon and trout in Oregon, USA. Aquatic Protected Areas: What Works Best and How Do We Know? Proceedings of the World Congress on Aquatic Protected Areas, Cairns, Australia, Austrain Society of Fish Biology, North Beach, WA, Australia.
Burnett, K. M. and D. Miller (2007). "Streamside Policies for Headwater Channels: An Example Considering Debris Flows in the Oregon Coastal Province." Forest Science 53(2): 239-253.
Burnett, K. M. and D. J. Miller (2007). "Streamside Policies for Headwater Channels: An Example Considering Debris Flows in the Oregon Coast Range Province." Forest Science 53(239-253).
Burnett, K. M. and D. J. Miller (2007). "Streamside policies for headwater channels: an example considering debris flows in the Oregon Coastal Province." Forest Science 53(2): 239-253.
Management policies are increasingly debated for headwater channels given their prevalence and ecological importance in many landscapes. Quantitative differences among headwater channels may offer an objective basis for prioritizing streamside protection. Here, we examine differences among headwater channels as potential transport corridors for debris flows. Specifically, we model differences among hill slopes and headwater channels in probabilities of initiating and being traversed by debris flows that deliver to fish-bearing channels. We develop an approach to rank these probabilities and apply the ranks in delineating alternative streamside management zones. Initiation and traversal probabilities are estimated from an empirically calibrated debris-flow model using regionally available 10-m digital elevation data. Alternatives are delineated by encompassing 25%, 50%, and 75% of debris-flow susceptible hill slopes and headwater channels. Highest initiation and traversal probabilities were contained in a relatively small percentage of the study area. Encompassing lower probabilities required disproportionately larger areas. Substituting delineated alternatives for currently prescribed headwater riparian management zones decreased the total area encompassed on federal lands but increased it on private and state lands. Our intent is not to advocate for any particular alternative but to demonstrate that knowledge about how headwater channels differ over large areas can help tailor riparian policies.
 
Burnett, K. M., et al. (2007). "Distribution of salmon-habitat potential relative to landscape characteristics and implications for conservation." Ecological Applications 17(1): 66-80.
Burnett, K. M., et al. (2006). Comparing Riparian and Catchment Influences on Stream Habitat in a Forested, Montane Landscape. Landscape Influences on Stream Habitats and Biological Assemblages. R. M. Hughes, L. Wang and P. W. Seelbach. Bethesda, MD, American Fisheries Society: 175-197.
Burnett, K. M., et al. (2003). A first step toward broad-scale identification of freshwater protected areas for Pacific Salmon and Trout in Oregon, USA. Aquatic Protected Areas: what works best and how do we know? Proceedings of the World Congress on Aquatic Protected Areas, Cairns, Australia, August 2002. A. Grant and D. C. Smith. North Beach, WA, Australia, Australian Society for Fish Biology: 144-154.
Burnett, K. M., et al. (2007). "Distribution of salmon-habitat potential relative to landscape characteristics and implications for conservation." Ecological Applications 17(1): 66-80.
The geographic distribution of stream reaches with potential to support high-quality habitat for salmonids has bearing on the actual status of habitats and populations over broad spatial extents. As part of the Coastal Landscape Analysis and Modeling Study (CLAMS), we examined how salmon-habitat potential was distributed relative to current and future (+100 years) landscape characteristics in the Coastal Province of Oregon, USA. The intrinsic potential to provide high-quality rearing habitat was modeled for juvenile coho salmon (Oncorhynchus kisutch) and juvenile steelhead (O. mykiss) based on stream flow, valley constraint, and stream gradient. Land ownership, use, and cover were summarized for 100-m analysis buffers on either side of stream reaches with high intrinsic potential and in the overall area encompassing the buffers. Past management seems to have concentrated nonindustrial private ownership, agriculture, and developed uses adjacent to reaches with high intrinsic potential for coho salmon. Thus, of the area in coho salmon buffers, 45% is either non-forested or recently logged, but only 10% is in larger-diameter forests. For the area in steelhead buffers, 21% is either non-forested or recently logged while 20% is in larger-diameter forests. Older forests are most extensive on federal lands but are rare on private lands, highlighting the critical role for public lands in near-term salmon conservation. Agriculture and development are projected to remain focused near high-intrinsic-potential reaches for coho salmon, increasing the importance of effectively addressing nonpoint source pollution from these uses. Percentages of larger-diameter forests are expected to increase throughout the province, but the increase will be only half as much in coho salmon buffers as in steelhead buffers. Most of the increase is projected for public lands, where policies emphasize biodiversity protection. Results suggest that widespread recovery of coho salmon is unlikely unless habitat can be improved in high-intrinsic-potential reaches on private lands. Knowing where high-intrinsic-potential stream reaches occur relative to landscape characteristics can help in evaluating the current and future condition of freshwater habitat, explaining differences between species in population status and risk, and assessing the need for and feasibility of restoration.
 
Burnett, M. R., et al. (1998). "The Influence of Geomorphological Heterogeneity of Biodiversity: I. A Patch-Scale Perspective." Conservation Biology 12(2): 363-370.
Burnham, K. P. and D. R. Anderson (2002). Model Selection and Multimodel Inference. New York, Springer.
Burroughs, E., Jr. (1985). Landslide Hazard Rating for the Oregon Coast Range. Watershed Management for the Eighties, Denver, Colorado.
Burroughs, E. R., Jr. (1984). Landslide hazard rating for portions of the Oregon Coast Range. Symposium on Effects of Forest Land Use on Erosion and Slope Stability, Univ. of Hawaii, Honolulu, Environment and Policy Institution.
Burroughs, E. R. and B. R. Thomas (1977). "Declining root strength in Douglas-fir after felling as a factor in slope stability." U.S.D.A. Forest Service
Research Paper INT-190: 27.
Burroughs, E. R. J. and B. R. Thomas (1977). Declining Root Strength in Douglas-fir After Felling as a Factor in Slope Stability. Odgen, UT, US Department of Agriculture, Forest Service: 27.
Bursik, M. I. and A. R. Gillespie (1993). "Late Pleistocene Glaciation on Mono Basin, California." Quaternary Research 39: 24-35.
Burton, C. L. a. M. S. R. (2003). "Decision support to assist environmental sedimentology modelling." enivonmental geology 43: 457-465.
Burton, T. A. (2005). "Fish and stream habitat risks from uncharacteristic wildfire: observations from 17 years of fire-related disturbances on the Boise National Forest, Idaho." Forest Ecology and Management 211: 140-149.
Several large, uncharacteristic wildfires occurred on the Boise National Forest in Southwest Idaho, from 1986 to 2003. From
1987 to 1994, severe wildfires burned almost 50% of the ponderosa pine forest types (about 200,000 ha). The intensity of the
fires varied across the landscape, with a mix of low to moderate severity, and lesser amounts of high burn severity. After the fires,
localized debris flows favored smaller order streams in watersheds less than 4000 ha in size, where there had been mostly high
severity burning. Locally, areas experiencing high heat and post-fire debris flows had reduced fish numbers and altered fish
habitats. Uncharacteristic wildfires on the managed portions of the Boise National Forest appeared to have more pronounced,
short-term effects on fish habitats as compared with characteristic wildfires in the Central Idaho Wilderness. Even in the most
severely impacted streams, habitat conditions and trout populations improved dramatically within 5–10 years. Post-fire floods
apparently rejuvenated stream habitats by exporting fine sediments and by importing large amounts of gravel, cobble, woody
debris, and nutrients, resulting in higher fish productivities than before the fire. These observations suggest that important
elements of biodiversity and fish productivity may be influenced, or even created by fire-related disturbances. In some cases,
habitats that were completely devoid of salmonid fishes just after the debris floods, were later re-colonized with migrants
returning from downstream or nearby tributary rearing habitats. Re-population was likely enhanced by higher fecundity, homing
instinct, and greater mobility of the larger migratory fish. Ecosystem restoration activities that reduce both short- and long-term
threats of uncharacteristic wildfire on imperiled fishes could be emphasized in areas where local populations may be weak and/or
isolated, but potentially recoverable. But forest ecosystem restoration alone may not reduce risks to fish if existing habitat
conditions and isolation are limiting the population.
 
 
Bush, G. (1983). Landslide Survey - 1981-1982, Summary sheet and worksheets. Corvallis, Oregon, Siuslaw National Forest Headquarters.
Bush, G., et al. (1997). Assessment of the effects of the 1996 flood on the Siuslaw National Forest. Corvallis, OR, USDA Forest Service, Siuslaw National Forest: 47.
Buss, K. (1989). Some General Aspects of Debris Flows (DRAFT).
Butler, D. R. (1989). "Snow Avalanche-Dams and Resultant Hazards in Glacier National Park, Montana." Northwest Science 63(3): 109-115.
Butler, J. J. J., et al. (2007). "A field investigation of phreatophyte-induced fluctuations in the water table." Water Resources Research 43.
Hydrographs from shallow wells in vegetated riparian zones frequently display a
distinctive pattern of diurnal water table fluctuations produced by variations in plant water
use. A multisite investigation assessed the major controls on these fluctuations and the
ecohydrologic insights that can be gleaned from them. Spatial and temporal variations in
the amplitude of the fluctuations are primarily a function of variations in (1) the
meteorological drivers of plant water use, (2) vegetation density, type, and vitality, and
(3) the specific yield of sediments in the vicinity of the water table. Past hydrologic
conditions experienced by the riparian zone vegetation, either in previous years or earlier
within the same growing season, are also an important control. Diurnal water table
fluctuations can be considered a diagnostic indicator of groundwater consumption by
phreatophytes at most sites, so the information embedded within these fluctuations should
be more widely exploited in ecohydrologic studies.
 
Butler, P. R. (1977). "Movement of cobblesin a gravel-bed stream during a flood season." GSA Bull 88: 1072-1074.
In a gravel-bed perennial stream in western Wyoming, 159 cobbles were tagged and placed on the stream bed. Thirty-five percent were recovered after one flood season during which the peak flow was 34 m3/s. Sixty-one percent of those recovered had been buried. The b axis of the tagged rocks ranged from 34 mm to 116 mm. The distance transported ranged from 0 to 420 m; 95% of the recovered cobbles had moved. There is no clear relationship between particle size and distance transported; distance of transport seems to be more closely related to position (stream edge or center) at time of entrainment. Susceptibility to burial seems to be related to both particle size and position at time of entrainment.
 
Buttle, J. M. and D. J. McDonald (2002). "Coupled vertical and lateral preferential flow on a forested slope." Water Resources Research 35(5): 1060, doi:1010,1029/2001 WR000773,002002.
Butturini, A., et al. (2002). "The influence of riparian-hyporheic zone on the hydrological responses in an intermittent stream." Hydrology and Earth System Sciences 6(3): 515-525.
Stream and riparian groundwater hydrology has been studied in a small intermittent stream draining a forested catchment for a system
representative of a Mediterranean climate. The relationship between precipitation and stream runoff and the interactions between stream
water and the surrounding riparian groundwater have been analysed under a wide spectrum of meteorological conditions. The hypothesis that
the hydrological condition of the near-stream groundwater compartment can regulate the runoff generation during precipitation events was
tested. Stream runoff is characterised by a summer dry period, and precipitation input explained only 25% of runoff variability over the study
period (r2=0.25, d.f.=51, p<0.001). The variability of precipitation v. stream runoff is explained partly by the hydrogeological properties of
the riparian near-stream zone. This zone is characterised by high hydrological conductivity values and abrupt changes in groundwater level
in summer. The summer dry period begins with a rapid decrease in near-stream groundwater level, and ends just after the first autumnal rain
when the original groundwater level recovers suddenly. Within this period, storms do not cause major stream runoff since water infiltrates
rapidly into the riparian compartment until it is refilled during the subsequent winter and spring; then the precipitation explains the 80% of
the stream runoff variability (r2=0.80, d.f.=34, p<0.001). These results suggest that the hydrological interaction between the riparian groundwater
compartment and the stream channel is important in elucidating the hydrological responses during drought periods in small Mediterranean
streams.
 
 
Byron, E. R. and C. R. Goldman (1989). "Land-Use and Water Quality in Tributary Streams of Lake Tahoe, California-Nevada." Journal of Enivironmental Quality 18: 84-88.
C. H. Luce, T. A. B. (1999). "Sediment production from forest roads in western Oregon." Water  Resources Research 35(8): 2561-2570.
C.R.W.Q.C.B. (2002). Mattole River Watershed Technical Support Document for the TMDL for Sediment (Draft). North Coast region, California Regional Water Quality Control Board.
Cafferata, P. H. and T. E. Spittler (1998). Logging Impacts of the 1970's vs. the 1990's in the Caspar Creek Watershed, US Forest Service: 103-112.
Cafferata, P. H., et al. (1998). Logging impacts of the 1970's vs. the 1990's in the Caspar Creek watershed. General Technical Report PSW, Report: PSW-GTR-168: 103-115.
The Caspar Creek watershed study provides resource professionals with information regarding the impacts of timber operations conducted under varying forest practices on sensitive aquatic habitats. In the South Fork watershed, roads were constructed near watercourse channels in the 1960's, and the watershed was selectively logged using tractors during the early 1970's. Subwatersheds in the North Fork were clearcut from 1985 to 1991 using predominantly cable yarding and roads located high on ridges. Numerous landslides were documented after road construction and logging in the South Fork owing to inadequate road, skid trail, and landing design, placement, and construction. In contrast, the size and number of landslides after timber operations in the North Fork to date have been similar in logged and unlogged units. Considerably more hillslope erosion and sediment yield have also been documented after logging operations in the South Fork, when compared to the North Fork. An analysis of the storm events associated with documented landslides showed that high 3-day or 10-day precipitation totals in combination with moderately high 1-day amounts have been more important than very high 1-day totals alone in triggering debris sliding at Caspar Creek. Storm sequences meeting the criteria required for causing documented landslides were found to have occurred in all phases of the 36-year study, with the greatest number occurring in water year 1998. Numerous large landslides associated with the road system in the South Fork occurred in early 1998, indicating that "legacy" roads continue to be significant sources of sediment decades after they were constructed.
 
Cain, D. H., et al. (1997). "A multi-scale analysis of landscape statistics." Landscape Ecology 12: 199-212.
Caine, N. (1980). "The rainfall intensity - duration control of shallow landslides and debris flows." Geografiska Annaler 62: 23-27.
Caine, N. (1990). "Rainfall intensity - duration control on shallow landslides and debris flows." Geografika Annaler 62A: 23-27.
Cairns, J., Jr. (1990). "Lack of Theoretical Basis for Predicting Rate and Pathways of Recovery." Environmental Management 14(5): 517-526.
CALFED (1998). Ecosystem Restoration Program Plan, CALFED: 40-46.
California Department of Conservation (1997). Factors affecting landslides in forested terrain California Department of Conservation: 1-5.
California Department of Water Resources (1975). Van Duzen River Basin Environmental Atlas, Prepared in cooperation with Humboldt County.
Caminiti, J. E. (2004). "Catchment modelling - a resource manager's perspective." Environmental Modelling & Software 19(11): 991-996.
Campbell, A. J. and R. C. Sidle (1985). "Bedload transport in a pool-riffle sequence of a coastal Alaska stream." Water Resources Bulletin 21(4): 579-590.
Campbell, C. S. (1989). "Self-lubrication for long runout landslides." The Journal of Geology 97(6): 653-665.
Campbell, D. H. (1984). The transport of road derived sediment as a function of slope characteristics and time. Department of Earth Resources. Fort Collins, CO, M.S. Thesis, Colorado State University: 46.
Campbell, R. H. (1975). Soil slips, debris flows, and rainstorms in the Santa Monica mountains and vicinity, Southern California, US Geologic Survey: 1-51.
Campo, J., V. Andreu, E. Gimeno-Garcia, O. Gonzalez, and J.L. Rubio (2006). "Occurrence of soil erosion after repeated experimental fires in a Mediterranean environment." Geomorphology 82(376-386).
Cannon, S. H. (1993). An empirical model for the volume-change behavior of debris flows. Proceedings, Hydraulic Engineering '93. H. W. Shen, S. T. Su and F. Wen. New York, American Society of Civil Engineers. 2: 1768-1777.
Cannon, S. H. (2001). "Debris-flow generation from recently burned watersheds." Environmental & Engineering Geoscience 7(4): 321-341.
Cannon, S. H. and S. Ellen "Rainfall conditions for abundant debris avalanches San Francisco bay region, California." California Geology: 267-272.
Cannon, S. H. and S. Ellen (1985). "Abundant Debris Avalanches." California Geology: 267-272.
Cannon, S. H. and J. E. Gartner (2005). Wildfire-related debris flows from a hazards perspective (Chapter 15). Heidelberg, Springer-Praxis.
Cannon, S. H., et al. (2007). Methods for the emergency assessment of debris-flow hazards from basins burned by the fires of 2007, southern California: U.S. Geological Survey Open File Report 2007-1384: 10.
This report describes the approach used to assess potential debris-flow hazards from basins burned by the Buckweed, Santiago, Canyon, Poomacha, Ranch, Harris, Witch, Rice, Ammo, Slide, Grass Valley and Cajon Fires of 2007 in southern California. The assessments will be presented as a series of maps showing a relative ranking of the predicted volume of debris flows that can issue from basin outlets in response to a 3-hour duration rainstorm with a 10-year return period. Potential volumes of debris flows are calculated using a multiple-regression model that describes debris-flow volume at a basin outlet as a function of measures of basin gradient, burn extent, and storm rainfall. This assessment provides critical information for issuing basin-specific warnings, locating and designing mitigation measures, and planning of evacuation timing and routes.
 
Cannon, S. H., et al. (2003). Emergency assessment of debris-flow hazards from basins burned by the Grand Prix and Old Fires of 2003, Southern California. U.S.G.S. Open File Report 03-475.
Cannon, S. H., et al. (2008). "Storm rainfall conditions for floods and debris flows from recently burned areas in southwestern Colorado and southern California." Geomorphology 96: 250-269.
Debris flows generated during rain storms on recently burned areas have destroyed lives and property throughout the
Western U.S. Field evidence indicate that unlike landslide-triggered debris flows, these events have no identifiable
initiation source and can occur with little or no antecedent moisture. Using rain gage and response data from five fires in
Colorado and southern California, we document the rainfall conditions that have triggered post-fire debris flows and
develop empirical rainfall intensity–duration thresholds for the occurrence of debris flows and floods following wildfires in
these settings. This information can provide guidance for warning systems and planning for emergency response in similar
settings.
Debris flows were produced from 25 recently burned basins in Colorado in response to 13 short-duration, high-intensity
convective storms. Debris flows were triggered after as little as six to 10 min of storm rainfall. About 80% of the storms that
generated debris flows lasted less than 3 h, with most of the rain falling in less than 1 h. The storms triggering debris flows ranged in
average intensity between 1.0 and 32.0 mm/h, and had recurrence intervals of two years or less. Threshold rainfall conditions for
floods and debris flows sufficiently large to pose threats to life and property from recently burned areas in south-central, and
southwestern, Colorado are defined by: I=6.5D− 0.7 and I=9.5D− 0.7, respectively, where I=rainfall intensity (in mm/h) and
D=duration (in hours).
Debris flows were generated from 68 recently burned areas in southern California in response to long-duration frontal
storms. The flows occurred after as little as two hours, and up to 16 h, of low-intensity (2–10 mm/h) rainfall. The storms
lasted between 5.5 and 33 h, with average intensities between 1.3 and 20.4 mm/h, and had recurrence intervals of two
years or less. Threshold rainfall conditions for life- and property-threatening floods and debris flows during the first
winter season following fires in Ventura County, and in the San Bernardino, San Gabriel and San Jacinto Mountains of
southern California are defined by I= 12.5D−0.4, and I=7.2D−0.4, respectively. A threshold defined for flood and debrisflow
conditions following a year of vegetative recovery and sediment removal for the San Bernardino, San Gabriel and
San Jacinto Mountains of I=14.0D−0.5 is approximately 25 mm/h higher than that developed for the first year following
fires.
 
Cannon, S. H., et al. (2001). "Wildfire-related debris-flow initiation processes, Storm King Mountain, Colorado." Geomorphology 39: 171-188.
Cannon, S. H. and S. I. Reneau (2000). "Conditions for generation of fire-related debris flows, Capulin Canyon, New Mexico." Earth Surface Processes and Landforms 25: 1103-1121.
Cannon, S. H. and W. Z. Savage (1988). "A mass-change model for the estimation of debris-flow runout." Journal of Geology 96: 221-227.
Canon, J., Dominguez, F., and J.B. Valdez ( (submitted)). "Downscaling climate variability associated with quasi-periodic climate signals: a new statistical approach. ." Climate Change.
Canon, J. J., Gonzalez, and J. Valdez (2007). "Precipitation in the Colorado River Basin and its low frequency associations with PDO and ENSO signals." Journal of Hydrology 333(24): 252-264.
Capra, A., L.M. Mazzara, and B. Scicolone (2005). "Application of the EGEM model to predict gully erosion in Sicily, Italy." Catena 59: 133-146.
Carling, P. A. (1983). "Threshold of coarse sediment transport in broad and narrow natural steams." Earth Surface Processes and Landforms 8: 1-18.
Carling, P. A. (1987). "Hydrodynamic Interpretation of a Boulder Berm and Associated Debris-Torrent Deposits." Geomorphology 1: 53-67.
Carling, P. A. (?). In-stream Hydraulics and Sediment Transport. ?: 101-125.
Carling, P. A., et al. (1998). "Coarse Bedload Transport in a Mountain River." Earth Surface Processes and Landforms 23: 141-157.
Carling PA, J. W., A Kelsey, MS Glaister, and HG Orr (1998). "Coarse bedload transport in a mountain river." Earth Surface Processes and Landforms 23: 141-157.
Carrara, A., et al. (1991). "GIS techniques and statistical models in evaluating landslide hazard." Earth Surface Processes and Landforms 16: 427-445.
Carson, B. (1985). Erosion and Sedimentation Processes in the Nepalese Himalaya. Kathmandu, Nepal, International Centre for Integrated Mountain Development: 39.
Carson, M. A. (1966). "The magnitude of variability in samples of certain geomorphic characteristics drawn from valley-side slopes."?: 93-100.
Carson, M. A. (1987). "Measures of flow intensity as predictors of bedload." Journal of Hydraulic Engineering 113(11): 1402-1421.
Carson, M. A. and G. A. Griffiths (1985). "Tractive stress and the onset of bed particle movement in gravel stream channels: different equations for different purposes." Journal of Hydrology 79: 375-388.
Carson, M. A. and G. A. Griffiths (1987). "Bedload Transport in Gravel Channels." Journal of Hydrology 26(1): 1-151.
Carson, W. W. and S. E. Reutebuch (1997). A rigorous test of the accuracy of USGS digital elevation models in forested areas of Oregon and Washington. 1997 ACSM/ASPRS Annual Convention & Exposition. Seattle, WA, American Congress of Surveying and Mapping; American Society for Photogrammetry & Remote Sensing. 1: 133-143.
A procedure for performing a rigorous test of elevational accuracy of DEMs using independent ground coordinate data digitized photogrammetrically from aerial photography is presented. The accuracy of a sample set of 23 DEMs covering National Forests in Oregon and Washington was evaluated. Accuracy varied considerably between eastern and western parts of Oregon and Washington, and to a lesser extent, by DEM production method. The elevational root mean square errors (RMSE) computed from independent ground data for DEMs produced using the line-trace method were on average 4 times larger than RMSEs published by the USGS for both eastside and westside DEMs. Computed RMSEs for all Westside line-trace DEMs exceeded 7m; whereas,80% of eastside line-trace DEMs had computed RMSEs of 7m or less.
 
Casadei, M., et al. (2003). "Testing a model for predicting the timing and location of shallow landslide initiation in soil-mantled landscapes." Earth Surface Processes and Landforms 28(DOI:10.1002/esp.470): 925-950.
Casali, J., J.J. Lopez, and J.V. Giraldez (1999). "Ephemeral gully erosion in southern Navarra (Spain)." Catena 36: 65-84.
Cashman, S. M., et al. (1995). Geology of the Redwood Creek Basin, Humboldt County, California. Geomorphic Processes and Aquatic Habitat in the Redwood Creek Basin.
Castro, J. M. and P. L. Jackson (2001). "Bankfull discharge recurrence intervals and regional hydraulic geometry relationships: patterns in the Pacific Northwest, USA." Journal of the American Water Resources Association 37(5): 1249-1262.
Caswell, H. (1976). The Validation Problem. Systems Analysis and Simulation in Ecology. B. C. Patton, Academic Press: 313-325.
Cederholm, C. J., et al. (1997). "Response of juvenile coho salmon and steelhead to placement of large woody debris in a coastal Washington stream." North American Journal of Fisheries Management 17: 947-963.
Many fish habitats have been altered in Pacific Northwest streams and rivers over the
past century by a variety of land use practices, including forestry, urbanization, agriculture, and channelization.
There are research and management needs for evaluation of the effectiveness of rehabilitation
projects intended to enhance stream fish habitat recovery. The response of populations of juvenile coho
salmon Oncorhynchus kisutch and steelhead O. mykiss to addition of large woody debris (LWD) was
tested in North Fork Porter Creek (NFPC). a small coastal tributary of the Chehalis River. Washington.
The NFPC was divided into three 500-m study sections; two sections were altered with two approaches
(engineered and logger's choice) to adding LWD. and the third was kept as a reference site. Immediately
after LWD addition, the abundance of LWD pieces was 7.9 limes greater than the pretreatment level
in the engineered site and 2.7 times greater in the logger's choice site; abundance was unchanged in
the reference site. Subsequent winter storms brought additional LWD into all three study sites. In the
years that followed, the amount of pool surface area increased significantly in both the engineered and
logger's choice sites, while it decreased slightly in the reference site. After LWD addition, winter
populations of juvenile coho salmon increased significantly in the engineered and logger's choice sites,
while they remained the same in the reference site. There were no significant differences in the coho
salmon populations during spring and autumn within the reference, engineered, or logger's choice sites.
The coho salmon smolt yield from the engineered and logger's choice sites also increased significantly
after LWD addition, while it decreased slightly in the reference site. After LWD addition, the reference
site and the engineered site both exhibited increases in age-0 steelhead populations; however, the
population in the logger's choice site did not change. There was no difference in age-1 steelhead
abundance among sites, or before and after enhancement during any season. Winter populations of
juvenile coho salmon and age-0 steelhead were related inversely to maximum and mean winter discharge.
 
 
Cederholm, C. J. and L. M. Reid (1987). Impact of forest management on coho salmon (Oncorhynchus kisutch) populations of the Clearwater River, Washington: A project summary. Streamside Management, Forestry and Fishery Interactions. E. O. Salo and T. W. Cundy. Seattle, Institute of Forest Resources, University of Washington: 373-398.
Cederholm, C. J., et al. (1980). Effects of Forest Road Erosion on Salmonid Spawning Gravel Composition and Populations of the Clearwater River, Washington. Salmon-Spawning Gravel: A Renewable Resource in the Pacific Northwest?, Seattle, Washington.
Cenderelli, D. A. and J. S. Kite (1998). "Geomorphic effects of large debris flows on channel morphology at North Fork Mountain, Eastern West Virginia, USA." Earth Surface Processes and Landforms 23: 1-19.
Cendrero, A. and F. Dramis (1996). "The contribution of landslides to landscape evolution in Europe." Geomorphology 15: 191-211.
Center for Streamside Studies (1988). Slope Stability and Forest Management. Seattle, Washington, University of Washington.
CEQ (1997). Considering cumulative effects under the National Environmental Policy Act. C. o. E. Quality. Washington, DC, Council on Environmental Quality: 64pp.
Cerdan, O., et al. (2002). "Rill erosion on cultivated hillslopes during two extreme rainfall events in Normandy, France." Soil & Tillage Research 67(1): 99-108.
Cerdan, O., et al. (2002). "Modelling interrill erosion in small cultivated catchments." Hydrological Processes 16(16): 3215-3226.
Cerdan, O., et al. (2002). "Sediment concentration in interrill flow: Interactions between soil surface conditions, vegetation and rainfall." Earth Surface Processes and Landforms 27(2): 193-205.
Cerveny, P. F., et al. (1988). History of Uplift and Relief of the Himalaya During the Past 18 Million Years: Evidence from Fission-Track Ages of Detrital Zircons from Sandstones of the Siwalik Group. New Perspectives in Basin Analysis. K. L. Kleinsphen and C. Paola, Springer-Verlag.
Chakraborti, R. K., et al. "Evaluation of aggregate properties in suspensions from in-situ measurements."
Size and density of flocs are important to evaluate the settling rate in natural waters. A study was conducted to determine the floc properties in a coagulation-flocculation mechanism of fine-grained suspended particles from the analysis of aggregate structure. The properties of large suspended flocs were determined from the images using a newly developed non-intrusive imaging technique. This analysis provides a direct description of the shape and size of floc. These flocs tend to be large, highly porous and irregularly shaped, and are described by fractal dimensions. Several important characteristics of the aggregates, including density, porosity and number of primary particles in an aggregate are calculated using fractal approach. For larger aggregates with smaller fractal dimensions, floc density and porosity demonstrate strong negative and positive relationships, respectively. A power law relationship between floc density and porosity has been found which lead to floc settling speed. A relationship among the solid content, floc density and the number of primary particles are established. Greater understanding of the relationship between the geometrical properties of aggregates and their behavior in flocculating systems increases our ability to understand flocculation mechanisms and the resulting sedimentation rate.
 
Chamberlin, T. W. (1982). Influence of Forest and Rangeland Management on Anadromous Fish Habitat in Western North America: Timber Harvest. Portland, Oregon, Forest Service, U.S. Department of Agriculture: 30.
Chang, H. H. (1984). "Modeling of River Channel Changes." Journal of Hydraulic Engineering 110(2): 157-172.
Chang, H. H. (1994). "Selection of Gravel-Transport Formula for Stream Modeling." Journal of Hydraulic Engineering 120(5): 646-651.
Chapman, D. W. (1988). "Critical Review of Variables Used to Define Effects of Fines in Redds of Large Salmonids." Transactions of the American Fisheries Society 117(1): 1-21.
Chapman, D. W., et al. (1986). "Effects of River Flow on the Distribution of Chinook Salmon Redds " Transactions of the American Fisheries Society 115: 537-547.
Chapman, T. G. (1996). "Common unitgraphs for sets of runoff events. Part 1: Unitgraph identification from streamflow data." Hydrological Processes 10: 773-782.
Charron, I. and E. A. Johnson (2006). "The importance of fires and floods on tree ages along mountainous gravel-bed streams." Ecological Applications 16(5): 1757-1770.
This paper examines the commonly accepted assumption in the riparian
literature that areas adjacent to streams do not burn.Using time-since-fire distributions,
derived from stand-origin maps for a watershed in the front ranges of the Canadian Rocky
Mountains,we found that the areas adjacent to streams and the whole study watershed have
similar fire frequencies.In addition,the relative importance of fires and floods is regulated by a
change in channel morphology associated with the creation of bars.The results demonstrate
that fires solely control tree establishment along straight streams without bars,while the
in fluence of floods is observed at the onset of lateral-and point-bar formation.This occurs
because bars are formed in-channel and require smaller discharges in order to be flooded,
compared to higher terraces.Consequently,bars are the only surfaces being flooded more
frequently than they are being burned.Thus,overall the results indicate that,on this
watershed,areas adjacent to streams are not less likely to burn than the uplands,except for
lateral and point bars.The generality of these results to other systems should be tested as they
have important implications for current forest ecological de finition of ‘‘riparian zones,’’which
typically include all fluvially derived landforms,from the channel banks to the terraces.
Indeed,this study suggests that along smaller,headwater,gravel-bed mountain watersheds,
the forests found on terraces are only in fluenced by fire and not fluvial processes and should
therefore not be included in the riparian zone,while the forests on bars are the only surfaces
currently being in fluenced by fluvial processes.Such a change in de finition has implications for
both ecologists and forest managers aiming to protect areas along streams as they now must
take into account the effects of two disturbances on these small gravel-bed streams.
 
Chatwin, S. C., et al. (1994). A Guide for Management of Landslide-Prone Terrain in the Pacific Northwest. Victoria, British Columbia, Canada, Crown Publications Inc.
546 Yates Street
Victoria, B.C., V8W 1K8.
Chaudhury, P. and J. Tumblin (2003). The trilateral filter for high contrast images and meshes. Eurographics Symposium on Rendering. P. Christensen and D. Cohen-Or: 1-11.
We present a new, single-pass nonlinear  lter for edge-preserving smoothing and visual detail removal for N
dimensional signals in computer graphics, image processing and computer vision applications. Built from two
modi ed forms of Tomasi and Manduchi's bilateral  lter, the new .trilateral.  lter smoothes signals towards
a sharply-bounded, piecewise-linear approximation. Unlike bilateral  lters or anisotropic diffusion methods that
smooth towards piecewise constant solutions, the trilateral  lter provides stronger noise reduction and better outlier
rejection in high-gradient regions, and it mimics the edge-limited smoothing behavior of shock-forming PDEs by
region  nding with a fast min-max stack. Yet the trilateral  lter requires only one user-set parameter,  lters an
input signal in a single pass, and does not use an iterative solver as required by most PDE methods. Like the
bilateral  lter, the trilateral  lter easily extends to N-dimensional signals, yet it also offers better performance
for many visual applications including appearance-preserving contrast reduction problems for digital photography
and denoising polygonal meshes.
 
Chen, J. (1998). "The simulation of runoff using a time-area graph derived from DTM." ITC Journal 2: 113-117.
Chen, J. C. and C. D. Jan (2003). Probabilistic equation of critical slope for debris-flow occurrence. Debris-Flow Hazards Mitigation: Mechanics, Prediction, and Assessment. D. Rickenmann and C. Chen. Rotterdam, Millpress. 1: 83-89.
Chen, X., et al. (2008). "Effects of large woody debris on surface structure and aquatic habitat in forested streams, southern interior British Columbia, Canada." River Research and Applications 24: 862-875.
It is well known that large woody debris (LWD) plays an important functional role in aquatic organisms’ life. However, the
influence of LWD on channel morphology and aquatic environments at watershed levels is still unclear. The relationships
between wood and surface structure and aquatic habitat in 35 first through fifth order streams of southern interior British
Columbia were investigated. Study streams in the channel networks of the study watersheds were classified into four size
categories based on stream order and bankfull width: Stream size I: bankfull width was less than 3m, Stream size II: 3–5 m,
Stream size III: 5–7 m, Stream size IV: larger than 7m.We found the number of functional pieces increased with stream size and
wood surface area in stream sizes I, II and III (24, 28 and 25m2/100m2, respectively) was significantly higher than that in stream
size IV (12m2/100m2). The contribution of wood pieces to pool formation was 75% and 85% in stream sizes II and III,
respectively, which was significantly higher than those in stream size I (50%) and size IV (25%). Between 21% and 25% of wood
pieces were associated with storing sediment, and between 20% and 29% of pieces were involved in channel bank stability in all
study streams. Due to long-term interactions, LWD in the intermediate sized streams (Size II and III) exhibited much effect on
channel surface structure and aquatic habitats in the studied watersheds. Copyright
 
Chen, Y. D. (2004). "Watershed modeling: Where are we heading?" Environmental Informatics Archives 2: 132-139.
Cheng, J. D. (1989). "Streamflow Changes After Clear-Cut Logging of a Pine Beetle-Infested Watershed in Southern British Columbia, Canada." Water Resources Research 25(3): 449-456.
Chernick, M. R. (1999). Bootstrap Methods. A Practitioner's Guide. New York, John Wiley & Sons, Inc.
Chorley, R. J. and B. A. Kennedy (1971). Physical Geography, A Systems Approach. London, Prentice Hall.
Christiansen, N. L., et al. (1989). "Interpreting the Yellowstone Fires of 1988." BioScience 39(10): 678-685.
ChuanWen, L., et al. Construction of GIS forest themes and their application to stream protective forest planning. Dongbei Linye Daxue Xuebao 28 (2), 4-8 pp.;   4 ref.; 2000.
GIS (Geographic information system) was developed for the forest of the Maor Mountain range in [Heilongjiang Province] China, with details of the characteristics and forest resources of each compartment and subcompartment from remote sensing data. A serial number system was allocated to the streams, compartments and subcompartments. The data in the GIS was used to classify and plan stream protective forest, based on the land occupation rate of protective forest per metre (LOR). The protective forest of the streams in the Beilin range of Maor Mountain forest farm was compartmentalized according to LOR, and a planning map was plotted directly.
 
Chung, C. J. and A. G. Fabbri (2003). "Validation of spatial prediction models for landslide hazard mapping." Natural Hazards 30: 451-472.
Chung, C. J. and A. G. Fabbri (2005). Systematic procedures of landslide-hazard mapping for risk assessment using spatial prediction models. Landslide Hazard and Risk. T. Glade, M. G. Anderson and M. J. Crozier. Chichester, England, John Wiley & Sons, Ltd.: 139-174.
Church, M. (1980). Records of recent geomorpholgical events, in Cullingford. Time Scales in Geomorphology. R. A. D. Cullingford, D.A.,  and J. Lewin, John Wiley: 13-29.
Church, M. (1983). Pattern of instability in a wandering gravel bed channel. Modern and ancient fluvial systems. J. D. Collinson and J. Lewin. Oxford, Blackwell. 6: 169-180.
Church, M., J.F. Wolcott, and W.K. Fletcher (1991). "A test of equal mobility in fluvial sediment transport: behavior of the sand fraction." Water Resources Research 27(11): 2941-2951.
Church, M. (1998). The Landscape of the Pacific Northwest. Carnation Creek and Queen Charlotte Islands Fish/Forestry Workshop: Applying 20 Years of Coastal Research to Management Solutions, Queen Charlotte City, B. C., B.C. Ministry of Forests.
Church, M. (2002). "Geomorphic thresholds in riverine landscapes." Freshwater Biology 47(4): 541-557.
Church, M. (2006). "Bed material transport and the morphology of alluvial river channels." Annual Reviews of Earth and Planetary Sciences 34: 325-354.
The morphology of an alluvial river channel is the consequence of sediment trans-
port and sedimentation in the river.Morphological style is determined chie fly by
the caliber and quantity of sediment delivered to the channel,although modulated
by channel scale.Yet the relations between sediment transport and river morphol-
ogy have received only limited,qualitative attention.In this review,the problem is
studied by de fining sediment transport regimes on the basis of the Shields number,a
nondimensional measure of the capacity of the channel to move sediment of a given
caliber.The problem is also approached from an inverse perspective by which the
quantity and character of sediment deposits are used to infer details about the vari-
ation of sediment transport and sedimentation along a channel.Coupling the two
approaches establishes a basis to gain new insights into the origins of alluvial channel
morphology.
 
Church, M., et al. (1987). "Late Quaternary Paleohydrology of Canada." Canadian Journal of Earth Sciences 26: 1809-1859.
Church, M. and M. A. Hassan (1992). "Size and distance of travel of unconstrained clasts on a streambed." Water Resources Research 28(1): 299-303.
Church, M. and M. A. Hassan (2002). "Mobility of bed material in Harris Creek." Water Resources Research 38(11): 10.1029/2001WR000753.
Church, M., et al. (1998). "Stabilizing self-organized structures in gravelbed stream channels: field and experimental observations." Water Resources Research 34(11): 3169-3179.
Church, M. and M. A. Hassen (1992). "Size and distance of travel of unconstrained clasts on a streambed." Water Resources Research 28(1): 299-303.
Church, M. and R. Kellerhals (1978). "On the statistics of grain size variation along a gravel river." Canadian Journal of Earth Science 15: 1151-1160.
Church, M. and J. M. Ryder (1972). "Paraglacial Sedimentation: A Consideration of Fluvial Processes Conditioned by Glaciation." Geological Society of America Bulletin 83: 3059-3072.
Church, M. and O. Slaymaker (1989). "Disequilibrium of Holocene sediment yield in glaciated British Columbia." Nature 337(2): 452-454.
Church, M. A., et al. (1987). River bed gravels: sampling and analysis. Sediment Transport in Gravel-bed Rivers. C. R. Thorne, J. C. Bathurst and R. D. Hey, John Wiley & Sons Ltd.: 43-88.
The characterization of coarse bed material in rivers is difficult because the range of grain sizes is so wide that it is impractical to maintain a single method of measurement. Less obvious problems include the establishment of criteria for sample size, accounting for the layered nature of deposits and characterizing 'structural' features of the bed. The latter include imbrication, particle clustering and shadow deposits.
 
In this paper, standards for sample size in various measurement techniques are considered. Measurement methods are reviewed and the notion of 'truncated samples' is introduced as a means to overcome sample bias at the extremes of the size range. Replicability in sampling and analysis, fundamental desiderata for scientific measurements, are examined. Strategies are considered to achieve representative samples of satisfactory precision for various pruposes, and the problem posed by layers, particularly surface/subsurface contrasts, is introduced.
 
Because bed material samples are drawn from a very large and variable population, it is customary to interpret their meaning statistically. Since statistics is based upon number counts, this procedure may be applied within samples when individual clasts are measured: however, customary sieve procedures yield frequency by weight, which is not a well-conditioned statistical entity. Transformations are unhelpjul in practice since the numbers become absurdly large for small clasts. Strategies for interpreting these measurements are considered.
 
Cipra, B. (2000). "Revealing Uncertainties in Computer Models." Retrieved 6/6/00, 2000, from http://www.sciencemag.org/cgi/content/full/287/5455/960.
Cissel, J. H., et al. A disturbance-based landscape plan for a managed forest ecosystem: the Augusta Creek Study.
Cissel, J. H., et al. (1999). "Landscape Management Using Historical Fire Regimes: Blue River, Oregon." Ecological Applications 9(4): 1217-1231.
Clapp, E. M., et al. "Using super(10)Be and super(26)Al to determine sediment generation rates and identify sediment source areas in an arid region drainage basin."
We measured super(10)Be and super(26)Al in 64 sediment and bedrock samples collected throughout the arid, 187 km super(2) Yuma Wash drainage basin, southwestern Arizona. From the measurements, we determine long-term, time-integrated rates of upland sediment generation (81 plus or minus 5 g m super(-2) year super(-1)) and bedrock equivalent lowering (30 plus or minus 2 m Ma super(-1)) consistent with other estimates for regions of similar climate, lithology, and topography. In a small ( similar to 8 km super(2)), upland sub-basin, differences in nuclide concentrations between bedrock outcrops and hillslope colluvium suggest weathering of bedrock beneath a colluvial cover is a more significant source of sediment (40 x 10 super(4) kg year super(-1)) than weathering of exposed bedrock surfaces (10 x 10 super(4) kg year super(-1)). Mixing models constructed from nuclide concentrations of sediment reservoirs identify important sediment source areas. Hillslope colluvium is the dominant sediment source to the upper reaches of the sub-basin channel; channel cutting of alluvial terraces is the dominant source in the lower reaches. Similarities in nuclide concentrations of various sediment reservoirs indicate short sediment storage times (< 10 super(3) years). Nuclide concentrations, measured in channel sediment from tributaries of Yuma Wash and in samples collected along the length of the Wash, were used to construct mixing models and determine sediment sources to the main stem channel. We find an exponential decrease in the channel nuclide concentrations with distance downstream, suggesting that as much as 40% of sediment discharged from Yuma Wash has been recycled from storage within basin fill alluvium. Sediment generation and denudation rates determined from the main stem are greater (25%) than rates determined from upland sub-basins suggesting that, currently, sediment may be exported from the basin more quickly than it is being generated in the uplands. Independence of nuclide concentration and sediment grain size indicates that channels transport sediment in discrete pulses before rapidly depositing poorly sorted material, suggesting that differences in transport times for different size materials are minimal.
 
Clark, J. S. (1989). "Ecological disturbance as a renewal process: theory and application to fire history." Oikos 56: 17-30.
Clarke, S. and K. Burnett (2003). "Comparison of digital elevation models for aquatic data development." Photogrammetric Engineering and Remote Sensing 69(12): 1367-1375.
Clarke, S., et al. (2008). "Modeling streams and hydrogeomorphic attributes in Oregon from digital and field data." Journal of American Water Resources Association 44(2): 459-477.
Clarke, S. E., et al. (2008). "Modeling streams and hydrogeomorphic attributes in Oregon from digital and field data." Journal of the American Water Resources Association 44(2): 459-477.
Clarkson, R. W. and J. R. Wilson (1995). "Trout Biomass and Stream Habitat Relationships in the White Mountains Area, East-Central Arizona." Transactions of the American Fisheries Society 124: 599-612.
Clarr, W. P., et al. (1993). Sediment Entrapment by Stream Channel Vegetation. Management of Irrigation and Drainage Systems, Park City, Utah.
Clary, W. P., et al. (1993). Sediment Entrapment by Stream Channel Vegetation. Management of Irrigation and Drainage Systems, Park City, Utah.
Clayton, J. L. and W. F. Megahan (1997). "Natural erosion rates and their prediction in the Idaho batholith." Journal of the American Water Resources Association 33(3): 689-703.
Natural rates of surface erosion on forested granitic soils in central Idaho were measured in 40 m(2) bordered erosion plots over a period of four years. In addition, we measured a variety of site variables, son properties, and summer rainstorm intensities in order to relate erosion rates to site attributes. Median winter erosion rates are approximately twice summer period rates, however mean summer rates are nearly twice winter rates because of infrequent high erosion caused by summer rainstorms. Regression equation models and regression tree models were constructed to explore relationships between erosion and factors that control erosion rates. Ground cover is the single factor that has the greatest influence on erosion rates during both summer and winter periods. Rainstorm intensity (erosivity index) strongly influences summer erosion rates, even on soils with high ground cover percentages. Few summer storms were of sufficient duration and intensity to cause rilling on the plots, and the data set was too small to elucidate differences in rib vs. interrill erosion. The regression tree models are relatively less biased than the regression equations developed, and explained 70 and 84 percent of the variability in summer and winter erosion rates, respectively.
 
Cleland, D. T., et al. (1997). National hierarchical framework of ecological units. Ecosystem Management. M. S. B. a. A. Haney. New Haven, CT, Yale University: 361 p.
Clifton, H. E. Tips on talks or how to keep an audince attentive, alert, and around for the conclusions at a scientific meeting. American Association of Petroleum Geologists.
Coates, R. and L. Collins (1984). Streamside landsliding and channel change in a suburban forested watershed: effects of an extreme event. Proceedings of Symposium on Effects of Forest Land Use on Erosion and Sloppe Stability. Honolulu, East-West Center: 155-164.
Coats, R., et al. (1985). "Channel Change, Sediment Transport, and Fish Habitat in a Coastal Stream: Effects of an Extreme Event." Environmental Management 9(1): 35-48.
Cobourn, J. (1989). An application of cumulative watershed effects (CWE) analysis on the Eldorado National Forest in California. In Proceedings of the symposium on Headwaters Hydrology, Bethesda, MD, American Water Resources Association.
Cochran, W. G. (1977). Sampling Techniques. New York, John Wiley & Sons, Inc.
Cochrane, T. A., and D.C. Flanagan (1999). "Assessing water erosion in small watersheds using WEPP with GIS and digital elevation models." Journal of Soil and Water Conservation 4: 678-685.
Cochrane, T. A., and D.C. Flanagan (2003). "Representative hillslope models for applying the WEPP model with DEMs and GIS." Transactions of the ASAE 46(4): 1041-1049.
Cochrane, T. A., and D.C. Flanagan (2005). "Effect of DEM resolutions in the runoff and soil loss predictions of the WEPP watershed model." Transactions of the ASAE 48(1): 109-120.
Coe, D. (2006). Sediment production and delivery from forest roads in the Sierra Nevada, CA.  M.S. thesis. Dept. of FRWS, Colorado State University, Fort Collins, CO.
Coe, D. B., and L.H. MacDonald (2005). Road sediment production and delivery in the California Sierra Nevada. European Geophysical Union Meeting, Vienna, Austria.
Coe, J., et al. (2008). "Initiation conditions for debris flows generated by runoff at Chalk Cliffs, central Colorado." Geomorphology 96: 270-297.
We have monitored initiation conditions for six debris flows between May 2004 and July 2006 in a 0.3 km2 drainage basin at Chalk
Cliffs; a band of hydrothermally-altered quartz monzonite in central Colorado. Debris flows were initiated by water runoff from
colluvium and bedrock that entrained sediment from rills and channels with slopes ranging from about 14° to 45°. The availability of
channel material is essentially unlimited because of thick channel fill and refilling following debris flows by rock fall and dry ravel
processes. Rainfall exceeding I=6.61(D)−0.77, where I is rainfall intensity (mm/h), and D is duration (h), was required for the
initiation of debris flows in the drainage basin. The approximate minimum runoff discharge from the surface of bedrock required to
initiate debris flows in the channels was 0.15 m3/s. Colluvium in the basin was unsaturated immediately prior to (antecedent) and
during debris flows. Antecedent, volumetric moisture levels in colluvium at depths of 1 cmand 29 cm ranged from 4–9%, and 4–7%,
respectively. During debris flows, peak moisture levels in colluvium at depths of 1 cm and 29 cm ranged from 10–20%, and 4–12%,
respectively. Channel sediment at a depth of 45 cm was unsaturated before and during debris flows; antecedent moisture ranged from
20–22%, and peak moisture ranged from 24–38%. Although we have no measurements from shallow rill or channel sediment, we
infer that it was unsaturated before debris flows, and saturated by surface-water runoff during debris flows.
Our results allow us to make the following general statements with regard to debris flows generated by runoff in semi-arid to
arid mountainous regions: 1) high antecedent moisture levels in hillslope and channel sediment are not required for the initiation of
debris flows by runoff, 2) locations of entrainment of sediment by successive runoff events can vary within a basin as a function of
variations in the thickness of existing channel fill and the rate of replenishment of channel fill by rock fall and dry ravel processes
following debris flows, and 3) rainfall and simulated surface-water discharge thresholds can be useful in understanding and
predicting debris flows generated by runoff and sediment entrainment.
 
Coe, J. A., et al. (2004). Landslide susceptibility from topography in Guatemala. Landslides: Evaluation and Stabilization. W. Lacerda, M. Ehrlich, S. A. B. Fontoura and A. S. F. Sayao. London, England, Taylor & Francis Group: 69-78.
We present a landslide susceptibility map of a 980 sq. km area in east-central Guatemala that
was impacted by Hurricane Mitch in 1998. Susceptibility was based on a ratio between topographic parameters
at Mitch landslide DEM cells to those of a random sampling of DEM cells within the study area. Ratios
for four topographic parameters (elevation, slope angle, planform curvature, and aspect) were calculated, but
we found that two of these, a combination of elevation (a proxy for rainfall) and slope angle, best portrayed
landslide susceptibility. The map produced from these two parameters includes 80 percent of landslides initiated
by Mitch within susceptible zones, and classifies 71 percent of the study area as having no preferred susceptibility.
Seventy–five percent of pre-Mitch landslides, and 90 percent of post-Mitch landslides, also fall
within susceptible zones. Similar results probably could be expected for other geographic areas where there
is a correlation between elevation and rainfall accumulation.
 
Coffin, B. A. and R. D. Harr (1992). Effects on forest cover on colume of water delivery to soil during rain-on-snow. Seattle, Washington: 106.
Coghlan, M. (1984). A climatologically-based analysis of the storm and flood history of Redwood Creek. Arcata, California, Redwood National Park: 1-47.
Cogle, A. L., et al. (2003). "Testing the hillslope erosion model for application in India, New Zealand and Australia." Environmental Modelling & Software 18(8-9): 825-830.
Cohen, P., et al. (1998). "Towards a regionalization of aquatic habitat: distribution of mesohabitats at the scale of a large basin." Regulated Rivers: Research & Management 14: 391-404.
Cohen, W. B., et al. (in press). "Modeling forest cover attributes as continuous variables in a regional context with Thematic Mapper data." International Journal of Remote Sensing.
Cohen, W. B., et al. (2002). "Characterizing 23 years (1972-1995) of stand replacement disturbance in western Oregon forests with Landsat Imagery." Ecosystems 5: 122-137.
Coho, C. and S. J. Burges (1993). Dam-break floods in low-order mountain channels of the Pacific Northwest. Seattle, WA, Department of Civil Engineering, University of Washington: 68.
Colby, B. R. (1964). Scour and Fill in Sand-Bed Streams, US Geological Survey: 32.
Cole, M. B., et al. (2006). "Change in the upper extent of fish distribution in eastern Washington streams between 2001 and 2002." Transactions of the American Fisheries Society 135: 634-642.
The upper extent of fish distribution in streams has received increasing attention in recent years
because fish-bearing streams in the Pacific Northwest receive greater protection from timber harvest than do
non-fish-bearing waters. Significant amounts of time and funding have been spent surveying streams for the
upper extent of fish distribution and, more recently, in developing models that predict where fish use ends in
streams. The reliability of each approach rests in part on the assumption that the upper extent of fish
distribution can be accurately determined by a single survey and that fish distribution boundaries (FDBs) do
not shift appreciably through time. To examine changes in the upper extent of fish distribution, streams were
surveyed throughout 10 forested watersheds in eastern Washington in the summers of 2001 and 2002. In
2002, 308 FDBs were resurveyed. Resurveys resulted in the establishment of 172 terminal boundaries
(occurring mid-channel or at the confluence of two non-fish-bearing streams) and 136 lateral boundaries
(occurring where a non-fish-bearing channel laterally intersects a fish-bearing channel). The number of
terminal boundary changes was approximately evenly distributed among upstream shifts, no change, and
downstream shifts. Excluding two relatively large downstream shifts, the mean distance between 2002 and
2001 terminal boundaries was  2.5 m, which is not statistically or biologically significant. Terminal
boundaries most often occurred immediately below small impasses created by large woody debris. Of the 136
streams occurring above lateral boundaries in 2001, 134 were again found to support no fish. The upper extent
of fish distribution was similar between these 2 years, perhaps because of similar streamflows during the two
sampling sessions. The number of terminal-to-lateral boundary shifts equaled the estimated number of lateralto-
terminal boundary shifts, indicating that no net change in the use of small tributaries occurred between the
2 years.
 
 
Coleman, S. E. and B. W. Melville (1994). "Bed-Form Development." Journal of Hydraulic Engineering 120(4): 544-560.
Collins, B. D. (1997). Effects of Land Use on the Stillaguamish River, Washington, -1870 to 1990: Implications for Salmonid Habitat and Water Quality and their Restoration. Seattle, Washington, report to the Tulalip Tribes Natural Resources Department, Snohomish County Department of Public Works, Stillaguamish Tribe of Indians, State of Washington Department of Ecology.
Collins, B. D. (1997). Effects of Land Use on the Stillaguamish River, Washington, ~1870 to ~1990: Implications for Salmonid Habitat and Water Quality and their Resotration, The Tulalip Tribes Natural Resources Department.
Collins, B. D., et al. (1994). Watershed assessment and salmonid habitat restoration strategy for Deer Creek, North Cascades of Washington. Seattle, Washington, 10,000 Years Institute.
Collins, B. D., et al. (1994). Watershed assessment and salmonid habitat restoration strategy for Deer Creek, North Cascades of Washington. Seattle, !0,000 Years Institute: 231.
Collins, B. D., et al. (1994). Watershed Assessment and Salmonid Habitat Restoration Strategy for Deer Creek, North Cascades of Washington. Seattle, Stillaguamish Tribe of Indians: 232.
Collins, B. D. and T. Dunne (1989). "Gravel Transport, Gravel Harvesting and Channel-Bed Degradation in Rivers Draining the Southern Olympic Mountains, Washington, U.S.A." Eniviron Geol Water Sci 13(3): 213-224.
Collins, B. D. and T. Dunne (1990). Fluvial Geomorphology and River-Gravel Mining: A Guide for Planners, Case Studies Included. Seattle, Washington, Department of Conservation: Division of Mines and Geology: 31.
Collins, B. D. and D. R. Montgomery (2001). Importance of archival and process studies to characterizing pre-settlement riverine geomorphic processes and habitat in the Puget Lowland. Geomorphic Processes and Riverine Habitat. J. M. Dorava, D. R. Montgomery, B. B. Palcsak and F. A. Fitzpatrick. Washington D.C., American Geophysical Union. Water Science and Application 4: 227-243.
Collins, B. D. and D. R. Montgomery (2002). "Forest development, wood jams and restoration of floodplain rivers in the Puget Lowland." Restoration Ecology 10(2): 237-247.
Collins, B. D., et al. (2002). "Historical changes in the distribution and functions of large wood in Puget Lowland rivers." Canadian Journal of Fisheries and Aquatic Science 59: 66-76.
Abstract: We examined changes in wood abundance and functions in Puget Lowland rivers from the last ~150 years of land use by comparing field data from an 11-km-long protected reach of the Nisqually River with field data from the Snohomish and Stillaguamish rivers and with archival data from several Puget Lowland rivers. Current wood abundance is one to two orders of magnitude less than before European settlement in the Snohomish and Stillaguamish basins. Most importantly, wood jams are now rare because of a lack of very large wood that can function as key pieces and low rates of wood recruitment. These changes in wood abundance and size appear to have fundamentally changed the morphology, dynamics, and habitat abundance and characteristics of lowland rivers across scales from channel unit to valley bottom. Based on our field studies, rivers had substantially more and deeper pools historically. Archival data and field studies indicate that wood jams were integral to creating and maintaining a dynamic, anastomosing river pattern
with numerous floodplain channels and abundant edge habitat and routed floodwaters and sediment onto floodplains. Establishing the condition of the riverine landscape before European settlement sets a reference against which to
evaluate contemporary conditions and develop restoration objectives.
 
Collins, B. D., et al. (2003). Reconstructing the Historical Riverine Landscape of the Puget Lowland. Restoration of Puget Sound Rivers. D. R. Montgomery, S. Bolton, D. B. Booth and L. Wall. Seattle, University of Washington Press: 79-128.
Collins, B. D. and G. R. Pess (1997). "Critique of Washington's Watershed Analysis Program." Journal of the American Water Resources Association 33(5): 997-1010.
Collins, B. D. and G. R. Pess (1997). Critique of Washington's Watershed Analysis Program.
Collins, B. D. and G. R. Pess (1997). "Critique of Washington's Watershed Analysis Programs " Journal of American Water Resources Association 33(5): 997-1010.
Collins, B. D. and G. R. Pess (1997). "Evaluation of Forest Practices Prescriptions from Washington's Watershed Analysis Program." Journal of American Water Resources Association 33(5): 969-996.
Collins, B. D. and G. R. Pess (1997). Evaluation of Forest Practices Prescriptions from Washington's Watershed Analysis Program.
Colman, S. M. (1986). Levels of Time Information in Weathering Measurements, with Examples from Weathering Rinds on Volcanic Clasts in the Western United States. Rates of Chemical Weathering of Rocks and Minerals, Academic Press, Inc.: 379-393.
Colson, T. P. (2006). Stream network delineation from high-resolution digital elevation models. Forestry. Raleigh, North Carolina, North Carolina State University. PhD: 241.
Colson, T. P., et al. (2006). Comparison of stream extraction models using LiDAR DEMs. Geographic Information Systems and Water Resources IV, AWRA Spring Specialty Conference. Houston, Texas, AWRA: 7.
Columbia, P. o. B. (1996). "Terrain stability mapping in British Columbia." from http://www.for.gov.bc.ca/ric/PUBS/EARTHSCI/Terrain2/.
Combs, P. G., et al. (1977). "Application of Flow-Sediment Model to Red River." Journal of the Hydraulics Division 103(1): 11-18.
Conference Coordinating Committee (2000). Abstracts. International Conference on Wood in world Rivers, Corvallis, Oregon.
Conference Coordinating Committee (2000). Conference Program. Corvallis, Oregon, International Conference on Wood in World Rivers: 39.
Connelly, B. A. and T. W. Cundy (1992). Cumulative Effects of Forest Management on Peak Streamflows During Rain-on-Snow Events. Interdisciplinary Approaches in Hydrology and Hydrogeology. Seattle, Washinton, University of Washington: 470-484.
Connelly, J. H. (1978). "Diversity in Tropical Rain Forests and Coral Reefs." Science 199(24): 1302-1310.
Conrad, R. H., et al. (2003). The development and assessment of the preliminary model for identifying fish habitat in western Washington. Olympia, WA, Cooperative Monitoring Evaluation & Research: 150.
Conroy, W. J. and J. C. Barrett (2002). Controlling sediment erosion from forest roads in coastal watersheds. American Water Resources Association Technical Publication Series TPS, vol.02-1. J. R. Lesnik: 237-242.
The Pacific Lumber Company (PALCO) owns 225,000 acres of industrial timberlands in coastal Northern California watersheds that drain to tidal estuaries, or salt-water bays. During the development of its Habitat Conservation Plan (HCP), PALCO recognized that roads were an important management-related sediment source, a conclusion supported by subsequent watershed studies. Consequently, the HCP includes, and PALCO has instituted, a state-of-the-art road management plan for the 1500 miles of permanent and seasonal roads on its ownership. The road management plan includes numerous measures for controlling sediment inputs to watercourses: removal or repair of "at-risk" stream crossings, removal of unneeded roads, "storm-proofing" of all permanent roads, diversion of road drainage away from watercourses, reduction or elimination of wet-weather road use and construction, and the upgrading of all seasonal roads. Monitoring and assessment studies have been used to document the effectiveness of these approaches in limiting road-related sediment delivery to streams.
 
Cooke, R. U. D., J. C. (1990). Geomorphology In Environmental Management, Oxford Press.
Coombs, H. A. (1989). "The Baker Project." Engineering Geology in Washington 1: 174-186.
Cooney, T. and D. M. Holzer (2007). Interior Columbia Basin viability Criteria Review Draft 2007. Appendix C: habitat intrinsic potential analysis: 21.
Cooper, R. M. (2005). Estimation of peak discharges for rural, unregulated streams in western Oregon. Scientific Investigations Report. Reston, Virginia, U.S. Geological Survey: 134.
Cooperative Instream Flow Service Group (1982). A Guide to Stream Habitat Analysis Using the Instream Flow Incremental Methodology, Fish and Wildlife Service, Environmental Protection Agency, Soil Conservation Service, Geological Survey: 163-169.
Corripio, J. G. (2003). "Vectorial algebra algorithms for calculating terrain parameters from DEMs and solar radiation modelling in mountainout terrain." Int. J. Geographical Information Science 17(1): 1-23.
Terrain parameters derived from digital elevation models (DEMs),
such as slope gradient, aspect and cell surface area, are represented as a vector
normal to the surface and calculated using the minimum areal unit of the DEM,
that is enclosed between four data points. The position of the Sun is calculated
by applying rotational matrices to a unit vector defined at noon as a function of
latitude and declination. The direct component of insolation intercepted by the
cell surface is then calculated as a dot product between the unit vector in the
direction of the Sun and the unit vector normal to surface, multiplied by direct
normal irradiation. Hillshading is computed by scanning the projection of cells
onto a solar illumination plane perpendicular to the Sun direction. Horizon angles
and estimated sky view factor are calculated using a more economical algorithm
than a rigorous evaluation of all the angles subtended by every grid cell to each
other. The performance of the slope algorithm is evaluated using a synthetic
surface and real world examples are given for the Mont Blanc Massif, in the
French Alps.
 
 
Costa, J. E. (1983). "Paleohydraulic reconstruction of flash-flood peaks from boulder deposits in the Colorado Front Range." Geological Society of America Bulletin 94: 986-1004.
Costa, J. E. (1983). The Physical Geomorphology of Debris Flows (DRAFT), University of Denver.
Costa, J. E. (1985). Floods from Dam Failures, U.S. Geological Survey: 1-54.
Costa, J. E. (1998). Flood Geomorphology. Rheologic, geomorphic, and sedimentologic differentiation of water floods, hyperconcentrated flows, and debris flows. V. R. Baker, R. C. Kochel and B. C. Patton, Wiley: 113-122.
Costa, J. E. and R. D. Jarrett (1981). "Debris Flows in Small Mountain Stream Channels of Colorado and their Hydrologic Implications." Bulletin of the Association of Engineering Geologists 18(3): 309-322.
Costa, J. E. and R. L. Schuster (1988). "The formation and failure of natural dams." Geological Society of America Bulletin 100: 1054-1068.
Costa-Cabral, M. C., and S.J. Burges (1994). "Digital elevation model networks (DEMON): a model of flow over hillslopes for computation of contributing and dispersal areas." Water Resources Research 30(6): 1681-1692.
Costa-Cabral, M. C. and S. J. Burges (1994). "Digital elevation model networks (DEMON): A model of flow over hillslopes for computation of contributing and dispersal areas." Water Resources Research 30(6): 1681-1692.
Costantini, A., et al. "Sediment generation from forest roads: bed and eroded sediment size distributions, and runoff management strategies."
A rainfall simulator and overland flow study was conducted to determine in situ and eroded sediment size distributions for a range of forest road surfaces at 2 important commercial plantation centres in subtropical south-east Queensland, Australia; and parameters necessary for running the CREAMS model to assess erosion and sediment transport from road/table drain systems. Results revealed very low concentrations of fine particles in the surface of gravel roads, and somewhat higher proportions in ungravelled (dirt) road surfaces. However, there was considerable enrichment of fine particles in sediment eroded under simulated rain, with concentrations of particles <0.02 mm in diameter being up to 8 g/L. Table drains were generally resistant to scour by overland flows, with the only exception being a drain bordering a newly gravelled road. This drain was bare of vegetation and contained significant quantities of loose gravel from which the fine component was easily eroded. It demonstrated the need to construct both roads and table drains at the end of the wet season when consolidation and re-vegetation can occur under lighter rains during the dry season. CREAMS model runs for a 'standard' road and drain configuration predicted considerable enrichment of fine particles in sediment from all road surfaces. The major factor controlling predicted concentrations of fine particles was the rate of erosion from the road surface, with gravelled surfaces showing considerably less erosion than ungravelled surfaces. Because road surfaces will be significant sources of fine sediment during erosive rains, a second part of this study was designed to model whether hillslopes could be used to infiltrate runoff, thereby controlling sediment movement. For the modelled hillslopes - typical of those used to support commercial forest plantations in south-east Queensland - design runoffs from forest road turn-out drains could be infiltrated. It is suggested that forest managers use hillslope infiltration as the primary tool for managing flows and sediments from road turn-out drains, and that vegetative filter strips be used only as a secondary support tool.
 
Coulombe-Pontbriand, M. and M. F. Lapointe (2004). "Landscape controls on boulder-rich, winter habitat availability and their effects on Atlantic salmon (Salmo salar) parr abundance in two fifth-order mountain streams." Canadian Journal of Fisheries and Aquatic Science 61: 648-658.
We test the effect at river reach and segment scales of landscape controls on the distribution of Atlantic
salmon (Salmo salar) parr densities, as well as associated variations in boulder (diameter ≥ 256 mm) abundance and
potential overwintering habitat. This study encompasses data from 45 km of fifth-order mainstem channels along two
neighbouring river catchments in the Gaspé region, Québec. At both scales, winter habitat availability was correlated
with boulder availability. At the river segment scale (1–5 km), parr densities significantly correlated (P < 0.05) with
boulder availability along the Bonaventure River, which presented significant intersegment variations in boulder abundances.
In contrast, segment-scale boulder and parr abundances were uniformly low along the Petite Cascapédia River.
At the reach scale (600 m), positive but less strongly significant boulder – parr abundance correlations were observed
in both the Bonaventure and Petite Cascapédia rivers. Spatial variations in boulder abundances in these systems reflected
variations in the degree of channel to valley walls coupling and imposed channel formative shear stresses. In
similarly boulder-poor segments with comparable fry abundances, parr abundances were significantly greater along the
Bonaventure than the Petite Cascapédia River, possibly because of the presence in the former system of nearby
boulder-rich refugia segments.
 
Cox, N. J. (1980). "On the Relationship Between Bedrock Lowering and Regolith Thickness." Earth Surface Process and Landforms 5: 271-274.
Cox, N. J. (1989). "Review of "Biogeomorphology"." Progress of Physical Geography 13: 620-624.
Cox, N. J. (1992). Precipitation statistics for geomorphologists: Variations on a theme by Frank Ahnert. Functional Geomorphology. K. M. Schmidt and J. d. Ploey. Cremlingen-Destedt, Catena Verlag. Catena Supplement 23: 189-212.
Craft, C. B., et al. (1993). "Vertical Accretion in Microtidal Regularly and Irregularly Flooded Estuarine Marshes " Estuarine Coastal and Shelf Science 37: 371-386.
Crandell, D. R. (?). "The Glacial History of Western Washington and Oregon "?: 341-353.
Creque, S. M., et al. (2005). "Use of GIS-derived landscape-scale habitat features to explain spatial patterns of fish density in Michigan rivers." North American Journal of Fisheries Management 25: 1411-1425.
Both site- and landscape-scale processes play important roles in the biological communities
of rivers. Understanding the influences of these processes on fish abundance can help
direct management and research efforts toward appropriate habitat variables and scales. We used
multiple linear regression analysis of a regional fish and habitat database to determine the feasibility
of using geographical information systems (GIS)–derived landscape-scale habitat variables to
explain the spatial variation in the density of five sport fish species (Chinook salmon Oncorhynchus
tshawytscha, steelhead O. mykiss, brown trout Salmo trutta, brook trout Salvelinus fontinalis, and
white sucker Catostomus commersonii) in the rivers of Michigan’s Lower Peninsula. We compared
these models with those developed using site-scale variables traditionally measured in the field.
Landscape-scale riverine habitat variables obtained through GIS analysis and modeling of catchment
characteristics accounted for 18–69% of the variation in fish density. Landscape estimates
of mean July water temperature were negatively correlated with the density of brook trout, brown
trout, and Chinook salmon. Drainage area was negatively correlated with the density of steelhead
and white suckers, and 90% exceedence flow yield (a measure of flow stability) was positively
correlated with the density of Chinook salmon and steelhead. Site-scale habitat variables explained
less (12–57%) of the variation in fish density than landscape-scale variables. In the site-scale
models, depth was negatively related to all species’ densities, and the percentage of soft substrates
was positively correlated only with white suckers. Although there was still much unexplained
variation in density, our models provide insight into key habitat variables that influence fish density
patterns on a large scale.
 
Crisp, D. T. (1993). "The ability of U.K. salmonid alevins to emerge through a sand layer " Journal of Fish Biology 43: 656-658.
Crisp, D. T. and P. A. Carling (1989). "Observations on siting, dimensions and structure of salmonid redds " J. Fish Biol. 34: 119-134.
Critchfield, W. B. (1985). "The late Quaternary history of lodgepole and jack pines." Review unknown: 749-772.
Croke, J., P. Hairsine,  and P. Fogarty (1999). "Runoff generation and redistribution in logged eucalytus forests, southeastern Australia." Journal of Hydrology 216: 55-77.
Croke, J., P. Hairsine,  and P. Fogarty (1999). "Sediment transport, redistribution and storage on logged forest hillslopes in south-eastern Australia." Hydrological Processes 13(17): 2705-2720.
Croke, J., and S. Mockler (2001). "Gully initiation and road-to-stream linkage in a forested catchment, southeastern Australia." Earth Surface Processes and Landforms 26: 205-217.
This study reports the nature of sediment delivery pathways and road-to-stream linkage in a forested catchment in southeastern Australia, and evaluates the causal factors associated with this linkage. Detailed field surveying of approximately 20 per cent of the 75 km road network reveals that 18 per cent of road drains show complete channel linkage via gully development from a road outlet to a stream. An additional 11 per cent of road drains show evidence of partial channel linkage where the gully does not extend the full hillslope length. Inclusion of the full range of road-to-stream linkage categories, including direct linkage at stream crossings, road bridges and fords, results in a 6 per cent increase in drainage density since initial road construction in 1964. The majority of this linkage is associated with relief culverts draining cut- and-fill roads in mid-valley positions. These drainage structures have contributing road lengths that are on average three times longer than those draining mitre drains on ridgetop roads. Runoff from these roads is also discharged onto hillslopes that are at least twice as steep as those used on ridgetop roads. Contributing road length (m) and the gradient of the discharge hillslope (tan theta ) are successfully used here in a linear discriminant analysis to separate channelled and non-channelled flow pathways within the catchment. The successful delineation of these pathways using two easily measured variables suggests that this approach has potential in the planning and rehabilitation of forest roads.
 
Croke, J., P. Hairsine, and P. Fogarty (2001). "Soil recovery from track construction and harvesting changes in surface infitration, erosion and delivery rates with time." Forest Ecology and Management 143: 3-12.
Croke, J., S. Mockler, P. Fogarty, and I. Takken (2005). "Sediment concentration changes in runoff pathways from a forest road network and the resultant spatial pattern of catchment connectivity." Geomorphology 68: 257-268.
Croke, J., and M. Nethery (2006). "Modelling runoff and soil erosion in logged forests: scope and application of some existing models." Catena 67: 35-49.
Croke, J., and P.B. Hairsine (2006). "Sediment delivery in managed forests: a review." Environmental Reviews 14: 59-87.
Croley, T. E. (2004). Spatially distributed model of interacting surface and groundwater storages. World Water and Environmental Resources Congress. June, 2004. Salt Lake City, UT Environmental Water Resources Institute, ASCE, Washington, DC.
Crookston, N. L. and G. E. Dixon (2005). "The forest vegetation simulator: a review of its structure, content, and applications." Computers and electronics in agriculture 49: 60-80.
The Forest Vegetation Simulator (FVS) is a distance-independent, individual-tree forest growth
model widely used in the United States to support management decisionmaking. Stands are the
basic projection unit, but the spatial scope can be many thousands of stands. The temporal scope
is several hundred years at a resolution of 5–10 years. Projections start with a summary of current
conditions evident in the input inventory data. FVS contains a self-calibration feature that uses
measured growth rates to modify predictions for local conditions. Component models predict the
growth and mortality of individual trees, and extensions to the base model represent disturbance
agents including insects, pathogens, and fire. The component models differ depending on the
geographic region represented by regionally specific model variants. The differences are due to
data availability and the applicability of existing models. The model supports specification of
management rules in the input, such as thinning if density is too high. The rules can be extended
to represent other factors. For example, the effect of climate change on stand development
by entering rules that specify how growth and mortality will change in response to changing
climate.
Applications range from development of silvicultural prescription for single stands to landscape
and large regional assessments. Key issues addressed with FVS include forest development, wildlife
habitat, pest outbreaks, and fuels management. The predictions are used to gain insights into how
forested environments will respond to alternative management actions. Broad-scale forest management
policies have been studied with FVS.
For the 30 years since the model was initially introduced, the development team has anticipated and
provided needed enhancements and maintained a commitment toworking with and training users. The
existence of an adequate user interface and the continued use of the original programming language
are often overlooked factors for the success of this model.
Future work will focus on improving FVS by adopting recent biometric techniques and including
newinformation linking geomorphology to mortality and growth. Extending the model to more closely
represent biophysical processes and adapting the model so that it is more relevant to management
questions related to predicted climate change are also foci. Providing ways to dynamically link FVS
to other models is our current strategy for providing major new capabilities.
 
Crosta, G. B. and P. Frattini (2003). "distributed modelling of shallow landslides triggered by intense rainfall." Natural Hazards and Earth System Sciences 3: 81-93.
Hazard assessment of shallow landslides represents
an important aspect of land management in mountainous
areas. Among all the methods proposed in the literature,
physically based methods are the only ones that explicitly includes
the dynamic factors that control landslide triggering
(rainfall pattern, land-use). For this reason, they allow forecasting
both the temporal and the spatial distribution of shallow
landslides. Physically based methods for shallow landslides
are based on the coupling of the infinite slope stability
analysis with hydrological models. Three different gridbased
distributed hydrological models are presented in this
paper: a steady state model, a transient “piston-flow” wetting
front model, and a transient diffusive model. A comparative
test of these models was performed to simulate landslide occurred
during a rainfall event (27–28 June 1997) that triggered
hundreds of shallow landslides within Lecco province
(central Southern Alps, Italy). In order to test the potential
for a completely distributed model for rainfall-triggered landslides,
radar detected rainfall intensity has been used. A new
procedure for quantitative evaluation of distributed model
performance is presented and used in this paper. The diffusive
model results in the best model for the simulation of
shallow landslide triggering after a rainfall event like the one
that we have analysed. Finally, radar data available for the
June 1997 event permitted greatly improving the simulation.
In particular, radar data allowed to explain the non-uniform
distribution of landslides within the study area.
 
Crouch, R. J. and T. Novruzi (1989). "Threshold conditions for rill initiation on a vertisol, Gunnedah, NSW, Australia." Catena 16(1): 101-110.
Crowe, W. P. a. J. (2003). "Surface-based transport model for mixed-size sediment." Journal of Hydraulic Engineering-ASCE 129(2): 120-128.
Crozier, M. J. (1999). "Prediction of rainfall-triggered landslides: A test of the antecedent water status model." Earth Surface Processes and Landforms 24: 825-833.
Crozier, M. J., et al. (1990). "Relative instability of colluvium-filled bedrock depressions." Earth Surface Processes and Landforms 15: 329 - 339.
Cruden, D. M., et al. (1993). "The landslide dam on the Saddle River near Rycroft, Alberta." Can. Geotech. J. 30: 1003-1015.
Cruden, D. M., et al. (1997). "The 1939 Montagneuse River landslide, Alberta." Can. Geotech. J. 34: 799-810.
Cruden, D. M. and S. Thomson (1997). "The 1939 Montagneuse River landslide, Alberta." Canadian Geotechnical Journal 34: 799-810.
Cruden, D. M. and D. J. Varnes (1996). Landslide types and processes. Landslides Investigation and Mitigation. A. K. Turner and R. L. Schuster. Washington, D.C., National Academy Press: 36-75.
Cui, Y., et al. (2003). "Sediment pulses in mountain rivers: 1. Experiments." Water Resources Research 39(9): 1239.
Cui, Y., et al. (2003). "Sediment pulses in mountain rivers: 2. Comparison between experiments and numerical predictions." Water Resources Research 39(9).
Mountain rivers in particular are prone to sediment input in the form of pulses
rather than a more continuous supply. These pulses often enter in the form of landslides
from adjacent hillslopes or debris flows from steeper tributaries. The activities of
humans such as timber harvesting, road building, and urban development can increase the
frequency of sediment pulses. The question as to how mountain rivers accommodate
pulses of sediment thus becomes of practical as well as academic significance. In part 1
[Cui et al., 2003], the results of three laboratory experiments on sediment pulses are
reported. It was found there that the pulses were eliminated from the flume
predominantly by dispersion of the topographic high. Significant translation was
observed only when the pulse material was substantially finer than the ambient load in
the river. Here the laboratory data are used to test a numerical model originally devised
for predicting the evolution of sediment pulses in field-scale gravel bed streams. The
model successfully reproduces the predominantly dispersive deformation of the
experimental pulses. Rates of dispersion are generally underestimated, largely because
bed load transport rates are underestimated by the transport equation used in the
model. The model reproduces the experimental data best when the pulse is significantly
coarser than the ambient sediment. In this case, the model successfully predicts the
formation and downstream progradation of a delta that formed in the backwater zone of
the pulse in run 3. The performance of the model is less successful when the pulse is
composed primarily of sand. This is likely because the bed load equation used in the
study is specifically designed for gravel. When the model is adapted to conditions
characteristic of large, sand bed rivers with low Froude numbers, it predicts substantial
translation of pulses as well as dispersion.
 
Culling, W. E. H. (1986). "Equifinality: modern approaches to dynamical systems and their potential for geographical thought." Trans. Instr. Br. Geogr. N.S. 12: 57-72.
Cummans, J. E., et al. (1975). Magnitude and Frequency of Floods in Washington. Tacoma, Washington, US Geological Survey: 14-20.
Cupp, C. E. (2005). Water typing model field performance assessment pilot study. Olympia, Cooperative Monitoring Evaluation & Research: 21.
Curry, R. (?). "Coupling Marine and Terrestrial Wateshed Processes ". Retrieved 10/12/98, 1998, from http://color.mlml.calstate.edu/www/mbnms/docs92/sympcurr.htm.
Curry, R. R. (1966). "Observation of Alpine Mudflows in the Tenmile Range, Central Colorado." Geological Society of America Bulletin 77: 771-776.
Czarnomski, N. M., et al. (2008). "Dynamics of wood in stream networks of the western Cascades Range, Oregon." Canadian Journal of Forest Research 38: 2236-2248.
We develop and test a conceptual model of wood dynamics in stream networks that
considers legacies of forest management practices, floods, and debris flows. We
combine an observational study of wood in 25 km of 2nd- through 5th-order
streams in a steep, forested watershed of the western Cascade Range of Oregon
with whole-network studies of forest cutting, roads, and geomorphic processes
over the preceding 50 years. Statistical and simple mass balance analyses show
that natural process and forest management effects on wood input, transport
processes, and decomposition account for observed patterns of wood in the stream
network. Forest practices reduced wood amounts throughout the network; in
headwater streams these effects are fixed in stream segments bordered by cuts
and roads, but in larger channels they are diffused along the channel by fluvial
transport of wood. Landforms and roads limited delivery of wood by debris flows
to mainstern channels. Network dynamics studies and watershed management plans
should include spatial patterns of debris flow initiation and runout, flood
redistribution, and reduction of wood in the network by forest cutting and
intentional wood removal from channels on time scales of forest succession and
recurrence of major floods.
 
Dade, W. B. (2000). "Grain size, sediment transport and alluvial channel pattern " Geomorphology 35: 119-126.
Dade, W. B. and P. F. Friend (1998). "Grain-Size Sediment-Transport Regime, and Channel Slope in Alluvial Rivers." The Journal of Geology 106: 661-675.
Dahlström, N., et al. (2005). "Long-term dynamics of large woody debris in a managed boreal forest stream." Forest Ecology and Management 210: 363-373.
Little is known about how past forest management in Sweden influenced the quantity and quality of large woody debris
(LWD) in streams. The present study provides information of the long-term dynamics of LWD in a reach of a boreal stream
intersecting a managed forest. Dendrochronological methods were used to reconstruct mortality years of the pieces of LWD and
the general history of fire and cuttings of the surrounding riparian forest. Today, spruce dominates among the living trees,
whereas the LWD is dominated by birch in the forest and by pine in the stream. Fire frequency prior to active fire suppression was
similar to values reported from boreal forests. Pine trees were more abundant in the riparian forest before selective logging
operations and active fire suppression began in the 1800s. Many of the pieces of LWD found in the stream today died more than
200 years ago and derived from a cohort of pines that generated in the early 1600s. Pine LWD in stream channels is highly
resistant to decomposition and can reside for more than 300 years. A substantial amount of the LWD found today in managed
forest streams in boreal Sweden most likely derives from the time before extensive human influence and is likely to decrease
further in the future. Management of riparian forests to ascertain future supply of long-lived LWD in streams should target to
increase the proportion of pine trees.
 
Dai, F. C. and C. F. Lee (2001). "Terrain-based mapping of landslide susceptibility using a geographical information system: a case study." Canadian Geotechnical Journal 38: 911-923.
This paper deals with the development of a technique for mapping landslide susceptibility using a geographical information system (GIS), with particular reference to landslides on natural terrain. The method has been applied to Lantau Island, the largest outlying island within the territory of Hong Kong. Landslide susceptibility in the study area is related to a number of terrain variables, viz., lithology, slope gradient, slope aspect, elevation, land cover, and distance to drainage line. Multiple correspondence analysis (MCA) was carried out to generate the principal axes that are linear combinations of these terrain variables using occurrence data of landslides and terrain variables. A GIS is used to project the values of the principal axes, and subsequently to relate these principal axes to landslide susceptibility by logistic regression modeling. The spatial landslide susceptibility response in the study area can then be obtained by applying this logistic regression model to the study area. The results from this study indicate that such a GIS-based model is useful and suitable for the scale adopted in this study.
 
Daily, G. C. and P. R. Ehrlich (1999). "Managing Earth's Ecosystems: An Interdisciplinary Challenge." Ecosystems 2: 277-280.
Daly, C., R.P.Neilson, and D.L. Phillips (1994). "A statistical-topographic model for mapping climatological precipitation over mountainous terrain." Journal of Applied Meteorology 33: 140-160.
Daly, C., et al. (2008). "Physiographically sensitive mapping of climatological temperature and precipitation acroos the conterminous United States." International Journal of Climatology.
Spatial climate data sets of 1971–2000 mean monthly precipitation and minimum and maximum temperature
were developed for the conterminous United States. These 30-arcsec (∼800-m) grids are the official spatial climate data
sets of the U.S. Department of Agriculture. The PRISM (Parameter-elevation Relationships on Independent Slopes Model)
interpolation method was used to develop data sets that reflected, as closely as possible, the current state of knowledge of
spatial climate patterns in the United States. PRISM calculates a climate–elevation regression for each digital elevation
model (DEM) grid cell, and stations entering the regression are assigned weights based primarily on the physiographic
similarity of the station to the grid cell. Factors considered are location, elevation, coastal proximity, topographic facet
orientation, vertical atmospheric layer, topographic position, and orographic effectiveness of the terrain. Surface stations
used in the analysis numbered nearly 13 000 for precipitation and 10 000 for temperature. Station data were spatially quality
controlled, and short-period-of-record averages adjusted to better reflect the 1971–2000 period.
PRISM interpolation uncertainties were estimated with cross-validation (C-V) mean absolute error (MAE) and the 70%
prediction interval of the climate–elevation regression function. The two measures were not well correlated at the point
level, but were similar when averaged over large regions. The PRISM data set was compared with the WorldClim and
Daymet spatial climate data sets. The comparison demonstrated that using a relatively dense station data set and the
physiographically sensitive PRISM interpolation process resulted in substantially improved climate grids over those of
WorldClim and Daymet. The improvement varied, however, depending on the complexity of the region. Mountainous and
coastal areas of the western United States, characterized by sparse data coverage, large elevation gradients, rain shadows,
inversions, cold air drainage, and coastal effects, showed the greatest improvement. The PRISM data set benefited from a
peer review procedure that incorporated local knowledge and data into the development process. Copyright  2008 Royal
Meteorological Society
 
 
Daly, C., et al. (1994). "A statistical-topographic model for mapping climatological precipitation over mountainous terrain." Journal of Applied Meteorology 33: 140-158.
Das, B. M. (2001). Slope Stability. Principles of Geotechnical Engineering. Boston, PWS Engineering: 429-487.
Davey, C. and M. Lapointe (2007). "Sedimentary links and the spatial organization of Atlantic salmon (Salmo salar) spawning habitat in a Canadian Shield river." Geomorphology 83: 82-96.
The segmenting of gravel-bed rivers flowing through mountain valleys into a number of discrete ‘sedimentary links’, each
characterized by downstream fining of alluvium, is a relatively recent concept which offers promise to model the large-scale spatial
organisation of many types of aquatic habitat (reproductive, feeding, refuge, etc), strongly dependent on dominant bed sediment
calibre. Although, so far, the ecological application of the concept has mainly focused on benthic invertebrates, here we illustrate its
application to fish (Atlantic salmon; Salmo salar). Moreover, the link concept has also been primarily applied to alpine river
environments where link formation is triggered by point sources (mainly tributaries) supplying coarser sediment. However,
somewhat lower relief, mountain valley landscapes of North Eastern Canada are often structured into sedimentary links triggered
by non-point, ‘supply zones’ of coarse sediments, originating in bedrock canyon reaches or valley bottom deposits of glacial drift.
Here, we propose an adaptation and extension of the original, sedimentary link concept to such landscapes and test its utility along
one such system, the Ste Marguerite River (SMR), a salmon river draining the Canadian Shield in the Saguenay region of Québec.
We first discuss a simple field and office based method of link delineation. Then we discuss potential sources of minor, sublink
scale grain size variability and their effects on how sedimentary links are defined. Lastly, we demonstrate the usefulness of the link
structure to model the distribution of Atlantic salmon spawning habitat (a habitat that depends critically on bed texture). Our results
indicate that a revised sedimentary link typology is needed to describe longitudinal grain size patterns where non-point, valleysegment
scale sources of coarse sediment are important and that consideration of the research purpose and scale is important in
defining meaningful link units. We also show that salmon spawning zones can be directly predicted from the link structure: along
the SMR, spawning activity is apportioned within each of the discrete links, in those sub-zones where surface sediment size and
sand content are optimal for reproduction.
 
Davies, J. R., et al. (2007). "Modeling stream channel characteristics from drainage-enforced DEMs in Puget Sound, Washington, USA." Journal of the American Water Resources Association 43(2): 414-426.
Mapping stream channels and their geomorphic attributes is an important step in many watershed
research and management projects. Often insufficient field data exist to map hydromorphologic attributes across
entire drainage basins, necessitating the application of hydrologic modeling tools to digital elevation models
(DEMs) via a geographic information system (GIS). In this article, we demonstrate methods for deriving synthetic
stream networks via GIS across large and diverse basins using drainage-enforced DEMs, along with techniques
for estimating channel widths and gradient on the reach scale. The two-step drainage enforcement
method we used produced synthetic stream networks that displayed a high degree of positional accuracy relative
to the input streams. The accuracies of our estimated channel parameters were assessed with field data, and
predictions of bankfull width, wetted width and gradient were strongly correlated with measured values
(r2 = 0.92, r2 = 0.95, r2 = 0.88, respectively). Classification accuracies of binned channel attributes were also
high. Our methodology allows for the relatively rapid mapping of stream channels and associated morphological
attributes across large geographic areas. Although initially developed to provide salmon recovery planners with
important salmon habitat information, we suggest these methodologies are relevant to a variety of research and
management questions.
 
Davies, T. R. H. (1987). Problems of bed load transport in braided gravel-bed rivers. Sediment Transport in Gravel-bed (this the right source, or not full?). C. R. Thorne, J. C. Bathurst and R. D. Hey, John Wiley & Son Ltd.: 793-827.
Davies-Colley, R. J., et al. (2009). "Modelling the time course of shade, temperature, and wood recovery in streams with riparian forest restoration." New Zealand Journal of Marine and Freshwater Research 43: 673-688.
Action is increasingly being taken in
New Zealand and elsewhere to restore ecological
function to streams through planting of riparian
zones. We used simulation modelling to explore the
relative performance of three strategies to restore
the riparian zone of a pastoral stream to native
forest by: (1) passive regeneration; (2) planting
then abandonment of a Pinus radiata plantation;
and (3) active restoration by planting selected native
trees. We linked the forest model LI NKNZ with a
shade and temperature model (sWAIO RA), and a
wood model (OSU_STREA MWOO D) to simulate
recovery trajectories for key forest stream attributes
in hypothetical streams (1.3–14.0 m channel width)
in the central North Island, New Zealand. Both
active restoration strategies outperformed passive
regeneration in shade, temperature and stream wood
volume for most of the simulation time (800 years).
Although the abandoned pine plantation provided
greatest shade initially (<100 years), active native
planting provided the greatest benefits overall. In
general, recovery of stream shade (and temperature) is
expected within decades, is accelerated by deliberate
planting, and is fastest in small streams in which
thermal stress from sunlight exposure is greatest.
However, full recovery of stream and riparian function
may take centuries, being dependent on large trees
providing wood to structure the channel.
 
Davison, A. C. and D. V. Hinkley (1997). Bootstrap Methods and Their Application. Cambridge, Cambridge University Press.
Dawson, C. W., M.R.Brown, and R.L.Wilby (2000). "Inductive learning approaches to rainfall-runoff modelling." International Journal of Neural Systems 10(1): 43-57.
Dawson, C. W. a. R. L. W. (2001). "Hydrological modelling using artificial neural networks." Progress  in Physical Geography 25(1): 80-108.
de Boer, D. H. (1992). "Hierarchies and spatial scale in process geomorphology: a review." Geomorphology 4: 303-318.
De Groot, J. D., et al. (2007). "Effects of logging second-growth forests on headwater populations of coastal cutthroat trout: a 6-year, multistream, before-and-after field experiment." Transactions of the American Fisheries Society 136: 211-226.
To understand how logging of second-growth forests affects populations of coastal cutthroat
trout Oncorhynchus clarkii clarkii, we examined trout relative abundance, body condition (mass relative to
length), and physical and thermal habitat in the summer and winter in four headwater streams (two treatment
streams and two nonlogged control streams) over a 6-year period (2 years prelogging [1997–1998] and 4
years postlogging [1999–2002]). This is one of the first efforts to conduct a multiyear, replicated stream,
before-and-after experiment on this scale to assess the effects of logging on fish and habitat. In the treatment
streams, 21% of the watershed area was logged by clear-cutting (no scarification or slash-burning). Careful
logging approaches were employed to remove most of the riparian overstory (i.e., no machines were used
within 5 m of stream, logs were felled and yarded away from riparian zones, all shrubs were left behind, and
large wood was left in streams). Because a cooler summer climate occurred coincidentally with our
postlogging period (the mean daily average summer air temperature was 1–28C cooler than the temperature
during the prelogging period), the mean average and mean maximum daily stream temperatures declined after
the logging period in the control streams and remained the same in the treatment streams. After accounting for
the effects of climate, logging had warmed treatment streams by about 18C. We could not detect any logging
treatment effects on summer or winter relative abundance or condition, nor were any changes evident to
instream physical habitat associated with the logging treatment. These results were probably attributable to the
careful logging approaches employed and the cooler climate that occurred during the postlogging period.
 
de la Rosa, D., F. Mayol, J.A.Moreno, T. Bonson, and S. Lozano. (1999). "An expert system/neural network model (ImpelERO) for evaluating agricultural soil erosion in Andalucia region, Southern Spain." Agriculture Ecosystems & Environment 73(1999): 211-226.
de Lugt, J. and I. A. Campbell (1992). Mass movements in the badlands of Dinosaur Provincial Park, Alberta, Canada. Functional Geomorphology. K. H. Schmidt and J. de Ploey. Cremlingen, Catena Verlag. Catena Supplement 23: 75-100.
De Roo, A. P. J. (2000). Modelling runoff and sediment transport in catchments using GIS. In Hydrological Applications of GIS. A. M. G. a. D. R. M. (editors). Wiley, England 176p.
De'ath, G., and K.E. Fabricius (2000). "Classification and regression trees: a powerful yet simple technique for ecological data analysis." Ecology 81(11): 3178-3198.
De'ath, G. (2002). "Multivariate regression trees: a new technique for modeling species-environment relationships." Ecology 83(4): 1105-1117.
Multivariate regression trees (MRT) are a new statistical technique that can
be used to explore, describe, and predict relationships between multispecies data and environmental
characteristics. MRT forms clusters of sites by repeated splitting of the data,
with each split defined by a simple rule based on environmental values. The splits are
chosen to minimize the dissimilarity of sites within clusters. The measure of species dissimilarity
can be selected by the user, and hence MRT can be used to relate any aspect of
species composition to environmental data. The clusters and their dependence on the environmental
data are represented graphically by a tree. Each cluster also represents a species
assemblage, and its environmental values define its associated habitat. MRT can be used
to analyze complex ecological data that may include imbalance, missing values, nonlinear
relationships between variables, and high-order interactions. They can also predict species
composition at sites for which only environmental data are available. MRT is compared
with redundancy analysis and canonical correspondence analysis using simulated data and
a field data set.
 
De'ath, G. (2007). "Boosted trees for ecological modeling and prediction." Ecology 88(1): 243-251.
Accurate prediction and explanation are fundamental objectives of statistical analysis, yet they seldom coincide. Boosted trees are a statistical learning method that attains both of these objectives for regression and classification analyses. They can deal with many types of response variables (numeric, categorical, and censored), loss functions (Gaussian, binomial, Poisson, and robust), and predictors (numeric, categorical). Interactions between predictors can also be quantified and visualized. The theory underpinning boosted trees is presented, together with interpretive techniques. A new form of boosted trees, namely, “aggregated boosted trees” (ABT), is proposed and, in a simulation study, is shown to reduce prediction error relative to boosted trees. A regression data set is analyzed using ABT to illustrate the technique and to compare it with other methods, including boosted trees, bagged trees, random forests, and generalized additive models. A software package for ABT analysis using the R software environment is included in the Appendices together with worked examples.
 
De'ath, G. and K. E. Fabricius (2000). "Classification and regression trees: a powerful yet simple technique for ecological data analysis." Ecology 81(11): 3178-3192.
Classification and regression trees are ideally suited for the analysis of complex
ecological data. For such data, we require flexible and robust analytical methods,
which can deal with nonlinear relationships, high-order interactions, and missing values.
Despite such difficulties, the methods should be simple to understand and give easily
interpretable results. Trees explain variation of a single response variable by repeatedly
splitting the data into more homogeneous groups, using combinations of explanatory variables
that may be categorical and/or numeric. Each group is characterized by a typical
value of the response variable, the number of observations in the group, and the values of
the explanatory variables that define it. The tree is represented graphically, and this aids
exploration and understanding.
Trees can be used for interactive exploration and for description and prediction of
patterns and processes. Advantages of trees include: (1) the flexibility to handle a broad
range of response types, including numeric, categorical, ratings, and survival data; (2)
invariance to monotonic transformations of the explanatory variables; (3) ease and robustness
of construction; (4) ease of interpretation; and (5) the ability to handle missing
values in both response and explanatory variables. Thus, trees complement or represent an
alternative to many traditional statistical techniques, including multiple regression, analysis
of variance, logistic regression, log-linear models, linear discriminant analysis, and survival
models.
We use classification and regression trees to analyze survey data from the Australian
central Great Barrier Reef, comprising abundances of soft coral taxa (Cnidaria: Octocorallia)
and physical and spatial environmental information. Regression tree analyses showed that
dense aggregations, typically formed by three taxa, were restricted to distinct habitat types,
each of which was defined by combinations of 3–4 environmental variables. The habitat
definitions were consistent with known experimental findings on the nutrition of these taxa.
When used separately, physical and spatial variables were similarly strong predictors of
abundances and lost little in comparison with their joint use. The spatial variables are thus
effective surrogates for the physical variables in this extensive reef complex, where information
on the physical environment is often not available.
Finally, we compare the use of regression trees and linear models for the analysis of
these data and show how linear models fail to find patterns uncovered by the trees.
 
DeBano, L. (1981). Water repellent soils: a state of the art. USDA Forest Service GTR-PSW-46. USDA Forest Service GTR-PSW-46, Berkely, California: 21pp.
DeBano, L. (2000). Fire-induced water repellency: an erosional factor in wildland environments. USDA Forest Service Proceedings RMRS-P-13, P.307-310.
DeBano, L. F., R.M. Rice, and C.E. Conrad (1979). Soil heating in chaparral fires: effects on soil properties, plant nutrients, erosion, and runoff. P. S. R. S. USDA Forest Service, RP-PSW-145 21p.
DeBano, L. F. (1981). Water Repellent Soils: a state of the art. General Technical Report, U.S. Forest Service, Pacific Southwest Forest and Range Experiment Station. PSW-46: 21.
Delfino, K. L. (1986). "Landscape Patterns."
Delmas, R. J. (?). "Climatic and Environmental Information from Ice Cores."
Dengler, L., et al. (1987). Bedrock geometry of unchannelized valleys. B. e. a. Beschta. Corvallis, Oregon, IAHS. 165: 81-90.
Dengler, L., et al. (1987). Bedrock geometry of unchannelized valleys. Erosion and Sedimentation in the Pacific Rim. B. Beschta, T. Blinn, G. E. Grant, G. G. Ice and F. J. Swanson. Washington, D.C., IAHS Press. 165: 81-90.
Dengler, L. and D. R. Montgomery (1989). "Estimating the Thickness of Colluvial Fill in Unchanneled Valleys from Surface Topography." Bulletin of the Association of Engineering Geologists 26(3): 333-342.
Denlinger, R. P. and R. M. Iverson (2001). "Flow of variably fluidized granular masses across three-dimensional terrain, 2, Numerical predictions and experimental tests." Journal of Geophysical Research 106(B1): 553-566.
Densmore, A. L. a. N. H. (2000). "Topographic fingerprints of bedrock landslides." Geology 28(4): 371-374.
Deschênes, J., et al. (2007). "Context-dependent responses of juvenile Atlantic salmon (Salmo salar) to forestry activities at multiple spatial scales within a river basin." Canadian Journal of Fisheries and Aquatic Science 64: 1069-1079.
We used classification trees and regression trees to relate the incidence and density of juvenile Atlantic salmon (Salmo salar) to forestry activities measured at four spatial scales (subbasin and 8, 2, and 0.5 km radii up-stream of study sites) and environmental features in 120 stream reaches of the Cascapedia River basin, Quebec, Canada. At all scales, incidence increased with reach size and accessibility to the reach from the river mainstem. Incidence declined with areal coverage of logging at all scales, but only in larger reaches. The time horizon over which logging effects were detected increased with spatial scale. At all scales, density in salmon-bearing reaches increased with accessibility. Density in more accessible reaches was negatively related to logging over the preceding 9 years at the subbasin and 8 km scales, but no effects of logging on density were detected at the 2 and 0.5 km scales. Overall, apparent effects of logging activities on salmon incidence and density were mostly negative and strong, but were both markedly scale-dependent and conditional on environmental context.
 
Desmet, P. J. J., J. Poesen, G.Govers, and K.Vandaele (1999). "Importance of slope gradient and contributing area for optimal prediction of the initiation of ephemeral gullies." Catena 37: 377-392.
Dess, G. (1992). The effects of fine sediment deposition in gravel bed rivers on Salmonid (Onchorhynchus spp.) productivity, ?: 1-34.
DeVries, P., et al. (2001). "Measurement of the temporal progression of scour in a pool-riffle sequence in a gravel bed stream using an electronic scour monitor." Water Resources Research 37(11): 2805-2816.
DeVries, P. E. (1997). "Riverine salmonid egg burial depths: review of published data and implications for scour studies " Canadian Journal of Fisheries and Aquatic Sciences 54: 1685-1698.
DeVries, P. E. (2000). Scour in Low Gradient Gravel Streams: Patterns, Processes, and Implications for the Survival of Salmonid Embryos University of Washington: 250.
DeVries, P. E. (2000). Scour Processes and Pre-emergent Survival of Salmonids in Gravel Bed Streams Seattle, Washington, University of Washington: 44.
DeVries, P. E. (2002). "Bedload Layer Thickness and Disturbance Depth in Gravel Bed Streams " Journal of Hydraulic Engineering 128(11): 983-991.
DeVries, P. E., et al. (2001). "Measurement of the temporal progression of scour in a pool-riffle sequence in a gravel bed stream using an electronic scour monitor " Water Resources Research 37(11): 2805-2816.
Dewberry, C., et al. (1998). The Effects of the Storms of 1996 on a Creek Restoration Project in Oregon, University of Wisconsin: 174-180.
Dewberry, C., et al. (?). The Effects of the Storms of 1996 on Anadromous Fish-Bearing Reaches in the Knowles Creek Watershed, prepared for John Hancock Timber Management Group, Pacific Rivers Council, US Forest Service: 16.
Dhakal, A. S., and R.C.Sidle (2003). "Long-term modelling of landslides for different forest management practices." Earth Surface Processes and Landforms 28(853-868).
Dhakal, A. S. and R. C. Sidle (2004). "Pore water pressure assessment in a forest watershed: Simulations and distributed field measurements related to forest practices." Water Resources Research 40.
A distributed shallow groundwater model related to slope stability is described to
assess the spatial distribution of pore water pressure in steep forested terrain in British
Columbia. Additionally, effects of timber harvesting and roads on measured changes in
pressure head during rainstorms were evaluated for the first time to assess the need for
incorporating different hydrological components in the event-driven distributed model.
Although explicit spatial quantification of pore water pressure requires many
measurements for accurate prediction, model performance using average parameter values
was reasonable when compared with pressure heads measured at nine spatially distributed
sites. Increases in maximum pressure head (varying from 9 to 28 cm) between
preharvesting (after road construction) and postharvesting rainstorm events were observed
in seven of nine sites. The remaining two sites showed either a small decrease ( 5 cm) or
similar peak pressure heads following harvesting. Peak pressure head evaluated at one
piezometer located 46 m downslope of the road decreased substantially ( 50 cm) after
road construction during moderate rainstorms and then recovered following harvesting.
Piezometric responses in sites upslope of the road were not affected by road construction
but did increase after harvesting. Moderate storms caused the largest relative increases in
pressure head between preharvesting (after roads) and postharvesting conditions; such
increases were small during large storms, lending support to the idea that timber
harvesting in temperate forests enhances hydrologic response only during small and
moderate storms. Since landslides in coastal Pacific Northwest are typically caused by
large winter rainstorms, it appears more justified to include better spatial representation of
soil physical and engineering parameters in the distributed shallow groundwater model
compared to specifying evapotranspiration; road hydrology may, however, need to be
included.
 
Di Stefano, C. and V. Ferro (2002). "Linking clay enrichment and sediment delivery processes." Biosystems Engineering 81(4): 465-479.
Di Stefano, C., et al. (2000). "Length slope factors for applying the Revised Universal Soil Loss Equation at basin scale in southern Italy." Journal of Agricultural Engineering Research 75(4): 349-364.
Diaz, G. E. and T. C. Brown (1997). AQUARIUS: A Modeling System for River Basin Water Allocation. Fort Collins, Colorado, US Forest Service: 160.
Dieterich, M. and N. H. Anderson (1998). "Dynamics of abiotic parameters, solute removal and sediment retention in summer-dry headwater streams of western Oregon." Hydrobiologia 379: 1-15.
Dietrich C.R., T. R. G., and A.J. Jakeman (1999). "An analytical model for stream sediment transport: application to Murray and Murrumbidgee river reaches, Australia." Hydrological Processes 13: 763-776.
Dietrich, W. E., D. Bellugi, and R.R. de Asua (2001). Validation of the shallow landslide model, SHALSTAB for forest management. Land Use and Watersheds: Human Influence on Hydrology and Geomorphology in Urban and Forest Areas. 2: 195-227.
Dietrich, W. E., et al. (2001). Validation of the Shallow Landslide Model, SHALSTAB, for Forest Management. Land Use and Watersheds. M. S. Wigmosta and S. J. Burges. Washington, D.C., American Geophysical Union: 195 - 227.
Dietrich, W. E. and R. Dorn (1984). "Signifigance of thick deposits of colluvium on hillslopes: a case study involving the use of pollen analysis in the coastal mountains of Northern California." Journal of Geology 92: 147-156.
Dietrich, W. E. and T. Dunne (1978). "Sediment budget for a small catchement mountainous terrain." Zietshrift fur Geomorphologie: 191-206.
Dietrich, W. E. and T. Dunne (1978). "Sediment budget for a small catchment in mountainous terrain." Zietshrift fur Geomorphologie Suppl. Bd. 29: 191-206.
Dietrich, W. E., et al. (1982). Construction of Sediment Budgets for Drainage Basins Sediment Budgets and Routing in Forested Drainage Basins: Proceedings of the Symposium, Corvallis, Oregon.
Dietrich, W. E., et al. (?). Construction of Sediment Budgets for Drainage Basins: 5-22.
Dietrich, W. E., et al. (1989). "Sediment supply and the development of the coarse surface layer in gravel-bedded rivers." Nature 340: 215-217.
Dietrich, W. E. and D. R. Montgomery (1998). SHALSTAB: A digital terrain model for mapping shallow landslide potential, to be published by NCASI.
Dietrich, W. E., et al. (1995). "A process-based model for colluvial soil depth and shallow landsliding using digital elevation data." Hydrological Processes 9: 383-400.
Dietrich, W. E., et al. (1987). Overview: "Zero-order basins" and problems of drainage density, sediment trasport and hillslope morphology. Erosion and Sedimentation in the Pacific Rim, Corvalilis, OR, International Association of Hydrological Sciences.
Dietrich, W. E. and J. D. Smith (1984). "Bed Load Transport in a River Meander " Water Resources Research 20(10): 1355-1380.
Dietrich, W. E., et al. (1979). "Flow and Sediment Transport in a Sand Bedded Meander " Journal of Geology 87: 305-315.
Dietrich, W. E., et al. (1993). "Analysis of erosion thresholds, channel networks, and landscape morphology using a digital terrain model." Journal of Geology 101: 259-278.
Dietrich, W. E., et al. (1992). "Erosion thresholds and land surface morphology." Geology 20: 675-679.
Dietrich, W. E., et al. (1986). Hollows, colluvium, and landslides in soil-mantled landscapes. Hillslope Processes, Allen and Unwin. A. D. Abrahams. London: 361-388.
Dietrich, W. E., et al. (1990). Hollows, colluvium, and landslides in soil-mantled landscapes. Hillslope Processes, 16th Annual Geomorphology Symposium, Allen and Unwin Ltd.
Dietrich, W. R., J.W. Kirchner, H. Ikeda, and F. Iseya (1989). "Sediment supply and the development of the coarse surface layer in gravel-bedded rivers." Nature, Letters to 340: 215-217.
Diez, J. C., B. Alvera, J. Puigdefabregas, and F. Gallart (1988). Assessing sediment sources in a small drainage basin above timberline in the Pyrenees. Sediment Budgets: Proceedings of the Porto Alegre Symposium, IAHS Publ. Number 174.
Diez, J. R., et al. (2001). "Woody Debris in North Iberian Streams: Influence of Geomorphology, Vegetation, and Management." Environmental Management 28(5): 687-698.
Diez, J. R., et al. (2000). "Effect of removal of wood on streambed stability and retention of organic matter." Journal of the North American Benthological Society 19(4): 621-632.
Dilts, T. E., et al. (2010). "Mapping riparian vegetation with Lidar Data." ArcUser Winter 2010: 18-21.
Dinehart, R. L. (1992). Evolution of Coarse Gravel Bed Forms: Field Measurements at Flood Stage Vancouver, Washington, US Geological Survey: 2667-2689.
Dinehart, R. L. (1992). "Evolution of coarse gravel bed forms: field measurements at flood stage." WRR 28(10): 2667-2689.
Field measurements to investigate the origin and growth of mesoscale gravel bed forms in deep flows were made in the North Fork Toutle River, Washington. Sonar observations of the gravel streambed at a stationary point were recorded during two storm flows in December 1989 and January 1990 with concurrent bed load sampling and continuous velocity measurements. Mean diameter of bed load was about 3 cm, flow depths were 1.4-2.4 m, and bed shear stresses were 2-5 times the critical stress of mean bed load diameter, as computed from the depth-slope product. These records document the hydrodynamic conditions under which dunelike coarse gravel bed forms were observed. Coarse gravel dunes (height, 20cm; length, 6-15 m) evolved more than 24 hours after peak stage, primarily by accretion, as inferred from bed form changes revealed in dual sonar records. Dune heights increased to 40 cm as mean trough elevation rose about 50 cm over several hours. Smaller dunes (wavelength, 1-3 m), transitional from bed load sheets, migrated on the backs of the large dunes. The superposed dunes finally became indistinguishable from the large dunes, which diminished in height by increasing the mean level of troughs. Gravel deposition occurred at the observation point in conjunction with migration of gravel dunes. The direct comparison of known bed form regimes and gravel bar facies provides alternative interpretations of gravelly deposits.
 
Dinehart, R. L. (1992). "Gravel-bed deposition and erosion by bedform migration observed ultrasonically during storm flow, North Fork Toutle River, Washington " Journal of Hydrology 136: 51-71.
Dingman, S. L. (1984). Fluvial Hydrology New York, W. H. Freeman and Company.
Dingman, S. L. (2002). Physical Hydrology, Prentice Hall, New Jersey.
Dinhart, R. L. (1992). "Evoluion of coarse gravel bed forms: field measurements at flood stage." WRR 28(10): 2667-2689.
Field measurements to investigate the origin and growth of mesoscale gravel bed forms in deep flows were made in the North Fork Toutle River, Washington. Sonar observations of the gravel streambed at a stationary point were recorded during two storm flows in December 1989 and January 1990 with concurrent bed load sampling and continuous velocity measurements. Mean diameter of bed load was about 3 cm, flow depths were 1.4-2.4 m, and bed shear stresses were 2-5 times the critical stress of mean bed load diameter, as computed from the depth-slope product. These records document the hydrodynamic conditions under which dunelike coarse gravel bed forms were observed. Coarse gravel dunes (height, 20cm; length, 6-15 m) evolved more than 24 hours after peak stage, primarily by accretion, as inferred from bed form changes revealed in dual sonar records. Dune heights increased to 40 cm as mean trough elevation rose about 50 cm over several hours. Smaller dunes (wavelength, 1-3 m), transitional from bed load sheets, migrated on the backs of the large dunes. The superposed dunes finally became indistinguishable from the large dunes, which diminished in height by increasing the mean level of troughs. Gravel deposition occurred at the observation point in conjunction with migration of gravel dunes. The direct comparison of known bed form regimes and gravel bar facies provides alternative interpretations of gravelly deposits.
 
Diplas, P. Interaction of Fines with a Gravel Bed  (DRAFT): 5-9 - 5-16.
Diplas, P. (1987). "Bedload Transport in Gravel-Bed Streams " Journal of Hydraulic Engineering 113(3): 277-291.
Diplas, P. and A. J. Sutherland (1988). "Sampling techniques for gravel sized sediments." Journal of Hydraulic Engineering(5): 484-501.
DNR Staff (1998). Proposed Forest Practices Rules for implementing the Forestry Module. Olympia, Washington, Department of Natural Resources: 1-61.
Dobbie, M. J., et al. (2008). "Sparse sampling: spatial design for monitoring stream networks." Statistics Surveys 2: 113-153.
Spatial designs for monitoring stream networks, especially e-
phemeral systems, are typically non-standard, ‘sparse’ and can be very
complex, reflecting the complexity of the ecosystem being monitored, the
scale of the population, and the competing multiple monitoring objectives.
The main purpose of this paper is to present a review of approaches to
spatial design to enable informed decisions to be made about developing
practical and optimal spatial designs for future monitoring of streams
 
D'Odorico, P. and S. Fagherazzi (2003). "A probabilistic model of rainfall-triggered shallow landslides in hollows: A long-term analysis." Water Resources Research 39(9): 1262.
The long-term temporal evolution of soil thickness in hollows depends on the processes controlling the rates of colluvium accumulation and erosion. Accumulation is due to soil creep and mass-wasting processes from the adjacent slopes, while erosion of colluvial deposits is mainly due to debris flow and landsliding. An analysis of the long-term evolution of colluvial deposits is developed through a stochastic model of soil mass balance at a point accounting for colluvium infilling, expressed as a deterministic function of the deposit thickness, and soil erosion by shallow landslides, modeled as a random (Poisson) process. Landsliding is related to the characteristics of the triggering precipitation through an infinite-slope stability analysis, a kinematic model of hollow response to rainfall, and the intensity-duration-frequency curves characterizing the regime of extreme precipitation. This analysis provides a probabilistic representation of the long-term dynamics at a point of colluvium thickness as a function of the timescale of hollow infilling and of the frequency of triggering rainfalls. The model is solved both numerically and (under simplified conditions) analytically, showing the existence of different regimes in the temporal evolution of soil thickness. In the case of steep slopes (i.e., with slope angles, β, greater than the soil repose angle, φ) the hollow can be either in a supply-limited state or in event-limited conditions, depending on whether the dynamics are limited by the supply of sediment from the adjacent slopes or by the occurrence of rainstorms able to trigger landslides. Nevertheless, since the likelihood of landslide occurrence increases with increasing values of deposit thickness, colluvium accretion always leads to conditions favorable to landsliding. Vice versa, in the case of gentle slopes (i.e., β < φ) the probability of landsliding decreases with increasing values of soil thickness, and event-limited conditions may evolve into unconditionally stable states.
 
Dodov, B. and E. Foufoula-Georgiou (2006). "Floodplain morphometry extraction from a high resolution digital elevation model: a simple algorithm for regional analysis studies." IEEE Geoscience and Remote Sensing Letters 3(3): 410-413.
In this paper we introduce a fast algorithm for floodplain delineation. The underlying assumption of the proposed procedure is that the inundation depth used to define the lateral extent of the “geomorphologic floodplain” is proportional to the depth of the channel at bankfull flow, a relationship which emerged from a detailed and objective regional analysis. Bankfull and floodplain geometry extracted with the proposed algorithm over a wide range of scales revealed important connections between fluvial and hydrologic processes and showed that the nature of these connections is scale-dependent.
 
Dodson, J. J., et al. (1998). "Elements in the development of conservation plans for Atlantic salmon (Salmo salar)." Canadian Journal of Fisheries and Aquatic Sciences 55: 312-323.
Doe, W. W., D.S. Jones, and S.D. Warren (1999). The soil erosion model guide for military land managers: analysis of erosion models for natural and cultural applications. The Tri-Service CADD/GIS Technology Center, CSU CEMML, Fort Collins, CO,122p.
Dolloff, C. A. (1983). The relationships of wood debris to juvenile salmonid production and microhabitat selection in small southeast Alaska streams. Bozeman, MT, Montana State University: 100.
Dolloff, C. A. and G. H. Reeves (1989). "Microhabitat Partitioning among Stream-Dwelling Juvenile Coho Salmon Oncorhynchus kisutch, and Dolly Varden, Salvelinus malma " Canadian Journal of Fisheries and Aquatic Sciences 47: 2297-2306.
Dominguez, F., J. Canon, and J.B. Valdez, ((in review).). "IPCC-AR4 climate simulations for the Southwestern US: the importance of future ENSO projections." Climate Change.
Dooge, J. C. I. (1986). "Looking for hydrologic laws." Water  Resources Research 22(9): 46S-58S.
Dorn, R. I., et al. (1987). "Isotopic evidence for climatic influence on alluvial-fan development in Death Valley, California." Geology 15: 108-110.
Doten, C. O., L.C. Bowling, J.S. Lanini, E.P. Maurer, and D.P. Lettenmaier (2006). "A spatially distributed model for the dynamic prediction of sediment erosion and transport in mountainous forested watersheds." Water Resources Research 42: W04417, doi:04410.01029/02004WR003829.
Doten, C. O., et al. (2006). "A spatially distributed model for the dynamic prediction of sediment erosion and transport in mountainous forested waterseds." Water Resources Research 42.
Erosion and sediment transport in a temperate forested watershed are predicted with a
new sediment model that represents the main sources of sediment generation in forested
environments (mass wasting, hillslope erosion, and road surface erosion) within the
distributed hydrology-soil-vegetation model (DHSVM) environment. The model produces
slope failures on the basis of a factor-of-safety analysis with the infinite slope model
through use of stochastically generated soil and vegetation parameters. Failed material is
routed downslope with a rule-based scheme that determines sediment delivery to streams.
Sediment from hillslopes and road surfaces is also transported to the channel network. A
simple channel routing scheme is implemented to predict basin sediment yield. We
demonstrate through an initial application of this model to the Rainy Creek catchment, a
tributary of the Wenatchee River, which drains the east slopes of the Cascade Mountains,
that the model produces plausible sediment yield and ratios of landsliding and surface
erosion when compared to published rates for similar catchments in the Pacific Northwest.
A road removal scenario and a basin-wide fire scenario are both evaluated with the model.
 
Dott, R. H. J. (1983). "1982 SEPM Presidential Address: Episoic Sedimentation - How Normal is Average? How Rare is Rare? Does it Matter?" Journal of Sedimentary Petrology 53(1): 0005-0023.
Douglas I (1999). "Hydrological investigations of forest disturbance and land cover impacts  in South-East Asia: a review." Phil. Trans.: The Royal Society London B 354: 1725-1738.
Douglass, J. E., and W.T. Swank (1972). Streamflow modification through management of eastern forests. USDA Forest Service, Research Paper SE-94, Southeastern Forest Experiment Station, Asheville, NC. 16pp.
Downs, P. W. and A. Simon (2001). "Fluvial geomorphological analysis of the recruitment of large woody debris in the Yalobusha River network Central Mississippi, USA." Geomorphology 37: 65-91.
Downs, P. W. and A. Simon (2001). "Fluvial geomorphological analysis of the recruitment of large woody debris in the Yalobusha River network, Cnetral Mississippi, USA." Geomorphology 37: 65-91.
Dragovich, D. and R. Morris (2002). "Fire intensity, run-off and sediment movement in eucalypt forest near Sydney, Australia." Geomorphology Publication - International Association of Geomorphologists 10: 145-164.
Dragovich, D. and R. Morris (2002). "Fire intensity, slopewash and bio-transfer of sediment in eucalypt forest, Australia." Earth Surface Processes and Landforms 27(12): 1309-1319.
Sediment movement (slopewash and bio-transfer), runoff, and organic matter movement (mainly leaf litter, ash and charcoal) were monitored on ten plots for a six-month period following bushfires in eastern Australia in the 1990s. Plots were installed in areas which had experienced high, moderate and low intensity burns. Although between-plot variability was high, slopewash and organic matter amounts were greatest on the intensely burnt areas, and progressively less on moderate and low intensity burns. In contrast, most leaf litter input from scorched leaf fall was derived from, and collected in, moderately burnt areas. Bio-transfer of sediment (direct downslope surface movement produced by faunal activity) occurred mainly by ant mounding, and was more than ten times greater on moderately burnt areas than on plots having high or low intensity burns. Bio-transfer accounted for approximately 36 per cent of total sediment collected, with this material being recorded least often and in smallest quantities on areas of high intensity burns. Bio-transfer by ant mounding and animal scratchings contributed loose surface sediment for transport by overland flow, disproportionately increasing total sediment movement to plot aprons in areas of moderate intensity burns. Abstract Copyright (2002), Wiley Periodicals, Inc.
 
Dragovich, J. D. and M. J. Brunengo (1995). Landslide Map and Inventory, Tilton River - Mineral Creek Area Lewis County, Washington, Washington Division of Geology and Earth Resources.
Dragovich, J. D. and M. J. Brunengo (1995). Landslide map of the Tilton River - Mineral Creek area, Lewis County, Washington. Olympia, Washington Division of Geology and Earth Resources.
Dragovich, J. D., et al. (1993). "Landslide inventory and analysis of the Tilton Creek-Mineral River area, Lewis County, Washington: Part 2: Soils, Harvest Age, and Conclusions." Washington Geology 21(4): 1-32.
Dragovich, J. D., et al. (1993). "Landslide Inventory and Analysis of the Tilton River - Mineral Creek Area, Lewis County, Washington. Part 1: Terrain and Geologic Factors." Washington Geology 21(3): 9-18.
Dragovich, J. D., et al. (1993). "Landslide Inventory and Analysis of the Tilton River - Mineral Creek Area, Lewis County, Washington. Part 2: Soils, Harvest Age, and Conclusions." Washington Geology 21(4): 18-30.
Dragovich, J. D., et al. (1993). "Landslide Inventory and Analysis of the Tilton River-MIneral Creek Area Lewis County Washington." Washington Geology 21(4): 18-30.
Drury, T. A. (2001). Steelhead Haven Landslide Remediation Feasibility Study, Prepared for The Stillaguamish Tribe of Indians by GeoEngineers, Inc.: 23.
Duan, J., et al. (1995). A coupled hydro-geomorphic and ecologic model for examining effects of forest cutting on watershed process. Eos, Transactions, American Geophysical Union. 76: 261.
Duan, J. G. and G. D. Gylsson (2001). "An analytical approach to estimate soil erosion from roadside ditches in mountain forest." Proceedings - Federal Inter-Agency Sedimentation Conference 7, Volume 2: V55-V62.
Road construction has increased the run-off erosion and silt-producing potential from road cut banks and ditches along highways. Road ditches are designed to transport flow and sediment coming from bank cuts. Erosion occurs when flow-induced shear stress is sufficient to entrain bed or bank material. This paper reports the development of an analytical approach to estimate soil erosion from road cut banks and ditches. The result indicated that sediment yield could be calculated based on runoff, the slope and length of road cut banks, the longitudinal slope of road, and pavement in road ditches. Highway engineers can use this approach to estimate soil erosion in the design of roadside ditches.
 
Duck Creek Associates, Inc. (2008). Wilson River Watershed Analysis. O. D. o. Forestry. Salem, OR.
Dudziak, J. (1974). "Observations of development of slope rills in forest clearings of the Tatra Mountains.(In Polish with English summary)." Czasopismo geograficzne; kwartalnick Zrzeszenia Pol. Nauczycieli Geografji, Towarzystwa Geograficznego we Lwowie i Towarzystwa Geograficznego w Pozna 45(1): 31-45.
Duijsings, J. J. H. M. (1987). "A Sediment Budget for a Forested Cathment in Luxembourg and it's Implications for Channel Development." Earth Surface Process and Landforms 12: 173-184.
Duncan, J. M. and S. G. Wright (1980). "The accuracy of equilibrium methods of slope stability analysis." Engineering Geology 16(1): 5-17.
Duncan, S. H., R.E. Bilby, L.W. Ward, and J.T. Heffner (1987). "Transport of road surface sediment through ephemeral stream channels." Water Resources Bulletin 23(1): 113-119.
Duncan, S. H. (?). "Slope Stability Analysis in Timber Harvest Planning: Smith Creek, Pacific County, Washington." Engineering Geology in Washington (is this right?) II: 927-932.
Duncan, S. H., et al. (1987). "Transport of road-surface sediment through ephermal stream channels." Water Resources Bulletin 23(1): 114-119.
Duncan, S. H. and J. V. Ward (1985). "The influence of watershed geology and forest roads on the composition of salmon spawning gravel." Northwest Science 59(3): 204-212.
The composition of stream gravels used for spawning by Pacific salmon (Oncorhynchus spp.) was examined in 12 streams in southwestern Washington State, whose watersheds have been actively managed for commercial timber production. The amount of fine sediment (< 2 mm in size) was more closely correlated to the lithology and soils of the watershed than to forest management practices, specifically forest roads. None of the sampled stream reaches contained sediment in amounts considered to impair salmonid survival. Streams in watersheds dominated by basalt rock had a higher percentage of sand particles 0.5-2.0 mm and lower percentages of silt and clay (particles < 0.063 mm) than did streams in watersheds dominated by siltstone and graywacke sandstone rock. A significant positive correlation was obtained between percentage of watershed area in sedimentary rock, and percentage of medium and fine sand and silt and clay particles in spawning gravels. Assessment of the potential of a watershed to produce fine sediment is possible by examination of basin geology and soils.
 
Duncan, S. H. and J. V. Ward (1985). "A technique for measuring scour and fill of salmon spawning riffles in headwater streams." Water Resources Bulletin 21(3): 507-511.
Duncan, S. H., et al. (1987). "A Method for Assessing Landslide Potential as an Acid in Forest Road Placement." Northwest Science 61(3): 152-162.
Dunham, J. B., et al. (2002). "Influences of spatial and temporal variation on fish-habitat relationships defined by regression quantiles." Transactions of the American Fisheries Society 131: 86-98.
We used regression quantiles to model potentially limiting relationships between the standing crop of cutthroat trout Oncorhynchus clarki and measures of stream channel morphology. Regression quantile models indicated that variation in fish density was inversely related to the width:depth ratio of streams but not to stream width or depth alone. The spatial and temporal stability of model predictions were examined across years and streams, respectively. Variation in fish density with width:depth ratio (10th–90th regression quantiles) modeled for streams sampled in 1993–1997 predicted the variation observed in 1998–1999, indicating similar habitat relationships across years. Both linear and nonlinear models described the limiting relationships well, the latter performing slightly better. Although estimated relationships were transferable in time, results were strongly dependent on the influence of spatial variation in fish density among streams. Density changes with width:depth ratio in a single stream were responsible for the significant (P < 0.10) negative slopes estimated for the higher quantiles (>80th). This suggests that stream-scale factors other than width:depth ratio play a more direct role in determining population density. Much of the variation in densities of cutthroat trout among streams was attributed to the occurrence of nonnative brook trout Salvelinus fontinalis (a possible competitor) or connectivity to migratory habitats. Regression quantiles can be useful for estimating the effects of limiting factors when ecological responses are highly variable, but our results indicate that spatiotemporal variability in the data should be explicitly considered. In this study, data from individual streams and stream-specific characteristics (e.g., the occurrence of nonnative species and habitat connectivity) strongly affected our interpretation of the relationship between width:depth ratio and fish density.
 
Dunham, J. B., et al. (1999). "Local and geographic variability in the distribution of stream-living Lahontan Cutthroat Trout." Transactions of the American Fisheries Society 128: 875-889.
We investigated local and geographic variability in the up- and downstream distribution
limits of threatened Lahontan cutthroat trout (Oncorhynchus clarki henshawi) in stream
habitats of the eastern Lahontan basin in northern Nevada and southeastern Oregon. At a geographic
scale, elevations of upstream distribution limits were significantly correlated with latitude and
longitude, suggesting a potential influence of climatic gradients. Elevations of upstream distribution
limits also were positively correlated with maximum basin elevation, which suggested topographic,
rather than climatic constraints may be important. Upstream distribution limits were not signifi-
cantly affected by local variation in stream size or presumptive dispersal barriers. Stream gradient
was related to upstream distribution limits, but this was again confounded by maximum basin
elevation. Stream gradients used by Lahontan cutthroat trout at upstream limits were considerably
steeper than those observed for other subspecies of cutthroat trout in other areas. Geographic
variation in elevations of downstream distribution limits was also apparent, paralleling variability
in summer air temperatures and presumed restriction of Lahontan cutthroat trout from lower
elevation habitats by unsuitably warm temperatures in summer. Previous regional models assumed
downstream distribution limits for salmonids correspond to isotherms of 22–248C (mean July air
temperature), but we found the distribution of Lahontan cutthroat trout to be considerably more
restricted, most closely corresponding to a mean July air temperature of 188C. Occurrence of
nonnative brook trout Salvelinus fontinalis was associated with a significant upstream shift of the
downstream distribution of Lahontan cutthroat trout, suggesting a highly variable, but negative
interaction between these species.
 
 
Dunham, J. B. and B. E. Rieman (1999). "Metapopulation structure of Bull Trout: influences of physical, biotic, and geometrical landscape characteristics." Ecological Applicatons 9(2): 642-655.
Metapopulation structure of species in fragmented landscapes is ultimately the result of spatial variability in demographic processes. While specific information on demographic parameters is desirable, a more practical approach to studying metapopulations in fragmented landscapes may begin with analyses of species’ occurrence in relation to large-scale habitat variability. Here, we analyzed occurrence of stream-living bull trout (Salvelinus confluentus) in relation to physical, biotic, and geometrical characteristics of habitats. Bull trout occurrence was analyzed at several spatial (10x m) scales. Data were from nested sampling of 720 sites (10 m), 179 reaches (102 m), and 81 patches (103 m) of stream habitats within the Boise River basin of central Idaho. Based on previous findings, patches were defined as stream catchments with suitable conditions for spawning and rearing of bull trout (>1600 m elevation). Patch-scale bull trout occurrence was significantly related to patch area and isolation (stream distance between occupied patches). Lack of spatial autocorrelation between patches indicated that isolation effects were more likely the result of limited interaction among habitats (such as dispersal), rather than of correlated environmental conditions. A third factor, human disturbance in the form of roads, was associated with reduced bull trout occurrence at the patch-scale. Analyses of occurrence among reaches within occupied patches showed bull trout may select larger (>2 m width) stream habitats. Occurrence of bull trout was not associated with nonnative brook trout (Salvelinus fontinalis) at large (patch), intermediate (reach), or small (site) spatial scales. Definition of a metapopulation structure for bull trout in the Boise River basin was complicated by uncertainties in the frequency and magnitude of dispersal. From the distribution of patch sizes and isolation among occupied patches, we suggest that the metapopulation is a complex mosaic of several elements found in conceptual models. This complexity poses a challenge to empirical and theoretical attempts to study stream-living bull trout. Future work to define the structure of bull trout metapopulations must relate temporal and spatial patterns of patch occupancy with complex patterns of dispersal that likely interact with habitat spatial structure, life history variability, and the historical context of regional climate changes. Results of this work suggest that conservation of bull trout should involve protection of larger, less isolated, and less disturbed (as indexed by road densities) habitats that may serve as important refugia or sources of recolonization. Bull trout populations in smaller, isolated, and more disturbed habitats may be at risk of extinction. Finally, metapopulation structure implies the existence of suitable, but presently unoccupied habitat, which should be managed carefully to facilitate potential natural recolonization or reintroductions of bull trout
 
Dunham, J. B., et al. (2002). Patch-based models to predict species occurrence: lessons from salmonid fishes in streams. Predicting species Occurrences: Issues of Accuracy and Scale. J. M. Scott, P. J. Heglund, F. Samson et al. Covelo, CA, Island Press: 327-334.
Dunham, J. B., et al. (2003). "Effects of fire on fish populations: landscape perspectives on persistence of native fishes and nonnative fish invasions." Forest Ecology and Management 178: 183-196.
Our limited understanding of the short and long-term effects of fire on fish contributes to considerable uncertainty in
assessments of the risks and benefits of fire management alternatives. A primary concern among the many potential effects of fire
is the effects of fire and fire management on persistence of native fish populations. Limited evidence suggests vulnerability of
fish to fire is contingent upon the quality of affected habitats, the amount and distribution of habitat (habitat fragmentation), and
habitat specificity of the species in question. Species with narrow habitat requirements in highly degraded and fragmented
systems are likely to be most vulnerable to fire and fire-related disturbance. In addition to effects of fire on native fish, there are
growing concerns about the effects of fire on nonnative fish invasions. The role of fire in facilitating invasions by nonnative fishes
is unknown, but experience with other species suggests some forms of disturbance associated with fire may facilitate invasion.
Management efforts to promote persistence of fishes in fire-prone landscapes can take the form of four basic alternatives: (1) prefire
management; (2) post-fire management; (3) managing fire itself (e.g. fire fighting); and (4) monitoring and adaptive
management. Among these alternatives, pre-fire management is likely to be most effective. Effective pre-fire management
activities will address factors that may render fish populations more vulnerable to the effects of fire (e.g. habitat degradation,
fragmentation, and nonnative species). Post-fire management is also potentially important, but suffers from being a reactive
approach that may not address threats in time to avert them. Managing fire itself can be important in some contexts, but negative
consequences for fish populations are possible (e.g. toxicity of fire fighting chemicals to fish). Monitoring and adaptive
management can provide important new information for evaluating alternatives, but proper implementation is often hampered by
inadequate study designs and inconsistent financial and institutional support. The challenge for providing better management
guidelines will be to add solid empirical data and models to assess the relevance of emerging concepts and theories, and provide
a sense of where and when fires pose significant risks and/or benefits to fishes.
 
Dunkerley, D. L. (1994). "Discussion: Bulk  sampling of coarse clastic sediments for particle-size analysis." Earth Surface Processes and Landforms 19: 255-261.
Dunne, J., et al. (1999). "Scaling factors for the rates of production of cosmogenic nuclides for geometric shielding and attenuation at depth on sloped surfaces." Geomorphology 27: 3-11.
Dunne, T., and Leopold, L. (1978). Water in environmental planning, W. H. Freeman & Company. San Francisco.
Dunne, T. (1980). "Formation and controls of channel networks." Progress in Physical Geography 4: 211.
Dunne, T. (1984). Effects of the Twin Falls and Weeks Falls Projects on Sedimentation Along the Snoqualmie River System. Seattle, Washington, HYDRA Geotechnical Group: 1-55.
Dunne, T. (1984). The prediction of erosion in forests.  Keynote address. Symposium on Effects of Forest Land Use on Erosion and Slope Stability., East-West Center, University of Hawaii, Honolulu, East West Center.
Dunne, T. (1988). Contributions of geomorphology to flood-control planning. Flood Geomorphology. R. C. K. V.R. Baker, and P.C. Patton. New York, Wiley and Sons.
Dunne, T. (1989). "Debris-Flow Fans (DRAFT)." 22.
Dunne, T. (1989). Unsolicited Proposal Submitted to the National Science Foundation Earth Sciences Division.
Dunne, T. (1991). "Stochastic Aspects of the Relations between Climate, Hydrology and Landform Evolution." Transactions of the Japanese Geomorphological Union 12(1): 1-24.
Dunne, T. (1998). "Critical data requirements for prediction of erosion and sedimentation in mountain drainage basins." Journal of the American Water Resources Association 34(4): 795-808.
Dunne, T. (2001). "A Scientific Basis for the Prediction of Cumulative Watershed Effects." Retrieved 6/3/2002, 2002, from http://www.cnr.berekely.edu/forestry/cwe/cwe_i.html.
Dunne, T. (?). Hydrology, Mechanics and Geomorphic Implications of Erosion by Subsurface Flow (DRAFT). Seattle, Washington, Department of Geological Sciences, University of Washington: 89.
Dunne, T. (?). The Prediction of Erosion in Forests. Symposium on Effects of Forest Land Use on Erosion and Slope Stability, Honolulu, Hawaii.
Dunne, T., et al. (2001). A scientific basis for the prediction of cumulative watershed effects. University of California Wildland Resource Center Report No. 46. Berkeley, CA, University of California Wildland Resource Center.
Dunne, T. and B. F. Aubry (?). Evaluation of Horton's theory of sheetwash and rill erosion on the basis of field experiments ? Chapter 2.
Dunne, T. and R. D. Black (1970). "An experimental investigation of runoff production in permeable soils." Water Resources Research 6: 478-490.
Dunne, T. and R. D. Black (1970). "An experimnetal investigation of runoff production in permeable soils." Water Resources Research 6: 478-490.
Dunne, T. and L. H. Fairchild "Estimation of flood and sedimentation hazards around Mt. St. Helens."? 36(4): 12-22.
Dunne, T. and L. B. Leopold (1978). Water in Environmental Planning. New York, W. H. Freeman and Company.
Dunne, T., et al. (1995). Microtopography of hillslopes and initiation of channels by horton overland flow. Natural and Anthropogenic Influences in Fluvial Geomorphology, The Wolman Volume. J. E. Costa, A. J. Miller, K. W. Potter and P. R. Wilcock. Washington, D.C., American Geophysical Union. Geophysical Monograph 89: 27-44.
Dunwiddie, P. W. (1986). "A 6000-year record of forest history on Mount Rainier, Washington." Ecological Society of America 67(1): 58-68.
Duvall, A., et al. (2004). "Tectonic and lithologic controls on bedrock channel profiles and processes in coastal California." Journal of Geephysical Research 109(F03002): 18.
Recent theoretical models suggest that topographic characteristics of bedrock
channels are products of interactions among tectonics, substrate resistance, and the
climatically modulated erosive ability of the river. The degree to which these factors
influence the form of channel profiles is poorly quantified at present. Here we investigate
bedrock channels developed across the southern flank of the Santa Ynez Mountains,
California. Uniform climate and systematic variations in lithology and rock uplift rate
along the range allow comparison of channel morphology between (1) channels eroding
rocks of uniform and nonuniform strength and (2) channels experiencing differences in
tectonic forcing. We combine field observations, surveys, and analysis of digital data
to determine topographic and hydraulic characteristics of bedrock channels. At a constant
rock uplift rate, streams flowing from resistant to less resistant bedrock exhibit highly
concave profiles and increased gradients along lower reaches relative to channels
developed in uniform bedrock. These effects are interpreted as responses to (1) an increase
in substrate resistance to channel incision in the upper reaches and (2) transport-limited
gradients along lower reaches. Comparisons of channels developed across uniform
lithology but experiencing an approximately sevenfold difference in rock uplift rate reveal
an approximately twofold increase in gradient and an approximately threefold decrease
in width. In this landscape the combined channel adjustments of gradient and width are
consistent with a fluvial incision model in which channel incision rate is linearly
proportional to mean bed shear stress.
 
Dwire, K. A. and J. B. Kauffman (2003). "Fire and riparian ecosystems in landscapes of the western USA." Forest Ecology and Management 178: 61-74.
Despite the numerous values of riparian areas and the recognition of fire as a critical natural disturbance, few studies have
investigated the behavior, properties, and influence of natural fire in riparian areas of the western USA. Riparian areas frequently
differ from adjacent uplands in vegetative composition and structure, geomorphology, hydrology, microclimate, and fuel
characteristics. These features may contribute to different fire environments, fire regimes, and fire properties (frequency,
severity, behavior, and extent) in riparian areas relative to uplands. In certain forested riparian areas, fire frequency has generally
been lower, and fire severity has been more moderate than in adjacent uplands, but in other areas, fires have appeared to burn
riparian areas with comparable frequency. Impacts of land use and management may strongly influence fire properties and
regimes in riparian areas. Fire suppression, livestock grazing, logging, damming and flow regulation, agricultural diversions,
channel modifications, and introduction of invasive species have led to shifts in plant species composition, structure and
distribution of fuel loads, and changes in microclimate and areal extent of riparian areas. Cumulative impacts of human
alterations are likely to exert the most pronounced influence on fire behavior during periods of drought and under conditions of
extreme fire weather. Riparian plant species possess adaptations to fluvial disturbances that facilitate survival and reestablishment
following fires, thus contributing to the rapid recovery of many streamside habitats. Given the critical resource values of
riparian zones, additional data are needed to understand interactions between fire and riparian ecosystems, and how riparian
zones affect spatial and temporal patterns of fires at the landscape scale. An improved understanding of fire ecology and effects
in riparian areas is needed to prescribe ecologically sound rehabilitation projects following fire.
 
 
Dwyer, W. P., et al. (1993). "Influence of Electroshock and Mechanical Shock on Survival of Trout Eggs " North American Journal of Fisheries Management 13: 839-843.
Dyrness, C. T. (1967). Mass soil movement in the H. J. Andrews Experimentat Forest, U.S.D.A Forest Service: 12.
Dyrness, C. T. (1967). Mass soil movements, US Forest Service: 1-12.
Eagleson, P. S. (1972). "Dynamics of flood frequency." Water Resources Research 8: 878-898.
Earle, C. J., et al. (1996). "Charcoal in northcentral Alaskan lake sediments: relationships to fire and late-Quaternary vegetation history." Review of Palaeobotany and Palynology 92: 83-95.
Earth Systems Institute (2002). River Network Model. Seattle, WA, Earth Systems Institute: Software.
Easterbrook, D. J. (1999). Quaternary Climatic Changes and the Ice Ages Surface Processes and Landforms. Upper Saddle River, New Jersey, Prentice Hall. Second Edition: 365-399.
Eaton, L. S., et al. (2003). "Role of debris flows in long-term landscape denudation in the central Appalachians of Virginia." Geology 31(4): 339-342.
Four major storms that triggered debris flows in the Virginia{ndash}West Virginia Appalachians provide new insights into the role of high-magnitude, low-frequency floods in long-term denudation and landscape evolution in mountainous terrain. Storm denudation in the Blue Ridge Mountain drainage basins is approximately an order of magnitude greater compared to basins located in the mountains of the Valley and Ridge province. This difference is probably the result of higher storm rainfall from the Blue Ridge storms. Radiocarbon dating of debris-flow deposits in the Blue Ridge indicates a debris-flow return interval of not more than 2{ndash}4 k.y. in mountainous river basins. This finding, combined with measurements of basin denudation, suggests that approximately half of the long-term denudation from mechanical load occurs episodically by debris-flow processes. Although floods of moderate magnitude are largely responsible for mobilizing sediment in low-gradient streams, our data suggest that high-magnitude, low-frequency events are the most significant component in delivering coarse-grained regolith from mountainous hollows and channels to the lowland floodplains.
 
Ebersole, J. L., et al. (2006). "Juvenile coho salmon growth and survival across stream network seasonal habitats." Transactions of the American Fisheries Society 135: 1681-1697.
Abstract.—Understanding watershed-scale variation in juvenile salmonid survival and growth can provide
insights into factors influencing demographics and can help target restoration and mitigation efforts for
imperiled fish populations. We assessed growth, movement, and apparent overwinter survival of individually
tagged juvenile coho salmon Oncorhynchus kisutch in a coastal Oregon watershed from June 2002 to June
2003 and related growth and survival parameters to stream characteristics. Fall body size of juvenile coho
salmon was a good predictor of smolt size and survival, but smolt size was also influenced by overwintering
location. This was due to strong spatial patterns in winter growth rates associated with residency and
movement into a small intermittent tributary. Though nearly dry in midsummer, this stream supported high
densities of spawning coho salmon in the fall, and juveniles rearing there exhibited relatively high growth
rates and emigrated as larger smolts. Improved winter growth and survival of juvenile coho salmon utilizing
tributary habitats underscore the importance of maintaining connectivity between seasonal habitats and
providing a diversity of sheltering and foraging opportunities, particularly where main-stem habitats have
been simplified by human land uses.
 
Edwards, R. T. (1998). The hyporheic zone. River ecology and management: lessons from the Pacific coastal ecosystem. R. Naiman and R. E. Bilby. New York, Springer: 399-429.
Effects, T. U. o. C. C. o. C. W. (2001). A scientific basis for the prediction of cumulative watershed effects. University of California Wildland Resource Center Report No. 46, 98pp.: 98pp.
Eide, J. (1990). A 48 year sediment budget (1942-1989) for Deer Creek basin, Washington, Western Washington University.
Eiler, J. H., et al. (1992). "Riverine spawning by sockeye salmon in the Taku River, Alaska and British Columbia." Transactions of the American Fisheries Society 121(6): 701-708.
Abstract.— Radio telemetry was used to determine the distribution of sockeye salmon Onco- rhynchus nerka returning to spawn in the glacial Taku River in 1984 and 1986,and to locate and characterize spawning areas used by this species.During the study,253 sockeye salmon were tracked as they moved upriver;204 of these were followed to spawning areas.Only 37%of the 204 fish traveled to areas associated with lakes;the remaining 63%returned to "riverine"areas- river areas without lakes (42%to the Taku River main stem,17%to the Nakina River,and 4% to other rivers).Sockeye salmon spawning in riverine areas used a variety of habitat types,including main-river channels,side channels,tributary streams,and upland sloughs.Most (55%)of the radio- tagged fish that returned to the Taku River main stem were tracked to side-channel spawning areas.Half of the 471 adult sockeye salmon sampled in main-stem spawning areas had migrated to sea as juveniles before their first winter.This study showed that many sockeye salmon returning to the Taku River do not depend on lakes,and that riverine sockeye salmon make up a major portion of the run in some river systems.
 
Einstein, H. A. and R. B. Banks (1950). "Fluid resistance of composite roughness " Transactions of the American Geophysical Union 31(4): 603-610.
Eisbacher, G. H. (1982). "Mountain Torrents and Debris Flows." Episodes 4: 12-17.
Ellen, S. D. and R. K. Mark (1993). Mapping debris-flow hazard in Honolulu using a DEM. Proceedings, Hydraulic Engineering '93. H. W. Shen, S. T. Su and F. Wen. New York, American Society of Civil Engineers. 2: 1774-1779.
Ellen, S. D., et al. (1993). Preliminary map of debris-flow hazard in the Honolulu District of Oahu, Hawaii: U.S. Geological Survey Open-File Report 93-0213., U.S. Geological Survey.
Elliot, E. (2007). "WEPP Internet interfaces for forest erosion prediction." Journal of American Water Resources Association 40(2): 299-309.
Elliot, W., R. Foltz, and C. Luce (1995). Validation of the water erosion prediction project for low volume forest roads. Sixth International Conference on Low Volume Roads, Minneapolis, MN.
Elliot, W. J., R.B. Foltz, and P.R. Robichaud (1994). A tool for estimating disturbed forest site sediment production. Interior cedar-hemlock-white pine forests; ecology and management: symposium proceedings., Spokane, WA.
Elliot, W. J., K. Hyde, L. MacDonald, and J. McKean (2006). Tools for analysis. CWE of fuel management. W. J. Elliot, and L.J. Andrei.
Elliot, W. J. and D. E. Hall (1997). Water Erosion Prediction Project (WEPP) forest applications. Moscow, ID, Intermountain research Station, USDA Forest Service: 11.
Elliot, W. J., et al. (?). "Predicting Sedimentation from Forest Roads " Journal of Forestry: 23-29.
Elliot, W. J., et al. (2000). The Forest Service WEPP interfaces. 2000 ASAE Summer Meeting, St. Joseph, MI, ASAE.
Elliott, J. G., et al. (1999). Estimation of fire/flood chronology from alluvial fan stratigraphy in the Buffalo Creek watershed, Colorado. Abstracts with Programs - Geological Society of America. 31: 441.
Stratigraphic and geomorphic evidence indicates that sequential forest fire/flood events have been intermittent but common in many mountainous areas during the past several thousand years. The recurrence of fire/flood events may reflect the joint probability between the recurrence of fires and the recurrence of large precipitation events. In the first several months following the May 1996 Buffalo Creek, Colorado, forest fire, serious flood and sedimentation damage occurred to stream channels, roadways, bridges, domestic and commercial structures, and water-supply facilities. Precipitation amounts and intensities that generated very little surface runoff outside of the burn area resulted in hillslope erosion, floods, and streambed sediment entrainment in the rugged, severely-burned, 48 square kilometer area. Stratigraphic interpretation, dendrochronologic dating, and carbon-14 dating of Buffalo Creek alluvial deposits and of similar deposits in other forested foothills regions can be used to develop a recurrence probability for large-magnitude fire/flood events. Forested alluvial fans are present at the confluence of the main stem of Buffalo Creek and most tributaries. Following the 1996 fire, floods added sediment to many existing alluvial fans, while simultaneously incising other fans. Incision of older fans revealed alternating sequences of gravel and fine-grain sediments representing periods of high sediment transport and fan aggradation, followed by intervals of quiescence. An alluvial-fan sequence at the mouth of a tributary draining a 1.3 square kilometer area indicated at least five previous fire/flood cycles in the watershed. Dendrochronologic and radiocarbon dating of material in this fan spans approximately 2,900 years. An initial estimate of the interval between large fire/flood events ranged from approximately 1.5 to 9 centuries.
 
Elliott, W. J., A.V.Elliot, Qiong, W., and J.M. Laflen (1991). "Validation of the WEPP model with rill erosion plot data." ASAE.
Elliott, W. J., and I.S. Miller (2002). Estimating erosion impacts from implementing the national fire plan. Paper Number 02-5011. ASAE Annual International Meeting / CIGR XVth World Congress, Chicago, Ill.
Embrechts, J. and M. d. Dapper (1987). "Morphology and genesis of hillslope pediments in the Febe area (South-Cameroon)." Catena 14: 31-43.
Emmett, W. a. M. W. (2001). "Effective discharge and gravel-bed rivers." Earth Surface Processes and Landforms 26: 1369-1380.
England, P. and P. Molnar (1990). "Surface uplift, uplift of rocks, and exhumation of rocks." Geology 18: 1173-1177.
Environmental Protection Agency, E. (2000). "National Water Quality Inventory, 1998 report to Congress.". from http://www.epa.gov/305b/98report/. Accessed 5/5/06.
Erman, D. C., et al. (1988). "Effects of Winter Floods on Fishes in the Sierra Nevada " Canadian Journal of Fisheries and Aquatic Sciences 45: 2195-2200.
Erman, N. A. (1996). Status of Aquatic Invertebrates Davis, California, Unversity of California, Davis: 987-1008.
Eros, T., et al.
Eros, T., et al. (2011). "Network thinking in riverscape conservation - a graph based approach." Biological Conservation 144: 184-192.
Erskine, R. H., et al. (2006). "Comparison of grid-based algorithms for computing upslope contributing area." Water Resources Research 42.
Terrain attributes based on upslope contributing area, A, are used widely in distributed
hydrologic models. Several grid-based algorithms are available for estimating A. In this
study, five algorithms (D8, 8, MFD, DEMON, and D
1
) were compared quantitatively
on two undulating agricultural fields (63 and 109 ha) in northeastern Colorado. Global
positioning system (GPS) data (0.02-m accuracy) were used to generate grid digital
elevation models (DEMs) at 5-, 10-, and 30-m cell sizes. Relative differences between A
values estimated using single- and multiple-direction algorithms increased with
decreasing grid cell size. Relative differences were greatest along ridges and side slopes, and
differences decreased where the terrain became more convergent. Multiple-direction
algorithms (MFD, DEMON, and D
1
), allowing for flow divergence, are recommended on
these undulating terrains for 5- and 10-m grids where A is most sensitive to the algorithm
selection.
 
Erskine, W. D., et al. (2002). "Land use effects on sediment yields and soil loss rates in small basins of Triassic sandstone near Sydney, NSW, Australia." Catena 49(4): 271-287.
Espinosa, F. A., Jr. and K. M. Lee (1991). Natural propagation and Habitat Improvement Idaho: LoLo Creek and upper Lochsa, U.S. Bonneville Power Administration and U.S. Forest Service: 1-103.
ESRI (1998). ESRI Shapefile Technical Description, An ESRI White Paper, Environmental Systems Research Institute, Inc.: 34.
Etienne, R. S. and J. A. P. Heesterbeek (2001). "Rules of thumb for conservation of metapopulations based on a stochastic winking-patch model." The Amercian Naturalist 158: 389-407.
Evans, B. F., et al. (1993). "Distribution and abundance of coarse woody debris in some southern New Zealand streams from contrasting forest catchments." New Zealand Journal of Marine and Freshwater Research 27: 227-239.
Greater amounts of coarse woody debris (CWD) occurred in streams from old native forests than in streams from young native and pine forests in southern New Zealand. The size of CWD in the streams generally reflected the age of the surrounding vegatation. More wood was present in pool than in non-pool sections of old native forest streams and the frequency of pools per unit length formed by woody debris was greatest in these streams. The volumes of pools formed by wood and those formed by inorganic substrates were similar. Amounts of woody debris in these streams were relatively small compared to values recorded from North America.
 
Evans, B. M., D.W. Lehning, K.J. Corradini, G.W. Peterson, E. Nizeyimana, J.M. Hamlett, P.D. Robillard, and R.L. Day (2002). "A comprehensive GIS-based modeling approach for predicting nutrient loads in watershed." Journal o f Spatial Hydrology 2(2): 1-18.
Evans BM, S. S., and DW Lehning (2003). "A spatial technique for estimating  streambank erosion based on watershed characteristics." Journal of Spatial Hydrology 3(1): 2003.
Evans, K. G. (2000). "Methods for assessing mine site rehabilitation design for erosion impact." Australian Journal of Soil Research 38(2): 231-247.
Evans, S. (1994). "Draining Seattle - WPA Landslide Stabilization Projects, 1935-1941." Washington Geology 22(4): 3-10.
Everest, F. H. (1987). Salmonids of Western Forested Watersheds. Streamside Management: Forestry and Fishery Interactions. E. D. Salo and T. W. Cundy. Seattle, Washington. Chapter 1: 3-8.
Everest, F. H., et al. (1987). Fine Sediment and Salmonid Production: A Paradox. Streamside Management: Forestry and Fishery Interactions. E. O. Salo and T. W. Cundy. Seattle, WA, USA, College of Forest Resources, University of Washington: 98-142.
Everest, F. H. and W. R. Meehan (1981). "Forest management and anadromous fish habitat productivity." Transactions of the North American Wildlife and Natural Resources Conference 46: 521-530.
Everest, F. H. and W. R. Meehan (1981). Some effects of debris torrents on habitat of anadromous salmonids, National Council for the Paper Industry for Air and  Stream Improvement: 23-30.
Everest, F. H. and G. H. Reeves (2007). Riparian and aquatic habitats of the Pacific Northwest and Southeast Alaska: Ecology, Management History and Potential Management Strategies, Pacific Northwest Research Station: 130p.
Everitt, B. L. (1968). "Use of cottonwood in an investigation of the recent history of a floodplain." American Journal of Science 266: 417-439.
Fabricius, C. and K. Coetzee (1992). "Geographic information systems and artificial intelligence for predicting the presence or absence of mountain reedback." South Afr. Wildl. Res. 22: 80-86.
Fagan, W. F. (2002). "Connectivity, fragmentation, and extinction risk in dendritic metapopulations." Ecology 83: 3243-3249.
Fahnestock, R. K. (1963). Morphology and Hydrology of a Glacial Stream - White River, Mount Rainier Washington, Department of the Interior: 67.
Fahnestock, R. R. and J. K. Agee (1983). "Biomass Consumption and Smoke Production By Prehistoric and Modern Forest Fires In Western Washington." Journal of Forestry 81(10): 653-656.
Fairfield, J. and P. Leymarie (1991). "Drainage networks from grid digital elevation models." Water Resources Research 27(5): 709-717.
Famiglietti, J., et al. (2008). "Community modeling in hydrologic science." EOS, Transaction American Geophysical Union 89(32): 292.
Fannin, R. J. and J. Jaakkola (1999). "Hydrological response of hillslope soils above a debris-slide headscarp." Canadian Geotechnical Journal 36(6): 1111-1122.
The Jamieson Creek debris slide initiated in thin soils over a competent bedrock surface, on a planar
section of hillslope, during a heavy rainstorm in November 1990. An array of automated piezometers and tensiometers
was placed along a 22 m wide section of the headscarp in 1997 to monitor the temporal variation of pore-water
pressures. Interpretation of the data addresses the hydrologic response to the storms in October and November 1997.
The piezometers, which were designed for installation by driving, reveal very localized responses in what otherwise
appears to be a uniform soil matrix. Peak positive pressures occur at the time of maximum rainfall intensity. The
tensiometers indicate the hydrological response at the ground surface appears uncoupled from that at the bedrock
interface. Implications of the extreme spatial variability in pore-water pressure are evaluated for conceptual models of
hillslope hydrology. The assumption of parallel seepage flow is widely adopted in translational slope stability analyses,
imposing a linear distribution of pore-water pressure with depth. None of the reported field data are consistent with
such a linear distribution with depth or a uniform response across the slope.
 
Fannin, R. J. and T. P. Rollerson (1993). "Debris flows: some physical characteristics and behavior." Canadian Geotechnical Journal 30(1): 71-81.
Fannin, R. J. and M. P. Wise (2001). "An empirical-statistical model for debris flow travel distance." Canadian Geotechnical Journal 38(5): 982-994.
Farella, N., et al. (2001). "Deforestation modifying terrestrial organic transport in the Rio Tapajos, Brazilian Amazon." Organic Geochemistry 32(12): 1443-1458.
Farvolden, R. N. (1963). "Geologic Controls on Ground-water Storage and Base Flow." Journal of Hydrology 1: 219-249.
Fassnacht, S. R. (1996). "A multi-channel suspended sediment model for the MacKenzie Delta, Northwest Territories." Journal of Hydrology 197: 128-145.
Fausch, K. D. and T. G. Northcote (1992). "Large Woody Debris and Salmonid Habitat in a Small Coastal British Columbia Stream " Canadian Journal of Fisheries and Aquatic Sciences 49: 682-693.
Fausch, K. D., et al. (2002). "Landscapes to riverscapes: Bridging the gap between research and conservation of stream fishes." BioScience 52(6): 483-498.
Fausch, K. D., et al. (2002). "Landscapes to riverscapes: Bridging the gap between research and conservation of stream fishes." BioScience 52: 483-498.
Faustini, J. M. and J. A. Jones (2003). "Influence of large woody debris on channel morphology and dynamics in steep, boulder-rich mountain streams, western Cascades, Oregon." Geomorphology 51: 187-205.
This study used 20-year records of stream channel change and wood to test hypotheses about the long-term influence of large
woody debris (LWD) on channel morphology, channel stability, and sediment dynamics in a steep, boulder-rich mountain stream.
We compared two nearly adjacent reaches of third-order Mack Creek over the period 1978–1997 after virtually all wood was
removed from the channel of the lower reach in 1964. We assessed the long-term legacy of wood removal using repeated crosssection
surveys, streamflow data, LWD inventory data, and detailed mapping and longitudinal profile surveys. At each of 11
cross sections in the upper reach and 19 in the lower reach, we calculated areas of scour and fill in response to the two largest
floods in the record. We used quasi-likelihood logistic regression models to test the proportion of each reach that experienced
change between consecutive surveys over the entire record (1978–1997) as a function of flood return periods. The longitudinal
profile of the site without LWD was more variable than the reach with LWD at the finest scale (f1 m) due to a greater frequency
of boulder steps, but the reach with LWD was more variable at the channel unit scale. LWD-created steps 1 to 2.5 m high in the
wood-rich reach accounted for nearly 30% of the total channel fall and created low-gradient upstream channel segments one to
three channel widths long. As a result, both reaches have the same average slope (about 9%), but nearly three times as much of the
channel in the wood-rich reach had a slope of V5% as in the reach without wood (20.4% of total channel length vs. 7.5% of
channel length). The reach with abundant LWD was less responsive to moderate streamflow events (return period <f5 years),
but it responded similarly to peak flows with a return period of about 10 to 25 years. Although the average magnitude of crosssection
changes was the same during the largest flood in the record (25-year return period), the reach without LWD experienced
scour and coarsening of the bed surface, whereas the reach with LWD experienced aggradation upstream of LWD features. Mack
Creek may be representative of many steep mountain streams in which channel structure is strongly influenced by nonfluvial
processes: a legacy of large boulders from glacial or mass movement processes and a legacy of dead wood from ecological
processes. Sediment-limited mountain streams with large boulders, when deprived of LWD, appear to exhibit less morphological
variation at the channel unit scale, to store less sediment, and to release it more readily than those with LWD.
 
Favalli, M. and M. T. Pareschi (2004). "Digital elevation model construction from structured topographic data: The DEST algorithm." Journal of  Geophysical Research 109(F04004).
An algorithm, Determination of Earth Surface Structures (DEST), is presented to
reconstruct digital terrain models of complex landforms from topographic data, such as
contour lines and spot heights. The algorithm provides a triangular irregular network
(TIN) of the source data, based on a modified Delaunay approach. Delaunay triangulation
can introduce artificial terraces from a nonrandom distribution of input points such as a
sampled contour line. The algorithm proposed here constructs the three-dimensional
principal skeletons of these artificial flat areas, eliminating the unwanted effects of contour
lines. The algorithm can also be applied to topographic data from a variety of mixed
sources such as photogrammetric information, radar altimetry measurements, and
traditional contour lines. The sparse fine-surface structures present in the source data are
preserved, allowing accurate morphological evaluations, tectonic lineament extraction,
and volume estimation. A methodology (D-DEST) to easily derive, from a TIN computed
by DEST, the drainage path and the catchment areas is also presented. A comparison of
DEST with other methodologies is performed. It results that our approach does not
introduce sensible biased effects in slopes, aspects, drainage network, and catchment
areas. The evolution of the upper cone of Vesuvius volcano (Italy) during the last century,
as derived from historical cartography, is presented as an application of DEST. The
algorithm implemented in C can be requested at DEST_pareschi@pi.ingv.it.
 
Favis-Mortlock, D. (1998). "A self-organizing dynamic systems approach to the simulation of rill initiation and development on hillslopes." Computers & Geosciences 24(4): 353-372.
Whereas current erosion models are able to make quantitative estimates of rates of soil erosion by water on hillslopes with reasonable success, they are less competent when there is a need to apply a spatial dimension to their estimates. In particular, the initiation and development of rill networks is poorly modelled. Observations from field and laboratory suggest that it is possible to apply an evolutionary analogy to rill growth and development. Microrills, formed both by the runoff resulting from individual raindrops and from the overflow of ponded surface water through knickpoints, can be thought of as "competing". The most "successful" of these become discontinuous rills, which in their turn also compete, with a subset dominating to form continuous rills., Rill "success" will depend on both the position on the slope and microtopography. However, microtopography will itself be modified by erosional processes such as rill growth. This creates a feedback loop. Is it possible to use this perspective to model the evolution of hillslope rill networks? This study investigates the feasibility of such an approach. A novel model is described which applies simple rules to govern the iterative interaction between microtopography,: runoff routing and soil loss. Runoff is conceptualized as consisting of discrete "packets" so that a Lagrangian frame of reference is adopted, in contrast to the more usual Eulerian perspective. Results from several experiments are described, using both measured and synthetic microtopographic surfaces. The model appears able to reproduce many of the larger-scale "emergent" features of erosional systems. Planform rill networks appear realistic, as does rill depth, which increases both downslope and below confluences. Simulated rill spacing conforms with observed relationships to slope angle; the spatial balance between rill and interrill erosion similarly echoes observational evidence, as does total erosion and change in microtopographic roughness. (C) 1998 Published by Elsevier Science Ltd. All rights reserved.
 
Favis-Mortlock, D. T., et al. (2000). "Emergence and erosion: a model for rill initiation and development." Hydrological Processes 14(11-12): 2173-2205.
Soil erosion by overland flow, resulting from infiltration-excess rainfall, generates rill networks on hillslope areas. The way in which these networks emerge and develop suggests that hillslope erosion functions as a self-organizing dynamic system. Based upon this argument, a model for soil erosion (RillGrow 1) has been developed: this operates at the spatial scale of raindrops and microtopography. In this paper the second generation of the model (RillGrow 2) is described and applied to four different soil surfaces. Results suggest that, even at this early stage in its development, RillGrow 2 is capable of replicating the successes of the earlier model and in some cases of extending them. The success of both models suggests that this self-organizing view of rill generation may capture some fundamental aspects of the operation of real erosional systems. Copyright (C) 2000 John Wiley & Sons, Ltd.
 
Fedora, M. A. and R. L. Beschta (1989). "Storm runoff simulation using an Antecedent Precipitation Index (API) model." Journal of Hydrology 112: 121-133.
Feist, B. E., et al. (2003). "The influence of scale on salmon habitat restoration priorities." Animal Conservation 6: 271-282.
Habitat loss and alteration is the leading cause of species ’ declines world--wide,therefore habitat
restoration and protection is a prominent conservation strategy.Despite obvious connections between
habitat and threatened or endangered species,conservationists have been hard pressed explicitly to link
abundance or population health with habitat attributes.Given that habitat relationships with species are
often characterized at a spatial scale that does not account for the functional relationships between habitat
and populations,it is not surprising that the habitat –population conundrum persists.In order to explore
the in fluence of spatial scale on the apparent relationship between habitat and populations,we examined
the relationship between GIS-based habitat data and spring/summer chinook salmon (Oncorhynchus
tshawytscha )redd (spawning nests built by females)densities in the Salmon River basin,Idaho,at two
very different spatial scales:stream reach and watershed.Redd density was strongly correlated with
climate,geology,wetlands and terrain.However,our stream-reach scale models provided poor predictive
power compared with the watershed scale models.Based on these results,we conclude that our perception
of which habitat attributes were important was clearly a function of our scale of observation,and that
restoration efforts should focus on conditions at the watershed or landscape scale when attempting to do
local or reach scale restoration projects.
 
Feller, W. (1966). An Introduction to Probability Theory and its Applications. New York, John Wiley.
Fenn, C. R. and A. M. Gurnell (1987). Proglacial Channel Processes. Glacio-fluvial Sediment Transfer. A. M. Gurnell and M. J. Clark, John Wiley and Sons Ltd.: 423-495.
Ferguson, R. (1987). Hydraulic and sedimentary controls of channel pattern. River Channels: Environment and Process. K. Richards. New York, Basil Blackwell Inc.: 129-158.
Ferguson, R. (2005). "Estimating critical stream power for bedload transport calculations in gravel-bed rivers." Geomorphology 70(2005): 33-41.
Ferguson, R. I. (1986). "Hydraulics and hydraulic geometry." Progress  in Physical Geography 10(1): 1-31.
Ferguson, R. I. (1994). "Critical discharge for entrainment of poorly sorted gravel." Earth Surface Processes and Landforms 19: 179-186.
Ferguson, R. I. (1998). "Tracer-pebble movement along a concave river profile: Virtual velocity in relation to grain size and shear stress " Water Resources Research 34(8): 2031-2038.
Ferguson, R. I., et al. (2002). "Mobility of river tracer pebbles over different timescales." Water Resources Research 38(5).
Ferguson, R. I., et al. (2006). "River system discontinuities due to lateral inputs: generic styles and controls." Earth Surface Processes and Landforms 31: 1149-1166.
In alluvial river systems, lateral inputs of water and/or sediment at junctions or undercut
hillsides can disrupt what would otherwise be smooth downstream trends in mainstream
bed elevation, channel gradient, and bed grain size. Generic styles of mainstream res-ponse
to lateral inputs are investigated using a one-dimensional sediment routing model
with multiple grain size fractions. Numerical experiments isolate the effects of three para-meters:
ratio of tributary to mainstream water flux (QR), ratio of tributary to mainstream
bedload flux (FR), and ratio of tributary to mainstream bedload diameter (DR). The
findings are not unduly sensitive to the choice of initial conditions or to approximations
made in the model. The primary distinction is between junctions that aggrade, causing
local profile convexity with interrupted downstream fining, and junctions that degrade.
The immediate effects of aggradation extend further upstream than downstream, whereas
degradation is much more subdued and has no upstream impact. Aggradation is typical of
coarse inputs (DR >>>>>2), and degradation of fine inputs (DR <<<<<1), but very high ratios of QR
to FR also promote degradation. Both aggrading and degrading junctions can lead to a
change in mainstream bed grain size well below the junction, with higher ratios of QR to
FR producing a coarser distal bed. The effect of a tributary reflects the interplay between
additional bed load and additional discharge to transport it.
 
Ferguson, R. I. and T. B. Hoey (2002). "Long-term slowdown of river tracer pebbles: Generic models and implications for interpreting short-term tracer studies." Water Resources Research 38(8): doi:10.1029/2001WR000637, 002002.
Ferguson, R. I. and C. Paloa (1997). "Bias and precision of percentiles of bulk grain size distributions." Earth Surface Processes and Landforms 22: 1061-1077.
Percentiles such as D 50 and D 84 , calculated from weights retained on different sieves, are widely used to characterize grain size distributions (GSDs) of bulk samples of sedimentary deposits or sediment fluxes. The sampling variability of such percentiles is not well known, and few sampling guidelines exist for reliable characterization of GSDs. We report results from computer sampling experiments on the variability of sample percentiles in different-sized samples from populations with a log-normal GSD by weight and different sorting coefficients. Sample sizes are scaled by the volume of a median-sized grain so that results can be applied to any log-normal GSD. Sampling is random for the GSD by number that is equivalent to a specified GSD by weight. Results show important differences from standard sampling theory applicable to pebble-count GSDs. In small bulk samples all percentiles, including the median, are underestimated (more so for smaller samples, coarser percentiles and poorer sorting), and precision does not improve with the square root of sample size until fairly large sample sizes are exceeded. Non-dimensional equations fitted by eye to the results give good approximations to expected bias and precision in any percentile from 50 to 95 for any given sample size and population sorting coefficient. They are inverted to estimate the sample size required to avoid significant bias, or achieve specified precision, in any percentile of interest given estimates of the population D 50 and sorting coefficient. Target sample sizes are sometimes considerably smaller, but in other circumstances larger, than suggested by previous guidelines relating to estimation of the entire grain size distribution. Bias is likely in small samples of river bedload and good precision requires very large samples of poorly sorted gravel deposits.
 
Ferguson, R. I. and S. J. Wathen (1998). "Tracer-pebble movement along a concave river profile: Virtual velocity in relation to grain size and shear stress." Water Resources Research 34(8): 2031-2038.
Ferreira, A. J. D., et al. (1997). "Sediment and solute yield in forest ecosystems affected by fire and rip-ploughing techniques, central Portugal; a plot and catchment analysis approach." Physics and Chemistry of the Earth 22(3-4): 309-314.
Ferrier, K. L., et al. (2005). "Erosion rates over millennial and decadel timescales at Caspar Creek and Redwood Creek, Northern California Coast Ranges." Earth Surface Process and Landforms 30: 1025-1038.
Ferro, V. (1997). "Further remarks on a ditributed approach to sediment delivery." Hydrological Sciences Journal-Journal Des Sciences Hydrologiques 42(5): 633-647.
Ferro, V., et al. (2001). "Monitoring and predicting sediment yield in a small Sicilian basin." Transactions of the ASAE 44(3): 585-595.
Ferro, V. and M. Minacapilli (1995). "Sediment delivery processes at basin scale." Hydrological Sciences Journal/Journal des Sciences Hydrologiques 40(6): 703-718.
Since eroded sediments are produced from different sources distributed throughout a basin, sediment delivery processes at basin scale have to be modelled by a spatially distributed approach. In this paper a new theoretically based relationship is proposed for evaluating the sediment delivery ratio, SDR sub(i), of each morphological unit, i, into which a basin is divided. Then, using the sediment balance equation written for the basin outlet, a relationship between the basin sediment delivery ratio, SDR sub(w) and the SDR sub(i) is deduced. This relationship is shown to be independent of the soil erosion model used. Finally, a morphological criterion for estimating a coefficient, beta , is proposed.
 
Ferro, V. and P. Porto (2000). "Sediment Delivery Distributed (SEDD) Model."
Because eroded sediments are produced from different sources throughout a basin, it is often advantageous to model sediment delivery processes at basin scale using a spatially distributed approach. In this paper, a sediment delivery distributed (SEDD) model applicable at morphological unit scale, into which a basin is divided, is initially proposed. The model is based on the Universal Soil Loss Equation (USLE), in which different expressions of the erosivity and topographic factors are considered, coupled with a relationship for evaluating the sediment delivery ratio of each morphological unit. Then the SEDD model is calibrated by sediment yield and rainfall and runoff measurements carried out, at annual and event scales, in three small Calabrian experimental basis. At event scale, the analysis showed that a good agreement between measured and calculated basin sediment yields can be obtained using the simple rainfall erosivity factor; the agreement is independent of the selected equations for estimating the topographic factors. The analysis developed at annual scale showed that the model reliability increases from the event scale to the annual scale. Finally, a Monte Carlo technique was used for evaluating the effects of the uncertainty of the model parameters on calculated sediment yield.
 
Ferro, V., et al. (1998). "Testing a distributed approach for modelling sediment delivery." Hydrological Sciences Journal-Journal Des Sciences Hydrologiques 43(3): 425-442.
Fiksdal, A. J. and M. J. Brunengo (1981). Forest Slope Stability Project, Washington Department of Natural Resources: 1-61.
Finlay, P. J., et al. (1999). "Landslide risk assessment: prediction of travel distance." Canadian Geotechnical Journal 36: 556-562.
The Geotechnical Engineering Office of Hong Kong database of landslides for the period 1984–1993,
containing over 3000 landslide records, and data for major landslides before this period were made available for a
landslide risk assessment study. Data from over 1100 of these landslides in man-modified slopes were collated,
processed, and analyzed statistically. Multiple regression models, based on slope geometry, were developed for the
prediction of landslide travel on a horizontal surface below the slope. Although developed for slopes in Hong Kong,
the models can be applied in similar geological situations (weathered granite and other weathered volcanics) and are
suited for landslide risk zoning in areas below man-modified slopes and for the assessment of risks from individual
cuts and fills.
 
Finlayson, D. P., and D. R. Montgomery (2003). "Modeling large-scale fluvial erosion in geographic information systems." Geomorphology 53: 147-164.
Finnegan NJ, G. R., DR Montgomery, and B Hallet (2005). "Controls on the channel width of rivers: implications for modeling fluvial incision of bedrock." Geology 33(3): 229-232.
Finney, M. A. (1995). "The Missing Tail and Other Considerations for the Use of Fire History Models." Int. J. Wildland Fire 5(4): 197-202.
Fisher, R. V. (1971). "Features of Coarse-Grained, High Concentration Fluids and their Deposits." Journal of Sedimentary Petrology 41: 916-927.
Fisher, S. G. (1997). "Creativity, idea generation, and the functional morphology of streams." Journal of North American Benthological Society 16(2): 305-318.
Fisher, S. G. (2000). "Book Review of "River ecology and management: Lessons from the Pacific coastal ecoregion"." Limnol. Oceanogr. 45(6): 1445.
Fisher, S. G., et al. (2007). "Functional Ecomorphology: Feedbacks between Form and Function in Fluvial Landscape Ecosystems." Geomorphology 89: 84-96.
Fisher, S. G., et al. (2002). "Landscape challenges to ecosystem thinking: creative flood and drought in the American Southwest." Scientia Marina 65(suppl 2): 181-192.
Flanagan, D. C., J.S. Ascough, A. D. Nicks, M.A. Nearing, and J.M. Laflen (1995). "Ch.1 An overview of the WEPP model."
Flanagan, D. C., C.S. Renschler, and T.A. Cochrane (2000). Application of the WEPP model with digital geographic information. 4th International Conference  on Integrating GIS and Environmental Modeling: Problems, Prospects and Research Needs., Banff, Alberta, Canada, September 2-8, 2000.
Flanagan, D. C., and M.A. Nearing (2000). "Sediment particle sorting on hillslope profiles in the WEPP model." Transactions of the ASAE 43(3): 573-583.
Flanagan, D. C., J.R. Frankenburger, C.S. Renschler, J.M. Laflen, and B.A. Engel (2001). Simulating small watersheds with Water Erosion Prediction Project technology. Soil Erosion Research for the 21st Century, Honolulu, HI, ASAE.
Flanagan, S. A., et al. (1998). Methods for Inventory and Environmental Risk Assessment of Road Drainage Crossings, U.S. Department of Agriculture, Forest Service: Technology and Development Program: 45.
Flannigan, M., K. Logan, B. Amiro, W. Skinner, and B. Stocks (2005). "Future area burned in Canada." Climatic Change 72(1-2): 1-16.
Fleener, G. B. (1997). Boulder, University of Colorado.
Fleishman, E. and R. Mac Nally (2002). "Topographic Determinants of Faunal Nestedness in Great Basin Butterfly Assemblages: Applications to Conservation Planning." Conservation Biology 16(2): 422-429.
Fleming, C. A. (1973). The Quaternary Record of New Zealand and Australia Quaternary Studies. R. P. Suggate and M. M. Cresswell: 1-20.
Flenniken, M., et al. (2001). "Hydrologic responses of a montane riparian ecosystem following cattle use." Journal of Range Management 54(5): 567-574.
Riparian areas link streams with their terrestrial catchments and decrease water pollution by trapping sediments from upland sources before they reach streams or lakes. Livestock grazing in riparian areas is a controversial practice. If not properly managed, cattle can cause degradation to both the riparian zone and adjacent water body. Vegetative, soil microtopographical, microchannel and hydrograph parameters were measured in a montane riparian community in northern Colorado to quantify the effects of cattle on overland flow and runoff characteristics. Treatments were cattle grazing plus trampling. cattle trampling, mowing, and a control. Water was applied to plots (3 m x 10 in) at a rate of 100 mm hr(-1) using a rainfall simulator. Concurrently, overland flow was introduced at the upper end of the plots at an equivalent rate of 25 mm hr(-1). A high intensity-short duration grazing treatment was used for the cattle-treated plots. Reduction in vegetation stem density and aboveground biomass by cattle decreased microchannel sinuosity and drainage density. Cattle-treated plots had greater now velocities and depths in microchannels compared with mowed and control plots. Reduced stem density and aboveground biomass by grazing left fewer obstacles to divert flows, which decreased microchannel sinuosity and drainage density. Flows were concentrated into fewer microchannels with deeper flows. Microchannel characteristics were not significant factors affecting total runoff. Stem density and rainfall intensity were the most important factors in predicting runoff characteristics and total runoff. Results from this study have improved our understanding of now and runoff processes following cattle use of a riparian ecosystem.
 
Flint, R. F. (?). Overall View of Late-Cenozoic Climate and Glaciation Glacial and Quaternary Geology. New York, London, Sydney, Toronto, John Wiley and Sons, Inc.: 1-26.
Flitcroft, R. L. (2007). Regions to Streams: Spatial and Temporal Variation in Stream Occupancy Patterns of Coho Salmon (Oncorhynchus kisutch) on the Oregon Coast. Department of Fisheries and Wildlife. Corvalis, OR, Oregon State University. Ph.D.: 189.
Aquatic ecological investigation is expanding to encompass considerations of
multiple scales across large landscapes. Much of the analysis included in this work
focuses specifically on coho salmon (Oncorhynchus kisutch) in multiple subbasins on
the Oregon coast. Coho salmon were chosen for an investigation of spatial scales,
network connections, and life history stages due to their broad distribution on the
Oregon coast, and abundant data describing their distribution, habitat needs, behavior,
and survival. Chapter 2 introduces dynamic network topology (DNT) as a framework
for analysis and interpretation of aquatic obligate species. DNT is based on the
premise that in-stream habitats change in form and organization over time, and native
aquatic species are adapted to those changes through movement and life history
diversity. Chapter 3 analyzes juvenile coho salmon density and stream network
occupancy at three spatial scales (site, patch, and subbasin). The site scale analysis
indicated that combining network and traditional in-stream habitat metrics (i.e.,
substrate and habitat juxtaposition variables) are most effective at describing juvenile
coho salmon density. Patch sizes of juvenile coho salmon were defined using
variograms. Variogram shape indicated that a nested spatial structure may be present
in larger subbasins, indicating overlapping patterns of juvenile stream use. At the
subbasin scale, stream network occupancy by juvenile coho salmon was shown to vary
over time within subbasins, and appeared to increase or decrease similarly to the size
of the adult spawning run. In chapter 3, two-tier Bayesian hierarchical models were
applied to adult (subbasin and basin scales) and juvenile (site and subbasin scales)
coho salmon in an attempt to combine spatial scales that might be influential at each
life history stage. The best fitting adult model included the percent of large trees in the
riparian zone at the subbasin scale with mean annual precipitation at the basin scale.
The best fitting juvenile model included three variables, percent sand, stream order,
and network distance to spawning habitat which mirrors the result of modeling efforts
in Chapter 3. Multiple spatial scales and the framework of a stream network were
informative at detecting patterns and interactions among scales and life history stages
of coho salmon.
 
Floras, S. A., and I.D. Sgouras (1999). "Use of geoinformation techniques in identifying and mapping areas of erosion in a hilly landscape of central Greece." International Journal of Applied Earth Observation and Geoinformation 1(1): 68-77.
Flores, A. N., B.P. Bledsoe, C.O.Cuhciyan, and E.E. Wohl (2006). "Channel-reach morphology dependence on energy, scale, and hydroclimatic processes with implicaltions for prediction using geospatial data." Water Resources Research 42: WO6412, doi:6410.1029/2005WR004226.
Flores, A. N., et al. (2006). "Channel-reach morphology dependence on energy, scale, and hydroclimatic processes with implications for prediction using geospatial data." Water Resources Research 42: W06412.
Channel types found in mountain drainages occupy characteristic but intergrading
ranges of bed slope that reflect a dynamic balance between erosive energy and channel
boundary resistance. Using a classification and regression tree (CART) modeling
approach, we demonstrate that drainage area scaling of channel slopes provides better
discrimination of these forms than slope alone among supply- and capacity-limited sites.
Analysis of 270 stream reaches in the western United States exhibiting four common
mountain channel types reveals that these types exist within relatively discrete ranges of
an index of specific stream power. We also demonstrate associations among regional
interannual precipitation variability, discharge distribution skewness, and means of the
specific stream power index of step-pool channels. Finally, we discuss a conceptual
methodology for predicting ecologically relevant morphologic units from digital elevation
models at the network scale based on the finding that channel types do not exhibit equal
energy dissipation.
 
Florinsky, I. V. (1998). "Accuracy of local topographic variables derived from digital elevation models." Int. J. Geographical Information Science 12(1): 47-61.
We study the accuracy of data on some local topographic attributes
derived from digital elevation models (DEMs). First, we carry out a test for the
precision of four methods for calculation of partial derivatives of elevations. We
® nd that the Evans method is the most precision algorithm of this kind. Second,
we produce formulae for root mean square errors of four local topographic
variables (gradient, aspect, horizontal and vertical landsurface curvatures), provided
that these variables are evaluated with the Evans method. Third, we demonstrate
that mapping is the most convenient and pictorial way for the practical
implementation of the formulae derived. A DEM of a part of the Kursk Region
(Russia) is used as an example. We ® nd that high errors of data on local topographic
variables are typical for ¯ at areas. Results of the study can be used to
improve landscape investigations with digital terrain models.
 
 
Florsheim, J. L., E.A. Keller, and D.W. Best (1991). "Fluvial sediment transport in response to moderate storm flows following chaparral wildfire, Ventura County, southern California." Geological Society of America Bulletin 103(504-511).
Florsheim, J. L. (2004). "Side-valley tributary fans in high-energy river floodplain environments: sediment sources and depositional processes, Navarro River basin, California." Geological Society of America Bulletin 116(7/8): 923-937.
Side-valley tributary fans within the
main axial valley river fl oodplain system are
integral components of the sediment assemblage
characterizing fl oodplains in the highenergy
Navarro River basin, California.
Laboratory analyses of subsurface core and
auger samples from four study sites reveal
that sediment-size distributions within the
fl oodplain are controlled by spatial and
temporal variation in sediment sources
and depositional processes. Over 75% of
sediment within individual lower-fan strata
are poorly sorted and include gravel clasts
inferred to be contributions from the sidevalley
tributary, whereas remaining strata
are composed of fi ne sediment less than 2
mm without gravel. High-water marks indicate
that portions of the side-valley tributary
fans are episodically inundated by axial valley
overbank fl ow, a factor that infl uences
fan relief and profi le. Thus, two dominant
sources supply sediment to this transitional
environment: side-valley tributaries and
the axial valley river. The transport and
depositional processes that supply sediment
to the fans operate under three conditions:
(1) Tributary debris or fl uvial fl ows, and
selective entrainment of fi nes by fan surface
fl ows, deposit sediment when there is
no overbank fl ow from the axial channel;
(2) overbank fl ow deposits sediment derived
from the axial valley river when there is
no tributary input; and (3) tributary fl ows
deposit sediment while the fl oodplain is
inundated by fl oods from the axial valley
river. The mixing and interlayering of sediment
from these disparate sources within the
active fl oodplain illustrate the signifi cance of
both hillslope-fl oodplain and channel-fl oodplain
interactions in high-energy fl uvial
systems with fl oodplains.
 
Flugel, W. A. (2000). System-related development of regional hydrological model systems. Wasser and Boden 52(3): pp. 14-17;   2000, Blackwell Wissenschaft-Verlag GmbH;   Blackwell Verlag GmbH: Kurfurstendamm 57, D-10707 Berlin, Germany.
An understanding of process dynamics, hydrological systems analyses, an appropriate regionalization strategy for aerial distribution of a river basin, and a development toolkit are prerequisites for the design of hydrological modelling systems using an object oriented approach. These components and the concept of Response Units (RUs) applied and introduced within the Object Modelling System (OMS) are discussed. Rus are three dimensional modelling entities delineated within a GIS and are preserving the catchment's physiographic heterogeneity controlling its vertical and lateral dynamics of water and solute transport. Applied within OMS they represent distributed object classes and the downscaling regionalization is based on the transfer of object information from the Rus down to their single areas. The hydrological process dynamics is thus represented in OMS by process objects applied on the RUs and fluxes of water and solutes are routed from the RU-areas towards the receiving stream. Realised by OMS process modules the RU concept open up object oriented synergetic potential such as the programming language Java for hydrological modelling and the design of decision support system (DSS) for the sustainable management of scare and valuable water resources.
 
Folt, C. L., et al. (1998). "Implications of  temporal and spatial scale for Atlantic salmon (Salmo salar) research." Can. J. Fish. Aquat. Sci 55(Suppl. 1): 9-12.
Fonda, R. W. (1974). "Forest succession in relation to river terrace development in Olympic National Park, Washington." Ecology 55: 927-942.
Fontana, G. D. a. L. M. (2003). "Slope-area relationships and sediment dynamics in two alpine streams." Hydrological Processes 17: 73-87.
Foody, G. M., and P.M. Atkinson (2002). Uncertainty in remote sensing and GIS, John Wiley & Sons Ltd, Chicester, England.
Ford, E. D. (2000). Scientific Methods for Ecological Research. Cambridge (United Kingdom), Cambridge University Press.
Forest Practices Code of British Columbia (1996). Channel Assessment Procedure Guidebook Forest Practices Code of British Columbia: 37.
Forest Practices Code of British Columbia (1999). Coastal Watershed Assessment Procedure Guidebook (CWAP) Interior Watershed Assessment Procedure Guidebook (IWAP) Forest Practices Code of British Columbia: 41.
Forest Practices Code of British Columbia (1999). Mapping and Assessing Terrain Stability Guidebook Forest Practice Code of British Columbia: 36.
Forman, R. T. T. and M. Gordon (1986). Landscape Ecology. New York, John Wiley and Sons.
Foster, D. R., et al. (1998). "Landscape patterns and legacies resulting from large, infrequent forest disturbances." Ecosystems 1: 497-510.
We review and compare well-studied examples of
five large, infrequent disturbances (LIDs)—fire, hurricanes,
tornadoes, volcanic eruptions, and
floods—in terms of the physical processes involved,
the damage patterns they create in forested landscapes,
and the potential impacts of those patterns
on subsequent forest development. Our examples
include the 1988 Yellowstone fires, the 1938 New
England hurricane, the 1985 Tionesta tornado, the
1980 eruption of Mount St. Helens, and the 1993
Mississippi floods. The resulting landscape patterns
are strongly controlled by interactions between the
specific disturbance, the abiotic environment (especially
topography), and the composition and structure
of the vegetation at the time of the disturbance.
The very different natures of these interactions yield
distinctive temporal and spatial patterns and demand
that ecologists increase their knowledge of
the physical characteristics of disturbance processes.
Floods and fires can occur over a long period,
whereas volcanic eruptions and wind-driven events
often last for no more than a few hours or days.
Tornadoes and floods produce linear patterns with
sharp edges, but fires, volcanic eruptions, and hurricanes
can affect broader areas, often with gradual
transitions of disturbance intensity. In all cases, the
evidence suggests that LIDs produce enduring legacies
of physical and biological structure that influence
ecosystem processes for decades or centuries.
 
Foster, G. R., T.E. Toy, and K.G. Renard (2003). Comparison of the USLE, RUSLE1.06 and RUSLE2 for applications to highly disturbed lands. First interagency Conference on Research in Watersheds, U.S. Department of Agriculture, Agricultural Research Service, Washington, D.C.
Fox, M. and S. Bolton (2007). "A regional and geomorphic reference for quantities and volumes of instream wood in unmanaged basins in Washington State." Journal of Fisheries Management, American Fisheries Society 27: 342-359.
Fox, M. and S. Bolton (2007). "A regional and geomorphic reference for quantities and volumes of instream wood in unmanaged forested basins of Washington state." North American Journal of Fisheries Management 27: 342-359.
We collected field data on instream wood quantities and volumes from 150 stream segments
draining unmanaged basins within Washington State to develop reference conditions for restoration and
management. The wood loads in these streams provide a reference for management since it is assumed that
they incorporate the range of conditions to which salmonids and other species have adapted. We also used
these data to evaluate existing standards for large wood in streams. Large wood is an important component of
salmonid habitat, and stream channel assessments and restoration and enhancement efforts often associate
habitat quality for salmon Oncorhynchus spp. with the quantity and volume of woody debris; however, the
wood targets currently used to assist resource managers typically do not account for variations in quantity or
volume owing to differences in geomorphology, forest zones, or disturbance regimes. For restoring the
appropriate range of conditions in salmon habitat, we offer a percentile wood distribution of natural and
unmanaged wood-loading ranges based on regional and geomorphic variation for the purpose of
reestablishing central tendencies. We recommend that streams in a degraded state (e.g., below the 25th
percentile) be managed for an interim target at or above the 75th percentile until the basin-scale wood loads
achieve these central tendencies. Based on the sample distribution, these reference conditions are applicable to
streams with bank-full widths between 1 and 100 m, gradients between 0.1% and 47%, elevations between 91
and 1,906 m, drainage areas between 0.4 and 325 km2, glacial and rain- or snow-dominated origins, forest
types common to the Pacific Northwest, and several other distinguishing physical and regional classifications.
 
Fraidenburg, M. E. (1989). "The New Politics of Natural Resources: Negotiating a Shift Toward Privatization of Natural Resource Policymaking in Washington State." The Northwest Environmental Journal 5: 211-240.
Franchini, M., Galeati, G., and Berra, S. (1998). "Global optimization techniques for the calibration of conceptual rainfall-runoff models." Hydrological Sciences Journal/Journal des Sciences Hydrologiques 43(3): 443-458.
Franklin, J. F. (1979). Vegetation of the Douglas fir region. Forest soils of the Douglas fir region, Pullman, WA, Washington State University Cooperative Extension.
Franklin, J. F. and C. T. Dyrness (1973). Natural vegetation of Oregon and Washington, Pacific Northwest Forest Range Experimental Station: 417.
Franklin, J. F. and C. T. Dyrness (1988). Natural Vegetation of Oregon and Washington. Corvallis, Oregon, Oregon State University Press.
Franklin, J. F. and R. T. T. Forman (1987). "Creating landscape patterns by forest cutting: Ecological consequences and principles." Landscape Ecology 1(1): 5-18.
Franklin, J. F., et al. (1987). "Tree Death as an Ecological Process." BioScience 37(8): 550-556.
Franklin, J. F. and T. A. Spies (?). Composition, function, and structure of Old-Growth Douglas-Fir Forests, U.S. Forest Service: 71-80.
Franks, C. A., et al. (2002). Hydraulic habitat in confluences: an ecological perspective on confluence hydraulics. The Structure, Function and Management Implications of Fluvial Sedimentary Systems. F. J. Dyer, M. C. Thoms and J. M. Olley, International Association of Hydrological Sciences. 276: 61-67.
The near-bed hydraulic characteristics of a confluence were monitored over a range of stages to assess the variability of hydraulic habitat and its influence over distributions of benthic macroinvertebrate communities. Spatial variability of hydraulic characteristics was observed in the confluence, and there is some suggestion from macroinvertebrate data that the hydraulic conditions influence the distribution of macroinvertebrates in the confluence
 
Fransen, B. R., et al. (2006). "A logistic regression model for predicting the upstream extent of fish occurrence based on geographical information systems data." North American Journal of Fisheries Management 26: 960-975.
Regulations governing human activities in streams and riparian zones frequently differ
depending on whether or not a stream reach supports fish. Fish presence or absence is usually determined by
sampling or by assuming the presence of fish if the stream exhibits certain physical characteristics. Field
surveys of fish occurrence in streams are time consuming and expensive. Inference of fish presence from
simple thresholds of physical attributes, such as gradient or channel width alone, is inaccurate. We attempted
to improve the accuracy and efficiency of this determination by developing a geographical information
systems (GIS)-based predictive model. A 10-m digital elevation model incorporated field data on fish
distribution from 517 streams in western Washington State and GIS-derived representations of the physical
characteristics of stream networks. A model predicting the upstream extent of fish occurrence was derived
using logistic regression models coupled with a heuristic ‘‘stopping rule.’’ Candidate variables included stream
gradient, upstream basin area, elevation, and mean annual precipitation. When assessed against independent
survey data, 91.9% of the occupied fish habitat was correctly classified by the model. Errors were generally
small, but occasional large errors did occur and were most frequently associated with barriers to fish
movement. Smaller errors occurred in marginal habitats, streams of low topographic relief, and streams that
originated from headwater ponds. Use of this type of model, coupled with targeted field survey in areas most
likely to be associated with model error, would greatly improve the efficiency and accuracy of current
classification schemes.
 
Franti, T. G., et al. (1996). "Modeling the effects of incorporated residue on rill erosion .1. Model development and sensitivity analysis." Transactions of the ASAE 39(2): 535-542.
Frazer, L. N. and M. K. Sen (1980). "Kirchhoff-Helmholtz reflection seismograms in a laterally inhomogeneous multi-layered elastic medium - I. Theory." Geophys. J.R. astr. Soc. 80: 121-147.
Fread, D. L. (1988). Breach: An Erosion Model for Earthen Dam Failures (DRAFT). Silver Spring, Maryland, National Weather Service: 1-29.
Fread, D. L. (?). "Dam-Breach Erosion Modeling."
Fredlund, D. G. and J. Krahn (1977). "Comparison of slope stability methods of analysis." Canadian Geotechnical Journal 14: 429-439.
Fredriksen, F. L. (1970). Erosion and Sedimentation Following Road Construction and Timber Harvest on Unstable Soils in three small Western Oregon Watersheds. Portland, Oregon, U.S. Foreest Service: 1-15.
Fredriksen, R. L. (1970). Erosion and sedimentation following road construction and timber harvest on unstable soils in three small western Oregon watersheds, U.S. Department of Agriculture
Pacific Northwest Forest and Range Experiment Station Forest Service: 1-15.
Freeman, M. C., et al. (2007). "Hydrologic connectivity and the contribution of stream headwaters to ecological integrity at regional scales." Journal of the American Water Resources Association 43(1): 5-14.
Cumulatively, headwater streams contribute to maintaining hydrologic connectivity and ecosystem
integrity at regional scales. Hydrologic connectivity is the water-mediated transport of matter, energy and
organisms within or between elements of the hydrologic cycle. Headwater streams compose over two-thirds of
total stream length in a typical river drainage and directly connect the upland and riparian landscape to the
rest of the stream ecosystem. Altering headwater streams, e.g., by channelization, diversion through pipes,
impoundment and burial, modifies fluxes between uplands and downstream river segments and eliminates distinctive
habitats. The large-scale ecological effects of altering headwaters are amplified by land uses that alter
runoff and nutrient loads to streams, and by widespread dam construction on larger rivers (which frequently
leaves free-flowing upstream portions of river systems essential to sustaining aquatic biodiversity). We discuss
three examples of large-scale consequences of cumulative headwater alteration. Downstream eutrophication and
coastal hypoxia result, in part, from agricultural practices that alter headwaters and wetlands while increasing
nutrient runoff. Extensive headwater alteration is also expected to lower secondary productivity of river systems
by reducing stream-system length and trophic subsidies to downstream river segments, affecting aquatic communities
and terrestrial wildlife that utilize aquatic resources. Reduced viability of freshwater biota may occur
with cumulative headwater alteration, including for species that occupy a range of stream sizes but for which
headwater streams diversify the network of interconnected populations or enhance survival for particular life
stages. Developing a more predictive understanding of ecological patterns that may emerge on regional scales as
a result of headwater alterations will require studies focused on components and pathways that connect headwaters
to river, coastal and terrestrial ecosystems. Linkages between headwaters and downstream ecosystems
cannot be discounted when addressing large-scale issues such as hypoxia in the Gulf of Mexico and global losses
of biodiversity.
 
 
Freer, J., et al. (2002). "The role of bedrock topography on subsurface storm flow." Water Resources Research 38(12): doi:10.1029/2001WR000872, 002002.
Freese, F. (1960). "Testing Accuracy." Forest Science 6(2): 139-145.
Freeze, R. A., and R.L. Harlan (1969). "Blueprint for a physically based, digitally simulated hydrologic response model." Journal of Hydrology 9: 237-258.
Freeze, R. A. (1982). Hydrogeological Concepts in Stochiastic and Deterministic Rainfall-Runoff Predictions. Vancouver, British Columbia, University of British Columbia: 63-79.
Freeze, R. A. (1987). Modelling Interrelationships Between Climate, Hydrology, and Hydrogeology and the Development of Slopes. Slope Stability. M. G. Anderson and K. S. Richards. Vancouver, Canada, John Wiley & Sons Ltd.: 381-403.
Friedman, J. H. (2002). "Stochastic gradient boosting." Computional Statistics and Data Analysis 38(4): 367-378.
Gradient boosting constructs additive regression models by sequentially fitting a simple parameterized function (base learner) to current “pseudo”-residuals by least squares at each iteration. The pseudo-residuals are the gradient of the loss functional being minimized, with respect to the model values at each training data point evaluated at the current step. It is shown that both the approximation accuracy and execution speed of gradient boosting can be substantially improved by incorporating randomization into the procedure. Specifically, at each iteration a subsample of the training data is drawn at random (without replacement) from the full training data set. This randomly selected subsample is then used in place of the full sample to fit the base learner and compute the model update for the current iteration. This randomized approach also increases robustness against overcapacity of the base learner.
 
Friedman, J. M. and G. T. Auble (1999). "Mortality of riparian box elder from sediment mobilization and extended inundation." Regulated Rivers-Research & Management 15(5): 463-476.
To explore how high flows limit the streamward extent of riparian vegetation we quantified the effects of sediment mobilization and extended inundation on box elder (Acer negundo) saplings along the cobble-bed Gunnison River in Black Canyon of the Gunnison National Monument, Colorado, USA. We counted and aged box elders in 144 plots of 37.2 m(2), and combined a hydraulic model with the hydrologic record to determine the maximum shear stress and number of growing-season days inundated for each plot in each year of the record. We quantified the effects of the two mortality factors by calculating the extreme values survived during the lifetime of trees sampled in 1994 and by recounting box elders in the plots following a high flow in 1995. Both mortality factors can be modeled as threshold functions; box elders are killed either by inundation for more than 85 days during the growing season or by shear stress that exceeds the critical value for mobilization of the underlying sediment particles. Construction of upstream reservoirs in the 1960s and 1970s reduced the proportion of the canyon bottom annually cleared of box elders by high flows. Furthermore, because the dams decreased the magnitude of high flows more than their duration, flow regulation has decreased the importance of sediment mobilization relative to extended inundation. We use the threshold functions and cross-section data to develop a response surface predicting the proportion of the canyon bottom cleared at any combination of flow magnitude and duration. This response surface allows vegetation removal to be incorporated into quantitative multi-objective water management decisions. Copyright (C) 1999 John Wiley & Sons, Ltd.
 
Frissell, C. A., W.J. Liss, C.E. Warren, and M.D. Hurley (1986). "A hierarchical framework for stream habitat classification: viewing streams in a watershed context." Environmental Management 10(2): 199-214.
Frissell, C. A. and D. Bayles (1996). "Ecosystem Management and the Conservation of Aquatic Biodiversity and Ecological Integrity." Water Resources Bulletin 32(2): 229-240.
Frissell, C. A., et al. (1986). "A Hierarchial Framework for Stream Habitat Classification: Viewing Streams in a Watershed Context " Enivironmental Management 10(2): 199-214.
Frissell, C. A., et al. (1986). "A hierarchical framework for stream habitat classification: Viewing streams in a watershed context." Environmental Management 10: 199-214.
Frissell, C. A. and R. K. Nawa (1992). "Incidence and Causes of Physical Failure of Artificial Habitat Structures in Streams of Western Oregon and Washington " North American Journal of Fisheries Management 12: 182-197.
Fritts, H. C. (1981). Statistical climatic reconstructions from tree-ring widths. Climatic variations and variability: facts and theories, D. Reidel Publishing Company: 135-153.
Fritz, K., M., et al. (2006). Field Operations Manual for Assessing the Hydrologic Permanence and Ecological Condition of Headwater Streams. EPA/600/R-06/126. Washington DC, US Environmental Protection Agency, Office of Research and Development.
Fritz, K., M., et al. (2008). "Physical indicators of hydrologic permanence in forested headwater streams." Journal of the American Benthological Society 27(3): 690-704.
Recent court cases have questioned whether all headwater streams are jurisdictional waters
under the US Clean Water Act. Rapid field-based indicators of hydrologic permanence are needed for
making jurisdictional determinations. Our study objectives were to: 1) identify physical characteristics of
forested headwater streams that best distinguish perennial, intermittent, and ephemeral reaches and 2)
assess the applicability of existing rapid field-based tools for classifying hydrologic permanence across a
wide geographic range. We surveyed reach- and drainage-scale characteristics at 113 sites across 10 study
forests in the US. Streams in 4 core forests (61 core sites) were sampled over 2 consecutive years and were
used in model construction. Streams in 6 satellite forests (72 satellite sites) were used to validate the models
over a broader geographic range. Discriminant function models successfully differentiated hydrologic
permanence categories at core sites. Drainage area, the Ohio Environmental Protection Agency Headwater
Habitat Evaluation Index (HHEI), and the North Carolina Department of Water Quality Stream
Classification Method (NCSC) were strongly correlated with the discriminant function that separated
ephemeral from perennial and intermittent sites. Entrenchment ratio was the most consistent variable
discriminating intermittent from perennial sites across the core forests. The models had mixed results when
applied to the validation data set, but did classify correctly most intermittent and ephemeral sites.
Classification trees were used to assess broad regional applicability of existing rapid field-based protocols
and to identify important metrics. Scores from the Rapid Bioassessment Protocol Habitat Assessment, HHEI,
and NCSC all clearly distinguished ephemeral from intermittent and perennial sites, but no differences were
detected between intermittent and perennial sites across all sites. However, data from core sites do indicate
that a suite of physical variables can be used successfully to identify hydrologic permanence at regional
scales.
 
 
Froehlich, H. A. (?). Accumulation of Large Debris in Forest Streams Before and After Logging (DRAFT), Oregon State University: 1-8.
Front Range Round Table (2006). Living with Fire: protecting communities and restoring forests.
Frothingham, K., M. (2002). "A Multiscale Conceptual Framework for Integrated Ecogeomorphological Research to Support Stream Naturalization in the Agricultural Midwest." Environmental Management 29(1): 16-33.
Frye, J. C. (1973). "Pleistocene Succession of the Central Interior United States." Quaternary Research 3: 275-283.
Fryxell, F. M. and L. Horberg (1943). "Alpine Mudflows in Grand Teton National Park, Wyoming." Bulletin of the Geological Society of America 54: 457-472.
Fukushima, M. (2001). "Salmonid habitat -- geomorphology relationships in low-gradient streams." Ecology 82(5): 1238-1246.
A link between stream geomorphology and lotic ecosystems was demon-strated
by quantitatively examining the precise locations of salmonid redds with respect to
the planform geometry of streams using a differential global positioning system. A total
stream distance of 59 km was surveyed in 17 streams, in which a total of 309 redds of
Sakhalin taimen (Hucho perryi) were recorded. The average size (6SD) of these redds was
227 6 60 cm in length and 122 6 42 cm in width. A meta-analysis of these data showed
that channel sinuosity was significantly greater at sites where Sakhalin taimen redds were
constructed than the average stream sinuosity. This salmonid preference for highly sinuous
reaches was detected when the sinuosity index was calculated at 50-m increments and
became insignificant at greater distance increments. This habitat–sinuosity relationship will
be more pronounced in streams with only moderately sinuous channels, less abundant large
woody debris, and higher spawner densities.
 
Fukushima, M., et al. (1998). "Estimation of eggs lost from superimposed pink salmon (Oncorhynchus gorbushcha) redds." Canadian Journal of Fisheries and Aquatic Sciences 55: 618-625.
Fuller, M. (1991). Forest Fires. New York, Wiley.
Fullerton, A., et al. (in press). "Hydrologic connectivity for riverine fishes: measurement challenges and research opportunites." Freshwater Biology.
Fullerton, A. H., et al. (2006). "Regional patterns of riparian characteristics in the interior Columbia River basin, Northwestern USA: applications for restoration planning." Landscape Ecology 21: 1347-1360.
Recent declines in anadromous Pacific
salmonids (Oncorhynchus spp.) have been attributed,
in part, to degradation of freshwater habitat.
Because riparian areas directly affect instream habitat,
assessing riparian characteristics is essential for
predicting salmon habitat quality and for prioritizing
restoration projects. We quantified land use modification
of anadromous fish-bearing streams in the
interior Columbia River basin at multiple resolutions.
We identified riparian areas in several land
use and land cover classes using remotely sensed
data. We then interpreted aerial photographs at
random locations within each class to quantify
riparian modifications at a local (stream reach)
scale. Riparian areas in agricultural and urban areas
were significantly narrower ( 30 m, median) than
those in forested or shrub/grass areas ( 70 m). The
largest proportion of modified riparian areas occurred
in low-gradient streams with floodplains in
semi-arid ecoregions. Riparian vegetation in these
areas is unlikely to provide adequate in-stream
functions, making these areas a natural starting point
for restoration prioritization. We investigated how
existing riparian restoration projects were spatially
related to riparian land use and found that restoration
effort varied among subwatersheds. Effective
strategies for restoring high quality salmon habitat
will be watershed-specific and must restore natural
watershed processes. By using a hierarchical analysis
to identify regional strategies, restoration or
conservation activity can be focused in specific basins
and thereby increase the likelihood that efforts
will significantly improve habitat conditions for
listed salmonids.
 
Fullerton, A. H., et al. (2008). "How certain are salmon recovery forecasts? A watershed-scale sensitivity analysis." Environmental Modelling and Assessment.
Complex relationships between landscape and
aquatic habitat conditions and salmon (Oncorhynchus spp.)
populations make science-based management decisions
both difficult and essential. Due to a paucity of empirical
data, models characterizing these relationships are often
used to forecast future conditions. We evaluated uncertainties
in a suite of models that predict possible future habitat
conditions and fish responses in the Lewis River Basin,
Washington, USA. We evaluated sensitivities of predictions
to uncertainty in model parameters. Results were sensitive
to 60% of model parameters but substantially so (|partial
regression coefficients| >0.5) to <10%. We also estimated
accuracy of several predictions using field surveys. Observations
mostly fell within predicted ranges for riparian
shade and fine-sediment deposition, but large woody debris
estimates matched only half the time. We provide suggestions
to modelers for improving model accountability, and
describe how managers can incorporate prediction uncertainty
into decision-making, thereby improving the odds of
successful salmon habitat recovery
 
 
Fuquay, D. M. (1980). Forecasting Lighting Activity Level and Associated Weather. Ogden, Utah, US Forest Service: 30.
Furbish, D. J. (1978). The Stochiastic Structure of a High Mountain Stream. Geological Sciences, University of Colorado: 245.
Furbish, D. J. and R. M. Rice (1983). "Predicting landslides related to clearcut logging, northwestern California, U.S.A." Mountain Research and Development 3(3): 253-259.
Landslides related to clearcut logging are a significant source of erosion in the mountains of northwestern California. Forest managers, therefore, frequently must include assessments of landslide risk in their land-use plans. A quantitative method is needed to predict such risk over large areas of rugged mountainous terrain. From air photographs, data were collected of conditions associated with a sample of logging-related slides and randomly located stable sites. Discriminant analyses were used to develop an equation that distinguishes the two types of sites-slide and non-slide-with 81 percent accuracy. The equation can be used to provide an assessment of risk for undisturbed terrain. Results showed that post-logging failure is most likely to occur near actively scouring streams, just below major convex breaks of slope and within drainage depressions.
 
Furniss, M. J., et al. (1998). Response of Road-Stream Crossings to Large Flood Events in Washington, Oregon, and Northern California. San Dimas, California, USDA Forest Service: 14.
Gabet, E. J., and T. Dunne (2002). "Landslides on coastal sage-scrub and grssland hillslopes in a severe El Nino winter: The effects of vegetation conversion on sediment delivery." GSA Bulletin 114: 983-990.
Gabet, E. J. (2003). "Post-fire thin debris flows: sediment transport and numerical modelling." Earth Surface Processes and Landforms 28: 1341-1348.
Gabet, E. J. (2003). "Sediment transport by dry ravel." Journal of Geophysical Research 108(No.B1).
Gabet, E. J. and A. Bookter (2008). "A morphometric analysis of gullies scoured by post-fire progressively bulked debris flows in soutwest Montana, USA." Geomorphology 96: 298-309.
In the fall of 2001, an intense thunderstorm in southwest Montana triggered many debris flows in the burned area of Sleeping
Child Creek. In most instances, the debris flows cut deep gullies into previously unchannelized colluvial hollows and deposited
large volumes of sediment onto the valley floor. The presence of rill networks above the gullies as well as the absence of landslide
features indicate that the gullies were scoured by progressively bulked debris flows, a process in which dilute surface runoff
becomes increasingly more laden with sediment until it transforms into a debris flow. In this contribution, we present a
morphometric analysis of six of the gullies to better understand this relatively understudied process. We find that the locations of
the rill heads and gully heads conform to slope-area thresholds that are characteristic of erosion by overland flow. Our data also
suggest that the volumes of the debris flows increase exponentially with normalized drainage area, thus lending support to an
assumption used in a recently proposed debris flow incision law. Finally, the debris flow fans have been relatively unaltered since
deposition, suggesting that the valley may be currently aggrading while the hillslopes are being denuded.
 
Gabet, E. J. and T. Dunne (2002). "Landslides on coastal sage-scrub and grassland hillslopes in a severe El Niño winter: The effects of vegetation conversion on sediment delivery." Geological Society of America Bulletin 114(8): 983-990.
During the 1997–1998 El Niño, record rainfall triggered >150 shallow landslides within a 9.5 km2 area near Santa Barbara, California. They were studied to analyze the sediment delivery to valley floors from landslides in coastal sage scrub and converted grasslands. The conversion of coastal sage to grasslands, primarily to provide pasturage for cattle, is common in the region, and the landscape's response may affect water quality, reservoir infilling, and debris flow hazards. We explore the relationship between lateral- root reinforcement and landslide volume by developing a slope-stability analysis that incorporates root cohesion along the sides of the failure. The stability analysis correctly predicts an inverse relationship between landslide volume and hillslope angle in the sage. The volumes of failures in the grasslands do not vary systematically with slope and are generally smaller than those in the sage. From aerial-photograph analysis and field mapping, we find that there are 22.9 failures per square kilometer in the grasslands compared to 13.2 failures per square kilometer in the sage. Despite the lower failure density in the coastal sage, greater failure volumes and longer transport distances delivered more sediment to valley floors, with a specific volumetric flux of 2.8 × 10–2 m3·m–1 for this El Niño compared to 1.7 × 10–2 m3·m–1 in the grasslands. We conclude that the conversion from vegetation with stronger and deeper roots (coastal sage) to vegetation with weaker and shallower roots (grass) has caused a pulse of increased landsliding in the grasslands because the soils are currently too thick for the prevailing root reinforcement. We suggest that, over time, soils in the grassland hollows will become thinner as the evacuation by landslides is repeated until the landsliding rate declines to balance the soil supplied from local colluvium production and diffusive processes upslope.
 
Gabet, E. J. and T. Dunne (2003). "A stochastic sediment delivery model for a steep Mediterranean landscape." Water  Resources Research 39(9): 1237-1248.
Gabet, E. J. and P. Sternberg (2008). "The effects of vegetative ash on infiltration capacity, sediment transport, and the generation of progressively bulked debris flows." Geomorphology 101: 666-673.
Through the alteration of the physical characteristics of a landscape, such as the destruction of vegetation and
the formation of a hydrophobic layer, a fire can dramatically amplify erosion rates. On the basis of field
observations, it has been proposed that the deposition of a layer of ash on the ground surface can enhance the
erosion of mountainous terrain by surface runoff and might even be a necessary condition for the generation of
progressively bulked debris flows. In this study, a flume was constructed to investigate the role of ash in
increasing both the volume and the transport capacity of runoff. The experiments demonstrated that the
presence of ash on the soil surface reduces the ability of flowing water to infiltrate; this effect is even greater
when the ash has been pre-wetted. In addition, the ability of ash slurries to infiltrate decreases with increasing
ash concentration. The results also indicate that the transport capacity of runoff is enhanced by the
incorporation of ash into the flow because of the increased fluid density. However, the addition of ash reduces
the boundary Reynolds number such that, at high ash concentrations and with fine-grained sediment,
sediment transport declines as the flow becomes hydraulically smooth. The experimental results were also
used to evaluate the ability of steep flow fronts, a common characteristic of debris flows and flash floods, to
increase sediment transport rates. Finally, it is proposed that ash slurries may evolve into progressively bulked
debris flows through a positive feedback between fluid density, transport capacity, and erosivity.
 
Gabris, et al. (2003). "Land use change and gully formation over the last 200 years in a hilly catchment." Catena 50(2-4): 151-164.
Gacek Associates (1990). Smith Creek Timber Harvest Plan. Whatcom County, Washington, prepared by Gacek Associates for The Trillium Corporation: 96.
Galay, V. J., et al. (1987). River Bed Scour and Construction of Stone Riprap Protection Sediment Transport Gravel-Bed Rivers. C. R. Thorne, J. C. Bathurst and R. D. Hey, John Wiley and Sons Ltd. Chapter 12.
Gallagher, M. R., and J. Doherty (2007). "Parameter interdependence and uncertainty induced by lumping in a hydrologic model." Water Resources Research 43(W05421, doi:10.1029/2006WR005347, 2007).
Gallant, J. C. and J. P. Wilson (1996). "Tapes-G: A grid-based terrain analysis program for the environmental sciences." Computers & Geosciences 22(7): 713-722.
Gallino, G. L. and T. C. Pierson (1984). The 1980 Polallie Creek Debris Flow and Subsequent Dam-break Flood, East Fork Hood River Basin, Oregon. Portland, Oregon, U.S. Geological Survey: 11-36.
Galster, R. W. (? (No year needed)). "Dams of Western Washington: Introduction - Early Projects." Engineering Geology 1: 165-172.
Galster, R. W. (? (No year needed)). "Erosion." Engineering Geology 1: 60-90.
Gangodagamage, C., et al. (2007). "Scaling in river corridor widths depicts organization in valley morphology." Geomorphology 91: 198-215.
Landscapes have been shown to exhibit numerous scaling laws from Horton's laws to more sophisticated scaling in topography
heights, river network topology and power laws in several geomorphic attributes. In this paper, we propose a different way of examining
landscape organization by introducing the “river corridor width” (lateral distance from the centerline of the river to the left and right
valley walls at a fixed height above the water surface) as one moves downstream. We establish that the river corridor width series,
extracted from 1 m LIDAR topography of a mountainous river, exhibit a rich multiscale statistical structure (anomalous scaling) which
varies distinctly across physical boundaries, e.g., bedrock versus alluvial valleys.We postulate that such an analysis, in conjunction with
field observations and physical modeling, has the potential to quantitatively relate mechanistic laws of valley formation to the statistical
signature that underlying processes leave on the landscape. Such relations can be useful in guiding field work (by identifying physically
distinct regimes from statistically distinct regimes) and advancing process understanding and hypothesis testing
 
Gangodagamage, C., et al. (2011). "Revisiting scaling laws in river basins: New considerations across hillslope and fluvial regimes." Water Resources Research 47(W07508): 12.
Increasing availability of high‐resolution (1 m) topography data and enhanced
computational processing power present new opportunities to study landscape
organization at a detail not possible before. Here we propose the use of “directed distance
from the divide” as the scale parameter (instead of Horton’s stream order or upstream
contributing area) for performing detailed probabilistic analysis of landscapes over a broad
range of scales. This scale parameter offers several advantages for applications in
hydrology, geomorphology, and ecology in that it can be directly related to length‐scale
dependent processes, it can be applied seamlessly across the hillslope and fluvial regimes,
and it is a continuous parameter allowing accurate statistical characterization (higher‐order
statistical moments) across scales. Application of this scaling formalism to three basins
in California demonstrates the emergence of three distinct geomorphic regimes of
divergent, highly convergent, and moderately convergent fluvial pathways, with notable
differences in their scaling relationships and in the variability, or spatial heterogeneity,
of topographic attributes in each regime. We show that topographic attributes, such as
slopes and curvatures, conditional on directed distance from the divide exhibit less
variability than those same attributes conditional on upstream contributing area,
thus affording a sharper identification of regime transitions and increased accuracy in
the scaling analysis.
 
Gangodagamage, C., et al. (2005). "River corridor geometry: Scaling relationships and confluence controls." EOS Transactions, Fall Meeting Suppliment 82(52): Abstract H33E-1421.
River corridor (valley) geometry is a topographic signature of tectonic and fluvial processes in a watershed. It is significantly affected by the tributary basins at confluences and its geometry influences the vegetation pattern and riverine habitat heterogeneity along the river reaches. An algorithm was developed to extract from LIDAR data the valley geometry and vegetation height along the whole river network as a function of distance downstream of the confluence, distance from the stream, and depth from the water surface. Based on those data and field observations from the Angelo Coast Reserve in the Eel river basin, CA, we propose metrics that quantify the "geomorphologic significance" of a tributary and we explore scaling relationships in valley geometry, riparian vegetation, and in the effect of confluences on the downstream channel morphometry.
 
Ganio, L. M., et al. (2005). "A geostatistical approach for describing spatial pattern in stream networks." Frontiers in Ecology and the Environment 3(3): 138-144.
The shape and configuration of branched networks influence ecological patterns and processes. Recent investigations
of network influences in riverine ecology stress the need to quantify spatial structure not only in a
two-dimensional plane, but also in networks. An initial step in understanding data from stream networks is
discerning non-random patterns along the network. On the other hand, data collected in the network may be
spatially autocorrelated and thus not suitable for traditional statistical analyses. Here we provide a method
that uses commercially available software to construct an empirical variogram to describe spatial pattern in
the relative abundance of coastal cutthroat trout in headwater stream networks. We describe the mathematical
and practical considerations involved in calculating a variogram using a non-Euclidean distance metric to
incorporate the network pathway structure in the analysis of spatial variability, and use a non-parametric
technique to ascertain if the pattern in the empirical variogram is non-random.
 
Garbrecht, J. (1991). "Effects of Spatial Accumulation of Runoff on Watershed Response." Journal of Enivironmental Quality 20: 31-35.
Garbrecht, J., and L.W. Martz (1999). Digital elevation model issues in water resources modeling. 19th ESRI International User Conference, San Diego, CA.
Garbrecht, J. and L. W. Martz (1997). "The assignment of drainage direction over flat surfaces in raster digital elevation models." Journal of Hydrology 193: 204-213.
Gardner, T. W., et al. (1987). "Geomorphic and tectonic process rates: Effects of measured time interval." Geology 15: 259-261.
Garrigues, S., et al. (2006). "Quantifying spatial heterogeneity at the landscape scale using variogram models." Remote Sensing of Environment 103: 81-96.
The monitoring of earth surface dynamic processes at a global scale requires high temporal frequency remote sensing observations which are
provided up to now by moderate spatial resolution sensors. However, the spatial heterogeneity within the moderate spatial resolution pixel biases
non-linear estimation processes of land surface variables from remote sensing data. To limit its influence on the description of land surface
processes, corrections based on the quantification of the intra-pixel heterogeneity may be applied to non-linear estimation processes. A
complementary strategy is to define the proper pixel size to capture the spatial variability of the data and minimize the intra-pixel variability.
This work provides a methodology to characterize and quantify the spatial heterogeneity of landscape vegetation cover from the modeling of
the variogram of high spatial resolution NDVI data. NDVI variograms for 18 landscapes extracted from the VALERI database show that the land
use is the main factor of spatial variability as quantified by the variogram sill. Crop sites are more heterogeneous than natural vegetation and forest
sites at the landscape level. The integral range summarizes all structural parameters of the variogram into a single characteristic area. Its square
root quantifies the mean length scale (i.e. spatial scale) of the data, which varies between 216 and 1060 m over the 18 landscapes considered. The
integral range is also used as a yardstick to judge if the size of an image is large enough to measure properly the length scales of the data with the
variogram. We propose that it must be smaller than 5% of the image surface. The theoretical dispersion variance, computed from the variogram
model, quantifies the spatial heterogeneity within a moderate resolution pixel. It increases rapidly with pixel size until this size is larger than the
mean length scale of the data. Finally based on the analysis of 18 landscapes, the sufficient pixel size to capture the major part of the spatial
variability of the vegetation cover at the landscape scale is estimated to be less than 100 m. Since for all the heterogeneous landscapes the loss of
NDVI spatial variability was small at this spatial resolution, the bias generated by the intra-pixel spatial heterogeneity on non-linear estimation
processes will be reduced.
 
Gartner, J. E., et al. (2005). Compilation of data relating to the erosive respose of 606 recently-burned basins in the Western U.S., USGS Open File Report 2005-1218.
This report presents a compilation of data on the erosive response, debris-flow initiation processes, basin morphology, burn severity, event-triggering rainfall, rock type, and soils for 608 basins recently burned by 53 fires located throughout the Western United States.  The data presented here are a combination of those collected during our own field research and those reported in the literature.  In some cases, data from a Geographic Information System (GIS) and Digital Elevation Models (DEMs) were used to supplement the data from the primary source.  Due to gaps in the information available, not all parameters are characterized for all basins.
This database provides a resource for researchers and land managers interested in examining relations between the runoff response of recently burned basins and their morphology, burn severity, soils and rock type, and triggering rainfall.  The purpose of this compilation is to provide a single resource for future studies addressing problems associated with wildfire-related erosion.  For example, data in this compilation have been used to develop a model for debris flow probability from recently burned basins using logistic multiple regression analysis (Cannon and others, 2004).  This database provides a convenient starting point for other studies.  For additional information on estimated post-fire runoff peak discharges and debris-flow volumes, see Gartner and others (2004).
 
Gartner, J. E., et al. (2008). "Empirical models to predict the volumes of debris flows generated by recently burned basins in the western U.S." Geomorphology 96: 339-354.
Recently burned basins frequently produce debris flows in response to moderate-to-severe rainfall. Post-fire hazard assessments of
debris flows are most useful when they predict the volume of material that may flow out of a burned basin. This study develops a set of
empirically-based models that predict potential volumes of wildfire-related debris flows in different regions and geologic settings.
The models were developed using data from 53 recently burned basins in Colorado, Utah and California. The volumes of debris
flows in these basins were determined by either measuring the volume of material eroded from the channels, or by estimating the
amount of material removed from debris retention basins. For each basin, independent variables thought to affect the volume of the
debris flow were determined. These variables include measures of basin morphology, basin areas burned at different severities, soil
material properties, rock type, and rainfall amounts and intensities for storms triggering debris flows. Using these data, multiple
regression analyses were used to create separate predictive models for volumes of debris flows generated by burned basins in six
separate regions or settings, including the western U.S., southern California, the Rocky Mountain region, and basins underlain by
sedimentary, metamorphic and granitic rocks.
An evaluation of these models indicated that the bestmodel (theWestern U.S. model) explains 83% of the variability in the volumes
of the debris flows, and includes variables that describe the basin area with slopes greater than or equal to 30%, the basin area burned at
moderate and high severity, and total storm rainfall. This model was independently validated by comparing volumes of debris flows
reported in the literature, to volumes estimated using the model. Eighty-seven percent of the reported volumes were within two residual
standard errors of the volumes predicted using the model. This model is an improvement over previous models in that it includes a
measure of burn severity and an estimate ofmodeling errors. The application of thismodel, in conjunction withmodels for the probability
of debris flows, will enable more complete and rapid assessments of debris flow hazards following wildfire.
 
Gary, H. L. (1974). "Snow Accumulation and Snowmelt as Influenced by a Small Clearing in a Lodgepole Pine Forest." Water Resources Research 10(2): 348-353.
Gasparini, N. M., et al. (1999). "Downstream fining through selective particle sorting in an equilibrium drainage network." Geology 27(12): 1079-1082.
The phenomenon of downstream fining has been attributed to both particle abrasion and
selective particle sorting; the latter is generally considered to play the dominant role within resistant
lithologies. It has been recognized that tributaries can disrupt fining patterns; however,
few downstream-fining studies have considered the entire fluvial network structure. Here we
combine a theory for selective transport with a model of river-basin evolution in order to simulate
the dynamics of selective sorting throughout a drainage network. Previous numerical modeling
studies of single-thread or braided channels have treated downstream fining as a phenomenon
driven by differential deposition rates. We show, however, that in an eroding drainage
network, downstream fining emerges as a natural dynamic adjustment to variable water, sediment,
and energy inputs, even under conditions of uniform size distribution in sediment flux.
Thus, although selective deposition and abrasion clearly can and do play a role in some fluvial
systems, neither is necessary to produce downstream fining within a drainage network.
 
Gautam, M. R., K. Watanabe, and H. Saegusa. (2000). "Runoff analysis in humid forest catchment with artificial neural network." Journal of Hydrology 235(2000): 117-136.
Gedalof, Z., D. Peterson, and N. Mantua (2005). "Atmospheric, climatic, and ecological controls on extreme wildfire years in the northwestern United States." Ecological Applications 15: 545: 154-174.
Gee, G. W., and J.W. Bauder (1986). Particle size analysis. In Methods of soil analysis: Part 1. E. A. Klute. Madison, WI, American Society of Agronomy: pp. 383-411.
Geertsema, M. and J. J. Clague (2006). "1,000-year record of landslide dams at Halden Creek, northeastern British Columbia." Landslides: 11.
Large, rapid, low-gradient landslides are common in
clay-rich glacial sediments in northeastern British Columbia. Many
of the landslides create upstream impoundments that may persist
for years in small watersheds in the region. We have documented
such events in the Halden Creek watershed, 60 km southeast of
Fort Nelson. The events are recorded geologically in two ways.
First, trees are drowned in lakes dammed by the landslides and
subsequently buried by deltaic sediments, where they are protected
from decay. Bank erosion later exhumes the drowned trees.
Second, landslide deposits with entrained wood are exposed along
stream banks. We have reconstructed the recent history of
landslide damming at Halden Creek by performing radiocarbon
dating on exhumed trees and wood in and beneath landslide
deposits at 13 sites in the watershed. Drowned trees range in age
from 169±59 to 274±49 14C year BP. Wood in and below landslide
deposits yielded radiocarbon ages ranging from modern to 965±49
14C year BP
 
Geertsema, M. and J. J. Pojar (2006). "Influence of landslide on biophysical diversity - A perspective from British Columbia." Geomorphology.
Landslides have long been overlooked or underestimated as important natural disturbance agents. In particular the ecological
role of landslides in maintaining biological diversity has been largely ignored. Here we provide a western Canadian (British
Columbian) perspective on the influences of landslides on biophysical diversity, which is related in several ways to biological
diversity. We recognize several types of biophysical/ecological diversity: site diversity, soil diversity, and the derivative habitat or
ecosystem (including aquatic ecosystems) diversity. There are also a variety of landslide types, depending on materials and on the
rate and style of movement. We discuss the roles of different landslide types on various aspects of terrestrial diversity. Landslides
are simultaneously depositional and erosional processes that influence sites by redistributing materials and changing surface
expression —usually creating a complex microtopography that can include very dry ridges and hummocks, and sometimes
depressions with standing water. Landslide impacts to site also influence soil and soil development. Portions of landslides with
exposed parent material are set back to the initial stages of soil development and ecological succession. Landslides can also change
soil density, structure, porosity, surface texture, chemistry and microclimate. By changing site and soil, landslides also influence
habitat. Landslides influence habitat diversity by engendering a mosaic of seral stages (often both primary and secondary), and in
overwhelmingly forested landscapes often create nodes or hotspots of non-forested habitat and biota. In some areas, like the boreal
forest, there is an important interplay between landslides and fire, while on the coast of British Columbia debris and snow
avalanches can be the dominant disturbance agent. Low-gradient and deep-seated landslides are often opportunistically colonized
by beaver and other water and shrub-loving fauna. Sag ponds and impounded streams provide aquatic habitat —often with
standing dead trees. Landslide rubble and scarps provide denning/nesting habitat, escape terrain, and cliff habitat for vertebrates.
 
Geist, D. R. (2000). "Hyporheic discharge of river water into fall chinook salmon (Oncorhynchus tshawytscha) spawning areas in the Hanford Reach, Columbia River." Canadian Journal of Fisheries and Aquatic Science 57: 1647-1656.
Fall chinook salmon (Oncorhynchus tshawytscha) spawned predominantly in areas of the Hanford Reach of
the Columbia River where hyporheic water discharged into the river channel. This upwelling water had a dissolved sol-ids
content (i.e., specific conductance) indicative of river water and was presumed to have entered highly permeable
riverbed substrate at locations upstream of the spawning areas. Hyporheic discharge zones composed of undiluted
ground water or areas with little or no upwelling were not used by spawning salmon. Rates of upwelling into spawning
areas averaged 1200 L·m –2 ·day –1 (95% CI = 784–1665 L·m –2 ·day –1 ) as compared with approximately 500 L·m –2 ·day –1
(95% CI = 303–1159 L·m –2 ·day –1 ) in nonspawning areas. Dissolved oxygen content of the hyporheic discharge near
salmon spawning areas was about 9 mg·L –1 (±0.4 mg·L –1 ) whereas in nonspawning areas, dissolved oxygen values
were 7 mg·L –1 (±0.9 mg·L –1 ) or lower. In both cases, dissolved oxygen of the river water was higher (11.3 ±
0.3 mg·L –1 ). Physical and chemical gradients between the hyporheic zone and the river may provide cues for adult
salmon to locate suitable spawning areas. This information will help fisheries managers to describe the suitability of
salmon spawning habitat in large rivers.
 
Geist, D. R. and D. D. Dauble (1998). "Redd Site Selection and Spawning Habitat Use by Fall Chinook Salmon: The Importance of Geomorphic Features in Large Rivers " Environmental Management 22(5): 655-669.
Geist, D. R., et al. (2002). "Physicochemical characteristics of the hyporheic zone affect redd site selection by chum slamon and fall chinook salmon in the Columbia River." North American Journal of Fisheries Management 22: 1077-1085.
Chum salmon Oncorhynchus keta and fall chinook salmon O. tshawytscha spawned
at separate locations in a side channel near Ives Island, Washington, in the Columbia River
downstream of Bonneville Dam. We hypothesized that measurements of water depth, substrate
size, and water velocity would not sufficiently explain the separation in spawning areas and began
a 2-year investigation of physicochemical characteristics of the hyporheic zone. We found that
chum salmon spawned in upwelling water that was significantly warmer than the surrounding river
water. In contrast, fall chinook salmon constructed redds at downwelling sites, where there was
no difference in temperature between the river and its bed. An understanding of the specific factors
affecting chum salmon and fall chinook salmon redd site selection at Ives Island will be useful
to resource managers attempting to maximize available salmonid spawning habitat within the
constraints imposed by other water resource needs.
 
Geist, D. R., et al. (2000). "Suitability criteria analyzed at the spatial scale of redd clusters improved estimates of fall chinook salmon (Oncorhynchus tshawytscha) spawning habitat use in the Hanford Reach, Columbia River." Canadian Journal of Fisheries and Aquatic Science 57: 1636-1646.
We improved our predictions of fall chinook salmon (Oncorhynchus tshawytscha) habitat use by analyzing
spawning habitat at the spatial scale of redd clusters. Spatial point pattern analyses indicated that redd clusters in the
Hanford Reach, Columbia River, were consistent in their location from 1994 to 1995. Redd densities were 16.1 and 8.9
redds·ha –1 in 1994 and 1995, respectively, and individual redds within clusters were usually less than 30 m apart. Pat-tern
analysis also showed strong evidence that redds were uniformly distributed within the clusters where interredd dis-tances
ranged from 2 to 5 m. Redd clusters were found to occur predominantly where water velocity was between 1.4
and 2 m·s –1 , water depth was 2–4 m, and lateral slope of the riverbed was less than 4%. This habitat use represented a
narrower range of use than previously reported for adult fall chinook salmon. Logistic regression analysis determined
that water velocity and lateral slope were the most significant predictors of redd cluster location over a range of river
discharges. Overestimates of available spawning habitat lead to nonachievable goals for protecting and restoring critical
salmonid habitat. Better predictions of spawning habitat may be possible if cluster-specific characteristics are used.
 
Gell-Man, M. (1994). The Quark and the Jaguar: Adventures in the Simple and the Complex. New York, W. H. Freeman Co.
Gell-Mann, M. (1995). The Quark and the Jaguar. New York, W. H. Freeman and Company.
Geppert, R. R., et al. (1984). Cumulative Effects of Forest Practices on the Environment. Olympia, Washington, prepared by Ecosystems, Inc. for the Washington Forest Practices Board: 208.
Germanoski, D. and M. D. Harvey (1993). Asychronous Terrace Development in Degrading Braided Channels. Physical Geography, V.H. Winston & Son. 14: 16-38.
Gershenfeld, N. (1999). The Nature of Mathematical Modeling. Cambridge, Cambridge University Press.
Gerson, R., et al. (1985). Stages in the creation of a large rift valley - geomorphic evolution along the southern Dead Sea Rift. Tectonic Geomorphology. J. T. Hack and M. Morisawa. London, Allen and Unwin.
Gertner, G., G. Wang, S. Fang, and A.B. Anderson (2002). "Effect and uncertainty of digital elevation model spatial resolutions on predicting the topographic for soil loss estimation." Journal of Soil and Water Conservation m/j: 164-174.
Gesch, D., et al. (2002). "The National Elevation Dataset." Photogrammetric Engineering and Remote Sensing 68(1): 5-11.
The National Elevation Dataset (NED) is a seamless raster product produced by the U.S. Geological Survey (USGS). The NED provides elevation data coverage of the continental United States, Alaska, Hawaii, and the island territories in a seamless format with a consistent projection, resolution, elevation units, and horizontal and vertical datums. The NED is the result of the maturation of the USGS elevation production program (Osborn and others, 2001), in which national coverage of quadrangle-based digital elevation models (DEM) has been completed.
 
Gharrett, J. T. and J. I. Hodges (1950). "Salmon Fisheries of the Coastal Rivers of Oregon South of the Columbia." Oregon Fish Commission Contribution #13: 3-30.
Gholz, H. L. (1982). "Environmental Limits on Aboveground Net Primary Production, Leaf Area, and Biomass in Vegetation Zones of the Pacific Northwest." Ecology 63(2): 469-481.
Giasson, E., et al. GIS-based spatial indices for identification of potential phosphorous export at watershed scale. Journal of Soil and Water Conservation 57 (6), pp. 373-381;   15 ref.; 2002.
Spatial indices for identifying potential pollution resulting from manure spread on agricultural lands were developed for evaluating lands in support of decision- and policy-making. An existing nutrient-delivery ratio was modified by calculating actual distance that water would have to travel to reach a stream and was further tailored to better represent runoff source areas in New York state by incorporating soil drainage class. The Animal Manure Potential Pollution Index (AMPPI) was derived from this modified delivery ratio and animal population census data. The AMPPI and other derived indices use geographical information and nutrient application data to identify and rank geographical areas with respect to potential nutrient export to streams. These indices were applied in the Cannonsville Reservoir Basin in Delaware County, New York, USA. Results demonstrate the potential for using the AMPPI and its derivative indices for selecting priority areas for implementing conservation practices or enrolling lands in programmes such as the Conservation Reserve Program. For example, conservation practices would result in large reductions of potential pollution per unit of area when implemented in identified areas of croplands with high AMPPI. Additionally, an efficient way to reduce total nutrient concentration in those streams that have high nutrient loading would be to enroll in conservation programmes those farms located in subbasins with high nutrient export per unit area, which correspond to areas with high animal density. Farms located in subbasins that have high ratios of nutrient loss per animal unit would benefit from improved manure-management practices, such as improved manure allocation.
 
Gibbons, A. B. and J. D. Mageath (1984). "Probability of moraine survival in a succession of glacial advances " Geology 12: 327-330.
Gifford Pinchot National Forest (?). Gifford Pinchot National Forest Area Cumulative Effects Analysis Final Report to the Forest Management Team, Gifford Pinchot National Forest.
Gilbert, G. K. (1917). Hydraulic-mining debris in the Sierra Nevada, United States Geological Survey: 154.
Gillian, S. E. (1989). Storage Dynamics of Fine Woody Debris for Two Low-order Coastal Streams in Southeast Alaska. ? ?, ?: 85.
Gillilan, S. E. (1989). Storage dynamics of fine woody debris for two low-order coastal streams in southeast Alaska. Forest Science. Corvallis, OR, Oregon State University.
Gilpin, M. E. (1987). Minimum viable populations: a restoration ecology perspective. Restoration Ecology. W. R. Jordan. Cambridge, MA, Cambridge University Press: 301-305.
Gippel, C. (1985). "Changes in stream channel morphology at tributary junctions, Lower Hunter Valley, New South Wales." Australian Geographical Studies 23: 291-307.
Many empirically derived models of downstream changes in channel form are based on the concept of hydraulic geometry, whereby changes in channel morphological variablies are related by a continuous power function to increasing discharge or its surrogate. However, changes in discharge throughout a drainage basin are normally concentrated at the junctions of stream channels. At junctions, the relationship between discharge and channel morphology is discontinuous. Some Lower Hunter Valley stream channels were surveyed upstream and downstream of tributary junctions. While significant regional downstream relationships were established, consistent trends at tributary junctions were not apparent from the data. It is possible that considerable scatter evident in ratios of change on either side of eachh junction was due to local within-site variability. In addition, the relative timing of discharge inputs and the nature of introduced sediment loads could be important in determining the magnitude and direction of channel change at tributary junctions. Interrelated changes in cross-section, profile and planform variables may be expected below junctions. The resultant channel morphology, in terms of these variables, might not be wholly determinate.
 
Glade, T. (2003). "Landslide occurrence as a response to land use change: a review of evidence from New Zealand." Catena 51: 297-314.
Glade, T. (2005). "Linking debris-flow hazard assessments with geomorphology." Geomorphology 66: 189-213.
Debris-flow hazard assessment schemes are commonly based on empirical, physical, or numerical methods and techniques.
Inherent in all methods is generally the assumption of unlimited sediment supply. This study compares model inputs of
sediment requirements for debris flows with estimated sediment reproduction from both solifluction and rockwall retreat. The
analysis is carried out in Bý´ldudalur, a community in the Westfjords of Iceland. Geomorphic techniques are applied to determine
the set of natural processes acting in this landscape to estimate spatial distribution of relevant processes, to approximate level of
processes activity, and to provide information for scenario modeling. Debris-flow volumes are determined by coupling rainfall
magnitudes and catchment sizes with average sediment contents. Rockwall retreat and solifluction rates are based on literature
reviews.
For a rainstorm with a 10-year return period, debris-flow volumes are calculated for 12 different creeks. Rates are assumed
for solifluction with a velocity of 0.25 m/yr at an average depth of 0.5 m and for rockwall retreat with 2 mm/yr. Comparing
sediment requirements with estimated sediment reproduction leads to a factor of deficit ranging between 6.2 and 8.5. Thus, the
sediment storage is not refilled as fast as the next potential triggering rainfall occurs. Consequently, if a debris flow has occurred
in the past, all sediment is removed, and the following rainstorm event is djustT causing a flood, which is by far less destructive
than a debris-flow event. The challenge of future debris-flow hazard-assessment schemes is to include geomorphic analysis to
be able to obtain more sustainable results.
 
 
Glass, C. E. and R. Klimmek (2001). "Routing Debris Flows." Environmental & Engineering Geoscience 7(2): 177-191.
Goetz, S. J. (2006). "Remote sensing of riparian buffers: past progress and future prospects." Journal of the American Water Resources Association 42(1): 133-143.
Riparian buffer zone management is an area of increasing
relevance as human modification of the landscape continues unabated.
Land and water resource managers are continually challenged to maintain
stream ecosystem integrity and water quality in the context of
rapidly changing land use, which often offsets management gains.
Approaches are needed not only to map vegetation cover in riparian
zones, but also to monitor the changes taking place, target restoration
activities, and assess the success of previous management actions. To
date, these objectives have been difficult to meet using traditional
techniques based on aerial photos and field visits, particularly over
large areas. Recent advances in remote sensing have the potential to
substantially aid buffer zone management. Very high resolution
imagery is now available that allows detailed mapping and monitoring
of buffer zone vegetation and provides a basis for consistent assessments
using moderately high resolution remote sensing (e.g., Landsat).
Laser-based remote sensing is another advance that permits even
more detailed information on buffer zone properties, such as refined
topographic derivatives and multidimensional vegetation structure.
These sources of image data and map information are reviewed in this
paper, examples of their application to riparian buffer mapping and
stream health assessment are provided, and future prospects for
improved buffer monitoring are discussed.
 
Gol'din, B. M. and Lyubashevskiy (1966). "Computation of the velocity of mudflows for Crimean Rivers." Soviet Hydrology 2: 179-181.
Gol'din, B. M. and L. S. Lyubashevskiy (1966). "Computation of the Velocity of Mudflows for Crimean Rivers." Transactions of the Ukrainian Hydrometeorological Research Institute 60: 73-75.
Gomez, B. (1983). "Temporal variations in bedload transport rates: the effect of progressive bed armouring." Earth Surface Processes and Landforms 8: 41-54.
Gomez, B., and M. Church (1989). "An assessment of bed load sediment transport formulae for gravel bed rivers." Water Resources Research 25(6): 1161-1186.
Gomez, B. (1991). "Bedload transport." Earth-Sc. Rev. 31: 89-132.
The complete understanding of bedload transport requires the development of reliable apparatus and techniques for predicting, measuring, and sampling bedload in rivers. This review presents an overview of the development of knowledge and research into bedload transport during the past century. Particular emphasis is placed on the development of methods of predicting and estimating bedload discharge. Problems involved in using field data to calibrate bedload transport fomulae and bedload samplers, and factors that affect the design of sampling programmes are specifically addressed. Reluctance to acknowledge the inherently unstable nature of bedload transport has been a principal factor in limiting progress. Some causes of temporal and spatial instability in bedload transport are isolated, and their effect upon the transport process elucidated.
 
Gomez, B. (1995). Bedload transport and changing grain size distributions. Changing River Channels. A. Gurnell and G. Petts, John Wiley & Sons Ltd.: 177-199.
Gomez, B., et al. (2003). "Gully erosion and sediment production: Te Weraroa Stream, New Zealand." Water Resources Research 39(7): 1187.
We derive a sediment budget for Te Weraroa Stream, New Zealand, the principal
drainage in a small (29 km2) steepland catchment where gully erosion, triggered by
conversion to pasture early in the twentieth century, was ameliorated by reforestation that
commenced in 1962. Estimates of sediment production were made using the change in
gully area observed in sequential aerial photographs. Channel storage was assessed from
stream cross-section surveys. At its peak, gully erosion affected  6% of the total catchment
area. The amount of sediment contributed from gullies declined by 62% as the forest
became established, but of the 28.7 Mt of sediment generated by gully erosion between
1950 and 1988, 48% was stored in the channel along the lower 8 km of TeWeraroa Stream.
Even if the amount of sediment generated by gully erosion continues to decline, it likely
will be many decades before the gravel is released from storage.
 
Gomez, B. and M. Church (1989). "An assessment of bed load sediment transport formulae for gravel bed rivers." WRR 25(6): 1161-1186.
The performance is tested of 12 bed load sediment transport formulae developed for use in gravel bed channels. The formulae are applied in the manner intended by the original authors. To this end, the test data are restricted to ones obtained in approximately steady flow when the material in motion was similar to that present on the bed; that is, transport was not size selective. This represents the nearest approach to "equilibrium sediment transport" that can be realized with available data. Four sets of river data and three sets of flume data were chosen for the test, covering a range of eight orders of magnitude in unit bed load transport rate. The test data were not used in the development of the formulae. The analysis separates mean bias and local bias. No formula performs consistently well. Limitations of the test data, the constraints imposed by an operationally realistic test, and reasons based upon the physics of the transport phenomenon all may be adduced for this. To estimate the magnitude of transport with limited hydraulic information, stream power equations should be used because they provide the most straightforward scale correlation of the phenomenon. In particular, the approach of Bagnold deserves further study. When local hydraulic information is available, a formula should be selected that is sensitive to bed state or grain size distribution and, in this context, the formulae of Einstein, Parker, and Ackers-White-Day bear continued examination.
 
Gomez, B., et al. (1998). "Floodplain construction by recent, rapid vertical accretion: Waipaoa River, New Zealand." Earth Surface Processes and Landforms 23: 405-413.
Gomez, B., et al. (1989). "Temporal variations in bedload transport rates associated with the migration of bedforms." Earth Surface Processes and Landforms 14: 135-156.
Gomi, T., R.D. Moore, M.A. Hassan (2005). "Suspended sediment dynamics in small forest streams of the Pacific Northwest." Journal of the American Water Resources Association 41(4): 877-898.
Gomi, T. and R. C. Sidle (2003). "Bed load transport in managed steep-gradient headwater streams of southeastern Alaska." Water Resources Research 39(12): DOI: 1029/2003WR002440, 002003.
Bed load transport was investigated in four headwater streams in southeastern Alaska subjected to different management and disturbance regimes. Bed load yield was positively correlated to peak discharge during the fall 1999 monitoring period. Fine bed load materials (1–11 mm) that were supplied from hillslope sources were equally mobilized during most storm events. Medium-sized bed materials (11–200 m) were only partially mobilized even during large storms, whereas large particles (>200 mm) were immobile and often formed interlocking channel structures. The transport distances of medium-size materials depended on amount of channel obstructions (e.g., woody debris) and sediment supply conditions; both of these factors were influenced by the occurrence of mass movement, timber harvesting, and the related recovery processes. The highest total bed load yield was observed in a channel affected by a debris flow in 1993. Total sediment yields are similar among channels with old-growth, clear cut (logged 4 years before monitoring), and young alder (affected by landslides and debris flows in 1961) riparian stands. By comparing the old-growth and young alder channels, it appears that bed load yield recovers from debris flow disturbances in about 40 years; however, recovery of channel conditions (e.g., reach types and woody debris) may take much longer. Effects of timber harvesting on bed load transport are controlled by sediment linkages between hillslopes and channels related to the occurrence of mass movement.
 
Gomi, T., et al. (2001). "The characteristics of woody debris and sediment distribution in headwater streams, southeastern Alaska." Canadian Journal of Forest Research 31: 1386-1399.
Large woody debris (LWD), fine woody debris (FWD), fine organic debris (FOD), and sediment deposition
were measured in 15 steep headwater streams with five management and disturbance regimes. Clear-cut channels
logged in 1995 contained large accumulations of logging residue that initially provided sites for sediment storage. Half
of the LWD in clear-cut channels was recruited during and immediately after logging. Woody debris from logging activities
remains in young growth conifer channels 37 years after logging. Numbers of LWD in clear-cut and young conifer
channels were significantly higher than in old-growth channels, although numbers of FWD pieces were not
significantly different because of higher recruitment from old-growth stands. Channels that experienced recent (1979
and (or) 1993) and earlier (1961 and (or) 1979) scour and runout of landslides and debris flows contained less LWD
and FWD, although large volumes of LWD and FWD were found in deposition zones. The volumes of sediment stored
in young alder and recent landslide channels were higher than in the other channels. Because of the recruitment of
LWD and FWD from young alder stands, the ratio of sediment stored behind woody debris to total sediment volume
was higher in young alder channels compared with recent landslide channels. Numbers of LWD and FWD pieces in all
streams were significantly correlated with the volumes of sediment stored behind woody debris. Timber harvesting and
soil mass movement influence the recruitment, distribution, and accumulation of woody debris in headwater streams;
this modifies sediment storage and transport in headwater channels.
 
Gomi, T., et al. (2002). "Understanding Processes and Downsteam Linkages of Headwater Systems." BioScience 52(10): 905-916.
Gomi, T., et al. (2002). "Understanding processes and downstream linkages of headwater systems." BioScience 52(10): 905-916.
Gomi, T., et al. (2004). "Hydrogeomorphic linkages of sediment transport in headwater streams, Maybeso Experimental Forest, southeast Alaska." Hydrological Processes 18: 667-683.
Hydrogemorphic linkages related to sediment transport in headwater streams following basin wide clear-cut logging
on Prince of Wales Island, southeast Alaska, were investigated. Landslides and debris flows transported sediment
and woody debris in headwater tributaries in 1961, 1979, and 1993. Widespread landsliding in 1961 and 1993
was triggered by rainstorms with recurrence intervals (24 h precipitation) of 7Ð0 years and 4Ð2 years respectively.
Occurrence, distribution, and downstream effects of these mass movements were controlled by landform characteristics
such as channel gradient and valley configuration. Landslides and channelized debris flows created exposed bedrock
reaches, log jams, fans, and abandoned channels. The terminus of the deposits did not enter main channels because
debris flows spread and thinned on the unconfined bottom of the U-shaped glaciated valley. Chronic sediment input to
channels included surface erosion of exposed till (rain splash, sheet erosion, and freeze–thaw action) and bank failures.
Bedload sediment transport in a channel impacted by 1993 landslides and debris flows was two to ten times greater and
relatively finer compared with bedload transport in a young alder riparian channel that had last experienced a landslide
and debris flow in 1961. Sediment transport and storage were influenced by regeneration of riparian vegetation,
storage behind recruited woody debris, development of a streambed armour layer, and the decoupling of hillslopes
and channels. Both spatial and temporal variations of sediment movement and riparian condition are important factors
in understanding material transport within headwaters and through channel networks.
 
Gonor, J. J., et al. (1988). What we know about large trees in estuaries, in the sea, and on coastal beaches. From the Forest to the Sea: A Story of Fallen Trees. C. Maser, R. F. Tarrant, j. M. Trappe and J. M. Franklin. Portland, OR, U.S.D.A., Forest Service. PNW-GTR-229: 83-112.
Goodrich, D. C., et al. (1997). "Linearity of basin response as a function of scale in a semiarid watershed." Water Resources Research 33(12): 2951-2965.
Goodwin, P. (2004). "Analytical solutions for estimating effective discharge." Journal of Hydraulic Engineering-ASCE 130(8): 729-738.
Gooseff, M. N., et al. (2008). On the relationships among temporal patterns of evapo-transpiration, stream flow and riparian water levels in headwater catchments during baseflow. Integrating Groundwater Science and Human Well Being. Yoyama, Japan: 842-851.
Diel fluctuations in stream flow during baseflow discharge and coincident fluctuations in riparian
water levels have been observed in many streams and are typically attributed to water losses from evapotranspiration
(ET). Riparian water level fluctuation may be due to 1) changes in stream stage, 2) changes in
supply of subsurface hillslope water to the riparian zone, and/or 3) ET from riparian vegetation. We analyze data
collected over the summer of 2004 from a previously-established network of wells located across a corridor of a
headwater stream (~100 ha contributing area) in the central Cascade Mountains of Oregon, USA. Our findings
indicate that groundwater-stream interactions link drawdown of riparian aquifers with streams in very complex
spatial and temporal patterns even at the local scale (m to 10s of m). This is coincident with the effects of basinwide
limitation of hillslope water yield to riparian zones (and streams) that cause local changes in stream stage.
 
Gordon, S. J. and R. D. Klousner, Jr. (1986). "Using Landslide Hazard Information in Planning." APA Journal 52(4): 431-442.
Gorsline, D. S. (1980). Sediment delivery to California margin basins over the past 1000 years. International Geological Congress, Abstracts--Congres Geologique Internationale, Resumes. location varies, International Geological Congress. 26: 477.
Gottesfeld, A. S., et al. (2004). "Sediment dispersion in salmon spawning streams: the influence of floods and salmon redd construction." Journal of the American Water Resources Association 40(4): 1071-1086.
Magnetically tagged particles were used to investigate
the effects of sockeye salmon (Oncorhynchus nerka) and floods on
the dispersion of coarse bed material in the Stuart-Takla region,
British Columbia, Canada. The dominant annual sediment transporting
event in the channels is the snowmelt flood events, with
lesser activity usually accomplished during summer floods. Annually
in August, the channel bed material is reworked by the Early
Stuart sockeye salmon spawning, as the fish excavate the
streambed to deposit and bury their eggs. These nesting cavities
are called redds. Results from 67 tracer recovery experiments over
five years were highly variable, subject to the magnitude of floods
and the returning population of salmon. Overall, the depositional
pattern from nival flood events usually demonstrated a high degree
of clast mobilization, long travel distances (up to 150 m), and mean
depths of burial up to 18 cm. Summer flood events showed somewhat
lower rates of mobilization, distances of travel, and depths of
burial. Although the fish did not move the tracers very far, their
effect on the bed was generally quite pervasive – up to 100 percent
of the clasts were mobilized, and the depth of burial was considerable
(mean burial depths up to 14 cm). The amount of vertical mixing
of sediment by salmon was often on the same order of
magnitude as flood events. The significant vertical mixing of sediments
by the fish has important implications for the mobility of
sediment in the stream. Since any armoring layer formed during
high flows throughout the year is subject to the bioturbation of
salmon, this suggests that the transport threshold in the creeks
remains relatively low. Salmon likely play an integral role in the
sediment transport dynamics and annual sediment budget of the
lower reaches of these creeks.
 
 
Govindaraju, R. S., and A. Ramachandra, Ed. (2000). Artificial neural networks in hydrology. Dordrecht, The Netherlands, Kluwer Academic Publlishers.
Graf, W. L. (1977). "The rate law in fluvial geomorphology." American Journal of Science 277: 178-191.
Graf, W. L. (1979). Catastrophe Theory as a Model for Change in Fluvial Systems. Adjustments of the Fluvial System. Kendall and Hunt: 13-32.
Graham, A. L. and J. O. Annexstad (1989). "Antarctic meteorites." Antarctic Science 1(1): 3-14.
Granger, D. E., et al. (1996). "Spatially Averaged Long-term Erosion Rates Measured From in Situ-produced Cosmogenic Nuclides in Alluvial Sediment." The Journal of Geology 104: 249-257.
Grant, E. H. C., et al. (2007). "Living in the branches: population dynamics and ecological processes in dendritic networks." Ecology Letters 10: 165-175.
Spatial structure regulates and modifies processes at several levels of ecological
organization (e.g. individual/genetic, population and community) and is thus a key
component of complex systems, where knowledge at a small scale can be insufficient for
understanding system behaviour at a larger scale. Recent syntheses outline potential
applications of network theory to ecological systems, but do not address the implications
of physical structure for network dynamics. There is a specific need to examine how
dendritic habitat structure, such as that found in stream, hedgerow and cave networks,
influences ecological processes. Although dendritic networks are one type of ecological
network, they are distinguished by two fundamental characteristics: (1) both the branches
and the nodes serve as habitat, and (2) the specific spatial arrangement and hierarchical
organization of these elements interacts with a species  movement behaviour to alter
patterns of population distribution and abundance, and community interactions. Here,
we summarize existing theory relating to ecological dynamics in dendritic networks,
review empirical studies examining the population- and community-level consequences
of these networks, and suggest future research integrating spatial pattern and processes
in dendritic systems.
 
 
Grant, G. (1986). The Rapid Technique: A New Method for Evaluating Downstream Effects of Forest Practices on Stream Channels (draft). Corvallis, OR, Forestry Sciences Laboratory: 51.
Grant, G. (1997). A geomorphic basis for interpreting the hydrologic behavior of large river basins. Boca Raton, FL, CRC Press Inc.
Grant, G. E. (1977). A geomorphic basis for interpreting the hydrologic behavior of large river basins. River Quality: Dynamics & Restoration. A. Laenen and D. A. Dunnette. Boca-Raton, Florida, CRC Press: 105-116.
Grant, G. E. (1986). Assessing Effects of Peak Flow Increases on Stream Channels: A Rational Approach. California Watershed Management Conference, West Sacramento, California.
Grant, G. E. (1986). Downstream Effects of Timber Harvest Activities on the Channel and Valley Floor Morphology of Western Cascade Streams. Baltimore, Maryland, Johns Hopkins University.
Grant, G. E. (1989). Cumulative effects of forest management on sediment production, transport, and storage; an overview. Agronomy Abstracts, vol.81. E. C. A. Runge: 303.
Grant, G. E. (1997). "Critical flow constrains flow hydraulics in mobile-bed streams: A new hypothesis." Water Resources Research 33(2): 349-358.
Grant, G. E. (?). The changing environment of debris flow management in the Pacific Northwest, U.S. and Japan. Corvallis, Oregon, US Forest Service: 11.
Grant, G. E. (?). Morphology of high gradient streams at different spatial scales, Western Cascades, Oregon. Workshop on "Channel Geomorphological Change and Sediment Control at Devastated River".
Grant, G. E., et al. (1992). XSPRO: A channel cross-section analyzer, U.S. Department of the Interior, Bureau of Land Management
 
U.S. Department of Agriculture
Forest Service: 3-53.
Grant, G. E., et al. (2008). Effects of forest practices on peak flows and consequent channel response: A state-of-science report for western Oregon and Washington. Portland, OR, U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station: 76.
This is a state-of-the-science synthesis of the effects of forest harvest activities on
peak flows and channel morphology in the Pacific Northwest, with a specific focus
on western Oregon and Washington. We develop a database of relevant studies
reporting peak flow data across rain-, transient-, and snow-dominated hydrologic
zones, and provide a quantitative comparison of changes in peak flow across both a
range of flows and forest practices. Increases in peak flows generally diminish with
decreasing intensity of percentage of watershed harvested and lengthening recurrence
intervals of flow. Watersheds located in the rain-dominated zone appear to be
less sensitive to peak flow changes than those in the transient snow zone; insufficient
data limit interpretations for the snow zone. Where present, peak flow effects on
channel morphology should be confined to stream reaches where channel gradients
are less than approximately 0.02 and streambeds are composed of gravel and finer
material. We provide guidance as to how managers might evaluate the potential risk
of peak flow increases based on factors such as presence of roads, watershed drainage
efficiency, and specific management treatments employed. The magnitude of
effects of forest harvest on peak flows in the Pacific Northwest, as represented by
the data reported here, are relatively minor in comparison to other anthropogenic
changes to streams and watersheds.
 
 
Grant, G. E. and T. Mizuyama (1991). Origin of step-pool sequences in high gradient streams: a flume experiment. Japan-U.S. Workshop on Snow Avalanche, Landslide, Debris Flow Prediction and Control.
Grant, G. E., et al. (1990). "Pattern and origin of stepped-bed morphology in high-gradient streams, Western Cascades, Oregon." GSA Bull 102: 340-352.
A general hierarchical framework for viewing steped-bed morphology in high-gradient channels is presented. We emphasize channel units - bed features that are one or more channel widths in length - as a particularly important scale of variation. Field strdies in two streams in the Cascade Range in Oregon indicated that pool, riffle, rapid, cascade, and step channel units had distinct bed slope ranges, with average slopes of 0.005, 0.011, 0.029, 0.055, and 0.173, rspectively. Steeper units (rapids and cascades) are composed of step-pool sequences created by particles representing the 90th or larger percentile size fraction of bed material. Step spacing is inversely proportional to bed slope.
  The distribution of channel units along a stream is influenced by bedrock and processes that introduce coarse sediment. Cascade and pool units dominate where landslide and debris-flow deposits constrict channel width and deliver large immobile boulders to the channel, whereas riffle and rapid units dominate in broad valley flats where deposition of finer sediment occurs. Markov chain analysis indicates that channel units occur in nonrandom two-unit sequences with the slope of the upstream unit inversely proportional to the slope of the next downstream unit. Pool-to-pool spacings average two to four channel widths, but variability in spacing is high, owing to uneven distribution of bedrock outcrops and boulder deposits within the channel.
  Hydraulic reconstruction indicates that channel units form during high-magnitude, low-frequency events with recurrence intervals of about 50 yr. Comparison of channel-unit morphology to high-gradient flume experiments with heterogenous bedload mixtures indicated that unit morphogenesis is linked to factors that cause congestion of large particles during bedload transport events; these include local constrictions in channel width, immobile bed material, and abrupt fluctions in velocity due to hydraulic jumps that promote deposition. Channel units appear to be a two-dimensional bar form found in streams where gradients exceed 2%, bedload is widely sorted, and width-to-depth ratios and sediment supply are low - condiditons found in many mountain environments.
 
Grant, G. E., et al. (1990). "Pattern and origin of stepped-bed morphology in high-gradient streams, Western Cascades, Oregon." Geological Society of America 102: 340-352.
A general hierarchical framework for viewing steped-bed morphology in high-gradient channels is presented. We emphasize channel units - bed features that are one or more channel widths in length - as a particularly important scale of variation. Field strdies in two streams in the Cascade Range in Oregon indicated that pool, riffle, rapid, cascade, and step channel units had distinct bed slope ranges, with average slopes of 0.005, 0.011, 0.029, 0.055, and 0.173, rspectively. Steeper units (rapids and cascades) are composed of step-pool sequences created by particles representing the 90th or larger percentile size fraction of bed material. Step spacing is inversely proportional to bed slope.
  The distribution of channel units along a stream is influenced by bedrock and processes that introduce coarse sediment. Cascade and pool units dominate where landslide and debris-flow deposits constrict channel width and deliver large immobile boulders to the channel, whereas riffle and rapid units dominate in broad valley flats where deposition of finer sediment occurs. Markov chain analysis indicates that channel units occur in nonrandom two-unit sequences with the slope of the upstream unit inversely proportional to the slope of the next downstream unit. Pool-to-pool spacings average two to four channel widths, but variability in spacing is high, owing to uneven distribution of bedrock outcrops and boulder deposits within the channel.
  Hydraulic reconstruction indicates that channel units form during high-magnitude, low-frequency events with recurrence intervals of about 50 yr. Comparison of channel-unit morphology to high-gradient flume experiments with heterogenous bedload mixtures indicated that unit morphogenesis is linked to factors that cause congestion of large particles during bedload transport events; these include local constrictions in channel width, immobile bed material, and abrupt fluctions in velocity due to hydraulic jumps that promote deposition. Channel units appear to be a two-dimensional bar form found in streams where gradients exceed 2%, bedload is widely sorted, and width-to-depth ratios and sediment supply are low - condiditons found in many mountain environments.
 
Grant, G. E. and A. L. Wolff (1991). Long-term patterns of sediment transport after timber harvest, western Cascade Mountains, Oregon, USA. Sediment and Stream Water Quality in a Changing Environment. N. E. Peters and D. E. Walling, International Association of Hydrological Sciences. 203: 31-40.
Suspended and bedload sediments were sampled from 1958-1988 on three small watersheds in the western Cascade Range in Oregon. Annual sediment yields varied greatly among watersheds, and the pattern of long-term sediment production reflects their timber harvest and mass movement histories. Total yields from 1958-1988 were 5100 t km-2 in the clearcut watersheds (WS1), 21000 t km-2 in the patchcut watershed with roads (WS3), and 800 t km-2 in the forested control (WS 2). More than 85% of the total sediment yield in WS 3 occurred during a storm in 1964 when a series of debris flows scoured the channel to bedrock. Excluding that event, post-logging annual export from WS 1 has been more than twice that from WS 3. The importance of episodic mass erosion events in this landscape limits the effectiveness of small watershed studies for analyzing long-term sediment yields.
 
Grant, G. G. and F. J. Swanson (1995). Morphology and processes of valley floors in mountain streams, Western Cascades, Oregon. Natural and Anthropogenic Influences in Fluvial Geomorphology, Geophysical Monograph 89. J. E. Costa, a. J. Miller, K. W. Potter and P. R. Wilcock. Washington, D. C., American Geophysical Union: 83-101.
Grant, G. T., and A.L. Wolff (1991). Long-term patterns of sediment transport after timber harvest, Western Cascade Mountains, Oregon. Sediment and Stream Water Quality in a Changing Environment: Trends and Explanation. Proceedings of the Vienna Symposium. IAHS Publication no. 203.
Grant, J. W. A., et al. (1998). "Implications of territory size for the measurement and prediction of salmonid abundance in streams." Canadian Journal of Fisheries and Aquatic Science 55(Suppl. 1): 181-190.
Information about territory size is useful for both the measurement and prediction of salmonid abundance.
Percent habitat saturation (PHS), the percentage of the stream area occupied by the territories of salmonid fishes, is a
better measure of abundance than population density because the former integrates the effects of (a) several age-classes
or species in a stream, and (b) variation in growth rate or sampling date. “Effective density” or “effective PHS,”
calculated by weighting crude density (no.·m-2) or PHS by the number of organisms in the sampling unit, more
accurately reflects density from the organism’s point of view than does crude density or PHS. Effective density and
PHS of Atlantic salmon (Salmo salar) in Catamaran Brook, New Brunswick, increased by 0.4 fish per m2 and 4%,
respectively, for each order of magnitude decrease in the area of the sampling unit. Literature data suggested that
territory size is inversely related to food abundance and can be used to predict changes in salmonid abundance that
accompany changes in food abundance. The allometry of territory size was a better predictor of the decline in density
of a cohort of Atlantic salmon in Catamaran Brook than the allometry of metabolic requirements.
 
Graumlich, L. J. (1987). "Precipitation Variation in the Pacific Northwest (1675-1975) as Reconstructed from Tree Rings." Annals of the Association of American Geographers 77(1): 19-29.
Graumlich, L. J. and L. B. Brubaker (1986). "Reconstruction of Annual Temperature (1590-1979) for Longmire, Washington, Derived from Tree Rings." Quaternary Research 25: 223-234.
Graves-Morfin, S. R. (2000). Moscow, University of Idaho.
With the growing public use of the national forests, forest engineers, managers, and planners have had the challenge of upgrading and maintaining extensive road networks, while minimizing the ecological impacts of those road systems on the environment. Properly designing road drainage for local soils, topography, and slope conditions can reduce erosion from road surfaces, subsequently reducing sedimentation risk elsewhere. Knowledge about the amount of soil eroded and delivered from the road system is valuable when making decisions about road drainage design or redesign. The Water Erosion Prediction Project (WEPP) model, a process-based erosion model, was validated and then used to create an output database of erosion amounts for 79,200 different road configurations, locations, and soil types. This output then was used to derive equations for erosion and runoff from low volume roads. Two approaches were used to derive these equations. The first was a Revised Universal Soil Loss Equation (RUSLE) approach where soil erodibility (K) factors and length-slope (LS) factors were derived using the WEPP database and the other RUSLE factors. This method resulted in equations that estimated road erosion within + or -50% of the database values. The second approach using the WEPP database was based on multivariate regressions and the development of equations for road soil loss, road runoff, and sediment yield. Single interaction, double interaction, and quadratic interaction multivariate analyses were done using the Statistical Analysis System (SAS) and the WEPP database to develop the regression equations. This approach resulted in road soil loss and runoff values within + or -20% of the original database values. This approach was also attempted for sediment delivery across a buffer, however, it was less successful due to complications from such factors as antecedent moisture, vegetative cover, and sediment loading in runoff.
 
Gray, D. H. and W. F. Megahan (1981). Forest vegetation removal and slope stability in the Idaho Batholith. Ogden, Utah, USDA Forest Service: 23.
Gray, L. G. (1934). "Long-period fluctuations of some meteorological elements in relation to California forest-fire problems." Monthly Weather Review: 231-235.
Green, D. G. (1994). "Connectivity and complexity in landscapes and ecosystems." Pacific Conservation Biology: 1-10.
Green TR, S. B., CR Dietrich, and AJ Jakeman (1999). "Relating stream-bank erosion to in-stream transport of suspended sediment." Hydrologic Processes 13: 777-787.
Gregory, K. J. (1992). Vegetation and River Channel Process Interactions. River Conservation and Management. P. J. Boon, P. Calow and G. E. Petts, John Wiley and Sons Ltd. 16: 255-269.
Gregory, K. J., et al. (1985). "The permance of debris dams related to river channel processes." Hydrological Sciences 30(3, 9): 371-381.
Gregory, K. J., et al. (1985). "The permanence of debris dams related to river channel processes." Hydrological Sciences 30(9): 371-381.
Gregory, S. and R. Wildman (1992). Aquatic Ecosystem Restoration Project. Corvallis, Oregon, Oregon State University: 1-29.
Gregory, S. V., F. J. Swanson, W.A. McKee, and K.W. Cummins (1991). "An ecosystem perspective of riparian zones." BioScience 41(8): 540-551.
Gregory, S. V., et al., Eds. (2003). The Ecology and Management of Wood in World Rivers. Bethesda, Maryland, American Fisheries Society.
Gregory, S. V., et al. (2003). Modeling the dynamics of wood in streams and rivers. The Ecology and Management of Wood in World Rivers, American Fisheries Symposium 37. S. V. Gregory, K. L. Boyer and A. M. Gurnell. Bethesda, Maryland, American Fisheries Society: 315-335.
Gregory, S. V., et al. (1991). "An Ecosystem Perspective of Riparian Zones." BioScience 41(8): 540-551.
Gregory, S. V. a. n. a. (2000). Bibliography: World literature on wood in streams, rivers, estuaries and riparian areas version 1.0. International Conference on Wood in World Rivers, Corvallis, Oregon.
Gresswell, R. E. (1999). "Fire and aquatic ecosystems in forested biomes of North America." Transactions of the American Fisheries Society 128: 193-221.
Gresswell, R. E. (1999). Fire and Aquatic Ecosystems in Forested Biomes of North America (DRAFT). Corvallis, Oregon, Oregon State University: 50.
Gresswell, R. E., et al. (2006). A spatially explicit approach for evaluating relationships among coastal cutthroat trout, habitat, and disturbance in small Oregon streams. Landscape Influences on Stream Habitats and biological Assemblages. R. M. Hughes, L. Wang and P. W. Seelbach. Bethesda, Maryland, American Fisheries Society. Symposium 48: 457-471.
Small stream systems are complex networks that form a physicochemical template
governing the persistence of aquatic species such as coastal cutthroat trout Oncorhynchus clarkii
clarkii. To gain new insight into these interactions, we initiated an integrated program of landscape-
scale sampling that is focused on fine- and broad-scale relationships among upslope
landscape characteristics, physical stream habitat, and the spatial patterns of cutthroat trout
abundance. Our sample of 40 catchments (500–1,000 ha) represented approximately 15% of
the 269 barrier-isolated catchments in western Oregon that support populations of cutthroat
trout. Because data were collected in a spatially contiguous manner throughout each catchment,
it was possible to collect biological and geographic information necessary to assess the
spatial structure of cutthroat trout abundance. Results underscore the influence of the physical
habitat template at a variety of spatial scales. For example, cutthroat trout move throughout
the accessible portions of small streams. Some cutthroat trout congregate in areas of suitable
habitat and form local populations that may exhibit unique genetic attributes. At times, some
cutthroat trout move into larger downstream portions of the network where they may contribute
to the genetic character of anadromous or local potamodromous assemblages. Results
underscore the advantages of viewing habitats that are critical to the fitness and persistence of
cutthroat trout populations as matrices of physical sites that are linked by movement. It is
apparent that human activities that impede movement among suitable habitat patches can
have unanticipated consequences for metapopulations of cutthroat trout and may ultimately
affect their persistence.
 
Gresswell, R. E. and S. R. Hendricks (2007). "Population-scale movement of coastal cutthroat trout in a naturally isolated stream network." Transactions of the American Fisheries Society 136: 238-253.
To identify population-scale patterns of movement, coastal cutthroat trout Oncorhynchus
clarkii clarkii tagged and marked (35 radio-tagged, 749 passive integrated transponder [PIT]-tagged, and
3,025 fin-clipped) were monitored from June 1999 to August 2000. The study watershed, located in western
Oregon, was above a natural barrier to upstream movement. Emigration out of the watershed was estimated
with a rotating fish trap. Approximately 70% of recaptured coastal cutthroat trout with PIT tags and 86% of
those with radio tags moved predominantly at the channel-unit scale (2–95 m); fewer tagged fish moved at
the reach scale (66–734 m) and segment scale (229–3,479 m). In general, movement was greatest in April as
spawning peaked and lowest in October, when discharge was at its lowest. Only 63 (,1% of tagged and
marked fish) coastal cutthroat trout were captured in the fish trap. Trap efficiency was about 33%, and the
expanded estimate of emigrants between February and June was 173 fish. These results suggest that unitscale
movement is common throughout the year and that reach- and segment-scale movements are important
during the winter and spring. Although movement in headwater streams is most common at the channel-unit
scale, restoration of individual channel units of stream may not benefit the population at the watershed scale
unless these activities are undertaken in the context of the greater whole. Individual coastal cutthroat trout
move great distances, even within the small watersheds in the Oregon Coast Range, and although these
movements may be infrequent, they may contribute substantially to recolonization after stochastic extirpation
events (e.g., landslides and debris flows). Management strategies that focus on maintaining and restoring
connectivity in a watershed represent an important step toward protecting the evolutionary capacity of stream
salmonids.
Most animals have the
 
Gresswell, S., et al. (1979). Mass movement response to forest management in the Central Oregon Coast Ranges, US Forest Service. 84: 1-26.
Gretener, P. E. (1967). "Signifigance of the rare event in geology." The American Association of Petroleum Geologists Bulletin 51(11): 2197-2206.
Griffiths, G. A. (1979). "Recent Sedimentation history of the Waimakariri River, New Zealand." Journal of Hydrology (New Zealand) 18: 6-28.
Griffiths, G. A. (1980). "Stochastic Estimation of Bed Load Yield in Pool-and-Riffle Mountain Streams." Water Resources Research 16(5): 931-937.
In many mountain streams the supply of gravel, transported as bed load, is predominantly infuenced by the stbility of a restricted area of riparian land. During stroms this zone contributes sediment to the stream through random mass movements and stream bank collapse with the result that floods, having similar hydrographs, produce different bed load yields. A stochastic model is developed which predicts total expected bed load yield from a given basin, in a prescribed interval, through the convolution of a Poisson number of events with an exponential probabiity of bed load yield. The exponential distribution is defined from regional flood frequency relations, an empirical bed load transport formula, and flow records from a nearby basin within the same hydrologic regime as the one under consideration; the expected number of events in an interval is also determined from these flow records. The model may be applied to small mountain watersheds, ungauged for either water or sediment outflows.
 
Griffiths, G. A. (1989). "Form Resistance in Gravel Channels with Mobile Beds " Journal of Hydraulic Engineering 115(3): 340-355.
Griffiths, G. A. (?). "Recent Sedimentation History of the Waimakariri River, New Zealand." 6-27.
Griffiths, G. A. and D. M. Hicks (1980). "Transport of sediment in mountain streams; performance of a measurement system during a two year storm (note)." Journal of Hydrology. New Zealand 19(2): 131-136.
Dry Acheron Stream, a small tributary of Rakaia River, drains a 21 km (super 2) area of eastern Southern Alps, New Zealand. Alpine and montane grassland, scrub, bare rock and scree occupy the mountainous portion of its catchment except for a 0.6 km (super 2) area of hardwood forest (Notofagus), near the basin centre. Altitude ranges from 650 to 1540 m and mean annual rainfall is approximately 2000 mm. Within the upper basin the stream channel is deeply entrenched, and is characterised by a pool-and-rifle geometry defined by channel boundaries of boulders and cobbles. Average stream gradient is 0.167. To provide quantitative assessments of sediment yields and transport patterns, a water and sediment discharge recording station was recently constructed just downstream of where the stream exits from its mountainous upper catchment (6 km (super 2) ). The station comprises three structures: (1) a weir to furnish a stable, rated, water discharge record and a convenient site to sample sediment load; (2) a debris-retention dam to trap most of the sediment load; and (3) a platform on the downstream side of the dam to allow sampling of material escaping through or over the dam. The purpose of this note is to describe the operation of the station and to present results from a recent flood which provided a test of the measuring system.
 
Griffiths, P. G., et al. (1996). Initiation and frequency of debris flows in Grand Canyon, Arizona. Tucson, AZ, U. S. Geological Survey: 40.
Griffiths, P. G., et al. (2004). "Frequency and initiation of debris flows in Grand Canyon, Arizona." Journal of Geophysical Research 109(doi:10.1029/2003JF000077): 14.
Griggs, G. B. (1987). The production, transport, and delivery of coarse-grained sediment by California's coastal streams. N. C. Kraus. New York, Am. Soc. Civ. Eng.
Grimm, M. M., et al. (1995). "Coarse-sediment deposition as evidence of an elevation limit for flash flooding, Bear Creek, Colorado." Geomorphology 14: 199-210.
Grimm, N. B. (1987). "Nitrogen dynamics during succession in a desert stream." Ecology 68: 1157-1170.
Griswold, J. and R. M. Iverson (2002). Inundation-area statistics and mobility equations for debris flows and rock avalanches. Geological Society of America, Cordilleran Section - 98th Annual Meeting. Corvallis, Oregon.
Inundation-area statistics have been used previously to establish equations that predict the maximum valley cross-sectional areas (A) and the total planimetric areas (B) inundated by lahars with various lahar volumes (V). The predictive equations obey 2/3 power laws and have the scale-invariant forms A=a1V2/3 and B=a2V2/3, where the a coefficients are calibrated using a database of lahar geometries around the world 1. Here, a new database is used to show that similar 2/3 power law equations apply to non-volcanic debris flows and rock avalanches. With recalibration of a coefficients, two new sets of equations are established to predict the cross-sectional and planimetric areas inundated by these types of rapid mass movements. These new equations provide an objective methodology for discriminating the mobilities of lahars, non-volcanic debris flows, and rock avalanches on the basis of a coefficients. Moreover, the mobility equations can be embedded in a GIS-based computer program, LAHARZ 1 , to generate hazard maps that depict inundation limits for debris flows and avalanches with a range of hypothetical volumes. In depicting hazards from comonplace flows or avalanches with volumes less than about one million cubic meters, satisfactory implemetation of LAHARZ requires high-accuracy, high-resolution digital topographic data. Such data are not widely available at present but will likely become more available in the future.
 
 
Grizzel, J. D. (2006). Geotechnical Assessment, "East Fork Tarboo" Timber Sale. Olympia, WA, Washington Department of Natural Resources: 22.
Grizzel, J. D. and N. Wolff (1998). "Occurrence of Windthrow in Forest Buffer Strips and its Effect on Small Streams in Northwest Washington." Northwest Science 72(3): 214-223.
Gross, M. R., et al. (1988). "Aquatic Productivity and Evolution of Diadromous Fish Migration." Science 239: 1291-1293.
Group, T. G. S. (1997). California Precipitation.
Group, T. G. S. (1997). "California Precipitation." from http://gis.ca.gov/casil/gis.ca.gov/teale/precipa/.
Grunwald, S. a. L. D. N. (1999). "An AGNPS-based runoff and sediment yield model for two small watersheds in Germany." Transactions of the ASAE 42(6): 1723-1731.
Guisan, A., et al. (2002). "Generalized linear and generalized additive models in studies of species distributions: setting the scene." Ecological Modelling 157: 89-100.
An important statistical development of the last 30 years has been the advance in regression analysis provided by
generalized linear models (GLMs) and generalized additive models (GAMs). Here we introduce a series of papers
prepared within the framework of an international workshop entitled: Advances in GLMs/GAMs modeling: from species
distribution to environmental management, held in Riederalp, Switzerland, 6 /11 August 2001.We first discuss some
general uses of statistical models in ecology, as well as provide a short review of several key examples of the use of
GLMs and GAMs in ecological modeling efforts. We next present an overview of GLMs and GAMs, and discuss some
of their related statistics used for predictor selection, model diagnostics, and evaluation. Included is a discussion of
several new approaches applicable to GLMs and GAMs, such as ridge regression, an alternative to stepwise selection of
predictors, and methods for the identification of interactions by a combined use of regression trees and several other
approaches. We close with an overview of the papers and how we feel they advance our understanding of their
application to ecological modeling.
 
Gunten, H. R. v. and R. N. Moser (1993). "How reliable is the Pb^210 dating method? Old and new results from Switzerland " Journal of Paleolimnology 9: 161-178.
Gupta, A. "Large Floods as Geomorphic Events in the Humid Tropics."
Gupta, S. K. and S. S. I. (1977). Distributed Numerical Model for Estimating Runoff and Sediment Discharge of Ungaged Rivers: 1. The Information System. Gupta and Solomon: Modeling Runoff and River Transport: 613-634.
Gupta, V. K. and E. Waymire (1989). "Statistical Self-Similarity in River Networks Parameterized by Elevation." Water Resources Research 25(3): 463-476.
Gupta, V. K., et al. (1980). "A representation of an instantaneous unit hydrograph from geomorphology." Water Resources Research 16(5): 855-862.
Gurnell, A. M. and M. J. Clark (1987). Glacio-Fluvial Sediment Transfer: An Alpine Perspective, John Wiley and Sons Ltd.
Gustafson, E. J. (1998). "Quantifying Landscape Spatial Pattern: What is the State of the Art?" Ecosystems 1: 143-156.
Gustafson, E. J. and G. R. Parker (1992). "Relationships between landcover proportion and indices of landscape spatial pattern." Landscape Ecology 7(2): 101-110.
Gustafson-Greenwood, I. and J. R. Moring (1990). "Territory size and distribution of newly-emerged Atlantic salmon (Salmo Salar) "? 206: 125-131.
Gustafson-Marjanen, K. I. and J. R. Moring (1984). "Construction of Artificial Redds for Evaluating Survival of Atlantic Salmon Eggs and Alevins " North American Journal of Fisheries Management 4: 455-461.
Guthrie, R. H. and S. G. Evans (2004). "Analysis of landslide frequencies and characteristics in a natural system, coastal British Columbia." Earth Surface Processes and Landforms 29: 1321-1339.
Guyette, R. P., et al. (2002). "Perspectives on the age and distribution of large wood in riparian carbon pools." Canadian Journal of Fisheries and Aquatic Sciences 59(3): 578-585.
Guzzetti, F., et al. (2000). "Comparing landslide maps: a case study in the Upper tiber Rivere Basin, Central Italy." Environmental Management 25: 247-263.
Guzzetti, F., et al. (1999). "Landslide hazard evaluation: a review of current techniques and their application in a multi-scale study, Central Italy." Geomorphology 31: 181-216.
Gyssels, G. and J. Poesen (2003). "The importance of plant root characteristics in controlling concentrated flow erosion rates." Earth Surface Processes and Landforms 28(4): 371-384.
While it has been demonstrated in numerous studies that the aboveground characteristics of the vegetation are of particular importance with respect to soil erosion control, this study argues the importance of separating the influence of vegetation on soil erosion rates into two parts: the impact of leaves and stems (aboveground biomass) and the influence of roots (belowground biomass). Although both plant parameters form inseparable constituents of the total plant organism, most studies attribute the impact of vegetation on soil erosion rates mainly to the characteristics of the aboveground biomass. This triggers the question whether the belowground biomass is of no or negligible importance with respect to soil erosion by concentrated flow. This study tried to answer this question by comparing cross-sectional areas of concentrated flow channels (rills and ephemeral gullies) in the Belgian Loess Belt for different cereal and grass plant densities. The results of these measurements highlighted the fact that both an increase in shoot density as well as an increase in root density resulted in an exponential decrease of concentrated flow erosion rates. Since protection of the soil surface in the early plant growth stages is crucial with respect to the reduction of water erosion rates, increasing the plant root density in the topsoil could be a viable erosion control strategy. Copyright (C) 2003 John Wiley Sons, Ltd.
 
Gyssels, G., et al. (2002). "The impact of sowing density of small grains on rill and ephemeral gully erosion in concentrated flow zones." Soil & Tillage Research 64(3-4): 189-201.
Despite the fact that soil erosion by water causes considerable on-site and off-site problems, farmers in Europe are reluctant to adopt prevention and control measures when such measures require additional labour and material inputs. This paper documents the impact of multiple sowing of small grains on concentrated flow erosion rates and grain production for a winter triticale field (X Triticosecale Wittmack ex. A. Camus) in the Belgian loess belt. Multiple sowing refers to drilling more than once in zones of concentrated flow erosion in order to increase the total root mass in this zone. Multiple sowing strongly altered the morphology of erosion channels and reduced soil loss significantly. Statistical analysis confirmed that differences in channel dimensions could be explained by the seedling density. Doubling the root mass in the topsoil by multiple sowing' resulted on average in a reduction of soil loss by 42% for the whole growing season. For the winter period, soil loss reduction, mainly attributed to the triticale roots, amounted even to 53%, showing the tremendous impact of seedling roots on soil erosion by concentrated overland flow in the early stages of vegetation growth. Furthermore, total grain yield in the multiple drilled zones was not significantly smaller compared to the conventionally drilled parts of the field. Grain size of cereals was slightly smaller. The results of this case study indicate that double sowing in concentrated overland flow zones may be a viable soil erosion control technique. (C) 2002 Elsevier Science B.V. All rights reserved.
 
Gyssels G, a. J. P. (2003). "The importance of plant root characteristics in controlling concentrated flow erosion rates." Earth Surface Processes and Landforms 28: 371-384.
H. M. Gucinski, H. B., M. J. Furniss, R. R. Ziemer (2001). "Forest roads: a synthesis of scientific information." U.S. Department of Agriculture, Forest Service, General Technical Report PNW-GRT-509, Portland Oregon: 103.
Haan, C. T. (2003). Analysis of Hydrologic Time Series. Statistical Methods in Hydrology, The Iowa State University Press: 1-13.
Hack, J. T. (1957). "Studies of longitudinal stream profiles in Virginia and Maryland." United States Geological Survey Professional Paper 294B.
Hack, J. T. (1957). Studies of longitudinal stream profiles in Virginia and Maryland, US Geological Survey: 97.
Hack, J. T. (1976). Dynamic equilibrium and landscape evolution. Theories of landform development: Publications in geomorphology. W. N. Melhorn and R. C. Flemal. Binghamton, N.Y., State University of New York: 87-102.
Hack, J. T. and J. C. Goodlett (1960). Geomorphology and Forest Ecology of a Mountain Region in the Central Appalachians, United States Geological Survey: 1-63.
Hack, J. T. and J. C. Goodlett (1960). Geomorphology and Forest ecology of a Mountain Region in the Central Appalachians, US Geological Survey Professional Paper 347: 66.
Hack, J. T. and J. C. Goodlett (1960). Geomorphology and forest ecology of a mountain region in the central Appalachians, US Geological Survey Professional Paper 347: 66.
Haga, H., et al. (2002). "Transport and retention of coarse woody debris in mountain streams: An in situ field experiment of log transport and a field survey of coarse woody debris distribution." Water Resources Research 38(8): 1126.
Although coarse woody debris (CWD) is an important component of stream
ecosystems in forested areas, the processes of CWD distribution, transport, and retention
have not been clarified. In this study the distribution process of CWD pieces shorter than
the bankfull width (S-CWD) is discussed using an in situ field experiment of log
transport and a field survey of CWD distribution in mountain streams. The transport
experiment showed that transport distance has a close relation to flow depth and also
implied that the magnitude and sequence of a series of flows were important factors for
S-CWD transport and retention in streams. The survey of CWD distribution indicated
that in-stream obstructions played an important role in the S-CWD retention in deeper
channels where S-CWD pieces were potentially transported distances more than spacing
between trapping sites of CWD. Overall, the in situ field experiment and the segment- to
reach-scaled analysis using h* (=depth/diameter) helped us understand the actual
movement and distribution of CWD.
 
Haggerty, R., et al. (2002). "Power-law residence time distribution in the hyporheic zone of a 2nd-order mountain stream." Geophysical Research Letters 29(13): 18-11 - 18-14.
Hairsine, P. "Comparing grass filter strips and near-natural riparian zones for trapping sediment and sorbed nutrients."
In relating hillslope sediment and sorbed pollutant fluxes to in-stream water quality, it is often assumed that vegetation on the lower portions of hillslope can act as buffer or sink by storing sediment and pollutant. The study reported here compares the sediment and sorbed phosphorus filtering capabilities of grass filters with those of near-natural riparian zones. Fluxes of water, sediment and sorbed nutrients were measured at the inlet and outlet of the grass filter strips, riparian zones and a combination of both, with a range of sediment fluxes and runoff rates from an intense upslope source-area. Aggregate size fractionation of the sediment load found that both the filter strips and riparian zones were effective in trapping sediment for a large part of the size band. Although riparian zones did have lower overall trapping efficiencies, they were capable of significantly reducing sorbed nutrient transport where the pollutant was carried by sediment in aggregate size classes greater than 20 mm. It was concluded that in this environment, a well-configured combination of a grass filter strip and riparian zone provides a very effective means of reducing the ingress of sediment and sorbed pollutants into a stream, in the short term.
 
Hairsine, P. B., J.C. Croke, H. Mathews, P. Fogarty, and S. P. Mockler (2002). "Modelling plumes of overland flow from logging tracks." Hydrological Processes 16(12): 2311-2327.
Most land-based forestry systems use extensive networks of unsealed tracks to access the timber resource. These tracks are normally drained by constructing cross-banks, or water bars, across the tracks immediately following logging. Cross-banks serve three functions in controlling sediment movement within forestry compartments:1. they define the specific catchment area of the snig track (also known as skid trails) so that the overland flow does not develop sufficient energy to cause gullies, and sheet and rill erosion is reduced;2. they induce some sediment deposition as flow velocity reduces at the cross-bank;3. they redirect overland flow into the adjacent general harvesting area (GHA) so that further sediment deposition may take place.This paper describes a simple model that predicts the third of these functions in which the rate of runoff from the track is combined with spatial attributes of the track and stream network. Predictions of the extent of the overland flow plumes and the volume of water delivered to streams is probabilistically presented for a range of rainfall-event scenarios with rainfall intensity, time since logging and compartment layout as model inputs. Generic equations guiding the trade-off between intercross-bank length and flow path distance from cross-bank outlet to the stream network needed for infiltration of track runoff are derived. Copyright (C) 2002 John Wiley Sons, Ltd.
 
Hairsine, P. B., et al. "Unsteady soil erosion due to rainfall impact: A model of sediment sorting on the hillslope."
A new method is presented for predicting sediment sorting associated with soil erosion by raindrop impact for non-equilibrium conditions. The form of soil erosion considered is that which results from raindrop impact in the presence of shallow overland flow itself where the flow is not capable of eroding sediment. The method specifically considers early time runoff and erosion when sediment leaving an eroding area is generally finer and thus may have a higher potential for transport of sorbed pollutants. The new mechanism described is the formation of a deposited layer on the soil surface, which is shown to lead to sediment sorting during an erosion event. The deposited layer is taken to have two roles in this process: to temporarily store sediment on the surface between successive trajectories, and to shield the underlying soil from erosive stresses. Equations describing the dynamics of the suspended sediment mixture and the deposited layer are developed. By integrating these equations over the length of eroding land element and over the duration of the erosion event, an event-based solution is proposed which predicts total sediment sorting over the event. This solution is shown to be consistent with experimentally observed trends in enrichment of fine sediment. Predictions using this approach are found to only partly explain measured enrichment for sets of experimental data for two quite different soils, but to be in poor agreement for an aridsol of dispersive character. It is concluded that the formation of the deposited layer is a significant mechanism in the enrichment of fine sediment and associated sorbed pollutants, but that processes in the dispersive soil are not as well described by the theory presented.
 
Halbert, C. L. (1993). "How Adaptive is Adaptive Management? Implementing Adaptive Management in Washington State and British Columbia." Reviews in Fisheries Science 1(3): 261-283.
Hales, Z. L., et al. (1970). "Riverbed Degradation Prediction." Water Resources Research 6(2): 549-556.
Hall, D. E., et al., Eds. (1994). Slope Stability Reference Guide for National Forests in the United States. Washington D.C., United States Department of Agriculture, Forest Service.
Hall, J. E., et al. (2007). "Predicting river floodplain and lateral channel migration for salmon habitat conservation." Journal of the American Water Resources Association 43(3): 786-797.
In this article, we describe a method for predicting floodplain locations and potential lateral channel
migration across 82,900 km (491 km2 by bankfull area) of streams in the Columbia River basin. Predictions
are based on channel confinement, channel slope, bankfull width, and bankfull depth derived from digital elevation
and precipitation data. Half of the 367 km2 (47,900 km by length) of low-gradient channels (£ 4% channel
slope) were classified as floodplain channels with a high likelihood of lateral channel migration (182 km2, 50%).
Classification agreement between modeled and field-measured floodplain confinement was 85% (j = 0.46,
p < 0.001) with the largest source of error being the misclassification of unconfined channels as confined (55%
omission error). Classification agreement between predicted channel migration and lateral migration determined
from aerial photographs was 76% (j = 0.53, p < 0.001) with the largest source of error being the misclassification
of laterally migrating channels as non-migrating (35% omission error). On average, more salmon populations
were associated with laterally migrating channels and floodplains than with confined or nonmigrating channels.
These data are useful for many river basin planning applications, including identification of land use impacts to
floodplain habitats and locations with restoration potential for listed salmonids or other species of concern.
 
Hallin, W. (1934). "Fast growing redwood." Journal of Forestry 32: 612-613.
Halwas, K., et al. (1997). Stream Hydroriparian Ecosystem Classification: Accurate Representation of Aribitrary Convenience? Vancouver, British Columbia, University of British Columbia: 1-40.
Ham, D. G. (1996). Patterns of Channel Change on Chilliwach River, British Columbia. Department of Geography, University of British Columbia.
Ham, D. G., and M. Church (2000). "Bed-material transport estimated from channel morphodynamics: Chilliwack River, British Columbia." Earth Surface Processes and Landforms 25: 1123-1142.
Hammond, C., et al. (1992). Level 1 Stability Analysis (LISA) Documentation for Version 2.0. Ogden, Utah, United States Department of Agriculture, Forest Service, Intermountain Research Station.
Hammond, H. (1983). Soil disturbance levels in ground skidded clearcuts in Southeastern British Columbia. New Forests for a Changing World, Portland, Oregon.
Hammond, H. L. (1988). Soil Degradation: Costs of Rehabilitation Versus Costs of  Prevention. Degradation of Forested Land: "Forest soils at risk", British Columbia Ministry of Forests, Land Management.
Hampton, M. A. (1975). "Competence of fine-grained debris flows." Journal of Sedimentary Petrology 45(4): 834-844.
Hancock, G. R. and K. G. Evans (2006). "Channel head location and characteristics using digital elevation models." Earth Surface Processes and Landforms 31: 809-824.
The drainage network is the conduit through which much surface water and sediment
are routed within a catchment. In a catchment, the position of where hillslopes begin and
channels end has long been considered the position of transition between diffusive processes
upslope and the more incisive fluvial processes downslope. Consequently, understanding
channel head location is an important issue in understanding catchment hydrology and
geomorphology. This study examines channel head position and characteristics in a catchment
in Arnhem Land, Northern Territory, Australia. In this study the position of channel
heads was mapped within the catchment and plotted on a reliable digital elevation model of
the catchment. It was found that the majority of channel heads have relatively small source
areas and that graphical catchment descriptors, such as the area–slope relationship and
cumulative area distribution, can provide reliable measures of the field position of the heads
of first-order streams and the transition from hillslope to channel. The area–slope relationship
and cumulative area distribution are also shown to be good tools for determining digital
elevation model grid size which can capture hillslope detail and the transition from hillslope
to channel.
 
Hancock, G. S., et al. (1999). "Dating fluvial terraces with ^10Be and ^26Al profiles: application to the Wind River Wyoming." Geomorphology 27: 41-60.
Hankin, S. (1984). Soil Strength Reinforcement form Distributions of Roots: A Mathematical Approach. ?, ?: 32.
Hankin, S. (1984). "Soil Strength Reinforcement from Distributions of Roots: A Mathematical Approach (DRAFT)." 32.
Hanks, T. C. (?). The age of scarpelike landforms from diffusion-equation analysis Dating and Earthquakes: Review of Quaternary Geochronology and its Applications to Paleoseismology: 497-535.
Hann, C. T. (1977). Statistical Methods in Hydrology. Ames, Iowa, Iowa State University Press.
Hansen, A. J., et al. (1991). "Conserving Biodiversity in Managed Forests: Lessons from natural forests." BioScience 41(6): 382-391.
Hansen, W. F. (2001). "Identifying stream types and management implications." Forest Ecology and Management 143: 39-46.
Characteristics of stream types were surveyed within the Chattooga River watershed in the Blue Ridge Mountains of southeastern US. The 728 km2 watershed is contained within the states of Georgia, South Carolina and North Carolina. The stream types of about 190 streams of various sizes were determined as perennial, intermittent or ephemeral using a combination of physical and biological indicators. The key indicators were defined as channel erosion into the landscape and aquatic insects. Comparisons were made between the measured stream characteristics and their hydrologic orders. The stream data indicated that most headwater order 1 streams were either ephemeral or intermittent. Order 2 streams were divided among intermittent, perennial and ephemeral. All order 3 streams and larger ones were perennial, except for one small, third order intermittent stream with an 18% channel gradient. The data suggest that the total 4670 km stream network was approximately equally divided between defined channels (28% perennial and 17% intermittent) and undefined channels (55% ephemeral). Changes in stream order produced exponential changes in the total number, mean segment length and total length of streams. The fine detail included about 15,630 order 1 channels with a combined length of 2740 km. Comparisons are provided between the estimated extent of stream types and the ‘blue-line’ streams identified on the US Geological Survey topographic contour maps at both 1:24,000 and 1:100,000 scales. Streams indicated on the 1:100,000 scale maps identified about 650 km of streams, while the 1:24,000 scale mapped the blue-line stream extent at 970 km. Of the estimated 1300 km of perennial streams, the topographic maps identified 50–75%, depending on scale. However, only about 14–21% of the entire stream network which includes perennial, intermittent and ephemeral streams were marked on the topographic contour maps. Recommendations are included for improving the consistency of how ‘stream orders’ and ‘stream types’ are delineated. Since water quality protection measures are often related to the stream type, improvement in identifying streams should improve stream protection design and help to prevent impacts on water quality and aquatic habitat.
 
Harcombe, P. A. (1987). "Tree Life Tables." BioScience 37(8): 557-568.
Harden, D. R. (1995). A comparison of flood-producing storms and their impacts in northwestern California. Geomorphic Processes and Aquatic Habitat in the Redwood Creek Basin, Northwestern California., US Geological Survey: D1 - D9.
Hare, P. W. and T. W. Gardner (1985). Geomorphic indicators of vertical neotectonism along converging plate margins, Nicoya Peninsula, Costa Rica. Tectonic Geomorphology. J. Hack and M. Morisawa: 75-104.
Hare, S. R., et al. (1999). "Inverse Production Regimes: Alaska and West Coast Pacific Salmon." Fisheries 24(1): 6-14.
Hargis, C., et al. (1998). "The behavior of landscape metrics commonly used in the study of habitat fragmentation." Landscape Ecology 13: 167-186.
Harmon, M. E. (1992). Long-term experiments on Log Decomposition at the H.J. Andrews Experimental Forest, U.S. Forest Service: 1-28.
Harmon, M. E., et al. (1986). "Ecology of Coarse Woody Debris in Temperate Ecoystems." Advances in Ecological Research 15: 133-302.
Harmon, M. E. and C. Hua (1991). "Coarse woody debris dynamics in two old-growth ecosystems." BioScience 41(9): 604-610.
Harr, R. D. (1977). "Water Flux in Soil and Subsoil on a Steep Forested Slope." Journal of Hydrology 33: 37-58.
Harr, R. D. (1979). Effects of Timber Harvest on Streamflow in the Rain-Dominated Portion of the Pacific Northwest. Workshop on Scheduling Timber Harvest for Hydrologic Concerns.
Harr, R. D. (1981). "Some Characteristics and Consequences of Snowmelt During Rainfall in Western Oregon." Journal of Hydrology 53: 277-304.
Harr, R. D. (1986). "Effects of Clearcutting on Rain-on-Snow Runoff in Western Oregon: A New Look at Old Studies." Water Resources Research 22(7): 1095-1100.
Harr, R. D. (1992). Flooding in West Cascade Mountain River Basins: A Perspective on the November 1990 Flooding of Western Washington. Seattle, Washington, Pacific Northwest Research Station.
Harr, R. D., et al. (1989). Effects of timber harvest on rain-on-snow runoff in the transient snow zone of the Washinton Cascades. Seattle, Washington, University of Washington: 29.
Harr, R. D. and T. W. Cundy (1992). The November 1990 floods in Western Washington, U.S.A. Interpraevent 1992, Bern.
Harr, R. D., et al. (1979). Changes in Streamflow Following Timber Harvest in Southwestern Oregon. Portland, Oregon, US Forest Service: 22.
Harr, R. D., et al. (1975). "Changes in Storm Hydrographs After Road Building and Clear-Cutting in the Oregon Coast Range." Water Resources Research 11(3): 436-444.
Harr, R. D., et al. (1982). "Streamflow Changes After Logging 130-year-Old Douglas Fir in Two Small Watersheds." Water Resources Research 18(3): 637-644.
Harrelson, C. C., et al. (1994). Stream Channel Reference Sites: An Illustrated Guide to Field Technique US Forest Service: 61.
Harris, D. D. (1977). Hydrologic Changes After Logging in Two Small Oregon Coastal Watersheds, U.S. Geological Survey: 31.
Harris, D. D., et al. (1979). Magnitude and frequency of floods in western Oregon. Portland, U.S. Geological Survey.
Harris, D. D. and R. C. Williams (1971). Streamflow, Sediment-Transport, and Water-Temperature Characteristics of Three Small Watersheds in the Alsea River Basin, Oregon, U. S. Department of the Interior.
Harris, R. M., et al. (1999). Upper Umatilla River sediment analysis. American Water Resources Association Technical Publication Series TPS, vol.99-3: 157-166.
The Umatilla River in Northeastern Oregon provides multiple beneficial uses, including native salmon habitat, water supply, and recreation. Water quality issues include developing TMDLs for sediment and temperature. The Forest Service operates three monitoring sites in the upper watershed with 34 years of streamflow records and 11 years of suspended sediment records. These stations were analyzed, with a focus on quantifying the annual and seasonal sediment loads and on the relationships between suspended sediment and discharge. We found: 1) high spatial and temporal variability in annual sediment loads between stations, and at the same station year to year, and 2) a lack of correlation between streamflow and sediment indicating complex streamflow and sediment supply relationships. Recommendations include future analysis to evaluate the reliability of automatic fixed point sampling by collecting concurrent depth integrated samples; future analysis to determine the frequency of sampling needed to characterize the sediment parameters; and, adding bedload sampling to determine bedload contribution to total load.
 
Hartley, C. (1958). Evaluation of wood decay in experimental work. Madison, Wisconsin, USDA Forest Service.
Hartman, G. F. and T. G. Brown (1987). "Use of Small Temporary, Floodplain Tributaries by Juvenile Salmonids in a West Coast Rain-Forest Drainage Basin, Carnation Creek, British Columbia " Canadian Journal of Fisheries and Aquatic Sciences 44: 262-270.
Hartman, G. F., et al. (1996). "Impacts of logging in Carnation Creek, a high-energy coastal stream in British Columbia, and their implication for restoring fish habitat." Canadian Journal of Fisheries and Aquatic Science 53(Suppl. 1): 237-251.
The land form, surficial geology, and hydrometeorology of the west coast of British Columbia cause streams in
the region to be highly variable in flow and vulnerable to land-use disturbance. Carnation Creek, a small drainage in this
region, was studied intensively for > 20 yr to examine the impacts of forest harvesting. Landslides and debris torrents
modified steep slope tributaries and the mainstem of the creek. Bank erosion also altered the stream channel on the alluvial
flood plain. These effects were additive in the system and reduced the quality of spawning and rearing habitat for juvenile
salmonids. In streams like Carnation Creek, it is necessary to restore some stability to the hill slopes and gullies before
attempting fish habitat improvements in the main channel. Salmonid production was limited by combinations of processes
and conditions that were different for each species and life-history stage. Knowledge of the processes that limit fish
production must be applied in habitat improvement work or the projects risk failure. Programs intended to restore natural
function to systems or to improve habitat for fish must be planned, evaluated, and reported methodically if they are to
succeed and provide information of use to future programs.
 
Harvey, A. M. (1991). "The influence of sediment supply on the channel morphology of upland streams: Howgill Fells, Northwest England." Earth Surface Processes and Landforms 16: 675-684.
Previous work on stream channels in upland areas of Britain has demonstrated a close control over channel morphology and stability by the rate of coars sediment supply form the hillslopes of the chatchment. Streams fed by large amounts of coarse sediment develop unstable, wide, often braided channels, whereas those with limited coarse sediment supply develop stable, much narrower, often meandering channels. The sediment supply from hillslopes is controlled by thresholds of hillslope stability, storm event frequency, and the coupling between the hillslopes and the channel. Climatically induced changes in any of these three factors may have implications for channel morphology and stability. This paper examines these implications in British upland fluvial systems, with particular reference to the Howgill Fells, Cumbria, in the contexts of the adjuxtment of stream channels to sediment supply from erosional gully systems, and their response to and recovery from major flood events.
 
Harvey, A. M. (1997). The role of alluvial fans in arid zone fluvial systems. Arid Zone Geomorphology. D. S. G. Thomas. Chichester, Wiley: 231-259.
Harvey, B. C. and T. E. Lisle (1999). "Scour of Chinook Salmon Redds on Suction Dredge Tailings " North American Journal of Fisheries Management 19: 613-617.
Harvey, B. C. and R. J. Nakamoto (1996). "Effects of Steelhead Density on Growth of Coho Salmon in a Small Coastal California Stream " Transactions of the American Fisheries Society 125: 237-243.
Harvey, B. C., et al. (2002). "Habitat relationships and larval drift of native and nonindigeneous fishes in neighboring tributaries of a coastal California river." Transactions of the American Fisheries Society 131: 159-170.
Motivated by a particular interest in the distribution of the nonindigenous, piscivorous Sacramento pikeminnow Ptychocheilus grandis, we examined fish–habitat relationships in small tributaries (draining 20–200 km2) in the Eel River drainage of northwestern California. We sampled juvenile and adult fish in 15 tributaries in both the summer and fall of 1995 and attempted to relate the densities of the most abundant species to physical variables. To determine which species used small tributaries for spawning, we also collected drifting larval fish during the spring of 1996 and 1997. Water temperature, as measured by maximum weekly average temperature, dominated the relationships between physical variables and the densities of age-0 Sacramento pikeminnow, age-0 steelhead Oncorhynchus mykiss, California roach Hesperoleucus symmetricus (also known as Lavinia symmetricus), and Sacramento sucker Catostomus occidentalis. Of these groups, only age-0 steelhead were most abundant in cool tributaries. In contrast to results for these groups, temperature regime, instream cover, summer discharge, and water depth contributed approximately equally to the best-fitting models of post-age-0 steelhead abundance. Drift samples revealed widespread use of tributaries for reproduction by both native species and the nonindigenous California roach. We also found drifting larval Sacramento pikeminnows in five streams that ranged widely in size. Temperature regimes in many Eel River tributaries have been affected by both human activities and large floods over the last 50 years. This study suggests that (1) these changes in temperature regime enhanced invasion of the drainage by nonindigenous fishes and (2) management efforts that alter temperature regimes in Eel River tributaries will have significant consequences for the composition of fish assemblages in general and for the effects of Sacramento pikeminnow in particular.
 
Harvey, J. W. and K. E. Bencala (1993). "The effect of streambed topography on surface-subsurface water exchange in mountain catchments." Water Resources Research 29(1): 89-98.
Harvey, M. D., et al. (1992). Estimating sediment delivery and yield on alluvial fans. Special Publication - Natural Hazards Research and Applications Information Center, vol.24. Boulder, Natural Hazards Research and Applications Information Center: 238.
Harvey, M. D. and C. C. Watson (1986). "Fluvial Processes and Morphological Thresholds in Incised Channel Restoration." Water Resources Bulletin 22(3): 359-368.
Haschenburger, J. K. (1996). Scour and fill in a gravel-bed channel: observations and stochastic models. Department of Geography, University of British Columbia.
Haschenburger, J. K. (1996). Scour and fill in a gravel-bed channel: observations and stochiastic models Department of Geography, University of British Columbia: 144.
Haschenburger, J. K., and M. Church (1998). "Bed material transport estimated from the virtual velocity of sediment." Earth Surface Processes and Landforms 23: 791-808.
Haschenburger, J. K. (1998). Channel scour and fill in coastal streams. Carnation Creek and Queen Charlotte Islands Fish/Forestry Workshop: Applying 20 years of Coastal Research to management Solutions: Land Management Handbook 41. D. L. Hogan, P. J. Tschaplinski and S. Chatwin. Victoria, B.C., B.C. Ministry of Forests: 109-117.
Haschenburger, J. K. (1999). "A probability model of scour and fill depths in gravel-bed channels." Water Resources Research 35(9): 2857-2869.
Haschenburger, J. K. and M. Church (1998). "Bed material transport estimated from the virtual velocity of sediment." Earth Surface Processes and Landforms 23: 791-808.
Hassan, M. A. (1990). "Scour, fill, and burial depth of coarse material in gravel bed streams " Earth Surface Processes and Landforms 15: 341-356.
Hassan, M. A., M. Church, T.E. Lisle, F. Brardinoni, L. Benda, G.E. Grant (2005). "Sediment transport and channel morphology in small forested streams." Journal of the American Water Resources Association 41(4): 853-876.
Hassan, M. A. (?). Sediment transport, channel morphology and dynamics of small-forested streams. Vancouver, B.C., University of British Columbia: 1-34.
Hassan, M. A. and M. Church (1992). The movement of individual grains on the streambed. Dynamics of Gravel-bed Rivers. P. Billi, R. D. Hey, C. R. Thorne and P. Tacconi, John Wiley & Sons Ltd.: 159-175.
Hassan, M. A. and M. Church (1994). "Vertical mixing of coarse particles in gravel bed rivers: A kinematic model." Water  Resources Research 30(4): 1173-1185.
Hassan, M. A., et al. (1992). "Virtual rate and mean distance of travel of individual clats in gravel-bed channels." Earth Surface Processes and Landforms 17: 617-627.
Travel distances in gravel-bed rivers of tagged particles of various sizes are related to excess stream power estimated for peak discharge. Mean distance of movement, irrespective of grain size, is weakly correlated with stream power, especially near the threshold of movement. There may be several reasons for the weak correlation, including variable effects of bed structure, varying magnitudes of sediment mobilizing events and sampling problems. Grain size itself is of marginal significance. Thevirtual rate of travel calculated using total time for which the flow is larger than that needed to initiate clast movement also bears a weak relation to the excess stream powere over the period. Better results are obtained by relating the virtual rate of travel to the first peak of the flow event only. This implies that the initial seeding of the tagged particles dominates the observations. Nonetheless, an underlying general relation is indicated by the data, which are derived from a wide range of flow regime types.
 
Hassan, M. A., et al. (2005). "Sediment transport and channel morphology of small, forested streams." Journal of the American Water Resources Association 41: 853-876.
Hassan, M. A., et al. (1991). "Distance of Movement of Coarse Particles in Gravel Bed Streams." Water Resources Research 27(4): 503-511.
Hassan, M. A. and R. Egozi (2001). "Impact of wastewater discharge on the channel morphology of ephemeral streams." Earth Surface Processes and Landforms 26: 1285-1302.
Hassan, M. A., et al. (2005). "Spatial and temporal dynamics of wood in headwater streams of the Pacific Northwest." Journal of the American Water Resources Association 41(4): 899-919.
This paper synthesizes information on the spatial and
temporal dynamics of wood in small streams in the Pacific Northwest
region of North America. The literature on this topic is somewhat
confused due to a lack of an accepted definition of what
constitutes “small” streams and what is the relative size of woody
debris contained within the channel. This paper presents a matrix
that defines woody debris relative to channel size and then discusses
the components of a wood budget. Headwater streams are in
close proximity to wood sources and, in steeplands, are often tightly
constrained by steep hillslopes. Special consideration is given to
ecosystem characteristics and to management practices that affect
the wood dynamics in this context. Knowledge gaps and uncertainties
that can be used to guide future research are identified. Very
little is currently known about the role of mass wasting in wood
recruitment and storage relative to other processes, such as bank
erosion and mortality, in larger streams. Further, very little work
has addressed the relative importance of different wood depletion
processes, especially those associated with wood transport. The
effect of other ecosystem variables on wood dynamics locally across
a watershed (from valley bottom to mountaintop) and regionally
across the landscape (from maritime to continental climates) is not
addressed. Finally, the scientific community has only begun to deal
with the effects of management practices on wood quantity, structure,
and movement in small streams
 
Hassan, M. A. and M. Klein (2002). "Fluvial adjustment of the Lower Jordan River to a drop in the Dead Sea Level." Geomorphology 45: 21-33.
Hassen, M. A. and M. Church (1992). "Virtual rate and mean distance of travel of individual clasts in gravel-bed channels." Earth Surface Processes and Landforms 17: 617-627.
Hatten, J. (1991). The Effects of Debris Torrents on Spawning Gravel Quality in the Tributary Basins and Side-channels of the Hoh River, Washington (DRAFT). Forks, Washington, Hoh Indian Tribe: 19.
Hauer, F. R., et al. (1999). "Large woody debris in bull trout (Salvelinus confluentus) spawning streams of logged and wilderness watersheds in northwest Montana." Canadian Journal of Fisheries and Aquatic Science 56: 915-924.
We measured large woody debris (LWD) in 20 known bull trout (Salvelinus confluentus) spawning stream
reaches from logged and wilderness watersheds in northwestern Montana. Mean bankfull width of stream reaches was
14.1 m ranging from 3.9 to 36.7 m. Streams were large enough to move LWD and form aggregates. We determined
the characteristics of individual pieces of LWD that were interactive with the stream channel. Large, short pieces of
LWD attached to the stream bank were the most likely to be positioned perpendicular to stream flow, while large, long
pieces either tended to be parallel to the flow or, when attached, were most apt to extend across the channel thalweg.
Observations indicated that the majority of pools were formed as scour pools by either very large LWD pieces that
were perpendicular to the stream or multipiece LWD aggregates. Among reaches in wilderness watersheds, ratios of
large to small LWD, attached to unattached LWD, and with and without rootwads were relatively consistent. However,
among reaches with logging in the watershed, these ratios varied substantially. These results suggest that logging can
alter the complex balance of delivery, storage, and transport of LWD in northern Rocky Mountain streams, and
therefore, the likely substantive change in stream habitats.
 
Haugerud, R. A. and Anonymous (2002). "High-resolution public-domain topography for western Washington; the Puget Sound Lidar Consortium." Abstracts with Programs - Geological Society of America 34(5): 31.
Haugerud, R. A., et al. (2003). "High-resolution Lidar topography of the Puget Lowland, Washington - A bonanza for Earth Science." GSA Today(6): 4-10.
Havis (1993). "A Mathematical Model of Salmonid Spawning Habitat " Water Resources Bulletin 29(3): 435-444.
Hawkins, C. P. (1993). "A Hierarchical Approach to Classifying Stream Habitat Features " Fisheries 18(6): 3-12.
Hawkins, C. P., et al. (1983). "Density of Fish and Salamanders in Relation to Riparian Canopy and Physical Habitat in Streams of the Northwestern United States." Canadian Journal of Fisheries and Aquatic Sciences 40(8): 1173-1185.
Hay, L. E., R.L. Wilby, and G. H. Leavesley (2000). "A comparison of delta change and downscaled GCM scenarios for three mountainous basins in the United States." Journal of the American Water Resources Association 36(2): 387-397.
Hays, W. H. and R. L. Schuster (1987). Maps Showing Ground-Failure Hazards in the Columbia River Valley Between Richland and Priest Rapids Dam, South-Central Washington.
He, C. (2003). "Integration of geographic information systems and simulation model for watershed management." Environmental Modelling & Software 18(8-9): 809-813.
He, H. S. (1999). "Spatially Explicit and stochastic simulation of forest-landscape fire disturbance and succession." Retrieved 3-14-02, 2002, from http://www.findarticles.com/cf_dls/m2120/1_80/53870314/print.jhtml.
He, Q. and D. E. Walling (2003). "Testing distributed soil erosion and sediment delivery models using Cs-137 measurements." Hydrological Processes 17(5): 901-916.
Healey, M. C. (2009). "Resilient salmon, resilient fisheries for British Columbia, Canada." Ecology and Society 14(1).
Salmon are inherently resilient species. However, this resiliency has been undermined in British Columbia by a century of centralized, command-and-control management focused initially on maximizing yield and, more recently, on economic efficiency. Community and cultural resiliency have also been undermined, especially by the recent emphasis on economic efficiency, which has concentrated access in the hands of a few and has disenfranchised fishery-dependent communities. Recent declines in both salmon stocks and salmon prices have revealed the systemic failure of the current management system. If salmon and their fisheries are to become viable again, radically new management policies are needed. For the salmon species, the emphasis must shift from maximizing yield to restoring resilience; for salmon fisheries, the emphasis must shift from maximizing economic efficiency to maximizing community and cultural resilience. For the species, an approach is needed that integrates harvest management, habitat management, and habitat enhancement to sustain and enhance resilience. This is best achieved by giving fishing and aboriginal communities greater responsibility and authority to manage the fisheries on which they depend. Co-management arrangements that involve cooperative ownership of major multistock resources like the Fraser River and Skeena River fisheries and community-based quota management of smaller fisheries provide ways to put species conservation much more directly in the hands of the communities most dependent on the well-being and resilience of these fisheries.
 
Heaton, T. H. and S. H. Hartzell (1987). "Earthquake Hazards on the Cascadia Subduction Zone." Science 236: 162-168.
Hedger, R. D., et al. (2006). "Improving models of juvenile Atlantic salmon habitat use through high resolution remote sensing." Ecological Modelling 197: 505-511.
The effect of spatial variation in substrate size on juvenile Atlantic salmon density in the
Sainte-Marguerite, a major salmon-spawning river in QC, Canada, was investigated using
ground surveys and remote sensing. Densities of both fry and parr were determined by
single-pass electro-fishing at up to 48 sites along the length of the river in August of each
of the years from 1997 to 2004. Substrate size was determined by two techniques. Firstly, a
ground survey of median substrate size (D50) was conducted concurrently with the electrofishing,
providing information on the substrate actually experienced by the salmon at the
point of capture (a parcel area of 5m×20m for each site). Secondly, an airborne survey was
conducted along the length of the river in August 2002 using a helicopter-mounted digital
camera. Images from this survey were then processed to show the mean D50 in the area of
each image (an area from bank to bank, 60m in length along the river) encompassing each
parcel. Relationships between juvenile salmon density and substrate size were determined
using preference models where the model partitioning had been determined by regression
tree analysis. For comparison, preference models were also produced where themodel partitioning
was determined arbitrarily. The shape of the relationships between juvenile salmon
density and parcel D50 were similar to those between juvenile salmon density and mean
image D50. However, the relationship was stronger with mean image D50, suggesting that
the habitat surrounding the location in which juvenile salmon were found had a direct
influence on their density. Additionally, preference models determined from regression tree
analysis had greater explanatory power than those determined using an arbitrary partitioning
approach. These results suggest that remote sensing, alongside a robust approach
for determining preference models, may be an effective tool in modelling juvenile Atlantic
salmon habitat use.
 
Hedman, C. W., et al. (1996). "In-stream large woody debris loading and riparian forest seral stage associations in the southern Appalachian Mountains." Canadian Journal of Forest Research 26: 1218-1227.
Large woody debris (LWD) is an important ecological component of mountain streams. However,
the relation of LWD loading and riparian forest composition is poorly understood in the southern Appalachians.
In this study, 500-m reaches of 11 riparian forest-stream systems representing a 300-year sere were inventoried
and measured to obtain quantitative estimates and descriptions of in-stream LWD. Loading volumes ranged from
7.1 to 31.2 m3/100 m of stream, or between 3.6 and 13.2 kg/m2. LWD loadings were highly variable during
midseral stages of plant community succession, primarily because of the wide range in loading of American
chestnut (Castanea dentata (Marsh.) Borkh.). Loadings increased linearly in late-successional through old-growth
systems over a 165-year interval. Eastern hemlock (Tsuga canadensis (L.) Carriere) and American chestnut were
the most dominant carry-over LWD species in midsuccessional stream systems. Loading of eastern hemlock
LWD increased from midsuccessional through old-growth stages as the species became dominant in the riparian
forest. Without carry-over debris, LWD loadings would be extremely low in midsuccessional stream systems.
American chestnut was a major component of LWD in midsuccessional stream systems, despite the fact that it
has been unavailable for recruitment for decades.
 
Hedman, C. W., et al. (1996). "In-stream large woody debris loading and riparian forest seral stage associations in the southern Appalachian Mountains." Canadian Journal of Forest Research 26: 1218-1227.
Heede, B. H. (1972). Flow and Channel Characteristics of Two High Mountain Streams, USDA Forest Service: 1-12.
Heede, B. H. (1972). "Influences of a forest on the hydraulic geometry of two mountain streams." Water Resources Bulletin 8(3): 523-530.
Heede, B. H. (1985). "Channel adjustments to the removal of log steps: an experiment in a mountain stream." Environmental Management 9(5): 427-432.
Heede, B. H. (1987). "Overland flow and sediment delivery five years after timber harvest in a mixed conifer forest, Arizona, U.S.A." Journal of Hydrology 91(3-4): 205-216.
Heede, B. H., et al. (1988). "Sediment Delivery Linkages in a Chaparral Watershed Following a Wildfire." Environmental Management 12(3): 349-358.
Hegge, B. J. and G. Masselink (1996). "Spectral analysis of geomorphic time series: auto-spectrum." Earth Surface Processes and Landforms 21: 1021-1040.
Heggenes, J., et al. (1990). "Comparison of Three Methods for Studies of Stream Habitat Use by Young Brown Trout and Atlantic Salmon " Transactions of the American Fisheries Society 119: 101-111.
Heikurainen, L. "Effect of cutting on the ground-water level on drained peatlands." 345-354.
Heimann, D. C. (1988). Recruitment trends and physical characteristics of large organic debris in Oregon Coast Range streams. Department of Forest Science. Corvallis, OR, Oregon State University.
Heimsath, A. M. and Anonymous (2001). Sediment production and transport on hilly landscapes. Abstracts with Programs - Geological Society of America. 33: 285.
The production and transport of sediment is the focus of extensive study across diverse landscapes. Landscape form has long been recognized to be shaped by processes of erosion, and model quantification of how landscape form may change due to external forcing depends on the laws chosen to characterize the dominant erosion. These laws are rarely based on field data, or direct observation, but are used widely to drive landscape evolution models. Here I focus on hilly soil mantled landscapes to demonstrate how two relatively new tools can be coupled and used to quantify erosion rates and mechanisms. Soil production rates are quantified using the in situ produced cosmogenic radionuclides, (super 10) Be and (super 26) Al, extracted from bedrock beneath the soil mantle. Nuclide concentrations are used to infer both local and catchment-averaged (when measured from stream sediments) erosion rates as extensive studies have demonstrated elsewhere. Hillslope sediment transport processes have been quantified by other studies using segmented rods, short-lived isotopes (under very constrained conditions), and fallout (super 10) Be. I show how both short-lived isotopes and a new method of using optically stimulated luminescence can be used where soil production or erosion rates are known to determine the dominant transport mechanisms. Specific examples from southeastern Australia (using OSL) and the Hubbard Brook experimental watershed in New Hampshire (using the short-lived isotopes (super 210) Pb, (super 7) Be, and (super 137) Cs) show how thoroughly the soil profiles are bioturbated and emphasize the role of overland flow in shaping the landscape. A Monte Carlo model helps connect the nuclide results with the transport results to suggest that grain movements are independent and random. The model also suggests that 20% of the transported soil was removed by overland flow over roughly the last 10 ka. This helps explain the divergence of field observations from a prediction of slope morphology and soil depth based on a simple diffusion model. These field results further support a new transport model that combined expressions for simple creep, depth-dependent creep and overland flow to predict soil thickness and suggest how a landscape evolves in response to climatic and tectonic forces.
 
Heimsath, A. M., et al. "Stochastic processes of soil production and transport: Erosion rates, topographic variation and cosmogenic nuclides in the Oregon Coast Range."
Landscapes in areas of active uplift and erosion can only remain soil-mantled if the local production of soil equals or exceeds the local erosion rate. The soil production rate varies with soil depth, hence local variation in soil depth may provide clues about spatial variation in erosion rates. If uplift and the consequent erosion rates are sufficiently uniform in space and time, then there will be tendency toward equilibrium landforms shaped by the erosional processes. Soil mantle thickness would adjust such that soil production matched the erosion. Previous work in the Oregon Coast Range suggested that there may be a tendency locally toward equilibrium between hillslope erosion and sediment yield. Here results from a new methodology based on cosmogenic radionuclide accumulation in bedrock minerals at the base of the soil column are reported. We quantify how soil production varies with soil thickness in the southern Oregon Coast Range and explore further the issue of landscape equilibrium. Apparent soil production is determined to be an inverse exponential function of soil depth, with a maximum inferred production rate of 268 m Ma super(-1) occurring under zero soil depth. This rate depends, however, on the degree of weathering of the underlying bedrock. The stochastic and large-scale nature of soil production by biogenic processes leads to large temporal and spatial variations in soil depth; the spatial variation of soil depth neither supports nor rejects equilibrium morphology. Our observed catchment-averaged erosion rate of 117 m Ma super(-1) is, however, similar to that estimated for the region by others, and to soil production rates under thin and intermediate soils typical for the steep ridges. We suggest that portions of the Oregon Coast Range may be eroding at roughly the same rate, but that local competition between drainage networks and episodic erosional events leads to landforms that are out of equilibrium locally and have a spatially varying soil mantle.
 
Heimsath, A. M., et al. (1999). "Cosmogenic nuclides topography and the spatial variation of soil depth." Geomorphology 27: 151-172.
Heimsath, A. M., et al. (2001). "Stochastic processes of soil production and transport: erosion rates, topographic variation and cosmogenic nuclides in the Oregon Coast Range." Earth Surface Processes and Landforms 26: 531-552.
Landscapes in areas of active uplift and erosion can only remain soil-mantled if the local production of soil equals or exceeds
the local erosion rate. The soil production rate varies with soil depth, hence local variation in soil depth may provide clues
about spatial variation in erosion rates. If uplift and the consequent erosion rates are sufficiently uniform in space and time,
then there will be tendency toward equilibrium landforms shaped by the erosional processes. Soil mantle thickness would
adjust such that soil production matched the erosion. Previous workin the Oregon Coast Range suggested that theremaybe a
tendency locally toward equilibrium between hillslope erosion and sediment yield. Here results from a new methodology
basedoncosmogenic radionuclide accumulation in bedrock minerals at the base of the soil column are reported.Wequantify
howsoil production varies with soil thickness in the southern Oregon Coast Range and explore further the issue of landscape
equilibrium. Apparent soil production is determined to be an inverse exponential function of soil depth, with a maximum
inferred production rate of 268 m Ma 1 occurring under zero soil depth. This rate depends, however, on the degree of
weathering of the underlying bedrock.The stochastic and large-scale nature of soil production bybiogenic processes leads to
large temporal and spatial variations in soil depth; the spatial variation of soil depth neither supports nor rejects equilibrium
morphology. Our observed catchment-averaged erosion rate of 117 m Ma 1 is, however, similar to that estimated for the
region by others, and to soil production rates under thin and intermediate soils typical for the steep ridges. We suggest that
portions of the Oregon Coast Range may be eroding at roughly the same rate, but that local competition between drainage
networks and episodic erosional events leads to landforms that are out of equilibrium locally and have a spatially varying soil
mantle.
 
Heine, R. A., et al. (2004). "Development and comparison of approaches for automated mapping of stream channel networks." Annals of the Association of American Geographers 94(3): 477-490.
Accurate mapping of stream channel networks is important for measuring hydrologic parameters, for site planning
in construction projects, and for use in hydrologic models. This article compares five existing and two new
methods for extracting stream channel networks for use in topographic mapping. In order of increasing accuracy,
these methods are: (1) blue lines on USGS 1:24,000 topographic maps (64.6 percent underrepresentation), (2)
placing stream heads using a constant flow-accumulation area to mimic USGS blue lines (47.8 percent underrepresentation),
(3) constant flow-accumulation area equal to the mean for identified channel heads (30.3 percent
combined under- and overrepresentation), (4) variable flow-accumulation area estimated by multiple linear
regression (28.9 percent combined under- and overrepresentation), (5) variable flow-accumulation area estimated
by a slope-power relationship (23.6 percent combined under- and overrepresentation), (6) identifying
stream cells using logistic regression (12.7 percent combined under- and overrepresentation), and (7) extracting
stream channel head locations from digital orthophotoquads (DOQs) (nearly 100 percent accurate, but only
applicable under ideal conditions). Methods 2–6 require 10 m resolution digital elevation models that can be
acquired directly in many areas or can be derived from 1:24,000 hypsography where available; Methods 4 and 6
are new methods developed in this paper.
Using DOQs, while extremely accurate, is labor intensive and can be applied only in a small minority of
locations where vegetation cover does not obscure channel head locations. We conclude that identifying stream
cells using logistic regression has the broadest applicability because it can be implemented in an automated
fashion using only DEMs while still achieving accuracies for mapping low-order streams that are far superior to
existing USGS maps.
 
 
Heller, P. (1981). "Small Landslide Types and Controls in Glacial Deposits: Lower Skagit River Drainage, Northern Cascade Range, Washington." Environmental Geology 3: 221-228.
Helley, E. J. (?). "The Red Bluff Pediment - A Datum Plane for locating Quaternary Structures in the Sacramento Valley, California "?: 1-6.
Helmlinger, K. R., et al. (1993). "On the use of digital elevation model data for hortonian and fractal analyses of channel networks." Water Resources Research 29(8): 2599-2613.
Hemphill-Haley, M. A., et al. (1998). Paleoseismicity of the Alvord fault, Steens Mountain, Southeastern Oregon Dating and Earthquakes: Review of Quaternary Geochronology an Its Application to Paleoseismology (Bigelow). J. M. Sowers, J. S. Noller and W. R. Lettis: 89-95.
Hemstrom, M. A. and J. F. Franklin (1982). "Fire and Other Disturbances of the Forests in Mount Rainier National Park." Quaternary Research 18: 32-51.
Henderson, F. M., T.F. Hart Jr., L. Orlando, B. Heaton, J. Portolese, and R. Chasan (1998). "Application of C-CAP protocol land cover data to nonpoint source water pollution potential spatial models in a coastal environment." Photogrammetric Engineering & Remote Sensing 64(10): 1015-1020.
Henderson, J. A. and D. Peter (1983). Preliminary Plant Associations and Habitat Types of the Hoodsport and Quilcene Ranger Districts Olympic National Forest, US Forest Service: 2-25.
Henkin, S. (1984). Soil Strength Reinforcement from Distributions of Roots: A Mathematical Approach: 32.
Hennon, P. E., et al. (1999). Comparing deterioration and ecosystem function of decay-resistant and decay-susceptible species of dead trees. Symposium on the ecology and management of dead wood in western forests, Reno, Nevada.
Hennon, P. E. and M. H. McClellan (2003). "Tree mortality and forest structure in the temperate rain forests of southeast Alaska." Canadian Journal of Forest Research 33: 1621-1634.
Heppner, C. S., Q. Ran, J.E. VanderKwaak, and K. Loague (2006). "Adding sediment transport to the integrated hydrology model (InHM): development and testing." Advances in Water Resources 29: 930-943.
  model, sediment yield, erosion, plot-scale, calibration, validation, sensitivity analysis
 
Hergarten, S. and H. J. Neugebauer (1998). "Self-organized criticality in a landslide model." Geophysical Research Letters 25(6): 801-804.
From landslide mapping it is known that the
frequency of landslide occurence as a function of their magnitude
can be described by a power law in many regions.
In order to investigate the magnitude distribution of landslides
from a theoretical point of view, we present a physically
based landslide model combining aspects of slope stability
and mass movement. If the long term driving processes
(fluvial or tectonic) are integrated, the model shows
self{organized criticality (SOC). The results coincide with
results obtained from landslide mapping, so that our model
suggests that landsliding may be seen as a SOC process.
In contrast to other models showing SOC that are mostly
based on cellular automata, our model is based on partial
dierential equations. The results show that SOC is not a
fashion of cellular automata, but can also occur in di
erential equation models.
 
Hermanns, R. L., et al. (2001). "Neotectonics and catastrophic failure of mountain fronts in the southern intra-Andean Puna Plateau, Argentina." Geological Society of America 29(7): 619-623.
Hernandez, M., et al. (2000). "Modeling runoff response to land cover and rainfall spatial variability in semi-arid watersheds." Environmental Monitoring and Assessment 64(1): 285-298.
This paper describes a procedure for evaluating the effects of land cover change and rainfall spatial variability on watershed response. Two hydrological models were applied on a small semiarid watershed in Arizona, USA; one model is event-based with a one-minute time step (KINEROS), and the second is a continuous model with a daily time step (SWAT). The inputs to the models were derived from Geographic Information System (GIS) theme layers of USGS digital elevation models, the State Soil Geographic Database (STATSGO) and the Landsat-based North American Landscape Characterization classification (NALC) in conjunction with available literature and look up tables. Rainfall data from a network of 10 raingauges and historical stream flow data were used to calibrate runoff depth using the continuous hydrologic model from 1966 to 1974. No calibration was carried out for the event-based model, in which six storms from the same period were used in the calculation of runoff depth and peak runoff. The assumption on which much of this study is based is that land cover change and rainfall spatial variability affect the rainfall-runoff relationships on the watershed. To validate this assumption, simulations were carried out wherein the entire watershed was transformed from the 1972 NALC land cover, which consisted of a mixture of desertscrub and grassland, to a single uniform land cover type such as riparian, forest, oak woodland, mesquite woodland, desertscrub, grassland, urban, agriculture, and barren. This study demonstrates the feasibility of using widely available data sets for parameterizing hydrological simulation models. The simulation results show that both models were able to characterize the runoff response of the watershed due to changes of land cover.
 
Herrera Environmental Consultants Inc. (2004). Review of the available literature related to wood loading dynamics in and around streams in Eastern Washington forests. Cooperative monitoring Evaluation and Research 03-308.
Hessburg, P. F., et al. (1999). Using Estimates of Natural Variation to Detect Ecologically Important Change in Forest Spatial Patterns: A Case Study, Cascade Range, Eastern Washington, US Department of Agriculture
Forest Service
Pacific Northwest Research Station: 1-65.
Heusser, C. J. (1974). "Quaternary vegetation, climate and glaciation in the Hoh River valley, Washington." Geological Society of America Bulletin 85: 1547-1560.
Heusser, C. J. (1977). "Quaternary palynology of the Pacific slope of Washington." Quaternary Research 8: 282-306.
Heuvelink, G. B. M. (1998). "Uncertainty analysis in environmental modelling under a change of spatial scale." Nutrient Cycling in Agroecosystems 50(255-264).
Hewawasam, T.
Hewawasam, T., et al. (2003). "Increase of human over natural erosion rates in tropical highlands constrained by cosmogenic nuclides." Geology 31(7): 597-600.
We quantify the difference between the human-caused sediment yield and the natural rates of soil production and bedrock erosion in a now largely deforested tropical highland. The present-day rate of soil loss in the Upper Mahaweli catchment, Sri Lanka, is calculated by using suspended river-load fluxes. These data provide spatially averaged sediment yields of 130{ndash}2100 t{middot}km{minus}2{middot}yr{minus}1. Local rates of soil loss from agricultural plots on hillslopes are as high as 7000 t{middot}km{minus}2{middot}yr{minus}1. By comparison, natural rates of sediment generation, as determined by measuring cosmogenic 10Be in quartz from sediments and soils, are only 13{ndash}30 t{middot}km{minus}2{middot}yr{minus}1. The natural rates presented here provide a benchmark against which recent erosion rates, determined by various sediment gauging techniques, can be referenced. In the Sri Lankan highlands, these results suggest that soil is now being lost 10{ndash}100 times faster from agriculturally utilized areas than it is being produced.
 
Hewitt, K. (1988). "Castastrophic Landslide Deposits in the Karakoram Himalayas." Science 242: 64-67.
Hewlett, J. D., and A.R. Hibbert (1961). "Increases in water yield after several types of forest cutting." IAHS-AISH Publication 6: 5-17.
Hewlett, J. D., and R. Doss (1984). "Forests, floods, and erosion: a watershed experiment in the southeastern Piedmont." Forest Science 30(2): 424-434.
Hey, R., D. (1988). Mathematical models of channel morphology. Modelling Geomorphological Systems. M. G. Anderson. Chichester, John Wiley & Sons: 99-125.
Hey, R. D. (1978). "Determinate Hydraulic Geometry of River Channels." Journal of the Hydraulics Division 6: 869-883.
Hey, R. D. (1979). "Dynamic Process-Response Model of River Channel Development." Earth Surface Processes 4: 59-72.
Hey, R. D. (1979). "Flow resistance in gravel-bed rivers." J. Hydr. Div., Amer. Soc. Civ. Eng. 105: 365-379.
Heyerdahl, E. K. and J. K. Agee (1993). Fire History Database of the Western United States: Preliminary Report Covering Oregon and Washington. Seattle, Washington, USDA Forest Service: 37.
Hibbs, D. E. and P. A. Giordano (1996). "Vegetation characteristics of alder-dominated riparian buffer strips in the Oregon Coast Range." Northwest Science 70(3): 213-222.
Hickin, E. J. (1984). "Vegetation and river channel dynamics." Canadian Geographer 28: 111-126.
Hicks, B. J., et al. (1991). "Long-term changes in streamflow following logging in western Oregon and associated fisheries implications." Water Resources Bulletin 27(3): 217-226.
Hicks, B. J. and J. D. Hall (2003). "Rock Type and Channel Gradient Structure Salmonid Populations in the Oregon Coast Range." Transactions of the American Fisheries Society 132: 468-482.
Hicks, B. J., et al. (1991). Responses of salmonids to habitat changes. Influences of Forest and Rangeland Management on Salmonid Fishes and Their Habitats. W. R. Meehan. Bethesda, Maryland, American Fisheries Society. Publication 19: 483-518.
Hicks, D. and H. Omura (1992). Landslides and Debris Flows at Shimoda, Japan. Lower Hutt, New Zealand, DSIR Land Resources: 11.
Hicks, D. and H. Omura (1995). Channel Morphology, Landslides, and Debris Flow Occurence at Tokyo, Japan. International Sabo Symposium, Tokyo, Japan.
Hicks, D. M., et al. (2000). "Erosion thresholds and suspended sediment yields, Waipaoa River Basin, New Zealand." Water Resources Research 36(4): 1129-1142.
Higgins, J. J. (2004). An introduction to modern nonparametric statistics, Brooks/Cole,Pacific Grove, CA, p366.
Higgins, J. L. (2003). Determination of upstream boundary points on southeastern Washington streams and rivers under the requirements of the Shoreline Management Act of 1971. U. S. G. Survey. Tacoma, WA.
Hilderbrand, R. H., et al. (1997). "Effects of large woody debris placement on stream channels and benthic macroinvertebrates." Canadian Journal of Fisheries and Aquatic Science 54: 931-939.
Large woody debris (LWD) was added as an experimental stream restoration technique in two streams in southwest
Virginia. Additions were designed to compare human judgement in log placements against a randomized design and an
unmanipulated reach, &d also to compare effectiveness in a low- and a high-gradient stream. Pool area increased 146% in the
systematic placement and 32% in the random placement sections of the low-gradient stream, lending support to the notion that
human judgement can be more effective than placing logs at random in low-gradient streams. Conversely, the high-gradient
stream changed very little after LWD additions, suggesting that other hydraulic controls such as boulders and bedrock
override LWD influences in high-gradient streams. Logs oriented as dams were responsible for all pools created by additions
regardless of stream or method of placement. Multiple log combinations created only two pools, while the other seven pools
were created by single LWD pieces. Total benthic macroinvertebrate abundance did not change as a result of LWD additions
in either stream, but net abundances of Plecoptera, Coleoptera, Trichoptera, and Oligochaeta decreased, while Ephemeroptera
increased significantly with the proportional increase in pool area in the low-gradient stream.
 
 
Hillman, G. R. and J. P. Verschuren (1988). "Simulation of the effects of forest cover, and its removal, on subsurface water." Water Resources Research 24(2): 305-314.
Hirano, M. and M. Aniya (1988). "A rational explanation of cross-profile morphology for glacial valleys and of glacial valley development." Earth Surface Processes and Landforms 13: 707-716.
Hitt, N. P. (2007). Effects of stream network topology on fish assemblage structure and bioassessment sensitivity in the mid-Atlantic highlands, USA. Fisheries and Wildlife Sciences. Blacksburg, Virginia, Virginia Polytechnic Institute and State University. PhD: 242.
Stream fish assemblages exist within stream networks defined by the size and
proximity of connected streams (i.e., stream network topology). The spatial position of
sites within stream networks may therefore regulate opportunities for fish dispersal to
access distant resources or colonize “new” habitats. Such inter-stream dispersal
dynamics will influence local fish assemblage structure and the vulnerability of local
assemblages to anthropogenic stressors. In this dissertation, I explored the effects of
stream network topology on fish assemblage structure in the mid-Atlantic highlands,
USA and tested the hypothesis that dispersal would affect the sensitivity of fish-based
environmental quality assessments (i.e., bioassessments).
In chapter 1, I evaluated the effects of stream networks by comparing fish
assemblages between sites with and without large downstream confluences (>3rd order) in
western Virginia, USA (i.e., mainstem tributaries and headwater tributaries, respectively).
I found that local species richness was higher in mainstem tributaries than headwater
tributaries and that these effects could not be explained by variation in local
environmental habitat conditions. In chapter 2, I developed and applied a continuous
model of stream network topology to explore the effects of downstream size and
proximity on local fish assemblage structure within the mid-Atlantic highlands. I found
that fish assemblage structure (i.e., Bray-Curtis distances in species abundance) was
significantly related to variation in stream network topology up to approximately 9 fluvial
km from sites.
Chapters 3 and 4 explored the implications of inter-stream dispersal for fish
bioassessments. In Chapter 3, I identified 10 fish metrics that corresponded predictably
to environmental stressors in the mid-Atlantic highlands. However, headwater tributary
assemblages showed stronger relations to local environmental quality than mainstem
tributaries, consistent with the hypothesis of riverine dispersal. In Chapter 4, I compared
the effects of stream network topology on fish and benthic macroinvertebrate
assemblages. Fish metrics were influenced by the size and proximity of connected
streams but benthic macroinvertebrate metrics were not. This finding suggests that
stream fishes may complement benthic macroinvertebrate bioassessments by indicating
environmental conditions at larger spatial grains.
 
Hjort, J. and M. Marmion (2008). "Effects of sample size on the accuracy of geomorphological models." Geomorphology 102: 341-350.
Commonly, the most costly part of geomorphological distribution modelling studies is gathering the data.
Thus, guidance for researchers concerning the quantity of field data needed would be extremely practical.
This paper scrutinises the relationship between the sample size (the number of observations varied from 20
to 600) and the predictive ability of the generalized linear model (GLM), generalized additive model (GAM),
generalized boosting method (GBM) and artificial neural network (ANN) in two data settings, i.e.,
independent and split-sample approaches. The study was performed using empirical data of periglacial
processes from an area of 600 km2 in northernmost Finland at grid resolutions of 1 ha (100×100 m) and
25 ha (500×500 m). A rather sharp increase in the predictive ability of the models was observed when the
number of observations increased from 20 to 100, and the level of robust predictions was reached with 200
observations. The result indicates that no more than a few hundred observations are needed in
geomorphological distribution modelling at a medium scale resolution (ca. 0.01–1 km2).
 
Hobbs, S. and S. Till (1991). COPE Program Progress Report for FY 91. COPE Advisory Council Meeting, Salem, Oregon.
Hodgson, M. E. and P. Bresnahan (2004). "Accuracy of airborne LiDAR-derived elevation: empirical assessment and error budget." Photogrammetric Engineering and Remote Sensing 70(3): 331-339.
As part of a countywide large-scale mapping effort for
Richland County, South Carolina, an accuracy assessment of
a recently acquired lidar-derived data set was conducted.
Airborne lidar (2-m nominal posting) was collected at a flying
height of 1207 meters above ground level (AGL) using an
Optech ALTM (Airborne Laser Terrain Mapper) 1210 system.
Unique to this study are the reference point elevations. Rather
than using an interpolation approach for gathering observed
elevations at reference points, the x-y coordinates of lidar
points were located in the field and these elevations were
surveyed. Using both total-station-based and rapid-static GPS
techniques, observed vertical heights were measured at each
reference lidar posting. The variability of vertical accuracy
was evaluated for six land-cover categories. Root-meansquared
error (RMSE) values ranged from a low of 17 to 19 cm
(pavement, low grass, and evergreen forests) to a high of
26 cm (deciduous forests). The unique error assessment of
lidar postings also allowed for the creation of an error budget
model. The observed lidar elevation error was decomposed
into errors from lidar system measurements, horizontal displacement,
interpolation error, and surveyor error. A crossvalidation
approach was used to assess the observed interpolated
lidar elevation error for each field-verified reference
point. In order of decreasing importance, the lidar system
measurements were the dominant source of error followed
by interpolation error, horizontal displacement error, and
surveyor error. Observed elevation error in steeper slopes
(e.g., 25°) was estimated to be twice as large as those on low
slopes (e.g., 1.5°).
 
 
Hodgson, M. E., et al. (2005). "An evaluation of Lidar-derived elevation and terrain slope in leaf-off conditions." Photogrammetric Engineering and Remote Sensing 17(7): 817-823.
The effects of land cover and surface slope on lidar-derived
elevation data were examined for a watershed in the piedmont
of North Carolina. Lidar data were collected over the
study area in a winter (leaf-off) overflight. Survey-grade
elevation points (1,225) for six different land cover classes
were used as reference points. Root mean squared error
(RMSE) for land cover classes ranged from 14.5 cm to 36.1 cm.
Land cover with taller canopy vegetation exhibited the
largest errors. The largest mean error (36.1 cm RMSE) was in
the scrub-shrub cover class. Over the small slope range (0° to
10°) in this study area, there was little evidence for an
increase in elevation error with increased slopes. However,
for low grass land cover, elevation errors do increase in a
consistent manner with increasing slope. Slope errors
increased with increasing surface slope, under-predicting
true slope on surface slopes  2°. On average, the lidarderived
elevation under-predicted true elevation regardless
of land cover category. The under-prediction was significant,
and ranged up to  23.6 cm under pine land cover.
 
 
Hodgson, M. E., et al. (2003). "An evaluation of LIDAR- and IFSAR-derived digital elevation models in leaf-on conditions with USGS Level 1 and Level 2 DEMs." Remote Sensing of Environment 84: 295-308.
An assessment of four different remote sensing based methods for deriving digital elevation models (DEMs) was conducted in a floodprone
watershed in North Carolina. New airborne LIDAR (light detecting and ranging) and IFSAR (interferometric synthetic aperture radar
(SAR)) data were collected and corresponding DEMs created. These new sources were compared to two methods: Gestalt Photomapper
(GPM) and contour-to-grid, used by the U.S. Geological Survey (USGS) for creating DEMs. Survey-grade points (1470) for five different
land cover classes were used as reference points. One unique aspect of this study was the LIDAR and IFSAR data were collected during leafon
conditions. Analyses of absolute elevation accuracy and terrain slope were conducted. The LIDAR- and contour-to-grid derived DEMs
exhibited the highest overall absolute elevation accuracies. Elevation accuracy was found to vary with land cover categories. Elevation
accuracy also decreased with increasing slopes—but only for the scrub/shrub land cover category. Appreciable terrain slope errors for the
reference points were found with all methods.
 
 
Hoey, T. (1992). "Temporal variations in bedload transport rates and sediment storage in gravel-bed rivers." Prog. in Phys. Geography 16(3): 319-338.
Temporal variability in bedload transport rates and spatial variability in sediment storage have been reported with increasing frequency in recent years. A spatial and temporal classification for these features is suggested based on the gravel bedform classificatyion of Church and Jones (1982). The identified scales, meso-, macro-, and mega- are each broad, and within each there is a wide range of processes acting to produce bedload fluctuations. Sampling the same data set with different sampling intervals yields a near linear relationship between sampling interval and pulse period. A range of modelling strategies has been applied to bed waves. The most successful have been those which allow for the three-dimensional nature of sediment storage processes, and which allow changes in the width and depth of stored sediment. The existence of bed waves makes equilibrium in gravel-bed rivers necessarily dynamic. Bedload pulses and bed waves can be regarded as equilibrium forms at sufficiently long timescales.
 
Hoey, T. and R. Ferguson (1994). "Numerical simulation of downstream fining by selective transport in gravel bed rivers: Model development and illustration." Water Resources Research 30(7): 2251-2260.
Hoey, T. B. and R. Ferguson (1994). "Numerical simulation of downstream fining by selective transport in gravel bed rivers: Model development and illustration." Water Resources Research 30(7): 2251-2260.
Hoey, T. B. and A. J. Sutherland (1991). "Channel Morphololgy and Bedload Pulses in Braided Rivers: A Laboratory Study." Earth Surface Processes and Landforms 16: 447-462.
Hof, J. and M. Bevers (2001). "A spatial linear program for optimally scheduling forest management to meet stormflow objectives." Journal of the American Water Resources Association 37(3): 571-584.
A spatial linear program that strategically arranges and schedules forest treatments so as to meet peak stormflow objectives is formulated and demonstrated. The approach uses simulated spatial routing of stormflows nested as short-term time schedules within longer-term forest planning time periods. A simple case example is used to demonstrate the formulation and explore its spatial sensitivity.
 
Hoffman, D. F. and E. J. Gabet (2007). "Effects of sediment pulses on channel morphology in a gravel-bed river." Geological Society of America Bulletin 119(1): 116-125.
Hofmeister, R. J. (2000). Slope Failures in Oregon. GIS Inventory for Three 1996/97 Storm Events. Portland, Oregon Department of Geology and Mineral Industries: 20.
Hofmeister, R. J. and D. J. Miller (2003). GIS-based modeling of debris-flow initiation, transport and deposition zones for regional hazard assessments in western, Oregon, USA. Debris-Flow Hazards Mitigation: Mechanics, Prediction, and Assessment. Reickenmann and Chen. Rotterdam, Millpress: 1141-1149.
Hofmeister, R. J., et al. (2002). Hazard map of potential rapidly moving landslides in western Oregon, Interpretive Map Series - 22. Portland, Oregon Department of Geology and Mineral Industries.
Hogan, D. (1984). The Influence of Large Organic Debris on Channel Morphology in Queen Charlotte Island Streams. Proceedings of the Western Association of Fish and Wildlife Agencies and the Western Division American Fisheries Society, Victoria, British Columbia.
Hogan, D. L. (1989). Channel Response to Mass Wasting in the Queen Charlotte Islands British Columbia: Temporal and Spatial Changes in Stream Morphology. Watershed 89: a conference on the stewardship of soil, air, and water resources, Juneau, Alaska.
Hogan, D. L., et al. (1998). Spatial and temporal evolution of small coastal gravel-bed streams: The influence of forest management on channel morphology and fish habitats. Gravel-Bed Rivers in the Environment, Gravel Bed Rivers IV. P. C. Klingeman, R. L. Beschta, P. D. Komar and J. B. Bradley, Water Resources Publications. Chapter 17: 365-392.
Hogan, D. L., et al. (1999). Spatial and Temporal Evolution of Small Coastal Gravel-bed Streams: Influence of Forest Management on Channel Morphology and Fish Habitat. Gravel-Bed Rivers in the Environment. P. C. Kingman and R. L. Beschta: 365-392.
Hogan, D. L., et al. (1998). Stream channel morphology and recovery processes. Carnation Creek and Queen Charlotte Islands Fish/Forestry Workshop: Applying 20 Years of Coast Research to Management Solutions. Land management handbook 41. D. L. Hogan, P. J. Tschaplinski and S. Chatwin. Victoria, B.C., Crown Publications, Inc.: 77-96.
Hogan, D. L. and M. Church (1989). "Hydraulic Geometry in Small, Coastal Streams: Progress Toward Quantification of Salmonid Habitat." Canadian Journal of Fisheries and Aquatic Sciences 46.
Hogan, D. L. and D. J. Wilford (1989). A Sediment Transfer Hazard Classification System: Linking Erosion to Fish Habitat. Watershed 89': A Conference on the Stewardship of Soil, Air, and Water Resources, Juneau, Alaska.
filed (2)
 
Hoh Indian Tribe (1990). ? (DRAFT). Forks, Washington, Hoh Indian Tribe.
Holling, C. S. and G. K. Meffe (1996). "Command and Control and the Pathology of Natural Resource Management." Conservation Biology 10(2): 328-337.
Hollingsworth, R. and G. S. Kovacs (1981). "Soil Slumps and Debris Flows: Prediction and Protection." Bulletin of the Association of Engineering Geologists 28(1): 17-27.
Holmes, K. L. and P. C. Goebel (2011). "A functional approach to riparian area delineation using geospatial methods." Journal of Forestry 109(4): 233-241.
Riparian areas are diverse ecotones that provide numerous, valuable ecosystem functions. However,
many riparian delineation methods use a fixed minimum width to create a riparian buffer or setback
that may not adequately protect actual riparian function. A method for riparian area delineation across
landscapes is presented that incorporates riparian function and moves beyond the fixed-width buffer
approach. Using geospatial data and tools, riparian areas were delineated functionally for the Cuyahoga
Valley National Park in northeastern Ohio and compared to fixed-width buffers in terms of extent and
protection of riparian function. We suggest that functional riparian area delineation be incorporated into
watershed management planning to improve protection and restoration of the valuable ecological
functions provided by riparian areas across landscapes.
 
Holmes, K. W., et al. (2000). "Error in a USGS 30-meter digital elevation model and its impact on terrain modeling." Journal of Hydrology 233: 154-173.
Calculations based on US Geological Survey (USGS) digital elevation models (DEMs) inherit any errors associated with that particular representation of topography. We investigated the potential impact of error in a USGS 30 m DEM on terrain analysis over 27 km2. The difference in elevation between 2652 differential Global Positioning Systems measurements and USGS 30-m DEM derived elevations provided the comparative error dataset. Analysis of this comparative error data suggested that although the global (average) error is small, local error values can be large, and also spatially correlated. Stochastic conditional simulation was used to generate multiple realizations of the DEM error surface that reproduce the error measurements at their original locations and sample statistics such as the histogram and semivariogram model. The differences between these alternative error surfaces provide a model of uncertainty for the unknown DEM error spatial distribution. These DEM errors had a signicant impact on terrain attributes which compound elevation values of many grid cells (e.g. slope, wetness index, etc.). A case study using terrain modeling demonstrates that the result of error propagation is most dramatic in valley bottoms and along streamlines.
 
Holt, R. A., et al. (1975). "Relation of Water Temperature to Flexibacter columnaris Infection in Steelhead Trout (Salmo gairdneri), Coho (Oncorhynchus kisutch) and Chinook (O. tschawytscha) Salmon " J. Fish Res. Board Can. 32: 1553-1559.
Holtby, L. B. and M. C. Healey (1986). "Selection for Adult Size in Female Coho Salmon (Onchorhynchus kisutch) " Canadian Journal of Fisheries and Aquatic Sciences 43: 1946-1959.
Holtby, L. B. and J. C. Scrivener (1989). Observed and Simulated Effects of Climatic Variability, Clear-Cut Logging and Fishing on the Numbers of Chum Salmon (Oncorhynchus keta) and Coho Salmon (O. kisutch) Returning to Carnation Creek, British Columbia. National Workshop on Effects of Habitat Alteration on Salmonid Stocks.
Hooke, J. (2003). "Coarse sediment connectivity in river channel systems: a conceptual framework and methodology." Geomorphology 56: 79-94.
Hooke, J. M. (1980). "Magnitude and distribution of rates of river bank erosion." Earth Surface Processes 5: 143-157.
Horn, B. K. P. (1981). "Hill shading and the reflectance map." Proceedings of the IEEE 69(1): 14-47.
Hornbeck, J. W., M.B. Adams, E.S. Corbett, E.S. Verry, and J.A. Lynch (1995). A summary of water yield experiments on hardwood forested watersheds in the northeastern U.S. 10th Central Hardwood Forest Conference.
Horton, J. S., et al. (1964). Guide for Surveying Phreatophyte Vegetation. Washington, D.C., U.S.D.A., Forest Service: 37.
Horton, R. E. (1945). "Hydrophysical approach to the morphology of hillslopes and drainage basins." Geological Society of American bulletin 56: 275-370.
Hosmer, D. W. and S. Lemeshow (2000). Applied Logistic Regression. New York, John Wiley & Sons, Inc.
Hovius, N., et al. (Landslide).
Hovius, N., et al. (1997). "Sediment flux from a mountain belt derived by landslide mapping." Geology 25(3): 231-234.
Hovius, N., et al. (2000). "Supply and Removal of Sediment in a Landslide-dominated Mountain Belt: Central Range, Taiwan." The Journal of Geology 108: 73-89.
Hovius, N., et al. (2000). "Supply and removal of sediment in a landslide-dominated mountain belt: Central Range, Taiwan." Journal of Geology 108: 73-89.
A strong coupling between hillslope and valley systems is often inferred for mountain landscapes dominated by bedrock landsliding. We reveal the nature of this link using data sets on landsliding and sediment transport from two montane catchments draining the eastern Central Range of Taiwan. Here, the magnitude-frequency distribution of landslides can be modeled by a robust power law, but this scale invariance is not mirrored in the sediment discharge at the mountain front. Instead, downstream sediment loads reflect a complex response to both sediment supply and ambient hydraulic conditions. The rivers do not transport significant amounts of sediment unless it is provided by hillslope mass wasting in the catchment. Removal of landslide debris is a function of the transport capacity of the stream at the site of entry; thus, there is a dual supply and transport control on sediment loads in bedrock-floored streams. Over a monitoring period of >25 yr, the bulk of the sediment leaving the mountain belt was supplied by climate-triggered mass wasting. Peaks in water discharge were always closely followed by sediment load maxima, and the rapid decay of the latter indicates an effective removal of most supply. Where an important part of a catchment's sediment yield is derived from interfluves, sediment transport cannot simply be estimated from known water discharge time series, using a sediment rating curve, but requires instead a detailed knowledge of the spatial and temporal patterns of hillslope mass wasting and sediment transfer into the fluvial system.
 
Hovius, N., et al. (1998). "Landslide-driven drainage network evolution in a pre-steady-state mountain belt: Finisterre Mountains, Papua New Guinea." Geology 26(12): 1071-1074.
Howard, A. D. (1967). "Drainage analysis in geologic interpretation: a summation." American Association of Petroleum Geologists Bulletin 51: 2246-2259.
Howard, A. D. (1988). Equilibrium models in geomorphology. Modelling geomorphological systems. M. G. Anderson. New York, Wiley: 49-72.
Howard, A. D., et al. (1994). "Modeling fluvial erosion on regional to continental scales." Journal of Geophysical Research 99(B7): 13971-13986.
Howard, A. D. and T. R. Knutson (1984). "Sufficient Conditions for River Meandering: A Simulations Approach." Water Resources Research 20(11): 1659-1667.
Howard, A. D. and C. F. McLane, III (1988). "Erosion of Cohesionless Sediment by Groundwater Seepage." Water Resources Research 24(10): 1659-1674.
Howard, A. D. and C. F. McLane III (1988). "Erosion of Cohesionless Sediment by Groundwater Seepage." Water Resources Research 24(10): 1659-1674.
Howell, P. J. (2006). "Effects of wildfire and subsequent hydrologic events on fish distribution and abundance in tributaries of North Fork John Day River." North American Journal of Fisheries Management 26: 983-994.
Abstract.—Recent large wildfires in western states have fueled increasing concerns of resource managers
and the public about the effects of fire, including risks to fish, particularly endangered species. However, there
are few empirical studies on the response of fish to fire and none that include anadromous species. The Tower
Fire was one of four large fires in the upper John Day River basin in Oregon in 1996. Much of the area burned
at moderate to high severity, consistent with the pattern of increasing fire severity and size projected for the
region. Intense spring storms in 1997 and 1998 triggered large floods, landslides, and debris torrents that
affected streams within and downstream of the fire. We investigated the effects of the fire and ensuing floods
on fish distribution and abundance in three streams immediately after the fire through 2003. Immediately after
the fire, no fish were found in moderate- and high-intensity burn areas. Fish began to repopulate defaunated
reaches the year after the fire, and within 4 years distribution of juvenile steelhead (anadromous rainbow trout
Oncorhynchus mykiss) and resident rainbow trout was similar to that before the fire. Juvenile spring Chinook
salmon O. tshawytscha also began to use lower reaches of one of the streams after the flood, which had
eliminated a culvert near the mouth of the stream suspected to be a barrier. Densities in most burned reaches
and in unburned reaches downstream of the fire have rebounded to levels similar to or greater than densities in
reference streams outside of the fire. An isolated introduced population of brook trout Salvelinus fontinalis
also recovered. Thus, despite the size and severity of the fire, postfire hydrologic events, and human-induced
changes to watersheds, fish populations were highly resilient.
 
 
Howes, D. E. and M. Sondheim (1989). Quantitative definitions of stability classes as related to post-logging clearcut landslide occurrence, B.C. Min. For. Land Man.: 167-187.
Hsu, H.-L., et al. (2010). "Bedrock detection using 2D electrical resistivity imaging along the Peikang River, central Taiwan." Geomorphology 114: 406-414.
Hsu, K. J. (1975). "Catastrophic Debris Streams (Sturzstroms) Generated by Rockfalls." Geological Society of America Bulletin 86: 129-140.
Huang, C. (1998). "Sediment regimes under different slope and surface hydrologic conditions." Soil Science Society of America Journal 62(2): 423-430.
During rainfall events, many physical processes occur simultaneously. Conceptual frameworks for infiltration, runoff, surface sealing, sediment detachment, transport, and deposition processes have long been recognized. Nevertheless, interactions between hydrologic and sediment regimes and how they are affected by changes in slope steepness, soil erodibility, and rainfall erosivity have not been evaluated. A laboratory study on the Glynwood clay loam (fine, illitic, mesic Aquic Hapludalf) was designed to demonstrate interactions between hydrologic and sediment regimes. Four rainfall intensities, from 30 to 120 mm h(-1), and two inflows, equivalent to 30 and 60 mm h(-1) rainfall, were applied to create different combinations of rainfall detachment and shallow flow transport. Sediment delivery was quantified from two rainfall events: first on freshly packed soil under free drainage and later on the eroded surface under seepage. During the first rainfall event, average sediment deliveries from 30 to 100 mm h(-1) runoff were 1.8, 3.4, and 3.3 kg m(-2) h(-1) for slopes of 5, 10, and 15%, respectively. The lack of slope effects from 10 to 15% slope indicates the shift from a transport-limiting sediment regime at low slopes to a detachment-limiting sediment regime at high slopes, Under the potentially more erodible seepage condition, the second rainfall event produced only 1.3, 1.8, and 1.5 kg m(-2) h(-1) sediment from the same range of slopes and runoff rates. Reduction in sediment delivery during the second run was caused by removal of the easily erodible fraction during the first run, Increased profile drainage at the 15% slope further limited sediment detachment, These results show interactions between surface hydrologic conditions and erosion processes and the importance of understanding dominant sediment regimes in developing erosion prediction models.
 
Huang, C., et al. (1999). "Sediment transport capacity and erosion processes: Model concepts and reality." Earth Surface Processes and Landforms 24(6): 503-516.
Huang, C. H. (1995). "Empirical analysis of slope and runoff for sediment delivery from interrill areas." Soil Science Society of America Journal 59(4): 982-990.
Abstract: Slope steepness (S) and runoff discharge ((qw)) are two major factors in determining sediment delivery rates (q(s)) from interrill areas. Under the current interrill erosion model concept, these two factors are assumed to have independent effects on q(s); thus, each factor can be quantified individually if the other factor is kept constant. This study was conducted to show the effects of S and q(w) on q(s) and their interdependency. Sediment discharge rates, measured under different rainfall intensities and slope gradients, for eight soils from two laboratory studies were analyzed empirically with curve-fitting procedures. Results showed that there was a pair of empirical equations for each soil: q(s) = A(1) q(w)(2) + A(2) q(w) + A(3) and q(s) = B-1 S-2 + B-2 S + B-3 where A(1), A(2), and A(3) are functions of S and B-1, B-2, and B-3 are functions of q(w). In other words, effects of slope steepness and runoff on sediment delivery are dependent on each other. When S and q(w) were combined together as stream power, Omega, and plotted against sediment concentration, 4,/4,, a unique nonlinear relationship existed for each soil: q(s)/q(q) = D-1 Omega(2) + D-2 Omega + D-3 ,where D-1, D-2, and D-3 are soil-dependent coefficients. The stream power, which encompasses both slope and runoff effects, may provide improved estimates for interrill erosion. Although not based on theory, this result may be useful in process-based erosion models.
 
Huang, C. H. and J. M. Bradford (1993). "Analyses of slope and runoff factors based on the Wepp erosion model." Soil Science Society of America Journal 57(5): 1176-1183.
Abstract: Under process-based erosion model development, the source of sediment transported off a field is separated into that from interrill and rill areas, and separate detachment equations are developed. Limitation of this spatial separation is that, in many conditions, rill and interrill areas are not clearly defined a priori. We interpreted the erosion and deposition equations used in the current Water Erosion Prediction Project (WEPP) in an alternative fashion such that spatial separation of rill and interrill areas is no longer required. Analytic solutions were derived for the WEPP erosion and deposition equations under general conditions. Simplified solutions for a specific case, uniform rain on uniform slope, were examined closely for slope and runoff effects on sediment delivery. Under both erosion and deposition conditions, analytic solutions show a linear relationship between sediment yield, q(s) and slope, S. The dependency of q(s) on runoff, q(w), is either linear or quadratic depending on whether the system is dominated by an erosion or deposition regime. These analytic findings explain results obtained from laboratory studies in which sediment yield was collected under variable slope and rain intensities.
 
Huang, C. H., et al. (1996). "Evaluation of the detachment-transport coupling concept in the WEPP rill erosion equation." Soil Science Society of America Journal 60(3): 734-739.
The rill erosion equation in the current Water Erosion Prediction Project (WEPP) model is based on the coupled detachment and transport processes concept proposed by Foster and Meyer in 1972. The first-order detachment-transport coupling states that the rill detachment rate, D-r, is proportional to the difference between transport capacity, T-c, and sediment load, q(s): D-r = alpha (T-c - q(s)), where alpha is a rate control constant. A held experiment was designed to examine the validity of this model. Flow channels, 0.2 m wide, with clear water introduced at the upslope end were used in the study. Sediment delivery from different channel lengths and inflow rates was measured. Data were collected for three soils: Russell silt loam (fine-loamy, mixed, mesic Typic Hapludalf), Saybrook silt loam (fine-silty, mixed, mesic Typic Argiudoll), and Sharpsburg silty clay (fine, montmorillonitic, mesic Typic Argiudoll). Results showed that rill detachment and transport are not coupled processes. In the upper reach of a channel, q(s) is limited by a soil-dependent detachment rate. For longer channels, transport capacity controls the sediment delivery. Experimental data supported the Meyer and Wischmeier model concept, in which detachment and transport processes are separated and sediment delivery is limited to the lesser of the two. A slight modification to the Meyer and Wischmeier concept is the inclusion of an ''overshoot'' situation (i.e., q(s) > T-c) when the sediment regime is shifted from a detachment-dominated to a transport-dominated condition. Additional work is required to expand the database to develop a validated rill detachment and transport model.
 
Huang, H. Q. and G. C. Nanson (1998). "The influence of bank strength on channel geometry: an integrated analysis of some observations." Earth Surface Processes and Landforms 23: 865-876.
Hubbart, J. A., T.E. Link, J.A. Gravelle, and W.J. Elliot (2007). "Timber harvest impacts on water yield in the Continental/Maritime hydroclimate region of the United States." Forest Science 53(2): 169-180.
Hubbert, K. R., and V. Oriol (2005). "Temporal fluctuations in soil water repellency following wildfire in chapparral steeplands, southern California." International Journal of Wildland Fire 14(439-447).
Huber, O. L. (1992). "Sedimentation in a Highly Erosive Watershed." California Geology: 187-191.
Hudson, H. R. (2002). Linking the physical form and processes of rivers with ecological response. The Structure, Function and Management Implications of Fluvial Sedimentary Systems. F. J. Dyer, M. C. Thoms and J. M. Olley, International Association of Hydrological Sciences. 276: 121-139.
Fluvial eco-geomorphology seeks to link the physical form and processes of rivers with ecological responses by adopting an ecosystem perspective in research and management. It is recognized that the physical habitat is not static, and there are numerous and complex interconnections at various scales that determine fluvial form and processes. In turn, the physical environment exerts a strong control on river biota, but the effects can be indirect and mediated through complex interactions. As a consequence, there are considerable uncertainties in manipulating fluvial systems; hence an adaptive approach to management is required. Contributions from papers presented from the session on eco-geomorphology in the IAHS–ICCE–UNESCO International Symposium on the Structure, Function and Management Implications of Fluvial Sedimentary Systems are discussed in the context of two major themes of ecosystem management: physical habitat improvement and restoration of flow regimes. Challenges and opportunities for fluvial eco-geomorphology are outlined.
 
Hudson, R., and G. Horel (2007). An operational method of assesing hydrologic recovery for Vancouver Island and south coastal BC. Forest Research Technical Report TR-032, 16pp., Coast Forest Region, 2100 Labieux Road, BC, Canada, V9T 6E9. 250 751 7001.
Huene, R. v. (1984). "Tectonic processes along the front of modern convergent margins - research of the past decade." Ann. Rev. Earth Planet. Sci. 12: 359-381.
Huggett, R. J. (1988). "Dissipative systems: implications for geomorphology." Earth Surface Processes and Landforms 13: 45-49.
Hughes, R. M., et al., Eds. (2006). Landscape Influences on Stream Habitats and Biological Assemblages. Bethesda, Maryland, American Fisheries Society.
Hughes, V. and M. C. Thoms (2002). Associations between channel morphology and large woody debris in a lowland river. The Structure, Function and Management Implications of Fluvial Sedimentary Systems. F. J. Dyer, M. C. Thoms and J. M. Olley, International Association of Hydrological Sciences: 11-18.
Associations between channel morphology and the distribution and character of large woody debris (LWD) within a 95-km reach of the River Murray, Australia were examined at different scales. At the reach scale there was a uniform pattern of LWD distribution along the river. Most LWD was associated with eroding sites, close to the bank and aligned at 90° or less to the flow. At the sub-reach scale (0.5–1.5 km) strong associations were found between the curvature of the river channel and LWD distribution. Distribution patterns at this scale suggest that LWD is mainly recruited by bank erosion and falls into the river perpendicular to the flow. It subsequently remains close to where it falls and is realigned rather than actively moved by the river. Within meander bends there was twice as much LWD along the outer bank as there was along the inner bank, and while the amount on the inner bank declined with increasing distance into the bend the reverse was true for the outer bank.
 
Humphrey, N. F. (1983). Pore pressures in debris failure initiation. Department of Geology, University of Washington.
Hungr, O. (1987). "An extension of Bishop's simplified method of slope stability analysis to three dimensions." Geotechnique 37: 113-117.
Hungr, O. (1995). "A model for the runout analysis of rapid flow slides, debris flows, and avalanches." Canadian Geotechnical Journal 32: 610-623.
Hungr, O., et al. (2008). "Magnitude-frequency relationships of debris flows and debris avalanches in relation to slope relief." Geomorphology 96: 355-365.
Probability of occurrence, hazard intensity and encounter probability are key parameters in the quantitative risk analysis (QRA)
of landslides. All are strongly dependent on magnitude of the landslides. As a result, magnitude–frequency analysis should be a
part of QRA. Deriving representative magnitude–frequency relationships for debris avalanches and debris flows, however, is
difficult. One key problem is illustrated with the example of a unique database from the coastal region of British Columbia,
Canada, which was compiled entirely from detailed ground investigations. The magnitude of debris avalanches and debris flows is
not an independent statistical quantity, but a function of the scale of a given slope, as characterized by the slope length. Thus,
attempting to derive probability and magnitude for a given location or sub-region from a regionally-derived magnitude–frequency
curve may lead to incorrect predictions. The same problem is pertinent to the application of the same approach to any type of
landslide in which the largest combined dimension of the source volume (including entrainment) is of the same order as the length
of the slope. It is recommended that greater emphasis be placed on site-specific geological observations, at the expense of
generalized statistics.
 
Hungr, O., et al. (1984). "quantitative analysis of debris torrent hazards for design of remedial measures." Canadian Geotechnical Journal 21: 663-677.
Hunter, G. and R. Fell (2003). "Travel distance angle for "rapid" landslides in constructed and natural soil slopes." Canadian Geotechnical Journal 40: 1123-1141.
Hunter, M. (1998). Watershed-level patterns amoung stream amphibians in the Blue river watershed, west-central Cascades of Oregon. Corvallis, OR, Oregon State University: 97.
Hunter, M. A., et al. (2005). "Low flow spatial characteristics in forested headwater channels of southwest Washington." Journal of the American Water Resources Association 41(3): 503-516.
Patterns of dry season surface flow in forested headwater
channels of southwest Washington were observed during
August to September 2001 and July to October 2002. In 2001, 17
channels were sampled once, and the uppermost points of continuous
flow (CF) and surface water (SW) were located. In 2002, sampling
was replicated three to five times at each of 21 channels.
Annual and seasonal data suggested that the location of SW varied
less than CF. In most channels, SW remained at or near the channel
head year around. The pattern of surface flow between CF and
the channel head was used to test alternative hypotheses describing
dry season recession patterns: (A) surface flow consistently
retreats in a downstream direction, and (B) flow comes from fixed
sources along the channel, thus surface flow retreats up-channel
towards these sources. The dominant surface flow spatial pattern
in streams less than 30 percent slope was increased intermittency
without a clear pattern of retreat, and thus inconsistent with either
hypothesis. High gradient channels (> 30 percent slope) exhibited a
combination of increased intermittency, and extensive upward
retreats of surface water consistent with Hypothesis B. Differences
between 2001 and 2002 suggest late summer flows in small headwater
basins were controlled by spring precipitation, rather than
the typically greater winter precipitation.
(KEY TERMS: forest hydrology; surface water hydrology; headwater
hydrology; headwater channels; dry-season flow recession; spatial
flow characteristics; spatially intermittent flow; watershed
function.)
 
Hupp, C. R. Plant Ecological Aspects of Flood Geomorphology and Paleoflood History: 335-356.
Hupp, C. R. (1986). "The Headward Extent  of Fluvial Landforms and Associated Vegetation on Massanutten Mountain, Virginia." Earth Surface Processes and Landforms 11: 545-555.
Hupp, C. R. and W. R. Osterkamp (1985). "Bottomland vegetation distribution along Passage Creek, Virginia, in relation to fluvial landforms." Ecology 66: 670-681.
Hurst, H. E. (1950). "Long-term storage capacity of reservoirs." Proc. Amer. Soc. Civil Eng. 76(11).
Hurst, H. E. (1957). "A suggested statistical model of some time series which occur in nature." Nature 180: 494?
Huryn, A. D. and J. B. Wallace (1987). "Local geomorphology as a determinant of macrofaunal production in a mountain stream." Ecology 68(6): 1932-1942.
Huston, M. A. (1994). Biological diversity: The coexistence of species on changing landscapes. Cambridge, Cambridge University Press.
Hutchinson, M. F., and T.I. Dowling (1991). "A continental hydrological assessment of a new grid-based digital elevation model of Australia." Hydrological Processes 5: 45-58.
DEM, erosion,  GIS,  models, topographic analysis, watershed, basin, paleodrainage
 
Hyatt, T. L. and R. J. Naiman (2001). "The residence time of large woody debris in the Queets River, Washington, USA." Ecological Applications 11(1): 191-202.
Hyatt, T. L., et al. (2004). "A watershed scale assessment of riparian forests, with implications for restoration." Restoration Ecology 12(2): 175-183.
A combination of air-photo interpretation, field data, and
Geographic Information System (GIS) analysis was used to
map riparian areas that are likely to provide wood and shade
to small- and medium-sized streams and where, conversely,
restoration might be most beneficial. The analysis encompassed
all salmonid-bearing waters of the Nooksack River
basin, in northwest Washington State, plus small tributaries
that were thought to contribute wood or effective shading.
The size and composition of each riparian stand was examined
to determine whether trees were large enough to contribute
logs that would form pools in the adjacent channel, with poolforming
size of wood a function of channel width. Riparian
stands were classified according to whether they passed this
pool-forming test. Model results were an exactmatch to actual
conditions in 69% of field-verified stands. A large proportion
(74%) of the stands failing the test in reaches of anadromous
fish use were in agricultural areas. Passing stands typically had
high shade levels, because both stream shade and effective
large woody debris size are a function of the size of the trees
relative to the size of the stream. The GIS layer of passing and
failing riparian stands can be combined with layers depicting
property ownership, threatened fish distribution, and other
information to objectively prioritize riparian restoration
locations and strategies.
 
Hyde, K., et al. (2007). "Predicting gully rejuvenation after wildfire using remotely sensed burn severity data." Geomorphology 86: 496-511.
The loss of surface vegetation and reduced infiltration caused by wildfires can trigger gully rejuvenation, resulting in damage to
downstream aquatic resources and risk to human life and property. We developed a spatially explicit metric of burn severity —the
Burn Severity Distribution Index (BSDI) — and tested its ability to predict post-fire gully rejuvenation in 1st and 2nd order basins
burned in the 2000 Valley Complex fires in the Sapphire Mountains of western Montana. The BSDI was derived from burn severity
data interpreted from Landsat 7 satellite imagery using the Normalized Burn Ratio (NBR) method, and ranged from 0.0 for
completely unburned basins to 4.0 for basins burned entirely at high severity. In July 2001 rainstorms with peak 30-minute
intensities of up to 17 mm h−1 triggered gully rejuvenation in 66 of the 171 basins examined. The frequency of gully rejuvenation
was higher in basins with higher BSDI values, increasing from zero for basins with a BSDI less than 1.3 to 67% for basins with a
BSDI greater than 3.0. Binary logistic regression indicated that BSDI was a more significant predictor of gully rejuvenation than
basin morphometric variables. The absence of gully rejuvenation in several basins with a high BSDI was attributed to low gradient,
dense riparian vegetation, or concentration of high burn severity at lower elevations in the basin. The presence of gully rejuvenation
in several basins with a low BSDI was associated with false negative NBR classification errors in northwest aspects, and
concentration of severe burn impacts in the drainage headslopes. BSDI is a useful metric for predicting gully rejuvenation after
wildfire. The use of the BSDI in Burned Area Emergency Response team assessments could improve the planning,
implementation, and monitoring of burned area recovery treatments.
 
Hyyppă, H., et al. (2005). Factors affecting the quality of DTM generation in forested areas. Laser Scanning 2005. Enschede, the Netherlands, ISPRS. 3: 85-90.
Airborne laser scanning (ALS) has become an established tool for acquiring digital terrain models (DTM) in forested areas. Even
though, there have been several empirical studies on DTM quality with laser scanning, a few studies have focused on factors
affecting the quality of DTM generation. This paper analyses especially the effects of the date, flight altitude, pulse mode, terrain
slope, forest cover and plot variation on the DTM accuracy at boreal forest zone. The boreal test site was collected with Toposys I
and Toposys II in 1998, 2000 and 2003. Since the measurements were recorded at various time of the season, i.e. May 14th 2003
(leaf-off), June 14th 2000 (leaf-on, low development of undergrowth), and September 2nd 1998 (leaf-on, high undergrowth), it was
possible to estimate the effect of leaves and undergrowth. In 2003 the flight altitudes of 400, 800 and 1500 m above ground level
were used providing nominal pulse densities of 8-10, 4-5 and 2-3 pulses per m2. At boreal forest zone, the random errors of less than
20 cm were obtained in most conditions for non-steep terrain. The increase of flight altitude 400 to 1500 m increased the random
error of DTM derivation by 50%. The difference of using first or last pulse caused a similar random error difference. There were
systematic shifts in the elevation models derived at various flight altitudes. It is expected that the beam size and sensitivity of the
laser system determine this systematic behaviour. Additionally, the systematic shifts between last and first pulse were significant. The
difference of DTMs derived at optimum and non-optimal season conditions were typically less than 5 cm for high-density data. In
stands consisting of deciduous trees, the seasonal effects were the highest. The random error increased with increasing terrain slope.
The effect of forest cover was higher when moving closer to the trunk. The results were site dependent, i.e. the obtained accuracy
varied strongly as a function of site conditions.
 
 
Ibbitt, R. P., et al. (1998). "Taieri River data to test channel network and river basin heterogeneity concepts." Water Resources Research 34(8): 2085-2088.
Ice, G., and R. Whittemore (1998). Alternatives for evaluating water quality and BMP effectiveness at the watershed scale, www.nwqmc.org/98proceedings/Papers/08-ICE.html.
Ice, G. (?). Lessons Learned about Watershed Assessments. ?, ?: 32-44.
Ice, G. G. (1983). Landslides inventories, current research and pending forest practice act rule changes for mass wasting associated with forest management practices. Forest Management Practices and Natural Events - Their Relation to Landslides and Water Quality Protection, National Council of the Paper Industry for Air and Stream Improvement, Inc.
Ice, G. G., et al. (2006). "Estimating areas and timber values of riparian management on forest lands." Journal of the American Water Resources Association 42(1): 115-124.
Buffers, filter strips, and other riparian protection zones are widely accepted practices used to minimize water quality impacts from forest management and other land use activities. Riparian management prescriptions are often developed with limited or inconsistent consideration of the impact they will have on land management opportunities or economics. A combination of factors influences the area and value of riparian management zones (RMZs). A simple tool can determine the percentage of a watershed in RMZs where they comprise only a small fraction of the watershed. For a more detailed analysis, the Oak Creek Watershed in Oregon served as an example of a geographic information system (GIS) assessment. The same watershed can have dramatically different areas and values of timber in RMZs depending on the resolution used to determine the stream network, different stream types (perennial fish-bearing, perennial nonfish bearing, and nonperennial; small, medium, and large), and different RMZ widths and management restrictions for each stream type. The areas of watershed put into riparian protection for two drainage densities and three RMZ prescriptions were determined. Finally, the volumes of merchantable timber in the RMZs and their subsequent values were determined.
 
Ichoku, C. and J. Chorowicz (1994). "A numerical approach to the analysis and classification of channel network patterns." Water  Resources Research 30(2): 161-174.
Iida, T. (1983). "Development of hillslopes due to landslides." Z. Geomorph. N.F. 46: 67-77.
Iida, T. (1984). "A Hydrological Method of Estimation of the Topographic Effect on the Saturated Throughflow." Transactions, Japanese Geomorphical Union 5(1): 1-11.
Iida, T. (1999). "A stochastic hydro-geomorphological model for shallow landsliding due to rainstorm." Catena 34: 293-313.
A prediction model of shallow landsliding is proposed. It considers not only the deterministic aspects containing slope stability, saturated throughflow and a soil (regolith) depth development, but also the stochastic aspects of intensity and duration of rainfall. It turns out that the probability of saturated throughflow, which is the direct trigger mechanism to shallow landsliding, can be expressed by a log-normal distribution. The short term probability of landsliding is defined as the excess probability that the depth of saturated throughflow surpasses the critical value. The average recurrence interval Tav of landsliding can be calculated as the expected value. This model was applied to a test field where a lot of shallow landsliding occurred at a heavy rainstorm in 1988. Then a DEM of 5-m grid interval was utilized to calculate Tav at every grid point. Consequently, it was found that the percentage of the landslide grid number to the total grid number for every Tav rank increases when Tav decreases. Therefore, it is confirmed that Tav is an index of the susceptibility to shallow landsliding and the distribution map of Tav can be regarded as a kind of hazard map. The spatial distribution of Tav reveals its significant dependence on the topography
 
Iida, T. (?). A stochiastic hydro-geomorphological model for shallow landsliding due to rainstorm (DRAFT).
Ijjasz-Vasquez, E. J., et al. (1992). "On the multifractal characterization of river basins." Geomorphology 5: 297-310.
Ikeda, S., et al. (1981). "Bend theory of river meanders, part 1: linear development." Journal of Fluid Mechanics 112: 363-377.
Ikeya, H. (1981). A method of designation for area in danger of debris flow. Erosion and Sediment Transport in Pacific Rim Steeplands Publ. No. 132, International Association of Hydrological Studies.
Impara, P. C. (1997). Spatial and temproal patterns of fire in the forests of the Central Oregon Coast Range. Corvallis, OR, Oregon State University.
Inbar, M., M. Tamir, and L. Wittenberg (1998). "Runoff and erosion processes after a forest fire in Mount Carmel, a Mediterranean area." Geomorphology 24: 17-33.
Independent Multidisciplinary Science Team (1999). Recovery of Wild Salmonids in Western Oregon Forests: Oregon Forest Practices Act Rules and the Measures in the Oregon Plan for Salmon and Watersheds. Salem, Oregon, Governor's Natural Resources Offices: 85.
Innes, J. L. (1983). "Debris flows." Physical Geography 7(4): 470-501.
Irwin, A. and B. Wynne (1996). Conclusions. Misunderstanding Science? The public reconstruction of science and technology, Cambridge University Press: 149-213.
Isaak, D. J., et al. (1999). "Accuracy and precision of stream reach water surface slopes estimated in the field and from maps." North American Journal of Fisheries Management 19: 141-148.
The accuracy and precision of five tools used to measure stream water surface slope
(WSS) were evaluated. Water surface slopes estimated in the field with a clinometer or from
topographic maps used in conjunction with a map wheel or geographic information system (GIS)
were significantly higher than WSS estimated in the field with a surveying level (biases of 34,
41, and 53%, respectively). Accuracy of WSS estimates obtained with an Abney level did not
differ from surveying level estimates, but conclusions regarding the accuracy of Abney levels and
clinometers were weakened by intratool variability. The surveying level estimated WSS most
precisely (coefficient of variation [CV] 5 0.26%), followed by the GIS (CV 5 1.87%), map wheel
(CV 5 6.18%), Abney level (CV 5 13.68%), and clinometer (CV 5 21.57%). Estimates of WSS
measured in the field with an Abney level and estimated for the same reaches with a GIS used in
conjunction with 1:24,000-scale topographic maps were significantly correlated (r 5 0.86), but
there was a tendency for the GIS to overestimate WSS. Detailed accounts of the methods used to
measure WSS and recommendations regarding the measurement of WSS are provided.
 
Isaak, D. J., et al. (2007). "Chinook salmon use of spawning patches: relative roles of habitat quality, size, and connectivity." Ecological Applications 17(2): 352-364.
Declines in many native fish populations have led to reassessments of
management goals and shifted priorities from consumptive uses to species preservation. As
management has shifted, relevant environmental characteristics have evolved from traditional
metrics that described local habitat quality to characterizations of habitat size and
connectivity. Despite the implications this shift has for how habitats may be prioritized for
conservation, it has been rare to assess the relative importance of these habitat components.
We used an information–theoretic approach to select the best models from sets of logistic
regressions that linked habitat quality, size, and connectivity to the occurrence of chinook
salmon (Oncorhynchus tshawytscha) nests. Spawning distributions were censused annually
from 1995 to 2004, and data were complemented with field measurements that described
habitat quality in 43 suitable spawning patches across a stream network that drained 1150 km2
in central Idaho. Results indicated that the most plausible models were dominated by
measures of habitat size and connectivity, whereas habitat quality was of minor importance.
Connectivity was the strongest predictor of nest occurrence, but connectivity interacted with
habitat size, which became relatively more important when populations were reduced.
Comparison of observed nest distributions to null model predictions confirmed that the
habitat size association was driven by a biological mechanism when populations were small,
but this association may have been an area-related sampling artifact at higher abundances.
The implications for habitat management are that the size and connectivity of existing habitat
networks should be maintained whenever possible. In situations where habitat restoration is
occurring, expansion of existing areas or creation of new habitats in key areas that increase
connectivity may be beneficial. Information about habitat size and connectivity also could be
used to strategically prioritize areas for improvement of local habitat quality, with areas not
meeting minimum thresholds being deemed inappropriate for pursuit of restoration activities.
 
Isaak, D. J. and R. F. Thurow (2006). "Network-scale spatial and temporal variation in Chinook salmon (Oncorhynchus tshawytscha) redd distributions: patterns inferred from spatially continuous replicate surveys." Canadian Journal of Fisheries and Aquatic Science 63: 285-296.
Spatially continuous sampling designs, when temporally replicated, provide analytical flexibility and are
unmatched in their ability to provide a dynamic system view. We have compiled such a data set by georeferencing the
network-scale distribution of Chinook salmon (Oncorhynchus tshawytscha) redds across a large wilderness basin
(7330 km 2 ) in central Idaho for 9 years (1995–2003). During this time, the population grew at a rate of 5.3 recruits
per spawner, and redd numbers increased from 20 to 2271. As abundances increased, fish expanded into portions of the
stream network that had recently been unoccupied. Even at the highest escapements, however, distributions remained
clustered, and a limited portion of the network contained the majority of redds. The importance of the highest density
spawning areas was greatest when abundances were low, suggesting these areas may serve as refugia during demo-graphic
bottlenecks. Analysis of variance indicated that redd numbers were strongly affected by local habitats and
broad climatic controls, but also revealed a space–time interaction that suggested temporal instability in spatial patterns.
Our results emphasize the importance of maintaining habitats with high densities of individuals, but also suggest that
broader views may be needed to accommodate the dynamics of natural salmonid populations.
 
Isaak, D. J., et al. (2003). "Temporal variation in synchrony among chinook salmon (Oncorhynchus tshawytscha) redd counts from a wilderness area in central Idaho." Canadian Journal of Fisheries and Aquatic Science 60: 840-848.
Metapopulation dynamics have emerged as a key consideration in conservation planning for salmonid fishes.
Implicit to many models of spatially structured populations is a degree of synchrony, or correlation, among populations.
We used a spatially and temporally extensive database of chinook salmon (Oncorhynchus tshawytscha) redd counts
from a wilderness area in central Idaho to examine patterns in synchrony as these fish underwent a sixfold decrease in
abundance. Our results suggested that populations became strongly synchronous as abundances decreased and that the
range, or diversity of correlations, exhibited among populations also decreased. These changes indicate that the likeli-hood
of simultaneous extirpations has increased, which could have long-term detrimental consequences for
metapopulation persistence. Implications for management are that the resilience of many metapopulations to large-scale
disturbance and anthropogenic suppression may not depend solely on attempts to maintain large and productive compo-nent
populations, but also on efforts to desynchronize populations that have become strongly correlated. Such efforts
could entail promoting the existence of a broad distribution and diversity of habitats that support a wide array of life-history
forms and ensuring that some habitats are sufficiently spatially disjunct so that risks from catastrophic stochas-tic
events are minimized.
 
Isaaks, E. H. and R. M. Srivastava (1989). Applied Geostatistics. New York, Oxford University Press.
Iseya, F. and H. Ikeda (1987). "Pulsations in bedload transport rates induced by a longitudinal sediment sorting: a flume study using sand and gravel mixtures." Geografiska Annaler 69 A(1): 15-27.
Ishikawa, Y., et al. (2003). "Suppression of debris movement by forests and damage to forests by debris deposition." Journal of Forest Research 8(1): 37-47.
Abstract Forest buffer zones have recently been introduced in an attempt to suppress debris movement caused by mass wasting. There are, however, many questions left unsolved regarding the role of forests in suppressing debris movement and damage to forests by debris deposition. A number of debris avalanches occurred in southern Fukushima Prefecture and Hiroshima Prefecture in August 1998 and June 1999, respectively. Suppression of debris movement by forests and damage to forests by debris deposition were investigated in this study using aerial photographic interpretation and topographic analysis. Of the debris avalanches delineated in the Fukushima and Hiroshima areas, 282 and 84 sites, respectively, were forested, and 43 and seven sites, respectively, were nonforested. Topographic parameters, land use, and forest type at each site were surveyed by aerial photographic interpretation and topographic analysis. Suppression of debris movement by forest vegetation was confirmed by higher equivalent coefficients of friction and shorter average deposition zone length (about 28% and 55% shorter in the Fukushima and Hiroshima areas, respectively) in forested sites compared with nonforested sites. While previous studies suggested that the width of the forest zone required to prevent sediment outflow by surface erosion increases with increasing slope gradient, no clear relationship was found in this study. While the length of deposition zone of debris material (this study) is strongly affected by the dynamic solid friction coefficient and fluid friction of debris material, the length of deposition zone of the outflow of surface wash (past studies) is mainly affected by the tractive force of water flow. Among the 65 trees that remained at the terminus of deposition zones at five sites, 36 were alive and 29 were dead. Damage to the forest increases with increasing thickness of deposits and decreases with increasing tree diameter. Such results are useful for designing forest buffer zones.
 
Istanbulluoglu, E. Ch. 3 A sediment transport model for incising gullies on steep topography
Ch. 4 Modeling approach on the interactions between forest vegeation, disturbances and sediment yields.
Istanbulluoglu, E. (2002). Quantification of stream sediment inputs from steep forested mountains. Civil Engineering. Logan, UT, Utah State University: 210.
Istanbulluoglu, E., et al. (2002). "A probabilistic approach for channel initiation." Water Resources Research 38(12): 1325, doi:1310.1029/2001WR000782.
Istanbulluoglu, E., et al. (2003). "A sediment transport model for incision of gullies on steep topography." Water Resources Research 39(4): doi:10.1029/2002WR001467.
Istanbulluoglu, E., et al. (2004). "Modeling of the interactions between forest vegetation, disturbances, and sediment yields." Journal of Geophysical Research 109.
The controls of forest vegetation, wildfires, and harvest vegetation disturbances on
the frequency and magnitude of sediment delivery from a small watershed ( 3.9 km2)
in the Idaho batholith are investigated through numerical modeling. The model
simulates soil development based on continuous bedrock weathering and the
divergence of diffusive sediment transport on hillslopes. Soil removal is due to
episodic gully erosion, shallow landsliding, and debris flow generation. In the model,
forest vegetation provides root cohesion and surface resistance to channel initiation.
Forest fires and harvests reduce the vegetation. Vegetation loss leaves the land
susceptible to erosion and landsliding until the vegetation cover reestablishes in time.
Simulation results compare well with field observations of event sediment yields
and long-term averages over  10,000 years. When vegetation is not disturbed by
wildfires over thousands of years, sediment delivery is modeled to be less frequent but
with larger event magnitudes. Increased values of root cohesion (representing denser
forests) lead to higher event magnitudes. Wildfires appear to control the timing of
sediment delivery. Compared to undisturbed forests, erosion is concentrated during the
periods with low erosion thresholds, often called accelerated erosion periods, following
wildfires. Our modeling suggests that drainage density is inversely proportional to
root cohesion and that reduced forest cover due to wildfires increases the drainage
density. We compare the sediment yields under anthropogenic (harvest) and natural
(wildfire) disturbances. Disturbances due to forest harvesting appear to increase
the frequency of sediment delivery; however, the sediment delivery following wildfires
seems to be more severe. These modeling-based findings have implications for
engineering design and environmental management, where sediment inputs to streams
and the fluctuations and episodicity of these inputs are of concern.
 
Iverson, R. M. (1984). "A constitutive equation for mass-movement behavior." Journal of Geology 93: 143-160.
Iverson, R. M. (1986). "Unsteady, nonuniform landslide motion: 1. Theoretical dynamics and the steady datum state." Journal of Geology 94: 1-15.
Iverson, R. M. (1986). "Unsteady, nonuniform landslide motion:  2. Linearized theory and the kinematics of transient response." Journal of Geology 94: 349-364.
Iverson, R. M. (1997). "Debris-flow mobilization from landslides." Annual Review of Earth and Planetary Sciences 25: 85-138.
Iverson, R. M. (1997). "The physics of debris flows." Reviews of Geophysics 35: 245-296.
Recent advances in theory and experimentation
motivate a thorough reassessment of the physics
of debris flows. Analyses of flows of dry, granular solids
and solid-fluid mixtures provide a foundation for a comprehensive
debris flow theory, and experiments provide
data that reveal the strengths and limitations of theoretical
models. Both debris flow materials and dry granular
materials can sustain shear stresses while remaining static;
both can deform in a slow, tranquil mode characterized
by enduring, frictional grain contacts; and both can
flow in a more rapid, agitated mode characterized by
brief, inelastic grain collisions. In debris flows, however,
pore fluid that is highly viscous and nearly incompressible,
composed of water with suspended silt and clay, can
strongly mediate intergranular friction and collisions.
Grain friction, grain collisions, and viscous fluid flow
may transfer significant momentum simultaneously.
Both the vibrational kinetic energy of solid grains (measured
by a quantity termed the granular temperature)
and the pressure of the intervening pore fluid facilitate
motion of grains past one another, thereby enhancing
debris flow mobility. Granular temperature arises from
conversion of flow translational energy to grain vibrational
energy, a process that depends on shear rates,
grain properties, boundary conditions, and the ambient
fluid viscosity and pressure. Pore fluid pressures that
exceed static equilibrium pressures result from local or
global debris contraction. Like larger, natural debris
flows, experimental debris flows of ;10 m3 of poorly
sorted, water-saturated sediment invariably move as an
unsteady surge or series of surges. Measurements at the
base of experimental flows show that coarse-grained
surge fronts have little or no pore fluid pressure. In
contrast, finer-grained, thoroughly saturated debris behind
surge fronts is nearly liquefied by high pore pressure,
which persists owing to the great compressibility
and moderate permeability of the debris. Realistic models
of debris flows therefore require equations that simulate
inertial motion of surges in which high-resistance
fronts dominated by solid forces impede the motion of
low-resistance tails more strongly influenced by fluid
forces. Furthermore, because debris flows characteristically
originate as nearly rigid sediment masses, transform
at least partly to liquefied flows, and then transform
again to nearly rigid deposits, acceptable models
must simulate an evolution of material behavior without
invoking preternatural changes in material properties. A
simple model that satisfies most of these criteria uses
depth-averaged equations of motion patterned after
those of the Savage-Hutter theory for gravity-driven flow
of dry granular masses but generalized to include the
effects of viscous pore fluid with varying pressure. These
equations can describe a spectrum of debris flow behaviors
intermediate between those of wet rock avalanches
and sediment-laden water floods. With appropriate pore
pressure distributions the equations yield numerical solutions
that successfully predict unsteady, nonuniform
motion of experimental debris flows.
 
 
Iverson, R. M. (2000). "Landslide triggering by rain infiltration." Water Resources Research 36(7): 1897-1910.
Iverson, R. M. and R. P. Denlinger (2001). "Flow of variably fluidized granular masses across three-dimensional terrain. 1. Coulomb mixture theory." Journal of Geophysical Research 106(B1): 537-552.
Iverson, R. M. and J. J. Major (1986). "Groundwater Seepage Vectors and the Potential for Hillslope Failure and Debris Flow Mobilization." Water  Resources Research 22(11): 1543-1548.
Iverson, R. M. and J. J. Major (1987). "Rainfall, ground-water flow, and seasonal movement at Minor Creek landslide, northwestern California: Physical interpretation of empirical relations." Geological Society of America Bulletin 99: 579-594.
Iverson, R. M. and M. E. Reid (1992). "Gravity-driven groundwater flow and slope failure potential, 1. Elastic effective-stress model." Water Resources Research 38(3): 925-938.
Iverson, R. M., et al. (2000). "Acute sensitivity of landslide rates to initial soil porosity." Science 290: 513-516.
Iverson, R. M., et al. (1997). "Debris-Flow Mobilization From Landslides." Annu. Rev. Earth Planet 25: 85-138.
Iverson, R. M., et al. (1998). "Objective delineation of lahar-inundation hazard zones." Geological Society of America Bulletin 110: 972-984.
Iverson, R. M. and J. W. Vallance (2001). "New views of granular mass flows." Geology 29(2): 115-118.
Concentrated grain-fluid mixtures in rock avalanches, debris flows, and pyroclastic
flows do not behave as simple materials with fixed rheologies. Instead, rheology evolves
as mixture agitation, grain concentration, and fluid-pressure change during flow initiation,
transit, and deposition. Throughout a flow, however, normal forces on planes parallel to
the free upper surface approximately balance the weight of the superincumbent mixture,
and the Coulomb friction rule describes bulk intergranular shear stresses on such planes.
Pore-fluid pressure can temporarily or locally enhance mixture mobility by reducing Coulomb
friction and transferring shear stress to the fluid phase. Initial conditions, boundary
conditions, and grain comminution and sorting can influence pore-fluid pressures and
cause variations in flow dynamics and deposits.
 
Ives, J. D. (1984). "The Himalaya-Ganges problem in the context of peace and resource-use conflict management." Mountain Research and Development 4(4): 363-365.
Ives, J. D. and B. Messerli (1984). "Stability and instability of mountain ecosystems: lessons learned and recommendations for the future." Mountain Research and Development 4(1): 63-71.
Jackson, C. R., D.P. Batzer, S.S.Cross, S.M. Haggerty, and C. A. Sturm (2007). "Headwater streams and timber harvest: channel, macroinvertebrate, and amphibian response and recovery." Forest Science 53(2): 356-370.
Jackson, C. R. and T. W. Cundy (1992). "A model of transient, topographically driven, saturated subsurface flow." Water Resources Research 28(5): 1417-1427.
Jackson, C. R., et al. (2000). "Timber Harvest Impacts on Small Headwater Stream Channels in the Coast Ranges of Washington." submitted to the Journal of American Water Resources Association: 1-22.
Jackson, C. R. and C. A. Sturm (2002). "Woody debris and channel morphology in first- and second-order forested channels in Washington's coast ranges." Water Resources Research 38(9): 16-11 - 16-14.
Jackson, D. A., et al. (2001). "What controls who is where in freshwater fish communities -- the roles of biotic, abiotic, and spatial factors." Canadian Journal of Fisheries and Aquatic Science 58: 157-170.
We examine evidence for the structuring of fish communities from stream and lake systems and the roles of
biotic, abiotic, and spatial factors in determining the species composition. Piscivory by fish is a dominant factor in both
stream and lake systems whereas evidence for the importance of competition appears less convincing. Within small
streams or lakes, the impact of predation may exclude other species, thereby leading to mutually exclusive distributions
and strong differences in community composition. Within a geographic region, abiotic effects frequently dictate the relative
importance of piscivory, thereby indirectly influencing the composition of prey species present. The spatial scale
of studies influences our perceived importance of biotic versus abiotic factors, with small-scale studies indicating a
greater importance of competition and large-scale studies emphasizing abiotic controls. The scale of the individual sites
considered is critical because smaller systems have higher variability and wider extremes of conditions than larger
lakes and rivers. The stability of physical systems and degree of spatial connectivity contribute to increased diversity in
both larger stream and larger lake systems. We identify challenges and needs that must be addressed both to advance
the field of fish community ecology and to face the problems associated with human-induced changes
 
Jackson, L. E., Jr., et al. (1982). "Paraglacial origin for terraced river sediments in Bow Valley, Alberta." Can. J. Earth Sci. Vol. 19: 2219-2231.
Jackson, R. C., et al. (2001). "Timber harvest impacts on small headwater stream channels in the Coast Ranges of Washington." Journal of the American Water Resources Association 37(6): 1533-1549.
We evaluated changes in channel habitat distributions,
particle-size distributions of bed material, and stream temperatures
in a total of 15 first- or second-order streams within and
nearby four planned commercial timber harvest units prior to and
following timber harvest. Four of the 15 stream basins were not
harvested, and these streams served as references. Three streams
were cut with unthinned riparian buffers; one was cut with a partial
buffer; one was cut with a buffer of non-merchantable trees;
and the remaining six basins were clearcut to the channel edge. In
the clearcut streams, logging debris covered or buried 98 percent of
the channel length to an average depth of 0.94 meters. The slash
trapped fine sediment in the channel by inhibiting fluvial transport,
and the average percentage of fines increased from 12 percent
to 44 percent. The trees along buffered streams served as a fence to
keep out logging debris during the first summer following timber
harvest. Particle size distributions and habitat distributions in the
buffered and reference streams were largely unchanged from the
pre-harvest to post-harvest surveys. The debris that buried the
clearcut streams effectively shaded most of these streams and protected
them from temperature increases. These surveys have documented
immediate channel changes due to timber harvest, but
channel conditions will evolve over time as the slash decays and
becomes redistributed and as new vegetation develops on the channel
margins.
 
Jackson, W. L. and R. L. Beschta (1982). "A model of two-phase bedload transport in an Oregon Coast Range stream." Earth Surface Process and Landforms 7: 517-527.
A descriptive model for bed material routing in small sand and gravel bedded channels with sequences of pools and armoured riffles was developed. Aspects of the model are demonstrated at Flynn Creek, a 202-ha, third-order drainage in the Oregon Coast Range. Assuming that channel geometry, sediment transport competence, and the availability of sediments for transport are all non-uniform in the downstream direction, the model described bedload transport as occurring in two phases. Phase I involves the transport of fine, predominantly sand-sized bed materials over stable gravel-armoured riffles. Phase II occurs at higher flows that can entrain riffle armour and transport riffle sediments, in addition to Phase I sediments. Phase I bedload transport was sampled during three moderate flow events. A power function related Phase I bedload transport to water discharge. The relationship was consistent between storms and between tow successive years. Phase II bedload transport, sampled during a 1.8-yr return period streamflow event in February, 1979, proved non-uniform in a downstream direction and unsteady over time at a given stream discharge. Transport rates closely correlated with the rapid scour and redeposition of the riffle at the bedload sampling cross-section; tranport peaks corresponded to scour and transport decreased greatly during deposition. The transport of large (> 12.5 mm diam.) bed material increased up to 12-fold during riffle scour.
 
Jacobson, R. B. (1995). Spatial Controls on Patterns of Land-use Induced Stream Disturbance at the Drainage. Natural and Anthropogenic Influences in Fluvial Geomorphology. J. E. Costa, A. J. Miller, K. W. Potter and P. R. Wilcock, American Geophysical Union. The Wolman Volume: 219-239.
Jacobson, R. B. (1995). Spatial controls on patterns of land-use induced stream disturbance at the drainage-basin scale - an example from gravel-bed streams of the Ozark Plateaus, Missouri. Natural and Anthropogenic Influences in Fluvial Geomorphology. J. E. Costa, A. J. Miller, K. W. Potter and P. R. Wilcock. Washington D.C., American Geophysical Union. Geophysical Monograph 89: 219-239.
Jacobson, R. B. and K. B. Gran (1999). "Gravel sediment routing from widespread, low intensity landscape disturbance, Current River basin, Missouri."
During the last 160 years, land-use changes in the Ozarks have had the potential to cause widespread, low-intensity delivery of excess amounts of gravel-sized sediment to stream channels. Previous studies have indicated that this excess gravel bedload is moving in wave-like forms through Ozarks drainage basins. The longitudinal, areal distribution of gravel bars along 160 km of the Current River, Missouri, was evaluated to determine the relative effects of valley-scale controls, tributary basin characteristics, and lagged sediment transport in creating areas of gravel accumulations. The longitudinal distribution of gravel-bar area shows a broad scale wave-like form with increases in gravel-bar area weakly associated with tributary junctions. Secondary peaks of gravel area with 1.8-4.1 km spacing (disturbance reaches) are superimposed on the broad form. Variations in valley width explain some, but not all, of the short-spacing variation in gravel-bar area. Among variables describing tributary drainage basin morphometry, present-day land use and geologic characteristics, only drainage area and road density relate even weakly to gravel-bar areal inventories. A simple, channel network-based sediment routing model shows that many of the features of the observed longitudinal gravel distribution can be replicated by uniform transport of sediment from widespread disturbances through a channel network. These results indicate that lagged sediment transport may have a dominant effect on the synoptic spatial distribution of gravel in Ozarks streams; present-day land uses are only weakly associated with present-day gravel inventories; and valley-scale characteristics have secondary controls on gravel accumulations in disturbance reaches.
 
Jaeger, K. (2004). Channel-Initiation and Surface Water Expression in Headwater Streams of Different Lithology. College of Forest Resources. Seattle, University of Washington: 66.
Jaeger, K. L., et al. (2007). "Channel and perennial flow initiation in headwater streams: management implications of variability in source-area size." Environmental Management 40: 775-786.
Despite increasing attention to management of
headwater streams as sources of water, sediment, and wood
to downstream rivers, the extent of headwater channels and
perennial flow remain poorly known and inaccurately
depicted on topographic maps and in digital hydrographic
data. This study reports field mapping of channel head and
perennial flow initiation locations in forested landscapes
underlain by sandstone and basalt lithologies in Washington
State, USA. Contributing source areas were delineated
for each feature using a digital elevation model (DEM) as
well as a Global Positioning System device in the field.
Systematic source area–slope relationships described
in other landscapes were not evident for channel heads in
either lithology. In addition, substantial variability in
DEM-derived source area sizes relative to field-delineated
source areas indicates that in this area, identification of an
area–slope relationship, should one even exist, would
be difficult. However, channel heads and stream heads,
here defined as the start of perennial flow, appear to be
co-located within both of the lithologies, which together
with lateral expansion and contraction of surface water
around channel heads on a seasonal cycle in the basalt
lithology, suggest a controlling influence of bedrock
springs for that location. While management strategies for
determining locations of channel heads and perennial flow
initiation in comparable areas could assign standard source
area sizes based on limited field data collection within that
landscape, field-mapped source areas that support perennial
flow are much smaller than recognized by current Washington
State regulations.
 
Jaeggi, M. N. R. (1987). Interaction of bed load transport with bars. Sediment Transport in Gravel Bed Rivers. C. R. Thorne, J. C. Bathurst and R. D. Hey, John Wiley & Sons Ltd. Chapter 26: 829.
only chapter 26 abstract on file
 
Jain, A., K.P.Sudheer, and S.Srinivasulu (2004). "Identification of physical processes inherent in artificial neural network rainfall-runoff models." Hydrologic Processes 18: 571-581.
Jain, S. K. (2001). "Development of integrated sediment rating curves using ANN." Journal of Hydraulic Engineering-ASCE 127(1): 30-37.
Jain, S. K. and F. Dolezal (2000). "Modeling soil erosion using EPIC supported by GIS, Bohemia, Czech Republic."
The Environmental Productivity Impact Calculator (EPIC), a complex semi-empirical environmental model with distributed parameters, was used to estimate water erosion on 18 fields of a small (1.42 km super(2)) agricultural catchment called Cernici in a foothills region of Central Bohemia, Czech Republic. Some input data for EPIC (areas, elevations, lengths and slopes) and the field-to-field sediment delivery ratios were prepared using a Geographic Information System. Average erosion rates predicted by EPIC were highest in May to September if the Uniform Soil Loss Equations (USLE) was used. The MUSLE (modified USLE) and AOF (Onstead-Foster method) also showed high erosion rates in December-January. The largest simulated soil erosion rates were found on a few ploughed fields on which crop rotation prone to erosion combined unfavorably with high field slopes and highly erosive weather. A change of crop rotation helped reduce the erosion.
 
Jain, V., et al. (2007). "Where do floodplains begin? The role of total stream power and longitudinal profile form on floodplain initiation processes." Geological Society of America Bulletin 120(1/2): 127-131.
Understanding downstream transitions in
river character and behavior is a basic concept
in fl uvial geomorphology. Downstream
patterns of depositional processes can be differentiated
between channel and fl oodplain
components. In this study a generic set of
methods is used to analyze fl oodplain initiation
and continuity in relation to downstream
changes in total stream power (slope and
discharge) and longitudinal profi le form for
river courses in the upper Hunter catchment,
Australia. Absolute values of these controlling
factors are shown to be poor indicators
of threshold conditions at which fl oodplains
begin to form along river courses. Catchment-
scale patterns of stream power and the
form of longitudinal profi les provide better
predictors of this transitional zone. The total
stream power plot derived along longitudinal
profi les represented by a second-order
exponential curve has a bimodal pattern.
In most cases, fl oodplains begin to form in
a transition zone characterized by a trough
area within the bimodal stream power distribution.
This bimodal stream power pattern
provides a better means to identify this
transition in depositional processes along
longitudinal profi les than more conventional
single peak stream power analyses based on
fi rst-order exponential longitudinal profi les.
Indirect controls such as basin geology and
accommodation space also infl uence the initiation
and pattern of fl oodplains.
 
Jakeman, A. J. and G. M. Hornberger (1993). "How much complexity is warrented in a rainfall-runoff model?" Water Resources Research 29(8): 2637-2649.
Jakeman AJ, T. G., SG Beavis, L Zhang, CR Dietrich, and PF Crapper (1999). "Modelling upland and instream erosion, sediment and phosphorus transport in a large catchment." Hydrological Processes 13: 745-752.
Jakob, M. and H. Weatherly (2002). "A hydroclimatic threshold for landslide initiation on the North Shore Mountains of Vancouver, British Columbia." Geomorphology 54: 137-156.
James, L. A. (1991). "Incision and morphologic evolution of an alluvial channel recovering from hydraulic mining sediment." Geological Society of America Bulletin 103: 723-736.
James, L. A. (1997). "Channel incision on the lower American River, California, from streamflow gage records." Water Resources Research 33(3): 485-490.
James, L. A., et al. (2007). "Using LiDAR data to map gullies and headwater streams under forest canopy: South Carolina, USA." Catena 71: 132-144.
The southeastern Piedmont of the USA was severely gullied during the early 20th century. A thick canopy established by reforestation in
many areas now inhibits the identification or mapping of gullies by aerial photography or other conventional remote sensing methods. An
Airborne Laser-Scanning (ALS or LiDAR) mapping mission flown for the U.S. Forest Service in April, 2004 acquired bare-Earth
topographic data. This paper tests the ability of the ALS topographic data to identify headwater channels and gullies for two branching gully
systems in forested areas and to extract gully morphologic information. Comparisons are made with field traverses using differential GPS and
reference cross sections measured by leveling surveys. At the gully network scale, LiDAR data provide accurate maps – the best available –
with robust detection of small gullies except where they are narrow or parallel and closely spaced. Errors in mapping channel location and
network topological connectivity under forest canopy increase with attempts to identify smaller features such as large rills. The ability of
LiDAR data to map gullies and channels in a forested landscape should improve channel-network maps and topological models. At the gully
reach scale, attempts to use LiDAR data to extract gully cross-section morphologic information under forest canopy were less successful due
to systematic underestimation of gully depths and overestimation of gully top widths. Limited morphologic accuracy of the data set at this
scale may be due to low bare-Earth point densities, shadowing of gully bottoms, and filtering of topographic discontinuities during postprocessing.
The ALS data used in this study are not suitable for detailed morphometric analysis or subtle change detection to monitor gullies
or develop sediment budgets. Data collection may be improved by orienting flights over gullies and with increased point densities through
improved scanner technology or better filtering and software capabilities to differentiate between vegetation and ground surfaces.
 
Janda, R. J., et al. (1975). Watershed conditions in the drainage basin of Redwood Creek, Humboldt County, California, as of 1973. Menlo Park, California, United States Geological Survey.
Jarret, R. D. (1984). "Hydraulics of High-Gradient Streams." Journal of Hydraulic Engineering 110(11): 1519-1539.
Jarret, R. D. (1985). Determination of Roughness Coefficients for Streams in Colorado. Lakewood, Colorado, U.S. Geological Survey: 54.
Jarret, R. D. (1987). "Errors in slope-area computations of peak discharges in mountain streams." Journal of Hydrology 96: 53-67.
Jarret, R. D. and H. E. Malde (1987). "Paleodischarge of the late Pleistocene Bonneville Flood, Snake River, Idaho, computed from new evidence." Geological Society of America Bulletin 99: 127-134.
Jarrett, R. D. (1984). "hydraulics of high-gradient streams." J. Hydraul. Eng. 110(11): 1519-1539.
Jarrett, R. D. (1990). "Hydrologic and hydraulic research in mountain rivers." Water. Res. Bull. WARBAQ 26(3): 419-429.
Jarrett, R. D., et al. (1999). "Geomorphic estimates of rainfall, floods, and sediment runoff; applied to the 1996 wildfire, Buffalo Creek, Colorado." Abstracts with Programs - Geological Society of America 31(7): 313.
A basin-wide, monitoring approach using geomorphic techniques provides important information to draw isohyetal maps of rainstorms, assess the effects of wildfire on water and sediment runoff, assess the effects of watershed rehabilitation activities, determine watershed recovery time, and help manage forest ecosystems. The approach is flexible, requires minimal resources, and complements comprehensive instrumented monitoring in a short amount of time. The approach was applied to the mountain community of Buffalo Creek, Colorado, which had a catastrophic wildfire on May 18, 1996. Subsequent rainstorms produced 10 floods larger than a 100-year (pre-fire conditions) flood as well as numerous smaller floods. Rates of flood runoff in 1997 and 1998 were about the same as in 1996, which indicate persistent effects from the fire and minimal watershed recovery. The largest rainstorm, on July 12, 1996, of about 130 mm in an hour was about a 1,000-year rainstorm. This extreme rainstorm produced a flood about 10 times larger than the 100-year (pre-fire) flood, which was larger than the 1976 flash flood in the Big Thompson River Colorado. On Sand Draw, and other small tributaries in burned basins near the center of the storm, peak discharges exceeded the 1,000-yr flood. Unburned basins in areas of maximum rainfall had minimal water and sediment runoff. Study results were used to help the National Weather Service determine threshold-rainfall amounts that could produce flash flooding in the Buffalo Creek area. Large quantities of sediment continue to be transported in burned-area streams since the fire. Most of this sediment is deposited in Strontia Springs Reservoir, which is a major water supply for Denver located a few km downstream from the burned area. Investigations of alluvial sediments indicate at least 10 wildfire-flood sequences during about the past 2,500 years in the Buffalo Creek area. Study results indicate that prehistoric fires and subsequent increased runoff prior to fire suppression of the last century contribute to cyclical, geomorphic instability.
 
Jarrett, R. D. and J. E. Costa (1986). Hydrology, Geomorphology, and Dam-Break Modeling of the July 15, 1982 Lawn Lake Dam and Cascade Lake Dam Failures, Larimer County, Colorado, U.S. Geological Survey: 1-78.
Jassby, A. D., et al. (1992). "Trend, seasonality, cycle, and irregular fluctuations in primary productivity at Lake Tahoe, California-Nevada, USA " Zoobiologia 246: 195-203.
Jassby, A. D., et al. (1994). "Atmospheric deposition of nitrogen and phosphorous in the annual nutrient load of Lake Tahoe (California-Nevada) " Water Resources Research 30(7): 2207-2216.
Jefferson, A., et al. (2006). "Influence of volcanic history on groundwater patterns on the west slope of the Oregon High Cascades." Water Resources Research 42(W12411).
Spring systems on the west slope of the Oregon High Cascades exhibit complex
relationships among modern topography, lava flow geometries, and groundwater flow
patterns. Seven cold springs were continuously monitored for discharge and temperature
in the 2004 water year, and they were periodically sampled for d18O, dD, tritium, and
dissolved noble gases. Anomalously high unit discharges suggest that topographically
defined watersheds may not correspond to aquifer boundaries, and oxygen isotope data
reveal that mean recharge elevations for the springs are coincident with extensive
Holocene lava fields. The 3He/4He ratios in most of the springs are close to atmospheric,
implying shallow flow paths, and aquifer thicknesses are estimated to be 30–140 m.
Estimates using 3H/3He data with exponential and gamma distributions yield mean transit
times of  3–14 years. Recharge areas and flow paths are likely controlled by the
geographic extent of lava flows, and some groundwater may cross the Cascade crest.
 
 
Jefferson, A., et al. (2007). A river runs underneath it: geological control of spring and channel systems and management implications, Cascade Range, Oregon. Advancing the Fundamental Sciences: proceedings of the Forest Service national earth sciences conference. PNW-GTR-689. M. J. Furniss, C. F. Clifton and K. L. Ronnenberg. Portland, OR, U.S. Department of Agriculture, Forest Service, Pacific Northwest Research Station: 391-400.
Long-term sustainable management of Cascade Range watersheds requires an understanding of water
sources and discharge patterns from tributary streams, particularly those sourced in large-volume cold springs
of the High Cascades geologic province. Focusing on the McKenzie River watershed, measurements of
discharge and stream temperature combined with laboratory analysis of spring water isotopes improve our
understanding of spatial and temporal recharge and discharge patterns. Summer streamflow in the McKenzie
is dominated by water from approximately ten spring-fed streams, which maintain 4 to 7°C spring water
temperatures and relatively steady flow throughout the summer. In winter months, streams in the Western
Cascades geologic province respond rapidly to rain and rain-on-snow events and become the major water
source to the McKenzie River. Spring-fed streams also respond to precipitation events, but show muted and
delayed hydrograph peaks. Summer flow behavior varies among springs, even between those that are located
near each other. Isotopic data reveal that recharge to large springs occurs between 1300-1800 m in elevation,
which is coincident with geologically young lava between McKenzie and Santiam Passes. Recharge elevations
also suggest some disagreement between recharge areas and topographic watersheds of the springs. Because
of their importance to summer streamflow, water quality, and habitat in the McKenzie River basin, water
resources decision-making must differentiate between spring-fed and runoff-dominated streams.
 
 
Jefferson, A., et al. (2008). "Hydrogelologic controls on streamflow sensitivity to climate variation." Hydrological Processes.
Climate models project warmer temperatures for the north-west USA, which will result in reduced snowpacks and decreased
summer streamflow. This paper examines how groundwater, snowmelt, and regional climate patterns control discharge at
multiple time scales, using historical records from two watersheds with contrasting geological properties and drainage
efficiencies. In the groundwater-dominated watershed, aquifer storage and the associated slow summer recession are responsible
for sustaining discharge even when the seasonal or annual water balance is negative, while in the runoff-dominated watershed
subsurface storage is exhausted every summer. There is a significant 1 year cross-correlation between precipitation and
discharge in the groundwater-dominated watershed (r D 0Ð52), but climatic factors override geology in controlling the interannual
variability of streamflow. Warmer winters and earlier snowmelt over the past 60 years have shifted the hydrograph,
resulting in summer recessions lasting 17 days longer, August discharges declining 15%, and autumn minimum discharges
declining 11%. The slow recession of groundwater-dominated streams makes them more sensitive than runoff-dominated
streams to changes in snowmelt amount and timing.
 
 
Jelinek, R., et al. (2002). "Mechanical properties and behavior of soils under the triaxial test, the Yamauchi landslide, Japan." Bulletin of Engineering, Geology, and the Environment 61: 207-212.
Jennings, S. B., et al. (1999). "Assessing forest canopies and understorey illumination: canopy closure, canopy cover and other measures." Forestry 72(1): 59-73.
Jenson, S. K. (1991). "Applications of hydrologic information automatically extracted from digital elevation models." Hydrological Processes 5: 31-44.
Jenson, S. K. and J. O. Domingue (1988). "Extracting topographic structure from digital elevation data for geographic information system analysis." Photogrammetric Engineering and Remote Sensing 54(11): 1593-1600.
Jian, L. (1985). Present Situation and Future Undertaking of Mudflow Research in China. Fourth International Conference and Workshop on Landslides, Tokyo, Japan.
Johnson, A. (?). "Rheological Properties of Debris, Ice, and Lava."
Johnson, A. C., et al. (2000). "Landslide initiation, runout, and deposition within clearcuts and old-growth forests of Alaska." Journal of American Water Resources Association 36(1): 17-30.
Johnson, A. M. (1965). A Model for Debris Flow (DRAFT), The Penn State University.
Johnson, A. M. (1970). Method and Geologic Problems Physical Processes in Geology: 28.
Johnson, A. M. and P. H. Rahn (1970). "Mobilization of debris flows." Zietschriff fur Geomorphologe 9: 168-186.
Johnson, A. M. and J. R. Rodine (1984). Debris Flow. Slope Instability. D. Brunsden and D. B. Prior, John Wiley and Sons Ltd. Chapter 8: 257-359.
Johnson, D. L. (1977). "The late quaternary climate of coastal California: evidence for an ice age refugium." Quaternary Research 8: 154-179.
Johnson, E. A. and S. L. Gutsell (1994). "Fire Frequency Models, Methods, and Interpretations." Advances in Ecological Research 25: 239-287.
Johnson, E. A. and C. E. Van Wagner (1984). "The theory and use of two fire history models." Can. J. For. Res. 15: 214-220.
Johnson, E. A. and C. E. Van Wagner (1985). "The theory and use of two fire history models." Canadian Journal of Forest Research 15: 214-220.
The objective of this paper is to explain the distributions, assumptions, interpretations, and relationships of the two compatible, stochastic models of fire history: the negative exponential and the Weibull. For each model the "fire interval" and "time-since fire" distributions are given. Both models apply to homogenous stationary stochastic processes. The negative exponential states that the instantaneous fire hazard rate is constant for all stand ages. The Weibull states that the instantaneous fire hazard rate increases with stand age when the shape parameter is > 1 (the negative exponential is a special case of the Weibull when shape = 1). An empirical method is given for separating from on observed fire history distribution, the pre- and post-fire suppression distributions. Four relationships are derived from the models and defined per study region (per stand): (i) the fire cycle (average fire interval), (ii) the annual percent burned area (fire frequency), (iii) the average age of the vegetation (average propective life-time), and (iv) the renewal rate.
 
Johnson, G. (1999, January 12, 1999). "Of Mice and Elephants: A Matter of Scale." Retrieved 10-22-2002, 2002, from http://www.nytimes.com/library/national/science/011299sci-scaling.html.
Johnson, M. H. and J. A. Elliott (1996). "Impacts of Logging and Wildlife on an Upland Black Spruce Community in Northwestern Ontario " Environmental Monitoring and Assessment 39: 283-297.
Johnson, P. A., et al. (1991). "Debris Basin Policy and Design." Journal of Hydrology 123: 83-95.
Johnson, P. A., et al. (1991). "Magnitude and frequency of debris flows." Journal of Hydrology 123: 69-82.
Johnson, S. L. (2004). "Factors influencing stream temperatures in small streams: substrate effects and a shading experiment." Canadian Journal of Fisheries and Aquatic Science 61: 913-923.
The temperature of stream water is an important control of many in-stream processes. To better understand
the processes and consequences of solar energy inputs to streams, stream temperature dynamics were examined before,
during, and after experimental shading of a 150-m reach of a second-order stream in the Oregon Cascade Range. Maxi-mum
water temperatures declined significantly in the shaded reach, but minimum and mean temperatures were not
modified. Heat budget calculations before shading show the dominance of solar energy as an influence of stream tem-perature.
The influence of substrate type on stream temperature was examined separately where the water flowed first
over bedrock and then through alluvial substrates. Maximum temperatures in the upstream bedrock reach were up to
8.6 °C higher and 3.4 °C lower than downstream in the alluvial reach. Better understanding of factors that influence
not only maximum but minimum temperatures as well as diurnal temperature variation will highlight types of reaches
in which stream temperature would be most responsive to changes in shading. Many apparent discrepancies in stream
temperature literature can be explained by considering variation in the relative importance of different stream tempera-ture
drivers within and among streams and over time.
 
Johnson, S. L., et al. (1997). Lessons from a Flood: An Integrated View of the February 1996 Flood in the McKenzie River Basin. Pacific Northwest Water Issues Conference, Portland, Oregon.
Johnson, S. L. and J. A. Jones (2000). "Stream temperature responses to forest harvest and debris flows in western Cascades, Oregon." Canadian Journal of Fisheries and Aquatic Science 57(Suppl. 2)(30-39).
Stream temperature controls the rates of many biotic and abiotic processes and is influenced by changes in
streamside land use practices. We compiled historic stream temperature data and reestablished study sites in three small
basins in the H.J. Andrews Experimental Forest in the western Cascades, Oregon, to reexamine the effects on and recovery
of stream temperatures following removal of riparian vegetation. Maximum stream temperatures increased 7°C
and occurred earlier in the summer after clear-cutting and burning in one basin and after debris flows and patch-cutting
in another. Diurnal fluctuations in June increased from approximately 2 to 8°C. Stream temperatures in both basins
gradually returned to preharvest levels after 15 years. The influence of the primary factor controlling stream temperatures,
shortwave solar radiation, was amplified following removal of riparian vegetation, and conduction between stream
water and nearby soils or substrates also appeared to be an important factor. Shifts in the timing of summer maxima
and greater increases in early summer stream temperatures could impact sensitive stages of aquatic biota.
 
Johnson, S. L., et al. (2005). "Effects of an increase in large wood on abundance and survival of juvenile salmonids (Oncorhynchus spp.) in an Oregon coastal stream." Canadian Journal of Fisheries and Aquatic Science 62: 412-424.
We examined the effect of an increase in large wood on the summer population size, smolt abundance, and
freshwater survival of steelhead (Oncorhynchus mykiss), coastal cutthroat trout (Oncorhynchus clarki clarki), and coho
salmon (Oncorhynchus kisutch). We examined these parameters for five brood years prior to the addition of wood and
five brood years after in Tenmile Creek, a direct ocean tributary on the Oregon coast. Over the same time frame, a
nearby reference stream, Cummins Creek, was also sampled for the same parameters. The input of large wood into
Tenmile Creek resulted from a planned habitat restoration project in 1996 and an unplanned addition of wood from a
winter storm the same year. Steelhead smolt abundance, steelhead freshwater survival, and coho salmon freshwater sur-vival
increased in Tenmile Creek after the input of large wood. Steelhead age-0+ summer populations and steelhead
smolt populations increased in the reference stream, although steelhead freshwater survival did not. Coho salmon popu-lations
remained unchanged in the reference stream. Our results illustrate the potential shortcomings of the before-after-control-
impact study design under field conditions and the potential for misinterpreting results had we employed a
more modest sampling plan.
 
Johnson, S. L., et al. (2000). "Riparian forest disturbances by a mountain flood - the influence of floated wood." Hydrological Processes 14: 3031-3050.
Large floods can have major impacts on riparian forests. Here we examine the variability and spatial distribution
of riparian forest responses along eight third- to fifth-order streams following a large flood ( ~ 100 year recurrence
interval) in the Cascade Mountain Range of Oregon. We categorized disturbance intensity (physical force) exerted
on riparian trees during floods into three classes: (i) purely fluvial (high water flow only); (ii) fluvial supplemented
by dispersed pieces of floating wood (uncongested wood transport); (iii) fluvial with movement of batches of wood
(congested wood transport). These types of material transport and associated classes of disturbance intensity
resulted in a gradient of biotic responses of disturbance severity ranging from standing riparian trees inundated by
high water, to trees toppled but still partially rooted, to complete removal of trees. High within-stream and among-
stream responses were influenced by pre-flood stream and riparian conditions as well as flood dynamics, especially
the availability of individual pieces or congested batches of wood.
Fluvial disturbance alone toppled fewer riparian trees than in reaches where floodwaters transported substantial
amounts of wood. Debris flows delivered additional wood and sediment to parts of reaches of four of these study
streams; riparian trees were removed and toppled for up to 1-5 km downstream of the debris flow tributary
channel. Congested wood transport resulted in higher frequency of toppled trees and greater deposition of new
wood levees along channel margins. The condition of the landscape at the time of a major flood strongly influenced
responses of riparian forests. Recent and historic land-use practices, as well as the time since the previous large
flood, influenced not only the structure and age of the riparian forests, but also the availability of agents of
disturbance, such as large pieces of floating wood, that contribute to disturbance of riparian forests during floods.
 
Johnson, S. L., et al. (2000). "Riparian forest disturbances by a mountain flood -- the influence of floated wood." Hydrological Processes 14: 3031-3050.
Johnson, S. Y. (1991). "Sedimentation and Tectonic Setting of the Chuckanut Formation, Northwest Washington." Washington Geology 19(4): 12-13.
Jones, J. (1992). Summary of the workshop on ecological principles for management of Pacific Northwest watersheds to protect fish, water, and riparian resources. Workshop on ecological principles for management of Pacific Northwest watersheds to protect fish, water, and riparian resources.
Jones, J. A., and G. Grant (1996). "Peak flow responses to clear-cutting and roads in small and large basins, western Cascades, Oregon." Water Resouces Research 32(4): 959-974.
Jones, J. A., F.J. Swanson, B.C. Wemple, and K.U. Snyder (2000). "Effects of roads  on hydrology, geomorphology, and disturbance patches in stream networks." Conservation Biology 14(1): 76-85.
Jones, J. A. (2000). "Hydrologic processes and peak discharge response to forest removal, regrowth, and roads in 10 small experimental basins, western Cascades, Oregon." Water Resources Research 36(9): 2621-2642.
Jones, J. A. and G. E. Grant (1993). "Peak flow responses to clear-cutting and roads in small and large basins, western Cascades, Oregon (DRAFT)." in presubmission review for WaterResources Research: 959-974.
Jones, J. A. and G. E. Grant (1996). "Peak flow responses to clear-cutting and roads in small and large basins, western Cascades, Oregon." Water  Resources Research 32(4): 959-974.
Jones, J. A., et al. (2000). "Effects of roads on hydrology, geomorphology, and disturbance patches in stream networks." Conservation Biology 14(1): 76-85.
Jones, J. L. (2006). "Side channel mapping and fish habitat suitability analysis using Lidar topography and orthophotography." Photogrammetric Engineering and Remote Sensing(11): 1202-1206.
Jones, K. H. (1998). "A comparison of two approaches to ranking algorithms used to compute hill slopes." GeoInformatica 2(3): 235-256.
The calculation of slope (downhill gradient) for a point in a digital elevation model (DEM) is a common
procedure in the hydrological, environmental and remote sensing sciences. The most commonly used slope
calculation algorithms employed on DEM topography data make use of a three-by-three search window or kernel
centered on the grid point (grid cell) in question in order to calculate the slope at that point. A comparison of eight
such slope calculation algorithms has been carried out using an arti®cial DEM, consisting of a smooth synthetic
test surface with various amounts of added Gaussian noise. Morrison's Surface III, a trigonometrically de®ned
surface, was used as the synthetic test surface. Residual slope grids were calculated by subtracting the slope grids
produced by the algorithms on test from true/reference slope grids derived by analytic partial differentiation of the
synthetic surface. The resulting residual slope grids were used to calculate root-mean-square (RMS) residual error
estimates that were used to rank the slope algorithms from ``best'' (lowest value of RMS residual error) to
``worst'' (largest value of RMS residual error). Fleming and Hoffer's method gave the ``best'' results for slope
estimation when values of added Gaussian noise were very small compared to the mean smooth elevation
difference (MSED) measured within three-by-three elevation point windows on the synthetic surface. Horn's
method (used in ArcInfo GRID) performed better than Fleming and Hoffer's as a slope estimator when the noise
amplitude was very much larger than the MSED. For the large noise amplitude situation the ``best'' overall
performing method was that of Sharpnack and Akin. The popular Maximum Downward Gradient Method (MDG)
performed poorly coming close to last in the rankings, for both situations, as did the Simple Method. A nonogram
was produced in terms of standard deviation of the Gaussian noise and MSED values that gave the locus of the
trade-off point between Fleming and Hoffer's and Horn's methods.
 
 
Jones, T. A. and L. D. Daniels (2008). "Dynamics of large woody debris in small streams disturbed by the 2001 Dogrib fire in the Alberta foothills." Forest Ecology and Management 256: 1751-1759.
We investigated the temporal dynamics of large woody debris (LWD) in five headwater streams before
and after the 2001 Dogrib fire in the foothills of Alberta. The density of LWD varied from 5 to 41 logs per
50 mof stream reach and accounted for 19.4   5.1 m3 ha 1 (mean   standard error) of wood in the riparian
zones and 114.1   30.1 m3 ha 1 of wood in the bankfull margins of the stream channel. Individual logs
averaged 18.9   1.15 cm in diameter, 5.5   0.7 m in length, and 0.2   0.02 m3 in volume. Logs became
significantly shorter in decay classes II–IV. Bridges were longer than partial bridges, which were longer than
loose and buried LWD. Individual log volume was greatest for bridges, but not significantly different among
other position classes. Bridges and loose LWD contributed little to stream morphology and function; however,
55% of partial bridges and all buried logs contributed to sediment storage, channel armouring, or riffles and
pools in the stream channel.
Using dendroecological methods, we estimated the year of death of 108 of 115 spruce logs. LWD
resulted from tree deaths that occurred between 1874 and 2001, so that time since death ranged from 5
to 132 years. Time since death increased from decay class II to III to IV and bridges were younger than
LWDin all other position classes. Due to high rates of recruitment after fire, 16.5% LWDrecruited between
2001 and 2006, most of which were bridges or partial bridges in decay class II. We anticipate a delay of
30–45 years before newly recruited logs contribute significantly to stream morphology and function.
Depletion rates of LWD were exponential, such that 50% of LWD would be lost to decay, erosion or
downstream transport within 30 years of tree death and<12% of LWDwould persistmore than 100 years.
Since recruitment of new LWD in post-fire lodgepole pine stands is delayed by ca. 40 years while trees
establish and stands develop, we anticipate periods of ca. 70 years between stand-replacing fires and
recruitment of new, functional LWD into stream channels. During this time, fire-killed snags are an
important source of LWD to small streams. For headwater streams in environments susceptible to floods
and erosion we recommend that buffer zones comprised of snags to be established after fires. The goal of
these post-fire buffers is to ensure a supply of LWD into streams for years to decades after a standreplacing
fire.
 
Joshi, J., et al. (1998). "Soil erosion and sediment disaster in Nepal-a review." J. Fac. Agr. 42(3.4): 491-502.
Joshi, J., et al. (1997). "Debris flow disaster in Harihara, Japan." Acta Geologica Universitatis Comenianae 52: 59-60.
Joy, M. K. and R. G. Death (2004). "Predictive modelling and spatial mapping of freshwater fish and decapod assemblages using GIS and neural networks." Freshwater Biology 49: 1036-1052.
1. We used stream fish and decapod spatial occurrence data extracted from a national
database and recent surveys with geospatial landuse data, geomorphologic, climatic, and
spatial data in a geographical information system (GIS) to model fish and decapod
occurrence in the Wellington Region, New Zealand.
2. To predict the occurrence of each species at a site from a common set of predictor
variables we used a multi-response, artificial neural network (ANN), to produce a single
model that predicted the entire fish and decapod assemblage in one procedure.
3. The predictions from the ANN using this landscape scale data proved very accurate
based on evaluation metrics that are independent of species abundance or probability
thresholds. The important variables contributing to the predictions included the latitudinal
and elevational position of the site reach, catchment area, average air temperature, the
vegetation type, landuse proportions of the catchment, and catchment geology.
4. Geospatial data available for the entire regional river network were then used to create a
habitat-suitability map for all 14 species over the regional river network using a GIS. This
prediction map has many potential uses including: monitoring and predicting temporal
changes in fish communities caused by human activities and shifts in climate, identifying
areas in need of protection, biodiversity hotspots, and areas suitable for the reintroduction
of endangered or rare species.
 
Junk, W. J., et al. (1989). The Flood Pulse Concept in River-Floodplain Systems International Large River Symposium.
Junk, W. J. and K. M. Wantzen (2004). The flood pulse concept: new aspects, approaches, and applications - an update. Proceedings of the Second International Symposium on the Management of Large Rivers for Fisheries, RAP Publication 2004/16. R. L. Welcomme and T. Petr. Bangkok, Food and Agriculture Organization & Mekong River Commission. FAO Regional Office for Asia and the Pacific. 2: 117-149.
The flood pulse concept (FPC), published in 1989, was based on the scientific experience of the authors
and published data worldwide. Since then, knowledge on floodplains has increased considerably, creating a
large database for testing the predictions of the concept. The FPC has proved to be an integrative approach for
studying highly diverse and complex ecological processes in river-floodplain systems; however, the concept
has been modified, extended and restricted by several authors. Major advances have been achieved through
detailed studies on the effects of hydrology and hydrochemistry, climate, paleoclimate, biogeography, biodiversity
and landscape ecology and also through wetland restoration and sustainable management of floodplains
in different latitudes and continents. Discussions on floodplain ecology and management are greatly
influenced by data obtained on flow pulses and connectivity, the Riverine Productivity Model and the Multiple
Use Concept. This paper summarizes the predictions of the FPC, evaluates their value in the light of recent
data and new concepts and discusses further developments in floodplain theory.
 
K. M. Burnett, D. J. M., Rick Guritz, Mark A. Meleason, Ken Vance-Borland, Rebecca Flitcroft, Matthew J. Nemeth, Justin Priest, Nicholas A. Som, Christian E. Zimmerman (2013). "2011 Arctic-Yukon-Kuskowkwin Sustainable Salmon Initiative " Project Final Report: 79.
Kahler, T. H., et al. (2001). "Summer movement and growth of juvenile anadromous salmonids in small western Washington streams." Canadian Journal of Fisheries and Aquatic Science 58: 1947-1956.
Movements of juvenile coho salmon (Oncorhynchus kisutch), cutthroat trout (Oncorhynchus clarki clarki), and
steelhead trout (Oncorhynchus mykiss) were studied by observations and recapture of marked individuals in three western
Washington streams to test the hypotheses that few fish would move, downstream movement would predominate, movers
would be initially smaller and grow slower after movement than residents, and habitat quality would influence movement.
Contrary to predictions, from 28 to 60% of marked fish moved at least one habitat unit, and immigration of unmarked
fish also indicated considerable movement. Upstream movement predominated but the stream with the step-pool/cascade
channel type had fewer upstream movers and greater distances moved downstream. Coho movers were not smaller than
nonmovers, as predicted based on assumptions that movement results from competitive exclusion. Habitat units that coho
left were smaller and shallower but lower in density than units where coho remained. Thus, movement is a common phenomenon
rather than an aberration, and may reflect habitat choice rather than territorial eviction. Moreover, movers grew
faster than nonmovers, so the “mobile fraction” of the population was not composed of competitively inferior fish but
rather individuals that thrived. The phenomenon of small-scale habitat- and growth-related movements should be considered
when planning and interpreting studies of juvenile salmonid ecology in streams.
 
Kalkhoff, S. J. (1993). "Using a Geographic Information System to Determine the Relation Between Stream Quality and Geology in the Roberts Creek Watershed, Clayton County, Iowa " Water Resources Bulletin 29(6): 989-996.
Kalliola, R. and M. Puharta (1988). "River dynamics and vegetation mosaicism: a case study of the River Kamajohka, northernmost Finland." Journal of Biogeography 15: 703-719.
Kampf, S. K., and S.J. Burges (2007). "A framework for classifying and comparing distributed hillslope and catchment hydrologic models." Water Resources Research 43: W05423, doi:05410.01029/02006WR005370.
Kappesser, G. (1991). The Use of the Wolman Counts and Channel Geometry to Define Channel Stability (Review Draft). Coeur d'Alene, Idaho, Idaho Panahandle National Forests: 4.
Kappesser, G. B. and P. Diplas (2002). "A riffle stability index to evaluate sediment loading to streams." Journal of the American Water Resources Association 38(4): 1069-1082.
Karlin, R. (1980). "Sediment Sources and Clay Mineral Distributions Off the Oregon Coast." Journal of Sedimentary Petrology 50(2): 543-560.
Karr, J. R. (1991). "Biological integrity: a long-neglected aspect of water resource management." Ecological Applications 1(1): 66-84.
Karr, J. R. (1996). "Thinking about Salmon Landscapes " The Osprey 26: 2-6.
Karr, J. R. and E. W. Chu (1997). "Biological Monitoring: Essential Foundation for Ecological Risk Assessment " Human and Ecological Risk Assessment 3(6): 993-1004.
Karwan, D. L., J.A. Gravelle, and  Hubbart, J.A., (2007). "Effects of timber harvest on suspended sediment loads in Mica Creek, Idaho." Forest Science 53(2): 181-188.
Kasahara, T. and S. M. Wondzell (2003). "Geomorphic controls on hyporheic exchange flow in mountain streams." Water Resources Research 39(1).
Hyporheic exchange flows were simulated using MODFLOW and MODPATH to
estimate relative effects of channel morphologic features on the extent of the hyporheic
zone, on hyporheic exchange flow, and on the residence time of stream water in the
hyporheic zone. Four stream reaches were compared in order to examine the influence of
stream size and channel constraint. Within stream reaches, the influence of pool-step or
pool-riffle sequences, channel sinuosity, secondary channels, and channel splits was
examined. Results showed that the way in which channel morphology controlled
exchange flows differed with stream size and, in some cases, with channel constraint.
Pool-step sequences drove hyporheic exchange in the second-order sites, creating
exchange flows with relatively short residence times. Multiple features interacted to drive
hyporheic exchange flow in the unconstrained fifth-order site, where pool-riffle sequences
and a channel split created exchange flows with short residence times, whereas a
secondary channel created exchange flows with long residence times. There was relatively
little exchange flow in the bedrock-constrained fifth-order site. Groundwater flow
models were effective in examining the morphologic features that controlled hyporheic
exchange flow, and surface-visible channel morphologic features controlled the
development of the hyporheic zone in these mountain streams.
 
 
Kashahara, T. and S. M. Wondzell (2003). "Geomorphic controls on hyporheic exchange flow in mountain streams." Water  Resources Research 38(0): 1-14.
Kates, R. W., et al. (2001). "Sustainability Science." Science 292: 641-642.
Kattelmann, R. (1999). Proposed fire management strategies and potential hydrologic effects in the Sierra Nevada. American Water Resources Association Technical Publication Series TPS, vol.99-2. E. Kendy. Bethesda, AWRA - American Water Resources Association: 43-48.
After several decades of remarkably thorough fire suppression, the mixed conifer forests of the Sierra Nevada are unnaturally dense and in a state of high susceptibility to stand-replacing fire. These large, high-intensity fires can greatly increase the delivery of water, sediment, and nutrients to streams and severely impact aquatic life. A variety of strategies to deal with the fire risk have been proposed in recent years. All the proposals have mixtures of possible consequences to the aquatic system. Most of the dense forest areas are probably producing less water and sediment than before the Gold Rush of 150 years ago. Various types and intensities of prescribed fire and mechanical removal will tend to increase streamflow and sediment delivery. Those effects must be weighed against the risk of more severe consequences of intense, widespread fires. The full range of issues involving vegetation management, fire risk, and related aquatic effects continues to be actively debated in the Sierra Nevada.
 
Kaufman, P. R. (1987). Channel Morphology and Hydraulic Characteristics of Torrent-Impacted Forest Streams in the Oregon Coast Range, U. S. A. Forest Engineering: 2.
Kavvas, M. L. (1999). "On the coarse-graining of hydrologic processes with increasing scales." Journal of Hydrology 217: 191-202.
Keane, R. E., et al. (2004). "A classification of landscape fire succession models: spatial simulations of fire and vegetation dynamics." Ecological Modeling 179: 3-27.
A classification of spatial simulation models of fire and vegetation dynamics (landscape fire succession models or LFSMs) is
presented. The classificationwas developed to provide a foundation for comparing models and to help identify the appropriate fire
and vegetation processes and their simulation to include in coarse scale dynamic global vegetation models. Other uses include a
decision tool for research and management applications and a vehicle to interpret differences between LFSMs. The classification
is based on the four primary processes that influence fire and vegetation dynamics: fire ignition, fire spread, fire effects, and
vegetation succession. Forty-four LFSMs that explicitly simulated the four processes were rated by the authors and the modelers
on a scale from 0 to 10 for their inherent degree of stochasticity, complexity, and mechanism for each of the four processes. These
ratings were then used to group LFSMs into similar classes using common ordination and clustering techniques. Another database
was created to describe each LFSM using selected keywords for over 20 explanatory categories. This database and the ordination
and clustering results were then used to create the final LFSM classification that contains 12 classes and a corresponding key.
The database and analysis results were used to construct a second classification key so managers can pick the most appropriate
model for their application based on computer resources, available modeling expertise, and management objective.
 
Keane, R. E., et al. (1996). FIRE-BGC - a mechanistic ecological process model for simulating fire succession on coniferous forest landscapes of the Northern Rocky Mountains. Fort Collins, CO, USDA Forest Service.
Keaton, J. R., et al. (1988). Assessing Debris Flow Hazards on Alluvial Fans in Davis County, Utah. 24th Symposium on Engineering Geology and Soils Engineering, Coeur D'Alene, Idaho, Department of Civil and Environmental Engineering.
Keefer, D. K. (1994). "The importance of earthquake-induced landslides to long-term slope erosion and slope-failure hazards in seismically active regions." Geomorphology 10(1994): 265-284.
Keefer, D. K., et al. (1987). "Real-Time Landslide Warning During Heavy Rainfall." Science 238: 921-925.
Keen, F. P. (1937). "Climatic cycles in Eastern Oregon as inicated by tree rings." Monthly Weather Review 65(5): 175-188.
Keim, R. F. and A. E. Skaugset (2002). "Physical aquatic habitat I. Errors associated with measurement and esitmation of residual pool volumes." North American Journal of Fisheries Management 22: 145-150.
Residual pools have been used as a measure of physical aquatic habitat, but mea-surement
techniques usually entail estimates or unquantified errors. In order for residual pools to
be a useful metric of stream channels, it is important for the errors associated with their mea-surement
to be quantified. We used precise, digital terrain models of a third-order mountain stream
in Oregon to identify the sensitivity of measurements to error and the consequences of those errors
in calculating the volumes of residual pools. Precise quantification of the elevations of the crests
of riffles is critical for precise measurement of the volume of individual residual pools because
they serve as control points for the water surface elevations of residual pools. Reach-level estimates
of the total volume of residual pools, however, are relatively insensitive to errors incurred in
measuring individual pools. This insensitivity indicates that reach-level measurements of residual
pool volume may be more appropriate than measurements of individual pools as metrics in mon-itoring
schemes that use low-precision measurement techniques.
 
Keim, R. F. and A. E. Skaugset (2003). "Modelling effects of forest canopies on slope stability." Hydrological Processes 17: 1457-1467.
We investigated the potential effects of rainfall intensity smoothing by forest
canopies on slope stability by modelling soil responses to measured rainfall
and throughfall during high-intensity rain. Field measurements showed that
maximum intensities of precipitation were generally reduced under forest
canopies at two sites in the Pacific Northwest, USA. Modelling soil water pore-pressure
responses of a hypothetical hillslope to the field data resulted in esti-mates
of slope stability that were generally greater under forest canopy than
for the same hillslope without forest canopy. Results indicate that smooth-ing
of precipitation intensities may translate into overall greater stability of
hillslopes under forest canopies.
 
Keim, R. F., et al. (2000). "Dynamics of coarse woody debris placed in three Oregon streams." Forest Science 46(1): 13-22.
Keim, R. F., et al. (2002). "Physical aquatic habitat II. Pools and cover affected by large woody debris in three western Oregon Streams." North American Journal of Fisheries Management 22: 151-164.
Abstract.—Large woody debris (LWD) is important in affecting stream channel morphology and
aquatic habitat. Although the greatest effects on streams of the Pacific Northwest have been by
LWD from large conifers, many riparian forests in the region are dominated by red alder Alnus
rubra. The effects of the small size and short life of LWD from red alders on channel morphology
may be different from that of conifers and are poorly understood. We added LWD (primarily red
alder) to three third-order streams in the Oregon Coast Range and used digital terrain models to
evaluate physical habitat for salmonids over 3 years. Total residual pool volume increased in two
streams, but in the one with the lowest gradient it did not change in the treated portion and even
decreased in the untreated portion. In all streams, both the relative proportion and absolute amount
of residual pool volume from deep pools increased from their pretreatment values. Cover from
LWD in pools increased after treatment and remained high, but the absolute amount of cover was
poorly predicted by the volume of LWD. Overall, the changes in stream channel morphology and
habitat were consistent with the effects of LWD, and these case studies indicate that small, red
alder LWD can effectively modify physical aquatic habitat.
 
Keinholz, H., et al. (1984). "Stability, instability, and conditional instability: Mountain ecosystem concepts based on a field survey on the Kakani area in the middle hills of Nepal." Mountain Research and Development 4(1): 55-62.
Keller, E. A., D.W. Valentine, and D.R. Gibbs (1997). "Hydrological response of small watersheds following the southern California Painted Cave Fire of June 1990." Hydrological Processes 11: 401-414.
Keller, E. A. and A. MacDonald (1995). River Channel Change: The Role of Large Woody Debris. Changing River Channels. A. Gurnell and G. Petts, John Wiley & Sons Ltd.: 217-235.
Keller, E. A. and T. Tally (1979). Effects of large organic debris on channel form and fluvial processes in the coastal redwood environment. Adjustments of the Fluvial System, Tenth Annual Geomorphology Symposium. D. D. Rhodes and G. P. Williams. SUNY, Binghamton, New York, Kendall Hunt Publishing Co., Dubuque, Iowa: 169-198.
Kellerhals, R. and D. I. Bray (1971). "Sampling procedures for coarse fluvial sediments." Journal of the Hydraulics Division 97(8): 1165-1180.
Kellerhals, R., et al. (1976). "Classification and analysis of river processes." Journal of Hydraulics Division 102: 813-829.
Kellert, S. H. (1994). In the Wake of Chaos: Unpredictable Order in Dynamical Systems. Chicago, University of Chicago Press.
Kelly, J. R. and M. A. Harwell (1990). "Indicators of ecosystem recovery." Environmental Management 14(5): 527-545.
Kelsey, H. M. (1980). "A sediment budget and an analysis of geomorphic process in the Van Duzen River basin, north coastal California, 1941-1975: Summary." Geological Society of America Bulletin 91: 190-195.
Kelsey, H. M., et al. (1987). "Stochastic model for the long-term transport of stored sediment in a river channel." Water Resources Research 23(9): 1738-1750.
Kelsey, H. M., et al. (1987). "Stochastic model for the long-term transport of stored sediment in a river channel." Water Resources Research 23(9): 1738-1750.
Kelsey, H. M., et al. (1988). Response of a Gravel-Bed River to a Debris Flow: Skagit County, Washington. Bellingham, WA, Western Washington University submitted to U.S. Forest Service: 18.
Kelsey, K. A. and S. D. West (1998). Riparian wildlife. River Ecology and Management: Lessons from the Pacific Coastal Ecoregion. R. N. a. R. Bilby, Springer: 235 - 252.
Kennard, P. and G. Pess (1994). Forest Management and Stream Degradation in Montague Basin - A Watershed Assessment. Marysville, Tulalip Tribes' Environmnetal Department.
Kennard, P., et al. (1997). Riparian-in-a-Box: A manager's tool to predict the impacts of riparian management on fish habitat.
Kennard, P., et al. (1998). Riparian-in-a-Box: A manager's tool to predict the impacts of riparian management on fish habitat. Forest-fish conference: land management practices affecting aquatic ecosystems, Calgary, Canada.
Kennedy, R. S. H. and T. A. Spies (2005). "Dynamics of hardwood patches in a conifer matrix: 54 years of change in a forested landscape in Coastal Oregon, USA." Biological Conservation 122: 363-374.
Changes to minor patch types in forested landscapes may have large consequences for forest biodiversity.The e ffects of forest
management and environment on these secondary patch types are often poorly understood.For example,do early-to-mid succes-
sional minor patch types become more expansive as late successional forest types are fragmented or do they also become more frag-
mented in managed landscapes?We evaluated the dynamics of early-to-mid successional hardwood patches in a conifer-dominated
landscape in relation to environment and land ownership in the central Coast Range of Oregon,USA,from the time of early logging
to the present-day using scanned and georeferenced aerial photographs and a GIS.Hardwood patches declined in size,number,total
area,and within-patch cover-type heterogeneity,and became more irregular in shape.Patch turnover and fragmentation was high,
with most patches present at the historical date disappearing by the present-day.Land ownership was important to hardwood patch
dynamics:hardwoods declined on lands owned by the USDA Forest Service,increased on non-industrial private lands,and were at
similar levels at both dates on private forest industry lands.Patch locations became more restricted to near-stream,lower elevation
areas where hardwoods are competitive.The relatively extensive distribution of hardwood patches at the historical date probably
resulted from earlier fire,selective logging,and grazing.In recent decades,forest management that includes fire suppression and
intensive management,and ecological constraints have resulted in a landscape in which early-to-mid successional hardwood patches
have been reduced in size,fragmented,and restricted to particular locales.
 
Keppeler, E. T. and R. R. Ziemer (1990). "Logging effects on streamflow: water yield and summer low flows at Caspar Creek in northwestern California." Water Resources Research 26(7): 1669-1679.
Streamflow data for a 21-year period were analyzed to determine the effects of selective tractor
harvesting of second-growth Douglas fir and redwood forest on the volume, timing, and duration of
low flows and annual water yield in northwestern California. The flow response to logging was highly
variable. Some of this variability was correlated with antecedent precipitation conditions. Statistically
significant increases in streamflow were detected for both the annual period and the low-flow season.
Relative increases in water yield were greater for the summer low-flow period than for annual flows,
but these summer flow increases generally disappeared within 5 years.
 
 
Ketcheson, G. and H. A. Froehlich (1978). Hydrologic factors and environmental impacts of mass soil movements in the Oregon Coast Range. Corvallis, OR, Water Resources Research Institute: 94.
Ketcheson, G. and H. A. Froelich (1978). Hydrologic factors and environmental impacts of mass soil movements in the Oregon Coast Range. Corvallis, Oregon, Water Resources Research Institute, Oregon State University: 94.
Ketcheson, G. L., and W. F. Megahan (1996). Sediment production and downslope sediment transport from forest  roads in granitic watersheds. Ogden, UT, USDA Forest Service: 16.
This paper presents the results of a study from highly erodible Idaho soils and provides information necessary to develop road design criteria and evaluate risks and trade-offs. Objectives of the study were to: quantify the volume of sediment deposition on slopes in relation to road features; determine the probability distribution of sediment travel distance; evaluate how sediment deposition occurs with respect to forest site conditions and downslope location of streams; determine the volume and particle size distribution of sediment deposits on slopes in relation to sediment travel distance; and investigate time trends in sediment deposition.
 
Ketcheson, G. L., et al. (1999). ""R1-R4" and "BOISED" sediment prediction model tests using forest roads in granitics." Journal of the American Water Resources Association 35(1): 83-98.
Kienzle, S. (2004). "The effect of DEM raster resolution on first order, second order and compound terrain derivatives." Transactions in GIS 8(1): 83-111.
Kiffney, P. M., et al. (2006). "Gradients in habitat heterogeneity, productivity, and diversity at tributary junctions." Canadian Journal Fisheries and Aquatic Sciences 63: 2518-2530.
Kiffney, P. M., et al. (2006). "Tributary streams create spatial discontinuities n habitat, biological productivity, and diversity in mainstem rivers." Canadian Journal of Fisheries and Aquatic Science 63: 2518-2530.
Kiffney, P. M. and J. S. Richardson (2001). "Interactions among nutrient, periphyton, and invertebrate and vertebrate grazers in experimental channels." Copeia: 422-429.
Kiffney, P. M., et al. (2004). "Establishing light as a causal mechanism structuring stream communities in response to experimental manipulation of riparian buffer width." Journal of the North American Benthologic Society 23(3): 542-555.
Previous studies on the effects of logging on streams have suggested that light and water
temperature were important variables structuring stream communities but, in many cases, these
effects were confounded. We observed pronounced gradients in the flux of solar energy and water
temperature in an earlier large-scale experiment in which we manipulated the width of riparian
buffers along headwater streams. Associated with these abiotic changes were increases in periphyton
biomass and primary consumer abundance. We present results from a study in streamside channels
that was designed to isolate the effects of light on stream communities, while holding water tem-perature
constant. Light treatments in the channel experiment simulated inputs of solar radiation
created during the prior watershed-scale experiment. Results from the present study suggested that
consumers limited periphyton biomass early in the study; however, a rainstorm midway through the
experiment reduced periphyton biomass and insect consumer abundance. Following this disturbance,
chlorophyll a biomass was 2 to 4 times higher in the full sunlight treatment compared to the 2 lowest
light treatments. At the end of the study, primary consumer abundance, biomass, survival, and growth
rate were positively related to light and periphyton resources. Therefore, we inferred biotic control
of periphyton during the early part of the channel study, whereas light appeared to control periph-yton
at the end of the study. Results from the large-scale and channel experiments suggested that
light was the primary constraint on periphyton biomass accrual. Moreover, both experiments, espe-cially
the channel study, showed that light indirectly influenced consumer performance as mediated
by increased primary production.
 
Kiffney, P. M. and P. Roni (2007). "Rrelationships between productivity, physical habitat, and aquatic invertebrate and vertebrate populations of forest streams: an information-theoretic approach." Transactions of the American Fisheries Society 136: 1088-1103.
Kilgore, B. M. (1973). "The Ecological Role of Fire in Sierran Conifer Forests: It's Application to National Park Management." Journal Quaternary Research 3(3).
Kilgore, B. M. (1979). The role of fire frequency and intensity in ecosystem distribution and structure: western forests and scrublands. Fire Regimes and ecosystem properties. H. A. Mooney, T. M. Bonnicksen, N. L. Christensen, J. E. Lotan and W. A. Reiners: 55-89.
Kim, H., et al. (2000). "Volumetric changes in fine fluvial material linking with forest disturbance of small catchments." Chikei = Transactions - Japanese Geomorphological Union 21(1): 57.
Kimble, L. A. and T. A. Wesche (1975). Relationships between selected physical parameters and benthic community structure in a small mountain stream. Laramie, Wyoming, Water Resources Research Institute: 60.
King, A. W. (?). Translating Models Across Scales in the Landscape. Chapter 19: 479-517.
King County (1987). Appendix II : Geology and Geomorphology of Western King County (DRAFT): 55.
King, J., et al. (1989). "The 1985 Bairaman landslide dam and resulting debris flow, Papua New Guinea." The Quarterly Journal of Engineering Geology 22: 257-270.
Kirchner, J. W., R.C.Finkel, C.S. Riebe, D.E. Granger, J.L. Clayton, J.G. King, and W.F. Megahan (2001). "Mountain erosion over 10 yr, 10 k.y., and 10 m.y. time scales." Geology 29(7): 591-594.
Kirchner, J. W., et al. (2001). "Mountain erosion over 10 yr, 10 k.y., and 10 m.y. time scales." Geology 29(7): 591-594.
Kirchner, J. W., et al. (2001). "Mountain erosion over 10 yr, 10 k.y., and 10 m.y. time scales." Geology 29(7): 591-594.
Kirkby, M. J. (1976). Hydrological Slope Models: The Influence of Climate. Geomorphology and Climate. E. Derbyshire, Wiley, John. Chapter 8: 247-267.
Kirkby, M. J. (1987). General Models of Long-Term Slope Evolution Through Mass Movement. Slope Stability. M. G. Anderson and K. S. Richards, John Wiley. Chapter 11: 359-379.
Kirkby, M. J. (1987). "The hurst effect and its implications for extrapolating process rates." Earth Surface Processes and Landforms 12: 57-67.
Kirkby, M. J. (1991). Sediment travel distance as an experimental and model variable in particulate movement. Erosion, Transport, Deposition Processes. H. R. Bork, J. d. Ploey and A. P. Schick. Cremlingen, Catena Verlag. Catena Supplement 19: 111-128.
Kirkby, M. J. (1992). An erosion-limited hillslope evolution model. Functional Geomorphology. K. M. Schmidt and J. de Ploey. Cremlingen, Catena Verlag. Catena Suppliment 23: 157-187.
Kirkby, M. J. (1993). Network Hydrology and Geomorphology. Channel Network Hydrology. K. Beven and M. J. Kirkby. Chichester, John Wiley & Sons: 1-11.
Kirkby, M. J., A.C.Imeson, G. Bergkamp, and L.H. Cammeraat (1996). "Scaling up processes and models from the field plot to the watershed and regional areas." Journal of Soil and Water Conservation 51(5): 391-398.
Kirkby, M. J. (1997). "Topmodel: A personal view." Hydrological Processes 11: 1087-1097.
Kirkby, M. J. (2001). From plot to continent: reconciling fine and coarse scale erosion models. Sustaining the Global Farm, 10th International Soil Conservation Organization Meeting, May 1999, Purdue University and USDA-ARS National Soil Erosion Laboratory.
Kirkby, M. J., et al. (2000). "The development of land quality indicators for soil degradation by water erosion."
The paper describes a proposed methodology for estimating water erosion risk for large areas. The estimates are based on a one-dimensional hydrological balance and a physically based sediment transport model. Estimates of risk associated with a given storm amount are integrated over the frequency distribution of daily rainfalls to provide a properly weighted estimate of the average annual risk. Soil factors are estimated from textual classes using qualitative pedo-transfer functions. Climatic data are taken from interpolated gridded data. Topographic data are taken from global or local DEMs. The estimates are for sediment delivery to stream channels. The method has been applied to provide a preliminary estimate for France at a resolution of 250 m, and could be applied globally at a resolution of 1 km.
 
Kitamura, Y. and S. Namba (1981). The function of tree roots upon landslide prevention presumed through the uprooting test. 313. Ibaraki, Japan., For. and For. Prod. res. Inst.: 175-208.
Klamt, R. R. (2000). The Garcia experience; a sediment TMDL case study. Water Resources Center Report - Centers for Water and Wildland Resources, Report: 98. C. W. Slaughter. Water Resources Center, University of California: 29-34.
Salmon and steelhead play a significant role in California north coastal economy, philosophy, and politics. Author Mark Twain is credited with saying that in the West, whiskey is for drinking and water is to fight over. More recently on the north coast, the fight has focused on salmon and land use. Continued concern for and anadromous fisheries has turned attention from the water itself to the riparian zone and hill slopes of the steep erosive coastal mountains of northern California. The Garcia River is a coastal tributary located about 100 miles north of San Francisco Bay. It is forested with commercial conifers and hardwoods and supports farming and cattle and sheep ranching. Historic waves of logging activity at different levels of regard for the land and water resources coupled with erosive soils on steep slopes and high winter rainfall resulted in significant erosion and sedimentation. Concern over declining anadromous salmonid populations brought attention to sediment impacts in the Garcia River watershed. That focus and threat of a lawsuit prompted the development of a sediment reduction strategy (TMDL) that addresses habitat and channel structure in the waterways by requiring land-owners to submit erosion control plans. In May of 1998 the North Coast Regional Water Quality Control Board adopted a TMDL and implementation plan with the assistance of the US Environmental Protection Agency. Controllable sediment discharges are prohibited, and reductions of sediment delivery to streams from roads, timber harvest, and agriculture are required on a 40 year time table. Instream numeric targets that describe the desired future condition of the riparian area, stream channel, and fish habitat are used as goals to measure the success of the reductions over time. The development of the TMDL and, especially, the implementation plan were contentious and involved numerous public workshops and hearings over a 2 year period. However, landowner response and attitude and, subsequently, the nature of land use activities is slowly changing. Landowner inventories and monitoring will provide a physical assessment of watershed recovery. The response of the fisheries will tell the ultimate story.
 
Klein, R. D. (1998). Recent and Historical Changes in Channel Cross Sections at Selected Sites in the Van Duzen River Basin prepared for Tetra Tech, Inc.: 1-25.
Klemes, V. (1974). "The Hurst Phenomenon: A Puzzle?" Water Resources Research 10(4): 675-688.
Klemes, V. (1986). "Dilettantism in Hydrology: Transition or Destiny?" Water Resources Research 22(9): 1775-1885.
Kline, T. C., Jr., et al. (1990). "Recycling of Elements Tranported Upstream by Runs of Pacific Salmon: I. 8^15N and 8^13C Evidence in Sashin Creek, Southeastern Alaska " Canadian Journal of Fisheries and Aquatic Sciences 47: 136-144.
Klock, G. O. and J. D. Helvey (1976). Debris flows following wildfire in north central Washington. Proc. Third Federal Interagency Sed. Conf., Denver, Colorado.
Knapp, R. A. and H. K. Preisler (1999). "Is it possible to predict habitat use by spawning salmonids? A test using California golden trout (Oncorhynchus mykiss aguabonite) " Canadian Journal of Fisheries and Aquatic Sciences 56: 1576 - 1584.
Knapp, R. A. and V. T. Vredenburg (1996). "Spawning by California Golden Trout: Characteristics of Spawning Fish, Seasonal and Daily Timing, Redd Characteristics, and Microhabitat Preferences " Transactions of the American Fisheries Society 125: 519-531.
Knight, D. W. (1989). Hydraulics of Flood Channels. Floods: Hydrological, Sedimentological, and Geomorphological Implications. K. Beven and P. Carling. Birmingham, John Wiley and Sons. Chapter 6: 83-105.
Knighton, A. D. (1980). "Longitudinal changes in size and sorting of stream bed material in four English rivers." Geological Society of America Bulletin 91: 55-62.
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 Process and Landforms 14: 333-359.
The mining of alluvial tin in the Ringarooma basin began in 1875, reached a peak in 1900-20, and had virtually ceased by 1982. During that time 40 million m3 of mining waste were supplied to the main river, quickly replacing the natural bed material and requiring major adjustments to the channel.
    Based on estimates of sediment supply from more than 50 widely scattered mines and the frequency of flows capable of transporting the introduced load, the river's transport history is reconstructed using a mass-conservation model. Because of the lengthy time period (110 years) and river distance (75 km) involved, the model cannot predict detailed change but it does reproduce the main pattern of sediment movement in which succissive phases of aggradation and degradation progress downstream. Peak storage is predicted in that part of the river where braiding and anastomosis are best developed.
    Aggradation was most rapid in the upper reaches close to major supply points, becoming slower and later with distance downstream. Channel width increased by up to 300% where the valley floor was broad and braiding became relatively common. Bridges had frequently to be replaced. While bed levels were still rising in lower reaches, degradation began in upper ones, notably after 1950, and by 1984 had progressed downriver over 30 km. Rates of incision reached 0.5 m/yr, especially in the early 1970s when record high flows occurred. As a result of degradation the bed material became gravelly through either reexposure of the original bed or lag concentration of coarse fractions. Also a narrower unbraided channel has developed. The river is beginning to heal itself and upper reaches now have reasonable stable beds but at least another 50 years will be required for the river to cleanse its channel of mining debris.
 
Knighton, A. D. (1999). "Downstream variation in stream power." Geomorphology 29: 293-306.
Knighton, D. (1998). Fluvial Forms and Processes: A New Perspective. New York, Oxford University Press Inc.
Knighton, D. (1998). Fluvial Forms and Processess, Oxford University Press US.
Knighton, M. D., et al. (1989). Applications and research in sediment delivery and routing models in the USDA-Forest Service. International symposium on Sediment transport modeling, New Orleans, LA, United States, Aug. 14-18, 1989, New York, Am. Soc. Civ. Eng.
Knopp, C. (1993). Testing indices of cold water fish habitat, North Coast Regional Water Quality Control Board.
Kobor, J. S. and J. J. Roering (2004). "Systematic variation of bedrock channel gradients in the central Oregon Coast Range: Implications for rock uplift and shallow landsliding." Geomorphology 62: 239-256.
In tectonically active regions, bedrock channels play a critical role in dictating the pace of landscape evolution. Models of
fluvial incision into bedrock provide a means of investigating relationships between gradients of bedrock channels and patterns of
active deformation. Variations in lithology, orographic precipitation, sediment supply, and erosional processes serve to complicate
tectonic inferences derived from morphologic data, yet most tectonically active landscapes are characterized by these
complexities. In contrast, the central Oregon Coast Range (OCR), which is situated above the Cascadia subduction zone, has
experienced rock uplift for several million years, did not experience Pleistocene glaciation, boasts a relatively uniform lithology,
and exhibits minor variations in precipitation. Although numerous process-based geomorphic studies suggest that rates of erosion
across the OCR are relatively constant, it has not been demonstrated that bedrock channel gradients in the region exhibit spatially
consistent values. Analysis of broadly distributed, small drainage basins (f5–20 km) in the central OCR enables us to explore
regional variability in bedrock channel gradients resulting from differential rock uplift or other sources. Consistent with previous
studies that have documented local structural control of deformed fluvial terraces in the western portion of our study area, our data
reveal a roughly 20-km-wide band of systematically elevated channel slopes (roughly twice the background value), roughly
coincident with the strike of N–S-trending mapped folds. Although many factors could feasibly generate this pattern, including
variable rock strength, precipitation gradients, or temporal or spatial variations in forearc deformation, the elevated bedrock
channel slopes likely reflect differential rock uplift related to activity of local structures. Importantly, our analysis suggests that
rock uplift and erosion rates may vary systematically across the OCR. Although our calculations were focused on the fluvialdominated
portion of study basins, our results have implications for upstream areas, including unchanneled valleys that often serve
as source areas for long-runout debris flows. Zero-order basins (or topographic hollows) within theN–S-trending band of elevated
channel slopes tend to be steeper than adjacent areas and may experience more frequent evacuation by shallow landsliding. Thus,
this region of the OCR may be highly sensitive to land use practices and high-intensity rainstorms.
 
Kochel, R. C., et al. (1987). "Role of tree dams in the construction of pseudo-terraces and variable geomorphic response to floods in Little River valley, Virginia." Geology 15: 718-721.
Kocik, J. F. and C. P. Ferreri (1998). "Juvenile production variation in salmonids: population dynamics, habitat, and the role of spatial relationships." Canadian Journal of Fisheries and Aquatic Science 55(Suppl. 1): 191-200.
Anadromous Atlantic salmon (Salmo salar) exhibit a complex life history that requires the use of habitats
that span several different temporal and spatial scales. While fisheries scientists have investigated the various elements
of habitat and how they affect Atlantic salmon growth and survival, these studies typically focus on requisite
requirements for a single life history stage. Current advances in our understanding of salmonid populations in lotic
systems indicates that ignoring the spatial positioning of different habitats and dispersal capabilities of fish between
them may affect estimates of habitat quality and production of juvenile Atlantic salmon. Using the concepts of
juxtaposition and interspersion, we hypothesize that discrete functional habitat units (FHU) occur within river systems
and that the spatial structure of FHU affects fish production. We present a method to delineate FHU using habitat
maps, fish ecology, and spatial habitat characteristics. Utilizing a simulation model, we illustrate how modeling FHU
structure of spawning and rearing habitat in a river system can improve our understanding of juvenile Atlantic salmon
production dynamics. The FHU concept allows a flexible approach to more comprehensive analyses of the impacts of
habitat alterations, seasonal habitat shifts, and spatial ecology of salmonids at various scales.
 
Kockel, R. C. (1990). Humid fans of the Appalachian Mountains. Alluvial Fans: A Field Approach. A. H. Rachocki and M. Church, John Wiley & Sons, Ltd.: 109-129.
Kodama, Y. (1994). "Downstream changes in the lithology and grain size of fluvial gravels, the Watarase River, Japan: Evidence of the role of abrasion in downstream fining." Journal of Sedimentary Research A64(1): 68-75.
Kodama, Y. (1994). "Experimental study of abrasion and its role in producing downstream fining in gravel-bed rivers." Journal of Sedimentary Research A64(1): 76-85.
Koehler, G. M. H., M. G. (1989). "Influences of seasons on bobcats in Idaho." U. S. A. J. Wildl. Manage 53: 197-202.
Koehler, R. D., et al. (2002). The role of stored historic sediment in short-term sediment production, south fork Noyo River, Jackson State Demonstration Forest, California. Abstracts with Programs - Geological Society of America. 34: 90.
In the South Fork Noyo River (SFNR) watershed, coastal northern California, large volumes of historic sediment were delivered to channels in response to past logging operations. This sediment presently is stored beneath historic terraces and in the present-day channels. Because current USEPA estimates of the Total Maximum Daily Load (TMDL) for the Noyo River were developed using office methods rather than field measurements, we conducted geomorphic mapping on the SFNR valley floor to quantify the volume of sediment stored in the watershed. We quantified sediment volumes associated with pre-historic terraces, historic terraces, and the active channel along four 1-mi-long stream reaches. Additionally, we established ten streamflow and suspended sediment sampling locations to monitor water and sediment discharges. We identified 158,000 cu. yds. of sediment stored in the active channel and 68,000 cu. yds. of sediment stored beneath historic terraces. These volumes are an order of magnitude less than the volumes calculated for pre-historic terraces. Based on analysis of surveyed cross sections we speculate that approximately 43 to 72% of the sediment originally stored beneath historic terraces has been transferred to the channel by erosion. The present-day channel sediment is stored presently in large gravel bars and is mobilized primarily during winter flood events. Based on our channel mapping and hydrologic data, we infer that the largest suspended sediment loads are spatially coincident with the location of the greatest amount of stored channel sediment. Re-mobilized historic sediment appears to increase suspended sediment load, and may be a significant, unrecognized, sediment source. Thus, accurately mapping and quantifying channel deposits is a critical step for assessing sediment budgets, especially in TMDL studies attempting to relate upslope management to suspended sediment production.
 
Koler, T. E. (1992). Literature search of effects of timber harvest to deep-seated landslides. Olympia, Washington, Cooperative, Monitoring, Evaluation, and Research Steering Committee: 52.
Koler, T. E. and K. G. Neal (?). "Chestershire and Backdrop Timber Sales: Case Histories of the Practice of Engineering Geology in the Olympic National Forest." Washington Division of Geology and Earth Resources Bulletin 78 II: 933-944.
Komar, P. D. (1989). Flow-competence evaluations of the hydraulic parameters of floods: an assessment of the technique. Floods: Hydrological, Sedimentological and Geomorphological Implications. K. Beven and P. Carling, John Wiley & Sons: 107-134.
Kondolf, G. M. (1988). Salmonid spawning gravels: A geomorphic perspective on their size distribution, modification by spawning fish, and criteria for gravel quality John Hopkinds University: 396.
Kondolf, G. M. (1997). "Application of the Pebble Count: Notes on Purpose, Method, and the Variants " Journal of the American Water Resources Association 33(1): 79-87.
Kondolf, G. M. (1998). "Lessons learned from river restoration projects in California " Aquatic Conservation: Marine and Freshwater Ecosystems 8: 39-52.
Kondolf, G. M. (2000). "Assessing salmonid spawning gravel quality." Transactions of the American Fisheries Society 129(262-281).
Abstract.—Much of the recent literature on salmonid spawning gravels has been devoted to the
search for a single statistic drawn or computed from the streambed particle size distribution to
serve as an index of gravel quality. However, a natural gravel mixture cannot be fully described
by any single statistic, because gravel requirements of salmonids differ with life stage, and thus
the appropriate descriptor will vary with the functions of gravel at each life stage. To assess
whether gravels are small enough to be moved by a given salmonid to construct a redd, the size
of the framework gravels (the larger gravels that make up the structure of the deposit) is of interest,
and the d50 or d84 of the study gravel (the sizes at which 50% or 84% of the sediments are finer)
should be compared with the spawning gravel sizes observed for the species elsewhere. To assess
whether the interstitial fine sediment content is so high as to interfere with incubation or emergence,
the percentage of fine sediment of the potential spawning gravel should be adjusted for probable
cleansing effects during redd construction, and then compared with rough standards drawn from
laboratory and field studies of incubation and emergence success. An assessment should also
consider that the fine sediment content of gravel can increase during incubation by infiltration,
that the gravels may become armored over time, or that downwelling and upwelling currents may
be inadequate. These considerations are incorporated in a nine-step, life-stage-specific assessment
approach proposed here.
 
Kondolf, G. M. (?). Learning from Stream Restoration Projects Davis, California, University of California, Davis: 107-110.
Kondolf, G. M., et al. (1991). "Distribution and Stability of Potential Salmonid Spawning Gravels in Steep Boulder-Bed Streams of the Eastern Sierra Nevada " Transactions of the American Fisheries Society 120: 177-186.
Kondolf, G. M., et al. (1989). "Spawning Gravels of Rainbow Trout in Glen and Grand Canyons, Arizona " Journal of Arizona-Nevada Academy of Science 23: 19-28.
Kondolf, G. M., et al. (1996). Status of Riparian Habitat Davis, California: University of California, Davis.
Kondolf, G. M. and W. V. G. Matthews (1991). "Unmeasured residuals in sediment budgets: A cautionary note." Water Resources Research 27(9): 2483-2486.
Kondolf, G. M., et al. (1993). "Modification of Fluvial Gravel Size by Spawning Salmonids " Water Resources Research 29(7): 2265-2274.
Kondolf, G. M. and M. G. Wolman (1993). "The sizes of salmonid spawning gravels." Water Resources Research 29(7): 2275-2285.
Korup, O. (2005). "Distribution of landslides in southwest New Zealand." Landslides 2: 43-51.
This study examines the size distribution of a regional
medium-scale inventory of 778 landslides in the mountainous
southwest of New Zealand. The spatial density of mapped landslides
per unit area can be expressed as a negative power–law
function of Landslide area AL spanning three orders of magnitude
( 10 2–101 km2). Although observed in other studies on landslide
inventories, this relationship is surprising, given the lack of absolute
ages, and thus uncertainty about the temporal observation
window encompassed by the data. Large slope failures (arbitrarily
defined here as having a total affected area AL>1 km2) constitute
83% of the total affected landslide area ALT. This dominance by
area affects slope morphology, where large-scale landsliding reduces
slope angles below the regional modal value of hillslopes,
jmod 39 . More numerous smaller and shallower failures tend to
be superimposed on the pre-existing relief. Empirical scaling relationships
show that large landslides involve >106 m3 of material.
The volumes VL of individual preserved and presumably prehistoric
(i.e. pre-1840) landslide deposits equate to 100–102 years of
total sediment production from shallow landsliding in the respective
catchments, and up to 103 years of contemporary regional
sediment yield from the mountain ranges. Their presence in an
erosional landscape indicates the geomorphic importance of
landslides as temporary local sediment storage.
 
Korup, O. (2006). "Effects of large deep-seated landslides on hillslope morphology, western Southern Alps, New Zealand." Journal of Geephysical Research 111.
Morphometric analysis and air photo interpretation highlight geomorphic imprints of
large landslides (i.e., affecting 1 km
2
) on hillslopes in the western Southern Alps (WSA),
New Zealand. Large landslides attain kilometer-scale runout, affect >50% of total basin
relief, and in 70% are slope clearing, and thus relief limiting. Landslide terrain shows
lower mean local relief, relief variability, slope angles, steepness, and concavity than
surrounding terrain. Measuring mean slope angle smoothes out local landslide
morphology, masking any relationship between large landslides and possible threshold
hillslopes. Large failures also occurred on low-gradient slopes, indicating persistent low-frequency/
high-magnitude hillslope adjustment independent of fluvial bedrock incision.
At the basin and hillslope scale, slope-area plots partly constrain the effects of landslides
on geomorphic process regimes. Landslide imprints gradually blend with relief
characteristics at orogen scale (10
2
km), while being sensitive to length scales of slope
failure, topography, sampling, and digital elevation model resolution. This limits means of
automated detection, and underlines the importance of local morphologic contrasts for
detecting large landslides in the WSA. Landslide controls on low-order drainage include
divide lowering and shifting, formation of headwater basins and hanging valleys, and
stream piracy. Volumes typically mobilized, yet still stored in numerous deposits despite
high denudation rates, are >10
7
m
3
, and theoretically equal to 10
2
years of basin-wide
debris production from historic shallow landslides; lack of absolute ages precludes further
estimates. Deposit size and mature forest cover indicate residence times of 10
1
–10
4
years.
On these timescales, large landslides require further attention in landscape evolution
models of tectonically active orogens.
 
Korup, O. (2006). "Rock-slope failure and the river long profile." Geology 34(1): 45-48.
This study examines the geomorphic effects of large (.106 m3) rock-slope failures on
long profiles of rivers in the Swiss Alps and the New Zealand Southern Alps. Regression
of channel slope versus drainage basin area objectively highlights knickpoints separating
incised from aggraded reaches that often correspond to locations of large rock-slope failures.
For a fixed concavity index, the highest values of the steepness index and erosion
index along a given profile spatially coincide with breach channels cut into formerly riverdamming
rockslide debris as old as 10 k.y. Assuming that the knickpoints do not predate
slope failure, data show that high profile steepness and inferred specific stream power are
not always the cause, but often a result, of large river-blocking rock-slope failure in mountain
basins. Omission of rockslide data from slope-area plots lowers the steepness index
and increases the concavity index on average, yet only in few cases more than one standard
deviation.
 
Korup, O., et al. (2006). "Fluvial response to large rock-slope failures: Examples from the Himalayas, the tien shan, and the Southern Alps in New Zealand." Geomorphology 78: 3-21.
We describe remnants of large (107–1010 m3) Late Pleistocene to Holocene rockslides and rock avalanches that block(ed) rivers
and are preserved in the Himalayas, the Tien Shan, and the New Zealand Southern Alps despite rates of uplift and erosion of up to 10
mm year−1. These natural dams control fluvial response on 101–104 year timescales by (a) storing and releasing sediment during
forced alluviation and fluvial re-incision; (b) relocating river channels through diversion or seepage; (c) inhibiting river erosion into
bedrock; (d) forming persistent long-profile knickpoints and knickslopes associated with steep high-energy (N103Wm−2) breach and
epigenetic bypass gorges and fluvial hanging valleys; and (e) shaping valley-floor morphology. Sediments indicate that rockslidedammed
lakes may persist up to 104 years, before being drained or infilled. Several short-lived (100–102 year) historical rockslide
dams in the Indian and Nepal Himalayas and the Southern Alps have had marked volumetric impacts on catchment sediment budgets
shortly following failure. Therefore, we caution against the linear extrapolation of sediment delivery from prehistoric rockslide dams
through time as a response variable. We find reach-scale changes to channel gradient to be prominent and persistent indicators of
fluvial response to large rock-slope failures.
 
Kostaschuk, R. A., et al. (1986). "Depositional process and alluvial fan-drainage basin morphometric relationships near Banff, Alberta, Canada." Earth Surface Processes and Landforms 11: 471-484.
Kothyari, U. C., et al. (2002). "Estimation of temporal variation of sediment yield using GIS." Hydrological Sciences Journal-Journal Des Sciences Hydrologiques 47(5): 693-706.
Kothyari, U. C., et al. (1994). "Prediction of sediment yield." Journal of Irrigation and Drainage Engineering [J. IRRIG. DRAIN. ENG.] 120(6): 1122-1131.
The prediction of sediment yield from catchments is necessary for studies of reservoir sedimentation, morphologic modeling, and soil-conservation planning. Though records on sediment yield are generally not available, methods that can be used to predict sediment yield from other commonly available hydrologic data, are available in the literature. In this study some of these methods are verified for their accuracy, using carefully collected data from experimental catchments. It is found that existing methods do not adequately account for the process of sediment delivery; hence these produce a less accurate prediction of sediment yield. A new method, based on the routing of surface erosion through time-area segments, is proposed. This method is found to produce the sediment yield more accurately than those it has been compared with. The method being proposed for the prediction of sediment yield has a more sound basis for further use in distributed models.
 
Kovalchik, B. L. and L. A. Chitwood (1990). "Use of geomorphology in the classification of riparian plant associations in mountainous landscapes of central Oregon, U.S.A." Forest Ecology and Management 33/34: 405-418.
Kovanen, D. J. and O. Slaymaker (2008). "The morphometric and stratigraphic framework for estimates of debris flow indicdnce in the North Cascades foothills, Wahsington State, USA." Geomorphology 99: 224-245.
Active debris flow fans in the North Cascade Foothills of Washington State constitute a natural hazard of importance to land managers, private
property owners and personal security. In the absence of measurements of the sediment fluxes involved in debris flow events, a morphologicalevolutionary
systems approach, emphasizing stratigraphy, dating, fan morphology and debris flow basin morphometry, was used. Using the
stratigraphic framework and 47 radiocarbon dates, frequency of occurrence and relative magnitudes of debris flow events have been estimated for
three spatial scales of debris flow systems: the within-fan site scale (84 observations); the fan meso-scale (six observations) and the lumped fan,
regional or macro-scale (one fan average and adjacent lake sediments). In order to characterize the morphometric framework, plots of basin area v.
fan area, basin area v. fan gradient and the Melton ruggedness number v. fan gradient for the 12 debris flow basins were compared with those
documented for semi-arid and paraglacial fans. Basin area to fan area ratios were generally consistent with the estimated level of debris flow
activity during the Holocene as reported below. Terrain analysis of three of the most active debris flow basins revealed the variety of modes of
slope failure and sediment production in the region.
Micro-scale debris flow event systems indicated a range of recurrence intervals for large debris flows from 106−3645 years. The spatial
variation of these rates across the fans was generally consistent with previously mapped hazard zones. At the fan meso-scale, the range of
recurrence intervals for large debris flows was 273−1566 years and at the regional scale, the estimated recurrence interval of large debris flows
was 874 years (with undetermined error bands) during the past 7290 years. Dated lake sediments from the adjacent Lake Whatcom gave
recurrence intervals for large sediment producing events ranging from 481−557 years over the past 3900 years and clearly discernible
sedimentation events in the lacustrine sediments had a recurrence interval of 67−78 years over that same period.
 
Kraft, C. E., et al. (2002). "Ice storm impacts on woody debris and debris dam formation in northeastern U.S. streams." Canadian Journal of Fisheries and Aquatic Sciences 59: 1677-1684.
Kraft, C. E. and D. R. Warren (2003). "Development of spatial pattern in large woody debris and debris dams in streams." Geomorphology 51: 127-139.
The spatial distribution of large woody debris (LWD) in streams was evaluated using Neighbor K statistics, following
extensive wood deposition from an ice storm in the eastern Adirondack Mountains (New York). Two years after wood
deposition, we surveyed individual pieces of LWD in one stream and surveyed debris dam locations in eight streams within the
ice storm area. To examine the linear pattern of debris dams within a stream, we used a one-dimensional version of Ripley’s K, a
second-order statistic that evaluates the spatial pattern of points within a landscape. Both aggregated and segregated (regularly
spaced) distributions of wood were identified. Individual pieces of LWD were aggregated at spatial extents ranging from 0 to 40
m and were segregated at spatial extents ranging from 80 to 100 m. In two streams, we found that debris dams were segregated
at distances ranging from 100 to 300 m relative to randomly chosen locations, but debris dams showed no significant spatial
pattern in six other study streams. Previous studies of wood distribution in streams have not observed segregated distribution
patterns. Spatial segregation of debris dams in the study area likely occurred in response to regularly spaced stream features or
processes that allow movement of individual pieces of LWD toward more stable accumulation points. Neighbor K statistics can
be used to identify and describe spatial pattern in large woody debris, and such patterns can be used to help evaluate and
identify processes responsible for their generation.
 
Krag, R. K., et al. (1986). A Forest Engineering Analysis of Landslides in Logged Areas on the Queen Charlotte Islands, British Columbia. Vancouver, British Columbia, Forest Engineering Research Institute of Canada.
Kramer, M. G., et al. (2001). "Abiotic controls on long-term windthrow disturbance and temperate rain forest dynamics in southeast Alaska." Ecology 82(10): 2749-2768.
Krammes, J. S. (1960). Erosion from mountain side slopes after fire in southern California. USDA For. Serv., Pac. Southwest For. and Range Exp. Station: 7.
Krammes, J. S. (1963). Seasonal debris movement from steep mountainside slopes in southern california. Sed. Conf., U.S. Department of Agriculture.
Kresch, D. L. (1998). Determination of upstream boundaries on western Washington streams and rivers under the requirements of the Shoreline Management Act of 1971, Water-Resources Investigations Report 96-4208. Tacoma, WA, U.S. Geological Survey.
Kreutzweiser, D. P., and S. S. Capell (2001). "Fine sediment deposition in streams after selective forest harvesting without riparian buffers." Canadian Journal of Forest Research 31(12): 2134-2142.
Krishna, J. H., et al. (1988). Predicting water and sediment yields from agricultural and grassland watersheds. ASAE Publication, vol.07-88: 324-329.
Krogstad, F. (1995). A Physiology and Ecology Based Model of Lateral Root Reinforcement of Unstable Hillslopes. Seattle, Washington, University of Washington: 44.
Krone, R. B. (1972). A field study of flocculation as a factor in estuarial shoaling processes, U. S. Army Corps of Engineers,: 62.
Kruse, C. G., et al. (1999). "Geomorphic Influences on the Distribution of Yellowstone Cutthroat Trout in the Absaroka Mountains, Wyoming." Transactions of the American Fisheries Society 126: 418-427.
Kufferberg, S. J. (1996). "Hydrologic and geomorphic factors affecting conservation of a river-breeding frog (Rana Boylii)! ." Ecological Applications 6(4): 1332-1344.
Kuhnle, R. A. (1989). "Bed-surface size changes in gravel-bed channel." Journal of Hydraulic Engineering(6): 731-743.
Kupferbert, S. J. (1996). "Hydrologic and geomorphic factors affecting conservation of a river-breeding frog (Rana boylii)." Ecol. Appl. 6: 1332-1344.
Kutzbach, J. E. and T. Webb, III (1993). Conceptual Basis for Understanding Late-Quaternary Climates Global Climates since the Last Glacial Maximum. H. E. Wright, J. E. Kutzbach, T. Webb, III et al., University of Minnesota Press. Chapter 2: 5-11.
L. E. Benda, N. L. P., C. Tague, M. A. Palmer, J. Pizutto, S. Cooper, E. Stanley, G.Moglen (2002). "How to avoid train wrecks when using science in environmental problem solving." BioScience 52(12): 1127-1136.
L., H. D., et al. (1995). Spatial and temporal evolution of small coastal gravel-bed streams: the influence of forest management on channel morphology and fish habitats. 4th International Workshop on Gravel-bed Rivers, Gold Bar, Washington.
La Marche, J. L., and D. P. Lettenmaier (2001). "Effects of forest roads on flood flows in the Deschutes River, Washington." Earth Surface Processes and Landforms 26: 115-134.
Lacey, S. T. (2000). "Runoff and sediment attenuation by undisturbed and lightly disturbed forest buffers." Water Air and Soil Pollution 122(1-2): 121-138.
A runoff plot experiment found that. ten metre undisturbed forest buffers removed 80-90% of runoff and over 95% of sediment produced by logging skid tracks. The study was carried out on 21 degrees slopes in a native forest in eastern New South Wales, Australia. The experiment included three replicates of four treatments including undisturbed control, skid track, skid track + undisturbed buffer and skid track + disturbed buffer. Skid track and control plots were 20 m long by 5 m wide. Buffer plots consisted of a 20 m by 5 m skid track directing runoff to a 10 m by 5 m naturally vegetated buffer that was either undisturbed or lightly disturbed. Runoff and sediment yields from plots were monitored over two successive summers. Undisturbed buffers greatly reduced overland flow and decreased sediment yields from around 100 Mg ha(-1) to less than 0.5 Mg ha(-1). Differences in both runoff and sediment yield between undisturbed buffer and control treatments were minimal and not statistically significant. Disturbed buffers achieved similarly large reductions in runoff and sediment yield in two out of three replicates. The third replicate yielded as much or more runoff and sediment than the skid track plots suggesting that disturbance increased the risk of buffer failure. The peak rate of outflow from buffer plots was generally not: related to peak buffer inflow until a threshold inflow of 1.6 L s(-1) was reached, after which peak outflow and peak inflow were linearly related.
 
Lackey, R. (1998). "Ecosystem Management: Desperately Seeking a Paradigm." Journal of Soil and Water Conservation 53(2): 92-94.
Lackey, R. T. (1994). Ecosystem Management:  Desparately Seeking a Paradigm. Corvallis, Oregon, EPA Environmental Research Laboratory: 30-46.
filed ( 2 copies)
 
Laflen J.M., D. C. F., and B.A. Engel. (2004). "Soil erosion and sediment yield prediction accuracy using WEPP." Journal of the American Water Resources Association 40(2): 289-297.
Laflen J.M., W. J. E., D.C. Flanagan, C.R.Meyer, and M.A. Nearing (1997). "WEPP-Predicting water erosion using a process-based model." Journal of Soil and Water Conservation m/j: 96-102.
Lague, D. (2010). "Reduction of long-term bedrock incision efficiency by short-term alluvial cover intermittency." Journal of Geophysical Research 115: 23pp.
Laituri, M. (200). "Marginal societies and geographic information systems." Varenius, UCSB.
Laituri, M. (2003). "The issue of access: an assessment guide for evaluating public participation geographic information science case studies." URISA Journal 15: 25-32.
Lajoie, K. R. (1986). Coastal Tectonics. Active Tectonics: Impact on Society, National Academies Press. Chapter 6: 95-124.
Lal, R. and B. Li (1980). Effects of forest clearing methods on sediment loss from agricultural catchments. Beijing, Guanghua Press.
Lamberti, G. A., et al. (1991). "Stream ecosystem recovery following a catastrophic debris flow." Canadian Journal of Fisheries and Aquatic Science 48: 196-208.
Recovery processes were studied for 3 yr in Quartz Creek (Cascade Mountains, Oregon), a third-order stream catastrophically affected by a February 1986 debris flow for which both predisturbance data and an upstream control reach were available. The debris flow altered channel geomorphology and destroyed riparian vegetation for 500 m, resulting in a reach with short, disordered channel units, low hydraulic retention, and an open canopy. High irradiance levels and reduced grazing by macroinvertebrates contributed to rapid accrual of benthic algae in the disturbed reach, which formed the bioenergetic basis for ecosystem recovery. Macroinvertebrates (mostly herbivores) recovered to upstream densities and taxonomic richness within 1 yr, although effects on community structure persisted into the second year. Cutthroat trout (Oncorhynchus clarki) populations were locally decimated by the disturbance, but by the following year, recruitment of young-of-the-year trout into the reach exceeded that of the upstream reach and populations had recovered to predisturbance densities. Despite the general rapid recovery of the biota within the disturbed reaches, most populations showed broad temporal fluctuations in abundance, suggesting that ecosystem stability was diminished by the debris flow.
 
Lancaster, S. T., S.K. Hayes, G.E. Grant, and Anonymous, (2002). "Modeling sediment and wood storage and dynamics in small mountainous watersheds." Abstracts with Programs - Geological Society of America 34(5): 89.
We examine controls on supply and transport of sediment and wood in a small (approximately two square kilometers) basin in the Oregon Coast Range, typical of streams at the interface between episodic sediment and wood delivery by mass movements and frequent fluvial sediment transport. We hypothesize that wood deposited by mass movements forms dams that lead to persistent sediment storage and inhibit coherent propagation of sediment pulses. Field data show that much sediment is stored behind such dams and in terraces after the dams breach. We developed a drainage basin-scale model driven by stochastic storm and fire sequences that combines empirical, stochastic and physical models of forest growth, tree fall, wood decay, soil production and diffusion, landslide initiation, debris flow runout, and fluvial sediment transport. In a 3000-year simulation of the study area, woody debris flow deposits form dams on the main channel and lead to steps in the channel profile and terraces on the valley floor that persist in place even after nearly all deposited wood has decayed. Simulated sediment output from the network is relatively steady and shows little evidence of episodic input. Our results suggest that abundant wood plays a key role in moderating sediment flux from small basins following debris flow events. Debris flow events coincident with a lack of abundant wood, such as might occur following forest harvest, could lead to more episodic sediment flux to downstream, fish-bearing reaches.
 
Lancaster, S. T. (2008). "Evolution of sediment accommodation space in steady state bedrock incising valleys subject to episodic aggradation." Journal of Geophysical Research 113: 1-17.
Lancaster, S. T. (2008). "Evolution of sediment accommodation space in steady state bedrock-incising valleys subject to episodic aggradation." Journal of  Geophysical Research 113: F04002.
Steepland valleys subject to debris flows incise bedrock even as episodic deposition
typically covers valley bottoms. This paper’s hypothesis is that, while continual fluvial
processes evacuate deposits, storage of episodic deposition drives valley widening and,
thereby, creation of accommodation space for sediment storage on the valley floor. Data
from three headwater valleys in the Oregon Coast Range show that valley-to-channel
width ratios and valley bottom deposit depths are variable, have little systematic variation
with respect to contributing area, and are similar on average among sites. A model of
valley cross-section evolution couples soil production, nonlinear diffusion, contrasting
rates of channel incision into deposits and bedrock, and stochastic valley bottom
deposition. The model reproduces observed flat, deposit-covered valley bottoms and
abrupt transitions to valley sides with oversteepened toe slopes. Simulations address
sensitivity of valley morphologies and incision rates to dimensionless numbers, the ratio
of instantaneous bedrock and deposit erosion rates (incision number), and the ratio of
deposition and evacuation rates (deposition number). For steady state simulations,
increasing deposition number by <101 leads to deposit depth and valley bottom width
increasing by 101 and 101.5, respectively, and valley bottom incision relative to the
instantaneous rate decreasing by 10 3. For incision number increasing by 103, valley
capacity (width times toe slope height) relative to mean deposit volume increases by 101.5.
Simulations, consistent with field data, imply that steady state valley widths are adjusted
to episodic deposition rates and respond more quickly to changes than profile gradients
because of contrasting limitations by instantaneous versus long-term lowering rates.
 
Lancaster, S. T. and N. E. Casebeer (2007). "Sediment s storage and evacuation in headwater valleys at the transition between debris-flow and fluvial processes." Geological Society of America 35(11): 1027-1030.
Lancaster, S. T. and N. E. Casebeer (2007). "Sediment storage and evacuation in headwater valleys at the transition between debris-flow and fluvial processes." Geology 35(11): 1027-1030.
Sediment from landscape disturbance often enters temporary storage in valleys and
evacuates over longer times, which in steeplands are poorly delimited. We hypothesize that,
across process transitions (e.g., debris fl ow versus fl uvial transport), distributions of sediment
transit times also change. We use fi eld surveys and extensive radiocarbon dating to assess
the distribution of transit (residence) times through the proxy measurement of ages of bank
deposits in two mainstem reaches of a 2.23 km2 watershed in the Oregon Coast Range. In the
downstream reach, debris fans impound fl uvial deposits; debris-fl ow, fi ne fl uvial, and coarse
fl uvial deposits compose nearly equal parts of the valley fi ll; and fl uvial erosion evacuates
deposits . Transit times have a sample mean of 1.22 × 103 14C yr and an exponential distribution,
indicating uniform probability of evacuation from storage. In the upstream reach, valleyspanning
debris jams impound debris-fl ow deposits composing >95% of the valley fi ll, which
is routinely scoured by debris fl ows. Transit times have a sample mean of 4.43 × 102 14C yr
and, if >100 14C yr, a power-law distribution, indicating preferential evacuation of younger
deposits and retention of older deposits. In both reaches, most sediment has short transit times
(<600 14C yr), but signifi cant volumes remain for millennia. Less than 20% of basin-wide
denudation passes through these reservoirs, but the latter are still signifi cant buffers between
hillslope disturbance and downstream aquatic habitat, especially for coarse sediment.
 
Lancaster, S. T. and G. E. Grant (2006). "Debris dams and the relief of headwater streams." Geomorphology: 14.
In forested, mountain landscapes where debris flows are common, their deposition commonly forms valley-spanning dams of
wood, boulders, or complex mixtures of both in headwater valleys. Sediment impoundment behind these dams causes alluviation in
what would otherwise be bedrock channels. In this paper, the effects of debris dams on the evolution of headwater valley profiles
over geologic time are examined. In the Oregon Coast Range, USA, longitudinal profiles of three headwater channels
(approximately 2 km 2 maximum contributing area), two of which are nominally bedrock, and all debris dams were surveyed.
Channel and valley widths were measured, and surface bed material measurements (pebble counts) were taken at several locations
along one stream. Cumulative relief in debris dams is highly variable within and among streams, reaching a maximum of 58% of
profile relief at one location. The surveyed dams comprise 9.8%, 19%, and 6.4% of total basin relief in the three corresponding
drainage basins, respectively. A model of bedrock erosion and channel profile evolution is derived that accounts for (a) fractional
coverage by the channel of total valley width at any given time and (b) shielding by sediment impounded behind debris dams. The
equation for incision rate, assumed constant, is solved for valley gradient as a function of contributing area, and parameter values
are estimated from or supplied by field data. Results are compared to stream-gradient敬汩灳楳contributing-area relationships obtained for
the field sites. These comparisons suggest a strong effect of network structure, which varies significantly among the sites, on profile
shape and relief because of the different susceptibilities to debris-flow deposition and, therefore, debris-dam formation.
Longitudinal profiles with and without the effects of debris dams are modeled for one forced-alluvial, i.e., nominally bedrock,
channel. For these model profiles, which extend only as far as the surveyed profile at this site, 55% of the profile relief is due to
debris dams and the concomitant shielding of the bed by impounded sediment. The results suggest that a significant fraction of the
relief of such forested, mountain landscapes is due to the effects of relatively immobile wood and boulders deposited in the valleys
by debris flows.
 
Lancaster, S. T., et al. (2000). Modeling sediment and wood storage and dynamics in small mountainous watersheds. Geomorphic Processes and Riverine Habitat. Washington, D. C., American Geophysical. 4: 85-102.
Lancaster, S. T., et al. (1999). "The interaction between trees and the landscape through debris flows [abs]." EOS, Transactions AGU 80(46), suppl.: F425.
Lancaster, S. T., et al. (2001). "Modeling Sediment and Wood Storage and Dynamics in Small Mountainous Watersheds." Geomorphic Processes and Riverine Habitat Water Science and Application: Water Science and Application 4: 85-102.
Lancaster, S. T., et al. (2001). Modeling sediment and wood storage and dynamics in small mountainous watersheds. Geomorphic Processes and Riverine Habitat. J. M. Dorava, D. R. Montgomery, B. B. Palcsak and F. A. Fitzpatrick. Washington, D.C., American Geophysical Union. 4: 85-102.
Lancaster, S. T., et al. (2003). "Effects of wood on debris flow runout in small mountain watersheds." Water Resources Research 39(6): doi:10.1029/2001WR001227.
Debris flows have typically been viewed as two-phase mixtures of sediment and
water, but in forested mountain landscapes, wood can represent a sizable fraction of total
flow volume. The effects of this third phase on flow behavior are poorly understood. To
evaluate whether wood can have a significant effect on debris flow runout in small
mountainous watersheds, we used a landscape-scale model combining empirical,
stochastic, and physical submodels of storms, fires, forest growth, tree fall, wood decay,
soil production and diffusion, landslide initiation, debris flow runout, and fluvial sediment
transport. We examined changes in the cumulative distribution function of debris flow
runout lengths in a small (2 km2) watershed in the Oregon Coast Range due to presence or
absence of two hypothesized effects of wood: (1) velocity reduction due to entrainment
of wood in the runout path and (2) velocity reduction due to changes in flow direction
angle. The model was calibrated such that the distribution for simulations including both
effects was similar to that measured in the study basin, and amounts of wood in the
simulation and the field, both fallen in small valleys and incorporated by debris flows,
were comparable. Removal of either effect, or both, significantly shifted runout length
distributions to longer lengths. Simulations and field observations indicate that with wood,
fluvial transport is a significant source of sediment output, few debris flows reach the
outlet, and debris flow deposits are widely distributed throughout the network.
Simulations indicate that without wood, basin sediment yield greatly increases, that yield
is dominated by longer-runout debris flows, and that debris flow deposits are concentrated
in the low-gradient reach near the outlet.
 
Lancaster, S. T., et al. (2003). Modeling sedmient and wood storage and dynamics in small mountainous watersheds, American Geophysical Unioni.
Lancaster, S. T., et al. (In Press). "Sediment reservoirs at mountain stream confluences: dynamics and effects of tributaries domionated by debris-flow and fluvial processes." Geological Society of America Bulletin.
Landres, P. B., et al. (1999). "Overview of the use of natural variability concepts in managing ecological systems." Ecological Applications 9(4): 1179-1188.
Lane, E. W. and W. M. Borland (1953). "River-Bed Scour During Floods." Transactions of the American Society of Civil Engineers 119: 1069-1089.
Lane, J. W. (1987). Relations between geology and mass movement features in a part of the East Fork Coquille River watershed, southern coast range, Oregon.
Lane, L. J. (1982). Development of a procedure to estimate runoff and sediment transport in ephemeral streams. Recent Developments in the Explanation and Prediction Erosion and Sediment Yield (Proceedings of the Exeter Symposium, July 1982) IAHS Publicaion No. 137.
Lane, L. J., T.E.Hakonson, and G.R.Foster (1986). Watershed erosion and sediment yield affecting contaminant transport. Proceedings of the USDOE Symposiom for environmental research on actinide elements, hilton Head, SC.
Lane, L. J., and Shirley E.D. (1988). Modelling erosion on hillslopes. Modelling geomorphological systems. M. G. Anderson, John Wiley and sons: 287-308.
Lane, L. J., M. Hernandez,  and M. Nichols (1997). "Processes controlling sediment yield from watersheds as functions of spatial scale." Environmental Modelling and Software 12(4): 355-369.
Lane, P. N., J. C. Croke, and P. Dignan (2004). "Runoff generation from logged and burnt convergent hillslopes: rainfall simulation and modelling." Hydrological Processes 18: 879-892.
Lane, P. N. J., and G.J. Sheridan (2002). "Impact of an unsealed forest road stream crossing: water quality and sediment sources." Hydrological Processes 16(13): 2599-2612.
Turbidity monitoring and rainfall and runoff simulation experiments were conducted at a newly constructed unsealed road stream crossing to determine the quantity and sources of sediment entering the stream. Continuous measurements of turbidity and estimation of total suspended solids (TSS) concentration upstream and downstream of the stream culvert were taken over a 5 month period. There was a statistically significant difference in turbidity and TSS downstream of the crossing during baseflow conditions, but the quality of the water column remained good during non-rain periods. Rainfall events comprised around 20% of the observation period and led to decreases in water quality downstream of the crossing. Water quality could be considered as degraded for 10% of the observations. This was during a period when the rainfall was 65% of the long-term average. Calculated suspended sediment loads were 0-78 t upstream and 2.77 t downstream, an increase of 3.5. It was estimated that at least 2-3 t of bedload material was also added to the stream during the crossing construction and from subsequent erosion. This material is a deposit on the cobble stream bed, and is most likely to degrade aquatic ecosystem values. Rainfall and runoff simulation revealed the principal sediment sources to be a fillslope that contributed coarse bedload material through rill erosion and unprotected toe scour, and the unmetalled road verge that provided fines. Although the quality of water column was good for the majority of the observations, the new Australian and New Zealand Water Quality Guidelines for Fresh and Marine Waters suggest this site exceeded 'trigger levels' that would warrant further investigation for both the water column and the bed deposits. Copyright (C) 2002 John Wiley Sons, Ltd.
 
Lane, S. N. and K. S. Richards (1997). "Linking river channel form and process: time, space and causality revisited." Earth Surface Processes and Landforms 22: 249-260.
Fluvial geomorphology has witnessed a continuing reduction in the time- and space-scales of research, with increasing emphasis on the dynamics of small site-specific river reaches. This shift can be regarded as part of a trend towards the understanding and explanation rather than description of how rivers change, which raises important questions regarding the relevance of such short time-scale and small space-scale research to understanding longer-term aspects of landform
behaviour. The methodological challenges that arise from such intensive case study research are illustrated here using a detailed investigation of a river reach. Morphological changes within this reach are shown to be driven by: (i) catchment-scale processes associated with the interaction of discharge and sediment supply waves; and (ii) modification of these processes through morphological controls on erosion and deposition patterns and hence net channel change. The ‘morphological conditioning’ of channel response reflects the configurational aspects of channel change, and the
importance of local characteristics in the understanding of system behaviour. Sensitivity to local conditions implies that short time-scale and small space-scale processes may be critical to channel behaviour, particularly if the system is interpreted in non-linear terms. Although it may be possible to identify statistically averaged stable states, non-linear system behaviour implies that system trajectories are sensitively dependent upon instantaneous system states. Thus, changes between average states can only be understood through an understanding of the sequence of configurational states through which the system evolves.
 
Lang, A., et al. (1999). "Classic and new dating methods for assessing the temporal occerrence of mass movements." Geomorphology 30: 33-52.
Langbein, W. B. and L. B. Leopold (1968). River meanders - theory of minimum variance, U. S. Geological Survey.
Lanka, R. P., et al. (1987). "Relations of Geomorphology to Stream Habitat and Trout Standing Stock in Small Rocky Mountain Streams " Transactions of the American Fisheries Society 116: 21-28.
Lapointe, M., et al. (2000). "Modelling the probability of salmonid egg pocket scour due to floods." Canadian Journal of Fisheries and Aquatic Science 57: 1120-1130.
Flood disturbance plays a key but complex role in structuring lotic ecosystems. Empirical models proposed
here allow salmonid resource managers to quantify the probability of egg pocket scour during floods and to predict
how the expected losses vary with flood strength and reach characteristics. The models are based on comparisons between
published salmonid egg pocket depth criteria and statistics on the intensity and spatial distribution of scour and
fill produced by three flood events of widely different magnitudes in three separate reaches of a gravel–cobble Atlantic
salmon (Salmo salar) river in the Saguenay region, Quebec. A simple substrate mobility index, based on reach-scale
geomorphic characteristics and flood hydraulics, was shown to provide useful predictions (R2 up to 74%) of the fraction
of the area of potential spawning zones undergoing flood scour greater than 30 cm. Any Atlantic salmon egg
pockets present in these deeply scoured areas would be destroyed. The models also predict the distribution of fill (net
rise in bed) potentially causing fry entombment at redds. The flood disturbance data suggest that average probability of
scour of an Atlantic salmon egg pocket in the study reaches ranges from under 5% for frequent-recurrence spring
floods to approximately 20% for an extreme, multicentenary-recurrence flood.
 
Laronne, J. B. and M. A. Carson (1976). "Interrelationships between bed morphology and bed-material transport for a small, gravel-bed channel." Sedimentology 23: 67-85.
Laronne, J. B., et al. (1994). "Scour chain employment in gravel bed rivers." Catena 22: 299-306.
The dynamics of river beds can be assessed from scour and fill data obtained from scour chains. Chains have not been used extensively in gravel bed rivers due to a variety of empolyment difficulties. Chain insertion, anchoring and relocation can be determined with minimal manpower by utilising a set of tested employment methodologies. These are described in sufficient detail to allow fluvial geomorphologists, ecologists and river engineers to more commonly utilise chains in studies of river stability and river response
 
Laronne, J. B., et al. (1995). "Selection of Gravel-Transport Formula for Stream Modeling." Journal of Hydraulic Engineering: 567-569.
Larsen, I. J., and L.H. MacDonald (2007). Does ash contribute to post-fire soil sealing and increased runoff rates? 27th Annual AGU Hydrology Days, Colorado State University, Fort Collins, CO.
Larsen, I. J., and L.H. MacDonald (2007). "Predicting postfire sediment yields at the hillslope scale: testing RUSLE and Disturbed WEPP." Water Resources Research 43(w11412, doi:10.1029/2006WR005560): 18p.
Larsen, I. J. and L. H. MacDonald (2007). "Predicting postfire sediment yields at the hillslope scale: Testing RUSLE and Distubed WEPP." Water Resources Research 43(W11412).
High-severity wildfires can increase hillslope-scale sediment yields by several orders
of magnitude. Accurate predictions of postfire sediment yields are needed to guide
management decisions and assess the potential impact of soil loss on site productivity and
downstream aquatic resources. The Revised Universal Soil Loss Equation (RUSLE)
and Disturbed WEPP are the most commonly used models to predict postfire sediment
yields at the hillslope scale, but neither model has been extensively tested against field
data. The objectives of this paper are to (1) compare predicted sediment yields from
RUSLE and Disturbed WEPP against 252 plot years of data from nine fires in the
Colorado Front Range; and (2) suggest how each model might be improved. Predicted and
measured sediment yields were poorly correlated for RUSLE (R2 = 0.16) and only slightly
better correlated for Disturbed WEPP (R2 = 0.25). Both models tended to over-predict
sediment yields when the measured values were less than 1 Mg ha 1 yr 1 and to
under-predict higher sediment yields. Model accuracy was not improved by increasing the
soil erodibility (K) factor in RUSLE and was only slightly improved by slowing the
vegetative recovery sequence in Disturbed WEPP. Both models much more accurately
predicted the mean sediment yields for hillslopes grouped by fire and severity (R2 = 0.54
to 0.66) than for individual plots. The performance of RUSLE could be improved by
incorporating an erosivity threshold and a nonlinear relationship between rainfall erosivity
and sediment yields. The performance of WEPP could be improved by reducing the
effective hydraulic conductivity in sites that have recently burned at high severity. The
results suggest that neither model can fully capture the complexity of the different
controlling factors and the resultant plot-scale variability in sediment yields.
 
Larsen, M. C. (2008). "Rainfall-triggered landslides, anthropogenic hazards, and mitigation strategies."
Larsen, M. C. and A. Simon (1993). "A rainfall intensity-duration threshold for landslides in a humid-tropical environment, Puerto Rico." Geografiska Annaler 75: 13-23.
Larsen, M. C., et al. (1999). "Slopewash, surface runoff and fine-litter transport in forest and landslide scars in humid-tropical steeplands, Luquillo Experimental Forest, Puerto Rico." Earth Surface Processes and Landforms 24(6): 481-502.
Larson, K. R. and R. C. Sidle (1980). Erosion and Sedimentation data catalog of the Pacific Northwest. Portland, Oregon, U.S. Forest Service.
Lascelles, B., et al. (2000). "Spatial and temporal variation in two rainfall simulators: Implications for spatially explicit rainfall simulation experiments." Earth Surface Processes and Landforms 25(7): 709-721.
Rainfall simulators are widely used yet there is little evidence in the literature to show that their spatial and temporal variability has been adequately taken into account. For experiments that are concerned only with some aggregate or mean effect of simulated rain then such variations may be unimportant. However, where rainfall simulation is being used to study (and perhaps model) small-scale processes that are themselves spatially variable (such as rill initiation) then knowledge of the simulator's inherent variability is vital. A first aim of this paper is therefore to examine this variability, and to appraise methodologies by which it may be quantified. A second aim is to evaluate the implications for spatially explicit rainfall simulation experiments.Two simulators were used, a portable drip-screen simulator and a laboratory-based full-cone nozzle simulator. Neither produced a spatially uniform distribution of rainfall depth: both produced distributional patterns that were fairly consistent despite varying intensities and run times. Small-scale, apparently random variations were superimposed on these more deterministic patterns. However, despite this marked spatial variability, calculation of uniformity coefficients (1-SD/mean) resulted in high values. Thus it appears that the uniformity coefficient gives little real indication of the spatial uniformity of simulated rainfall, despite its established usage in the literature. Additionally, spatial distributions of raindrop size -and hence kinetic energy -were calculated for the full-cone nozzle simulator. These show that zones of high rainfall amount do not necessarily relate to zones of high energy reaching the surface.The presence of such variability raises a number of issues for spatially explicit rainfall simulation experiments. While there has been little work on the spatial variability of natural rainfall at field scale and smaller, it appears that the spatial heterogeneity of simulated rainfall depths observed in this study does not differ greatly from that of natural rain. But since a major attraction of rainfall simulation experiments is additional control over rainfall's many variables, the spatial non-uniformity of depth observed in this Study is unwelcome. The existence of an apparently deterministic component lo this non-uniformity nonetheless suggests that it can, at least in principle, be corrected by calibration. Less easily handled is the discrepancy between spatial distributions of rainfall depth and energy, since this will certainly affect rainfall simulation experiments that are, for example, concerned with erosion processes due to raindrop impact. Copyright (C) 2000 John Wiley & Sons, Ltd.
 
Lashermes, B. and E. Foufoula-Georgiou (2007). "Area and width functions of river networks: new results on multifractal properties." Water Resources Research 43: W09405.
This paper investigates the multiscale statistical structure of the area and width
functions of simulated and real river networks via state-of-the-art wavelet-based
multifractal (MF) formalisms. First, several intricacies in performing MF analysis of these
signals are discussed, and a robust framework for accurate estimation of the MF
spectrum is presented. Second, it considers the following three questions: (1) Does the
topology of river networks leave a unique signature on the MF spectrum of area and width
functions? (2) How different are the MF properties of commonly used simulated trees
and those of real river networks? and (3) Are there differences between the MF properties
of width and area functions, and what can these tell us about the topology of hillslope
versus channelized drainage patterns in a river basin? The results indicate discrepancies
between the statistical scaling of the area functions of real networks (found to be
multifractal with a considerable spread of local singularities and the most prevailing
singularity ranging from 0.4 to 0.8) and that of several commonly used stochastic selfsimilar
networks (found to be monofractal with a single singularity exponent H in the
range of 0.5–0.65). Moreover, differences are found between the MF properties of width
and area functions of the same basin. These differences may be the result of distinctly
different branching topologies in the hillslope versus channelized drainage paths and need
to be further investigated.
 
Lashermes, B., et al. (2007). "Channel network extraction from high resolution topography using wavelets." Geophysical Research Letters 34(L23S04): 6.
The availability of high resolution topography from
LIDAR offers new opportunities for objectively extracting
the channels directly from a DEM using local topographic
information, instead of inferring them indirectly based on
global criteria, such as area or area-slope threshold
relationships. Here we introduce the use of wavelet filtering
to delineate threshold curvatures for defining valleys and
threshold slope-direction-change for defining probable
channeled portions of the valleys. This approach exploits
the topographic signatures uniquely found in high resolution
topography, and reveals the fuzzy topographic transition in
which local weakly convergent areas lie at the transition
between hillslopes and valleys. Citation: Lashermes, B.,
E. Foufoula-Georgiou, and W. E. Dietrich (2007), Channel network
extraction from high resolution topography using wavelets
 
Latham, D. J. and J. A. Schlieter (1989). Ignition Probabilities of Wildland Fuels Based on Simulated Lightning Discharges. Ogden, UT, Intermountain Research Station: 16.
Lathrop, G. R., Jr. and D. L. Peterson (1992). "Identifying structural self-similarity in mountainous landscapes." Landscape Ecology 6(4): 233-238.
Latterell, J. J. and R. J. Naiman (2007). "Sources and dynamics of large logs in a temperate floodplain river." Ecological Applications 17(4): 1127-1141.
Large logs, important agents of biophysical heterogeneity in temperate
floodplain rivers, have been virtually eliminated from modified systems. Our purpose was to
quantify the sources and dynamics of large logs ( 1 m diameter) in the mainstem of a nearly
pristine system: the Queets River, Washington, USA. Erosion of forests by the river supplies
0.40 logs (100 m) 1 yr 1 to the channel. Most (72%) are new logs entering the river for the first
time as the river undercuts mature fluvial terraces dominated by large conifers. Retrospective
airphoto analyses demonstrate that, over 63 years, the Queets River recruits 95% of new logs
from a riparian corridor extending 265 m laterally on both banks, mostly through channel
meandering. However, input rates are patchy, with 10% of the valley length supplying 38% of
the new logs. As the river moves laterally, remnant logs are left on channel surfaces that later
develop riparian forests and reenter the river when those forests erode. Remnant logs lying on
the floodplain forest floor surface or buried in alluvium constitute 21% and 7% of the annual
inputs from bank erosion, respectively. We estimate that 50% of logs deposited in the channel
in a given year, including those underpinning logjams, are transported downriver within five
years. Over the next 55 years, bank erosion reclaims an additional 23%, leaving 27% of the logs
stable for .60 years. Simulations indicate that recurrent transport is common, with half of the
large conifers being deposited in  3 locations and transported  1.5 km prior to disintegrating.
One in ten logs links distant reaches by occupying  7 locations spanning  12.0 km. Instream
supplies are therefore a mixture of new and old logs from nearby and upstream forests,
sustained by the recapture and transport of stockpiled remnant logs during periods when new
inputs are low. We propose that patchy input rates and the periodic rearrangement of large
logs are important drivers of temporal variation in river valley habitats, adding to the spatial
complexity created by stable logs. These findings underscore the importance of extensive
mature forests and connectivity in temperate floodplain rivers.
 
Lawson, D. E. (1982). "Mobilization, movement and deposition of active subaerial sediment flows, Matanuska glacier, Alaska." Journal of Geology 90: 279-300.
Lawson, P. W. (1993). "Cycles in Ocean Productivity, Trends in Habitat Quality, and the Restoration of Salmon Runs in Oregon." Fisheries 18(8).
Lawson, P. W., et al. (2004). Identification of Historical populations of Coho Salmon (Onchorhynchus kisutch) in the Oregon Coast Evolutionarily Significant Unit. Review Draft. , Oregon Northern California Coast Technical Recovery Team. NOAA/NMFS/NWFSC.: 129.
Lawson, P. W., et al. (2004). "Environmental factors influencing freshwater survival and smolt production in Pacific Northwest coho salmon (Oncorhynchus kisutch)." Canadian Journal of Fisheries and Aquatic Science 61: 360-373.
Climate variability is well known to affect the marine survival of coho salmon (Oncorhynchus kisutch) in
Oregon and Washington. Marine factors have been used to explain up to 83% of the variability in Oregon coastal
natural coho salmon recruitment, yet about half the variability in coho salmon recruitment comes from the freshwater
life phase of the life cycle. This seeming paradox could be resolved if freshwater variability were linked to climate and
climate factors influencing marine survival were correlated with those affecting freshwater survival. Effects of climate
on broad-scale fluctuations in freshwater survival or production are not well known. We examined the influence of
seasonal stream flows and air temperature on freshwater survival and production of two stock units: Oregon coastal
natural coho salmon and Queets River coho salmon from the Washington Coast. Annual air temperatures and second
winter flows correlated strongly with smolt production from both stock units. Additional correlates for the Oregon
Coast stocks were the date of first fall freshets and flow during smolt outmigration. Air temperature is correlated with
sea surface temperature and timing of the spring transition so that good freshwater conditions are typically associated
with good marine conditions.
 
LeBoutillier, D. W. and P. R. Waylen (1993). "A stochastic model of flow duration curves." Water Resources Research 29(10): 3535-3541.
Lecce, S. A. (1990). The Alluvial Fan Problem. Alluvial Fans: A Field Approach. A. H. Rachocki and M. Church, John Wiley and Sons Ltd. Chapter 1: 3-23.
Lee, H.-Y., et al. (1997). "Quasi-two-dimesional simulation of scour and deposition in alluvial channels " Journal of Hydraulic Engineering 123(7): 600-609.
Lee, K. N. and J. Lawrence (1986). "Adaptive management: learning from the Columbia River Basin Fish and Wildlife Program." Environmental Law 16: 431-460.
Lefsky, M. A., W.B. Cohen, G.G. Parker, and D.J. Harding (2002). "Lidar remote sensing for ecosystem studies." BioScience 52(1): 19-30.
Lefsky, M. A., et al. (1999). "Lidar remote sensing of the canopy structure and biophysical properties of douglas-fir western hemlock forests." Remote Sensing and the Environment 70: 339-361.
Lefsky, M. A., et al. (1999). "Surface lidar remote sensing of basal area and biomass in deciduous forests of eastern Maryland, USA." Remote Sensing and the Environment 67: 83-98.
Legates, D. R., and G.J. McCabe Jr. (1998). "Evaluating the use of "goodness-of-fit" measures in hydrologic and hydroclimatic model validation." Water Resources Research 35(1): 233-241.
Legleiter, C. J. and P. C. Kyriakidis (2006). "Forward and inverse transformations between cartesian and channel-fitted coordinate systems for meandering rivers." Mathematical Geology 38(8): 927-958.
The spatial referencing of river channels is complicated by their meandering planform, which dictates
that Euclidean distance in a Cartesian reference frame is not an appropriate metric. Channel-fitted
coordinate systems are thus widely used in application-oriented geostatistics as well as theoretical
fluid mechanics, where flow patterns are described in terms of a streamwise axis s along the channel
centerline and an axis n normal to that centerline. A means of transforming geographic (x, y) coordinates
to their equivalents in the (s, n) space and vice versa is needed to relate the two frames of
reference, and this paper describes a pair of transformation algorithms that are explicitly intended
for reach-scale studies of modern rivers. The forward transformation from Cartesian to channel-fitted
coordinates involves parametric description of the centerline using cubic splines, calculation of centerline
normal vectors and curvature using results from differential geometry, and an efficient local
search to find in-channel data points and compute their (s, n) coordinates. The inverse transformation
finds the nearest vertices of a discretized centerline and uses a finite difference approximation to the
streamwise rates of change of the centerline’s Cartesian coordinates to obtain the geographic equivalent
of a point in the (s, n) space. The performance of these algorithms is evaluated using: (i) field
data from a gravel-bed river to examine the effects of initial centerline digitization and subsequent
filtering; and (ii) analytically-defined centerlines and simulated coordinates to assess transformation
accuracy and sensitivity to centerline curvature and discretization. Any discrepancy between a point’s
known coordinates in one frame of reference and the coordinates produced via transformation from
the other coordinate system constitutes a transformation error, and our results indicate that these
errors are 2–4% and 0.2–0.5% of the channel width for the field case and simulated centerlines,
respectively. The primary sources of transformation error are the initial digitization of the centerline
and the relationship between centerline curvature and discretization.
 
Lehre, A. K. (1982). Sediment Budget of a Small Coast Range Drainage Basin in North-Central California Workshop on Sediment Budgets and Routing in Forested Mountain Drainage Basins, U.S.D.A. Forest Service.
Lehre, A. K. and G. Carver (?). Thrust Faulting and Earthflows: Speculations on the Sediment Budget of a Tectonically Active Drainage Basin Humboldt State University: 169-184.
Leman, V. N. (1993). "Spawning Sites of Chum Salmon, Oncorhynchus keta: Microhydrological Regime and Viability of Progeny in Redds (Kamchatka River Basin) " Journal of Ichthyology 33(2): 104-117.
Leopold, E. B., et al. (1982). "Pollen and lignin records of the late Quaternary vegetation, Lake Washington." Science 218: 1305-1307.
Leopold, E. B. and G. M. Wolman (1960). "River meanders." Geological Society of America Bulletin 71: 769-794.
Leopold, L. B., and Thomas Maddock, Jr. (1953). The hydraulic geometry of stream channels and some physiographic implications. Geological Survey Professional Paper 252, Geological Survey professional paper 252. U.S. Govt. Print. Office, Washington, D.C.: 53.
Leopold, L. B. (1970). "An improved method for size distribution of stream bed gravel." Water Resources Research 6(5): 1357-1366.
Leopold, L. B. and W. B. Bull (1979). "Base level, Aggradation, and Grade." Proc. Amer. Phil. Soc. 123(3): 168-202.
Leopold, L. B. and W. W. Emmett (1976). "Bedload measurements, East Fork River, Wyoming." Proc. Natl. Acad. Sci. USA 73(4): 1000-1004.
Leopold, L. B. and T. J. Maddock (1953). The hydraulic geometry of stream channels and some physiographic implications: USGS Professional Paper 252, United States Geological Survey.
Leopold, L. B. and T. J. Maddock (1953). The hydraulic geometry of stream channels and some physiographic implications: USGS Professional Paper 252, United States Geological Survey: 57.
Leopold, L. B. and D. L. Rosgen (1991). Selected readings from dynamic stream classification and restoration procedures (DRAFT).
Leopold, L. B. and C. Vita-Finzi (1998). "Valley Changes in the Mediterranean and America and Their Effects on Humans " Proceedings of the American Philosophical Society 1: 1-17.
Leopold, L. B., et al. (1964). Fluvial Processes in Geomorphology. San Francisco, H. W. Freeman.
Lepp, L. R., et al. (1993). "Channel Erosion in Steep Gradient, Gravel-Paved Streams." Bulletin of the Association of Engineering Geologists XXX(4): 443-454.
Lestelle, L. and M. L. Rowse (1995). Concepts and Tools for Planning Natural Coho Salmon Enhancement Measures: A Planning Guide (DRAFT). Vashton, Washington, Mobrand Biometrics Inc.: 1-40.
Lewin, J. and D. Hughes (1980). "Welsh Floodplain Studies: An Application of a Qualitative Inundation Model." Journal of Hydrology 46: 35-49.
Lewin, R. (1986). "In Ecology, Change Brings Stability." Science 234: 1071-1073.
Lewis, J., and R.R. Ziemer (1998). Evaluating the impacts of logging activities on erosion and suspended sediment transport in the Caspar Creek watersheds. General Technical Report Report: PSW-GTR-168, USDA Forest Service, Pacific South West Research Station, CA.: 55-69.
Suspended sediment has been sampled at both the North and South Fork weirs of Caspar Creek in northwestern California since 1963, and at 13 tributary locations in the North Fork since 1986. The North Fork gaging station (NFC) was used as a control to evaluate the effects of logging in the South Fork, in the 1970's, on annual sediment loads. In the most conservative treatment of the data, suspended loads increased by 212 percent over the total predicted for a 6-yr period commencing with the onset of logging. When the roles of the watersheds were reversed and the same analysis repeated to evaluate harvesting in the North Fork under California Forest Practice Rules in the 1990's, no significant increase was found at NFC in either annual suspended or bed load. With the advent of automatic pumping samplers, we were able to sample sediment concentration much more frequently in the 1980's. This allowed storm event loads from control watersheds in the North Fork to be used in a new regression analysis for NFC. According to this more sensitive analysis, for the 7-yr period commencing with the onset of logging, the sum of the suspended storm loads at NFC was 89 percent higher than that predicted for the undisturbed condition. The much greater increase after logging in the South Fork is too great to be explained by differences in sampling methods and in water years, and appears to be the result of differences in road alignment, yarding methods, and stream protection zones. Similar analyses of storm event loads for each of the treated subwatersheds in the North Fork suggested increased suspended loads in all but one of the tributaries, but effects were relatively small or absent at the main stem locations. Of watersheds with less than 50 percent cut, only one showed a highly significant increase. The greater increase in sediment at NFC, compared to other main-stem stations, is largely explained by a 3,600-m (super 3) landslide that occurred in 1995 in a subwatershed that drains into the main stem just above NFC. Differences among tributary responses can be explained in terms of channel conditions. Analysis of an aggregated model simultaneously fit to all of the data shows that sediment load increases are correlated with flow increases after logging. Field evidence suggests that the increased flows, accompanied by soil disruption and intense burning, accelerated erosion of unbuffered stream banks and channel headward expansion. Windthrow along buffered streams also appears to be important as a source of both woody debris and sediment. All roads in the North Fork are located on upper slopes and do not appear to be a significant source of sediment reaching the channels. The aggregated model permitted evaluation of certain types of cumulative effects. Effects of multiple disturbances on suspended loads were approximately additive and, with one exception, downstream changes were no greater than would have been expected from the proportion of area disturbed. A tendency for main-stem channels to yield higher unit-area suspended loads was also detected, but after logging this was no longer the case in the North Fork of Caspar Creek.
 
Lewis, J. (2002). "Quantifying recent erosion and sediment delivery using probability sampling: A case study." Earth Surface Processes and Landforms 27(5): 559-572.
Lewis, J. (?). Evaluating the Impacts of Logging Activites on Erosion and Suspended Sediment Transport in the Caspar Creek Watersheds (DRAFT)
Lewis, N. K. (1997). "use of the discharge-weighted avarage velocity in studies of the frictional energy loss of streamflow." Earth Surface Processes and Landforms 22: 329-336.
Lewis, N. K. (1997). "Use of the discharge-weighted average velocity in studies of the frictional energy loss of streamflow." Earth Surface Processes and Landforms 22: 329-336.
Lewis, S. and A. Nir (1978). "On tracer theory in geophysical systems in the steady and non-steady state: Part II. Non-steady state-theoretical introduction." Tellus 30(3): 260-271.
Lewis, T. D. e. a. (2000). Executive Summary: Regional Assessment of Stream Temperature Across Northern California and Their Relationship to Various Landscape Level and Site Specific Attributes. Forest Science Project. Arcata, CA, Humboldt State University Foundation: 14.p.
Li, Z. and P. D. Komar (1986). "Laboratory measurements of pivoting angles for applications to selective entrainment of gravel in a current." Sedimentology 33: 413-423.
Libohova, Z. (2004). Effects of thinning and a wildfire on sediment production rates, channel morphology, and water quality in the Upper South Platte River Watershed. Fort Collins, Colorado, M.S. Thesis, Colorado State University: 103.
Licznar, P. a. M. A. N. (2003). "Artificial neural networks of soil erosion and runoff prediction at the plot scale." Catena 51(2003): 89-114.
Liebault, F. and H. Piegay (2001). "Assessment of channel changes due to long-term bedload supply decrease, Roubion River, France." Geomorphology 36(3-4): 167-186.
Lienkaemper, G. W. and F. J. Swanson (1987). "Dynamics of large woody debris in streams in old growth Douglas-fir forests." Canadian Journal of Forest Research 17: 150-156.
Lienkaemper, G. W. and F. J. Swanson (1987). "Dynamics of large woody debris in streams in old-growth Douglas-fir forests." Canadian Journal of Forest Research 17(2): 150-156.
Transfer of large woody debris (> 10 cm diameter) from old-growth Douglas-fir (Pseudotsuga menziesii (Mirbel) Franco) forests into five first-to fifth-order stream reaches (drainage areas of 0.1 to 60.5 km super(2)) has ranged from 2.0 to 8.8 mg multiplied by ha super(-1) multiplied by year super(-1) in 7- to 9-year study periods. Amounts of large debris in these streams range from 230 to 750 mg multiplied by ha super(-1), with generally lower values in larger channels. The addition of woody debris is widely scattered in time and space and comes mainly from single trees rooted away from the streambank. The authors inter that wind is a major agent for entry of wood into these streams. Downstream movement of debris is strongly related to length of individual pieces; most pieces that moved were shorter than bankfull width.
 
Ligon, F., et al. (1995). "Downstream Ecological Effects of Dams " BioScience 45(3): 183-192.
Ligon, F., et al. (1999). Report of the Scientific Review Panel on California Forest Practice Rules and Salmonid Habitat Sacramento, California, prepared for The Resources Agency of California and the National Marine Fisheries Service.
Ligon, F. K., et al. (1995). "Downstream ecological effects of dams." BioScience 45(3): 183-192.
Likens, G. E. and R. E. Bilby (1988). Development, maintenance, and role of organic-debris dams in New England streams. Worskhop on Sediment Budgets and Routing in Forested Drainage Basins. F. J. Swanson, R. J. Janda, T. Dunne and D. N. Swanston. Portland, OR, U. S. D. A., Forest Service. PNW-141: 122-128.
Lindsay, J. B. (2003). "A physically based model for calculating contributing area on hillslopes and along valley bottoms." Water Resources Research 39(12): 1332.
Most existing methods of calculating contributing area are unable to accurately model the pattern of contributing area on hillslopes and along valley bottoms. This paper describes a new flow algorithm, the adjustable dispersion routing algorithm (ADRA). Rather than calculating contributing area using predetermined flow characteristics that are insensitive to location in the landscape, ADRA predicts channel location and adjusts flow characteristics accordingly. ADRA increases the degree of flow divergence downslope from divides until a channel head is detected. Channel head locations are estimated on the basis of a user-defined threshold of an area-slope function. Therefore the algorithm overcomes the problems of aggregated flow on hillslopes and divergent flow along valley bottoms. The pattern of catchment area produced by ADRA was compared with similar patterns calculated using a similar flow algorithm for a variety of natural landscapes. ADRA produced patterns of contributing area that were more consistent with the theory of channel initiation.
 
Lindsay, J. B. and I. F. Creed (2005). "Removal of artifact depressions from digital elevation models: towards a minimum impact approach." Hydrological Processes 19: 3113-3126.
Artifact depressions in digital elevation models (DEMs) interrupt flow paths and alter drainage directions. Techniques
for removing depressions should enforce continuous flow paths in a way that requires the least modification of the DEM.
Impacts on the spatial and statistical distributions of elevation and its derivatives were assessed for four methods of
removing depressions: (1) filling; (2) breaching; (3) a combination of filling and breaching, with breaching constrained
to a maximum of two grid cells; (4) a combination of filling and breaching based on an impact reduction approach
(IRA). The IRA removes each depression using either filling or breaching, depending on which method has the least
impact, in terms of the number of modified cells and the mean absolute difference in the DEM.
Analysis of a LiDAR DEM of a landscape on the Canadian Shield showed significant differences in the impacts
among the four depression removal methods. Depression filling, a removal method that is widely implemented in
geographical information system software, was found to impact terrain attributes most severely. Constrained breaching,
which relies heavily on filling for larger depressions, also performed poorly. Both depression breaching and the IRA
impacted spatial and statistical distributions of terrain attributes less than depression filling and constrained breaching.
The most sensitive landscapes to depression removal were those that contained large (i.e. >10%) flat areas, because
of the occurrence of relatively large depressions in these areas.
 
 
Lindsay, J. B., et al. (2008). "Mapping outlet points used for watershed delineation onto DEM-derived stream networks." Water Resources Research 44(W08442).
Outlet point positions taken from hydrometric stations commonly do not coincide
with stream locations extracted from digital elevation models (DEMs). This is a serious
problem for accurate watershed delineation of data sets containing numerous outlets,
which is critical in regional-scale studies that relate catchment characteristics to basin
responses. The advanced outlet repositioning approach (AORA), presented here, replicates
the processes involved in manual outlet placement while reducing inefficiency and
potential for blunders. The technique uses water body names to identify locations for
outlet repositioning that are consistent with nearby outlets. The AORA performance was
compared against two existing automated techniques using 993 stations in seven basins in
northwest England. The AORA had the fewest repositioning errors in each basin and
nearly halved the overall number of errors in the data set, compared with the second-best
method. This work highlights the potential errors that may be present in studies that have
employed existing automated watershed mapping methods.
 
 
Liquori, M. K. (2006). "Post harvest riparian buffer response: implications for wood recruitment and buffer design." Journal of American Water Resources Association 42(1): 177-189.
Liquori, M. K. (2006). "Post-harvest riparian buffer response: Implications for wood recruitment modeling and buffer design." Journal of the American Water Resources Association 42: 177-189.
Despite the importance of riparian buffers in providing
aquatic functions to forested streams, few studies have sought to cap-ture
key differences in ecological and geomorphic processes between
buffered sites and forested conditions. This study examines post-har-vest
buffer conditions from 20 randomly selected harvest sites within
a managed tree farm in the Cascade Mountains of western Washing-ton.
Post-harvest wind derived treefall rates in buffers up to three
years post-harvest averaged 268 trees/km/year, 26 times greater than
competition-induced mortality rate estimates. Treefall rates and stem
breakage were strongly tied to tree species and relatively unaffected by
stream direction. Observed treefall direction is strongly biased toward
the channel, irrespective of channel or buffer orientation. Fall direction
bias can deliver significantly more wood recruitment relative to ran-domly
directed treefall, suggesting that models that utilize the random
fall assumption will significantly underpredict recruitment. A simple
estimate of post-harvest wood recruitment from buffers can be
obtained from species specific treefall and breakage rates, combined
with bias corrected recruitment probability as a function of source dis-tance
from the channel. Post-harvest wind effects may reduce the
standing density of trees enough to significantly reduce or eliminate
competition mortality and thus indirectly alter bank erosion rates,
resulting in substantially different wood recruitment dynamics from
buffers as compared to unmanaged forests.
 
Lisle, T. (1989). "Sediment transport and resulting deposition in spawning gravels, North Coastal California." Water Resources Research 25(6): 1303-1319.
Incubating salmonid eggs in streambeds are often threatened by deposition of fine sediment within
the gravel. To relate sedimentation of spawning gravel beds to sediment transport, infiltration of fine
sediment (<2 mm in diameter) into clean gravel beds, bed material size distributions, scour-fill depths,
and sediment transport during 10 storm flow events were measured in three streams of north coastal
California. Although suspended sediment comprised most (75-94%) of the clastic load during storm
flows, bed load material (0.25-2 mm) accounted for most (70-78%) of the fine sediment accumulated in
experimental gravel implanted in the streambeds. Sand trapped in the interstices of the top several
centimeters formed a seal that impeded deeper deposition of very fine sand and finer material. The seal
was responsible at least in part for a decrease in the rate of fine-sediment accumulation with increasing
cumulative bed load transport. Areas of the streambeds commonly scoured or filled 0.1 m or more
during storm flows, and thus scour and fill commonly created a sandy layer at least as thick as the seal
formed by sediment infiltration. Scour could erode eggs laid in the bed and expose deeper levels of the
bed to infiltration by fine sediment, but at the same time could allow fine sediment to be winnowed
away. Great temporal and spatial variation in sedimentation in these streams suggests that individual
storms of moderate size pose a threat to eggs in many but not all areas selected by fish for spawning.
 
Lisle, T. and M. Church (2002). "Sediment transport-storage relations for degrading, gravel bed channels." Water Resources Research 38(11): 1029/2001WR001086.
Lisle, T. E. (1981). The Recovery of Stream Channels in North Coastal California from Recent Large Floods. Habitat Disturbance and Recovery, California Trout, Inc.
Lisle, T. E. (1982). "Effects of aggradation and degradation on pool-riffle morphology in natural gravel channels, northwestern California." Water Resources Research 18: 1643-1651.
Lisle, T. E. (1982). "Effects of aggradation and degradation on riffle-pool morphology in natural gravel channels, Northwestern California " Water Resources Research 18(6): 1643-1651.
Lisle, T. E. (1986). "Stabilization of a gravel channel by large streamside obstructions and bedrock bends, Jacoby Creek, northwestern California." Geological Society of America Bulletin 97: 999-1011.
Lisle, T. E. (1989). "Sediment Transport and Resulting Deposition in Spawning Gravels, North Coastal California." Water Resources Research 25(6): 1303-1319.
Lisle, T. E., and J. Lewis (1992). "Effects of sediment transport on survival of salmonid embryos in a natural stream: a simulation approach." Canadian Journal of Fisheries and Aquatic Sciences 49: 2337-2344.
Lisle, T. E. (1995). "Particle size variations between bed load and bed material found in natural gravel bed channels." Water Resources Research 31(4): 1107-1118.
Lisle, T. E. (1995). "Particle size variations between bed load and bed material in natural gravel bed channels." Water Resources Research 31(4): 1107-1118.
Lisle, T. E., and S. Hilton (1999). "Fine bed material in pools of natural gravel bed channels." Water Resources Research 35(4): 1291-1304.
Lisle, T. E., J.M.Nelson, J. Pitlick, M.A.Madej, and B.L. Barkett (2000). "Variability of bed mobility in natural, gravel-bed channels and adjustments to sediment load at local and reach scales." Water Resources Research 36(12): 3743-3755.
Lisle, T. E. (2002). "How much dead wood in stream channels is enough." USDA Forest Service Gen. Tech. Rep. PSW-GTR-181: 85-93.
Private forest managers often seek guidelines on how much dead wood should be retained in
streams in order to adequately fulfill ecosystem functions. There are three approaches to
answering this question for a particular reach of channel. The first approach uses an
understanding of ecologic functions of dead wood in streams to determine the amount needed
to fulfill ecologic and geomorphic functions. This approach fails because the complexities of
sizes, shapes, and arrangements of dead wood in a variety of lotic ecosystems overwhelm any
scientific specification of target loadings. Another approach uses reference loadings to
evaluate departures in amounts of dead wood in streams from reference amounts in unaltered
systems. A precise threshold cannot be defined using this approach because dead wood
volumes are highly variable, even within pristine channels in similar settings, and
distributions for managed and pristine channels overlap. A third approach constructs a wood
budget by evaluating past, present, and projected supplies in streams and riparian areas. This
is a cumulative-effects analysis that shifts the focus from channels to riparian forests. In
combination, the three approaches provide the best information to determine how much wood
is enough, but they do not offer simple, formulaic prescriptions. The demands for performing
the necessary analyses before harvesting riparian wood suggest that management of riparian
forests will continue to be guided most often by general prescriptions.
 
Lisle, T. E. (?). Effects of Woody Debris on Anadromous Salmonid Habitat, Prince of Wales Island, Southeast Alaska. Arcata, CA, USDA Forest Service: 27.
Lisle, T. E. (?). "How Much Dead Wood in Stream Channels Is Enough?".
Lisle, T. E. and R. E. Eads (1991). Methods to Measure Sedimentation of Spawning Gravels Berkeley, California, US Forest Service: 1-7.
Lisle, T. E. and S. Hilton (1992). "The volume of fine sediment in pools: an index of sediment supply in gravel-bed streams." Water Resources Bulletin 28(2): 371-383.
Lisle, T. E., et al. (1991). "Formation of stationary alternate bars in a steep channel with mixed-size sediment: a flume experiment." Earth Surface Processes and Landforms 16: 463-469.
Lisle, T. E. and H. H. Kelsey (1982). Effects of large roughness elements on the thalweg course and pool spacing. American Geomorphological Field Group Field Trip Guidebook, 1982 Conference, Pinedale, Wyoming, American Geophysical Union, Berkeley, CA.
Lisle, T. E. and J. Lewis (1992). "Effects of Sediment Transport on Survival of Salmonid Embryos in a Natural Stream: A Simulation Approach " Canadian Journal of Fisheries and Aquatic Sciences 49: 2337-2344.
Lisle, T. E. and M. A. Madej (1992). Spatial Variation in Armouring in a Channel with High Sediment Supply. Dynamics of Gravel-bed Rivers. P. Billi, R. D. Hey, C. R. Thorne and P. Tacconi, John Wiley and Sons: 277-292.
Lisle, T. E., et al. (1998). Effects of recent logging on the main channel of North Fork Caspar Creek. General Technical Report PSW, Report: PSW-GTR-168: 81-85.
The response of the mainstem channel of North Fork Caspar Creek to recent logging is examined by time trends in bed load yield, scour and fill at resurveyed cross sections, and the volume and fine-sediment content of pools. Companion papers report that recent logging has increased streamflow during the summer and moderate winter rainfall events, and blowdowns from buffer strips have contributed more large woody debris. Changes in bed load yield were not detected despite a strong correlation between total scour and fill and annual effective discharge, perhaps because changes in stormflows were modest. The strongest responses are an increase in sediment storage and pool volume, particularly in the downstream portion of the channel along a buffer zone, where large woody debris (LWD) inputs are high. The association of high sediment storage and pool volume with large inputs of LWD is consistent with previous experiments in other watersheds. This suggests that improved habitat conditions after recent blowdowns will be followed in future decades by less favorable conditions as present LWD decays and input rates from depleted riparian sources in adjacent clearcuts and buffer zones decline.
 
Lisle, T. E., et al. (2000). "Variability of bed mobility in natural, gravel-bed channels and adjustment to sediment load at local and reach scales." Water Resources Research 36(12): 3743-3755.
Lisle, T. E., et al. (2000). "Variability of bed mobility in natural, gravel-bed channels and adjustments to sediment load at local and reach scales " Water Resources Research 36(12): 3743-3755.
Lisle, T. E., et al. (1997). "Evolution of a sediment wave in an experimental channel." Water Resources Research 33(8): 1971-1981.
Lisle, T. E., et al. (2001). "The Dominance of Dispersion in the Evolution of Bed Material Waves in Gravel-Bed Rivers." Earth Surface Process and Landforms 26: 1409-1420.
Litschert, S. E., and L.H. MacDonald, (2004). "Connectivity of timber harvest units to the stream network in the Sierra Nevada,California." Eos Trans. AGU, 85(47), Fall Meet. Suppl., Abstract H51B-1124. San Francisco, California.
Liu, B., M.A. Nearing, C. Baffaut, and J.C. Ascough (1997). "The WEPP watershed model: III. Comparison to measured data from small watersheds." Transactions of the ASAE 40(4): 945-952.
Loague, K. M. and R. A. Freeze (1985). "A Comparison of Rainfall-Runoff Modeling Techniques on Small Upland Catchments." Water Resources Research 21(2): 229-248.
Loch, R. J. (2000). "Using rainfall simulation to guide planning and management of rehabilitated areas: Part I. Experimental methods and results from a study at the Northparkes Mine, Australia." Land Degradation & Development 11(3): 221-240.
A study using simulated rain and overland flows was carried out on a 6-km-long bund constructed at the NorthParkes Mine, New South Wales, Australia. The bund has been progressively constructed using 5-m lifts to achieve final heights of 10-20 m. The slope profile consisted of alternate benches and batters of approximately 30 per cent gradient.For the topsoil placed on the bund, properties assessed included infiltration, erodibility, effects of plant cover on infiltration and erosion, and of slope length on erosion. Studies also considered permeability of the underlying spoil material, and compared susceptibility to erosion by overland flows of the two spoil materials available for bund construction.For sites that had been revegetated for one and two years, infiltration increased and soil erodibility decreased with time under vegetation. There were large increases in infiltration with increasing vegetative cover, and high infiltration rates at 100 per cent cover. The soil loss ratio curve fitted to erosion data show ed larger effects of cover on erosion than predicted by the universal soil loss equation (USLE). Overland flow studies indicated no effect of slope length on erosion rates, and indicated 'target' levels of vegetative cover necessary to prevent rill development.Permeability of spoil underlying topsoil at one site was shown to be high, and unlikely to impede water entry and drainage. Overland flow studies showed marked differences between the two spoils studied in both critical shear for rill initiation and in rill erodibilities.Implications of the results obtained on rehabilitation planning and management are discussed. Copyright (C) 2000 John Wiley & Sons, Ltd.
 
Loewenherzlawrence, D. S. (1994). "Hydrodynamic Description for Advective Sediment Transport Processes and Rill Initiation." Water Resources Research 30(11): 3203-3212.
A physically based model for advective sediment transport during rill initiation is developed based on (1) a short-wavelength rescaling of the momentum balance equations for the surface water flux; (2) a deforming coordinate system which characterizes local morphological changes of the surface during erosion; and (3) a generalized transport function for the surface sediment flux which distinguishes the contributions of diffusive and advective transport. A linear stability analysis of the coupled system indicates that an intrinsic minimum transverse length scale for rill initiation is forced by the cross-slope pressure gradient in the surface water. The wavelength of this initial rill spacing is dependent upon the relative significance of the advective component of sediment transport and the hydraulic efficiency of the surface water flux and is independent of the total length of the hillslope. The hydrodynamic model presented can also be used to evaluate the nonlinear development of rills and rill networks, subsequent to initiation.
 
Loftis, J. C., et al. (2001). "Detecting cumulative watershed effects: the statistical power of pairing." Journal of Hydrology 251: 49-64.
Logan, R. L. (? (No year needed)). "Slope Stability and Rural Land Management: The Marblemount Landslide Case History." Engineering Geology II: 903-910.
Logan, R. L., et al. (1991). Prediction of Sediment Yield from Tributary Basins along Huelsdonk Ridge, Hoh River, Washington, Washington State Department of Natural Resources: 1-14.
Logerwell, E. A., et al. (2003). "Tracking environmental processes in the coastal zone for understanding and predicting Oregon coho (Oncorhynchus kisutch) marine survival." Fisheries Oceanography 12(6): 554-568.
To better understand and predict Oregon coho
(Oncorhynchus kisutch) marine survival, we developed
a conceptual model of processes occurring during four
sequential periods: (1) winter climate prior to smolt
migration from freshwater to ocean, (2) spring transition
from winter downwelling to spring/summer
upwelling, (3) the spring upwelling season and (4)
winter ocean conditions near the end of the maturing
coho’s first year at sea. We then parameterized a
General Additive Model (GAM) with Oregon Production
Index (OPI) coho smolt-to-adult survival
estimates from 1970 to 2001 and environmental data
representing processes occurring during each period
(presmolt winter SST, spring transition date, spring
sea level, and post-smolt winter SST). The model
explained a high and significant proportion of the
variation in coho survival (R2 ¼ 0.75). The model
forecast of 2002 adult survival rate ranged from 4 to
8%. Our forecast was higher than predictions based on
the return of precocious males (‘jacks’), and it won’t be
known until fall 2002 which forecast is most accurate.
An advantage to our environmentally based predictive
model is the potential for linkages with predictive climate models, which might allow for forecasts more
than 1 year in advance. Relationships between the
environmental variables in the GAM and others (such
as the North Pacific Index and water column stratification)
provided insight into the processes driving
production in the Pacific Northwest coastal ocean.
Thus, coho may be a bellwether for the coastal environment
and models such as ours may apply to populations
of other species in this habitat.
 
Londo, H. A., P.A.Glass, D.L.Evans, K.L. Belli, R.C. Parker, T.G. Matney, E. B. Schultz, and N. Roller (2002). Integration of remote sensing and GPS with traditional forest inventory procedures. Proceedings for the 3rd Southern Forestry GIS Conference, University of Georgia, Athens, GA.
Long, B. A. (1987). Recruitment and abundance of large woody debris in an Oregon coastal stream system. Geography, Oregon State University: 143.
Long, C. J. (1995). Fire history of the central coast range, Oregon: a CA. 9000 year record from Little Lake. Geography.
Long, C. J., et al. (1998). "A 9000-Year Fire History from the Oregon Coast Range, Based on a High-Resolution Charcoal Study." Canadian Journal of Forest Resources 28(5): 774-787.
Lopes, V. L., P.F. Ffolliott, and M.B. Baker, Jr. (2001). "Impacts of vegetative practices on suspended sediment from watersheds of Arizona." Journal of Water Resources Planning and Management 127(1): 41-47.
Effects of vegetative practices on suspended sediment discharge from ponderosa pine forests and pinon-juniper woodlands in north-central Arizona are examined. Sediment-rating curves were developed to analyze the impacts. Disturbance from vegetative practices generally increased suspended sediment transport above those of control (reference) watersheds. Completely cleared and strip-cut ponderosa pine watersheds produced higher sediment concentrations than did a control watershed. Likewise, cabled and herbicide-treated pinon-juniper watersheds yielded higher sediment-laden streamflows than did a control. Sediment transport regimes are also related to streamflow-generation mechanisms and hydrograph stages. Although about 85% of the data analyzed represented snowmelt-runoff events in both vegetative types, derivation of sediment-rating curves based on streamflow-generation mechanisms improved the sensitivity of the analysis. Sediment data collected during rising and falling hydrograph stages varied between the two vegetative types. Sediment concentrations were generally higher in the rising stage than in the falling stage for ponderosa pine watersheds. There was no clear evidence of higher sediment concentrations in the rising stage of the hydrograph as compared to the falling stage in the pinon-juniper watersheds.
 
Lopes, V. L., and H.E. Canfield (2004). "Effects of watershed representation on runoff and sediment yield modeling." Journal of the American Water Resources Association 40(2): 311-319.
Lopez, J. L. and M. A. Falcon (1999). "Calculation of Bed Changes in Mountain Streams " Journal of Hydraulic Engineering 125(3): 263-270.
Lorensen, T., et al. (1994). Water Classification. Salem, Oregon, Oregon Department of Forestry: 14.
Lorente, A., et al. (2003). "Debris flow characteristics and relationships in the Central Spanish Pyrenees." Natural Hazards and Earth System Sciences 3: 683-692.
Unconfined debris flows (i.e. not in incised channels)
are one of the most active geomorphic processes in
mountainous areas. Since they can threaten settlements and
infrastructure, statistical and physically based procedures
have been developed to assess the potential for landslide erosion.
In this study, information on debris flow characteristics
was obtained in the field to define the debris flow runout
distance and to establish relationships between debris flow
parameters. Such relationships are needed for building models
which allow us to improve the spatial prediction of debris
flow hazards. In general, unconfined debris flows triggered in
the Flysch Sector of the Central Spanish Pyrenees are of the
same order of magnitude as others reported in the literature.
The deposition of sediment started at 17.8 , and the runout
distance represented 60% of the difference in height between
the head of the landslide and the point at which deposition
started. The runout distance was relatively well correlated
with the volume of sediment.
 
Lorenz, J. M. and J. H. Eiler (1989). "Spawning Habitat and Redd Characteristics of Sockeye Salmon in the Glacial Taku River, British Columbia and Alaska " Transactions of the American Fisheries Society 118: 495-502.
Lorimer, C. G. (1984). "Methodological considerations in the analysis of forest disturbance history." Can. J. For. Res. 15: 200-213.
Lubowe, J. K. (262). "Stream junction angles in the dendritic drainage pattern." American Journal of Science 262: 325-339.
Luce, C. H. (1997). "Effectiveness of road ripping in restoring infiltration capacity of forest roads." Restoration Ecology 5(3): 265-270.
Luce, C. H., and T.A. Black (1999). "Sediment production from forest roads in western Oregon." Water Resources Research 35(8): 2561-2570.
Luce, C. H., and and T. A. Black (2000). Erosion over time from forest roads in the Oregon Coast Range. Water Resources Center Report - Centers for Water and Wildland Resources, Report: 98. C. W. Slaughter: 140.
Luce, C. H. and T. A. Black (2001). Spatial and temporal patterns in erosion from forest roads. Land use and watersheds: Human influences on hydrology and geomorphology in urban and forest areas, American Geophysical Union. 2: 165-178.
Luce, C. H., et al. (2001). Effects of traffic and ditch maintenance on forest road sediment production. Proceedings - Federal Inter-Agency Sedimentation Conference, vol.7, Volume 2: V67-V74.
Observations of sediment yield from road segments in the Oregon Coast Range show that either heavy traffic during rainfall or blading the road ditch will increase erosion from forest roads. For the fine soils and high quality aggregate surfacing on the study plots, ditch blading increased sediment yield more than traffic equivalent to 12 log trucks per day. The combination of ditch blading and heavy traffic did not produce significantly more sediment than simply blading the ditch, a finding with important implications for sediment modeling and erosion control design. Increases in sediment production caused by traffic persisted after traffic ceased.
 
Luckman, B. H. (1990). "Mountain areas and global change: a view from the Canadian Rockies." Mountain Research and Development 10(2): 183-195.
Ludwig, D. (2001). "The Era of Management Is Over." Ecosystems 4: 758-764.
Lugt, J. d. and I. A. Campbell (1992). Mass movements in the badlands of Dinosaur Provincial Park, Alberta, Canada. Functional Geomorphology. K. M. Schmidt and J. d. Ploey. Cremlingen, Catena Verlag. Catena Supplement 23: 75-100.
Lunetta, R. S., et al. (1997). "GIS-based evaluation of salmon habitat in the Pacific Northwest." Photogrammetric Engineering and Remote Sensing 63(10): 1219-1229.
Luo, W. and T. Stepinski (2008). "Identification of geologic contrasts from landscape dissection pattern: an application to the Cascade Range, Oregon, USA." Geomorphology 99: 90-98.
This paper demonstrates the plausibility of inferring the spatial variability of geology from topographically derived landscape dissection
patterns. This enables surveying large regions for spatial variability in geology, for which direct remote sensing is not feasible, by studying
variability in dissection pattern, a feature extracted straight off from digital elevation model data. Dissection pattern is obtained automatically by a
novel algorithm, especially designed to delineate the valleys with high accuracy in order to reflect spatial variability in dissection density. The
dissection pattern is encapsulated by a continuous map of drainage density, a raster variable best suited for showing spatial variability of
dissection. Such a map, constructed for the study area in the Cascade Range, Oregon, USA, shows a sharp discontinuity in the dissection pattern,
indicating change in underlying geology. Possible factors controlling the dissection pattern such as climate, local and regional slopes, vegetation,
and geology are examined, and geology has been found to be the dominant controlling factor. The dissection contrast coincides with the boundary
between the Western and High Cascades, two geologic provinces with different rock ages and types. The older and less permeable Western
Cascades are associated with denser dissection pattern, whereas the younger and more permeable High Cascades correspond to less dissected
pattern. This new mapping method can be applied to locations where topography is the only readily available data, and the generated map could be
used to extract previously unknown geologic or environmental information.
 
Lupton, R. (1993). Statistics in Theory and Practice. Princeton, N.J., Princeton University Press.
Lyon, S. W., et al. (2008). "Incorporating landscape characteristics in a distance metric for interpolating between observations of stream water chemistry." Hydrology and Earth System Sciences 12: 1229-1239.
Spatial patterns of water chemistry along stream
networks can be quantified using synoptic or “snapshot”
sampling. The basic idea is to sample stream water at many
points over a relatively short period of time. Even for intense
sampling campaigns, the number of sample points is limited
and interpolation methods, like kriging, are commonly used
to produce continuous maps of water chemistry based on the
point observations from the synoptic sampling. Interpolated
concentrations are influenced heavily by how distance between
points along the stream network is defined. In this
study, we investigate different ways to define distance and
test these based on data from a snapshot sampling campaign
in a 37-km2 watershed in the Catskill Mountains region (New
York State). Three distance definitions (or metrics) were
compared: Euclidean or straight-line distance, in-stream distance,
and in-stream distance adjusted according characteristics
of the local contributing area, i.e., an adjusted in-stream
distance. Using the adjusted distance metric resulted in a
lower cross-validation error of the interpolated concentrations,
i.e., a better agreement of kriging results with measurements,
than the other distance definitions. The adjusted
distance metric can also be used in an exploratory manner
to test which landscape characteristics are most influential
for the spatial patterns of stream water chemistry and, thus,
to target future investigations to gain process-based understanding
of in-stream chemistry dynamics.
 
Lyons, J. K. and R. L. Beschta (1983). "Land Use, Floods, and Channel Changes: Upper Middle Fork Williamette River, Oregon (1936-1980)." Water Resources Research 19(2): 463-471.
MacCracken, J. G. and W. C. Boyd (undated report). Spatial and temporal variation in stream origins on Longview Fibre Company timberlands in Washington, Longview Fibre Company.
MacDonald, A., et al. (1982). The Role of Large Organic Debris of Stream Channels Draining Redwood Forests Northwestern California. Late Cenozoic History and Forest Geomorphology of Humboldt County, California: Friends of the Pleistcene Pacific Cell Field Trip.
MacDonald, A. and K. W. Ritland (1989). Sediment Dynamics in Type 4 and 5 Waters: A Review and Synthesis. Olympia, Washington, PTI Environmental Services prepared for TFW/CMER Sediment, Hydrology and Mass Wasting Steering Committee and Washington Department of Natural Resources.
MacDonald, L. H. (1988). "An inexpensive portable system for drilling into subsurface layers." Soil Science of America Journal 52(6): 1817-1819.
MacDonald, L. H. (1992). Sediment Monitoring: Reality and Hope. EPA/USFS Technical Workshop on Sediments, Corvallis, Oregon.
MacDonald, L. H., D.M. Anderson, and W.E. Dietrich (1997). "Paradise threatened: land use and erosion on St. John, US Virgin Islands." Environmental Management 21: 851-863.
MacDonald, L. H. (2000). "Evaluating and managing cumulative effects: Process and constraints." Environmental Management 26(3): 299-315.
ABSTRACT / Cumulative effects (CEs) result from the com-bined
effect of multiple activities over space or time. This
implies a persistence through time and often a transmittal
mechanism through space. Environmental legislation often
requires a broader CE assessment in addition to the more
direct, project-specific impacts. Current efforts to evaluate
and manage CEs are hampered by the conceptual problems
of defining the key issues, specifying the appropriate spatial
and temporal scales, and determining the numerous interac-tions
and indirect effects. These problems can be greatly
alleviated by following an explicit process. The process pro-posed
in this paper includes a scoping phase, an analysis
phase, and a planning and management phase, with each
phase consisting of two to five discrete but interrelated
tasks.
Numerous approaches have been developed to assess
CEs, and these range from simple checklists to complex,
physically based models. The utility of each approach de-pends
on the resource of concern, relative risk to those re-sources,
information available, and time frame for the evalu-ation.
In nearly all cases the assessment and regulation of CEs is
severely hampered by the variability in site conditions and
management effects, inability to predict secondary or indi-rect
effects, lack of data on recovery rates, difficulty of vali-dating
predictive models, and uncertainty of future events.
 
Since any proposed activity could contribute to a wide range
of potential CEs at different spatial and temporal scales, a
tiered or nested approach should be followed to assess CEs.
The difficulty of assessing and predicting CEs also suggests
that in many cases the most efficient approach is to focus on
minimizing on-site impacts. Under some circumstances
adaptive management can also be a viable alternative to
detailed CE assessments. Regular monitoring and feedback
is critical to the successful management and regulation of CEs.
 
MacDonald, L. H. (2000). "Evaluating and managing cumulative effects: process and contraints." Environmental Management 26(3): 299-315.
MacDonald, L. H., R.W. Sampson, and D.M. Anderson (2001). "Runoff and road erosion at the plot and road segment scale, St. John, U.S. Virgin Islands." Invited paper for a special issue on roads, Earth Surface Processes and Landforms 26: 1-22.
MacDonald, L. H., and J.D. Stednick (2003). Forests and water: a state-of-the-art review for Colorado. CWRRI Completion Report No. 196, CWRRI Completion Report No. 196, Colorado State University, Fort Collins, Colorado: 65pp.
MacDonald, L. H., D.B. Coe, and S.E. Litschert (2004). Assessing cumulative watershed effects in the Central Sierra Nevada: hillslope measurements and catchment-scale modeling. Proceedings of the Sierra Nevada Science Symposium,, U.S.D.A. Forest Service Gen. Tech. Rep. PSW-GTR-193, Albany, CA,.
MacDonald, L. H., and J. Pietraszak, (2005). Post-fire erosion at the hillslope scale in the Colorado Front Range: rates and controls. Paper presented at AGU Hydrology Days,, Colorado State University, Fort Collins, CO.
MacDonald, L. H., and D. Coe (2007). "Influence of headwater streams on downstream reaches in forested areas." Forest Science In Press.
MacDonald, L. H. and D. Coe (2007). "Influence of headwater streams on downstream reaches in forested areas." Forest Science 52(2): 148-168.
MacDonald, L. H., et al. (2001). "Runoff and road erosion at the plot and road segment scales, St John, US Virgin Islands." Earth Surface Processes and Landforms 26(3): 251-272.
Previous studies have identified unpaved roads as the primary source of erosion on St John in the US Virgin Islands, but these studies estimated road erosion rates only as annual averages based primarily on road rill measurements. The goal of this project was to quantify the effect of unpaved roads on runoff and sediment production on St John, and to better understand the key controlling factors. To this end runoff and sediment yields were measured from July 1996 to March 1997 from three plots on naturally vegetated hillslopes, four plots on unpaved road surfaces and two cutslope plots. Sediment yields were also measured from seven road segments with contributing areas ranging from 90 to 700 m(2).With respect to the vegetated plots, only the two largest storm events generated runoff and there was no measurable sediment yield. Runoff from the road surface plots generally occurred when storm precipitation exceeded 6 mm, Sediment yields from the four road surface plots ranged from 0.9 to 15 kg m(-2) a(-1), and sediment concentrations were typically 20-80 kg m(-3). Differences in runoff between the two cutslope plots were consistent with the difference in upslope contributing area. A sprinkler experiment confirmed that cross-slope roads intercept shallow subsurface stormflow and convert this into surface runoff. At the road segment scale the estimated sediment yields were 0.1 to 7.4 kg m(-2) a(-1).Road surface runoff was best predicted by storm precipitation, while sediment yields for at least three of the four road surface plots were significantly correlated with storm rainfall, storm intensity and storm runoff. Sediment yields at the road segment: scale were best predicted by road surface area, and sediment yields per unit area were most strongly correlated with road segment slope. The one road segment subjected to heavy traffic and more frequent regrading produced more than twice as much sediment per unit area than comparable segments with no truck traffic. Particle-size analyses indicate a preferential erosion of fine particles from the road surface and a rapid surface coarsening of new roads. Published in 2001 by John Wiley & Sons, Ltd.
 
MacDonald, L. H. a. E. L. H. (2004). "Post-fire soil water repellency: persistence and soil moisture thresholds." Soil Sci Soc Am J 68: 1729-1734.
MacDonald, L. H. a. P. R. R. (2007). JFSP Final Report: Postfire erosion and effectiveness of emergency rehabilititation treatments over time, JFSP project number: 03-2-3-22.
MacIsaac, E., et al. (2002). Aquatic Communities in Small Streams: A Review of Their Ecology and Efffects of Forest Harvesting in the Pacific Northwest, Workshop on Small Stream Channels and their Riparian  Zones: Their Form, Function and Ecological Importance in a Watershed Context.
Mackin, J. H. (1941). "A geologic interpretation of the failure of the Cedar Reservoir, Washington." University of Washington Engineering Experiment Station Bulletin 107: 1-31.
Macklin, M. G. and J. Lewin (1989). "Sediment transfer and transformation of an alluvial valley floor: the River South Tyne, Northumbria, U.K." Earth Surface Processes and Landforms 14: 233-246.
The amounts of alluvial storage and the mechanisms responsible for the dispersal of sediment associated with historic metal mining are examined for a 115 year timespan on a 22 km reach of the River South Tyne, northern England. Analysis of lateral and vertical channel change over this period, the extent of actively reworked gravels, and identification of metals in alluvial units, show five 'sedimentation zones' separated by more stable reaches. Aggradation in the late nineteenth century was folowed by a period of storage and local reworking, and then incision. Some reaches show short-term storage of sediment injected from tributaries which may be dispersed by floods. There is some evidence of transfer of sediment bodies downvalley within sedimentation zones, and of a sediment 'wave' movement between zones in lower parts of the South Tyne investigated. Present-day incision, the variation in the area of actively reworked gravel spreads, and the widespread dispersal of fine-grained toxic metal mining wastes all have practical implications. The intricate response of the channel system to historic changes in sediment supply is not readily described by either a complex response model or a simple translatory wave.
 
Macnab, K., et al. (2006). "Spatial variability of controls on downstream patterns of sediment storage: a case study in the Lane Cove Catchment, New South Wales, Australia." Geographical Research 44(3): 255-271.
This study uses GIS techniques to examine the spatial distribution of stream
power along major streamlines in the Lane Cove catchment in northern Sydney,
Australia. Channel gradient estimates derived from a 5 m resolution digital
elevation model (DEM) are combined with streamflow data to estimate stream
power along river courses. Stream power and its constituent components are then
related to a detailed field-based assessment of sediment storage along the trunk
stream and primary tributaries. At the catchment scale, sediment storage per unit
length decreases as channel gradient and gross stream power increase. However,
local controls such as variability in valley width and occurrence of confluence
zones exert a greater influence upon sediment storage, disrupting systematic
catchment-wide relationships. The total volume of storage along each streamline
has a strong linear relationship to the area of the subcatchment, but the distribution
of sediment along streamlines varies between subcatchments. The GIS framework
employed in this project allows generation of continuous, empirical data, thereby
providing catchment-specific predictive capacity that can accompany theoretical
approaches to stream power modelling.
 
MacNally, R., et al. (2002). "Current Loads of Coarse Woody Debris on Southeastern Australian Floodplains:Evaluation of Change and Implications for Restoration." Restoration Ecology 10(4): 627-635.
Maddock, T., Jr. (1976). "A primer on floodplain dynamics " Journal of Soil and Water Conservation 31(2): 44-47.
Madej, M. A. (1978). Response of a stream channel to an increase in sediment load, University of Washington: 136.
Madej, M. A. (1990). Changes in Channel-Stored Sediment, Redwood Creek,Northwestern California, 1947 to 1980. Geomorphic Processes and Aquatic Habitat in the Redwood Creek Basin, Northwestern California. U.S. Geological Survey Professional Paper 1454. K. M. Nolan, H. M. Kelsey and D. C. Marron. Washington D.C.
Madej, M. A. (1992). Changes in Channel-Stored Sediment, Redwood Creek, Northwestern California, 1947 to 1980, U.S. Geological Survey: 38.
Madej, M. A., and V. Ozaki (1998). "Channel response to sediment wave propagation and movement, Redwood Creek, California." Earth Surface Processes and Landforms 21(10): 911-927.
Madej, M. A. (1999). "Temporal and spatial variability in thalweg profiles of a gravel bed river." Earth Surface Process and Landforms 24: 1153-1169.
Madej, M. A. (1999). "Temporal and spatial variability in Thalweg profiles of a gravel-bed river." Earth Surface Processes and Landforms 24: 1153-1169.
Madej, M. A. (2001). "Erosion and sediment delivery following removal of forest roads." Earth Surface Processes and Landforms 26(2): 175-190.
Madej, M. A., et al. (2006). "Assessing possible thermal rearing restrictions for juvenile coho salmon (Oncorhynchus kisutch) through thermal infrared imaging and in-stream monitoring, Redwood Creek, California." Canadian Journal of Fisheries and Aquatic Science 63: 1384-1396.
We quantified patterns in stream temperature in a northern coastal California river using thermal infrared
(TIR) imaging and in-stream monitoring and related temperature patterns to the historical and present distributions of
juvenile coho salmon (Oncorhynchus kisutch). In Redwood Creek, California, water temperature increased from the
headwaters to about 60 km downstream, then gradually decreased over the next 40 km as the river approaches the
Pacific Ocean. Despite the lack of fish migration barriers, juvenile coho are currently only observed in the downstream-most
20 km, whereas historically they were found in 90 km of river channel. Maximum daily temperatures and dura-tion
of elevated stream temperatures were not significantly different in the headwater and downstream reaches but were
significantly higher in the 50 km long intervening reach, where maximum weekly maximum temperatures ranged from
23 to 27 °C. An increase in stream temperatures in the middle basin during the last three decades as a result of chan-nel
aggradation, widening, and the removal of large riparian conifers may play an important role in restricting juvenile
coho to one-fifth of their historical range.
 
Madej, M. A. and V. Ozaki (1994). Channel response and partial recovery after large sediment inputs to Redwood Creek, California (DRAFT).
Madej, M. A. and V. Ozaki (1996). "Channel response to sediment wave propagation and movement, Redwood Creek, California, USA." Earth Surface Processes and Landforms 21: 911-927.
Madsen, H., et al. (2002). "Regional estimation of rainfall intensity-duration-frequency curves using generalized least squares regression of partial duration series statistics." Water Resources Research 38(11): 10.1029/2001WR001125.
Magana, A. E. M. (2001). "Litter input from riparian vegetation to streams: a case study of the Njoro River, Kenya." Hydrobiologia 458: 141-149.
Magee, J. P., et al. (1996). "Spatial variation in spawning habitat of cutthroat trout in a sediment-rich stream basin." Transactions of the American Fisheries Society 125: 768-779.
We examined distribution and habitat characteristics of spawning sites of cutthroat
trout Oncorhynchus clarki at various spatial scales to assess effects of sedimentation within a large
basin in Montana. Redd density varied widely across the basin; nearly all (99%) of the 362 redds
observed occurred in two high-elevation headwater tributaries. Redd density at the reach scale
was positively correlated (r2 = 0.72, P = 0.001) with abundance of spawning gravels. Other
habitat variables, such as gradient, width, depth, embeddedness, bank stability, and percent riffle,
were not significantly correlated to redd density. Taylor Fork redds contained some of the highest
proportions of fine sediments (<6.35 mm, mean = 41.6%; <0.85 mm, 17.9%) observed in egg
pockets of salmonid redds in the Rocky Mountain region. Cache Creek, a highly disturbed subbasin,
had significantly greater proportions of fine sediments smaller than 0.85 mm in redds than the
undisturbed Wapiti Creek subbasin. High fine-sediment levels in redds led to very low estimated
embryo survival (mean, 8.5%), but sedimentation did not appear to limit recruitment. Our data
suggest that compensatory juvenile survival and high embryo survival in the small proportion of
redds with good substrate quality may buffer the effects of the high sediment levels in the basin.
 
Magilligan, F. J., et al. (2003). "Scale-independent assessment of discharge reduction and riparian disconnectivity following flow regulation by dams." Geology 31(7): 569-572.
By using the established hydraulic relationships among flood frequency, flood magnitude, and river-channel capacity, we develop a scale-independent assessment of the hydrogeomorphic impacts of 21 dams across the United States that have broad ranges in function and contributing drainage area. On the basis of generalized extreme value (GEV) analysis of pre- and post-dam hydrologic records, our analysis indicates that the 2 yr discharge has decreased {sim}60{percnt} following impoundment, exceeding the magnitude of climatically triggered discharge reductions occurring during the Holocene. Reductions in the frequency of the pre-dam 2 yr discharge have been equally profound. The pre-dam 2 yr flood has occurred on average twice per site, whereas statistical analysis indicates that it should have occurred {sim}20 times. Furthermore, floods greater than bankfull have been essentially eliminated by dams, completely disconnecting the riparian zone from riverine influence. Our analyses herein suggest that a critical threshold of disconnectivity exists and corresponds approximately to the pre-dam 5 yr flood. This similar recurrence probability exists independent of region, dam type, or catchment size.
 
Magirl, C. S. and T. D. Olsen (2009). Navigability potential of Washington rivers and streams determined with hydraulic geometry and a geographic information system. U. S. G. Survey.
Mahoney, J. M. and S. R. Rood (1998). "Streamflow requirements for cottonwood seedling recruitment - an integrative model " Wetlands 18(4): 634-645.
Maidment, D. R., et al. (1996). "Unit hydrograph Derived from a Spatially Distributed Velocity Field." Hydrological Processes 10: 831-844.
Majitan, S., et al. (1999). "Catastrophic flood in eastern Slovakia." J. Fac. Agr. 44: 213-217.
Malamud, B. D., et al. (2004). "Landslide inventories and their statistical properties." Earth Surface Processes and Landforms 29(6): 687-711.
Landslides are generally associated with a trigger, such as an earthquake, a rapid snowmelt or a large storm. The landslide event can include a single landslide or many thousands. The frequency-area (or volume) distribution of a landslide event quantifies the number of landslides that occur at different sizes. We examine three well-documented landslide events, from Italy, Guatemala and the USA, each with a different triggering mechanism, and find that the landslide areas for all three are well approximated by the same three-parameter inverse-gamma distribution. For small landslide areas this distribution has an exponential roll-over and for medium and large landslide areas decays as a power-law with exponent -2·40. One implication of this landslide distribution is that the mean area of landslides in the distribution is independent of the size of the event. We also introduce a landslide-event magnitude scale mL = log(NLT), with NLT the total number of landslides associated with a trigger. If a landslide-event inventory is incomplete (i.e. smaller landslides are not included), the partial inventory can be compared with our landslide probability distribution, and the corresponding landslide-event magnitude inferred. This technique can be applied to inventories of historical landslides, inferring the total number of landslides that occurred over geologic time, and how many of these have been erased by erosion, vegetation, and human activity. We have also considered three rockfall-dominated inventories, and find that the frequency-size distributions differ substantially from those associated with other landslide types. We suggest that our proposed frequency-size distribution for landslides (excluding rockfalls) will be useful in quantifying the severity of landslide events and the contribution of landslides to erosion.
 
Malanson, G. P. (1993). Riparian landscapes. Riparian landscapes, Press Syndicate of the University of Cambridge: 80-227.
Malard, F., et al. (2002). "A landscape perspective of surface-subsurface hydrological exchanges in river corridors." Freshwater Biology 47(4): 621-640.
Malard, F., et al. (2000). "Physio-chemical heterogeneity in a glacial riverscape." Landscape Ecology 15: 679-695.
Malik, I. (2008). "Dating of small gully formation and establishing erosion rates in old gullies under forest by means of anatomical changes in exposed tree roots (Southern Poland)." Geomorphology 93: 421-436.
Malmaeus, J. M. and M. A. Hassan (2002). "Simulation of Individual Particle Movement in a Gravel Streambed." Earth Surface Process and Landforms 27: 81-97.
Mandelbrot, B. B. and J. R. Wallis (1968). "Noah, Joseph, and Operational Hydrology." Water Resources Research 4(5): 909-917.
Mandelbrot, B. B. and J. R. Wallis (1969). "Computer Experiments with Fractional Gaussian Noises. Part 1, Averages and Variances." Water Resources Research 5(1): 228-265.
Manga, M. and J. W. Kirchner (2000). "Stress partitioning in streams by large woody debris " Water Resources Research 36(8): 2373 - 2379.
Manners, R. B. and M. W. Doyle (2008). "A mechanistic model of woody debris jam evolution and its application to wood-based restoration and management." River Research and Applications 24: 1104-1123.
The natural tendency of woody debris to accumulate into complex debris jams has been adapted by the restoration industry
because of the morphological and ecological benefits of these structures. While much work has been done on woody debris, there
is a lack of understanding of the dynamics of debris jams including the controls on their formation and the associated changes in
hydraulics. Treatment of jams as static structures, whose hydraulics may be described by that of a single-solid object, prevents
optimal success of wood-based restoration projects. This paper reviews the state of the science on the initiation and accumulation
of wood forming a debris jam. This review is used to develop a conceptual model of the evolution of a single debris jam focussing on
the relationship between the structure and hydraulics and the feedback that exists between them. The proposed mechanisms
behind debris jam evolution are supported by a case-study of three natural jams. Incorporation of this model into restoration and
management plans will result in more successful and cost-efficient projects
 
Manouchehr Dadkhah, a. G. F. G. (2007). "INFLUENCE OF VEGETATION, ROCK COVER, AND TRAMPLING ON INFILTRATION RATES AND SEDIMENT PRODUCTION." Journal of the American Water Resources Association 16(6): 979-986.
Marcos, E., et al. (2000). "Comparative analysis of runoff and sediment yield with a rainfall simulator after experimental fire." Arid Soil Research and Rehabilitation 14(3): 293-307.
Changes produced in runoff and sediment levels before and after fire and during the revegetation process were examined using a rainfall simulator. The area was burned in an experimental fire, reaching temperatures from 35 degrees to 563 degrees C. Then it was revegetated using different species combinations. Fifteen permanent plots were established in the burnt area (4 treatments and a control replicated three times). Simulated rainfall of 15 mm per 5 min was applied in each treatment. No significant differences were found in sediment yield and runoff between treatments, but greatest runoff was observed to occur immediately after the fire. A significant relationship was found between runoff and woody cover, and a decrease in runoff can be observed as cover increases. The relationship between sediment yields and runoff rates was also positive. The low rates observed during rainfall simulation are due to the effect of natural vegetation rather than revegetation treatments. The high organic matter content also had an influence on the low rates of runoff and sediment.
 
Marcus, A. (1980). "First-Order Drainage Basin Morphology - Definition and Distribution." Earth Surface Processes 5: 389-398.
Marek, D. W., et al. (1996). "Aquatic sediments." Water Environment Research 68(4): 629-662.
Marion, A. and L. Fraccarollo (1997). "Experimental investigation of mobile armoring development." Water Resources Research 33(6): 1447-1453.
Marion, D. A., and S.J. Ursic (1992). Sediment production in Forests of the Coastal Plain, Piedmont, and Interior Highlands. Proceedings of the EPA/ ForestService Workshop; Corvallis, OR.
Marion, D. A., and J.A. Clingenpeel (2007?). "Methods used for analyzing the cumulative watershed effects of fuel management on sediment in the eastern United States." submitted.
Mark, D. M. (1988). Network models in geomorphology. Chichester, John Wiley & Sons Ltd.
Mark, R. K. and E. B. Newman (1988). Rainfall totals before and during the storm: Distribution and correlation with damaging landslides. Landslides, Floods, and Marine Effects of the Storm of January 3-5, 1982, in the San Francisco Bay Region, California. U.S. Geological Survey Professional Paper 1434. S. D. Ellen and G. F. Wieczorek. Washington, D. C.
Marques, M. A. and E. Mora (1992). "The Influence of Aspect on Runoff and Soil Loss in a Mediterranean Burnt Forest (Spain)." Catena 19(3-4): 333-344.
In order to evaluate the soil loss in a burnt forest and the role played by slope orientation in erosion, a study has been carried out in the Montserrat area (NE Spain) affected by a fire in summer 1986. Previous geomorphological surveys allowed us to propose the hypothesis that slope orientation acted as a macrofactor or integrating factor in erosion control. Aspect controls temperature, moisture and water supply and then vegetation and soil development.The evaluation was determined by monitoring two experimental plots 250 m apart developed on equal substrate and slope but one facing north and the other facing south. The plots, with a surface area of 200 m2 are equipped with collecting and sampling units (Multi-slot divisors) and storage tanks.The results (August 1987 - December 1988) confirm the contrast in runoff and soil losses according to slope orientation (2176 g/m2 south-facing plot and 352 g/m2 north-facing plot). This different erosive response is mainly due to: a) the state of the vegetation and soil development before the fire, b) the differences in vegetation recovery after fire.The north-facing slopes had a denser plant cover and thicker soil than the south-facing ones. After the fire the north-facing slopes were covered by a thicker blanket of ashes and burnt plant debris and, later, more vegetation giving the soil higher infiltration capacity and resistance.Moreover, the two major events displayed different behaviour according to slope orientation. The first of those events, due to the conditions on each slope (higher or lower protection and available erodible material) gave rise to a change in the threshold conditions: a) the rill formation on the south-facing plot enhanced the flow concentration and sediment transport (lowering of threshold) in the second event and b) on the north-facing one, the lack of rills and the decrease in erodible material caused by the first event hindered subsequent erosion (threshold rise).Finally, the progressive recovery of spontaneous vegetation caused erosion to drop, reaching low rates in 1988 (22 g/m2 and 2 g/m2 on each plot).
 
Marron, D. C. (?). Colluvium in bedrock hollows on steep slopes, Redwood Creek drainage basin Northwestern California, ?: 1-.
Martin, C. S. and J. J. Zovne (1971). "Finite-difference simulation of bore propagation." Journal of the Hydraulics Division 7: 993-1010.
Martin, C. W. and W. C. Johnson (1987). "Historical Channel Narrowing and Riparian Vegetation Expansion in the Medicine Lodge River Basin, Kansas, 1871-1983." Annals of the Association of American Geographers 77(3): 436-449.
Martin, D. and L. Benda (2003). Habitat Configuration, Population Spatial Structure, and Core Areas for Chinook Salmon in King County Rivers/Development of Salmonid Conservation Strategies Phase I, Project No T01426T, King Count, Water and Land Resources Division, Seattle, WA.
Martin, D., et al. (2004). Core Areas: A Framework for Identifying Critical Habitat for Salmon. Seattle, WA, King County Department of Natural Resources and Parks, Water and Land Resources Division.
Martin, D. and A. Grotefendt (2007). "Stand mortality in buffer strips and the supply of woody debris to streams in Southeast Alaska." Canadian Journal of Forest Research 37: 36-49.
Martin, D. A. (2000). "Studies of post-fire erosion in the Colorado Front Range benefit the Upper South Platte Watershed Protection and Restoration Project." Watershed Management Council Networker v. Winter:: 5-8.
Martin, D. J. (2001). "The influence of geomorphic factors and geographic region on large woody debris loading and fish habitat in Alaska coastal streams." North American Journal of Fisheries Management 21: 429-440.
Large woody debris (LWD) and channel data from three Alaska coastal regions with
varying geomorphic channel types were analyzed to document regional variability in LWD abundance,
define geomorphic characteristics affecting LWD abundance, and identify relationships
between LWD abundance and the formation of pools and gravel bars in streams. Large woody
debris abundance was significantly lower at the northern edge of the coastal coniferous forest than
in Southeast Alaska and was significantly greater in alluvial gravel-bedded channels than in contained
boulder–bedrock channels. More pools and gravel bars were formed by LWD in alluvial
channels than in contained channels. Pool spacing (the number of channel widths between pools)
decreased with increasing LWD abundance (pieces/km) and was significantly influenced by the
interaction between LWD abundance and channel width. As channel width increased, pool spacing
was more strongly influenced by changes in LWD abundance, but the relative change in pool
spacing diminished with increasing LWD load. The percentage of stream area in pools was insensitive
to changes in LWD abundance and was best predicted by channel type. The percentage
of habitat units with gravel as the dominant substrate was positively related to LWD abundance
and negatively related to stream gradient.
 
Martin, D. J. (2001). "The Influence of Geomorphic Factors and Geomorphic Region on Large Woody Debris Loading and Fish Habitat in Alaska Coastal Streams." North American Journal of Fisheries Management 21: 429-440.
Martin, D. J. and L. E. Benda (2001). "Patterns of instream wood recruitment and transport at the watershed scale." Transactions of the American Fisheries Society 130: 940-958.
Martin, D. J. and R. A. Grotefendt (2007). "Stand mortality in buffer strips and the supply of woody debris to streams in Southeast Alaska." Canadian Journal of Forest Research 37: 36-49.
Martin, D. J. and M. E. Robinson (1998). The Effectiveness of Riparian Buffer Zones for Protection of Salmonid Habitat in Alaska Coastal Streams. Seattle, WA
North Bend, WA, Martin Environmental and Grotefendt Photogrammetric Services, Inc. prepared for Sealaska Corporation and Alaska Forest Association: 85.
Martin, K. (1977). Forest management and landslide prone sites: the effectiveness of headwall leave areas and evaluation of two headwall risk rating methods. Department of Civil Engineering. Corvallis, Oregon, Oregon State University.
Martin, Y. "Modelling hillslope evolution: linear and nonlinear transport relations."
Many recent models of landscape evolution have used a diffusion relation to simulate hillslope transport. In this study, a linear diffusion equation for slow, quasi-continuous mass movement (e.g., creep), which is based on a large data compilation, is adopted in the hillslope model. Transport relations for rapid, episodic mass movements are based on an extensive data set covering a 40-yr period from the Queen Charlotte Islands, British Columbia. A hyperbolic tangent relation, in which transport increases nonlinearly with gradient above some threshold gradient, provided the best fit to the data. Model runs were undertaken for typical hillslope profiles found in small drainage basins in the Queen Charlotte Islands. Results, based on linear diffusivity values defined in the present study, are compared to results based on diffusivities used in earlier studies. Linear diffusivities, adopted in several earlier studies, generally did not provide adequate approximations of hillslope evolution. The nonlinear transport relation was tested and found to provide acceptable simulations of hillslope evolution. Weathering is introduced into the final set of model runs. The incorporation of weathering into the model decreases the rate of hillslope change when theoretical rates of sediment transport exceed sediment supply. The incorporation of weathering into the model is essential to ensuring that transport rates at high gradients obtained in the model reasonably replicate conditions observed in real landscapes. An outline of landscape progression is proposed based on model results. Hillslope change initially occurs at a rapid rate following events that result in oversteepened gradients (e.g., tectonic forcing, glaciation, fluvial undercutting). Steep gradients are eventually eliminated and hillslope transport is reduced significantly.
 
Martin, Y., and M. Church (2000). "Re-examination of Bagnold's empirical bedload formulae." Earth Surface Processes and Landforms 25: 1011-1024.
Martin, Y. and M. Church (1995). "Bed-material transport estimated from channel surveys: Vedder River, British Columbia." Earth Surface Processes and Landforms 20: 347-361.
This study investigates the possibility to estimate bed-material transfer in gravel-bed rivers by analysis of morphological changes along Vedder River, British Columbia. Data from repeated cross-section surveys are used to estimate volume changes along the length of an 8 km reach. Gravel budgets are based on a continuity approachl. An error analysis is performed to evaluate the uncertainty in the best estimate of transport rates. The mean annual gravel transport into the reach over a 9 year period was estimated to be 36,600 ±5,600 m3/yr. The sediment transport regime along the length of the river is evaluated and examined in relation to peak flood flows. Significant spatial and temporal variability in transport rates is demonstrated, making dubious the generalization of transport estimates from hydraulic calculations, or from sample measurements at a single cross-section. The assumptions, proceedures and limitations of the 'morphological approach' to sedimtne transport analysis are discussed. It is concluded that this approach provides information of quality comparable or superior to that of direct measurements of transport, yet requires less field effort. It also provides additional information about river morphological changes, making it a preferred method for geomorphological investigations and for many river management concerns.
 
Martin, Y., et al. (2002). "Sediment transfer by shallow landsliding in the Queen Charlotte Islands, British Columbia." Canadian Journal of Earth Science 39.
Despite the importance of landsliding in routing sediment through mountainous drainage basins, few studies have documented landsliding rates over extended time and space scales. We have investigated landsliding in surficial material in the Queen Charlotte Islands using a large inventory of events, derived from aerial photography, covering an area of 166.7 km2. The mean erosion rate for shallow landsliding is 0.10 mm·a–1, which is at the upper end of shallow landsliding rates observed in the Pacific Northwest and coastal British Columbia, but several orders of magnitude lower than rock-based landsliding rates reported in the literature. Probability distributions for landslide area and volume are somewhat convex in form. Flattening of the curve found at low magnitudes may be due to sampling bias or physical mechanisms inhibiting failure, and the steepening for high values may exist because the sampling period is not long enough to adequately represent large events. Landslides generally initiate on hillslope gradients greater than 0.50–0.60. The largest numbers of landslides occur on south- to southwest-facing slopes and east- to northeast-facing slopes. Most events occur on concave and straight hillslopes in upper-slope positions. Landsliding rates were found not to be affected by rock type. Hillslopes in the Queen Charlotte Islands are often mantled by weathered Quaternary deposits and, hence, landsliding events are not directly controlled by weathering of bedrock. About 31% of landslides identified in this study deposited material in stream reaches, with about 83% of these landslides deposited in reaches with gradients between 3% and 10%.
 
Martin, Y. a. D. H. (2005). "Testing bedload transport formulae using morphological transport estimates and field data: Fraser River, British Columbia." Earth Surface Processes and Landforms 30: 1265-1282.
Martinelli, M., Jr. (1986). "A Test of the Avalanche Runout Equations Developed By the Norwegian Geotechnical Institute." Cold Regions Science and Technology 13: 19-33.
Martinez-Casasnovas, J. A., et al. (2003). "Sediment production in large gullies of the Mediterranean area (NE Spain) from high-resolution digital elevation models and geographical information systems analysis." Earth Surface Processes and Landforms 28(5): 443-456.
Martz, L. W. and E. d. Jong (1987). "Using cesium-137 to assess the variability of net soil erosion and its association with topography in a Canadian prairie landscape." Catena 14: 439-451.
Marzocchi, W., et al. (1997). "Detecting low-dimensional chaos in geophysical time series." Journal of Geophysical Research 102(B2): 3195-3209.
Maser, C., et al. (1988). From the Forest to the Sea: A Story of Fallen Trees. Portland, Oregon, U.S.D.A. Forest Service.
Massong, T. M. and D. R. Montgomery (2000). "influence of sediment supply, lithology, and wood debris on the distribution of bedrock and alluvial channels." Geological Society of America Bulletin 112(5): 591-599.
Mathewson, C. C., et al. (1990). "Role of bedrock ground water in the initiation of debris flows and sustained post-flow stream discharge." Bulletin of the Association of Engineering Geologists 27(1): 73-83.
Matthael, C. D., et al. (1999). "Scour and fill patterns in a New Zealand stream and potential implications for invertebrate refugia " Freshwater Biology 42: 41-57.
Matthews, W. J. (1998). Patterns in Freshwater Fish Ecology. New York, Chapman & Hall.
Maurer, B. A. (1999). Untangling Ecological Complexity. Chicago, University of Chicago Press.
Maxwell, J. R., et al. (1995). "A hierarchical framework of aquatic ecological units in North America (Nearctic Zone)." USDA Forest Service, North Central Forest Experiment Station, General Technical Report NC-176.
May, C. L. (1998). Debris Flow Characteristics Associated with Forest Practices in the Central Oregon Coast Range. Forest Engineering. Corvallis, OR, Oregon State University: 121.
May, C. L. (2001). Spatial and Temporal Dynamics of Sediment and Wood in Headwater Streams in the Central Oregon Coast Range. Corvallis, OR, Oregon State University: 154.
May, C. L. (2002). "Debris Flows through Different Forest Age Classes in the Central Oregon Coast Range." Journal of the American Water Resources Association 38(4): 1097-1113.
May, C. L. (2007). "Sediment and wood routing in steep headwater streams: an overview of geomorphic processes and their topographic signatures." Forest Science 53(2): 119-130.
May, C. L. and W. E. Dietrich (2004). "Slope-area controls on the expression of reach-scale channel morphology, debris flow runout, and the spatial distribution of salmonids in steep mountain streams." EOS Transactions 85(47): H53C-1269.
Steepness and concavity indexes derived from the power function relationship between drainage area and channel slope provide a process-based characterization of river profiles. We propose that these geomorphic indexes provide a useful context for classifying basins that express different reach morphologies, fish habitat capacity, and responses to episodic disturbance. Strongly concave profiles that develop in steep terrain indicate that almost all of the relief in the drainage network occurs in small headwater streams. In these basins a large proportion of the drainage network has low-gradient morphologies, such as pool-riffle sequences, which provide favorable rearing habitat for many salmonid species. Complex metapopulation structures can develop within these networks because fish distribution expands into the tributaries, allowing for a spatial spreading of risk that may enhance a population's ability to persist during adverse conditions for survival and growth. The severity of pulse disturbances is also reduced because debris flows typically form discrete deposits where steep tributaries abruptly encounter low-gradient mainstem channels at tributary junctions. In contrast, less concave profiles in steep terrain indicate that the spatial extent of high gradient reaches morphologies, such as step-pool and cascade sequences, are more extensive. Metapopulation development in these basins is diminished because most tributaries are too steep to provide habitat, confining fish to mainstem channels. Furthermore, the change in slope at tributary junctions is less pronounced and debris flows rarely form discrete deposits. Instead, these mass flows continue to travel down steep mainstem channels and alter aquatic and riparian habitats for long distances. The combined influence of a limited spatial distribution and the increased severity of debris flows may result in more extreme fluctuations in population abundance because they are less resilient to pulse disturbances.
 
May, C. L. and R. E. Gresswell (2003). "Large wood recruitment and redistribution in headwater streams in the southern Oregon Coast Range, USA." Canadian Journal of Forest Research 33: 1352-1362.
Large wood recritment and redistribution mechanisms were investigated in a 3.9 km2 basin with an old-growth Pseudotsuga menziesii (Mirg.) Franco and Tsuga heterophylla (Raf.) Sarg. forest, located in the southern Coast Range of Oregon. Stream size and topographic setting strongly influenced processes that delivered wood to the channel network. In small colluvial channels draining steep hillslopes, processes associated with slope instability dominated large wood recruitment. In the larger alluvial channel, windthrow was the dominant recruitment process from the local riparian area. Consequently, colluvial channels received wood from further upslope than the alluvial channel. Input and redistribution processes influenced piece location relative to the direction of flow and thus, affected the functional role of wood. Wood recruited directly from local hillslopes and riparian areas was typically positioned adjacent to the channel or spanned its full width, and trapped sediment and wood in transport. In contrast, wood that had been fluvially redistributed was commly located in min-channel positions and was associated with scouring of the streambed and banks. Debris flows were a unique mechanism for creating large accumulations of wood in small streams that lacked the capacity for abundant fluvial transport of wood, and for transporting wood that was longer than the bank-full width of the channel.
 
May, C. L. and R. E. Gresswell (2003). "Processes and Rates of Sediment and Wood Accumulation in Headwater Streams of the Oregon Coast Range, USA." Earth Surface Process and Landforms 28: 409-424.
May, C. L. and R. E. Gresswell (2004). "Spatial and temporal patterns of debris-flow deposition in the Oregon Coast Range, USA." Geomorphology 57(3-4): 135-149.
May, C. L. and D. C. Lee (2004). "The relationships among in-channel sediment storage, pool depth, and summer survival of juvenile salmonids in Oregon Coast Range streams." North American Journal of Fisheries Management 24: 761-774.
May, D. W. (2003). "Properties of a 5500-year-old flood-plain in the Loup River Basin, Nebraska." Geomorphology 56(3-4): 243-254.
Flood-plain aggradation within the Loup River Basin of central Nebraska was episodic and alternated with incision throughout much of the Holocene. A widespread episode of flood-plain stability, however, occurred about 5700-5100 cal. year BP. The purpose of this paper is to describe the properties of this buried flood-plain at six sites in the basin, to consider why the properties of the buried flood-plain vary from site to site, and to evaluate possible reasons why the Loup River flood-plains stabilized 5500 years ago. Episodic valley-bottom aggradation was common during flood-plain formation at five of the six sites. The radiocarbon ages, particle-size data, and organic-carbon data for the buried flood-plain reveal that valley-bottom aggradation generally slowed between about 5 700 and 5 100 cal. year BR Erratic down-profile changes in percentages of sand, clay, and organic matter indicate flood-plain sedimentation and soil formation were often episodic. Sand and clay rarely show a steady fining-upward trend. Organic matter fluctuates with depth; at some sites multiple, incipient A horizons were buried during waning valley-bottom aggradation. At two localities, the buried flood-plain is evident as a clay-rich stratum that must have been deposited in a paleochannel. Flood-plain stabilization between 5700 and 5100 cal. year BP probably occurred in response to the effects of external climate forcing on vegetation and hydrologic changes. flood-plains of other rivers in the central Great Plains also stabilized at this time, further supporting a climatic explanation for slowing of valley aggradation and formation of a flood-plain at this time. Recognition of buried flood-plains is important to both soil mapping in valleys and to the discovery of cultural resources in valleys. (C) 2003 Elsevier Science B.V. All rights reserved.
 
McArdle, R. E., et al. (1961). The Yield of Douglas Fir in the Pacific Northwest, U. S. Department of Agriculture: 74.
McBain, S. and W. Trush (?). The Fluvial Geomorphology of the Tuolumne River: Implications for the Riverine Ecosystem and Salmonid Restoration Arcata, CA: 7.
McCaffery, M., et al. (2007). "Effects of road decommissioning on stream habitat characteristics in the South Fork Flathead River, Montana." Transactions of the American Fisheries Society 136: 553-561.
Previous studies have demonstrated the negative
effects of roads on stream characteristics important for fish
survival; however, few have examined whether decommissioning
reverses these adverse impacts. We examined the
relationships between the percentage of fine sediment in stream
substrate and roads and looked at whether decommissioning
had measurable effects on stream habitat in the Flathead
National Forest, Montana. We conducted habitat surveys and
substrate coring in 12 streams encompassing three watershed
treatment types: (1) roadless areas, (2) areas with roads in use,
and (3) areas with decommissioned roads. Significant positive
correlations were found between the percentage of fine
sediment in substrate and various measures of road impact
(road density, roads in use, and number of stream crossings).
Watersheds with roads in use had higher percentages of fine
sediment than those without roads and those with decommissioned
roads. Watersheds with high levels of vegetative
regrowth on decommissioned roadbeds had a lower percentage
of fines in stream sediment. Decommissioning efforts that
enhance regrowth may improve stream habitat, although
significant effects of these manipulations are difficult to detect
through spatial comparisons. Future studies using either
before–after or before–after–control designs to evaluate the
effects of decommissioning practices on fish and wildlife
habitat and populations are needed.
 
McCammon, B., et al. (1998). A Framework for Analyzing then Hydrologic Condition of Watersheds. U.S. Forest Service.
McCardle, R. E., et al. (1961). The Yield of Douglas Fir in the Pacific Northwest. Washington, D.C., United  States Department of Agriculture,.
McCarthy, E. (1997). "Lake Tahoe: A Watershed Management Study." Western Water: 4-13.
McCarthy, E. J., et al. (1991). "Experimental determination of the hydrologic components of a drained forest watershed." Transactions of the ASAE 34(5): 2031-2039.
McCleary, R. J. and M. A. Hassan (2008). "Predictive modeling and spatial mapping of fish distributions in small stream of the Canadian Rocky Mountain Foothills." Canadian Journal Fisheries and Aquatic Sciences 65: 319-333.
McCleary, R. J. and M. A. Hassan (2008). "Predictive modeling and spatial mapping of fish distributions in small streams of the Canadian Rock Mountain foothills." Canadian Journal Fisheries and Aquatic Sciences 65: 319-333.
McCleary, R. J. and M. A. Hassan (2008). "Predictive modeling and spatial mapping of fish distributions in small streams of the Canadian Rocky Mountain foothills." Canadian Journal Fisheries and Aquatic Sciences 65: 319-333.
McCleary, R. J., et al. (2011). "Spatial organization of process domains in headwater drainage basins of a glaciated foothills region with complex longitudinal profiles." Water Resources Research 47(W05505).
McClintock, K. and J. M. Harbor (1995). "Modeling potential impacts of land development on sediment yields." Physical Geography 16(5): 359-370.
McCold, L. N., and J.W. Saulsbury (1996). "Including past and present impacts in cumulative impact assessments." Environmental Management 20(5): 767-776.
McComb, W. C., et al. (2002). "Models for mapping potential habitat at landscape scales: an example using Northern Spotted Owls." Forest Science 48(2): 203-216.
We are assessing the potential for current and alternative policies in the Oregon Coast
Range to affect habitat capability for a suite of forest resources. We provide an example of a spatially
explicit habitat capability model for northern spotted owls ( Strix occidentalis caurina) to illustrate the
approach we are taking to assess potential changes in habitat capability for vertebrates across the
Coast Range. The model was based on vegetation structure at five spatial scales: the potential nest
tree, a 0.5 ha potential nest patch, 28 ha around a potential nest patch, 212 ha around a potential
nest patch, and a 1,810 ha home range area around a potential nest patch. Sensitivity analyses
indicated that the proportion of the 28 ha patch in large trees around a potential nest patch, and the
number of potential nest trees per ha in the nest patch, had the greatest influence on habitat capability
estimates. The model was verified using georeferenced locations of spotted owl nests from
systematically surveyed areas. Logistic regression analysis indicated that habitat capability scores
were significantly associated with the probability of a site having a nest. Alternative model structures
were tested during verification to test assumptions associated with four variables. The final model
allowed development of a map of habitat capability for spotted owl nesting. The model will be linked
to a model of forest dynamics to project changes in habitat capability under alternative land
management policies.
 
McConnell, R. J. and G. R. Snyder (1972). Key to Field Identification of Anadromous Juvenile Salmonids in Pacific Northwest. Seattle, WA, NOAA: 1-6.
McCuen, R. H., Z. Knight, and A. G. Cutter (2006). "Evaluation of the Nash-Sutcliffe Efficiency Index." Journal of Hydrologic Engineering 11(6): 597-602.
McCutcheon, S. C., M.B. Adams, W. Swank, J.L. Campbell, R.H. Hawkins, and C.R. Dye (2006). Rainfall-runoff relationships for selected eastern U.S. forested mountain watersheds: testing of th ecurve number method for flood analysis, Report prepared for the West Virginia Division of Forestry, Charleston, WV by Environmental and Hydrologic Engineering, 147 Spalding Court, Athens, Georgia 30605. (706) 543 6972; EnvironHyd@aol.com.
McDade, M. H., et al. (1990). "Source distances for coarse woody debris entering small streams in western Oregon and Washington." Canadian Journal of Forest Resources 20: 326-330.
2 copies on file
 
McDade, M. H., et al. (1990). "Source distances for coarse woody debris entering small streams in western Oregon and Washington." Canadian Journal of Forest Resources 20: 326-330.
McDonald, L., et al. (1990). ?
McDonald, L., et al. (1991). Monitoring Guidelines to Evaluate Effects of Forestry Activities on Streams in the Pacific Northwest and Alaska. Seattle, Washington, University of Washington: 166.
McDonnell, J. J. (2003). "Where does water go when it rains? Moving beyond the variable source area concept of rainfall-runoff response." Hydrological Processes 17: 1869-1875.
McDowell, P. F. (2001). Spatial Variations in Channel Morphology at Segment and Reach Scales, Middle Fork John Day River, Northeastern Oregon. Geomorphic Processes and Riverine Habitat. J. M. Dorava, D. R. Montgomery, B. B. Palcsak and F. A. Fitzpatrick. Washington, D.C., American Geophysical Union. 4: 159-172.
McEldowney, R. R., et al. (2002). "Sediment movement and filtration in a riparian meadow following cattle use." Journal of Range Management 55(4): 367-373.
Improper livestock grazing practices in western U.S. riparian areas may reduce the nutrient and pollutant removal function of riparian communities, resulting in degradation of surface water quality. Short duration-high intensity cattle use in 3 x 10 in plots was evaluated in a montane riparian meadow in northern Colorado to quantify livestock effects on sediment movement and filtration under simulated rainfall (approximate to100 mm hour(-1)) plus overland flow (approximate to25 mm hour(-1)) conditions. Four treatments: 1) control, 2) mowed to 10 cm stubble height, 3) trampled by cattle, and 4) cattle grazed plus trampled (grazed) were evaluated. Sixty kg of sediment was introduced to overland flow in each plot. Sediment movement was evaluated using sediment traps positioned in microchannels and on vegetation islands at 5 distances downslope from the upper end of the plots and by sediment front advancement. Most sediment deposition occurred within the first meter downslope from application. About 90% of the applied sediment was filtered from runoff within 10 in in the control and mowed treatments, while approximately 84 and 77% of the applied sediment was trapped in the trampled and grazed treatment plots, respectively. The primary variables that influenced sediment fittration were stem density and surface random roughness. Stem density was the most influential variable that affected sediment filtration. Cattle grazing reduced the stem density by 40%. Monitoring of stem density should aid land managers in regulating cattle use of riparian communities and facilitate the protection of surface water quality from sediment in overland flow.
 
McElhany, P., et al. (2009). Uncertainty in a complex salmon habitat model. Pacific Salmon Environmental and Life History Models, American Fisheries Society Symposium 71. E. E. Knudsen and J. H. Michael. Bethseda, Maryland, American Fisheries Society: 339-356.
McFadden, L. D. and P. L. K. Knuepfer (1990). "Soil geomorphology: the linkage of pedology and surficial processes " Geomorphology 3: 197-205.
McFadden, L. D., et al. (1989). "Use of Multiparameter Relative-Age Methods for Age Estimation and Correlation of Alluvial Fan Surfaces on a Desert Piedmont, Eastern Mojave Desert, California." Quaternary Research 32: 276-290.
McGarigal, K. and S. A. Cushman (2002). The Gradient Concept of Landscape Structure University of Massachusetts: 1-15.
McGlynn, B. L. and J. Seibert (2003). "Distributed assessment of contributing area and riparian buffering along stream networks." Water Resources Research 39(4): doi:10,1029/2002WR001521.
McGuire, K. J., et al. (2005). "The role of topography on catchment-scale water residence time." Water Resources Research 41(W05002).
The age, or residence time, of water is a fundamental descriptor of catchment
hydrology, revealing information about the storage, flow pathways, and source of water
in a single integrated measure. While there has been tremendous recent interest in
residence time estimation to characterize watersheds, there are relatively few studies
that have quantified residence time at the watershed scale, and fewer still that have
extended those results beyond single catchments to larger landscape scales. We
examined topographic controls on residence time for seven catchments (0.085–
62.4 km2) that represent diverse geologic and geomorphic conditions in the western
Cascade Mountains of Oregon. Our primary objective was to determine the dominant
physical controls on catchment-scale water residence time and specifically test the
hypothesis that residence time is related to the size of the basin. Residence times were
estimated by simple convolution models that described the transfer of precipitation
isotopic composition to the stream network. We found that base flow mean residence
times for exponential distributions ranged from 0.8 to 3.3 years. Mean residence
time showed no correlation to basin area (r2 < 0.01) but instead was correlated
(r2 = 0.91) to catchment terrain indices representing the flow path distance and flow
path gradient to the stream network. These results illustrate that landscape organization
(i.e., topography) rather than basin area controls catchment-scale transport. Results
from this study may provide a framework for describing scale-invariant transport
across climatic and geologic conditions, whereby the internal form and structure of the
basin defines the first-order control on base flow residence time.
 
 
McGurk, B. J., and D.R. Fong (1995). "Equivalent Roaded Area as a Measure of Cumulative Effect of Logging." Environmental Management 19(4): 609-621.
McHenry, M. L. (1990). Re: Preliminary Findings, Hoh River Macroinvertebrate Analysis. Hoh Cumulative Effects Group Members, K. Lyons, J. Calhoun and Interested Parties, Northwest Indian Fisheries Commission: ?
McHenry, M. L. and D. C. Morrill (1993). Spawning Gravel Quality, Watershed Characteristics and Egg to Alevin Survival of Coho Salmon and Steelhead in Five North Olympic Peninsula Watersheds (DRAFT), Lower Elwha S'Klallam Fisheries Department.
McHenry, M. L., et al. (1995). Assessment of Physical and Biological Conditions Within the Deep Creek Watershed, North Olympic Peninsula, Washington, and Recommendations for Watershed Restoration, prepared by the Deep Creek Working Group for the Washington Department of Natural Resources.
McHenry, M. L., et al. (1998). "Changes in the quantity and characteristics of large woody debris in streams of the Olympic Peninsula, Washington, U.S.A. (1982-1993)." Canadian Journal of Fisheries and Aquatic Sciences 55(6): 1395-1407.
McHenry, M. L., et al. (1998). "Changes in the quantity and the characteristics of large woody debris in streams of the Olympic Peninsula, Washington, U.S.A. (1982-1993)." Canadian Journal of Fisheries and Aquatic Science 55: 1395-1407.
We assessed the changes in large woody debris (LWD) abundance and composition at 28 sites in 27 low-gradient
Olympic Peninsula streams between 1982 and 1993. The average number of pieces of debris was virtually identical (P = 0.98)
in both years (50.7 versus 50.6). However, we found a significant (P £ 0.01) reduction in the total volume of LWD material in
the stream sites surveyed (51.7 m3×100 m–1 in 1982 to 38.2 m3×100 m–1 in 1993). While the mean volume of second-growth
derived LWD increased from 3.6 to 10.9 m3×100 m–1 (P < 0.01), the increase was insufficient to offset the loss of old-growth
derived LWD. The mean volume of old-growth derived LWD for all sites decreased from 48.1 to 27.4 m3×100 m–1 between
sample years (P < 0.01). The mean diameter of second-growth derived LWD was significantly larger in 1993 than in 1982,
although still smaller than old-growth derived pieces. We measured a significant increase in the percentage of LWD pieces
rated as highly decayed from 1982 to 1993. The results indicate that the loss of old-growth derived LWD following the
removal of old-growth riparian forests is initially very rapid, followed by a slower rate of depletion associated with watershed
destabilization. Inputs of LWD from second-growth riparian forests up to 73 years old were characterized by small diameter,
high mobility, and high decay rates.
 
McHugh, P. and P. Budy (2004). "Patterns of spawning habitat selection and suitability for two populations of spring Chinook salmon, with an evaluation of generic versus sit-specific suitability criteria." Transactions of the American Fisheries Society 133: 89-97.
We evaluated patterns of redd site selection in relation to physical habitat variables
(depth, velocity, and gravel size) using logistic regression and developed spawning habitat suitability
models for two populations of spring Chinook salmon Oncorhynchus tshawytscha in Idaho.
Additionally, we evaluated the validity of published, generic spawning habitat suitability criteria
relative to our stream-specific models. In Elk Creek, fish used sites with coarse gravel sizes and
shallow water depths; there was no difference in velocity between used and unused sites in this
stream. Salmon spawning in Sulphur Creek used sites with deeper and faster water irrespective
of gravel size. Spawning habitat suitability in Elk Creek was best modeled as a quadratic function
of gravel size; in Sulphur Creek, it was best modeled as a positive function of depth alone. The
best model fit to a pooled data set was a hybrid of the two single-stream models. In all cases,
models containing all of the habitat variables performed worse than the simplest models. Generic
criteria predicted site use poorly in Elk Creek but exceptionally well in Sulphur Creek: in Elk
Creek nearly half of all sampled sites were misclassified, while in Sulphur Creek nearly 75% were
correctly classified. As river-specific models are likely to outperform generic criteria in most cases,
we recommend that researchers carefully weigh the costs of parameterizing such models relative
to the costs of the potential misclassification errors from using preexisting generic models.
 
 
McIntyre, L. (1998). "Complexity: A Philosopher's Reflections." Complexity 3(6): 26-32.
McKean, J. and J. Roering (2004). "Objective landslide detection and surface morphology mapping using high-resolution airborne laser altimetry." Geomorphology 57: 331-351.
A map of extant slope failures is the most basic element of any landslide assessment. Without an accurate inventory of slope
instability, it is not possible to analyze the controls on the spatial and temporal patterns of mass movement or the environmental,
human, or geomorphic consequences of slides. Landslide inventory maps are tedious to compile, difficult to make in vegetated
terrain using conventional techniques, and tend to be subjective. In addition, most landslide inventories simply outline landslide
boundaries and do not offer information about landslide mechanics as manifested by internal deformation features. In an
alternative approach, we constructed accurate, high-resolution DEMs from airborne laser altimetry (LIDAR) data to characterize
a large landslide complex and surrounding terrain near Christchurch, New Zealand. One-dimensional, circular (2-D) and
spherical (3-D) statistics are used to map the local topographic roughness in the DEMs over a spatial scale of 1.5 to 10 m. The
bedrock landslide is rougher than adjacent unfailed terrain and any of the statistics can be employed to automatically detect and
map the overall slide complex. Furthermore, statistics that include a measure of the local variability of aspect successfully
delineate four kinematic units within the gently sloping lower half of the slide. Features with a minimum size of surface folds
that have a wavelength of about 11 to 12 m and amplitude of about 1 m are readily mapped. Two adjacent earthflows within the
landslide complex are distinguished by a contrast in median roughness, and texture and continuity of roughness elements. The
less active of the earthflows has a surface morphology that presumably has been smoothed by surface processes. The Laplacian
operator also accurately maps the kinematic units and the folds and longitudinal levees within and at the margins of the units.
Finally, two-dimensional power spectra analyses are used to quantify how roughness varies with length scale. These results
indicate that no dominant length scale of roughness exists for smooth, unfailed terrain. In contrast, zones with different styles of
landslide deformation exhibit distinctive spectral peaks that correspond to the scale of deformation features, such as the
compression folds. The topographic-based analyses described here may be used to objectively delineate landslide features,
generate mechanical inferences about landslide behavior, and evaluate relatively the recent activity of slides
 
McKean, J. and W. Wright (2005). "Mapping channel morphology and stream habitat with a full waveform green lidar." EOS Transactions, Fall Meeting Suppliment 86(52): Abstract H34B-05.
Basic topographic mapping and objective description of physical habitat in channels remains problematic using standard methods of cross section and thalweg surveying. Even with ground GPS surveys, detailed maps of bedforms and banks are normally only attempted for short stream reaches on the order of 100 m. The Experimental Advanced Airborne Research Lidar (EAARL) is a full waveform lidar, operating at 532 nm, with the potential to seamlessly map riparian vegetation, floodplain topography, and channel banks and bedforms over whole stream networks. We have used EAARL to map channel topography continuously over 85 km of streams in the watershed of the Middle Fork Salmon River, Idaho. The data are being compared to control field surveys in six reaches of plane bed and pool/riffle channels to assess their accuracy. In 15- to 50-meter-wide channels, EAARL data support fundamental hydrologic measures such as bankfull depth, pool and riffle definition, residual pool volumes, channel cross-sectional area and slope, bedform hydraulic roughness, and water turbidity. We are also using EAARL data to classify channels based on their morphometry and inferred processes and to objectively define the physical habitat for Chinook salmon and other species from derived metrics such as available spawning surface area, bedform amplitude, and locations, dimensions, and hydrologic connection of off-channel rearing areas. The continuous nature of the data allows us to study network scale habitat connectivity and explore controls on the spatio-temporal distribution of whole salmon populations. Continuous thalweg profiles have been interpreted in the frequency domain using wavelet analyses to investigate periodicity of channel topography at a variety of spatial scales. The wavelet analyses accurately map transitions in channel types and in one stream segment closely define historic preferred spawning sites. Future work is planned to exploit the three-dimensional bed topography mapped by EAARL and refine local predictions of bed shear stress and bed mobility, substrate median grain size, and hyporheic exchange over whole stream networks.
 
McKean, J. A., et al. (2008). "Geomorphic controls on salmon nesting patterns described by a new, narrow-beam terrestrial-aquatic lidar." Frontiers in Ecology and the Environment 6(3): 125-130.
Riverine aquatic biodiversity is rapidly being lost worldwide, but preservation efforts are hampered, in part
because studies of these dynamic environments are limited by cost and logistics to small local surveys. Full
understanding of stream ecosystems requires precise, high-resolution mapping of entire stream networks and
adjacent landforms. We use a narrow-beam, water-penetrating, green lidar system to continuously map 10 km
of a mountain stream channel, including its floodplain topography, and wavelet analyses to investigate spatial
patterns of channel morphology and salmon spawning. Results suggest the broadest fluvial domains are a
legacy of approximately 15 000 years of post-glacial valley evolution and that local pool–riffle channel topography
is controlled by contemporary hydraulics operating on this broad template. Salmon spawning patterns
closely reflect these hierarchical physical domains, demonstrating how geomorphic history can influence modern
distributions of aquatic habitat and organisms. The new terrestrial–aquatic lidar could catalyze rapid
advances in understanding, managing, and monitoring of valuable aquatic ecosystems through unprecedented
mapping and attendant analyses.
 
McKee, A., et al. (1987). Research Publications of the H. J. Experimental Forest, Cascade Range, Oregon, 1948 to 1986. Corvallis, OR, U.S. Department of Agriculture, Forest Service: 74.
McKenzie, D., et al. (1996). "Extrapolation problems in modeling fire effects at large spatial scales: A review." International Journal of Wildland Fire 6(4): 165-176.
McKerchar, A. I., et al. (1998). "Data for Ashley River to test channel network and river basin heterogeneity concepts." Water Resources Research 34(1): 139-142.
McKinley, M. (1997). Large Woody Debris Source Distances for Western Washington Cascade Streams. Seattle, Washington, University of Washington: 36.
McMaster, K. J. (2002). "Effects of digital elevation model resolution on derived stream network positions." Water Resources Research 38(4).
This study examines the effect of digital elevation model (DEM) resolution on the positional
accuracy of derived hydrologic networks and quantitatively confirms that DEM resolution should
be greater than the average hillslope length when used for hydrologic modeling. Seven hundred
kilometers of mapped stream networks are compared to stream networks derived from 17 DEMs
with resolutions ranging from 30 m to 3 km. Comparison between predicted and mapped streams
reveals that accuracy of predicted stream locations decays quickly beyond a DEM resolution of
180 m. A new application of Gyasi-Agyei et al.’s [1995] DEM resolution suitability test based on
average slope and vertical resolution indicates an average hillslope length of between 150 and 180 m.
Tarboton et al.’s [1991] method of determining average hillslope length based on slope and
accumulation areas reconfirms the length to be 150 m and verifies the link between network accuracy
and hillslope scale. Two algorithms are used to derive stream networks: the D 1 algorithm [Tarboton,
1997], which allows for flow dispersion, and the D8 algorithm [O’Callaghan and Mark, 1984],
which does not. A comparison between the D8 and D 1 algorithms shows that modeling flow
dispersion is not necessary in steep terrain, as both algorithms performed equally well. However, the
D 1 algorithm is found to be less susceptible to modeling erroneous hillside flow convergence,
making it preferable to the D8 algorithm when used by Tarboton et al.’s [1991] method of
determining the average hillslope lengths. Finally, field observation indicates that predicted first- and
second-order streams tend to exist when predicted; however, mapped streams have a higher
positional accuracy. INDEX TERMS: 1860 Hydrology: Runoff and streamflow; 3210
Mathematical Geophysics: Modeling; 1894 Hydrology: Instruments and techniques; KEYWORDS:
streamflow, terrain modeling, digital elevation model (DEM), resolution, hydrologic networks,
hillslope length
 
McMichael, G. A., et al. (1997). "Effects of Residual Hatchery-Reared Steelhead on Growth of Wild Rainbow Trout and Spring Chinook Salmon " Transactions of the American Fisheries Society 126: 230-239.
McNabb, D. H. and F. J. Swanson (1990). Effects on Fire on Soil Erosion. Natural and Prescribed Fire in Pacific Northwest Forests. J. D. Walstad. Corvallis, Oregon, Oregon State University Press. Chapter 14: 159-176.
McNally, R., et al. (2002). "Current Loads of Coarse Woody Debris on Southeastern Australian Floodplains: Evaluation of Change and Implications for Restoration." Restoration Ecology 10: 627-635.
McNamara, J. P., et al. (2005). "Soil moisture states, lateral flow, and streamflow generation in a semi-arid, snowmelt-driven catchment." Hydrological Processes 19(20): 4023-4038.
Hydraulic connectivity on hillslopes and the existence of preferred soil moisture states in a catchment have important controls on runoff generation. In this study we investigate the relationships between soil moisture patterns, lateral hillslope flow, and streamflow generation in a semi-arid, snowmelt-driven catchment. We identify five soil moisture conditions that occur during a year and present a conceptual model based on field studies and computer simulations of how streamflow is generated with respect to the soil moisture conditions. The five soil moisture conditions are (1) a summer dry period, (2) a transitional fall wetting period, (3) a winter wet, low-flux period, (4) a spring wet, high-flux period, and (5) a transitional late-spring drying period. Transitions between the periods are driven by changes in the water balance between rain, snow, snowmelt and evapotranspiration. Low rates of water input to the soil during the winter allow dry soil regions to persist at the soil-bedrock interface, which act as barriers to lateral flow. Once the dry-soil flow barriers are wetted, whole-slope hydraulic connectivity is established, lateral flow can occur, and upland soils are in direct connection with the near-stream soil moisture. This whole-slope connectivity can alter near-stream hydraulics and modify the delivery of water, pressure, and solutes to the stream.
 
McNamara, J. P., et al. (2006). "Channel head locations with respect to geomorphologic thresholds derived from a digital elevation model: a case study in northern Thailand." Forest Ecology and Management 224: 147-156.
The shape of a catchment is controlled by the interplay of different erosion processes acting within the catchment. It is therefore possible to
assess dominant erosion processes, and geomorphologic thresholds that spatially separate those processes, by evaluating catchment form. In this
paper, geomorphologic thresholds are detected in a digital elevation model of the Pang Khum Experimental Watershed in northern Thailand and
compared to the locations of field mapped channel heads. The intersection of thresholds in the slope–area relationship, the probability distribution
of drainage areas, and the probability distribution of energy index produce distinct domains in slope–area space that partition the landscape
according to erosion mechanisms. All mapped channel heads plot higher than an energy threshold defined by the product of slope and the square
root of drainage area. Above this threshold different types of channel heads are partitioned by independent slope or drainage area thresholds. For
example, channel heads formed at groundwater seeps plot higher than a drainage area threshold, independent of slope. Channel heads that originate
from landslides and overland flow erosion plot higher than a slope threshold, independent of drainage area. It is our interpretation that the channel
heads that did not initiate at groundwater seeps were affected by human disturbance (forest conversion for swidden-based agriculture), as they tend
to lay above seeps on highly disturbed hillslopes. This paper explores relationships between the shape of a catchment as defined by a digital
elevation model and the distribution of mapped channel heads. These relationships serve as a first-order means to identify locations of potentially
unstable areas in a landscape, thereby providing a basis to assess the potential impacts of future catchment disturbances.
 
McNeil, R. C. and D. B. Zobel (1980). "Vegetation and Fire History of a Ponderosa Pine-White Fir Forest in Crater Lake National Park." Northwest Science 54(1): 30-45.
McNulty, S. G., L. Swift, J. Hays, and A. Clingenpeel (1995). Predicting watershed erosionproduction and over-land sediment transport using a GIS., In: Carrying the torch for erosion control: an olympic task. Proceedings, XXVI, International Erosion Control Association Conference, Atlanta, GA.
McNulty, S. G., and W.T. Swank (1996). Forest ecosystem analysis using a GIS. Eco-Informa '96: Global networks for environmental information, Lake Buena Vista, Florida.
Meade, R. H. (1985). "Wavelike Movement of Bedload Sediment, East Fork River, Wyoming." Environ. Geol. Water Sci. 7: 215-225.
Meade, R. H., et al. (1990). Movement and storage of sediment in rivers of the United States and Canada Surface water hydrology: Boulder, Colorado. M. G. Wolman and H. C. Riggs. Chapter 11: 225-280.
Means, J., et al. (1995). "Past and present disturbance regimes, biodiversity and ecosystem dynamics in the Orgeon Coast Range (DRAFT)." 13.
Mears, A. I. (1979). "Flooding and sediment transport in a small alpine drainage basin in Colorado." Geology 7(1): 53-57.
Megahan, W. F., and Kidd, W. J. Jr. (1972). "Effects of logging and logging roads on erosion and sediment deposition from steep terrain." Journal of Forestry 70(3): 136-141.
Megahan, W. F. (1982). Channel sediment storage behind obstructions in forested drainage basins draining the granitic bedrock of the Idaho Batholith. Portland, OR, Pacific Northwest Forest and Range Experiment Station.
Megahan, W. F. (1982). Channel sediment storage behind obstructions in forested drainage basins draining the granitic bedrock of the Idaho Batholith. Sediment budgets and routing in forested drainage basins. R. J. J. Swanson F.J., T. Dunne, D.N. Swanston, USDA Forest Service. PNW-GTR-141.
Megahan, W. F. (1983). "Hydrologic Effects of Clearcutting and Wildfire on Steep Granitic Slopes in Idaho." Water Resources Research 19(3): 811-819.
Megahan, W. F. (2000). Cumulative watershed effects research needs for forested watersheds in the 21st century. Water Resources Center Report - Centers for Water and Wildland Resources, Report: 98. C. W. Slaughter: 61-68.
Key cumulative effects research needs dealing with forest watersheds include natural variability in watershed characteristics and in downstream responses, sediment routing, and effects of disturbance on streamflow. Important information needs relating to natural variability include better descriptions of spatial variability and a broader perspective on temporal variability. Large regional differences in landslide types and streamflow rates are used to illustrate the need for studies of spatial variability. Justification for increased research on temporal variability is based on recent studies documenting long-term (1000s of years) sediment supplies from mountain watersheds in Idaho averaging about an order of magnitude greater than present day (10s of years) sediment data indicate. Additional studies in western Oregon show that mountain channels exhibit a natural succession from energy limited (aggraded) to supply limited (degraded) conditions, with accompanying changes in aquatic habitat conditions. Important components of sediment routing research include delivery of landslide material to channels, downslope sediment from roads to channels, and downstream routing of bedload sediments in channels. Evaluation of the effects of disturbance on streamflow should include the effects of forest management and wildfire on all levels of streamflow, and the accompanying erosional and sedimentation response of channels and aquatic ecosystems. Physically-based, distributed models need to be developed and improved to predict effects of disturbance on streamflow and channel changes.
 
Megahan, W. F. (2000). Lessons Learned in Watershed Management: A Retrospective View, USDA Forest Service: 177-188.
Megahan, W. F. and C. C. Bohn (1989). Progressive, Long Term Slope Failure Following Road Construction and Logging on Noncohesive, Granitic Soils of the Idaho Batholith. Symposium on Headwaters Hydrology, American Water Resources Association.
Megahan, W. F. and G. L. Ketcheson (1996). "Predicting downslope travel of granitic sediments from forest roads in Idaho." Water Resources Bulletin 32(2): 371-382.
Megahan, W. F. and D. C. Molitor (1975). "Erosinoal effects of wildfire and logging in Idaho." ASCE, Journal Irrigation and Drainage Division: 423-444.
Megahan, W. F. and R. A. Nowlin (1976). Sediment storage in channels draining small forested watersheds in the mountains of central Idaho. Proceedings, Third Federal Interagency Sedimentation Conference, Water Resources Council, Washington, D.C.
Meier, M. F. (?). "Global ice volume and sea level duiring the coming century."? ?
Meleason, M. A. (2001). A Simulation Model of Wood Dynamics in Pacific Northwest Streams. Corvallis, OR, Oregon State University. PhD.
Meleason, M. A., et al. (2008). "Characterizing the variability of wood in streams: simulation modelling compared with multiple-reach surveys." Earth Surface Processes and Landforms 32: 1164-1173.
Wood abundance in streams is an indicator of its likely geomorphological and ecological
importance. However, wood volume estimates can be highly variable, due in part to natural
variability and the methodology used to characterize it. We measured wood volume in streams
of similar sizes and riparian forest conditions using extended field surveys and simulation
modelling. We surveyed a total of 3·1 km along four tributary streams of New Zealand’s
Waihaha River to obtain an estimate of wood volume in streams with similar basin positions
(second order), forest types (podocarp/hardwood forest), disturbance histories (post volcanic
eruption of Taupo ca. 180 AD) and stream sizes (2–3 m bankfull width). A ‘sliding window’
analysis was conducted whereby the wood volume was calculated for a ‘window’ (i.e., reach
survey of fixed length) that was progressively moved upstream in 10 m increments. The resulting
frequency distributions of wood volume were bimodal and represented the range and
relative proportion of reach volumes possible from the wood surveys. The wood volume based
on all streams surveyed (23 m3/100 m) was equivalent to the 64th percentile of the sliding window
distribution, suggesting that a randomly placed study reach would be likely to underestimate
wood volume. The bimodal distribution was attributed to the inclusion/exclusion of relatively
large (≥10 m3), but rare (0·3 logs/100 m) logs. We also examined the variability of
reach-level wood volume estimates (200 m and 400 m) for the Waihaha tributaries using the
model OSU StreamWood. The volume frequency distributions from the simulations were
similar to those from the empirical approach, except that they were unimodal. We attribute the
unimodal distribution to the greater number of reach-scale estimates used in the simulations
(n = 2000). The two independent approaches characterized the variation of wood volumes
possible for this forest type and stream width.
 
Meleason, M. A., et al. (2002). Simulation of stream wood source distance for small streams in the western Cascades, Oregon, USDA Forest Service: 457-466.
The model, STREAMWOOD, is an individual-based stochastic model designed to simulate
the dynamics of wood in small streams of the Pacific Northwest. We used STREAMWOOD
to examine source distance as a function of tree fall regime and stand age. Our results suggest
that source distance increased with stand age for the first 400 years of stand development and
then declined. Simulated source distance for mature conifer forests (81 to 200 years old) were
consistent with observed data, but simulated source distances for old-growth forests (201 to
1,000 years old) were below observed data. Further information on stand ages for the forests
used in the observational study would refine the compassion with simulated data.
 
Meleason, M. A., et al. (2003). "Implications of Riparian Management Strategies on Wood in Streams of the Pacific Northwest." Ecological Applications 13(5): 1212-1221.
Meleason, M. A. and G. M. J. Hall (2005). "Managing plantation forests to provide short- to long=term supplies of wood to streams: a simulation study using New Zealand's pine plantations." Environmental Management 36(2): 258-271.
ABSTRACT / Riparian functions such as the recruitment of
wood to streams take decades to recover after a clear-fell
harvest to the stream edge. The implications of two sets of
riparian management scenarios on the short- and long-term
recruitment of wood to a hypothetical stream (central North
Island, New Zealand) were compared through simulation
modeling. In the first set (native forest buffer), a designated
treeless riparian buffer was colonized by native forest species
after a pine crop (Pinus radiata) had been harvested to
the stream bank. In the second set (pine to native forest
buffer), native forest species were allowed to establish under
the pine canopy in a designated riparian buffer. In general,
the volume of wood was greater in streams with wider buffers
(5-m to 50-m) and this effect increased with forest age (800
years). The pine to native forest buffer supplied more wood
to the stream more quickly, and matched the long-term
supply to the stream from the native forest buffer. For the
native forest buffer, total wood volume was minimal for the
first 70 years and then increased uniformly for the remainder
of the simulation. In contrast, the pine to native forest buffer
produced a bimodal response in total wood volume with the
initial sharp peak at year 100 attributed to pine recruitment
and a second more gradual peak lasting for the rest of the
simulation, which was similar to levels in the native forest
simulations. These results suggest that existing plantations
could be an important source of wood to the stream during
the first 100+ years of native forest development.
 
Melelli, L. and A. Taramelli (2004). "An example of debris-flows hazard modeling using GIS." Natural Hazards and Earth System Sciences 4: 347-358.
We present a GIS-based model for predicting
debris-flows occurrence. The availability of two different
digital datasets and the use of a Digital Elevation Model (at
a given scale) have greatly enhanced our ability to quantify
and to analyse the topography in relation to debris-flows. In
particular, analysing the relationship between debris-flows
and the various causative factors provides new understanding
of the mechanisms. We studied the contact zone between
the calcareous basement and the fluvial-lacustrine infill adjacent
northern area of the Terni basin (Umbria, Italy), and
identified eleven basins and corresponding alluvial fans. We
suggest that accumulations of colluvium in topographic hollows,
whatever the sources might be, should be considered
potential debris-flow source areas. In order to develop a susceptibility
map for the entire area, an index was calculated
from the number of initiation locations in each causative factor
unit divided by the areal extent of that unit within the
study area. This index identifies those units that produce the
most debris-flows in each Representative Elementary Area
(REA). Finally, the results are presented with the advantages
and the disadvantages of the approach, and the need for further
research.
 
Melesse, A. M., and W. D. Graham (2005). "Storm runoff prediction based on a spatially distributed travel time method utilizing remote-sensing and GIS." Journal of the American Water Resources Association 40(4): 863-879.
Melis, T. S., et al. (1995). Magnitude and frequency data for historic debris flows in Grand Canyon National Park and Vicinity, Arizona, U. S. Geological Survey: 224.
Melosh, H. J. (1983). "Acoustic Fluidization." American Scientist 71: 158-165.
Merigliano, M. F. (1996). Ecology and Management of the South Fork Snake River Cottonwood Forest Missoula, Montana, University of Monana: 79.
Merritt, W. S., et al. (2003). "A review of erosion and sediment transport models." Environmental Modelling & Software 18(8-9): 761-799.
Merritts, D. J. (1996). "The Mendocino triple junction: active faults, coastal emergence and rapid uplift." Journal of Geophysical Research 101(B3): 6051-6070.
Merritts, D. J., et al. (1994). "Long river profiles, tectonism, and eustasy: A guide to interpreting fluvial terraces." Journal of Geophysical Research 99(B7): 14031-14050.
Mertz, J. E., et al. (2005). "Sediment budget for salmonid spawning habitat rehabilitation in a regulated river." Geomorphology 76: 207-228.
Bed elevation, feature adjustments, and spawning use were monitored at three Chinook salmon (Oncorhynchus tschawytscha) spawning habitat rehabilitation sites to measure project longevity in a regulated river. Sites enhanced with 649–1323 m3 of gravel lost from 3–20% of remaining gravel volume annually during controlled flows of 8–70 m3/s and 2.6–4.6% of placed material during a short-duration (19 days) release of 57 m3/s. The oldest site lost 50% of enhancement volume over 4 years. Of the mechanisms monitored, gravel deflation was the greatest contributor to volumetric reductions, followed by hydraulic scour. Spawning, local scour around placed features, and oversteepened slopes contributed to volumetric changes. As sites matured, volumetric reductions decreased. Sites captured as much large woody debris as was lost. While complexity is an extremely important aspect of ecological function, artificial production of highly diverse and complex habitat features may lead to limited longevity without natural rejuvenation.
 
Merz, W. and R. B. Bryan (1993). "Critical Conditions for Rill Initiation on Sandy Loam Brunisols - Laboratory and Field Experiments in Southern Ontario, Canada." Geoderma 57(4): 357-385.
In rainfall simulation experiments on 0.8 m x 10 m plots with agricultural soils, knickpoint development and incipient rill erosion was associated with distinct ranges of flow intensity, expressed by shear velocity, unit stream power, and unit discharge. Critical threshold shear velocity and unit stream power values were considerably higher than those identified in other studies, probably due to differences in experimental methodology. Rill initiation was found to be a function of local flow hydraulic conditions and independent of total plot discharge. This complicates prediction of this process for effective modelling as some index which links soil microtopography and flow concentration must also be incorporated in addition to runoff volume.Rill initiation could not be successfully linked to soil shear strength variations measured by a torsional shear vane. The vane shear strength values measured cannot be linked sufficiently precisely to a thin surface soil layer to provide a reliable indicator of rill and interrill erodibility.
 
Metailie, J.-P. (1987). The Degradation of the Pyrenees in the Nineteenth Century - An Erosion Crisis? International Geomorphology. V. Gardiner. II: 533-545.
Meyer, C. R., and D. C. Flanagan (1991). Applying case-based reasoning techniques to the WEPP soil erosion model. 1991 International winter meeting, ASAE, Chicago, IL, ASAE.
Meyer, G. A. and M. E. Leidecker (1999). Fluvial terraces along the Middle
Fork Salmon River, Idaho, and their relation to glaciation, landslide dams,
and incision rates: A preliminary analysis and river-mile guide. Guidebook to the Geology of Eastern Idaho. S. S. Hughes and G. D. Thackray. Pocatello, Idaho Museum of Natural History: 219-235.
Meyer, G. A. and J. L. Pierce (2003). "Climatic controls on fire-induced sediment pulses in Yellowstone National Park and Central Idaho: A long-term perspective." Forest Ecology and Management 178: 89-104.
Meyer, G. A., et al. (2001). "Fire, storms, and erosional events in the Idaho batholith." Hydrological Processes 15: 3025-3038.
In late December 1996, the South Fork Payette River basin in west-central Idaho experienced a prolonged storm that culminated on January 1, 1997, with intense rain on melting snow that triggered slide failures, producing debris flows and sediment-charged floods. Failures occurred in saturated, cohesionless, grussy colluvium derived from weathered Idaho batholith granitic rocks. Many failures along the South Fork Payette River originated in ponderosa pine forests burned in the 1989 stand-replacing Lowman fire. An example is the 0 times 49 km super(2) 'Jughead' Creek basin, where a single large colluvial failure produced almost 40% of the total volume eroded from the basin and generated a massive and rapid debris flow. Failures also occurred in steep, unburned, and unforested drainages such as Hopkins Creek. In this south-facing 0 times 58 km super(2) basin, 15 colluvial hollows failed, but no single failure produced more than 10% of the total eroded volume. Sediment transport in Hopkins Creek occurred by prolonged sediment-charged sheetflooding. Despite vegetation differences, sediment yields from the geomorphically similar Hopkins Creek ( similar to 42 000 Mg km super(-2)) and Jughead Creek ( similar to 44 000 Mg km super(-2)) basins were quite similar. These 1997 erosion events are equivalent to several thousand years of sediment yield at low rates (2 times 7-30 Mg km super(2) year super(-1)) measured by short-term sediment trapping and gauging in Idaho batholith watersheds. If similar large events were solely responsible for sediment export, recurrence intervals (RIs) of several hundred years would account for higher sediment yields averaged over similar to 10 super(4) year from Idaho batholith watersheds. Dating of small fire-induced sheetflooding events in an early Holocene tributary junction fan of Jughead Creek indicates that frequent small sedimentation events (RI approximately 33-80 year) occurred between 7400 and 6600 cal year BP, with an average yield not greatly exceeding 16 Mg km super(-2) year super(-1). Compared with the Holocene average, erosion rates during that 800 year period were unusually low, suggesting that sediment yields have not been constant over time, and that climatic variations and related fire regime changes may exert a strong influence on the probability of major erosional events.
 
Meyer, G. A., et al. (2001). "Fires, storms, and sediment yield in the Idaho batholith." Hydrological Processes 15: 3025-3038.
Meyer, G. A., et al. (1995). "Fire and alluvial chronology in Yellowstone National Park: Climatic and intrinsic controls on Holocene geomorphic processes." GSA Bulletin 107(10): 1211-1230.
Meyer, G. A. and S. G. Wells (1997). "Fire-related sedimentation events on alluvial fans, Yellowstone National Park, U.S.A." Journal of Sedimentary Research 67(5): 776-791.
Meyer, G. A., et al. (1992). "Response of alluvial systems to fire and climate change in Yellowstone, National Park." Nature 357(14): 147-150.
Meyer, G. A., et al. (1995). "Fire and alluvial chronology in Yellowstone National Park: Climatic and intrinsic controls on Holocene geomorphic processes." Geological Society of America Bulletin 107(10): 1211-1230.
Meyer, J. L., et al. (1988). "Elemental dynamics in streams." Journal of the North American Benthological Society.
Meyer, J. L. and J. B. Wallace (2001). Lost linkages and lotic ecology: rediscovering small streams. Ecology: Achievement and Challenge. N. H. a. S. L. M. Press. Oxford, UK, Blackwell Science. Chp. 14: 295-317.
Michael, D. (1998). Oregon Department of Forestry's Use of Geotechnical Specialists. Environmental, Groundwater and Engineering Geology: Applications from Oregon. S. Burns, Star Publishing Company: 493-496.
Michael, D. L. (1997). "Landslides From Forest Lands in Oregon A Proposal of Risk Based Management", ?
Michael, D. L. (1998). Landslide Lessons Learned From Recent Winters in Oregon Forests. Forest Grove, OR, Oregon Department of Forestry: 7.
Micheli, E. R., et al. (2003). "Quantifying the effect of riparian forest versus agricultural vegetation on river meandering migration rates, Central Sacramento River, California, USA." River Research and Applications 19: 1 - 12.
Micheli, E. R., et al. (2004). "Quantifying the effect of riparian forest versus agricultural vegetation on river meander migration rates, central Sacramento River, California, USA." River Research and Applications 20: 537-548.
Riparian forest vegetation is widely believed to protect riverbanks from erosion, but few studies have quantified the effect of
riparian vegetation removal on rates of river channel migration. Measured historical changes in a river channel centreline, combined
with mapped changes in floodplain vegetation, provide an opportunity to test how riparian vegetation cover affects the
erodibility of riverbanks. We analysed meander migration patterns from 1896 to 1997 for the central reach of the Sacramento
River between Red Bluff and Colusa, using channel planform and vegetation cover data compiled from maps and aerial photography.
We used a numerical model of meander migration to back-calculate local values for bank erodibility (i.e. the susceptibility
of bank materials to erosion via lateral channel migration, normalized for variations in near-bank flow velocities due to
channel curvature). A comparison of migration rates for approximately 50 years before and after the construction of Shasta dam
suggests that bank migration rates and erodibility increased roughly 50%, despite significant flow regulation, as riparian floodplains
were progressively converted to agriculture. A comparison of migration rates and bank erodibilities between 1949 and
1997, for reaches bordered by riparian forest versus agriculture, shows that agricultural floodplains are 80 to 150% more erodible
than riparian forest floodplains. An improved understanding of the effect of floodplain vegetation on river channel migration
will aid efforts to predict future patterns of meander migration for different river management and restoration scenarios.
 
Middleton, G. V. and P. R. Wilcock (1994). Mechanics in the Earth and Environmental Sciences. Cambridge, Cambridge University Press.
Mikos, M. (1995). "Fluvial abrasion: converting size reduction coefficients into weight reduction rates." Journal of Sedimentary Research A65(3): 472-476.
Miles, M. J. and R. Kellerhals (1981). Some engineering aspects of debris torrents. 5th Canadian Hydrotechnical Conference, Frederiction, New Brunswick.
Milhouse, R. T. (1973). The movement of individual particles in a gravel bottom stream. Unpublished paper given at the Northwest Regional American Geophysical Union meetings, Vancouver.
Milhouse, R. T. (1973). Sediment transport in a gravel-bottomed stream, Oregon State University.
Millar, R. G. (1999). "Grain and form resistance in gravel-bed rivers " Journal of Hydraulic Research 37: 303-312.
Millar, R. G. and C. D. Rennie (2001). "Comment on "Effects of hydraulic roughness on surface textures of gravel-bed rivers" by John M. Vuffington and David r. Montgomery." Water Resources Research 37(5): 1527-1528.
Millard, T. (1999). Debris flow initiation in coastal British Columbia gullies. Nanaimo, BC, Canada, British Columbia Ministry of Forests.
Millard, T., et al. (2002). Post-logging landslide rates in the Cascade Mountains, southwestern British Columbia. Nanaimo, BC, Canada, British Columbia Ministry of Forests.
Millard, T., et al. (2002). Post-logging landslide rates in the Cascade Mountains, southwestern British Columbia, Vancouver Forest Region, BC Ministry of Forests, Forest Research: 1-18.
Miller, A. J. (1990). "Fluvial response to debris associated with mass wasting during extreme floods." Geology 18: 599-602.
Miller, B. G. N. (2000). Two landslides and their dams, Peace River Lowlands, Alberta. Department of Earth and Atmospheric Sciences. Edmonton, University of Alberta: 131.
Miller, B. G. N. and D. M. Cruden (2002). "The Eureka River landslide and dam, Peace River Lowlands, Alberta." Canadian Geotechnical Journal 39(4): 863-878.
Miller, D. and L. Benda (1998). "Effects of mass wasting on channel morphology and sediment transport: South Fork Gate Creek, Oregon." Technical Bulletin No. XXX, NCASI.
Miller, D. and L. Benda (2000). "Effects of mass wasting on channel morphology and sediment transport: South Fork Gate Creek, Oregon." Bulletin of the Geological Society of America: 1814-1824.
MIller, D. and L. Benda (submitted). "Consequences of routing a punctuated sediment supply through branching networks." submitted to Water Resources Research.
Miller, D., et al. (2008). Wood recruitment inventory and analysis. Final report to Coeur d'Alene Tribe Fisheries Program. Plummer, Idaho, Coeur d'Alene Tribe Fisheries
Miller, D. C. (1985). "Holocene eruptions at the Inyo volcanis chain, California: Implications for possible eruptions in Long Valley, Caldera " Geology 13: 14-17.
Miller, D. J. "Damage in King County from the Storm of January 9, 1990." Washington Geology 19(1): 28-37.
Miller, D. J. (1960). Giant Waves in Lituya Bay Alaska, U.S. Department of the Interior: 51-81.
Miller, D. J. (1995). "Coupling GIS with Physical Models to Assess Deep-Seated Landslide Hazards." Environmental & Engineering Geoscience 1(3): 263-276.
Miller, D. J., and L.E. Benda (2000). "Effects of punctuated sediment supply on valley-floor landforms and sediment transport." GSA Bulletin 112(12): 1814-1824.
Miller, D. J. (2003). Programs for DEM Analysis. Landscape Dynamics and Forest Management, General Technical Report RMRS-GTR-101CD. Fort Collins, CO, USA, USDA Forest Service, Rocky Mountain Research Station.
Miller, D. J. (2003). Programs for DEM Analysis, available at http://www.fsl.orst.edu/clams/prj_wtr_millerprg.html. Seattle, WA, Earth Systems Institute: 38.
Miller, D. J. (2004). Landslide Hazards in the Stillaguamish Basin: A New Set of GIS Tools, The Stillaguamish Tribe of Indians: 50.
Miller, D. J. and L. E. Benda (1996). Sediment Routing in Steep Forested Mountain Basins, May 9th and 10th, 1996, National Council for Air and Stream Improvement: 1-81.
Miller, D. J. and L. E. Benda (2000). "Effects of punctuated sediment supply on valley-floor landforms and sediment transport." Geological Society of America Bulletin 112(12): 1814-1824.
Miller, D. J. and K. M. Burnett (2007). "Effects of forest cover, topography, and sampling extent on the measured  density of shallow, translational landslides." Water  Resources Research 43: 1-23.
Miller, D. J. and K. M. Burnett (2007). "A probabilistic model of debris-flow delivery to stream channels, demonstrated for the Coast Range of Oregon, USA." Geomorphology.
Miller, D. J. and K. M. Burnett (2007). "A probabilistic model of debris-flow delivery to stream channels, demonstrated for the Oregon Coast Range, USA." Geomorphology 94: 184-205.
Miller, D. J. and K. M. Burnett (2008). "A probabilistic model of debris-flow delivery to stream channels, demonstrated for the Coast Range of Oregon, USA." Geomorphology 94: 184-205.
Miller, D. J., et al. (2007). Facotrs controlling availability of spawning habitat for salmonids at the basin scale. Salmonid Spawning Habitat in RIvers: Physical Controls, Biological Responses, and Approaches to Remediation, American Fisheries Society.
Miller, D. J., et al. (2008). Factors controlling availability of spawning habitat for salmonids at the basin scale. Salmon Spawning Habitat in Rivers: Physical Controls, Biological Responses and Approaches to Remediation. D. Sear and P. DeVries. Bethesda, MD, American Fisheries Society. Symposium 65: 103-120.
Miller, D. J., et al. (2002). "A simple model for regional prediction of debris flow impacts. In http://www.fsl.orst.edu/clams/; Projects, Broad-Scale Models of Landslides and Debris Flows." from www.fsl.orst.edu/clams.
Miller, D. J., et al. (?). Sediment Delivery by Debris Flow to a Channel Network. ?, ?
Miller, D. J. and T. Dunne (1996). "Topographic perturbations of regional stresses and consequent bedrock fracturing." Journal of Geophysical Research 101(B11): 25, 523-525,536.
Miller, D. J., et al. (2003). "Time, space, and episodicity of physical disturbance in streams." Forest Ecology and Management 178(1-2): 121-140.
Miller, D. J. and J. Sias (1998). "Deciphering large landslides: linking hydrological, groundwater and slope stability models through GIS." Hydrological Processes 12: 923-941.
Miller, D. J. and J. C. Sias (1998). "Deciphering large landslides: linking hydrological, groundwater and slope stability models through GIS." Hydrological Processes 12: 923-941.
Miller, J. F., et al. (1973). NOAA ATLAS 2: Precipitation-Frequency Atlas of the Western United States, Volume 9, Washington, National Oceanic and Atmospheric Administration.
Miller, J. P. (1958). High Mountain Streams: Effects of Geology on Channel Characteristics and Bed Material. Socorro, New Mexico.
Miller, J. P. (1958). High mountain streams: effects of geology on channel characteristics and bed material, New Mexico Bureau of Mines and Mineral Resources: 53.
Miller, J. R., et al. (2004). "Spatial extrapolation: the science of predicting ecological patterns and processes." BioScience 54(4): 310-320.
Ecologists are often asked to contribute to solutions for broadscale problems. The extent of most ecological research is relatively limited, however,
necessitating extrapolation to broader scales or to new locations. Spatial extrapolation in ecology tends to follow a general framework in which
(a) the objectives are defined and a conceptual model is derived; (b) a statistical or simulation model is developed to generate predictions, possibly
entailing scaling functions when extrapolating to broad scales; and (c) the results are evaluated against new data. In this article, we examine the
application of this framework in a variety of contexts, using examples from the scientific literature.We conclude by discussing the challenges, limitations,
and future prospects for extrapolation.
 
Milliman, D. D. and J. P. M. Syvitski (1992). "Geomorphic/tectonic control of sediment discharge to the ocean: the importance of small mountain rivers." Journal of Geology 100: 525-544.
Mills, D. E. (?). "Flood Hazards in Washington." Engineering Geology 1: 65-70.
Mills, H. H. (1981). "Boulder deposits and the retreat of mountain slopes, or, "Gully Gravure" revisited." Journal of Geology 89: 649-660.
Mills, H. H. (1989). "Hollow form as a function of boulder size in the Valley and Ridge province, southwestern Virginia." Geology: 595-597.
Millspaugh, S. H. and C. Whitlock (1988). "A 750-year fire history based on lake sediment records in central Yellowstone National Park, USA." Environmental Management 12(3): 349-358.
Millspaugh, S. H., et al. (2000). "Variations in fire frequency and climate over the past 17000 yr in central Yellowstone National Park." Geology 28: 211-214.
Milne, J. A. and D. A. Sear (1997). "Modelling river channel topography using GIS " Int. J. Geographical Information Science 11(5): 499-519.
Milner, A. M., et al. (2000). "Colonization and development of stream communities across a 200-year gradient in Glacier Bay National Park, Alaska, U.S.A." Canadian Journal Fisheries and Aquatic Sciences 57: 2319-2335.
Mingguo, Z., C. Quiangguo, C. Quinjuan (2008). "Modelling the runoff-sediment yield relationship using a proportional function in hilly areas of the Loess Plateau, North China." Geomorphology 93: 288-301.
Minshall, G. W., et al. (1989). "Wildfires and Yellowstone's Stream Ecosystems." BioScience 39(10): 707-715.
Minshall, G. W., et al. (1985). "Developments in stream ecosystem theory." Canadian Journal of Fisheries and Aquatic Science 42: 1045-1055.
Minshall, G. W., et al. (1985). "Species richness in streams of different size from the same drainage basin." The American Naturalist 125(1): 16-38.
Minshall, G. W., et al. (1997). "Immediate and mid term responses of lotic ecosystems in Yellowstone National Park, USA to wildfire." Canadian Journal Fisheries and Aquatic Sciences 54: 2509-2525.
Minshall, W. G., et al. (1985). "Developments in stream ecosystem theory " Canadian Journal of Fisheries and Aquatic Sciences 42: 1045-1055.
Mitchell, C. E., et al. (1994). "Present-day veretical deformation of the Cascadia margin, Pacific Northwest, United States." Journal of Geophysical Research 99: 12257-12277.
Mitchell, J. K., et al. (1989). Validation of AGNPS for small watersheds using an integrated AGNPS/GIS system. Ionic Currents in Development., 1989: 833-842.
The AGNPS (AGricultural NonPoint Source) model was evaluated for predicting runoff and sediment delivery from small watersheds of mild topography. Fifty sediment yield events were monitored from two watersheds and five nested subwatersheds in East Central Illinois throughout the growing season of four years. Half of these events were used to calibrate parameters in the AGNPS model. Average calibrated parameters were used as input for the remaining events to obtain runoff and sediment yield data. These data were used to evaluate the suitability of the AGNPS model for predicting runoff and sediment yield from small, mild-sloped watersheds. An integrated AGNPS/GIS system was used to efficiently create the large number of data input changes necessary to this study. This system is one where the AGNPS model was integrated with the GRASS (Geographic Resources Analysis Support System) GIS (Geographical Information System) to develop a decision support tool to assist with management of runoff and erosion from agricultural watersheds. The integrated system assists with the development of input GIS layers to AGNPS, running the model, and interpretation of the results.
 
Mitchell, J. K., et al. (1993). "Validation of Agnps for Small Watersheds Using an Integrated Agnps/Gis System." Water Resources Bulletin 29(5): 833-842.
Mizuyama, T. (1981). An intermediate phenomenon between debris flow and bed load transport. Erosion and Sediment Transport in Pacific Rim Steeplands Publ. No. 132. Christchurch, International Association of Hydrological Sciences: 212-224.
Modde, T., et al. (1991). "Use of a Habitat-Based Stream Classification System for Catergorizing Trout Biomass " North American Journal of Fisheries Management 11: 305-311.
Moglen, G. E., et al. (1998). "On the sensitivity of drainage density to climate change." Water Resources Research 34(4): 855-862.
Moglen GE, a. R. B. (2002). "Spatially explicit hydrologic modeling of land use change." Journal of the American Water Resources Association 38(1): 241-253.
Mohr, J. A., et al. (2000). "Postglacial Vegetation and Fire History, Eastern Klamath Mountains, California, USA." The Holocene 10(4): 587-601.
Mohseni, O., et al. (1998). "A nonlinear regression model for weekly stream temperatures." Water Resources Research 34(10): 2685-2692.
Moir, H. J., et al. (1998). "Hydraulic and sedimentary characteristics of habitat utilized by Atlantic salmon spawning in the Girnock Burn, Scotland." Fisheries Management and Ecology 5: 241-254.
Moir, H. J., et al. (2002). "Hydraulic and sedimentary controls on the availability and use of Atlantic salmon (Salmo salar) spawning habitat in the River Dee system, north-east Scotland " Geomorphology 45: 291-308.
Molinas, A., and B. Wu (1998). "Effect of size gradation on transport of sediment mixtures." Journal of Hydraulic Engineering-ASCE 124(8): 786-793.
Molnar, P., et al. (1987). "Geologic Evolution of Northern Tibet: Results of an Expedition to Ulugh Muztagh." Science 235: 299-305.
Molnar, P., et al. (2002). "Integrated catchment assessment of riverine landscape dynamics." Aquatic Science 64: 129-140.
Abstract. The traditional approach to study riverine
environments focuses on the river reach scale, with
streamflow as a steady state driving force. Here, the ac-cent
is on the dynamic nature of streamflow. Impacts of
the hydrological regime, of floods and streamflow vari-ability,
on riverine landscapes are reviewed. To evaluate
such impacts, it is necessary to focus on the entire catch-ment
in an integrated fashion, so that local changes in
river morphology and river habitat can be evaluated in
context with upstream catchment processes. A framework
for an integrated physically-based catchment modelling
system, based on models of hydrology, hydrodynamics,
sedimentology and ecology, is presented. The hydrologi-
cal element addresses runoff response in a catchment on
a continuous basis in time and distributed in space, while
the hydrodynamic, sedimentological and ecological ele-ments
address the interactions and feedbacks between
water, sediment and the ecosystems at the scale of the
river corridor. The models are arranged in a nested
fashion, with long-term quantification of catchment and
river system dynamics as the main objective. A long-term
vision of catchment processes is important for the
evaluation of potential anthropogenic influences and
climate change effects, as well as for the evaluation of
river conservation projects.
 
Molnar, P. and P. England (1990). "Late Cenozoic uplift of mountain ranges and global climate change: chicken or egg?" Nature 346: 29-34.
Mondry, Z., et al. (2000). Influences of Sediment and Large Wood Inputs on 1997 Flood Effects in three Klamath Mountain Streams.
Montgomery D.R., T. B. A., J.M. Buffington, N.P. Peterson, K.M. Schmidt, and J.D. Stock (1996). "Distribution of bedrock and alluvial channels in forested mountain drainage basins." Nature 381: 587-589.
Montgomery, D. R. "Signifigance of Thick Deposits of Colluvium on Hillslopes: A Case Study Involoving the Use of Pollen Analysis in the Coastal Mountains of Northern California." Jour. Geol. 84: 147-158.
Montgomery, D. R., and J.M. Buffington (1993). Channel classification, prediction of channel response, and assessment of channel condition, Washington State Timber/Fish/Wildlife Agreement: 110.
Montgomery, D. R. (1994). "Road surface drainage, channel initiation, and slope instability." Water Resources Research 30(6): 1925-1932.
Montgomery, D. R. (1996). "Distribution of bedrock and alluvial channels in forested mountain drainage basins." Nature 381(13): 587-689.
Montgomery, D. R., and J.M. Buffington (1997). "Channel-reach morphology in mountain drainage basins." GSA Bulletin 109(5): 596-611.
Montgomery, D. R. (1999). "Process domains and the river continuum." Journal of the American Water Resources Association 35(2): 397-410.
Montgomery, D. R. (2000). "Coevolution of the Pacific salmon and Pacific Rim topography." Geology 28: 1107-1110.
Montgomery, D. R., K.M. Schmidt, H.M. Greenberg, and W.E. Dietrich (2000). "Forest clearing and regional landsliding." Geology 28(4): 311-314.
Montgomery, D. R., and K.B. Gran (2001). "Downstream variations in width of bedrock channels." Water Resources Research 37(6): 1841-1846.
Montgomery, D. R., B.D.Collins, J.M. Buffington, and T.B. Abbe (2003). Geomorphic effect of wood in rivers. American Fisheries Society Symposium 2003.
Montgomery, D. R. and T. B. Abbe (2006). "Influence of logjam-formed hard points on the formation of valley-bottom landforms in an old-growth forest valley, Queets River, Washington, USA." Quaternary Research 65: 147-155.
Field surveys and radiocarbon dating of buried logjams in the floodplain of an old-growth forest river demonstrate the formation of erosionresistant
‘‘hard points’’ on the floodplain of the Queets River, Washington. These hard points provide refugia for development of old-growth forest
patches in frequently disturbed riparian environments dominated by immature forest. Our surveys show that local bed aggradation associated with
logjams not only influences channel patterns and profiles but leads to development of a patchwork of elevated landforms that can coalesce to form
portions of the valley bottom with substantial (i.e., 1 to >4 m) relief above the bankfull elevation. In addition, logjam-formed hard points promote
channel avulsion, anastomosing morphology, and growth of mature patches of floodplain forest that, in turn, provide large logs needed to form
more logjam-formed hard points. Hence, our findings substantiate the potential for a feedback mechanism through which hard points sustain
complex channel morphology and a patchwork floodplain composed of variable-elevation surfaces. Conversely, such a feedback further implies
that major changes in riparian forest characteristics associated with land use can lead to dramatic simplification in channel and floodplain
morphology
 
Montgomery, D. R., et al. (1996). "Distribution of bedrock and alluvial channels in forested mountain drainage basins." Nature 381: 587-589.
Montgomery, D. R., et al. (1999). "Channel type and salmonid spawning distribution and abundance." Canadian Journal of Fisheries and Aquatic Science 56: 377-387.
Consideration of fundamental channel processes, together with map-based and field investigations, indicates
that stream channel type influences salmonid spawning distributions across entire channel networks and salmonid
abundance within channel reaches. Our analysis suggests that salmonid spawning patterns in mountain drainage basins
of the Pacific Northwest are adapted to, among other things, the timing and depth of channel bed mobility. We
hypothesize that because the bed of pool–riffle and plane–bed reaches scours to a variable fraction of the thickness of
alluvium, survival to emergence is favored by either burying eggs below the annual scour depth or avoiding egg burial
during times of likely bed mobility. Conversely, annual mobility of all available spawning gravel in steeper step–pool
and cascade channels favors either adaptations that avoid egg burial during times of likely bed mobility or selection of
protected microhabitats. Consistent with these expectations, we find that salmonid spawning distributions track channel
slope distributions in several west-slope Pacific Northwest watersheds, implying that spatial differences in channel
processes influence community structure in these rainfall-dominated drainage basins. More detailed field surveys
confirm that different channel types host differential use by spawning salmonids and reveal finer-scale influences of
pool spacing on salmonid abundance.
 
 
Montgomery, D. R. and S. M. Bolton (2003). Hydrogeomorphic variability and river restoration. Strategies for Restoring River Ecosystems: Sources of Variability and Uncertainty in Natural and Managed Systems. R. C. Wissmar and P. A. Bisson. Bethesda, Maryland, American Fisheries Society: 39-79.
Montgomery, D. R. and J. M. Buffington (1993). Channel classification, prediction of channel response, and assessment of channel condition Seattle, Washington, Washington State Timber Fish and Wildlife: 83.
Montgomery, D. R. and J. M. Buffington (1997). "Channel-reach morphology in mountain drainage basins." Geological Society of America Bulletin 106(5): 596-611.
Montgomery, D. R. and J. M. Buffington (1998). Channel Processes, Classification, and Response. River Ecology and Management. R. J. Naiman and R. E. Bilby. New York, Springer-Verlag: 13-42.
Montgomery, D. R., et al. (1996). "Stream-bed scour, egg burial depths, and the influence of salmonid spawning on bed surface mobility and embryo survival " Canadian Journal of Fisheries and Aquatic Sciences 53: 1061-1070.
Montgomery, D. R., et al. (1995). "Pool spacing in forest channels." Water Resources Research 31(4): 1097-1105.
Montgomery, D. R., et al. (1995). "Pool spacing in forest channels." Water Resources Research 31: 1097-1105.
Montgomery, D. R., et al. (2003). Geomorphic effects of wood in rivers. The Ecology and Management of Wood in World Rivers. S. V. Gregory, K. L. Boyer and A. M. Gurnell. Bethesda, Maryland, American Fisheries Society: 21-47.
Montgomery, D. R. and W. E. Dietrich (1988). "Where do channels begin?" Nature 336: 232-234.
Montgomery, D. R. and W. E. Dietrich (1989). "Source areas, drainage density, and channel initiation." Water Resources Research 25(8): 1907-1918.
Montgomery, D. R. and W. E. Dietrich (1992). "Channel initiation and the problem of landscape scale." Science 255: 826-830.
Montgomery, D. R. and W. E. Dietrich (1994). "A physically based model for the topographic control on shallow landsliding." Water Resources Research 30(4): 1153-1171.
Montgomery, D. R. and W. E. Dietrich (2002). "Runoff generation in a steep, soil-mantled landscape." Water Resources Research 38(9): 1168.
Montgomery, D. R., et al. (2002). "Piezometric response in shallow bedrock at CB1: Implications for runoff generation and landsliding." Water Resources Research 38(12): doi:10.1029/2002WR001429, 002002.
Montgomery, D. R., et al. (1998). The role of GIS in watershed analysis. Landform Monitoring, Modelling and Analysis. S. N. Lane, K. S. Richards and J. H. Chandler, John Wiley & Sons Ltd.: 241-261.
Watershed analysis methods characterize historical, current and potential geomorphological and ecological processes and conditions within a drainage basin in order to guide assessment of the ecological impacts of land management. Within the context of watershed analysis, Geographical Information Systems (GIS) off advantages for: (i) predicting the spatial distribution of geomorphological processes; (ii) formulating spatially explicit hypotheses against which to compare historical reconstructions and field observations; (iii) denoting central tendencies in landscape properties; (iv) generalising from point field data to an entire watershed; and (v) exploring and illustrating potential effects of specific land management strategies. Evaluation of watershed conditions and potential response to land management decisions requires field data, however, and cannot consist entirely of GIS-driven models because landscape conditions reflect the effects of historic variability in geomorphological processes. Watershed-scale predictions of areas prone to shallow landsliding, of stream channel type, and of bed-surface grain size illustrate how GIS-driven analytical models can contribute to the analysis of geomorphological processes in mountain drainage basins. Although such spatially explicit predictions can provide hypotheses for landscape sensitivity, digital elevation models do not always adequately represent relevant topographic attributes (e.g. fine-scale hollows, channel slope and channel confinement). Also, mountain drainage basins are characterised by temporal variability in landscape characteristics that arise from disturbance and recovery processes. The framework outlined here provides a model for integrating GIS-based analyses into procedures for generating a landscape-scale understanding of geomorphological processes, and for linking such an understanding to land use decision making.
 
Montgomery, D. R., et al. (1997). "Hydrologic response of a steep, unchanneled valley to natural and applied rainfall." Water Resources Research 33(1): 91-109.
Montgomery, D. R. and E. Foufoula-Georgiou (1993). "Channel network source representation using digital elevation models." Water Resources Research 29(12): 3925-3934.
Montgomery, D. R. and K. B. Gran (2001). "Downstream variations in the width of bedrock channels." Water Resources Research 37(6): 1841-1846.
Field surveys of channel width w and drainage area A in bedrock channel
reaches reveal relationships where w 5 cAb, similar to the classic hydraulic geometry of
alluvial channels. Data from five mountain channel networks support the assumption used
in many landscape evolution models that an alluvial hydraulic geometry relationship where
b 5 0.3–0.5 holds for bedrock channel systems. Although there is substantial local
variability in channel width in bedrock channel systems, there is no systematic difference
in width versus drainage area relations for the surveyed bedrock and alluvial reaches in
sedimentary lithologies in coastal Oregon and Washington. In contrast, bedrock channels
were narrower, and therefore had deeper flow, than alluvial channels with equal drainage
areas in the granite and limestone terrain of the Yuba River, California. In addition, data
from the Mokelumne River show that bedrock channel width decreases substantially
downstream at the contact between relatively weak limestone and more erosion-resistant
granite, but that channel slope does not change appreciably across contacts between these
two lithologies. Data from coastal Oregon drainage basins further show systematic channel
widening after flood flows and debris flow impacts. We conclude that downstream
variations in the width of bedrock channels generally follow traditional hydraulic geometry
relations but also reflect the local influence of longitudinal patterns of bedrock erosivity
and disturbance history.
 
Montgomery, D. R., et al. (1995). "Watershed Analysis as a Framework for Implementing Ecosysterm Management." Water Resources Bulletin 31(3): 369-386.
Montgomery, D. R. and L. H. MacDonald (2002). "Diagnistic approach to stream channel assessment and monitoring." Journal of American Water Resources Association 38: 1-16.
Montgomery, D. R., et al. (2002). "Influence of debris flows and log jams on the location of pools and alluvial channel reaches, Oregon Coast Range." Geological Society of America Bulletin 115(1): 78-88.
Montgomery, D. R., et al. (2003). "Influence of debris flows and log jams on the location of pools and alluvial channel reaches, Oregon Coast Range." Geological Society of America Bulletin 115(1): 78-88.
Montgomery, D. R., et al. (1999). "Channel-bed mobility response to extreme sediment loading at Mount Pinatubo." Geology 27(3): 271-274.
Montgomery, D. R., et al. (2000). "Forest clearing and regional landsliding." Geology 28(4): 311-314.
Montgomery, D. R., et al. (1998). "Regional test of a model for shallow landsliding." Hydrological Processes 12(6): 943-955.
Montgomery, D. R., et al. (1991). "Debris Flow Hazard Mitigation for Colluvium-Filled Swales." Bulletin of the Association of Engineering Geologists XXVIII(3): 303-324.
Montgomery, D. R., et al. (1991). "Debris flow hazard mitigation for colluvium-filled swales." Bulletin of the Association of Engineering Geologists 28(3): 303-323.
Moody, J. A., and D.A. Martin (2001). "Initial hydrologic and geomorphic response following a wildfire in the Colorado Front Range." Earth Surface Processes and Landforms 26: 1049-1070.
Moody, J. A. and D. A. Kinner (2006). "Spatial structures of stream and hillslope drainage networks following gully erosion after wildfire." Earth Surface Processes and Landforms 31(3): 319-337.
The drainage networks of catchment areas burned by wildfire were analysed at several scales. The smallest scale (1-1000 m2) representative of hillslopes, and the small scale (1000 m2 to 1 km2), representative of small catchments, were characterized by the analysis of field measurements. The large scale (1-1000 km2), representative of perennial stream networks, was derived from a 30-m digital elevation model and analysed by computer analysis.
Scaling laws used to describe large-scale drainage networks could be extrapolated to the small scale but could not describe the smallest scale of drainage structures observed in the hillslope region. The hillslope drainage network appears to have a second-order effect that reduces the number of order 1 and order 2 streams predicted by the large-scale channel structure. This network comprises two spatial patterns of rills with width-to-depth ratios typically less than 10. One pattern is parallel rills draining nearly planar hillslope surfaces, and the other pattern is three to six converging rills draining the critical source area uphill from an order 1 channel head. The magnitude of this critical area depends on infiltration, hillslope roughness and critical shear stress for erosion of sediment, all of which can be substantially altered by wildfire. Order 1 and 2 streams were found to constitute the interface region, which is altered by a disturbance, like wildfire, from subtle unchannelized drainages in unburned catchments to incised drainages. These drainages are characterized by gullies also with width-to-depth ratios typically less than 10 in burned catchments. The regions (hillslope, interface and channel) had different drainage network structures to collect and transfer water and sediment. 
 
Moody, J. A. and D. A. Martin (2001).
Moody, J. A., et al. (2008). "Post-wildfire erosion response in two geologic terrains in the western USA." Geomorphology 95: 103-118.
Volumes of eroded sediment after wildfires vary substantially throughout different geologic terrains across the western United
States. These volumes are difficult to compare because they represent the response to rainstorms and runoff with different
characteristics. However, by measuring the erosion response as the erodibility efficiency of water to detach and transport sediment
on hillslopes and in channels, the erosion response from different geologic terrains can be compared. Specifically, the erodibility
efficiency is the percentage of the total available stream power expended to detach, remobilize, or transport a mass of sediment.
Erodibility efficiencies were calculated for the (i) initial detachment, and for the (ii) remobilization and transport of sediment on the
hillslopes and in the channels after wildfire in two different geological terrains.
The initial detachment efficiencies for the main channel and tributary channel in the granitic terrain were 10±9% and 5±4%
and were similar to those for the volcanic terrain, which were 5±5% and 1±1%. No initial detachment efficiency could be
measured for the hillslopes in the granitic terrain because hillslope measurements were started after the first major rainstorm. The
initial detachment efficiency in the volcanic terrain was 1.3±0.41%. The average remobilization and transport efficiencies
associated with flash floods in the channels also were similar in the granitic (0.18±0.57%) and volcanic (0.11±0.41%) terrains. On
the hillslope the remobilization and transport efficiency was greater in the volcanic terrain (2.4%) than in the granitic terrain
(0.65%). However, this may reflect the reduced sediment availability after the first major rainstorm (30-min maximum rainfall
intensity ∼90 mm h−1) in the granitic terrain, while easily erodible fine colluvium remained on the hillslope after the first rainstorm
(30-min maximum rainfall intensity=7.2 mm h−1) in the volcanic terrain. The erosion response in channels and on hillslopes of the
granitic and volcanic terrains was similar when compared using erodibility efficiencies.
 
Moody, J. A., et al. (1999). "Ontogeny of a flood plain " GSA Bulletin 111(2): 291-303.
Moody, J. A., et al. (1999). "Ontogeny of a flood plain." Geological Society of America Bulletin 111(2): 291-303.
Moog, D. B. and P. J. Whiting (1998). "annual hysteresis in bed load rating curves." Water Resources Research 34(9): 2393-2399.
Moore, I. D., and G.J. Burch (1986). "Sediment transport capacity of sheet and rill flow: application of unit stream power theory." Water Resources Research 22(8): 1350-1360.
Moore, I. D., R.B.Grayson, and A.R.Ladson (1991). "Digital terrain modelling: a review of hydrolgical, geomorphological, and biological applications." Hydrological Processes 5: 3-30.
Moore, I. D. and R. B. Grayson (1991). "Terrain-based cathcment partitioning and runoff prediction using vector elevation data." Water Resources Research 27(6): 1177-1191.
Moore, P. (1997). "Life after logging." Forestry 80(1): 61-67.
Moore, P. (?). Green Spirit: Trees are the answer.
Moore, R. D. (2002). Small Stream Channels and their Riparian Zones: A Review of Hydrology and Climate-Related Processes Relevant to Riparian Management in a Pacific Northwest Forestry Context, University of British Columbia: 1-48.
Moore, R. D., et al. (2005). "Thermal regime of a headwater stream within a clear-cut, Coastal British Columbia, Canada." Hydrological Processes doi: 10.1002/hyp.5733.
Moore, R. D. and S. M. Wondzell (2005). "Physical hydrology and the effects of forest harvesting in the Pacific Northwest: a review." Journal of the American Water Resources Association 41(4): 763-784.
The Pacific Northwest encompasses a range of hydrologic regimes that can be broadly characterized as either coastal (where rain and rain on snow are dominant) or interior (where snowmelt is dominant). Forest harvesting generally increases the fraction of precipitation that is available to become streamflow, increases rates of snowmelt, and modifies the runoff pathways by which water flows to the stream channel. Harvesting may potentially decrease the magnitude of hyporheic exchange flow through increases in fine sediment and clogging of bed materials and through changes in channel morphology, although the ecological consequences of these changes are unclear. In small headwater catchments, forest harvesting generally increases annual runoff and peak flows and reduces the severity of low flows, but exceptions have been observed for each effect. Low flows appear to be more sensitive to transpiration from vegetation in the riparian zone than in the rest of the catchment. Although it appears that harvesting increased only the more frequent, geomorphically benign peak flows in several studies, in others the treatment effect increased with return period. Recovery to pre-harvest conditions appeared to occur within about 10 to 20 years in some coastal catchments but may take many decades in mountainous, snow dominated catchments.
 
 
Morantz, D. L., et al. (1987). "Selection of Microhabitat in Summer by Juvenile Atlantic salmon (Salmo salar) " Canadian Journal of Fisheries and Aquatic Sciences 44: 120-129.
Morgan, R. P. C., J.N. Quinton, R.E.Smith, G. Govers, J.W.A.Paesen, K. Auerswald, G. Chisci, D. Torri, and M.E. Styczen (1998). "The European soil erosion model (EUROSEM): a dynamic approach for predicting sediment transport from fields and small catchments." Earth Surface Processes and Landforms 23: 527-544.
The European Soil Erosion Model (EUROSEM) is a dynamic distributed model, able to simulate sediment transport, erosion and deposition over the land surface by rill and interill processes in single storms for both individual fields and small catchments. Model output includes total runoff, total soil loss, the storm hydrograph and storm sediment graph. Compared with other erosion models, EUROSEM has explicit simulation of interill and rill flow; plant cover effects on interception and rainfall energy; rock fragment (stoniness) effects on infiltration, flow velocity and splash erosion; and changes in the shape and size of rill channels as a result of erosion and deposition. The transport capacity of runoff is modelled using relationships based on over 500 experimental observations of shallow surface flows. EUROSEM can be applied to smooth slope planes without rills, rilled surfaces and surfaces with furrows. Examples are given of model output and of the unique capabilities of dynamic erosion modelling in general.
 
Morgan, R. P. C. (2001). "A simple approach to soil loss prediction: a revised Morgan-Morgan-Finney model." Catena 44(4): 305-322.
Morisawa, M. (1957). "Accuracy of determination of stream lengths from topographic maps." Transactions of the American Geophysical Union 38(1): 86-88.
Morrison, M., L., et al. (1998). Wildlife habitat relationships: concepts and applications. Madison, Wisconsin, USA, The University of Wisconsin Press.
Morrison, P. and F. Swanson (1990). Fire history in two forest ecosystems of the central western Cascades of Oregon, USDA Forest Service.
Morrison, P. H. (1975). Ecological and geomorphological consequences of mass movements in the Alder Creek watershed and implications for land management. Eugene, University of Oregon: 102.
Morrison, P. H. and F. J. Swanson (1990). Fire History and Pattern in a Cascade Range Landscape. Portland, Oregon, USDA Forest Service: 77.
Morrison, R. B. (1991). Quaternary stratigraphic, hydrologic, and climatic history of the Great Basin, with emphasis on Lakes Lahontan, Bonneville, and Tecopa The Geology of North America. Chapter 10: 283-320.
Morton, D. M., et al. (2008). "Contrasting rainfall generated debris flows from adjacent watersheds at Forest Falls, southern California, USA." Geomorphology 96: 322-338.
Debris flows are widespread and common in many steeply sloping areas of southern California. The San Bernardino Mountains
community of Forest Falls is probably subject to the most frequently documented debris flows in southern California. Debris flows at
Forest Falls are generated during short-duration high-intensity rains that mobilize surface material. Except for debris flows on two
consecutive days in November 1965, all the documented historic debris flows have occurred during high-intensity summer rainfall,
locally referred to as ‘monsoon’ or ‘cloudburst’ rains. Velocities of the moving debris range from about 5 km/h to about 90 km/h.
Velocity of a moving flow appears to be essentially a function of the water content of the flow. Low velocity debris flows are
characterized by steep snouts that, when stopped, have only small amounts of water draining from the flow. In marked contrast are highvelocity
debris flows whose deposits more resemble fluvial deposits. In the Forest Falls area two adjacent drainage basins, Snow Creek
and Rattlesnake Creek, have considerably different histories of debris flows. Snow Creek basin, with an area about three times as large
as Rattlesnake Creek basin, has a well developed debris flow channel with broad levees. Most of the debris flows in Snow Creek have
greater water content and attain higher velocities than those of Rattlesnake Creek.Most debris flows are in relative equilibrium with the
geometry of the channel morphology. Exceptionally high-velocity flows, however, overshoot the channel walls at particularly tight
channel curves. After overshooting the channel, the flows degrade the adjacent levee surface and remove trees and structures in the
immediate path, before spreading out with decreasing velocity. As the velocity decreases the clasts in the debris flows pulverize the upslope
side of the trees and often imbed clasts in them. Debris flows in Rattlesnake Creek are relatively slowmoving and commonly stop
in the channel. After the channel is blocked, subsequent debris flows cut a new channel upstream from the blockage that results in the
deposition of new debris-flow deposits on the lower part of the fan. Shifting the location of debris flows on the Rattlesnake Creek fan
tends to prevent trees from becoming mature. Dense growths of conifer seedlings sprout in the spring on the late summer debris flow
deposits. This repeated process results in stands of even-aged trees whose age records the age of the debris flows.
 
Morton, D. M. and R. H. Campbell (1974). "Spring mudflows at Wrightwood, Southern California." Q. Jl Engng Geol. 7: 377-384.
Mosley, M. P. (1976). "An experimental study of channel confluences." Journal of Geology 84: 535-562.
Mosley, M. P. (1978). "Bed material transport in the Tamaki River near Dannevirke, North Island, New Zealand." New Zealand Journal of Science 21: 619-626.
Mosley, M. P. (1981). "The influence of organic debris on channel morphology and bedload transport in a New Zealand forest stream." Earth Surface Processes and Landforms 6: 571-579.
Mossop, B. and M. J. Bradford (2004). "Importance of large woody debris for juvenile chinook salmon habitat in small boreal forest streams in the upper Yukon River basin, Canada." Canadian Journal of Forest Research 34: 1955-1966.
The importance of large woody debris (LWD) in forested stream ecosystems is well documented. However,
little is known about LWD in northern boreal forest streams. We investigated the abundance, characteristics, and function
of LWD in 13 small tributary streams of the upper Yukon River basin, Yukon Territory, Canada. LWD abundance
was similar to values reported from temperate regions, whereas LWD size and total volume were well below values for
the Pacific Northwest. LWD formed 28% of the pools, which provide important habitat for juvenile chinook salmon
(Oncorhynchus tshawytscha Walbaum). The median diameter of pool-forming pieces was 17 cm, and ring counts on
fallen riparian trees indicated that pool-forming pieces were likely 70–200 years old when downed. Juvenile chinook
salmon density was correlated with LWD abundance in our study reaches. We conclude that despite differences in climate
and forest type, LWD in Yukon streams and LWD in temperate regions appear to perform a similar function in
creating fish habitat. Resource managers should consider the relatively slow tree growth and thus potentially long recovery
times following human disturbances in these watersheds.
 
Mossop, B. and M. J. Bradford (2006). "Using thalweg profiling to assess and monitor juvenile salmon (Oncorhynchus spp.) habitat in small streams." Canadian Journal of Fisheries and Aquatic Science 63: 1515-1525.
Thalweg profiles are longitudinal profiles of the streambed elevation measured along the deepest portion of
the stream. This technique has recently been advocated as a tool to assess and monitor fish habitat in streams because
metrics calculated from thalweg profiles can provide useful information on habitat quality, and measurements are both
repeatable and independent of stream flow. Relations between thalweg metrics and land use have also been documented.
However, a relation between fish abundance and thalweg metrics has not been established. To develop this relation, we
surveyed thalweg profiles and sampled juvenile Chinook salmon (Oncorhynchus tshawytscha) density in 14 reaches of
small tributary streams of the upper Yukon River. Chinook salmon density was correlated with three thalweg metrics.
Two of these metrics — length in residual pool and mean maximum residual pool depth — provided useful measures
of pool extent and quality and useful information on Chinook salmon habitat. Thalweg metrics differed between these
undisturbed streams and reaches in streams affected by placer gold mining. These results suggest that thalweg profiling
provides a useful tool to assess and monitor fish habitat in small streams.
 
Motha, J. A., P.J. Wallbrink, P.B. Hairsine, and R.B. Grayson (2003). "Determining the sources of suspended sediment in a forested catchment in southeastern Australia." Water Resources Research 39(3).
Mount, J. F. (1995). California Streams and Rivers: The Conflict Between Fluvial Processes and Land Use. Berkeley, University of California Press.
Mouton, A. (2005). Generating Stream Maps Using LiDAR Derived Digital Elevation Models and 10-m USGS DEM. Forest Resources. Seattle, WA, University of Washington. Masters: 78.
The effects of digital elevation model (DEM) grid size for stream network predictions in
the northwestern United States were examined to test the accuracy of high-resolution
LiDAR (Light Detection And Ranging) digital elevation data. LiDAR elevation data
were gridded at 2-, 6-, and 10-m scales and flow paths were predicted by four common
routing algorithms known as D8, D-Infinity, Multiple Flow, and DEMON, D8 being the
least sophisticated. These routing algorithms were also applied to a 10-m USGS DEM to
compare LiDAR with the previously used data for hydrologic modeling. The analyses
indicated that as topographic detail increased, all LiDAR-derived models delineated more
streams and located streams in their topographically correct position when compared to a
10-m USGS DEM. Stream maps generated by either D8 or DEMON converged as the
DEM resolution was increased. The data suggests that increased DEM resolution
decreases the need for sophisticated models, reducing processing times required to create
accurate stream locations and attributes.
LiDAR digital elevation data also improved the modeling of perennial stream heads and
fish habitat potential in a direct comparison to a 10-m USGS DEM. Distances between
stream heads predicted using a LiDAR dataset and field verified stream heads were
significantly less than those predicted using a USGS dataset. This illustrates the potential
use of LiDAR to accurately predict perennial flow in a given landscape. The ability to
locate fish barriers based on landscape gradient also improved with LiDAR data. A
USGS dataset used to find fish barriers occasionally found barriers in places where none
existed or vice versa. As LiDAR datasets become more available, automated creation of
stream networks and their hydrologic features will become more feasible and the
accuracy of the results will be much improved.
 
 
Moyle, P. B. (?). "Status of Fish Amphibians, and Watersheds in the Sierra Nevada: Reflections on a Decline in Ecosystem Health ". Retrieved 10/13/98, 1998, from http://glinda.cnrs.humboldt.edu/WMChome/new/sum_97/fish_sn.html.
Mueller, J. E. (1972). "Re-evaluation of the relationship of master streams to drainage basins." Bulletin of the Geological Society of America 83: 3471-3474.
Mulvey, W. E. and M. Lowe (1992). 1991 Cameron Cove Debris Flow, North Ogden, Utah. Conference on Arid West Floodplain Management Issues, Las Vegas, Nevada.
Murphy, M. (1996). A Model of Coho Salmon Reponses to Buffer Zones (DRAFT): 1-10.
Murphy, M. L. and K. V. Koski (1989). "Input and depletion of woody debris in Alaska streams and implications for streamside management." North American Journal of Fisheries Management 9: 427-436.
Murphy, P. N. C., et al. (2008). "Stream network modelling using lidar and photogrammetric digital elevation models: a comparison and field verification." Hydrological Processes 22: 1747-1754.
A conventional, photogrammetrically derived digital elevation model (DEM; 10 m resolution) and a light detection and ranging
(lidar)-derived DEM (1 m resolution) were used to model the stream network of a 193 ha watershed in the Swan Hills of
Alberta, Canada. Stream networks, modelled using both hydrologically corrected and uncorrected versions of the DEMs and
derived from aerial photographs, were compared. The actual network, mapped in the field, was used as verification. The lidar
DEM-derived network was the most accurate representation of the field-mapped network, being more accurate even than the
photo-derived network. This was likely due to the greater initial point density, accuracy and resolution of the lidar DEM
compared with the conventional DEM. Lidar DEMs have great potential for application in land-use planning and management
and hydrologic modelling. The network derived from the hydrologically corrected conventional DEM was more accurate than
that derived from the uncorrected one, but this was not the case with the lidar DEM.
 
 
Murray, A. B. and C. Paola (2003). "Modelling the effect of vegetation on channel pattern in bedload rivers." Earth Surface Processes and Landforms 28(2): 131-143.
We modify a simple numerical stream-pattern model to examine the effect of sediment stabilization by roots on the channel pattern of bedload rivers. In the model, vegetation enhances bank resistance to erosion, causing the development of a single channel instead of a rapidly changing, multiple channel (braided) pattern. Net aggradation resulting from a high sediment supply, however, causes frequent avulsions that destroy vegetation locally, leading to the development of a multiple-channel pattern. A stability diagram representing multiple model runs predicts whether a river will exhibit single or multiple channels, based on plant-enhanced bank strength, and on the time scale of plant development relative to a time scale for change in unvegetated channels. A second stability diagram predicts the way in which the amplitude and period of a fluctuating imposed sediment load influence whether a single or multiple-channel pattern develops. Copyright (C) 2003 John Wiley Sons, Ltd.
 
Murray, B. A. (2002). "Seeking Explaination Affects Numerical Modeling Strategies." EOS, Transaction American Geophysical Union 83(37): 418-419.
Murray, J., et al. Development of a GIS database for ground-water recharge assessment of the Palouse Basin. Soil Science 168 (11), pp. 759-768;   41 ref.; 2003.
The advent of Geographic Information Systems (GIS) is bringing about the expansion of soil survey data into interdisciplinary research projects. A GIS database was developed for the Palouse Basin in northern Idaho and eastern Washington to identify areas where soil and geologic features are likely to impact ground-water recharge. Using GeoProcessing operations in ArcView, 1:24,000 soil survey data and surficial geology were combined. The resulting ArcView-based GIS coverage was used to delineate recharge mechanisms and classify recharge potential in the Palouse Basin. A database was developed using binary weighting and index overlay modelling methods to assign values to soil map units based on selected soil characteristics, including permeability rates, depth to bedrock, and the presence of perched water tables. These data were then linked to Basin recharge mechanisms to produce maps indicating the potential for deep percolation through soil and subsequent recharge to the local aquifer system. Results indicate that recharge through loess is the most spatially extensive recharge mechanism, operating over 71% of the total study area. Of this, approximately 2300 ha have high potential for recharge. Recharge through stream loss operates over 16% of the total study area, whereas percolation through localized fractures in bedrock is the least extensive recharge mechanism, operating over just 13% of the total study area. Although stream loss and infiltration along Basin margins occupy a limited spatial extent they have more land area rated as 'high potential' for recharge. There is lower recharge potential in the eastern portion of the Basin because of the presence of extensive hydraulically restrictive subsoil horizons.
 
N.L. Poff, B. P. B., C.O. Cuhaciyan (2006). "Hydrologic variation with land use across the contiguous United States: Geomorphic an ecological consequences for stream ecosystems." Geomorphology 79: 264=285.
Nachtergaele, J., et al. (2002). "Medium-term evolution of a gully developed in a loess-derived soil." Geomorphology 46(3-4): 223-239.
Field surveys in the Belgian loess belt revealed the presence in many forested areas of large, permanent gully systems, most of which are currently inactive. In cultivated areas, such gullies can only be observed in cross-sectional soil profiles through hollows, as virtually all such large gullies are currently infilled with colluvium. Little is known about the spatial distribution, initiation and temporal evolution of these large, permanent gully systems on loess-derived soils. Therefore, the medium-term evolution of a gully initiated in a cultivated area on loess-derived soils southwest of Leuven (Belgium) in May-June 1986 was studied over 13 years. Two intense rainfall events created this (ephemeral) gully, which was not erased by subsequent tillage. Between June 1986 and December 1999, eight field surveys were conducted to measure gully dimensions. During two surveys, topographic indices (e.g., slope and drainage area) were also measured. Daily rainfall for the measuring period were obtained from a rainfall station located some 10 km southwest of the gully. Analysis of rainfall data showed that no extreme rainfall event was required to initiate such large (permanent) gullies, as observed in forested areas and through cross-sectional profiles in cultivated fields in the Belgian loess belt. Return periods of the event that caused the gully varied between <1 year and 25 years, depending on the assumptions used for defining event rain intensity. Once established, length, surface area and volume of the studied gully evolved with time, cumulative rainfall or cumulative runoff, following a negative exponential relation. This accords with observations reported for gullies in Australia and the USA. This study shows that a degressive increase of gully extension, can be largely explained by the evolution of a "slope-drainage area" factor (SXA, which is proportional to stream power) with time. While gully length and gully surface area asymptotically evolve towards a final value, gully volume decreased at a given point in time. From this, it is inferred that sediment deposition will potentially infill the gully to such an extent that the farmer can drive across it. From this moment on, the combined effect of water and tillage erosion in the gully drainage area, will lead towards rapid infilling. This expected evolution of a gully in cultivated fields accords with observations of large infilled gully systems in cultivated areas in eastern Belgium. The permanent gullies observed under forest are attributed to the fact that after severe gully erosion, this area was reforested or abandoned. Therefore, the sediment source was cut off and the gully was not filled in by sediment deposition.
 
Naden, P. S. and D. M. Cooper (1999). "Development of a sediment delivery model for application in large river basins." Hydrological Processes 13(7): 1011-1034.
Nagle, G. The White Line. Working the Woods Working the Sea. F. Wilcox and J. Gorsline: 71-75.
Nagle, G. (1979). "Hoedads."? ?: 33-36.
Nagle, G. N., et al. (1999). "The management of sedimentation of tropical watersheds (DRAFT)." Environmental Management 23(4): 441-452.
Nagy, H., K. Watanabe, and M. Hirano (2000). "Determination of sediment conveyance rate in alluvial streams by using neural networks." Journal of hydroscience and hydraulic engineering 18(2): 93-103.
Nagy, H., K. Watanabe, and M. Hirano (2002). "Prediction of sediment load concentration in rivers using ANN model." Journal of Hydraulic Engineering-ASCE 128(6): 588-595.
Naiman, R. J. (1996). "Water, society, and landscape ecology " Landscape Ecology 11(4): 193-196.
Naiman, R. J. (1999). "A perspective on interdisciplinary science." Ecosystems 1999(2): 292-295.
Naiman, R. J., et al. (2002). Dead Wood Dynamics in Stream Ecosystems. Proceedings of the Symposium on the Ecology and Management of Dead Wood in Western Forests, Reno, VN. W. F. Laudenslayer Jr., P. J. Shea, B. Valentine, E., C. P. Weatherspoon and T. E. Lisle. Albany, CA, USDA Forest Service: 23-48.
Large woody debris (LWD: > 10 cm diameter and > 1 m in length) in stream channels of
forested regions in North America is an essential ecosystem component. This article
summarizes information from the literature on the spatial and temporal variability of LWD
abundance, distribution and age; the processes of LWD delivery and elimination; and the
influence of LWD on material retention, habitat formation, and productivity of streams.
Examples are drawn mostly from the Pacific Coastal Ecoregion, but the fundamental
principles learned from this region have application over the broad, forested regions of the
Temperate Zone. Key studies show that LWD is an integral component of stream and river
corridors, positively affecting material retention, habitat formation, and productivity. It is
abundant in streams of all sizes flowing through forested regions, although the density and
form of accumulation changes with forest type, landscape topography, and flow regime. The
management implications of maintaining natural stream LWD dynamics are significant.
Overall, LWD is a fundamental component of streams in many western states. This suggests
that measures assuring a continued supply of LWD of appropriate size, volume, and species
composition are essential for maintaining the long-term integrity of stream and river corridors
 
Naiman, R. J., et al. (1992). Fundamental elements of ecologically healthy watersheds in the Pacific Northwest coastal ecoregion. Watershed Management: Balancing Sustainability and Environmental Change. R. J. Naiman. New York, Springer-Verlag: 127-188.
Naiman, R. J. and R. E. Bilby (1998). River Ecology and Management in the Pacific Coastal Ecoregion. River Ecology and Management: Lessons From the Pacific Coastal Ecoregion. New York, Springer-Verlag. Chapter 1: 1-9.
Naiman, R. J., et al. (1998). Approaches to Management at the Watershed Scale Approaches to Management at Larger Spatial Scales: 239-253.
Naiman, R. J., et al. (1988). "The potential importance of boundaries to fluvial ecosystems." Journal of the North American Benthological Society 7(4): 289-306.
Naiman, R. J., et al. (1992). General Principles of Classification and the Assessment of Conservation Potential in Rivers. River Conservation and Management. P. C. a. G. E. P. P. J. Boon, John Wiley & Sons Ltd: 93-123.
Naiman, R. J., et al. (1995). "Freshwater Ecosystem's and Their Management: A National Initiative " Science 270(27): 584-584.
Naiman, R. J., et al. (1987). "Longitudinal patterns of ecosystem processes and community structure in a subarctic river continuum." Ecology 68(5): 1139-1156.
Naiman, R. J., et al. (1991). The Application of Ecological Knowledge to River Management: 1-48.
Nakamura, F. (1986). "Analysis of Storage and Transport Processes Based on Age Distribution of Sediment." Transactions of the Japanese Geomorphical Union 7(3): 165-184.
Nakamura, F. (1986). "Chronological study on the torrential channel bed by the age distribution of deposits." Research Bulletin Coll. Exp. For. Hokkaido University 43(1): 1-26.
Nakamura, F., et al. (1995). "Sediment routing analyses based on chronological changes in hillslope and riverbed morphologies." Earth Surface Processes and Landforms 20: 333-346.
Recent studies of sediment delievery and budgets in the United States indicate that upland erosion rates at a given time may not explain contemporaneous sediment yields from a drainage basin. This suggests temporal discontinuities in sediment delivery associated with hillslope and channel storage processes. Integration of sediment production, storage and transport is essential to understand sediment routing in basins. We analysed each process chronologically using aerial photographs, monitoring data of sediment movement and annual tree-rings, and then compared estimated temporal changes in sediment production from hillslopes, floodplain disturbance areas and sediment transport in river channels. Toeslopes, floodplains and alluvial fans together contained 59% of sediment eroded from uplands over the last 30 years. Monitoring results of riverbed changes showed that the volume of stored sediment on floodplains decreased exponentially with succeeding floods. The age distribution of floodplain deposits reflected the distrubance history of a river channel, and followed an exponential decrease with age. The results of this study may have important implications for sediment control plans for watersheds in steep regions.
 
Nakamura, F. and F. J. Swanson (1993). "Effects of coarse woody debris on morphology and sediment storage of a mountain stream system in Western Oregon." Earth Surface Processes and Landforms 18: 43-61.
Nakamura, F., et al. (2000). "Disturbance regimes of stream and riparian systems: a disturbance-cascade perspective." Hydrological Processes 14: 2849-2860.
Nakamura, F., et al. (2000). "Disurbance regimes of stream and riparian systems -  a disturbance-cascade perspective." Hydrological Processes 14: 2849-2860.
Nakamura, S., et al. (1991). An Investigation of Environmental Improvements for Fish Production in Developed Japanese Rivers. American Fisheries Society Symposium 10.
Nanson, G. C. (1974). "Bedload and suspended-load transport in a small, steep, mountain stream." American Journal of Science 274: 471-486.
Nanson, G. C. (1981). "New evidence for scroll-bar formation on the Beatton River." Sedimentology 28: 889-891.
Nanson, G. C. and J. C. Croke (1992). "A genetic classification of floodplains." Geomorphology 4: 459-486.
Nanson, G. C. and E. J. Hickin (1986). "A statistical analysis of bank erosion and channel migration in western Canada." Geological Society of America Bulletin 97: 497-504.
Nash, D. (1987). A comparative review of limit equilibrium methods of stability analysis. Slope Stability. M. G. Anderson and K. S. Richards. New York, John Wiley and Sons: 11-75.
Nash, D. B. (1994). "Effective Sediment-Transporting Discharge from Magnitude-Frequency Analysis." J. of Geol. 102: 79-95.
The long-term effectiveness of a geomorphic event of a particular magnitude is the product of the effect of a single event of that magnitude times the frequency with which it recurs (magnitude-frequency analysis). If discharge frequency of a stream is assumed to be log-normally distributed and if sediment transport rate is assumed to be apower function of discharge, then transport effectiveness attains a maximum value at some discharge, termed here the "effective discharge", that moves the most sediment over the long term. The effective discharge and its recurrence interval may be predicted if both assumptions are correct. Analysis of 55 U.S. streams demonstrates that the observed and the predicted recurrence intervals of the effective discharge are in poor agreement. This lack of agreement results from the failure of the power function of discharge to predict adequately the sediment transport rate above some high-discharge threshold or thresholds. The observed recurrence interval of the effective discharge for the studied streams is highly variable, ranging from a week to several decades, calling into question the widely held notion that effective fluvial discharge occurs about once a year.
 
Nash, J. E., and J.V. Sutcliffe (1970). "River flow forecasting through conceptual models: part 1 - a discussion of principles." Journal of Hydrology 10: 282-290.
Naslas, G. D., et al. (1994). "Sediment, Nitrate, and Ammonium in Surface Runoff From Two Tahoe Basin Soil Types " Water Resources Bulletin 30(3): 409-417.
Nawa, R. K. and C. A. Frissell (1993). "Measuring Scour and Fill of Gravel Streambeds with Scour Chains and Sliding-Bead Monitors." North American Journal of Fisheries Management 13: 634-639.
NCASI (1985). NCASI - Catalog of Landslide Inventories for the Northwest. New York, NCASI.
NCASI (National Council for Air and Stream Improvement) ( 1996). Sediment Routing in Steep, Forested Mountain Basins. Workshop May 9-10, 1996.
NCASI (National Council of the Paper Industry for Air and Stream Improvement, I. (1981). Research on the effects of mass wasting of forest lands on water quality and the impact of sediment on aquatic organisms, National Council of the Paper Industry for Air and Stream Improvement, Inc.: 56.
NCASI (National Council of the Paper Industry for Air and Stream Improvement, I. (1983). Forest Management Practices and Natural Events - Their Relation to Landslides and Water Quality Protection, NCASI: 40.
NCASI (National Council of the Paper Industry for Air and Stream Improvement, I. (1985). Industry, State, and Federal Programs Designed to Assess and Protect Water Quality Associated with Managed Western Forests, National Council of the Paper Industry for Air and Stream Improvement, Inc.: 60.
NCASI (National Council of the Paper Industry for Air and Stream Improvement, I. (1986). Assessing Forest Practice Rules and Managing to Avoid Landslides and Cumulative Effects, National Council of the Paper Industry for Air and Stream Improvement, Inc.: 44.
Nearing, M. A., F.F.  Pruski, and M.R. O'Neal (2004). "Expected climate change impacts on soil erosion rates: a review." Journal of Soil and Water Conservation 59(1): 43-50.
Nearing, M. A., et al. (1999). "Soil erosion by surface water flow on a stony, semiarid hillslope." Earth Surface Processes and Landforms 24: 677-686.
Neary, D. G., K.C. Ryan, and L.F. Debano (2005). Wildland fire in ecosystems: effects of fire on soil and water. USDA Forest Service General Technical Report RMRS-GTR-42-vol. 4. Ogden, Utah, 250 pp.
Neely, M. K. and R. M. Rice (1990). "Estimating risk of debris slides after timber harvest in northwestern California." Bulletin of the Association of Engineering Geologists 27(3): 281-289.
Neeson, T. M., et al. (2008). "Factors affecting accuracy of stream channel slope estimates derived from geographical information systems." North American Journal of Fisheries Management 28: 722-732.
Stream channel slope is often a critical component of geographical information systems (GIS)-
based models of preferred habitat of aquatic species, but the relative accuracy of various GIS slope derivation
methods is not well established. We examined the accuracy of GIS-derived stream slopes for a set of stream
reaches in Idaho and Ohio. We also used the Ohio data set to examine in more detail the effects of stream
reach length, source of GIS file representing the stream path (‘‘shapefile’’), and digital elevation model (DEM)
resolution on the accuracy of GIS-derived slopes. The accuracy of GIS-derived slopes in the Ohio dataset
improved with increasing reach length, but we could not draw any consistent conclusions about the effect of
DEM resolution or shapefile. We present a simple and efficient method for improving GIS-derived slopes by
identifying probable elevation errors in the GIS-derived longitudinal stream profiles. The resulting derived
slopes were improved in all cases; the slopes derived by using a 10-m DEM and a manually traced stream
shapefile were the most accurate. We demonstrate how our results can be used to evaluate the feasibility of
implementing a GIS-based habitat model of sea lampreys Petromyzon marinus.
 
Nehlsen, W., et al. (1991). "Pacific salmon at the crossroads: stocks at risk from California, Oregon, Idaho, and Washington." Fisheries 16(2): 4-21.
Neill, C. R. (1985). Sediment balance considerations linking long-term transport and channel processes. Edmonton, Alberta, Northwest Hydraulic Consultants Ltd.: 1-17.
Nelitz, M. A., et al. (2007). "A science-based approach for identifying temperature-sensitive streams for rainbow trout." North American Journal of Fisheries Management 27: 405-424.
To regulate human-induced changes to fish habitat, resource managers commonly set standards
based on maximum allowable changes. For example, new legislation in British Columbia (BC), Canada, calls
for restrictions on harvesting of trees and related activities near temperature-sensitive streams. However,
methods for designating such streams are still evolving. Our objective was to help develop such methods by
(1) improving understanding of the temperature-dependent responses of fish and (2) devising improved
methods for estimating the effects of forestry-related activities on stream temperature as well as the chance of
exceeding upper temperature limits. Using previously published models, we found that for rainbow trout
Oncorhynchus mykiss particular increases in stream temperature led to different effects on juvenile growth
rate, egg survival rate, and resistance to mortality from diseases. In a separate analysis, to evaluate the chance
that cumulative forestry activities will increase stream temperature by various amounts, we compiled summer
temperature data for 104 streams in central BC that reflected different watershed features, contrasting summer
climates, and various levels of land use. A classification and regression tree analysis of a summer maximum
weekly average temperature (MWAT) index grouped streams into six categories as a function of watershed
size, watershed elevation, and air temperature. We then analyzed the remaining unexplained variation among
stream temperature indices using Bayesian regression. We found high probabilities that increases in road
density and the density of road crossings of streams within watersheds are associated with increases in
residual temperature. For instance, a Bayesian regression indicated a 6-in-10 chance that the MWAT in our
study area will increase by 1.258C for a road density of 2 km/km2 of watershed area and by 3.258C for a road
density of 4 km/km2. These analyses illustrate some possible ways to help designate temperature-sensitive
streams.
 
Nelson, E. J. a. D. B. B. (2002). "Sediment sources in an urbanizing, mixed landuse watershed." Journal of Hydrology 264: 51-68.
Nelson, K. C., et al. (2009). "Forecasting the combined effects of urbanization and climate change on stream ecosystems: from impacts to management options." Journal of Applied Ecology 46: 154-163.
1.
Streams collect runoff, heat, and sediment from their watersheds, making them highly vulnerable
to anthropogenic disturbances such as urbanization and climate change. Forecasting the effects of
these disturbances using process-based models is critical to identifying the form and magnitude of
likely impacts. Here, we integrate a new biotic model with four previously developed physical
models (downscaled climate projections, stream hydrology, geomorphology, and water temperature)
to predict how stream fish growth and reproduction will most probably respond to shifts in climate
and urbanization over the next several decades.
2.
The biotic submodel couples dynamics in fish populations and habitat suitability to predict fish
assemblage composition, based on readily available biotic information (preferences for habitat,
temperature, and food, and characteristics of spawning) and day-to-day variability in stream conditions.
3.
We illustrate the model using Piedmont headwater streams in the Chesapeake Bay watershed of
the USA, projecting ten scenarios: Baseline (low urbanization; no on-going construction; and
present-day climate); one Urbanization scenario (higher impervious surface, lower forest cover,
significant construction activity); four future climate change scenarios [Hadley CM3 and Parallel
Climate Models under medium-high (A2) and medium-low (B2) emissions scenarios]; and the same
four climate change scenarios plus Urbanization.
4.
Urbanization alone depressed growth or reproduction of 8 of 39 species, while climate change
alone depressed 22 to 29 species. Almost every recreationally important species (i.e. trouts, basses,
sunfishes) and six of the ten currently most common species were predicted to be significantly
stressed. The combined effect of climate change and urbanization on adult growth was sometimes
large compared to the effect of either stressor alone. Thus, the model predicts considerable change
in fish assemblage composition, including loss of diversity.
5.
Synthesis and applications
. The interaction of climate change and urban growth may entail
significant reconfiguring of headwater streams, including a loss of ecosystem structure and services,
which will be more costly than climate change alone. On local scales, stakeholders cannot control
climate drivers but they can mitigate stream impacts via careful land use. Therefore, to conserve
stream ecosystems, we recommend that proactive measures be taken to insure against species loss or severe population declines. Delays will inevitably exacerbate the impacts of both climate change
and urbanization on headwater systems.
 
Newcombe, C. P. and J. O. T. Jensen (1996). "Channel Suspended and fisheries: A synthesis for Quantitative Assessment of Risk and Impact " North American Journal of Fisheries Management 16(4): 693-727.
Newham, L. T. H., et al. (2003). "Sensitivity analysis for assessing the behaviour of a landscape-based sediment source and transport model." Environmental Modelling & Software 18(8-9): 741-751.
Newman, W. A., et al. (1993). Pleistocene Geology of the Boston Basin and its Adjacent Surroundings Field Trip Guidebook for the Northeastern United States: 1993 Boston GSA. J. T. Cheney and J. C. Hepburn. Amherst, Massachusetts: U1-U11.
Nicholas, A. P., et al. (1995). "Sediment slugs: large-scale fluctuations in fluvial sediment transport rates and storage volumes." Progress in Physical Geography 19(4): 500-519.
Nichols, D. R. and L. A. Yehle (1985). Volcanic Debris Flows, Copper River Basin, Alaska. IVth International, Conference and Field Workshop on Landslides, Tokyo, Japan.
Nichols, W. F., et al. (1998). "The Influence of Geomorphological Heterogeneity of Biodiversity: II. A Landscape Perspective." Conservation Biology 12(2): 371-379.
Nickelson, T. E. (1998). A Habitat-Based Assessment of Coho Salmon Production Potential and Spawner Escapement Needs for Oregon Coastal Streams. Portland, OR, Oregon Department of Fish and Wildlife: 15.
Nickelson, T. E. and P. W. Lawson (1998). "Population viability of coho salmon, Oncorhynchus kisutch, in Oregon coastal basins: application of a habitat-based life cycle model." Canadian Journal of Fisheries and Aquatic Science 55(11): 2383-2392.
Nickelson, T. E., et al. (1992). "Seasonal Changes in Habitat Use by Juvenile Coho Salmon (Oncorhynchus kisutch) in Oregon Coastal Streams " Canadian Journal of Fisheries and Aquatic Sciences 49: 783-789.
Nicolau, J. M. (2002). "Runoff generation and routing on artificial slopes in a Mediterranean-continental environment: the Teruel coalfield, Spain." Hydrological Processes 16(3): 631-647.
The aim of this study was to identify the mechanisms of runoff generation and routing and their controlling factors at the hillslope scale, on artificial slopes derived from surface coal mining reclamation in a Mediterranean-continental area. Rainfall and runoff at interrill and microcatchment scales were recorded for a year on two slopes with different substrata: topsoil cover and overburden cover. Runoff coefficient and runoff routing from interrill areas to microcatchment outlets were higher in the overburden substratum than in topsoil, and greater in the most developed rill network. Rainfall volume is the major parameter responsible for runoff response on overburden, suggesting that this substratum is very impermeable - at least during the main rainfall periods of the year (late spring and autumn) when the soil surface is sealed. In such conditions, most rainfall input is converted into runoff, regardless of its intensity. Results from artificial rainfall experiments, conducted 3 and 7 years after seeding, confirm the low infiltration capacity of overburden when sealed. The hydrological response shows great seasonal variability on the overburden slope in accordance with soil surface changes over the year. Rainfall volume and intensities (I-30, I-60) explain runoff at the inter-rill scale on the topsoil slope, where rainfall experiments demonstrated a typical Hortonian infiltration curve, However, no correlation was found at the microcatchment level, probably because of the loss of functionality of the only rill as ecological succession proceeded, The runoff generation mechanism on the topsoil slope is more homogeneous throughout the year. Runoff connectivity, defined as the ratio between runoff rates recorded at the rill network scale and those recorded at the interrill area scale in every rainfall event, was also greater on the rilled overburden slope, and in the most developed rill network. The dense rill networks of the overburden slope guarantee very effective runoff drainage, regardless of rainfall magnitude. Rills drain overland flow from interrill-sealed areas, reducing the opportunity of reinfiltration in areas not affected by siltation. Runoff generation and routing on topsoil slopes are controlled by grass cover and soil moisture content, whereas on overburden slopes rill network density and soil moisture content are the main controlling factors. Copyright (C) 2002 John Wiley Sons, Ltd.
 
Nidora, V., et al. (1996). "On channel network fractal properties: A case study of the Hutt River basin, New Zealand." Water Resources Research 32(11): 3375-3384.
Nielson, J. L., et al. (1994). "Thermally Stratified Pools and Their Use by Steelhead in Northwern California Stream " Transactions of the American Fisheries Society 123: 613-626.
Niemann, K. O. and D. E. Howes (1991). "Applicability of digital terrain models for slope stability assessment." ITC Journal 3: 127-137.
Nierenberg, T. R. and D. E. Hibbs (2000). "A characterization of unmanaged riparian areas in the central Coast Range of western Oregon." Forest Ecology and Management 129: 195-206.
As an approach to providing baseline information about riparian ecosystems, this study characterized the dominant riparian
vegetation along unmanaged streams in central Oregon Coast Range forests. We systematically sampled along various reaches
of nine Ærst- to fourth-order streams, all of which were subject to stand-replacing Æres ca. 145 year ago. The near-stream
communities were divided into different vegetative and/or topographic units called landscape units (LUs); LU1s were closest
to the stream, and LU2s were farther from the stream. Fifty-two percent of LU1s had no trees, and among all LUs, red alder
was the most frequently found tree species. Although in some cases sample plots simply fell between widely spaced trees, we
hypothesize that red alder originally dominated many of the current treeless patches and has since senesced to release
understory shrubs. With increased distance from the stream, hardwoods decreased in compositional importance relative to
conifers, not because hardwood frequency changed, but because conifer frequency increased. Our results suggest that the
competitive advantage of hardwoods and shrubs is the biggest limiting factor of conifer growth in the near-stream microenvironment
and that without vigorous competition, conifers have the potential to grow over more of the riparian area than that
on which they occurred in unmanaged areas. Calculations of disturbance frequency, based on ages of shade-intolerant stand
dominants, indicate that along the stream reaches we sampled, a minimum of 2.6 disturbances per stream km per century
occurred since the last stand-resetting Ære. Riparian areas are spatially and temporally diverse, and any riparian management
model should incorporate this variability.
 
Nierenberg, T. R. and D. E. Hibbs (2000). "A characterization of unmanaged riparian ares in the central Coast Range of western Oregon." Forest Ecology and Management 129: 289-306.
Nilsson, C., et al. (1989). "Patterns of plant species richness along riverbanks." Ecology 70: 77-84.
Nilsson, C., et al. (2003). "Ecological forecasting and the urbanization of stream ecosystems: challenges for economists, hydrologists, geomorphologists, and ecologists." Ecosystems 6: 659-674.
Nir, A. and S. Lewis (1975). "On tracer theory in geophysical systems in the steady and non-steady state. Part I." Tellus XXVII 4: 372-283.
Noel, J. R. and R. C. Sidle (1989). "A Program to Calculate Channel Scour and Fill " Water Resources Bulletin 25(4): 733-741.
Nolan, K. M. and R. J. Janda (1981). "Erosion and sediment transport in Pacific Rim Steeplands." International Association of Hydrological Sciences 132: 414-437.
Nolan, K. M. and R. J. Janda (1981). Use of short-term water and suspended-sediment discharge observations to assess impacts of logging on stream-sediment discharge in the Redwood Creek basin, northwestern California, U. S. A. Erosion and Sediment Transport in Pacific Rim Steeplands. Christchurch, I.A.H.S. Publ. No. 132: 415-437.
Nolan, K. M. and D. C. Marron (1985). "Contrast in stream-channel response to major storms in two mountainous areas of California." Geology 13: 135-138.
The generally catastrophic effects of major stroms on the geometry of intermediate- and high-order channels in northwestern California contrast with localized storm-related channel changes found in the Santa Cruz Mountains. Sediment delivery in northwestern California during stroms overwhelms transport capacities throughout the length of most intermediate- and high-order channels and causes long-lasting changes in geometry. In contrast, sediment delivery during storms in the Santa Cruz Mountains overwhelms transport capacities only in localized reaches. This contrast in channel behyavior is at least partially caused by contrasts in the style of landsliding found in the two areas.
 
Nolan, K. M. and D. C. Marron (1988). Stream-Channel Response to the Storm in the Santa Cruz Mountains. Landslides, Floods, and Marine Effects of the Storm of January 3-5, 1982, in the San Francisco Bay Region, California. U.S. Geological Survey Professional Paper 1434. S. D. Ellen and G. F. Wieczorek. Washington D.C., U.S. Geological Survey.
Nolan, K. M. and D. C. Marron (1990). History, Causes, and Significance of Changes in the Channel Geometry of Redwood Creek, Northwestern California, 1926-82. Geomorphic Processes and Aquatic Habitat in the Redwood Creek Basin, Northwestern California. U.S. Geological Survey Professional Paper 1454. K. M. Nolan, H. M. Kelsey and D. C. Marron. Washington D.C., U.S. Geological Survey.
Nonaka, E. and T. A. Spies (2005). "Historical range of variability in landscape structure: a simulation study in Oregon, USA." Ecological Applications 15(5): 1727-1746.
We estimated the historical range of variability (HRV) of forest landscape
structure under natural disturbance regimes at the scale of a physiographic province (Oregon
Coast Range, 2 million ha) and evaluated the similarity to HRV of current and future
landscapes under alternative management scenarios. We used a stochastic fire simulation
model to simulate presettlement landscapes and quantified the HRV of landscape structure
using multivariate analysis of landscape metrics. We examined two alternative policy scenarios
simulated by two spatially explicit simulation models: (1) current management policies
for 100 years into the future and (2) the wildfire scenario with no active management
until it reached the HRV.
The simulation results indicated that historical landscapes of the province were dynamic,
composed of patches of various sizes and age classes ranging from 0 to .800 years including
numerous, small, unburned forest islands. The current landscape was outside the HRV. The
landscape did not return to the HRV in the 100 years under either scenario, largely because
of lack of old-growth forests and the abundance of young forests. Under the current policy
scenario, development of landscape structure was limited by the spatial arrangement of
different ownerships and the highly contrasting management regimes among ownerships.
As a result, the vegetation pattern after 100 years reflected the ownership pattern. Surprisingly,
the wildfire scenario initially moved the landscape away from the HRV during
the first 100 years, after which it moved toward the HRV, but it required many more centuries
to reach it. Extensive forest management and human-caused fires in the 20th century have
left legacies on the landscape that could take centuries to be obliterated by wildfire.
Departure from the HRV can serve as an indicator of landscape conditions, but results
depend on scale and quantification of landscape heterogeneity. The direct application of
the concept of HRV to forest policy and management in large landscapes is often limited
since not all ownerships may have ecological goals and future climate change is anticipated.
Natural disturbance-based management at large scales would not show the projected effects
on landscape structure within a typical policy time frame in highly managed landscapes.
 
Nonaka, E., et al. (2007). "Historical range of variability in live and dead wood biomass: a regional-scale simulation study." Canadian Journal of Forest Research 37: 2349-2364.
North Carolina Division of Water Quality (2008). Mapping headwater streams: intermittent and perennial headwater stream model development and spatial application, North Carolina division of Water Quality: 54.
North Coast Regional Quality Control Board in cooperation with the California Department of Forestry (1993). Testing Indices of Cold Water Fish Habitat North Coast Regional Quality Control Board in cooperation with the California Department of Forestry: 55.
Northwest Indian Fisheries Commission, et al. (1993). Summary: 1992 Salmon and Steehad Stock Inventory (SASSI), Northwest Indian Fisheries Commission, Washington Department of Fisheries, Washington Department of Wildlife.
Northwest Power Planning Council (1999). Ocean conditions and Management of Columbia River Salmon. Ocean conditions and Management of Columbia River Salmon, Portland, Oregon.
Norton, B. C. a. J. (2004). Regionalisation of rainfall-runoff models. Complexity and Integrated Resources Management, Transactions of the 2nd Biennial Meeting of the International Environmental Modelling and Software Society, iEMSs: Manno, Switzerland, 2004., University of Osnabrück, Germany.
Noson, L. L., et al. (1988). Washington State Earthquake Hazards. Olympia, WA, Washington State Department of Natural Resources: 77.
Noss, R. F. e. (1999). The redwood forest: history, ecology, and conservation of the coastal redwoods, Island Press.
O'Brien, J. S., et al. (1993). "Two-dimensional water flood and mudflow simulation." Journal of Hydraulic Engineering 119(2).
O'Callaghan, J. F., and D.M. Mark (1984). "The extraction of drainage networks from digital elevation data." Computer Vision, Graphics, and Image Processing 28: 328-344.
O'Callaghan, J. F. and D. M. Mark (1984). "The extraction of drainage networks from digital elevation data." Computer Vision, Graphics, and Image Processing 28(323-344).
O'Connell, P. E. (1971). A simple stochastic modeling of Hurst's law. International Symposium on Mathematical Models in Hydrology, Warsaw, International Association of Scientific hydrology.
O'Connor Environmental, I. (2000). Garcia River Large Woody Debris Instream Monitoring. Healdsburg, CA, Mendocino County Resource Conservation District: 1-21.
O'Connor, J. E. and V. R. Baker (1992). "Magnitudes and implications of peak discharges from glacial Lake Missoula " Geological Society of America Bulletin 104: 267-279.
O'Connor, J. E., et al. (2003). "Floodplain and channel dynamics of the Quinault and Queets Rivers, Washington, USA." Geomorphology 51: 31-59.
O'Connor, M. and Anonymous (1996). "Effects of forest management on sediment load, channel morphology and large woody debris load in a Northwest Washington river." Abstracts with Programs - Geological Society of America 28(5): 97.
O'Connor, M. and Anonymous (2001). Quantitative assessment of erosion and sedimentation effects of forest management in Northern California.
O'Connor, M. and T. Cundy (1993). North Fork Calawah River Watershed Condition Survey: Landslide Inventory and Geomorphic Analysis of Mainstem Alluvial System. Part I. Landslide Inventory and Geomorphic Analysis of Mass Erosion.
O'Connor, M. D. Bedload Transport Processes in Steep Tributary Streams, Olympic Peninsula, Washington, USA.
O'Connor, M. D. (1993). Bedload transport processes in steep tributary streams, Olympic Peninsula, Washington, U.S.A. Advances in Hydro-Science and -Engineering, Volume 1. S. S. Y. Yang: 243-250.
O'Connor, M. D. (1993). Bedload Transport Processes in Steep Tributary Streams, Olympic Peninsula, Washington, USA.
O'Connor, M. D. (1994). Sediment transport in steep tributary streams and the influence of large organic debris. College of Forestry. Seattle, Washington, University of Washington: 317.
O'Connor, M. D. and R. R. Ziemer (1988). Coarse woody debris ecology in a second-growth Sequoia sempervirens forest stream. California Riparian Systems Conference, Davis, California.
OHEPA, O. E. P. A. (2002). Field evaluation manual for Ohio's primary headwater habitat streams. Final Version Columbus, Ohio, Division of Surface Water, Ohio Environmental Protection Agency.
Ohmann, J. L. and M. J. Gregory (2002). "Predictive mapping of forest composition and structure with direct gradient analysis and nearest neighbor imputation in coastal Oregon, USA." Canadian Journal of Forest Resources 32: 725-741.
Ohmann, J. L., et al. (2007). "Influence of environment, disturbance, and ownership on forest vegetatin of coastal Oregon." Ecological Applications 17(1): 18-33.
Information about how vegetation composition and structure vary quantitatively
and spatially with physical environment, disturbance history, and land ownership is
fundamental to regional conservation planning. However, current knowledge about patterns
of vegetation variability across large regions that is spatially explicit (i.e., mapped) tends to be
general and qualitative. We used spatial predictions from gradient models to examine the
influence of environment, disturbance, and ownership on patterns of forest vegetation
biodiversity across a large forested region, the 3-million-ha Oregon Coast Range (USA).
Gradients in tree species composition were strongly associated with environment, especially
climate, and insensitive to disturbance, probably because many dominant tree species are longlived
and persist throughout forest succession. In contrast, forest structure was strongly
correlated with disturbance and only weakly with environmental gradients. Although forest
structure differed among ownerships, differences were blurred by the presence of legacy trees
that originated prior to current forest management regimes. Our multi-ownership perspective
revealed biodiversity concerns and benefits not readily visible in single-ownership analyses,
and all ownerships contributed to regional biodiversity values. Federal lands provided most of
the late-successional and old-growth forest. State lands contained a range of forest ages and
structures, including diverse young forest, abundant legacy dead wood, and much of the highelevation
true fir forest. Nonindustrial private lands provided diverse young forest and the
greatest abundance of hardwood trees, including almost all of the foothill oak woodlands.
Forest industry lands encompassed much early-successional forest, most of the mixed
hardwood–conifer forest, and large amounts of legacy down wood. The detailed tree- and
species-level data in the maps revealed regional trends that would be masked in traditional
coarse-filter assessment. Although abundant, most early-successional forests originated after
timber harvest and lacked legacy live and dead trees important as habitat and for other
ecological functions. Many large-conifer forests that might be classified as old growth using a
generalized forest cover map lacked structural features of old growth such as multilayered
canopies or dead wood. Our findings suggest that regional conservation planning include all
ownerships and land allocations, as well as fine-scale elements of vegetation composition and
structure.
 
Ojha, C. S. P. a. S., V.P. (2002). ANN modelling in watershed hydrology. Mathematical models in large wastershed hydrology. Highlands Ranch, CO, Water Resources,: 67-88.
Okuda, S. and H. Suwa (?). "Some relationships between debris flow motion and micro-topography for the Kamikomihori Fan, North Japan Alps."
O'Laughlin, E. M. (1986). "Prediction of surface saturation zones in natural catchments by topographic analysis." Water Resources Research 22: 794-804.
Olden, J. D., N. L. Poff, and B. P. Bledsoe (2006). "Incorporating ecological knowledge into ecoinformatics: an example of modeling hierarchically structured aquatic communities with neural networks." Ecological Informatics 1: 33-42.
Olden, J. D. and D. A. Jackson (2002). "A comparison of statistical approaches for modelling fish species distributions." Freshwater Biology 47: 1976-1995.
1. The prediction of species distributions is of primary importance in ecology and
conservation biology. Statistical models play an important role in this regard; however,
researchers have little guidance when choosing between competing methodologies
because few comparative studies have been conducted.
2. We provide a comprehensive comparison of traditional and alternative techniques for
predicting species distributions using logistic regression analysis, linear discriminant
analysis, classification trees and artificial neural networks to model: (1) the presence ⁄
absence of 27 fish species as a function of habitat conditions in 286 temperate lakes located
in south-central Ontario, Canada and (2) simulated data sets exhibiting deterministic,
linear and non-linear species response curves.
3. Detailed evaluation of model predictive power showed that approaches produced
species models that differed in overall correct classification, specificity (i.e. ability to
correctly predict species absence) and sensitivity (i.e. ability to correctly predict
species presence) and in terms of which of the study lakes they correctly classified.
On average, neural networks outperformed the other modelling approaches, although
all approaches predicted species presence ⁄absence with moderate to excellent
success.
4. Based on simulated non-linear data, classification trees and neural networks greatly
outperformed traditional approaches, whereas all approaches exhibited similar correct
classification rates when modelling simulated linear data.
5. Detailed evaluation of model explanatory insight showed that the relative importance of
the habitat variables in the species models varied among the approaches, where habitat
variable importance was similar among approaches for some species and very different for
others.
6. In general, differences in predictive power (both correct classification rate and identity
of the lakes correctly classified) among the approaches corresponded with differences in
habitat variable importance, suggesting that non-linear modelling approaches (i.e.
classification trees and neural networks) are better able to capture and model complex,
non-linear patterns found in ecological data. The results from the comparisons using
simulated data further support this notion.
7. By employing parallel modelling approaches with the same set of data and focusing on
comparing multiple metrics of predictive performance, researchers can begin to choose
predictive models that not only provide the greatest predictive power, but also best fit the
proposed application
 
Olden, J. D., et al. (2002). "Predictive models of fish species distributions: a note on proper validation and chance predictions." Transactions of the American Fisheries Society 131: 329-336.
The prediction of species distributions is a
primary goal in the study, conservation, and management
of fisheries resources. Statistical models relating
patterns of species presence or absence to multiscale
habitat variables play an important role in this regard.
Researchers, however, have paid little attention to how
improper model validation and chance predictions can
result in unfounded confidence in the performance and
utility of such models. Using simulated and empirical
data for 40 lake and stream fish species, we demonstrate
that the commonly employed resubstitution approach to
model validation (in which the same data are used for
both model construction and prediction) produces highly
biased estimates of correct classification rates and consequently
an inaccurate perception of true model performance.
In contrast, a jackknife approach to validation
resulted in relatively unbiased estimates of model performance.
The estimated rates of model correct classification
are also shown to be substantially influenced by
species prevalence (i.e., the proportion of sites at which
a species is present), and often result in poorly performing
models being viewed as powerful. We use simulated
data to show how the expected frequency of chance predictions
from models is a function of species prevalence
and sample size. Finally, we use empirical data to illustrate
a randomization approach for assessing whether
the performances of the fish habitat models are statistically
greater than expectations based on chance predictions.
In summary, we urge researchers to employ
proper and defensible methodologies for model validation
and prediction assessment; failing to do so will
only add to the accumulating number of published species
habitat models in the fisheries literature that are of
limited use and reliability.
 
O'Leary, S. J. and R. L. Beschta (1981). "Bedload transport in an Oregon coast range stream." Water Resources Bulletin 17(5): 886-894.
Oliferov, A. N. (1970). "Transport of large rocks by mudflows." State Hydrologic Institute, Collection of Papers on Hydrology 9: 137-141.
Oliver, C. D. (1980/1981). "Forest development in North America following major disturbances." Forest Ecology and Management 3: 153-168.
Oliver, C. D., et al. (1984). "Disturbance patterns and forest development in a recently deglaciated valley in the northwestern Cascade Range of Washington, U.S.A." Can.J. For. Res. 15: 221-232.
Oliver, C. D., et al. (1992). Integrating Management Tools, Ecological Knowledge, and Silviculture. Watershed Management: Balancing Sustainability and Environmental Change. New York, Springer-Verlag. chapter 13: 361-382.
Olivera, F. and D. Maidment (1999). "Geographic information systems (GIS)-based spatially distributed model for runoff routing." Water Resources Research 35(4): 1155-1164.
O'Loughlin, C. (1974). "The effect of timber removal on the stability of forest soils." Journal of Hydrology 13(2): 121-133.
O'Loughlin, C. and A. Watson (1979). "Root-wood strength deterioration in Radiata pine after clearfelling." N. Z. J. For. Sci. 9(3): 284-293.
O'Loughlin, C. L. (1974). "The effect of timber removal on the stability of forest soils." Journal Hydrology, N. Z. 13(2): 121-123.
Olson, C. M. and B. Orr Combining tree growth, fish and wildlife habitat, mass wasting, sedimentation, and hydrologic models in decision analysis and long-term forest land planning. Forest Ecology and Management 114 (2/3), 339-348 pp.;   13 ref.; 1999. B. P. Oswald.
This paper describes the ecosystem management planning approach that has been successfully implemented for landscape-scale sustained yield plans (SYPs) and is currently being used for multispecies habitat conservation plans (HCPs). The method was developed for large forest landowners in California (USA). A key aspect of the approach is its emphasis on preventative management, which involves matching land use activities to the capabilities of the landscape such that forest management can occur in an ecologically sustainable manner. The planning process targets a desired future condition using a geographical information system (GIS) to store and analyse site-specific resource data. Monitoring and adaptive management will be needed to evaluate the effectiveness of the selected plan, and to continue to improve the land management plan through time. The components of the planning model are: a digital terrain model, base map features, vegetation classification, timber growth and yield data (using the FREIGHTS simulation model), wildlife habitat classification, yarding methods, surface erosion hazard estimation and mapping, a hillslope sediment relative hazard rating system, stream channel network classification, fish-bearing channel type classification, a fish habitat and channel sensitivity model, a watershed relative risk index system, timber and regional economic vitality data, and decision analysis.
 
Olson, C. M. and B. Orr (1999). Combining Tree Growth, Fish and Wildlife Habitat, Mass Wasting, Sedimentation, and Hydrologic Models in Decision Analysis and Long-Term Forest Land Planning (DRAFT): 18.
Olson, S. A. and M. C. Brouillette (2006). A logistic regression equation for estimating the probability of a stream in Vermont having intermittent flow: U.S. Geological Survey Scientific Investigations Report 2006-5217, U.S. Geological Survey: 15.
A logistic regression equation was developed for estimating the probability of a stream flowing intermittently at unregulated, rural stream sites in Vermont. These determinations can be used for a wide variety of regulatory and planning efforts at the Federal, State, regional, county and town levels, including such applications as assessing fish and wildlife habitats, wetlands classifications, recreational opportunities, water-supply potential, waste-assimilation capacities, and sediment transport. The equation will be used to create a derived product for the Vermont Hydrography Dataset having the streamflow characteristic of “intermittent” or “perennial.” The Vermont Hydrography Dataset is Vermont’s implementation of the National Hydrography Dataset and was created at a scale of 1:5,000 based on statewide digital orthophotos.
The equation was developed by relating field-verified perennial or intermittent status of a stream site during normal summer low-streamflow conditions in the summer of 2005 to selected basin characteristics of naturally flowing streams in Vermont. The database used to develop the equation included 682 stream sites with drainage areas ranging from 0.05 to 5.0 square miles. When the 682 sites were observed, 126 were intermittent (had no flow at the time of the observation) and 556 were perennial (had flowing water at the time of the observation).
The results of the logistic regression analysis indicate that the probability of a stream having intermittent flow in Vermont is a function of drainage area, elevation of the site, the ratio of basin relief to basin perimeter, and the areal percentage of well- and moderately well-drained soils in the basin. Using a probability cutpoint (a lower probability indicates the site has perennial flow and a higher probability indicates the site has intermittent flow) of 0.5, the logistic regression equation correctly predicted the perennial or intermittent status of 116 test sites 85 percent of the time.
 
Omura, H. (1984). The resistance index map to forecast quick shallow landslide in Japan. Interpraevent 1984, Villach.
Omura, H. (1987). Erosion Phase at South Region of Tienshan Range, China. Special report on mountaineering expedition of Shizuoka University in Xin-jiang Province in 1987: 125-168.
Omura, H. (1987). Progressive failure model on the creep curve of landslide. 5th ECFL Fifth International Conference and Field Workshop on Landslides, Christchurch, New Zealand.
Omura, H. (1988). "Resistance index of various forest in the the gamma distribution model applied to forcast rapid shallow landslide." 152-160.
Omura, H. (1992). Estimation of Probable Maximum Daily Rainfall in Climatic Change Era ---Recent Increase Tendency of Maximum Daily Rainfall at Mt. Amagi---. Internationales Symposion Interpraevent 1992, Bern.
Omura, H. (1992). "Field excursion of soil and water conservation, Taiwan in R.O.C, April 5-14, 1992." Journal of Chinese Soil and Water Conservation 23(1): 85-97.
Omura, H. (1995). Fractal dimension analysis on spatial distribution of shallow landslides triggered by heavy rainfall. Internation Sabo Symposium, Tokyo, Japan.
Omura, H. (1999). Red Soil Problem in Okinawa Island (DRAFT). Soil Conservation in Large-Scale Land Use, Brattislava.
Omura, H. (2000). Risk assessment in high altitude forest (DRAFT), Kyushu University: 4.
Omura, H. (2002). Evolution of Mitigation Strategy of Debris Flow Disaster in Japan. First International Conference on Debris-Flow Disaster Mitigation Strategy.
Omura, H. and K. Hara (1984). A model experimental study on the fence effects of two trees against debris flow. Symposium on Effects of Land Use on Erosion and Slope Stability, University of Hawaii.
Omura, H. and D. Hicks (1992). Landslides and debris flows at Shimoda, Japan, DSIR Land Resources
Department of Scientific and Industrial Research: 5-11.
Omura, H. and D. Hicks (1994). Fractal dimension analysis on shallow landslide and channel net system in small watershed, River Ochiai. Proceedings of the International Symposium on Forest Hydrology, Tokyo, Japan.
Omura, H. and Y. Marumo (1986). An experimental study of the fence effects of protection forest on the interception of shallow mass movement. Shizuoka, Univeristy, Shizuoka University: 9.
Omura, H. and S. Mochizuki (?). "Forest distribution as the erosion indicator at large scale landslide."
Omura, H., et al. (?). "Environmental Adjustment Planning for Torrent - A Case Study of R. Abekawa, Japan." 14-20.
Omura, H., et al. (1995). Fractal dimension analysis on a ridge system in large slope collapses in the Akaishi Mountain Range, Japan. Proceedings of the International Sabo Symposium, Tokyo, Japan, August 1995.
Omura, H. and T. Sakurai (1996). Simulation by small scale model on deep seated landslide related with double ridges. Interpraevent 1996, Garmisch-Partenkirchen.
Onesti, L. J. and T. K. Miller (1978). "Topological Classifications of Drainage Networks: An Evaluation." Water Resources Research 14(1): 144-148.
Oniki, M. and T. Mizuyama (1987). Procedures of Scaled Model Experiments on Mudflows. 31st Hydraulic Lecture Meeting.
Oostwoud Wijdenes, D. J. and P. Ergenzinger "Erosion and sediment transport on steep marly hillslopes, Draix, Haute-Provence, France: an experimental field study."
Despite a thin regolith layer, measured sediment yields indicate denudation rates of more than 10 to 20 mm yr super(-1) on strongly denuded slopes in black marl (terres noires), near Draix, Alpes de Haute-Provence. Overland-flow and rainfall-simulation experiments were conducted on three slopes in the Draix Erosional Research Basin to determine the nature and intensity of hillslope erosion processes. The experiments showed that during a single storm, sediment transport includes water-solid mixtures with a wide range of concentrations, from highly charged rill flow (up to 400 g l super(-1)) to hyperconcentrated flow (up to 800 g l super(-1)) to debris flows (up to 1400 g l super(-1)). Miniature debris flows (MDFs), strongly controlled sediment transport as long as the sediment store was not depleted. Slope surfaces were completely reworked in the winter and the regolith layer was rebuilt. Temperature records from nine recent winter periods indicate that conditions are favourable for soil frost processes.
 
Oregon Department of Forestry (2005). Oregon Administrative Rules Guidance Manuals, Division 635, Water Protection Rules: Purpose, Goals, Classification and Riparian Management Areas. Salem, OR, Oregon Department of Forestry.
Oregon Department of Forestry (2006). Oregon Administrative Rules, Division 623: Shallow, rapidly moving landslides and public safety. Salem, Oregon, Oregon Department of Forestry.
Oregon Forest Industries Council (1993). 1993 Aquatic Inventory Pilot Project, Draft Stream Report for South Fork Gate Creek.
Oregon State University (1987). TEMP86 User's Guide, Oregon State, University: 15.
Oregon State University (1996). Summary of precipitation for central Oregon Coast Range, October 1995 to April 1996. Corvallis, Oregon State University.
Oreskes, N., et al. (1994). "Verification, validation, and confirmation of numberical models in the earth sciences." Science 263: 641-646.
Orlandini, S. and R. Rosso (1998). "Parameterization of stream channel geometry in the distributed modeling of catchment dynamics." Water Resources Research 34(8): 1971-1985.
Orme, A. R. (1989). "The Nature and Rate of Alluvial Fan Aggradation in a Humid Temperate Environment, Northwest Washington." Physical Geography 10(2): 131-146.
Orme, A. R. (1990). Recurrence of Debris Production under Coniferous Forest, Cascade Foothills, Northwest United States. Vegetation and Erosion. J. B. Thornes, John Wiley & Sons Ltd. chapter 6: 67-84.
Orr, E. L., et al. (1992). Geology of Oregon. Dubuque, Iowa, Kendall/Hunt.
Orsborn, J. F. and S. C. Ralph (1994). An Aquatic Resource Assessment of the Dungeness River System: Phase II - Physical Channel Analysis, Hydrology and Hydraulics and Phase II - Fisheries Habitat Survey (DRAFT), prepared for The Jamestown S'Klallam Tribe and The Quilcene Ranger District: 1 - 8.
Osborne, L. L. and M. J. Wiley (1992). "Influence of tributary spatial position on the structure of warm water fish assembledges." Canadian Journal of Fisheries and Aquatic Science 49: 671-681.
Osborne, L. L. and M. J. Wiley (1992). " Influence of tributary spatial position on the structure of warmwater fish communities." Canadian Journal of Fisheries and Aquatic Sciences 49: 671-681.
Osidele, O. O., et al. (2003). "Coping with uncertainty; a case study in sediment transport and nutrient load analysis." Journal of Water Resources Planning and Management 129(4): 345-355.
We present a computational approach for identifying the significance of uncertainty in assessing the consequences of sediment and nutrient transport in a section of the Chattahoochee River south of Lake Lanier, as it passes through Atlanta, Georgia. Specifically, our analysis aims at identifying the key control and management actions, and the key scientific uncertainties about the fluvial system, that govern the attainment of a set of water quality objectives for the downstream boundary of the study area. To this end, we present a computational framework that integrates a recently developed sediment-nutrient dynamics model with a Monte Carlo-based methodology for model uncertainty evaluation. Our results suggest that, in general, reliable execution of controls and management actions is more crucial to meeting the target values for flow, sediment, and phosphorus concentration, than the scientific uncertainties associated with fluvial processes within the river channel. We also discuss the potential utility of our framework for accommodating the various science-and policy-derived uncertainties in the total maximum daily load process.
 
Osterkamp, W. R. a. T. J. T. (1997). "Geomorphic considerations for erosion prediction." Environmental Geology 29(3/4): 152-157.
Ott, R. L., and M. Longnecker (2001). An introduction to statistical methods and data analysis. Pacific Grove, California, 1152 pp., Duxbury.
Ouchi, S. (1987). Tombi landslide and its impact on the Joganji River, Japan. Proceedings of the Erosion and Sedimentation in the Pacific Rim. B. e. al, IAHS Publ. 165: 135-136.
Ouimet, W. B., et al. (2008). "Epigenetic gorges in fluvial landscapes." Earth Surface Process and Landforms 33(13): 1993-2009.
Ouimet, W. B., et al. (2008). "Epigenetic gorges in fluvial landscapes." Earth Surface Process and Landforms 33: 1993-2009.
Ouimet, W. B., et al. (2007). "The infleunce of large landslides on river incision in a transient landscape: Eastern margin of the Tibetan Plateau (Sichuan, China)." Geological Society of America Bulletin 119: 1462-1476.
Overby, S., et al. (1995). Sediment and nutrient regime from a central Arizona chaparral watershed. Hydrology and Water Resources in Arizona and the Southwest, vol.22-25: 61-67.
Overton, C. K., et al. (1997). R1/R4 (Northern/ Intermountain Regions) Fish and Fish Habitat Standard Inventory Procedures Handbook Ogden, UT, US Forest Service.
Pabst, R. J., et al. (2008). "Calibrating and testing a gap model for simulating forest management in the Oregon Coast Range." Forest Ecology and Management 256: 958-972.
The complex mix of economic and ecological objectives facing today’s forest managers necessitates the
development of growth models with a capacity for simulating a wide range of forest conditions while
producing outputs useful for economic analyses. We calibrated the gap model ZELIG to simulate standlevel
forest development in the Oregon Coast Range as part of a landscape-scale assessment of different
forest management strategies. Our goal was to incorporate the predictive ability of an empirical model
with the flexibility of a forest succession model. We emphasized the development of commercial-aged
stands of Douglas-fir, the dominant tree species in the study area and primary source of timber. In
addition, we judged that the ecological approach of ZELIG would be robust to the variety of other forest
conditions and practices encountered in the Coast Range, including mixed-species stands, small-scale
gap formation, innovative silvicultural methods, and reserve areas where forests grow unmanaged for
long periods of time. We parameterized the model to distinguish forest development among two
ecoregions, three forest types and two site productivity classes using three data sources:
chronosequences of forest inventory data, long-term research data, and simulations from an empirical
growth-and-yield model. The calibrated model was tested with independent, long-term measurements
from 11 Douglas-fir plots (6 unthinned, 5 thinned), 3 spruce-hemlock plots, and 1 red alder plot. ZELIG
closely approximated developmental trajectories of basal area and large trees in the Douglas-fir plots.
Differences between simulated and observed conifer basal area for these plots ranged from  2.6 to
2.4 m2/ha; differences in the number of trees/ha  50 cm dbh ranged from  8.8 to 7.3 tph. Achieving
these results required the use of a diameter-growth multiplier, suggesting some underlying constraints
on tree growth such as the temperature response function. ZELIG also tended to overestimate
regeneration of shade-tolerant trees and underestimate total tree density (i.e., higher rates of tree
mortality). However, comparisons with the chronosequences of forest inventory data indicated that the
simulated data are within the range of variability observed in the Coast Range. Further exploration and
improvement of ZELIG is warranted in three key areas: (1) modeling rapid rates of conifer tree growth
without the need for a diameter-growth multiplier; (2) understanding and remedying rates of tree
mortality that were higher than those observed in the independent data; and (3) improving the tree
regeneration module to account for competition with understory vegetation.
 
Pabst, R. J. and T. A. Spies (1999). "Structure and composition of unmanaged riparian forests in the coastal mountains of Oregon, U.S.A." Canadian Journal of Forest Research 29: 1557-1573.
We characterized the structure and composition of unmanaged riparian forests in three river basins in
Oregon’s coastal mountains. Our objective was to evaluate stand attributes at three spatial scales: streamside (site),
drainage network (stream order), and basin (subregion). Data on basal area, species composition, snag density, canopy
cover, and tree regeneration were collected along transects at 124 sites. Conifer basal area increased with distance from
stream, a trend similar among subregions, and was highest at sites along first-order streams. Hardwood basal area was
relatively constant with distance from stream and was proportionally higher at sites along second- and third-order
streams than at sites along first-order streams. Conifer and hardwood tree regeneration occurred infrequently and varied
by topographic position, stream order, and subregion. Conifer regeneration was associated with basal area of shadetolerant
conifers and appeared to be limited by shrub competition. The unmanaged forests we studied were
characterized by a patchy mosaic of structure and composition. Hardwoods and shrubs were major components of the
near-stream environment in these forests, whereas dominance of conifers was limited to hillslopes. It appears that finescale
patterns associated with proximity to the stream are influenced by coarser scale factors such as valley-floor width
and climate.
 
Pabst, R. J. and T. A. Spies (2001). "Ten years of vegetation succession on a debris-flow deposit in Oregon." Journal of American Water Resources Association 37(6): 1693-1708.
Pack, R. T., et al. (1998). The SINMAP Approach to Terrain Stability Mapping. Eight Congress of the International Association of Engineering Geology, Vancouver, British Columbia.
Pack, R. T., et al. (1998). The SINMAP approach to Terrain Stability Mapping. 8th Congress of the International Association of Engineering Geology. Vancouver, British Columbia, Canadia.
Pain, C. F. and P. L. Hosking (1970). "The Movement of Sediment in a Channel in Relation to Magnitude and Frequency Concepts - A New Zealand Example." Earth Science Journal 4(1): 17-23.
Paine, A. D. M. (1985). "'Ergodic' reasoning in geomorphology: time for a review of the term?" Progress in Physical Geography 9: 1-15.
Paintal, A. S. (1971). "A stochastic model of bed load transport." Journal of Hydraulic Research(4): 527-554.
Palik, B., et al. (1998). "Geomorphic variation in riparian tree mortality and stream coarse woody debris recruitment from record flooding in a coastal plain stream." Ecoscience 5(4): 551-560.
Large floods are an important process controlling the structure and function of stream ecosystems. One of the ways
floods affect streams is through the recruitment of coarse woody debris from stream-side forests. Stream valley geomorphology
may mediate this interaction by altering flood velocity, depth, and duration. Little research has examined how floods and
geomosphic features interact to control debris recruitment from riparian forests. With this in mind, we examined debris
recruitment resulting from tree mortality during a record flood in a Georgia (U.S.A.) stream. We quantified debris characteristics
as related to riparian geomorphology, and we examined the influence of floods on the structure of strearn-side forests.
The flood killed, and recruited into the stream debris pool, an average of 22 treeslkm. Variation in recruitment was related to
geomorphology; mortality was highest in reaches having nanow valleys and high elevations of riparian landforms, while it
was lowest in reaches having wide valleys and low landform elevations. Species differed in probability of mortality; three
taxa, out of 47, contributed 75% of new debris to the stream. The structure of stream-side forests reflected the influence of
floods on tree mostality; forests along constrained reaches lack small individuals. Our results suggest that constrained reaches
are the primary sources of debris during large floods, while unconstrained reaches function as debris sinks. Debris characteristics
may be linked to floods through tree demography. Specifically, large floods have the potential to limit future recruitment
of larger-sized susceptible species into the stream, by limiting the number of small trees that grow into the canopy.
Keywords: coarse woody debris, riparian forest, coastal plain, flooding, tree mortality
 
Palladino, D. J. and R. B. Peck (1972). "Slope failures in an onverconsolidated clay, Seattle, Washington." Geotechnique 4: 563-595.
Palmer, M. A., et al. (1997). "Biodiversity and ecosystem processes in freshwater sediments." Ambio 26(8): 571-577.
Palmer, M. A. and L. N. Poff (1997). "Heterogeneity in Streams. The influence of environmental heterogeneity on patterns and processes in streams." Journal of the North American Benthological Society 16(1): 169-173.
Palmer, M. A., et al. (2002). "Ecological forecasting: hurdles and hopes for threatened landscapes." in review National Academy of Sciences: 15.
Palmer, R. N. and R. M. Snyder (1985). "Effects of Instream Flow Requirements on Water Supply Reliability." Water  Resources Research 21(4): 439-446.
Palmieri, A., F. Shah and A. Dinar (2001). "Economics of reservoir sedimentation and sustainable management of dams." Journal of Environmental Management 61: 149-163.
Palmquist, R. (2005). Type N Stream Demarcation Study, Phase 1: Pilot Results, CMER: 61.
Pannkuk, C. D. and P. R. Robichaud (2003). "Effectiveness of needle cast at reducing erosion after forest fires." Water Resources Research 39(12).
[1] Needle cast from partially burnt conifer trees commonly occurs after forest fires. The effectiveness of needles in reducing soil erosion was investigated in this study. Two needle types, ponderosa pine and Douglas fir needles, were used at four different cover amounts ( 0, 15, 40, and 70 percent) on granitic and volcanic derived soils. Simulated rainfall was used to examine interrill erosion; added inflow was used to determine rill erosion in a laboratory setting. After a series of "runs,'' data showed that sediment delivery was greater for the granitic soil compared with the volcanic soil. Douglas fir needles were more effective at reducing interrill erosion compared with the ponderosa pine needles. Ponderosa pine needles, because of their shape and being bundled together, often caused minidebris dams to form. The minidebris dams formed by ponderosa pine needles reduce flow within the rill, resulting in less rill erosion than the Douglas fir needles. A 50 percent cover of Douglas fir needles reduced interrill erosion by 80 percent and rill erosion 20 by percent. A 50 percent cover of ponderosa pine needles reduced interrill erosion by 60 percent and rill erosion by 40 percent. We also compared the effectiveness of using stream power, rather than shear stress, to model rill erosion. Stream power was a better predictor of sediment load than shear stress. Rill detachment rates based on stream power decreased with increasing cover for both needle types. These results challenge the use of shear stress detachment rates in current erosion models and provide insight into the use of stream power detachment rates.
 
Paola, C. (1997). "When streams collide." Nature 387: 232-233.
Paola, C. and R. Seal (1995). "Grain size patchiness as a cause of selective deposition and downstream fining." Water Resources Research 31(5): 1395-1407.
Pappenburger, F., and K.J. Beven (2006). "Ignorance is bliss: or seven reasons not to use uncertainty analysis." Water Resources Research 42: W05302, doi: 05310.01029/02005WR004820, 002006.
Paris, S., et al. (2007). A gentle introduction to bilateral filtering and its applications. International Conference on Computer Graphics and Interactive Techniques, ACM.
This course reviews the wealth of work related to bilateral filtering. The bilateral filter is ubiquitous in computational photography applications. It is increasingly common in computer graphics research papers but no single reference summarizes its properties and applications. This course provides a graphical, intuitive introduction to bilateral filtering, and a practical guide for image editing, tone-mapping, video processing and more.
 
Parish, R., et al. (2010). "Snag longevity of Douglas-fir, western hemlock, and western redcedar from permanent sample plots in coastal British Columbia." Forest Ecology and Management 259: 633-640.
Snags are important both as structural components and as animal habitat in forests, but abundance is
often low and their dynamics poorly understood in young, managed stands. Using a large data set of
19,622 snags from permanent plots in second-growth forests of coastal British Columbia, we modeled
snag longevity (time from tree mortality to snag fall) for three species: Douglas-fir (Pseudotsuga
menziesii), western hemlock (Tsuga heterophylla), and western redcedar (Thuja plicata). Snag longevity
was strongly related to species and snag size (diameter): the median snag longevity was 16 years for
Douglas-fir, 11 years for hemlock and 5 years for redcedar. Western redcedar was predominantly in the
subcanopy and its rapid fall rate was related to the small size of its snags. In addition to diameter, other
attributes (height to diameter ratio, height, and live crown ratio before death) contributed significantly
to models for one or two of the species. However, site level variables did not contribute significantly to
any of themodels. Snags greater than 50 cm diameter, especially Douglas-fir snags, have the potential for
persistence well beyond 20 years in these second-growth forests, and could be important for wildlife.
 
Park, C. (1977). "World-wide variations in hydraulic geometry exponents of stream channels: an analysis and some observations." Journal of Hydrology 33(1977): 133-146.
Park, C. C. (1995). Channel Cross-sectional Change. Changing River Channels. A. Gurnell and G. Petts, John Wiley & Sons, Ltd.: 117-145.
Parker, C., C. Thorne, R. Bingner, R. Wells, D. Wilcox (2007). "Automated Mapping of Potential for Ephemeral Gully Formation in Agricultural Watersheds." National Sedimentation Laboratory, Publication No. 56.
Parker, G. (1990). The "ACRONYM" series of PASCAL programs for computing bedload transport in gravel rivers. Minneapolis, Minnesota, University of Minnesota, St. Anthony Falls Hydraulic Laboratory.
Parker, G. (1990). "Surface-based bedload transport relation for gravel rivers " Journal of Hydraulic Research 28(4): 417-436.
Bedload transport in gravel-bed rivers is accomplished by means of the mobilization of grains exposed on the bed surface. this mobilization is due to the action of fluid forces on the exposed grains. Substrate particles can participate in the bedload only to the extent that local or global scour results in their exposure on the surface. It follows that a calculation of the bedload transport rate of mixutres should be based on the availability of each size range in the surface layer. Herein an existing empirical substrate-based bedload relation for gravel mixutres, developed soley with reference to field data, is transformed into a surface-based relation.
 
Parker, G. (1991). "Selective sorting and abrasion of river gravel. II. Applications." Journal of Hydraulic Engineering 117(2): 150-171.
Most gravel rivers show a tendency for characteristic grain size to decrease in the downstream direction over scales of tens or hundreds of kilometers. It has been surmised that this downstream fining is due to some combination of selective sorting, by which finer grains are preferentially transported downstream; and abrasion, by which individual particles are reduced in size. Here the framework for the simultaneous treatment of both phenomena, developed in a companion paper, is used to analyze several cases of interest. In particular, wavelike aggradational profiles of permanent form are considered. An application to the Red Deer River, in Alberta, Canada, as well as several hypothetical cases, suggests the following result. In the case of quartzite, selective sorting controls downstream fining; in the case of limestone, abrasion and selective sorting are of roughly equal importance; and in the case of a mixture of quartz and limestone, abrasion ceases to be important beyond some characteristic length scale required to grind the gravel-sized limestone out of existence.
 
Parker, G. and P. C. Klingeman (1982). "On why gravel bed streams are paved." Water Resources Research 18(5): 1409-1423.
Bedload transport in poorly sorted gravel bed streams is considered. Bedload and typical bed material (subpavement) size distributions are observed to be similar; it follows that the coarse half o the subpavement moves through a reach at a rate near that of the fine half. Since coarser grains are intrinsically less mobile than fine grains, it follows that some mechanism must act to nearly equalize mobility. It is hypothesized that the pavement seen in gravel bed streams at low flow is in fact in place during typical transport events capable of moving all available sizes. This pavement can provide the equalizing mechanism by exposing proportionally more coarse grains to the flow. Field data are used to quantify this concept and to develop a predictive relation for river pavement. The model indicates that pavement should be absent in most sand bed streams, in agreement with observations.
 
Parker, G., et al. (1982). "Bedload and size distribution in paved gravel-bed streams." Journal of the Hydraulics Division, Proceedings of the American Society of Civil Engineers 108(NY4): 544-571.
Parker, G. and A. W. Peterson (1980). "Bar Resistance of Gravel-Bed Streams " Journal of the Hydraulics Division 10: 1559-1575.
Parker, G. and A. J. Sutherland (1990). "Fluvial armor." Journal of Hydraulic Research 28(5): 529-544.
Mobile armor layers which form during bed load transport of non-uniform sediments are shown to be closely related to the static armor layers that form by selective erosion as a result of the action of clear water flows. Two previously published numerical models of the transport of non-uniform sediments are used as a basis for the discussion. Each model is inverted, so that it predicts surface layer compositions given the imposed flow conditions and bed load transport rate and composition. Static armor is then obtained in the limit of vanishing sediment transport, under the constraint that the bedload and substrate size distributions are identical. Good agreement is obtained between measured and calculated static armor compositions over the entire grain size distribution.
 
Parks, B. and R. J. Madison (1985). Estimation of Selected Flow and Water-Quality Characteristics of Alaskan Streams, Water-Resources Investigations Report 84-4247. U. S. G. Survey. Anchorage, Alaska: 64.
Parsons, A. J., et al. (1993). "Tracing sediment movement in interrill overland flow on a semi-arid grassland hillslope using magnetic susceptibility." Earth Surface Processes and Landforms [EARTH SURF. PROCESS. LANDFORMS] 18(8): 721-732.
Experiments were undertaken to determine the feasibility of tracing sediment movement in interrill overland flow. Crushed magnetite was introduced as a source-line 10 cm wide by 8 m long on a runoff plot 18 m wide by 29 m long located in southern Arizona. Initial magnetic susceptibilities along this source line, and along three transects located 0.25, 2.95 and 5 m downslope of the source-line, were measured. Movement of the magnetite in response to three rainfall simulation experiments was monitored. During the first two experiments, overland flow discharge was sampled at miniature flumes located along two cross sections on the plot downslope of the source-line, and at a supercritical flume at the plot outlet. Magnetic susceptibilities along the source-line and transects were measured after all three experiments. Results show that the magnetite moves very early in the experiments and that it reaches one of the flumes 2 m downslope of the source-line in 3 min. Most of the tracer moves a very short distance: 29.7 per cent is deposited within 25 cm of the source-line and only 2.2 per cent is deposited 2.95 m away. The deposition rate appears to decrease exponentially away from the source-line. Very little magnetite is recorded in the flow through the miniature flumes: in general it makes up less than 1 per cent of the total sediment load. No temporal pattern in these percentages is observed. Magnetite appears to be an effective tracer of sediment movement in interrill overland flow, though its higher density than natural soil may affect its detachment and transport.
 
Passalacqua, P., et al. (in review). "A geometric framework for channel network extraction from LiDAR: nonlinear diffusion and geodesic paths." Journal of  Geophysical Research.
Patton, P. C. Geomorphic Response of Streams to Floods in the Glaciated Terrain of Southern New England.
Patton, P. C. (1988). Drainage Basin Morphometry and Floods. Flood Geomorphology. V. R. Baker, R. C. Kochel and P. C. Patton. New York, Wiley: 51-64.
Paudel, P. P., et al. (2002). "A study of shallow landslides and properties of soil on the twigs of Chamaecyparis obtuse." Kysushu J. For. Res. 55: 81-85.
Paulsen, C. M., et al. (2007). "Measure twice, estimate once: Pacific salmon population viability analysis for highly variable populations." Transactions of the American Fisheries Society 136: 346-364.
Because many stocks of Pacific salmon Oncorhynchus spp. are listed under the U.S.
Endangered Species Act (ESA), research has focused on predicting the future population dynamics for these
low-abundance stocks. One method used to make predictions is known as population viability analysis.
Pacific salmon populations exhibit much higher apparent variability than other ESA-listed vertebrates, and
high variability increases the probability of extinction. If the high variability is primarily due to counting
methods, it could be reduced in model predictions by using methods that correct for measurement error,
sampling error, or both. Using data from British Columbia pink salmon O. gorbuscha and Snake River springor
summer-run Chinook salmon O. tshawytscha and several modeling approaches (Ricker, Dennis, and statespace
models), we compared repeated counts of the same population (e.g., spawner and fry, dam and redd
counts). We applied the methods to the first half of the time series and compared the predictions with the last
half of the time series. The results demonstrated that having counts of all life stages of a Pacific salmon
population is no guarantee that variability will be markedly reduced. Measurement error is not the primary
cause of high variability in empirical estimates of abundance or in predicted future abundance for the stocks
analyzed. The very wide bounds on predicted abundance limit the utility of the model predictions for making
management decisions. Furthermore, obtaining more accurate or complete measurements of population
abundance is unlikely to reduce the wide error bounds in predictions of future abundances.
 
Paustian, S. J., ed. (1992). "A channel type users guide for the Tongass National Forest, southeast Alaska." Alaska Region, Region 10, U.S. Forest Service, Technical Paper 26.
Paybins, K. S. (2003). Flow origin, drainage area, and hydrologic characteristics for headwater streams in the mountaintop coal-mining region of southern West Virginia, 2000-01. U.S. Geological Survey Water Resources Investigations Report 02-4300. U. G. Survey. Reston, Virginia, US Geological Survey.
Payne, B. A. and M. F. Lapointe (1997). "Channel morphology and lateral stability: effects on distribution of spawning and rearing habitat for Atlantic salmon in a wandering cobble-bed river." Canadian Journal of Fisheries and Aquatic Science 54: 2627-2636.
Pronounced downstream variations in channel morphology in the wandering, gravel–cobble Nouvelle River,
Quebec, provided an opportunity to assess certain effects of channel planform and stability on rearing and spawning habitat
for Atlantic salmon (Salmo salar). At summer low flows, weighted useable area per unit channel length differed significantly
among five reaches of contrasting morphology with identical discharge regimes. A braid-like reach, dominated by a wide,
dissected midstream bar, offered three to five times more potential habitat for juveniles (21 m usable width for fry, 12 m for
parr) than two sharply curved reaches with minor backchannels (4 m for both fry and parr) and two to three times more
potential habitat than two moderately curved reaches. Fining of potential spawning riffles downstream from eroding cutbanks
was not detected, in five unstable river reaches even where erosion rates reached 10 m/year, and the percentage sand content
of eroding banks was five times that of instream gravels. This finding challenges the assumption that large local inputs of sand
necessarily cause fining of instream spawning gravels, and suggests that hydraulic forces in a moderately powerful river can
be quite efficient in preventing the buildup of excess fines in midchannel riffle habitat.
 
Paz, A. R. and W. Collischonn (2007). "River reach length and slope estimates for large-scale hydrological models based on a relatively high-resolution digital elevation model." Journal of Hydrology(343): 127-139.
Application of regular grid-based distributed hydrological models requires
information related to river drainage networks, such as flow directions, flow
accumulated areas, basin delineation, and length and slope of river reaches for every
grid cell. Most of those data may be extracted automatically from Digital Elevation
Models (DEM). However, when dealing with large basins, available DEMs are usually
in higher spatial resolution than the model grid. Upscaling procedures have therefore
been developed to extract low-resolution flow directions and flow accumulated areas
from relatively high-resolution DEMs. This paper presents a new methodology, which
extends the upscaling method, to automatically extract length and slope of river
reaches for large-scale grid-based hydrological models. The proposed method is new
and is believed to be the first attempt to produce such information automatically
from DEMs. The methodology was applied to parts of the Uruguay river basin in
South-America, using the globally-wide available DEM produced by the Shuttle Radar
Topography Mission (SRTM-90m). Quality of results was assessed by comparing
calculated river lengths with distances measured over vectorized river networks, which
were assumed to be correct. It was shown that the proposed methodology adequately
assigns river lengths to every model cell, while ensuring that the whole length of the
streams is considered. The effect of DEM resolution on calculated river length errors
was analyzed by resampling the SRTM DEM to three different cell sizes, with best
results obtained for the higher resolution DEM. The modification of the original
high-resolution DEM through the process known as stream burning was also tested,
largely improving the quality of the results. Finally, optimal distance transforms were
used for the calculation of distance increments, instead of Euclidean local distances,
 
 
Peak Northwest, I. (1986). Nooksack River Basin Erosion and Fisheries Study. Talent, Oregon, prepared by Peak Northwest, Inc. for Lummi Tribal Fisheries Department: 109.
Pearce, A. J. (1986). "Geomorphic effectiveness of erosion and sedimentation events." Journal of Water Resources 5(1): 551-567.
Pearce, A. J. and A. Watson (1983). "Medium-term effects of two landsliding episodes on channel storage of sediment." Earth Surface Processes and Landforms 8: 29-39.
Pearce, A. J. and A. J. Watson (1986). "Effects of earthquake-induced landslides on sediment budget and transport over a 50-yr period." Geology 14: 52-55.
Pederson, J., et al. (2000). "Ancient hillslope deposits: Missing links in the study of climate controls on sedimentation." Geology 28(1): 27-30.
Pelletier, G. and D. Bilhimer (2003). Stillaguamish River Watershed Temperature Total Maximum Daily Load, Washington State Department of Ecology.
Peñas, F. J., et al. (2011). "Influence of data sources and processing methods on theoretical river network quality." Limnetica 30(2): 197-216.
Stream ecosystem research and water resource management need to be considered over broad spatial scales. Moreover,
the investigation of the spatial configuration and habitat characteristics of streams requires an accurate and precise spatial
framework to reflect a catchment’s physical reality that can successfully explain observed patterns at smaller scales. In
this sense, geographic information systems represent an essential tool to satisfy the needs of researchers and managers.
Specifically, theoretical river networks (TRNs) extracted from digital elevation models (DEMs) have become much more
common in recent years, as they can provide a suitable spatial network and hierarchical organisation to sort out river ecosystem
information from reach to catchment levels. Nevertheless the quality of the extracted TRN depends greatly on the spatial
resolution of the DEM and the methodology used in the network extraction processes.
In this study, we compare the quality of 9 TRNs extracted from DEMs with different spatial resolutions ranging from regional
(5 m) to national (25 m) and global scales (90 m) using the ArcHydro, Hec-GeoHMS and Netstream software packages. To
achieve our goal, we compared (i) the DEM-derived slope; (ii) the spatial accuracy of the TRNs in relation to a control river
network; (iii) the structure of the TRNs through analysis of the number of river segments, average river segment length and
total river length by stream order, drainage density and the mean upstream slope throughout the TRN; and (iv) the ability of
variables derived from TRNs to discriminate among stream types classified according to flow type and substrate composition.
We demonstrated that not only DEM spatial resolution but also the DEM data source and raster creation process exert
an important influence on terrain characteristics derived from DEMs and TRN properties. Moreover, TRNs extracted with
NetStream generally showed better performance than those extracted with ArchHydro and HecGeoHMS. Nevertheless, river
network extraction quality, DEM spatial resolution and extraction algorithms exhibit complex relationships due to the large
number of interacting factors.
 
Pentec Environmental Inc. (1991). Methods for Testing Effectiveness of Washington Forest Practice Rules and Regulations with Regard to Sediment Production and Transport to Streams. Edmonds, Washington, prepared by Pentec Environmental, Inc. for TFW/CMER Water Quality Steering Commitee: 124.
Perera, H. and G. Willgoose (1998). "A physical explanation of the cumulative area distribution curve." Water Resources Research 34(5): 1335-1343.
Pérez-Cabello, F., et al. (2006). "Mapping erosion-sensitive areas after wildfires using fieldwork, remote sensing, and geographic information systems techniques on a regional scale." Journal of  Geophysical Research G04S10.
Alterations in the hydrological cycle following wildfire due to the loss of ground
cover vegetation and changes in soil properties have been documented in many studies.
Nevertheless, the rapid process of vegetation recovery reduces such negative effects.
Vegetation cover before fire, fire severity, and geophysical properties are important factors
that control spatial discontinuities involved in the vegetation-covering process. The
objective of this study was to estimate the probability of high erosion in order to map
erosion-sensitive areas after fire. The analysis was carried out in different plant
communities burnt by summer wildfires in the pre-Pyrenean area (Spain). Three-year
Landsat Thematic Mapper (TM) images have been used for mapping wildfire areas and
severity levels. Conversion to spectral reflectance has been applied for radiometric
correction by normalizing topographic and atmospheric effects. Likewise, other physical
variables have also been incorporated into the geographic information system (GIS):
vegetation types, parent material, illumination, slope, aspect, and precipitation. The
dependent variable has been characterized by means of fieldwork and a
photointerpretation process based on high-resolution digital aerial orthophotographs taken
11–12 years after the fire. Different logistic regression models have been used for
mapping the probability of erosion. Results indicate that prefire normalized difference
vegetation index values and aspect are the most important variables for estimating erosionsensitive
areas after fire (Nagelkerke r2 = 0.66; Kappa values = 0.65). Finally, the use
of nonparametric models with environmental digital information based on GIS can
facilitate the management of burnt areas.
 
Perkins, S. (1993). Green River Channel Migration Study. Seattle, WA, King County Department of Public Works, Surface Water Management Division: 50.
Perkins, S. (2000). Geomorphic evaluation of gravel placement in the Green River, Washington. Seattle, WA, U. S. Army Corps of Engineers, Seattle District: 52.
Perkins, S. and B. D. Collins (1997). Landslide and Channel Response Inventory for the Stillaguamish Watershed, Snohomish and Skagit Counties, Washington, Stillaguamish Tribe of Indians, Department of Natural Resources.
Perkins, S. J. (1989). Interactions of landslide-supplied sediment with channel morphology in forested watersheds. Geological Sciences. Seattle, WA, University of Washington.
Perkins, S. J. (1989). "Landslide deposits in low-order streams - their erosion rates and effects on channel morphology." Headwaters Hydrology, American Water Resources Association.
Perron, J. T., et al. (2008). "Spectral signatures of characteristic spatial scales and non-fractal structure in landscapes." Journal of  Geophysical Research 113(F04003).
Landscapes are sometimes argued to be scale-invariant or random surfaces, yet qualitative observations suggest that they contain characteristic spatial scales. We quantitatively investigate the existence of characteristic landscape scales by analyzing two-dimensional Fourier power spectra derived from high-resolution topographic maps of two landscapes in California. In both cases, we find that spectral power declines sharply above a frequency that corresponds roughly to hillslope length, implying that the landscape is relatively smooth at finer scales. The spectra also show that both landscapes contain quasiperiodic ridge-and-valley structures, and we derive a robust measure of the ridge-valley wavelength. By comparing the spectra with the statistical properties of spectra derived from randomly generated topography, we show that such uniform valley spacing is unlikely to occur in a random surface. We describe several potential applications of spectral analysis in geomorphology beyond the identification of characteristic spatial scales, including a filtering technique that can be used to measure topographic attributes, such as local relief, at specific scales or in specific orientations.
 
Perry, C. D., et al. (1999). "Watershed-scale water quality impacts of riparian management." Journal of Water Resources Planning and Management 125(3): 117-125.
The SPANS (SPatial ANalysis System) Geographic Information System (GIS) and the Georgia Landcover Database were used to identify land use within the 333 km (super 2) Little River Research Watershed of Georgia and to perform a variety of spatial analyses with the purpose of determining the water quality impact of riparian forest management within the watershed. The watershed contains a dense dendritic stream network, with approximately 54% of a 30 m riparian buffer on either side of all streams of second order or larger currently in forest cover. Approximately 60% of the upland is in some type of agricultural production. GIS analyses were used to extrapolate field-scale measurements on riparian forest nonpoint-source-pollution nutrient buffering capacity to the watershed-scale, and to evaluate the water quality impact of three increasing deforestation and three increasing reforestation riparian forest management scenarios. Nitrogen (N) and phosphorus (P) loadings at the watershed outlet could potentially increase by 16.7 and 9.5%, respectively, under a 50% riparian buffer deforestation scenario. Decreases in potential N and P loadings of 13.2 and 7.4%, respectively, could potentially occur under a 50% reforestation scenario. Fragmentation analysis of the riparian buffer indicated that only six blocks of riparian forest at least 2 km long remained in the entire watershed, confirming the fragmented state of the riparian forest.
 
Personius, S. F. (1995). "Late Quaternary stream incision and uplift in the forearc of the Cascadia subduction zone, western Oregon." Journal of  Geophysical Research 100(B10): 20,193-120,210.
Documentation of a latest Pleistocene/earliest Holocene episode of strath formation and fluvial aggradation in the Oregon Coast Range provides a datum from which long-term bedrock stream incision rates are determined. Variations in long-term incision rates probably reflect cumulative differential uplift in the forearc of the Cascadia subduction zone, although factors such as bedrock and climatic controls and isostatic adjustments to erosion obscure the precise relationship between surface uplift and stream incision. Patterns of differential incision are most striking near the latitude of Newport, where a steep gradient divides a region of higher rates (∼0.6–0.9 mm/yr) in the northern Coast Range from a region of lower rates (∼0.1–0.3 mm/yr) in the central Coast Range. The steep incision gradient is nearly coincident with abrupt changes in marine terrace (∼80–125 kyr) uplift rates, the locations of Quaternary faults, and the southern flank of a saddle of low historic (∼40–70 years) uplift. The exact causes of these variable patterns of incision/uplift are unknown. Analogies with uplift patterns in other subduction zones and comparisons with other neotectonic data in the region indicate that patterns of differential incision probably are caused by variations in permanent strain accumulation along the Cascadia subduction zone. Such variations may be related to differences in seismic moment release during individual earthquakes, to changes in plate geometry or rates of wedge accretion, to segmentation of earthquake ruptures, and/or to deformation on active structures in the North American plate and accretionary wedge.
 
Personius, S. F., et al. (1993). "Evidence for Regional Stream Aggradation in the Central Oregon Coast Range during the Pleistocene-Holocene transition." Quaternary Research 40: 297-308.
Pess, G. R., et al. (1999). "Stream-reach and watershed-scale variables and salmonid spawning distribution and abundance in the Pudget Sound Region " Watershed Management to Protect Declining Species: 387-346.
Pess, G. R., et al. (1999). Historic and current factors that limit Coho salmon (Oncorhynchus kisutch) production in the Stillaguamish River basin, Washington State: implications for salmonid habitat protection and restoration., Prepared for Snohomish County Department of Public Works and The Stillaguamish Tribe of Indians.
Pess, G. R., et al. (2003). Anthropogenic alterations to the biogeography of Puget Sound salmon. Restoration of Puget Sound Rivers. D. R. Montgomery, S. Bolton, D. B. Booth and L. Wall. Seattle, University of Washington Press: 129-154.
Pess, G. R., et al. (2002). "Landscape characteristics, land use, and coho salmon (Oncorhynchus kisutch) abundance, Snohomish River, Wash., USA." Canadian Journal of Fisheries and Aquatic Science 59: 613-623.
We used temporally consistent patterns in the spatial distribution of returning adult coho salmon
(Oncorhynchus kisutch) to explore relationships between salmon abundance, landscape characteristics, and land use
patterns in the Snohomish River watershed, Wash. The proportion of total adult coho salmon abundance supported by a
specific stream reach was consistent among years, even though interannual adult coho salmon abundance varied substantially.
Wetland occurrence, local geology, stream gradient, and land use were significantly correlated with adult
coho salmon abundance. Median adult coho salmon densities in forest-dominated areas were 1.5–3.5 times the densities
in rural, urban, and agricultural areas. Relationships between these habitat characteristics and adult coho salmon abundance
were consistent over time. Spatially explicit statistical models that included these habitat variables explained almost
half of the variation in the annual distribution of adult coho salmon. Our analysis indicates that such models can
be used to identify and prioritize freshwater areas for protection and restoration.
 
Peterson, D. L. and K. C. Ryan (1986). "Modeling postfire conifer mortality for long-range planning." Environmental Management 10(6): 797-808.
Peterson, D. P. B. E. R., J. B. Dunham, K. D. Fausch, M. K. Young, Michael (2008). "Analysis of trade-offs between threats of invasion by nonnative brook trout (Salvelinus fontinalis) and intentional isolation for native westslope cutthroat trout (Oncorhynchus clarkii lewisi)." Canadian Journal of Fisheries and Aquatic Science 65: 557-573.
Peterson, E. E., et al. (2007). "Geostatistical modelling on stream networks: developing valid covariance matrices based on hydrologic distance and stream flow." Freshwater Biology 52: 267-279.
1. Geostatistical models based on Euclidean distance fail to represent the spatial
configuration, connectivity, and directionality of sites in a stream network and may not be
ecologically relevant for many chemical, physical and biological studies of freshwater
streams. Functional distance measures, such as symmetric and asymmetric hydrologic
distance, more accurately represent the transfer of organisms, material and energy through
stream networks. However, calculating the hydrologic distances for a large study area
remains challenging and substituting hydrologic distance for Euclidean distance may
violate geostatistical modelling assumptions.
2. We provide a review of geostatistical modelling assumptions and discuss the statistical
and ecological consequences of substituting hydrologic distance measures for Euclidean
distance. We also describe a new family of autocovariance models that we developed for
stream networks, which are based on hydrologic distance measures.
3. We describe the geographical information system (GIS) methodology used to generate
spatial data necessary for geostatistical modelling in stream networks. We also provide an
example that illustrates the methodology used to create a valid covariance matrix based on
asymmetric hydrologic distance and weighted by discharge volume, which can be
incorporated into common geostatistical models.
4. The methodology and tools described supply ecologically meaningful and statistically
valid geostatistical models for stream networks. They also provide stream ecologists with
the opportunity to develop their own functional measures of distance and connectivity,
which will improve geostatistical models developed for stream networks in the future.
5. The GIS tools presented here are being made available in order to facilitate the
application of valid geostatistical modelling in freshwater ecology.
 
Peterson, I. (1987). "Forest fires, barnacles and trickling oil." Science News 132: 220-223.
Peterson, J. T. and T. J. Kwak (1999). "Modeling the effects of land use and climate change on riverine smallmouth bass." Ecological Applicatons 9(4): 1391-1404.
Peterson, N. P. (1982). "Immigration of Juvenile Coho Salmon (Oncorhynchus kisutch) into Riverine Ponds " Canadian Journal of Fisheries and Aquatic Sciences 39: 1308-1310.
Peterson, N. P., et al. (1992). Assessment of Cumulative Effects on Salmonid Habitat: Some Suggested Parameters and Target Conditions. Seattle, Washington, prepared by Streamside Studies for the Washington Department of Natural Resources and The Coordinated Monitoring, Evaluation and Research Committee: 36-40.
Peterson, N. P. and T. P. Quinn (1995). "Spatial and temporal variation in dissolved oxygen in natural egg pockets of chum salmon, in Kennedy Creek, Washington ": 131- 143.
Peterson, N. P. and T. P. Quinn (1996). "Persistence of egg pocket architecture in redds of chum salmon, Oncorhynchus keta " Environmental Biology of Fishes 46: 243-253.
Peterson, N. P. and T. P. Quinn (?). Incubation Environment of Chum Salmon (Oncorhynchus keta) in Kennedy Creek: Part A. Persistence of Egg Pocket Architecture in Chum Salmon Redds: 9-25.
Petit, F. (1990). "Evaluation of grain shear stresses required to initiate movement of particles in natural rivers." Earth Surface Processes and Landforms 15: 135-148.
Peucker, T. K. and D. H. Douglas (1975). "Detection of surface-specific points by local parallel processing of discrete terrain elevation data." Computer Graphics and Image Processing 4: 375-387.
Philips, E. L. (?). Washington Climate for these counties: Northeast Clallam, Northeast Jefferson, Island, San Juan, Skagit, Snohomish, and Whatcom. Seattle, WA, U. S. Weather Bureau: 5-9,12.
Philips, F. M. (1995). Cosmogenic Chlorine-36 Accumulation: A method for Dating Quaternary Landforms Dating Methods for Quaternary Deposits. N. W. Rutter and N. R. Catto. St. John's, Newfoundland, Geological Association of Canada.
Philips, R. W., et al. (1975). "Some Effects of Gravel Mixtures on Emergence of Coho Salmon and Steelhead Trout Fry " Transactions of the American Fisheries Society 3: 461-466.
Piccolo, J. J. and M. S. Wipfli (2002). "Does red alder (Alnus rubra) in upland riparian forests elevate macroinvertebrate and detritus export from headwater streams to downstream habitats in southwestern Alaska?" Canadian Journal of Fisheries and Aquatic Sciences 59: 1-11.
Pichon, C., et al. (2008). "Spatial metrics and methods for riverscapes: quantifying variability in riverine fish habitat patterns." Environmetrics.
Defining the optimal configuration of all habitats required during a life cycle, called vital habitat, is a necessary step
for effective management of riverine fishes and restoration of river habitats. Landscape ecology provides many
metrics and methods to study the composition and configuration of habitats, but they need to be adapted for fishes
in river environments or riverscapes. For example, hydrographic distance seems more appropriate than Euclidean
distance for measuring distances between vital habitats in riverscapes.We adapted some metrics to assess habitats
patterns of a threatened cyprinid species (Barbus barbus) for natural and artificial riverscapes of the Seine river,
France. Composition metrics provided essential quantification of the relative abundance of the vital habitats,
whereas configuration metrics were relevant to quantify their spatial arrangement and spatial relationships.
Nearest-neighbor hydrographic distance was useful to evaluate the influence of flow variability in the natural
riverscape, but was not relevant to discriminate the artificial riverscape from the natural one. Conversely, a
proximity index revealed high fragmentation in the artificial riverscape. Spatial habitat relationships between
feeding and resting habitats, evaluated with a moving window analysis, provided a map of daily activity patches
and emphasized the gaps in the biological continuity of the riverscape. The spatial metrics and methods we adapted
to the particularities of the Seine river allowed us to detect natural and artificial variability in fish habitat patterns.
They should help in evaluating impacts of habitat alteration and isolation and prioritize preservation and
restoration policies in human-impacted rivers.
 
Pickett, S. T. A. and P. S. e. White (1985). The ecology of natural disturbance and patch dynamics. New York, New York, Academic Press.
Pickford, S. G., et al. (1980). "Weather, Fuels, and Lightning Fires in Olympic National Park." Northwest Science 54(2): 92-105.
Pickles, J. (1997). "Tool or science? GIS, technoscience, and the theoretical turn." Annals of the Association of American Geographers 87(2): 363-372.
Pickup, G., et al. (1983). "Modelling sediment transport as a moving wave - the transfer and deposition of mining waste." Journal of Hydrology 60: 281-301.
Pickup, G. and A. Marks (2001). "Regional-scale sedimentation process models from airborne gamma ray remote sensing and digital elevation data."
Airborne gamma ray survey data were used to provide information on potassium, thorium and uranium concentrations in surface soil and rock in arid central Australia. Spatial patterns in these radioelements allow tracing of paths of sediment at catchment scale. Survey elevation data are combined with contour data to produce digital elevation models for terrain analysis, tracing of sediment flow paths and modelling of extreme floods. Gamma ray data show consistent variation with slope, a limited range of drainage areas, and erosion/deposition models derived from the conservation of mass equation. Supply-limited sediment transport models give a reasonable reproduction of observed radioelement distribution but some elements of the distribution pattern reflect the area inundated by 500-1000 year floods rather than the effects of simple downslope movement. Partial area sediment supply models are derived by downstream accumulation of erosion and deposition rates calculated using the conservation of mass equation with transport laws based on slope alone and stream power. Comparison with observed radioelement patterns suggests that both transport laws apply in different parts of the landscape. Regional-scale sediment transport models will require a range of models depending on location in the landscape and event frequency. This approach may allow estimation of sediment delivery ratios.
 
Pickup, G. R., et al. (1983). "Modelling sediment transport as a moving wave - the transfer and deposition of mining waste." Journal of Hydrology 60: 281-301.
Piegay, H., et al. (1999). "Input, storage and distribution of large woody debris along a mountain river continuum, the Drome River, France." Catena 35: 19-39.
Piegay, H. and A. M. Gurnell (1997). "Large woody debris and river geomorphological pattern: examples from S.E. France and S. England." Geomorphology 19: 99-116.
Piégay, H., et al. (1999). "Input, storage and distribution of large woody debris along a mountain river continuum, the Drôme River, France." Catena 35: 19-39.
Large woody debris (LWD) input, storage and distribution were studied along the Droˆme
River, a French Alpine river with an active shifting channel and a well-developed riparian forest.
LWD input from the floodplain is low: 669.6 mg yeary1 between 1948 and 1971 and 569.3 mg
yeary1 between 1971 and 1991. Based on estimates of average LWD mass per study plot, a range
of 766–2122 mg yeary1 of LWD were stored within the active channel  e.g., unvegetated bars
and low-flow channel; 60 km course covering 492 ha.. LWD accumulations are mainly observed
on gravel bars at a limited number of preferential sites. They are relatively rare and are randomly
distributed in the low-flow channel. Consequently, LWD location is mainly associated with the
decrease of flow level in shallow sectors. In the low-flow channel, LWD stop-en-route is primarily
caused by in-channel structures such as boulders or vegetated islets. General geomorphological
factors e.g., pattern, slope, etc..provide less explanation of LWD distribution. Moreover, the
residence time of LWD accumulations on the Droˆme River is short  LWD storages1.3–3.7 times
the annual LWD input from the floodplain. and their morphogenic role is negligible: few of the
accumulations are buried or characterized by vegetation shoots and associated pools.
 
Pierce, J. L., et al. (2004). "Fire-induced erosion and millennial-scale climate change n northern ponderosa pine forests." Nature 432(11): 87-90.
Pierson, A. (2004). Geologic Report: Slope stability assessment of the Golden Goat Timber Sale. Olympia, WA, Washington Department of Natural Resources.
Pierson, F. B., et al. (2001). "Spatial and temporal effects of wildfire on the hydrology of a steep rangeland watershed." Hydrological Processes 15(15): 2905-2916.
Wildfire is a major ecological process and management issue on western rangelands. The impacts of wildfire on hydrologic processes such as infiltration, runoff, and erosion are not well understood. Small-plot rainfall simulation methods were applied in a rangeland wildfire setting to determine post-fire hydrologic response. Infiltration and interrill erosion processes were measured immediately post-fire and one year following the 1999 34 400 ha Denio fire in northwestern Nevada. Plot-scale spatial and temporal variability in fire impacts was compared with adjacent unburned areas. An index of water repellency was derived and used to quantify the influence of water-repellent soil conditions on infiltration. Results indicate the impact of the fire on infiltration was localized primarily on coppice microsites directly under shrubs characterized by high surface litter accumulations. Coppice microsites had very uniform fire-induced soil water repellency with 29 of 30 plots exhibiting at least a 10% reduction in initial infiltration with an average 28% reduction. Cumulative erosion was nearly four times higher on burned coppices compared with unburned coppices. The impact of the fire on infiltration and erosion was reduced, but still evident, 1 year after fire. Significant temporal variability in infiltration between years was observed on both burned and unburned areas, complicating the interpretation of fire impacts and hydrologic recovery following wildfire. Published in 2001 by John Wiley & Sons, Ltd.
 
Pierson, T. C. (1977). Factors controlling debris-flow initiation on forested hillslopes in the Oregon Coast Range. Geological Sciences. Seattle, WA, University of Washington: 167.
Pierson, T. C. (1980). "Erosion and deposition by debris flows at Mt. Thomas, North Canterbury, New Zealand." Earth Surface Processes and Landforms 5: 227-247.
Pierson, T. C. (1981). "Dominant particle support mechanisms in debris flows at Mt. Thomas, New Zealand, and implications for flow mobility." Sedimentology 28: 49-60.
Pierson, T. C. (1983). "Soil pipes and slope stability." Quaternary Journal of Engineering Geology, London 16: 1-11.
Pierson, T. C. (?). Flow behavior of two major lahars triggered by the May 18, 1980 Eruption of Mount St. Helens, Washington (DRAFT). Vancouver, Washington, U.S. Geological Survey: 16.
Pierson, T. C. and J. E. Costa (1987). "A rheologic classification of subaerial sediment-water flows." Reviews in Engineering Geology 7: 1, 10.
Pierson, T. C. and K. M. Scott (1985). "Downstream Dilution of a Lahar: Transition From Debris Flow to Hyperconcentrated Streamflow." Water  Resources Research 21(10): 1511-1524.
Pietraszek, J. H. (2006). Controls on post-fire erosion at the hillslope scale, Colarado Front Range. Department of Forest, Rangeland, and Watershed Stewardship. Fort Collins, Colorado, Colorado State University. Master of Science: 124.
Pike, R. G. and R. Scherer (2003). "Overview of the potential effects of forest management on low flows in snowmelt-dominated hydrologic regimes." BC Journal of Ecosystems and Management 3(1): 44-60.
This paper reviews potential effects of forest management on low flows in snowmelt-dominated hydrologic
regimes. The hydrologic response of low flows to forest management was found to be highly variable in
magnitude, time, and space. Forest management generally increases water volume—no case studies relevant
to snowmelt-dominated regimes reported a decrease in water quantity as a result of forest harvesting.
In areas where fog drip occurs, a decrease in water volume contributing to low flows might be observed.
The longevity of increased water quantity is infrequently discussed in the literature specific to snowmeltdominated
regimes. A few authors, however, have commented on expected longevity of response based
upon analysis of literature not specific to snowmelt-dominated regimes. These authors generally report
a return to pre-treatment low flow levels within 3–6 years with the re-establishment of vegetation.
The review identifies many knowledge, research, and extension needs. Knowledge of low flows is
hampered by an incomplete understanding of generation processes, particularly those relating to subsurface
flow, evapotranspiration, and the interrelated effects of forest practices and climate change. Forest management
is only one of many human activities that can potentially affect a watershed’s hydrologic regime.
Because natural processes and human activities occur simultaneously, studying the sole effects of forest
management on low flows is difficult. Limitations in low flow science around measurement methodologies,
scaling of results, and inadequate research design are noted.
 
 
Pike, R. J. (1988). "The geometric signature: Quantifying landslide-terrain types from digital elevation models." Mathematical Geology 20: 491-511.
Pilotti, M., et al. (1996). "Identification and analysis of natural channel networks from digital elevation models." Earth Surface Processes and Landforms 21: 1007-1020.
Pine, E. J., et al. (1999). Characteristics and evolution of rills on steep terrain after a forest fire. Abstracts with Programs - Geological Society of America. 31: 441.
The Buffalo Creek wildfire burned approximately 4800 hectares in the Rocky Mountain foothills of Colorado during May 18-20, 1996. One of the most heavily impacted watersheds was Spring Creek, in which nearly 79% was burned. Comparison of aerial and oblique photos show there were no rills present until after the intense rainstorms of July 1996. The geology of the watershed is characterized by bedrock of Pikes Peak granite, a highly erodable coarse-grained rock that produces shallow soils that support vegetation which is predominately Ponderosa Pine on south-facing slopes and Douglas Fir on north-facing slopes. The hillslopes selected for this study were convex, with both north- and south-facing aspects, with slopes ranging from 29-47%, and rill density ranging from 0.6-1.0 rills/meter. The top widths range from 0.23-0.68 meters and the depths range from 0.03-0.14 meters. The width and depth increase in the downslope direction as the rills join to form dendritic drainage patterns with nearly parallel channels. Width-to-depth ratios were about 9 near the ridge and decreased in the downslope direction to 4 near the bottom of the hillslope. The cross-sectional area of these rills is related to the hydraulic radius by a power law with an exponent ranging from 0.4 to 0.7 which is similar to that of rills on agricultural lands ( approximately 0.5). The geometry of these rills has changed with time. The width to-depth ratio has shown a slight increase by the second year after the fire as has the hydraulic radius. We hypothesized that the freeze-thaw process during winter increases rill widths and eroded soil collects in the bottom of the rill decreasing their depths. Sediment trap measurements indicate that the rills are transporting very little of this eroded soil downslope perhaps because the infiltration properties of the soil have changed since the fire and opportunistic vegetation has begun to stabilize the soil collecting in the rill bottoms.
 
Pinol, J., et al. (1997). "Modelling the hydrological response of Mediterranean catchments, Prades, Catalonia. The use of distributed models as aids to hypothesis formulation." Hydrological Processes 11: 1287-1306.
Pinter, N., et al. (1994). "Relative Dating of Terraces on the Owens River, Northern Owens Valley, and Correlation with Morraines of the Sierra Nevada " Quaternary Research 42: 266-276.
Pitlick, J. and M. M. Van Steeter (1998). "Geomorphology and endangered fish habitats of the upper Colorado River. 2. Linking sediment transport to habitat maintenance." Water Resources Research 34(2): 303-316.
Pizzuto, J. (1994). "Channel adjustments of the Powder river, Montana." Geological Society of America Bulletin 106: 1494-1501.
Pizzuto, J. E. (1992). "The morphology of graded gravel rivers: a network perspective." Geomorphology 5: 457-474.
Analytical hydraulic geometry equations are combined with a watershed model to quantify downstrema trends in equilibrium fluvial morphology in gravel streams. The hydraulic geometry equations relate fluvial morphology to discharge, bedload transport rate and grain size; these independent variables are in turn determined by routing gravel supplied at zero-order basins through a channel network. As gravel is routed downstream, it decreases in size according to an empirical power law. the resulting model specifies the discharge, bedload transport rate, grain size, width, depth, slope and river bed elevation at every point in the drainage basin.
When the downstream fining law is calibrated using data from central Pennsylvania, downstream hydraulic geometry exponents for the width and depth are accurately reproduced. Longitudianl profiles calculated from the model fit observed profiles in larger watersheds, but are inaccurate in smaller watersheds. The model also reproduces Hack's (1957) empirical result relating slope to the 0.6 power of A/Ds, where A is the drainage basin area and ds is the mean grain size.
Unfortunately, the model requires an equilibrium time scale which is unreasonably long, suggesting that the channels of the study area are best described as systems with multiple time scales and multiple rates of response. The model probably is successful because the width and depth have relatively short relaxation times, and also because downstream trends in discharge and sediment transport rate are controlled by network topology. Thus, downstrema hydraulic geometry equations do not necessarily provide strong evidence for the equilibrium of fluvial landscapes. Rather, they reflect the constancy of network topology and the rapid response times of width and depth.
 
Planchon, O. and F. Darboux (2001). "A fast, simple and versatile algorithm to fill the depressions of digital elevation models." Catena 46: 159-176.
The usual numerical methods for removing the depressions of a Digital Elevation Model
ŽDEM.gradually fill the depressions and merge the embedded ones. These methods are complex
to implement and need large computation time, particularly when the DEM contains a high
proportion of random noise. A new method is presented here. It is innovative because, instead of
gradually filling the depressions, it first inundates the surface with a thick layer of water and then
removes the excess water. The algorithm is simple to understand and to implement, requiring only
a few tens of code lines. It is much faster than usual algorithms. Moreover, this method is
versatile: depressions can be replaced with a surface either strictly horizontal, or slightly sloping.
The first option is used for the calculation of depression storage capacity and the second one for
drainage network extraction. The method is fully detailed and a pseudo-code is provided. Its
practical computation time, evaluated on generated fractal surfaces, is asymptotically proportional
to N1.2 where N is the number of grid points. The theoretical computation time is asymptotically
proportional to N1.5 in all cases, with the exception of some exotic ones with no practical interest.
By contrast, existing methods have a computation time asymptotically proportional to N2.
Applications are done for both generated and measured surfaces with 256 cells to 6.2 million cells.
 
Platt, J. R. (1964). "Strong inference." Science 146(3642): 347-353.
Platts, W. S. and R. L. Nelson (1988). "Fluctuations in trout populations and their implications for land-use evaluation." North American Journal of Fisheries Management 8: 333-345.
Platts, W. S., et al. ( 1989). "Changes in Salmon Spawning and Rearing Habitat from Increased Delivery of Fine Sediment to the South Fork Salmon River, Idaho." Transactions of the American Fisheries Society 118: 274-283.
Pleus, A. and D. Schuett-Hames (1998). Method Manual for Stream Segment Identification Olympia, Washington, NW Indian Fisheries Commission and Timber, Fish, and Wildlife: 103.
Pleus, A. E. and P. Goodman (2003). Type N stream demarcation study: 2002 tribal perennial stream survey data collection using CMER methods, Northwest Indian Fisheries Commission: 53.
Pleus, A. E. and D. Schuett-Hames (1998). "Method Manual for Stream."
Plotnikoff, R. W. (1993). Washington State Department of Ecology Instream Biological Assessment Monitoring Protocols: Benthic Macroinvertebrates (DRAFT). Olympia, Washington, Washington State Department of Ecology.
Poesen, J., J. Nachtergaele, G. Verstraeten, and C. Valentin (2003). "Gully erosion and environmental change: importance and research needs." Catena 50(2-4): 91-133.
Assessing the impacts of climatic and, in particular, land use changes on rates of soil erosion by water is the objective of many national and international research projects. However, over the last decades, most research dealing with soil erosion by water has concentrated on sheet (interrill) and rill erosion processes operating at the (runoff) plot scale. Relatively few studies have been conducted on gully erosion operating at larger spatial scales.Recent studies indicate that (1) gully erosion represents an important sediment source in a range of environments and (2) gullies are effective links for transferring runoff and sediment from uplands to valley bottoms and permanent channels where they aggravate off site effects of water erosion. In other words, once gullies develop, they increase the connectivity in the landscape. Many cases of damage (sediment and chemical) to watercourses and properties by runoff from agricultural land relate to (ephemeral) gullying. Consequently, there is a need for monitoring, experimental and modelling studies of gully erosion as a basis for predicting the effects of environmental change (climatic and land use changes) on gully erosion rates.In this respect, various research questions can be identified. The most important ones are:(1) What is the contribution of gully erosion to overall soil loss and sediment production at various temporal and spatial scales and under different climatic and land use conditions?(2) What are appropriate measuring techniques for monitoring and experimental studies of the initiation and development of various gully types at various temporal and spatial scales?(3) Can we identify critical thresholds for the initiation, development and infilling of gullies in different environments in terms of flow hydraulics, rain, topography, soils and land use?(4) How does gully erosion interact with hydrological processes as well as with other soil degradation processes?(5) What are appropriate models of gully erosion, capable of predicting (a) erosion rates at various temporal and spatial scales and (b) the impact of gully development on hydrology, sediment yield and landscape evolution?(6) What are efficient gully prevention and gully control measures? What can be learned from failures and successes of gully erosion control programmes?These questions need to be answered first if we want to improve our insights into the impacts of environmental change on gully erosion. This paper highlights some of these issues by reviewing recent examples taken from various environments. (C) 2003 Elsevier Science B.V. All rights reserved.
 
Poesen, J. W., et al. (1994). "Effects of Rock Fragments on Soil-Erosion by Water at Different Spatial Scales - a Review." Catena 23(1-2): 141-166.
This paper reviews the various effects of rock fragments on soil erosion by water. Since these effects are scale dependent, they are investigated at three different nested spatial scales: the microplot (4 x 10(-6) - 10(0) m2), the mesoplot (10(-2) - 10(2) m2) and the macroplot (10(1) - 10(4) m2). For each scale the corresponding process mechanisms are discussed. Particular attention is paid to the effects of rock fragment cover on the intensity of soil erosion processes. At the mesoplot scale, i.e. on interrill areas, rock fragments at the soil surface can have negative as well as positive effects on sediment yield. These ambivalent effects are conditioned by the type of fine earth porosity, soil surface slope, vertical position and size of rock fragments and by the occurrence of horseshoe vortex erosion. At the microplot scale, i.e. the soil surface area which is covered by a single rock fragment, and at the macroplot scale, i.e. upland areas where both interrill and rill erosion takes place, rock fragments at the soil surface have a negative effect on sediment yield. In these two scales rock fragments can thus be considered as natural soil surface stabilizers. At the macroplot scale the mean decrease of relative interrill and rill sediment yield with rock fragment cover can be expressed by an exponential decay function. The scatter of the data indicates that a given rock fragment cover can have different efficiencies in reducing interrill and rill sediment yield depending on the varying intensities of the hydrological and erosion subprocesses. These findings have implications for erosion modelling and soil conservation.
 
Poff, N. L. (1996). "A hydrogeography of unregulated streams in the United States and an examination of scale-dependence in some hydrological descriptors." Freshwater Biology 36: 71-91.
Poff, N. L. (1997). "Landscape filters and species traits: towards mechanistic understanding and prediction in stream ecology." Journal of North American Benthological Society 16(2): 391-409.
Poff, N. L. (1998). "Multi-scale determinants of secondary production in Atlantic salmon (Salmo salar) streams." Canadian Journal of Fisheries and Aquatic Sciences 55(Suppl. 1): 201-217.
Poff, N. L., B.P.Bledose, and C.O. Cuhaciyan (2006). "Hydrologic variation with land use across the contiguous United States: Geomorphic and ecological consequences for stream ecosystems." Geomorphology 79: 264-285.
Poff, N. L., J.D.Olden, D.M. Pepin and B.P. Bledsoe (2006). "Placing global streamflow variability in geographic and geomorphic contexts." River Research and Applications 22: 149-166.
Poff, N. L., et al. (1997). "The natural flow regime: A paradigm for river conservation and restoration." BioScience 47(11): 769-784.
Poff, N. L. and J. V. Ward (1990). "Physical habitat template of lotic systems: recovery in the context of historical pattern of spatiotemporal heterogeneity." Environmental Management 14(5): 629-645.
Pollock, M. (1998). Biodiversity. River Ecology and Management: Lessons from The Pacific Coastal Ecoregion. R. Naiman and R. Bilby, Springer: 430-448.
Pollock, M. M., et al. (2003). Hydrologic and geomorphic effects of beaver dams and their influence on fishes, American Fisheries Society.
Pollock, M. M. and P. Kennard (1998). A low-risk strategy for preserving riparian buffers needed to protect and restore salmonid habitat in forested watersheds of the Pacific Northwest. Bainbridge Island, Washington, 10,000 years Institute: 42.
Pollock, M. M., et al. (1998). "Plant species richness in riparian wetlands - a test of biodiversity theory." Ecology 79(1): 94-105.
Pollock, M. M., et al. (1998). "Predicting Plant Species Richness In Forested And Emergent Wetlands - A Test of Biodiversity Theory." Ecology 79: 94-105.
Pollock, M. M., et al. (2004). "The importance of beaver ponds to coho salmon production in the Stillaguamish River basin, Washington, USA." North American Journal of Fisheries Management 24: 749-760.
The use of beaver Castor canadensis ponds by juvenile coho salmon Oncorhynchus
kisutch and other fishes has been well established. However, the population-level effects on coho
salmon resulting from the widespread removal of millions of beaver and their dams from Pacific
Coast watersheds have not been examined. We assessed the current and historic distributions of
beaver ponds and other coho salmon rearing habitat in the Stillaguamish River, a 1,771-km2
drainage basin inWashington and found that the greatest reduction in coho salmon smolt production
capacity originated from the extensive loss of beaver ponds. We estimated the current summer
smolt production potential (SPP) to be 965,000 smolts, compared with a historic summer SPP of
2.5 million smolts. Overall, current summer habitat capacity was reduced by 61% compared with
historic levels, most of the reduction resulting from the loss of beaver ponds. Current summer
SPP from beaver ponds and sloughs was reduced by 89% and 68%, respectively, compared with
historic SPP. A more dramatic reduction in winter habitat capacity was found; the current winter
SPP was estimated at 971,000 smolts, compared with a historic winter SPP of 7.1 million smolts.
In terms of winter habitat capacity, we estimated a 94% reduction in beaver pond SPP, a 68% loss
in SPP of sloughs, a 9% loss in SPP of tributary habitat, and an overall SPP reduction of 86%.
Most of the overall reduction resulted from the loss of beaver ponds. Our analysis suggests that
summer habitat historically limited smolt production capacity, whereas both summer and winter
habitats currently exert equal limits on production. Watershed-scale restoration activities designed
to increase coho salmon production should emphasize the creation of ponds and other slow-water
environments; increasing beaver populations may be a simple and effective means of creating
slow-water habitat.
 
Poole, G. C. (2002). "Fluvial landscape ecology: addressing uniqueness within the river discontinuum." Freshwater Biology 47: 641-660.
Poole, G. C., et al. (2002). "Three-dimensional mapping of geomorphic controls on flood-plain hydrology and connectivity from aerial photos." Geomorphology 48: 329-347.
Porter, S. C. (1979). "Hawaiian Glacial Ages." Quaternary Research 12: 161-187.
Porter, S. C. (1986). "Pattern and Forcing of Northern Hemisphere Glacier Variations during the Last Millennium." Quaternary Research 26: 27-48.
Post, D. and A. J. Jakeman (1996). "Relationships between catchment attributes and hydrological response characteristics in small Australian mountain ash catchments." Hydrological Processes 10: 877-892.
Post, D. A. and J. A. Jones (2001). "Hydrologic regimes of forested, mountainous, headwater basins in New Hampshire, North Carolina, Oregon, and Puerto Rico." Advances in Water Resources 24: 1195-1210.
Postma, G. (1986). "Classification for sediment gravity-flow deposits based on flow conditions during sedimentation." Geology 14: 291-294.
Potts, D. and B. K. M. Anderson (1990). "Organic debris and the management of small stream channels." Western Journal of Applied Forestry 5(1): 25-28.
Potyondy, J. P., et al. (1991). A Procedure for estimating sediment yields from forested watersheds. 5th Federal Interagency Sedimentation Conference.
Powell D.M. (1998). "Patterns and processes of sediment sorting in gravel-bed rivers." Progress  in Physical Geography 22(1): 1-32.
Powers, M. E., et al. (1996). "Dams and Downstream Aquatic Biodiversity: Potential Food Web Consequences on Hydrologic and Geomorphic Change " Environmental Management 20(6): 887-895.
Press, W. H., et al. (1986). Numerical Recipes. Cambridge, Cambridge University Press.
Press, W. H., et al. (1992). Numerical Recipes. Cambridge, Cambridge University Press.
Press, W. H., et al. (1996). Numerical Recipes in Fortran 90. Cambridge, Cambridge University Press.
Prichard, D., et al. (1998). Riparian Area Management: A User Guide to Assessing Proper Functioning Condition and the Supporting Science for Lotic Areas, USDA Bureau of Land Management: 126.
Pringle, C. M., et al. (1988). "Patch dynamics in lotic systems: the stream as a mosaic." Journal of the North American Benthological Society.
Pringle, P. T., et al. (1998). "New radiocarbon ages of major landslides in the Cascade Range, Washington." Washington Geology 26(1): 31-39.
Prochaska, A. B., et al. (2008). "Debris-flow runout predictions based on the average channel slope (ACS)." Engineering Geology 98: 29-40.
Prediction of the runout distance of a debris flow is an important element in the delineation of potentially hazardous areas on alluvial fans and
for the siting of mitigation structures. Existing runout estimation methods rely on input parameters that are often difficult to estimate, including
volume, velocity, and frictional factors. In order to provide a simple method for preliminary estimates of debris-flow runout distances, we
developed a model that provides runout predictions based on the average channel slope (ACS model) for non-volcanic debris flows that emanate
from confined channels and deposit on well-defined alluvial fans. This model was developed from 20 debris-flow events in the western United
States and British Columbia. Based on a runout estimation method developed for snow avalanches, this model predicts debris-flow runout as an
angle of reach from a fixed point in the drainage channel to the end of the runout zone. The best fixed point was found to be the mid-point
elevation of the drainage channel, measured from the apex of the alluvial fan to the top of the drainage basin. Predicted runout lengths were more
consistent than those obtained from existing angle-of-reach estimation methods. Results of the model compared well with those of laboratory
flume tests performed using the same range of channel slopes. The robustness of this model was tested by applying it to three debris-flow events
not used in its development: predicted runout ranged from 82 to 131% of the actual runout for these three events. Prediction interval multipliers
were also developed so that the user may calculate predicted runout within specified confidence limits.
 
Proffitt, G. T. and A. J. Sutherland (1983). "Transport of non-uniform sediments." Journal of Hydraulic Research(1): 33-43.
Prosser, I. P. and B. Abernethy (1996). "Predicting the topographic limits to a gully network using a digital terrain model and process thresholds." Water Resources Research 32(7): 2289-2298.
A digital terrain model is used with process thresholds to predict the extent of a stable gully network in a 5 km(2) catchment of the southeastern highlands of Australia. The model, developed by Dietrich et al. [1992, 1993], predicts the topographic controls on channel networks and interprets these in terms of a critical shear stress for channel incision (tau(c)) applied by saturation overland flow. We adapt the model slightly to compare the shear stress applied by Hortonian overland flow to that applied by saturation overland flow. The limits to gully erosion in the catchment are controlled strongly by a topographic threshold that has an inverse relationship between upslope catchment area and local gradient. The topographic threshold for channel incision is reproduced using a simple model of Hortonian overland flow and a tau(c) appropriate for incision into a degraded grass surface (tau(c) = 245 dyn/cm(2)). This is consistent with historical evidence for the timing of gully erosion. The study confirms a strong topographic control on the extent of the channel network in a catchment significantly different from the western North America catchments where the topographic threshold was first demonstrated. Despite its simplicity, the model for incision by overland flow appears capable of distinguishing the hydrological processes responsible for channel incision when these are reflected in the relationship between channel network and landscape morphology. The model requires relatively simple inputs, suggesting it may be useful for mapping gully erosion hazard in actively eroding catchments.
 
Prosser, I. P. and W. E. Dietrich (1995). "Field Experiments on Erosion by Overland-Flow and Their Implication for a Digital Terrain Model of Channel Initiation." Water Resources Research 31(11): 2867-2876.
Dietrich et al. (1992, 1993) proposed a digital terrain model for predicting the location of channel heads on the basis of the assumption that they occur where saturation overland flow exerts a boundary shear stress in excess of a critical value. Flume experiments were conducted in the modeled field site to evaluate the threshold hypothesis and to constrain critical shear stress and flow resistance parameters. Under complete grass cover, microtopography and grass stems were found to prevent significant sediment transport at all but the highest flows. When the grass stems were cut close to the ground, flow resistance and critical shear stress for significant sediment transport were reduced by up to an order of magnitude, but the remaining dense root mat prevented deep scour. These field experiments support the threshold assumption and the model estimations of the critical shear stress if local topographic convergence of flow is taken into account. The experiments also support the interpretation that significant degradation of vegetation cover is required for channel incision.
 
Prosser, I. P., et al. (1995). "Flow Resistance and Sediment Transport by Concentrated Overland-Flow in a Grassland Valley." Geomorphology 13(1-4): 71-86.
Flow resistance and sediment transport data are needed from well vegetated humid environments to evaluate surface wash erosion and channel incision by overland flow. In humid environments, runoff in valley floors can reach depths of several centimetres but erosion is often limited by dense grass cover. Intense grazing reduces grass cover but the impacts of this on sediment transport processes are poorly understood, We conducted flume experiments in a grassed valley of coastal California to investigate flow resistance and sediment yield under natural conditions and with progressive clipping of grass cover. Flow resistance has a laminar-like relationship with Reynolds number but we attribute this to very low velocity beneath submerged stems, and not to the state of flow.The sediment transport relations provide support for the concept of a threshold shear stress below which erosion is effectively prevented by surface resistance. Shear stress partitioning suggests that on a densely grassed surface over 90% of flow resistance is exerted on plant stems. This effectively prevents sediment transport at boundary shear stresses as high as 1000-1800 dyn/cm(2). Complete clipping of the grass cover reduces the critical shear stress for sediment transport to 11-38% of that under natural conditions. Continued surface wash erosion and channel initiation are prevented, however, by strong soil cohesion provided by a dense root mat. Even with reduction of root density, boundary shear stresses of at least 250-430 dyn/cm(2) are required for channel incision.
 
Prosser, I. P. and P. Rustomji (2000). "Sediment transport capacity relations for overland flow." Progress in Physical Geography 24(2): 179-193.
Nearly all physically based models of soil erosion and landform evolution include an equation for the sediment transport capacity of flow, yet the choice of relation varies from one model to another. This raises the question of the empirical and theoretical support for particular choices. We review descriptions of the effect of discharge and slope on sediment transport capacity of overland flow. A limited range of relations is defined for which there is strong theoretical and empirical support. Nevertheless, suggestions are made for improved experimental designs that further evaluate theories of sediment transport capacity and reduce uncertainty in its application to catchment modelling of sediment transport.
 
Prosser, I. P. and M. Soufi (1998). "Controls on gully formation following forest clearing in a humid temperate environment." Water Resources Research 34(12): 3661-3671.
We have constructed a chronology of gully initiation, forest clearing, and rainfall events for gullies eroded into pine plantations near Bombala, southeastern Australia, over the last 15 years. The chronology suggests that daily rainfall of 80-100 mm, which has a recurrence interval of 1.4-2 years, can initiate gully erosion on areas cleared of native forest within the previous year. Massive gully erosion was experienced from a daily rainfall of 200 mm with a recurrence of 10-15 years. Resistance to channel initiation effectively recovers within a year of disturbance, allowing only a limited opportunity for erosion. Analysis of the spatial pattern of gully erosion, using a digital elevation model, shows that gullies were initiated across all landscape positions. In contrast to previous studies, there is no clear topographic threshold that limits the extent of the gully network. We infer that the weak topographic threshold results from low resistance to scour, allowing local flow convergence to dominate over topographic accumulation of flow. Although resistance to scour is low relative to previous studies, a process threshold for gully initiation is still a useful simplification of the erosion processes. For the soils that we studied, the threshold for gully erosion relates to intense scour exposing erodible subsoils.
 
Prosser IP, I. R., JM Olley, WJ Young, PJ Wallbrink, and CJ Moran (2001). "Large-scale patterns of erosion and sediment transport in river networks, with examples from Australia." Marine and Freshwater Research 52(1): 81-99.
Puigdefabregas, J., et al. "Scales and processes of water and sediment redistribution in drylands: results from the Rambla Honda field site in southeast Spain."
Arid lands are characterised by a combination of high temporal variability of rainfall and spatial heterogeneity of soil surface properties. In response to these environmental conditions, sources and sinks of runoff water and sediments tend to be organised in mosaics with distinct spatial attributes. These patterns can be identified at several scales, each with a predominance of a different set of processes. The dynamic relationships between these patterns and processes are an essential aspect of spatial connectivity in arid landscapes. During the last six years, part of the research at Rambla Honda, a field site in southeast Spain operating under the MEDALUS project, has been concerned with this subject. This paper reviews the results obtained up to date at the patch and the hillslope scales. The research at the patch scale focused on the role of vegetation as a source of spatial heterogeneity that affects short-range redistribution patterns of water and sediments. The approach has been to identify the dynamic relationships between plant clumps and bare ground in sparse vegetation mosaics, using field observations, experiments and simulation models. Field observations included runoff and sediment yield measurements on bounded plots and hillslope sectors, analysis of spatial correlation structures, as well as physiological and architectural properties of plant functional types. Experiments included rainfall simulation and runoff exclusion in the field, and soil fertility bioassays both in the field and the laboratory. A cellular automata model was built to explore the interactions between plant clumps and sediment movement. The research at the hillslope scale was concerned with the long-range transference of water and sediments from rocky upperslopes to their footslope sediment fill. The approach was based on an analysis of the available information about spatial patterns of soil moisture and discharge of runoff and sediments from plots and stream gauges in a first order catchment. Results show that, at the patch scale, in sparse vegetation, a range of positive feedback mechanisms lead to nucleation, or to the increase of spatial heterogeneity, by concentrating resources in the soil beneath plant clumps at the expense of the neighbouring bare ground. This spatial heterogeneity arises dynamically through the interaction between plant growth and hillslope fluxes of water and sediments. Within specific boundary conditions, this interaction is 'tuned' towards the formation of mosaics of bare and vegetated patches with patterns that minimise redistribution lengths of water and sediments. The boundary conditions that affect the 'tuning' process include factors that determine the potential distance and transport capacity of runoff, such as temporal variability of rainfall, slope angle, slope length, among others, and plant specific factors that affect the efficiency of plant clumps in trapping the resources that are redistributed on the hillslope. At the hillslope scale, the transference of sediment and water between hillslope elements requires very specific within-event temporal distributions of rainfall that allow for the widespread formation of a saturated layer at shallow depth and overland flow to reach first order channels. During most rainfall events these conditions are not met and, therefore, in most seasons; mean values of soil moisture do not increase downhill, and rather reflect variation in local soil properties than the effects of lateral redistribution processes. As a consequence, it may be expected that small changes of the frequency distribution of rainfall characteristics, in terms of within-storm temporal distribution of intensities, could lead to significant changes in soil moisture patterns and hydrologic connectivity between hillslope elements.
 
Purser, M. D., et al. (2003). Classification and Analysis of August 2001 Land Cover: Snohomish County, WA, Snohomish County, Surface Water Management: 14.
Pyles, M. R. and H. A. Froehlich (1987). "Discussion of "Rates of landsliding as impacted by timber managment activities in northwestern California", by M. Wolfe and J. Williams." Bulletin of the Association of Engineering Geologists 24(3): 425-431.
Pyles, M. R., et al. (1987). "Mechanics and stability of the Lookout Creek earth flow." Bulletin of the Association of Engineering Geologists XXIV(2): 267-280.
Pyles, M. R. and A. E. Skaugset (1998). Landslides and Forest Practice Regulation in Oregon. Environmental, Groundwater and Engineering Geology. S. Burns. Belmont, California, Star Publishing Co.: 481-488.
Pyne, S., P.A. Andrews, R. Laven (1996). Introduction to Wildland fire. New York, Wiley.
Quinn, P., K. Beven, P. Chevallier, and O. Planchon (1991). "The prediction of hillslope flow paths for distributed hydrological modelling using digital terrain models." Hydrological Processes 5: 59-79.
Quinn, T. P. and P. Peterson, N. (1996). "The influence of habitat complexity and fish size on over-winter survival and growth of individually makred juvenile coho salmon (Oncorhynchus kisutch) in Big Beef Creek, Washington." Canadian Journal of Fisheries and Aquatic Science 53: 1555-1564.
Wild juvenile coho salmon (Oncorhynchus kisutch) were individually marked in October 1990 and 1991 to
evaluate the effects of habitat complexity and fish size on over-winter survival in Big Beef Creek, Washington. Habitat
complexity was quantified for the habitat unit where the fish were collected and, in 1991, also for the 500-m reach
downstream from the collection site. Survival, estimated from recovery of marked smolts at the streamís mouth, differed
between years (25.4 and 46.2%) and also varied among habitat units and reaches within years. Survival was at most weakly
correlated with complexity of the habitat units but was strongly correlated with the quantity of woody debris and density of
habitat units in the 500-m reach, and distance from the estuary. Because distance covaried with habitat complexity, we could
not ascertain which factor had the primary influence on survival. In addition, larger fish generally survived at a higher rate
than smaller individuals. However, fish tagged above William Symington Lake were smaller in the fall but larger as smolts
and had higher survival rates than those tagged below the lake. Taken together, these results reveal complex relationships
between size, habitat, and growth that may affect over-winter survival and subsequent life-history events.
 
Railsback, S. and B. C. Harvey (2001). Individual-based model formulation for cutthroat trout, Little Jones Creek, California Albany, CA, US Forest Service: 64-68.
Railsback, S. and B. C. Harvey (2002). "Analysis of habitat-selection rules using an individual-based model  " Ecology 83(7): 1817-1830.
Raines, M. A. (1991). Sediment budget for the Grouse Creek Basin, Humboldt County, California Western Washington University: 90.
Ralph, S. C. (1971). Timber/Fish/Wildlife Stream Ambient Monitoring Field Manual. Seattle, Washington, Center for Streamside Studies: 40.
Ralph, S. C., et al. (1994). "Stream channel morphology and woody debris in logged and unlogged basins of western Washington." Canadian Journal of Fisheries and Aquatic Science 51: 37-51.
Ramirez, J. A. (2000). Prediction and modeling of flood hydrology and hydraulics. Inland flood hazards: human, riparian and aquatic communities. E. Wohl, Cambridge University Press.
Ramos, C. (1996). Quantification of Stream Channel Morphological Features: Recommended Procedures For Use In Watershed Analysis and TFW Ambient Monitoring prepared by University of California, Berkely for Northwest Indian Fisheries Commission: 89.
Ramos-Scharrón, C. E. and L. H. MacDonald (2006). "Development and application of a GIS-based sediment budget model." Journal of Environmental Management.
Accelerated erosion and increased sediment yields resulting from changes in land use are a critical environmental problem. Resource
managers and decision makers need spatially explicit tools to help them predict the changes in sediment production and delivery due to
unpaved roads and other types of land disturbance. This is a particularly important issue in much of the Caribbean because of the rapid
pace of development and potential damage to nearshore coral reef communities. The specific objectives of this study were to: (1) develop
a GIS-based sediment budget model; (2) use the model to evaluate the effects of unpaved roads on sediment delivery rates in three
watersheds on St. John in the US Virgin Islands; and (3) compare the predicted sediment yields to pre-existing data.
The St. John Erosion Model (STJ-EROS) is an ArcInfo-based program that uses empirical sediment production functions and delivery
ratios to quantify watershed-scale sediment yields. The program consists of six input routines and five routines to calculate sediment
production and delivery. The input routines have interfaces that allow the user to adjust the key variables that control sediment
production and delivery. The other five routines use pre-set erosion rate constants, user-defined variables, and values from nine data
layers to calculate watershed-scale sediment yields from unpaved road travelways, road cutslopes, streambanks, treethrow, and
undisturbed hillslopes.
 
STJ-EROS was applied to three basins on St. John with varying levels of development. Predicted sediment yields under natural
conditions ranged from 2 to 7Mgkm 2 yr 1, while yield rates for current conditions ranged from 8 to 46Mg km 2 yr 1. Unpaved roads
are estimated to be increasing sediment delivery rates by 3–6 times for Lameshur Bay, 5–9 times for Fish Bay, and 4–8 times for
Cinnamon Bay. Predicted basin-scale sediment yields for both undisturbed and current conditions are within the range of measured
sediment yields and bay sedimentation rates. The structure and user interfaces in STJ-EROS mean that the model can be readily adapted
to other areas and used to assess the impact of unpaved roads and other land uses sediment production and delivery.
 
Rapp, C. F. and T. B. Abbe (2003). A framework for delineating channel migration zones, Ecology Publication #03-06-027. Olympia, Washington, Washington State Department of Ecology.
Rastetter, E. B., A.W. King, B.J. Cosby, G.M. Hornberger, R.V. O'Neill, and J.E. Hobbie (1992). "Aggregating fine-scale ecological knowledge to model coarser-scale attributes of ecosystems." Ecological Applications 2(1): 55-70.
Rathburn, S. L., and E.E. Wohl (2001). "One-dimensional sediment transport modeling of pool recovery along a mouintain channel after a reservoir sediment release." Regulated Rivers: Research and Management 17: 251-273.
Rathburn, S. L., and E.E. Wohl (2003). "Predicting fine sediment dynamics along a pool-riffle mountain channel." Geomorphology 55: 111-124.
Read, W. W. and R. E. Volker (1993). "Series solutions for steady seepage through hillsides with arbitrary flow boundaries." Water Resources Research 29(8): 2871-2880.
Redding, J. M., et al. (1987). "Physiological Effects on Coho Salmon and Steelhead of Exposure to Suspended Solids " Transactions of the American Fisheries Society 116: 737-744.
Reed, W. J. (1994). "Estimating the historic probability of stand-replacement fire using the age-class distribution of undisturbed forest." Forest Science 40(1): 104-119.
Reedyk, S., et al. (1997). "Design for an inexpensive continuous digital output water level recorder." Water Resources Research 33(6): 1523-1526.
Reeves, G., et al. (2003). "Sources of large wood in the mainstem of a fourth-order watershed in coastal Oregon." Canadian Journal of Forest Research 33(8): 1352-1362.
Reeves, G., et al. (2006). "Postfire logging in riparian areas." Conservation Biology 20(4): 994-1004.
Reeves, G. H. (?). "The Impact of Catastrophic Disturbance on Freshwater Habitats of Anadromous Salmonids in the Pacific Northwest."?
Reeves, G. H., et al. (1995). A disturbance-based ecosystem approach to maintaining and restoring freshwater habitats of evolutionarily significant units of anadromous salmonids in the Pacific Northwest. Evolution and the Aquatic Ecosystem: Defining Unique Units in Population Conservation, American Fisheries Society Symposium 17. J. L. Nielson and D. A. Powers. Bethesda, Maryland, USA, American Fisheries Society: 334-349.
Reeves, G. H., et al. (1998). Fish Communities. River ecology and management: lessons from the Pacific coastal ecosystem. R. J. Naiman and R. E. Bilby. New York, Springer: 200-234.
Reeves, G. H., et al. (2006). "Postfire logging in riparian areas." Conservation Biology 20(4): 994-1004.
We reviewed the behavior of wildfire in riparian zones, primarily in the western United States, and
the potential ecological consequences of postfire logging. Fire behavior in riparian zones is complex, but many
aquatic and riparian organisms exhibit a suite of adaptations that allow relatively rapid recovery after fire.
Unless constrained by other factors, fish tend to rebound relatively quickly, usually within a decade after a
wildfire. Additionally, fire and subsequent erosion events contribute wood and coarse sediment that can create
and maintain productive aquatic habitats over time. The potential effects of postfire logging in riparian areas
depend on the landscape context and disturbance history of a site; however, available evidence suggests two
key management implications: (1) fire in riparian areas creates conditions that may not require intervention
to sustain the long-term productivity of the aquatic network and (2) protection of burned riparian areas gives
priority to what is left rather than what is removed. Research is needed to determine how postfire logging
in riparian areas has affected the spread of invasive species and the vulnerability of upland forests to insect
and disease outbreaks and how postfire logging will affect the frequency and behavior of future fires. The
effectiveness of using postfire logging to restore desired riparian structure and function is therefore unproven,
but such projects are gaining interest with the departure of forest conditions from those that existed prior to
timber harvest, fire suppression, and climate change. In the absence of reliable information about the potential
consequence of postfire timber harvest, we conclude that providing postfire riparian zones with the same
environmental protections they received before they burned is justified ecologically. Without a commitment to
monitor management experiments, the effects of postfire riparian logging will remain unknown and highly
contentious.
 
 
Reeves, G. H., et al. (2002). Fish and aquatic ecosystems of the Oregon Coast Range. Forest and Stream Management in the Oregon Coast Range. S. D. Hobbs, J. P. Hayes, R. L. Johnson et al. Corvallis, Oregon State University Press: 68-98.
Reeves, G. H., et al. (2003). "Sources of large wood in the main stem of a fourth-order watershed in coastal Oregon." Canadian Journal of Forest Research 33: 1363-1370.
We compared the contribution of large wood from different sources and wood distributions among channel
zones of influence in a relatively pristine fourth-order watershed in the central Coast Range of Oregon. Wood in the
main stem of Cummins Creek was identified as coming from either (i) streamside sources immediately adjacent to the
channel or (ii) upslope sources delivered by landslides or debris flows more than 90 m from the channel. About 65%
of the number of pieces and 46% of the estimated volume of wood were from upslope sources. Streamside sources
contributed about 35% of the number of pieces and 54% of the estimated volume of wood. The estimated mean volume
of upslope-derived pieces was about one-third that of streamside-derived pieces. Upslope-derived pieces were located
primarily in the middle stream reaches and in the zones of influence that had the most contact with the low-flow
channel. Streamside-derived pieces were more evenly distributed among the examined reaches and were predominately
in the influence zones that had the least contact with the low-flow channel. Our findings suggest that previous studies
that examined only streamside sources of wood have limited applications when designing and evaluating riparian management
approaches in landslide-prone areas. The failure to recognize the potential sources of wood from upslope areas
is a possible reason for the decline of large wood in streams in the Pacific Northwest.
 
Reeves, G. H., et al. (1989). Identification of Physical Habitats Limiting the Production of Coho Salmon in Western Oregon and Washington, USDA Forest Service: 1-17.
Reeves, G. H., et al. (1993). "Diversity of Juvenile Anadromous Salmonid Assemblages in Coastal Oregon Basins with Different Levels of Timber Harvest " Transactions of the American Fisheries Society 122(3): 309-317.
Refsgaard, J. C. (2000). Towards a formal approach to calibration and validation of models using spatial data. Spatial Patterns in catchment hydrology: observations and modeling. R. Grayson, and B.G. Bloschl. Cambridge, UK, Cambridge University Press: 404.
Regional Interagency Executive Committee (1995). Ecosystem Analysis at the Watershed Scale: Federal Guide for Watershed Analysis. Portland, Oregon, U.S. Forest Service, U.S. Bureau of Land Management, U.S. Bureau of Indian Affairs, U.S. Fish and Wildlife Service, U. S. Environmental Protection Agency, et al.: 26.
Reice, S. R. (1994). "Nonequilibrium determinants of biological community structure." American Scientist 82: 424-435.
Reid, I., and J.B. Laronne (1995). "Bedload sediment transport in an ephemeral stream and a comparison with seasonal and perennial counterparts." Water Resources Research 31(3): 773-782.
Reid, I., et al. (1985). "The incidence and nature of bedload transport during flood flows in coarse-grained alluvial channels." ESPL 10: 33-44.
A continuous record reveals that the incidence of bedload in a coarse-grained channel changes from flood to flood. Long periods of inactivity encourage the channel bed to consolidate sufficiently so that bedload is largely confined to the recession limb of the next flood-wave. But ewhen floods follow each other closely, the bed material is comparatively loose and offers less resistance to entrainment. In this case, substantial amounts of bedload are generated on the rising limb. This is confirmed by values of bed shear stress or stream power a the threshold of initial motion which can be up to five times the overall mean in the case of isolated floods or those which are the first of the season. This produces a complicated relationship between flow parameters and bedload and explains some of the difficulties in establishin bedload rating curves for coarse-grained channels. Besides this, the threshold of initial motion is shown to occur at levels of bed shear stress three times those at the thresholds of final motion. This adds further confusion to attempts at developing predictive bedload equations and clearly indicates at least one reason why equations currently in use are unsatisfactory. Bedload is shown to be characterized by a series of pulses with a mena periodicity of 1-7 hours. In the absence of migrating bedfoms, it is speculated that this weee-documented pattern reflects the pasage of kinematic waves of particles ina slow-moving traction carpet. The general pattern of bedload, including pulsations, is shown to occur more or less synchronously at different points across the stream channel.
 
Reid, L. (1998). Cumulative watershed effects and watershed analysis. Pacific Coastal Ecoregion. R. Naiman and R. Bilby: 476-501.
Reid, L. M. (1981). Sediment Production from gravel-surfaced forest roads, Clearwater Basin, Washington. Seattle, Washington, University of Washington: 247.
Reid, L. M. (1993). Research and cumulative effects. USDA Forest Service General Technical Report PSW-GTR-141. USDA Forest Service General Technical Report PSW-GTR-141. Albany, CA.: 118pp.
Reid, L. M. (1993). Research and Cumulative Watershed Effects US Forest Service: 118.
Reid, L. M. (1998). "Calculation of average landslide frequency using climatic records." Water Resources Research 34(4): 869-877.
Reid, L. M. (1998). Review of the Sustained Plan/Habitat Conservation Plan for the properties of the Pacific Lumber Company, Scotia Pacific Holding Company, and Salmon Creek Corporation G. Miller. Washington, DC.
Reid, L. M. (1999). Forest Practice Rules and cumulative watershed impacts in California F. Keeley. Sacramento, CA.
Reid, L. M. (2001). "Cumulative Watershed Effects: Then and Now." Watershed Management Council Networker 10(1): 24-33.
Reid, L. M. (2006). Understanding and evaluating cumulative watershed impacts. CWE for fuels management in the Western U.S. W. J. Elliot, and L.J. Audin.
Reid, L. M. and T. Dunne (1988). "Sediment production from forest road surfaces." Water Resources Research 20(11): 1753-1761.
Reid, L. M. and T. Dunne (1996). Rapid Construction of Sediment Budgets for Drainage Basins. Cremlingen, Germany, Catena-Verlag.
Reid, L. M. and T. Dunne (1996). Rapid Evaluation of Sediment Budgets. Reiskirchen, Germany, Catena Verlag.
Reid, L. M. and T. Dunne (?). "Rapid evaluation of sediment budgets."
Reid, L. M., et al. (1981). "Application of Sediment budget Studies to the Evaluation of Logging Road Impact." Journal of Hydrology 29: 49-62.
Reid, L. M. and M. J. Furniss (2002). On the use of regional channel-based indicators for monitoring, USDA Forest Service.
Reid, L. M. and S. Hilton (1998). Buffering the buffer, USDA Forest Service: 71-80.
Riparian buffer strips are a widely accepted tool for helping to sustain
aquatic ecosystems and to protect downstream resources and values in forested
areas, but controversy persists over how wide a buffer strip is necessary. The
physical integrity of stream channels is expected to be sustained if the
characteristics and rates of tree fall along buffered reaches are similar to those in
undisturbed forests. Although most tree-fall-related sediment and woody debris
inputs to Caspar Creek are generated by trees falling from within a tree’s height
of the channel, about 30 percent of those tree falls are triggered by trees falling
from upslope of the contributing tree, suggesting that the core zone over which
natural rates of tree fall would need to be sustained is wider than the one-treeheight’s-
width previously assumed. Furthermore, an additional width of “fringe”
buffer is necessary to sustain appropriate tree-fall rates within the core buffer.
Analysis of the distribution of tree falls in buffer strips and un-reentered streamside
forests along the North Fork of Caspar Creek suggests that rates of tree fall
are abnormally high for a distance of at least 200 m from a clearcut edge, a
distance equivalent to nearly four times the current canopy height. The
appropriate width of fringe buffer needed to protect the core zone will need to be
determined using an analysis of the long-term effects and significance of
accelerated tree-fall rates after logging.
 
Reid, L. M. and M. J. Page (2002). "Magnitude and frequency of landsliding in a large New Zealand catchment." Geomorphology 49: 71-88.
Knowledge of long-term average rates of erosion is necessary if factors affecting sediment yields from catchments are to be understood. Without such information, it is not possible to assess the potential influence of extreme storms, and, therefore, to evaluate the relative importance of various components of a sediment budget. A study of the sediment budget for the Waipaoa catchment, North Island, New Zealand, included evaluation of long-term rates of landsliding for six landslide-prone land systems in the catchment. The number of landslides per unit area generated by each of several storms was counted on sequential aerial photographs and correlated with the magnitude of the corresponding storm. The resulting relationships were combined with magnitude–frequency relationships derived for storms from 70- to 100-year rainfall records in the area to estimate a long-term magnitude–frequency relationship for landsliding for each land system. The long-term average values of the areal landslide frequency (number of slides per unit area per unit time) were then calculated from these relationships. The volumes of a sample of landslide scars were measured in the field, and the proportion of slides that deliver sediment to channels was determined from aerial photographs. These measurements then allowed calculation of the long-term average rate of sediment production to streams from landslides for different land systems and types of vegetation. Results suggest that shallow landslides currently contribute about 15% of the suspended sediment load in the Waipaoa River above the Kanakanaia gauging station, and that 75% of the sediment production from the landslides occurs during storms with recurrence intervals of less than 27 years. Reforestation of 6.3% (93 km2) of the slide-prone lands in the catchment between 1990 and 1995 resulted in a calculated decrease in slide-derived sediment of 10%. Calculations suggest that reforestation of an additional 3% (66 km2) of the catchment in areas with the most sensitive combinations of land system and storm regime could decrease the total sediment inputs from landsliding by about 20%.
 
Reid, L. M. and R. R. Ziemar (1988). Cumulative Watershed Effects: Problem and Approach (DRAFT), Pacific Southwest Forest and Range Experiment Station: 9.
Reid, L. M. and R. R. Ziemer (1998). Cumulative watershed effects; Caspar Creek and beyond. General Technical Report PSW, Report: PSW-GTR-168: 117-127.
Cumulative effects are the combined effects of multiple activities, and watershed effects are those which involve processes of water transport. Almost all impacts are influenced by multiple activities, so almost all impacts must be evaluated as cumulative impacts rather than as individual impacts. Existing definitions suggest that to be significant, an impact must be reasonably expected to have occurred or to occur in the future, and it must be of societally validated concern to someone or influence their activities or options. Past approaches to evaluating and managing cumulative watershed impacts have not yet proved successful for averting these impacts, so interest has grown in how to regulate land-use activities to reverse existing impacts. Approaches being discussed include requirements for "zero net increase" of sediment, linkage of planned activities to mitigation of existing problems, use of more protective best management practices, and adoption of thresholds for either land-use intensity or impact level. Different kinds of cumulative impacts require different kinds of approaches for management. Efforts are underway to determine how best to evaluate the potential for cumulative impacts, and thus to provide a tool for preventing future impacts and for determining which management approaches are appropriate for each issue in an area. Future impact analysis methods probably will be based on strategies for watershed analysis. Analysis would need to consider areas large enough for the most important impacts to be evident; to evaluate time scales long enough for the potential for impact accumulation to be identified; and to be interdisciplinary enough that interactions among diverse impact mechanisms can be understood.
 
Reid, L. M. B. P. M. (1993). "A procedure for watershed analysis. Review draft prepared for the Interagency Forest Ecosystem Management Assessment Team." 139.
Reid, M. E., S.D. Ellen, D.L. Brien, J. de la Fuente, J.N. Falls, B.G. Hicks, and T.E. Koler (2001). "Comparison of topographic models for predicting debris-slide locations." Eos Trans. AGU,  Fall Meet. Suppl., Abstract #H22G-09. San Francisco, California.
Reid, M. E., et al. (2003). Debris-flow initiation from large, slow-moving landslides. Debris-Flow Hazards Mitigation: Mechanics, Prediction, and Assessment. Rickenmann and Chen. Rotterdam, Millpress: 155-166.
In some mountainous terrain, debris flows preferentially initiate from the toes and margins of larger, deeper, slower-moving landslides. During the wet winter of 1997, we began real-time monitoring of the large, active Cleveland Corral landslide complex in California, USA. When the main slide is actively moving, small, shallow, first-time slides on the toe and margins mobilize into debris flows and travel down adjacent gullies. We monitored the acceleration of one such failure; changes in velocity provided precursory indications of rapid failure. Three factors appear to aid the initiation of debris flows at this site: 1) locally steepened ground created by dynamic landslide movement, 2) elevated pore-water pressures and abundant soil moisture, and 3) locally cracked and dilated materials. This association between debris flows and large landslides can be widespread in some terrain. Detailed photographic mapping in two watersheds of northwestern california illustrates that the areal density of debris-flow source landsliding is about 3 to 7 times greater in steep geomorphically fresher landslide deposits than in steep ground outside landslide deposits
 
Reid, M. E., et al. (2001). "Comparison of topographic models for predicting debris-slide locations." American Geophysical Union, Fall Meeting 2001, abstract #H22G-09.
Shallow, rapid debris slides often sculpt steep hillslopes and deliver sudden pulses of sediment to streams. In northwestern California, debris slides occur on steep slopes, on steep inner gorges near streams, in steep convergent swales, or on steep parts of larger landslides. As part of a study on mass-wasting in forested terrain, we tested four topographic models aimed at identifying locations of debris-slide sources. We selected three study areas covering a spectrum of geologic materials, topography, and landslide processes; all have experienced Holocene tectonic deformation, large rainstorms, strong earthquake shaking, and > one hundred years of timber-harvest activities. At least three separate inventories of non-road related debris-slides and a comprehensive air-photo inventory of larger landslides were conducted by different investigators for each study area. We implemented the models in a GIS using USGS 10m DEMs. The four models are: 1) local slope, 2) proximity to stream, 3) SHALSTAB with cohesionless, "standard" parameters, and 4) debris-slide landforms. The latter model identifies two landforms prone to debris slides: steep, concave areas (swales) that lead directly to streams, and steep slopes contiguous to streams. None of the models were calibrated to a specific lithology or terrain. We compare model success at predicting locations of past, non-road related, debris-slides using the ratio of two criteria, 1) the areal percentage of debris slides captured in a hazard class, and 2) the percentage of the landscape taken up in that class. Our method differs from that used in the current SHALSTAB version, wherein an entire slide area is attributed to the highest hazard class having any DEM cell within the slide. Our method makes no assumptions about initiation and allows rational comparison between models. All of the models demonstrate a similar pattern of predictive success, however there are significant differences between model performance in different terrain. For example, in topography with scalloped or amphitheater morphology, the slope model performs well, whereas in inner-gorge topography the proximity-to-stream model performs well. SHALSTAB predictive success is rarely superior in any of the settings; a LISA-based analysis available for one study area is likewise not superior. The debris-slide landforms model exhibits the highest predictive success (typically 10-30% better) for nearly all the debris-slide data sets in all three study areas. Also, for data sets containing only large-volume debris slides that delivered sediment to a stream, the landforms model has even better predictive success (up to 75% better), in part because the landforms incorporate pathways to the channel network. Thus this model performs well overall, and excels at predicting large sediment delivery to streams. In addition to the modeled debris-slide landforms, the presence of large, geomorphically recent landslides influences the propensity for debris sliding. In some terrain, steep areas contiguous to streams and located within these large landslides show more than twice the spatial density of historical debris-slide sources as similar areas outside these landslides. In northwestern California, it appears that the combination of debris-slide landforms and large landslides offers a strong predictor for locating future debris-slide activity.
 
Reid, M. E. and R. M. Iverson (1992). "Gravity-driven groundwater flow and slope failure potential, 2. Effects of slope morphology, material properties, and hydraulic heterogeneity." Water Resources Research 28(3): 939-950.
Reid, M. E., et al. (1988). "Hydrologic factors triggering a shallow hillslope failure." Bulletin of the Association of Engineering Geologists 25(3): 349-361.
Reid, S. C., S.N. Lane, D.R. Montgomery, and C. J. Brookes (2007). "Does hydrological connectivity improve modelling of coarse sediment delivery in upland environments." Geomorphology 90: 263-282.
Reinfelds, I., T. Cohen, P. Batten, and G. Brierley (2004). "Assessment of downstream trends in channel gradient, total and specific stream power: a GIS approach." Geomorphology 60: 403-416.
Reisenbichler, R. R. and J. D. McIntyre (1977). "Genetic Differences in the Growth and Survival of Juvenile Hatchery and Wild Steelhead Trout, Salmo gairdneri " J. Fish Res. Board Can. 34: 123-128.
Reiser, D. W. and T. C. Bjornn (1979). Habitat Requirements of Anadromous Salmonids. Portland, Oregon, USDA Forest Service: 54.
Reiser, D. W. and R. G. White (1983). "Effects of Complete Redd Dewatering on Salmonid Egg-Hatching Success and Development of Juveniles " Transactions of the American Fisheries Society 112: 532-540.
Reiser, D. W. and R. G. White (1988). "Effects of Two Sediment Size-Classes on Survival of Steelhead and Chinook Salmon Eggs " North American Journal of Fisheries Management 8: 432-437.
Renard, K. G., G.R. Foster, G.A. Weesies, D.K. McCool, and D.C. Yoder (1997). Predicting soil erosion by water: a guide to conservation planning with the revised universal soil loss equation (RUSLE).
Reneau, D. K. and W. E. Dietrich (1991). "Erosion Rates in the Southern Oregon Coast Range: Evidence for an Equilibrium Between Hillslope and Sediment Yield." Earth Surface Processes and Landforms 16: 307-322.
Reneau, S. I. (1993). "Manganese Accumulation in Rock Varnish on a Desert Piedmont, Mojave Desert, California, and Application to Evaluating Varnish Development." Quaternary Research 40: 309-317.
Reneau, S. I. and W. E. Dietrich (1987). "The importance of hollows in debris flow studies: Examples from Marin County, California." Reviews of Engineering Geology 7: 26.
Reneau, S. I. and W. E. Dietrich (1987). Size and location of colluvial landslides in a steep forested landscape. International Symposium on Erosion and Sedimentation in the Pacific Rim.
Reneau, S. I., et al. (1990). "late Quaternary history of colluvial deposition and erosion in hollows, central California Coast Ranges." Geological Society of America Bulletin 102: 969-982.
Reneau, S. L. (1988). Depositional and erosional history of hollows: application to landslide location and frequency, long-term erosion rates and the effects of climatic change, Berkeley, University of California: 328.
Reneau, S. L. and W. E. Dietrich (1987). Size and location of colluvial landslides in a steep forested landscape. International Symposium on Erosion and Sedimentation in the Pacific rim, Corvallis, Oregon, International Association of Hydrological Sciences Bulletin.
Reneau, S. L. and W. E. Dietrich (1987). Size and location of colluvial landslides in a steep forested landscape. International Symposium on Erosion and Sedimentation in the Pacific Rim. B. Beschta. Corvallis, Oregon, International Association of Hydrological Sciences Bulletin. 165.
Reneau, S. L. and W. E. Dietrich (1990). "Depositional History of Hollows on Steep Hillslopes, Coastal Oregon and Washington." National Geographic Research 6(2): 220-230.
Reneau, S. L. and W. E. Dietrich (1991). "Erosion rates in the southern Oregon Coast Range: Evidence for an equilibrium between hillslope erosion and sediment yield." Earth Surface Processes and Landforms 16: 307-322.
Reneau, S. L., et al. (1990). "Late Quaternary history of colluvial deposition and erosion in hollows, central California Coast Ranges." Geological Society of America Bulletin 102: 969-982.
Reneau, S. L., et al. (2007). "Sediment delivery after a wildfire." Geology 35(2): 151-154.
We use a record of sedimentation in a small reservoir within the
Cerro Grande burn area, New Mexico, to document postfi re delivery
of ash, other fi ne-grained sediment carried in suspension within fl oods,
and coarse-grained sediment transported as bedload over a fi ve-year
period. Ash content of sediment layers is estimated using fallout 137Cs
as a tracer, and ash concentrations are shown to rapidly decrease
through a series of moderate-intensity convective storms in the fi rst
rainy season after the fi re. Over 90% of the ash was delivered to the
reservoir in the fi rst year, and ash concentrations in suspended sediment
were negligible after the second year. Delivery of the remainder
of the fi ne sediment also declined rapidly after the fi rst year despite
the occurrence of higher-intensity storms in the second year. Fine
sediment loads after fi ve years remained signifi cantly above prefi re
averages. Deposition of coarse-grained sediment was irregular in time
and was associated with transport by snowmelt runoff of sediment
stored along the upstream channel during short-duration summer
fl oods. Coarse sediment delivery in the fi rst four years was strongly
correlated with snowmelt volume, suggesting a transport-limited system
with abundant available sediment. Transport rates of coarse sediment
declined in the fi fth year, consistent with a transition to a more
stable channel as the accessible sediment supply was depleted and the
channel bed coarsened. Maximum impacts from ash and other fi negrained
sediment therefore occurred soon after the fi re, whereas the
downstream impacts from coarse-grained sediment were attenuated
by the more gradual process of bedload sediment transport.
 
Rennie, C. D. (1998). Bed mobility of gravel rivers: Mobilization (scour) depth of chum salmon redds and equilibrium bedload transport Department of Civil Engineering, Water Resources, University of British Columbia: 169.
Rennie, C. D. and R. G. Millar (2000). "Spatial variability of stream bed scour and fill: a comparison of scour depth in chum salmon (Oncorhynchus keta) redds and adjacent bed " Canadian Journal of Fisheries and Aquatic Sciences 57: 928-938.
Renschler, C. S. a. J. H. (2002). "Soil erosion assessment tools from point to regional scales-the role of geomorphologists in land management research and implementation." Geomorphology 47: 189-209.
Resh, V. H., et al. (1988). "The role of disturbance in stream ecology." Journal of the North American Benthological Society 7: 433-455.
Resh, V. H., et al. (1988). "The role of disturbance in stream ecology." The North American Benthological Society.
Resources, W. D. o. N. (1997). Methods for Conducting Watershed Analysis, Washington Forest Practices Board.
Reuter, J. E., et al. (1992). "The Use of Wetlands for Nutrient Removal from Surface Runoff in a Cold Climate Region of California - Results from a Newly Constructed Wetland at Lake Tahoe " Journal of Environmental Management 36: 35-53.
Rey, F. (2003). "Influence of vegetation distribution on sediment yield in forested marly gullies." Catena 50(2-4): 549-562.
The cover provided by forest vegetation makes it possible to fight erosion efficiently. Furthermore, vegetation barriers can also play a major role by trapping eroded sediments and thus vegetation located downslope of a gully can be important. The objective of this study is to highlight the role of vegetation distribution in a marly gully in reducing sediment loss at the outlet. To do this, sediment yield in gullies with similar geological and geomorphological components, but with different vegetation cover and distribution, was studied for 2 years. Sediment traps installed at gully outlets permitted the distinction of active gullies from inactive ones. The results show that gully activity is not correlated with the percentage of total vegetation cover, but with the percentage of cover of low vegetation in the gully floor. This is due to sediment trapping processes and highlights the importance of the spatial distribution of forest vegetation in reducing sediment yield at the gully outlet. Thus, gullies with similar total vegetation cover can have very different activities, and inactive gullies exist that are only partially covered with vegetation. Inactive gullies with as low as 33% vegetation cover occur where vegetation is mainly present in the gully floor. The results also show that above 50% cover of low vegetation in the gully floor as a percentage of the gully floor surface, gullies are generally inactive.
 
Rhea, S. (1993). "Geomorphic observations of rivers in the Oregon Coast Range from a regional reconnaissance perspective." Geomorphology 6: 135-190.
Rhee, H., et al. (2000). A road sediment analysis tool. Water Resources Center Report - Centers for Water and Wildland Resources, Report: 98. C. W. Slaughter: 142.
Rhoads, B. L. (1987). "Changes in stream characteristics at tributary junctions." Physical Geography 8(4): 346-361.
Rhoads, B. L. (1987). "Stream power terminology." Professional Geographer 39(2): 189-195.
Rhodes, D. D. (1987). "The b-f-m diagram for downstream hydraulic geometry." Geografiska Annaler 69A(1): 147-161.
Rhodes, J. (1980). Soil and Water Characteristics Affecting Soil Liquefacation Potential (DRAFT): 9.
Rhodes, J. J., et al. (1994). A coarse screening process for evaluation of the effects of land management activities on salmon spawning and rearing habitat in ESA consultations, Columbia River Inter-Tribal Fish Commission: 1-127.
Rhodes, J. J. and M. Purser (1998). Thinning for Increased Water Yield in the Sierra Nevada: Free Lunch or Pie in the Sky? Albany, California, Pacific Rivers Council: 1-25.
Rial, J. A. (1999). "Pacemaking the Ice Ages by Frequency Modulation of Earth's Orbital Eccentricity " Science 285: 564-568.
Ribberink, J. S. (1987). Summary and Conclusions. Mathematical modelling of one-dimensional morphological changes in rivers with non-uniform sediment. Chapter 9: 159-166.
Rice, M. R., R.R. Ziemer, and J. Lewis (2000). Forest management effects on erosion, sediment, and runoff: lessons from Caspar Creek and northwestern California. Proceedings of the SAF 2000 National Convention, Washington D.C.
Rice, R. M. (1973). The hydrology of chaparral watersheds. Living with chaparral, Riverside, California, Sierra Club.
Rice, R. M. and P. A. Datzman (1981). Erosion associated with cable and tractor logging in northwestern California. Erosion and Sediment Transport in Pacific Rim Steeplands. Christchurch, I. A. H. S. Publ. No. 132: 362-373.
Rice, R. M., et al. (2008). The ecological importance of tributaries and confluences. River Confluences, Tributaries and the Fluvial Network. R. M. Rice, A. G. Roy and B. L. Rhoads, John Wiley and Sons Ltd.: 209-237.
Rice, R. M., et al. (1971). Erosional consequences of timber harvest: an appraisal. National Symposium on Watersheds in Transition, American Water Resources Association.
Rice, S. (1998). "Which tributaries disrupt downstream fining along gravel-bed rivers?" Geomorphology 22: 39-56.
Tributaries and other lateral sediment sources can have a significant impact on river bed sediment texture and, in turn, on
channel form. Sufficiently voluminous or distinct sediment inputs redefine the mainstem grain-size distribution, punctuating
downstream maturation and isolating a sequence of discrete sedimentary links. Within these links fining processes usually
dominate, such that models of sorting and abrasion, when applied to individual links, provide reasonable predictions of
grain-size change. Links represent the fundamental natural unit within which fining models can be tested, developed and
applied. Identification of significant lateral sources is therefore important, yet, beyond vague references to relative tributary
size, sediment load, and sediment calibre, no criteria exist for the a priori discrimination of such sources. In this paper a
procedure for identifying significant lateral (tributary) sources, without the benefit of grain-size information, is outlined. A
high-resolution characterisation of bed material texture along two Canadian gravel-bed rivers facilitated classification of all
their perennial tributaries as either significant or insignificant. Three absolute tributary basin parameters and their relative
counterparts, chosen tcl reflect the likely controls on tributary significance, are then used to develop a discriminant function
which isolates a large proportion of significant tributaries while minimising incorrect classifications. Examination of
consistently misclassifi.ed (anomalous) tributaries reveals the importance of lateral source spacing and of inconsistencies in
the geomorphic history of the contributing basins. In turn, a general tributary categorisation procedure is suggested which
includes a logistic regression model for attaching probability statements to individual classifications. The generality of the
discriminant and logistic functions cannot be assessed because of the lack of other suitable data sets.
 
Rice, S. and M. Church (1998). "Grainsize along two gravel-bed rivers: Statistical variation, spatial pattern and sedimentary links." Earth Surface Processes and Landforms 23: 345-363.
Rice, S. P., et al. (2006). "Tributary control of physical heterogeneity and biological diversity at river confluences." Canadian Journal Fisheries and Aquatic Sciences 63: 2553-2566.
Rice, S. P., et al. (2001). "Macroinvertebrate community changes at coarse sediment recruitment points along two gravel bed rivers." Water Resources Research 37(11): 2793-2803.
Rice, S. P., et al. (2001). "Tributaries, sediment sources, and the longitudinal organisation of macroinvertebrate fauna along river systems." Canadian Journal of Fisheries and Aquatic Sciences 58: 824-840.
Rich, C. F., Jr., et al. (2003). "Local-habitat, watershed, and biotic features associated with Bull Trout occurrence in Montana streams." Transactions of the American Fisheries Society 132: 1053-1064.
We evaluated the association of local-habitat features, large-scale watershed factors,
the presence of nonnative brook trout Salvelinus fontinalis, and connectivity to neighboring populations
with patterns of occurrence of threatened bull trout S. confluentus in 112 first-order to fourthorder
streams in the Bitterroot River drainage in western Montana. Species presence or absence was
estimated via single-pass electrofishing, local-habitat features were measured in 500-m sampling
reaches, watershed variables were obtained from topographic maps, and potential demographic support
from nearby bull trout populations occupying larger main-stem streams was estimated from
electrofishing data records. We defined a set of nine candidate models that represented various
combinations of these four main factors and used an information-theoretic approach to evaluate the
relative plausibility of competing models. Models combining local habitat (width, gradient, and woody
debris) with brook trout presence and the main-stem abundance of bull trout and a global model (all
variables) were the best approximating models. In contrast, watershed models based on elevation,
basin area, and tributary slope and models with local-habitat or biotic variables alone were poor
predictors of bull trout occurrence. Bull trout occurrence was positively associated with channel
width, large woody debris, and the presence of a ‘‘strong’’ neighboring main-stem population and
negatively associated with channel gradient and the presence of brook trout. Our findings suggest
that bull trout have increased resistance to invasion by brook trout in streams with high habitat
complexity and connectivity. Consideration of abiotic and biotic factors at multiple scales, along
with a means for ranking their relative importance, is needed to perform more comprehensive assessments
of landscape and local influences on species distribution patterns.
 
 
Richard, G. A., et al. (2005). "Statistical analysis of lateral migration of the Rio Grande, New Mexico." Geomorphology 71: 139-155.
The lateral migration rates of alluvial rivers are affected by changes in water and sediment regimes. The Rio Grande
downstream from Cochiti Dam exhibits spatial and temporal variability in lateral movement rates documented since 1918. A
tremendous database exists that documents planform, bed material size, channel geometry, and water and sediment regimes. A
statistical analysis reveals that migration rates primarily decreased with decreasing flow energy (R2N0.50, p b0.0001). The
addition of a second parameter describing total channel width increased the explained variance to N60%. The findings show that
lateral movement increases with increasing flow energy and with degree of braiding.
 
Richards, C., et al. (1996). "Landscape-scale influences on stream habitats and biota." Canadian Journal of Fisheries and Aquatic Science 53(Suppl. 1): 295-311.
The relative influence of geologic versus anthropogenic attributes of catchments on stream ecosystems was
examined in 45 catchments of a river basin in central Michigan. Each catchment was characterized by land use, surficial
geology, elevation, and hydrography, and summaries of these data were related to physical habitat characteristics that had the
greatest influence on macroinvertebrate assemblages. Partial redundancy analysis revealed that geologic and land-use
variables had similar magnitudes of influence on stream habitats. Of the geologic variables, catchment area, proportion of
lacustrine clays, and glacial outwash materials had the strongest influence on physical habitat, particularly on channel
dimensions. Row-crop agriculture and the presence of wetlands were the most important land-use variables, particularly
influencing amounts of woody debris. Stream buffers (100 m) were more important than whole catchment data for predicting
sediment-related habitat variables; however, channel morphology was more strongly related to whole catchments. Results
suggest that catchment-wide geology and land-use characteristics may be more important than stream buffers for maintaining
or restoring stream ecosystems. These techniques can be used to develop biologic signatures of catchment condition that
discriminate causal factors influencing the biodiversity and health of stream ecosystems.
 
Richards, K. (1982). Rivers: Form and Process in Alluvial Channel, Methuen and Company.
Richards, K., Ed. (1987). River Channels: Environment and Process. New York, Basil Blackwell Inc.
Richards, K., et al. (2002). "Geomorphic dynamics of floodplains: ecological implications and a potential modelling strategy." Freshwater Biology 47(4): 559-579.
Richards, K. S. (1980). "A Note on Changes in Channel Geometry at Tributary Junctions." Water Resources Research 16(1): 241-244.
Richardson, D. (1968). "Glacier Outburst Floods in the Pacific Northwest." U.S. Geological Survey 600-D: D79-D86.
Richardson, J. S., R. J. Naiman, P. A. Bisson (2012). "How did fixed-width buffers become standard practice for protecting freshwaters and their riparian areas from forest harvest practices." Freshwater Science 31(1): 232-238.
Richardson, J. S. and R. J. Danehy (2007). "A synthesis of the Ecology of Headwater Streams and their Riparian Zones in Temperate Forests." Forest Science 53(2): 131-147.
Richmond, A. D. and K. D. Fausch (1995). "Characteristics and function of large woody debris in subalpine Rocky Mountain streams in northern Colorado." Canadian Journal of Fisheries and Aquatic Sciences 52: 1789-1802.
Rickenmann, D. (1997). "Sediment transport in Swiss torrents." Earth Surface Processes and Landforms 22: 937-951.
Rickenmann, D. (1999). "Empirical relationships for debris flows." Natural Hazards 19: 47-77.
The assessment of the debris flow hazard potential has to rely on semi-quantitative methods. due to the complexity of the debris-flow process, numerical simulation models of debris flows are still limited with regard to practical applications. Thus, an overview is given of empirical relationships that can be used to estimate the most important parameters of debris-flow behavior. In a possible procedure, an assessment of a maximum debris-flow volume may be followed by estimates of the peak discharge, the mean flow velocity, the total travel distance, and the runout distance on the fan. the applicability of several empirical equations is compared with available field and laboratory data, and scaling considerations are used to discuss the vvariability of the parameters over a large range of values. some recommendations are made with regard to the application of the presented relationships by practicing engineers, apart from advocating field reconnaissance and searching for historic events wherever possible.
 
Ricklefs, R. E. (1987). "Community diversity: relative roles of local and regional processes." Articles 235: 167-171.
Ricks, C. and G. Chen (1990). Elk River Drainage Basin (DRAFT), USDA Forest Service, Siskiyou National Forest: 31.
Riebe, C. A., et al. (2000). "Erosional equilibrium and disequilbrium in the sierra Nevada, inferred from cosmogenic 26Al and 10Be in alluvial sediment." Geology 28(9): 803-806.
Riebe, C. A., et al. (2001). "Strong tectonic and weak climatic control of long-term chemical weathering rates." Geology 29(6): 511-514.
Riebe, C. S., J. W. Kirchner, D.E. Granger, and R.C. Finkel (2000). "Erosional equilibrium and disequilibrium in the Sierra Nevada, inferred from cosmogenic 26Al and 10Be in alluvial sediment." Geology 28(9): 803-806.
Rieger, W. (1998). "A phenomenon-based approach to upslope contributing area and depressions in DEMs." Hydrological Processes 12(6): 85-72.
Rieger, W. A., et al. (1982). The behaviour of sediment concentrations and solute concentrations in small forested catchments. National Conference Publication - Institution of Engineers, Australia. E. M. O'Loughlin and L. J. Bren. 82: 79-83.
Rieman, B. and J. Clayton (1999). "Wildfire and Native Fish: Issues of Forest Health and Conservation of Sensitive Species." Fisheries 22(11): 6-15.
Rieman, B., et al. (2003). "Status of native fishes in the western United States and issues for fire and fuels management." Forest Ecology and Management 178: 197-211.
Rieman, B. E. and J. B. Dunham (2000). "Metapopulations and salmonids: a synthesis of life history patterns and empirical observations." Ecology of Freshwater Fish 9: 51-64.
Abstract – Metapopulation theory has attracted considerable interest with reference to the salmonids. There has been little empirical evidence,
however, to guide the evaluation or application of metapopulation con-cepts.
From knowledge of salmonid life histories and our own work with
bull trout (Salvelinus confluentus), Lahontan cutthroat trout (Onco-rhynchus
clarki henshawi) and westslope cutthroat trout (Oncorhynchus
clarki lewisi), we suggest that simple generalizations of salmonid metapop-ulations
are inappropriate. Although spatial structuring and dispersal
mechanisms are evident, the relevance of extinction and colonization pro-cesses
are likely to vary with life history, species, scale, and landscape.
Understanding dispersal, the role of suitable but unoccupied habitats, and the potential for extinction debts in non-equilibrium metapopulations are
key issues. With regard to conservation of salmonids, we suggest that efforts to understand and conserve key processes likely to influence the persistence of populations or metapopulations will be more successful
than efforts to design minimal habitat reserves based on metapopulation
theory.
 
Rieman, B. E., et al. (2006). "Emerging concepts for management of river ecosystems and challenges to applied integration of physical and biological sciences in the Pacific Northwest, USA." Intl. J. River Basin Management 4(2): 85-97.
Integration of biological and physical concepts is necessary to understand and conserve the ecological integrity of river systems. Past attempts at
integration have often focused at relatively small scales and on mechanistic models that may not capture the complexity of natural systems leaving
substantial uncertainty about ecological responses to management actions. Two solutions have been proposed to guide management in the face of that
uncertainty: the use of “natural variability” in key environmental patterns, processes, or disturbance as a reference; and the retention of some areas
as essentially unmanaged reserves to conserve and represent as much biological diversity as possible. Both concepts are scale dependent because
dominant processes or patterns that might be referenced will change with scale. Context and linkages across scales may be as important in structuring
biological systems as conditions within habitats used by individual organisms. Both ideas view the physical environment as a template for expression,
maintenance, and evolution of ecological diversity. To conserve or restore a diverse physical template it will be important to recognize the ecologically
important differences in physical characteristics and processes among streams or watersheds that we might attempt to mimic in management or
represent in conservation or restoration reserves.
 
 
Rieman, B. E., et al. (1997). Does wildfire threaten extinction for salmonids? Responses of Redband Trout and Bull Trout following recent large fires on the Boise National Forest. Fire Effects on Rare and Endangered Species and Habitats. Coeur d' Alene, Idaho, IAWF: 47-57.
Rieman, B. E. and J. D. McIntyre (1995). "Occurrence of Bull Trout in naturally fragmented habitat patches of varied size." Transactions of the American Fisheries Society 124: 285-296.
Bull troul Salvelinus confluentus and other salmonids in the Pacific Northwest are
believed at risk of local and regional extinctions because of ongoing habitat loss and fragmentation.
Biologists have focused on defining and protecting critical stream channel characteristics, but there
is little information regarding the scale or spatial geometry of habitat that may be necessary for
the species' long-term persistence. We investigated the influence of habitat patch size on the
occurrence of bull trout by determining the presence or absence of fish in naturally fragmented
watersheds of the Boise River basin in Idaho. We defined patches of potential habitat for bull
trout as watersheds above 1,600 m elevation, a criterion based on the presumed restriction of local
populations by stream temperature. We used logistic regression to investigate the possible influence
of patch size as well as stream width and gradient on the occurrence of bull trout at reach, stream,
and patch scales of analysis. Both stream width and patch size were significant in the models, but
individual effects could not be clearly resolved because of collinearity. The predicted probability
of occurrence based on patch size alone was less than 0.10 for patches smaller than about 1,000
ha and more than 0.50 for patches larger than about 2,500 ha. Our results support the hypothesis
that area of available habitat influences the distribution of disjunct populations of bull trout. An
area effect is consistent with the predictions of island biogeography and metapopulation theory,
and our work suggests that larger-scale spatial processes may be important to the persistence of
species like bull trout.
 
 
Rieman, B. E. and J. D. McIntyre (1996). "Spatial and temporal variability in Bull Trout redd counts." North American Journal of Fisheries Management 16: 132-141.
Rieman, B. E., et al. (2006). "Have brook trout (Salvelinus fontinalis) dispolayed bull trout (Salvelinus confluentus) along longitudinal gradients in coastal Idaho streams?" Canadian Journal Fisheries and Aquatic Sciences 63: 63-78.
Riitters, L. H., et al. (1995). "A factor analysis of landscape pattern and structure metrics." Landscape Ecology 10(1): 23-39.
Riley, K. (2008). Innovations in Assessment of Wildfire Effects: Post-fire Debris Flows and Wildfire Severity from Sub-pixel to Landscape Scale, Doctoral Dissertaion Proposal. Department of Geosciences, University of Montana.
Riley, S. J., et al. (1999). "A terrain ruggedness index that quantifies topographic heterogeneity." Intermountain Journal of Sciences 5(1-4): 23-27.
Rinaldo, A. and i. Rodriguez-Iturbe (1996). "Geomorphological theory of the hydrological response." Hydrological Processes 10: 803-829.
Rinne, J. N. (1980). "Spawning Habitat and Behavior of Gila Trout, a Rare Salmonid of the Southwestern United States " Transactions of the American Fisheries Society 109: 83-91.
Rinne, J. N. (1996). "Short-Term Effects of Wildfire on Fishes and Aquatic Macroinvertebrates in the Southwestern United States."? 16: 653-658.
Rinne, J. N. (1996). "Short-Term Effects on Wildlife on Fishes and Aquatic Macroinvertebrates in the Southwestern United States " North American Journal of Fisheries Management 16: 653-658.
Ritchie, B. D., et al. (1999). "Three-dimensional numerical modeling of coarse-grained elastic deposition in sedimentary basins." Journal of Geophysical Research-Solid Earth 104(B8): 17759-17780.
Ritchie, J. C., et al. (1995). "Laser altimeter measurements at Walnut Gulch Watershed, Arizona." Journal of Soil and Water Conservation 50(5): 440-442.
Rivenbark, B. L., and C.R. Jackson (2004). "Concentrated flow breakthroughs moving through silvicultural streamside management zones: Southeastern Piedmont, USA." Journal of the American Water Resources Association 40(4): 1043-1052.
Rivenbark, B. L. and C. R. Jackson (2004). "Average discharge, perennial flow initiation, and channel initiation - small southern Appalachian basins." Journal of the American Water Resources Association 40(3): 639-646.
Robert, A. and K. S. Richards (1988). "On the modeling of sand bedforms using the semivariogram." Earth Surface Process and Landforms 13: 459-473.
Roberts, R. G. and M. Church (1986). "The sediment budget in severely disturbed watersheds, Queen Charlotte Ranges, British Columbia." Canadian Journal of Forest Research 16: 1092-1106.
Robertson, D. M. and D. A. Saad Environmental water-quality zones for streams: a regional classification scheme. Environmental Management [New York] 31(5): pp. 581-602;   2003, Springer-Verlag New York: 175 Fifth Avenue, New York, NY 10010, USA.
Various approaches have been used to classify large geographical areas into smaller regions of similar water quality or extrapolate water-quality data from a few streams to other unmonitored streams. A combination of some of the strengths of existing techniques is used to develop a new approach for these purposes. In this new approach, referred to here as SPARTA (SPAtial Regression-Tree Analysis), environmental characteristics for each monitored stream are first quantified using a Geographic Information System (GIS) and then regression-tree analysis is used to determine which characteristics are most statistically important in describing the distribution of a specific water-quality constituent. GIS coverages of only the most statistically significant environmental characteristics are then used to subdivide the area of interest into relatively homogeneous environmental water-quality zones. Results from the regression-tree analysis not only define the most important environmental characteristics, but also describe how to subdivide the coverage of the specific characteristic (e.g. areas with <26% soil clay content). The resulting regionalization scheme is customized for each water-quality constituent based on the environmental characteristics most statistically related to that constituent.   SPARTA was used to delineate areas of similar phosphorus, nitrogen, and sediment concentrations and areas of similar potential water quality. The SPARTA approach reduced the variability in water-quality concentrations (phosphorus, total nitrogen, Kjeldahl nitrogen, and suspended sediment) within similarly classified zones from that obtained using the US Environmental Protection Agency's nutrient ecoregions.
 
Robichaud, P. R., R.T. Graham, and R.D. Hungerford (1993). Onsite sediment production and nutrient losses from a low severity burn in the interior northwest. Interior cedar-hemlock-white pine forests; ecology and management., Spokane, WA, Pullman, WA, Dept. of Natural Resources.
Robichaud, P. R. (1997). Geostatistics: A new tool for describing spatially-varied surface conditions from timber harvested and burned hillslopes. ASAE International Meeting 1997, Minneapolis, Minnesota, ASAE, 2950 Niles Road, St. Joseph, MI 49085-9659.
Robichaud, P. R., and S.M. Miller (1999). "Spatial interpolation and simulation of post-burn duff thickness after prescribed fire." International Journal of Wildland Fire 9(2): 137-143.
Robichaud, P. R., and R.E. Brown (1999). "What happens after the smoke cleared: onsite erosion rates after a wildfire in eastern Oregon." Wildland hydrology(june/july): 419-426.
Recent fires have renewed interest in fire's effect on different components of the ecosystem, in particular fire's effect on erosion. Erosion is a natural process occurring at varying rates and scales depending on soil type, topography, vegetation, climate and type of disturbance. Erosion rates in a forest environment are generally small except when the ground surface is disturbed by human or natural causes. First year surface erosion rates after wildfire can vary from 0.1 to greater than 50 Mg/ha/yr, often decreasing by orders of magnitude in subsequent years. Wildfires may consume the forest floor, leaving the soil surface exposed to raindrop impact and overland flow. However, wildfires do not burn uniformly across the hillslope or landscape. Thus fire severity and erosion potential can vary spatially. Additionally, wildfires may leave the soil surface in a water repellent condition, reducing infiltration and increasing overland flow. These highly erosive conditions are temporary, because natural and artificial revegetation occur within several years after the fire. The two dominant hillslope erosion processes are interrill and rill erosion. Raindrop impact and shallow overland flow can detach soil particles, making them available for transport. Concentrated flow (rill erosion) can detach soil particles if the hydraulic energy is greater than the erosion resistance of the soil. In steep forest environments, ample hydraulic energy is available to move sediment downslope; thus erosion is generally limited by the amount of detached sediment available for transport. Over the past 10 years, rainfall simulation has been used to quantify interrill erosion and infiltration rates. Simulated concentrated flow techniques have been used to quantify the rill erosion rates. Silt fences quantified natural hillslope erosion and catchment-scale watersheds have been monitored to quantify runoff and erosion leaving the hillslope.
 
Robichaud, P. R., J.L. Beyers, and D.G. Neary (2000). Evaluating the effectiveness of postfire rehabilitation treatments. USDA Forest Service General technical report  RMRS-GTR-63, 59p.
Robichaud, P. R. (2000). "Fire effects on infiltration rates after prescribed fire in Northern Rocky Mountain forests, USA." Journal of Hydrology 231: 220-229.
Infiltration rates in undisturbed forest environments are generally high. These high infiltration rates may be reduced when forest management activities such as timber harvesting and/or prescribed fires are used. Post-harvest residue burning is a common site preparation treatment used in the Northern Rocky Mountains, USA, to reduce forest fuels and to prepare sites for natural and artificial tree regeneration. Prescribed burn operations attempt to leave sites with the surface condition of a low-severity burn. However, some of the areas often experience surface conditions associated with a high-severity burn which may result in hydrophobic or water repellent conditions. In this study, infiltration rates were measured after logging slash was broadcast burned from two prescribed burns. The two sites were in Northern Rocky coniferous forests of Douglas-fir/lodepole pine and 93 mm h(-1) within the three surface conditions found after the burn: unburned-undisturbed areas, low-severity burn areas and high-severity burn areas.Runoff hydrographs from the rainfall simulations were relatively constant from the plots that were in unburned-undisturbed areas and in areas subjected to a low severity burn. These constant runoff rates indicate constant hydraulic conductivity values for these surface conditions even though there was variation between plots. Hydrographs from the rainfall simulation plots located within areas of high-severity burn indicate greater runoff rates than the plots in low-severity burn areas especially during the initial stages of the first rainfall event. These runoff rates decreased to a constant rate for the last 10 min of the event. These results indicate hydrophobic or water I repellent soil conditions, which temporarily cause a 10-40% reduction in hydraulic conductivity values when compared to a normal infiltrating soil condition. Since variability was high for these forest conditions, cumulative distribution algorithms of hydraulic conductivity provide a means to account fur the inherent variability associated with these hillslopes and different surface conditions cause by fire. Published by Elsevier Science B.V.
 
Robichaud, P. R. (2000). "Forest Fire Effects on Hillslope Erosion: What We Know." Watershed Management Council Networker: 1,9-13.
Robichaud, P. R., L.H. MacDonald, and R. B. Foltz (2006). Fuel management and erosion. CWE for fuels management in the Western U.S. W. J. Elliot, and L.J. Andrei.
Robichaud, P. R. and R. E. Brown (1999). What happened after the smoke cleared: Onsite erosion rates after a wildfire in Eastern Oregon. Wildland hydrology, American Water Resources Association.
Robichaud, P. R., et al. (2007). Erosion Risk Management Tool (ERMiT) User Manual (version 2006.01.18), USDA Rock Mountain Research Station: 25.
Robichaud, P. R., et al. (1993). "Development of an on site sediment prediction model for forest roads and timber harvest areas." IAHS-AISH Publication 217: 135-140.
Robichaud, P. R. and S. M. Miller (2000). "Spatial interpolation and simulation of post-burn duff thickness after prescribed fire." International Journal of Wildland Fire 9(2): 137-143.
Prescribed fire is used as a site treatment after timber harvesting. These fires result in spatial patterns with some portions consuming all of the forest floor material (duff) and others consuming little. Prior to the burn, spatial sampling of duff thickness and duff water content can be used to generate geostatistical spatial simulations of these characteristics. Results from field studies indicated that spatial patterns of duff characteristics occurred, and they were then modeled by kriging, simulation and a trend-surface modeling techniques. The higher elevations of the study unit burned more severely than the lower portion. This is believed to be due to the heat generated by the fire drying out the upper portions of the units, thus consuming more duff material and thinner preburn duff thickness due to ground-based harvesting techniques. Attempts to predict duff consumption and subsequent post-burn duff thickness were successful using a trend-surface model developed for this site and a general duff consumption model. Knowledge of spatial patterns of duff remaining may help land managers adjust prescriptions and alter ignition patterns to reduce areas where total consumption of duff might occur.
 
Robichaud, P. R. and T. A. Waldrop (1994). "A Comparison of Surface Runoff and Sediment Yields from Low-Severity and High-Severity Site Preparation Burns." Water Resources Bulletin 30(1): 27-34.
Slash burning is a common site preparation technique used after timber harvest throughout the Southeastern United States. Little quantitative information exists on the hydrologic response to burn severity. This study compared the effects of low-severity and high-severity burns on runoff and sediment yields during rainfall simulation and during natural rainfall in the Southern Appalachian Mountains. Fire severity was largely determined by moisture conditions of the forest floor prior to ignition. Runoff and sediment yield variability was high between plots within the same treatment area due to differences in forest floor characteristics and infiltration rates. Conditions of high-severity resulted when burning was conducted with relatively dry fuels. Sediment yields were 40-times greater for the high-severity treatment areas than the low-severity treatment areas.
 
Robinson, J. S. and M. Sivapalan (1997). "Temporal scales and hydrological regimes: Implications for flood frequency scaling." Water Resources Research 33(12): 2981-29999.
Robinson, T. W. (1958). Phreatophytes. G. Survey, U.S. Geological Survey. Water-Supply Paper 1423.
Robison, E. G. (?). "Small Channel Impacts related to low frequency storm effects in forested areas of Western Oregon."
Robison, E. G. and R. L. Beschta (1989). Coarse woody debris and channel morphology of low gradient streams in southeast Alaska, U. S. A. Headwater Streams. Herndon, VA, Amercian Water Resources Association.
Robison, E. G. and R. L. Beschta (1990). "Characteristics of coarse woody debris for several coastal streams of southeast Alaska, USA." Canadian Journal of Aquatic Sciences 47: 1684-1693.
Robison, E. G. and R. L. Beschta (1990). "Coarse woody debris and channel morphology interactions for undisturbed streams in southeast Alaska, USA." Earth Surface Processes and Landforms 15: 149-156.
Robison, E. G. and R. L. Beschta (1990). "Identifying trees in riparian areas that can provide coarse woody debris to streams." Forest Science 36(3): 790-801.
Robison, E. G., et al. (1996). Oregon Department of Forestry: Storm Impacts and Landslides of 1996: Final Report, Oregon Department of Forestry: 145.
Robison, E. G., et al. (1999). "Oregon Department of Forestry Storm Impacts and Landslides of 1996: Draft Final Report."
Robison, G. E. and R. L. Beschta (1990). "Identifying trees in riparian areas that can provide coarse woody debris to streams." Forest Science 36(3): 790-801.
Robison, G. E., et al. (1999). Storm Impacts and Landslides of 1996: Final Report, Oregon Department of Forestry.
Rockwell, D. L. (2002). "The influence of groundwater on surface flow erosion processes during a rainstorm." Earth Surface Processes and Landforms 27(5): 495-514.
Surface erosion rates on a disturbed natural soil in a 10 in indoor flume increased by an order of magnitude when a water table developed at a 10 cm depth during simulated rainstorms. Erosion rate increases did not correlate well With surface hydraulic flow conditions, and all significant erosion increases began before the full soil depth was saturated, before the water table reached the soil surface, and before seepage was possible. Groundwater influenced erosion processes primarily by increasing unsaturated pore-water pressures and decreasing soil shear strength in surface rainflow, rather than through the direct entrainment of soil particles by seepage flow.There was no unique morphologic expression of the influence of groundwater during a rainstorm. Subsurface processes influencing Surface erosion were detected only by appropriate subsurface instrumentation, which included micropiezometers, tensiometers and time domain reflectometry. Erosion rate increases occur-red all along the slope, and were not concentrated at the base of slope due to a seepage zone. Soil depth was crucial to determining surface erosion increase. It is likely that confusing trends in surface flow erosion rates in past Studies have occurred due to unrecorded groundwater development or an emphasis on seepage effects. Groundwater must be monitored along hillslopes under all moisture and soil conditions in order to avoid misleading and inconsistent conclusions derived solely from surface How or seepage data. Copyright (C) 2002 John Wiley Sons, Ltd.
 
Rodhe, A., et al. (1996). "Transit times for water in a small till catchment from a step shift in the oxygen 18 content of the water input." Water Resources Research 32(12): 3497-3511.
Rodine, J. D. and A. M. Johnson (1976). "The ability of debris, heavily freighted with coarse elastic materials, to flow on gentle slopes." Sedimentology 23: 213-234.
Rodriguez-Iturbe, I., et al. (1979). "Discharge response analysis and hydrologic similarity: The interrelation between the geomorphologic IUH and the strom characteristics." Water Resources Research 15(6): 1435-1444.
Rodriguez-Iturbe, I., et al. (1972). "Streamflow simulation, 1, A new look at Markovian models, fractional Gaussian noises and crossing theory; 2) The broken line process as a potential model for hydrologic simulation." Water Resources Research 8(4): 921-941.
Rodriguez-Iturbe, I. and A. Rinaldo (1997). Fractal River Basins. Cambridge, Cambridge University Press.
Rodriguez-Iturbe, I. and J. B. Valdes (1979). "The geomorphologic structure of hydrologic response." Water Resources Research 15(6): 1409-1420.
Roering, J. J. (2005). "Fire-driven landscape disequilibrium at the hillslope and range scale." Eos Trans. AGU,  Fall Meet. Suppl., Abstract #H34A-03. San Francisco, California.
Roering, J. J., et al. (2002). "Soil transport driven by biological processes over millennial time scales." Geology 30(12): 1115-1118.
Roering, J. J. and M. Gerber (2005). "Fire and evolution of steep, soil-mantled landscapes." Geology 33(5): 349-552.
Recent burns in the western United States attest to the significant geomorphic impact of fire in mountainous landscapes, yet we lack the ability to predict and interpret fire-related erosion over millennial time scales. A diverse set of geomorphic processes is often invoked following fire; the magnitude of postfire erosional processes coupled with temporal variations in fire frequency dictate the extent to which fires affect sediment production and landscape evolution. In the Oregon Coast Range, several models for long-term rates of soil production and transport have been tested and calibrated, although treatment of fire-related processes has been limited. Following recent fires in the Oregon Coast Range, we observed extensive colluvial transport via dry ravel, localized bedrock emergence due to excess transport, and talus-like accumulation in adjacent low-order valleys. Soils exhibited extreme but discontinuous hydrophobicity, and no evidence for rilling or gullying was observed. Using a field-based data set for fire-induced dry ravel transport, we calibrated a physically based transport model that indicates that soil flux varies nonlinearly with gradient. The postfire critical gradient (1.03), which governs the slope at which flux increases rapidly, is lower than the previously estimated long-term value (1.27), reflecting the reduction of slope roughness from incineration of vegetation. By using a high-resolution topographic data set generated via airborne-laser swath mapping, we modeled the spatial pattern of postfire and long-term erosion rates. Postfire erosion rates exceed long-term rates (which average 0.1 mm·yr−1) by a factor of six, and subtle topographic variations generated local patches of rapid postfire erosion, commonly >1 mm·yr−1. Our simulations indicate that fire-related processes may account for 50% of temporally averaged sediment production on steep hillslopes. Our analysis provides a mechanistic explanation for the coincident early Holocene timing of increased fire frequency and regional aggradation in Oregon Coast Range drainage basins. Given the sensitivity of steep hillslopes to fire-driven transport, changes in climate and fire frequency may affect soil resources by perturbing the balance between soil transport and production.
 
Roering, J. J. and M. Gerber (2005). "Fire and the evolution of steep, soil-mantled hillslopes." Geology 33(No. 5): 349-352.
Roering, J. J., et al. "Evidence for nonlinear, diffusive sediment transport on hillslopes and implications for landscape morphology."
The mechanics of diffusive processes and their influence on hillslope morphology were studied. Hillslopes with profile and planform convexity occur in diverse climatic and tectonic regimes, suggesting that diffusive sediment transport processes control hillslope morphology in many different settings. The interaction between local disturbances, and frictional and gravitational forces resulted in a diffusive transport law that depends nonlinearly on hillslope gradient. This transport law was tested and calibrated using high-resolution topographic data from the Oregon Coast Range. At five small basins of the study area, hillslope curvature approaches zero with increasing gradient, consistent with the proposed transport law.
 
Roering, J. J., et al. (1999). "Evidence for nonlinear, diffusive sediment transport on hillslopes and implications for landscape morphology." Water Resources Research 35(3): 853-870.
Roering, J. J., et al. (2005). "Characterizing structural and lithologic controls on deep-seated landsliding: Implications for topographic relief and landscape evolution in the Oregon Coast Range, USA." Geological Society of America Bulletin 117: 654-668.
Roering, J. J., et al. (2003). "Shallow landsliding, root reinforcement, and the spatial distribution of trees in the Oregon Coast Range." Canadian Geotechnical Journal 40: 237-253.
Rogers, R. (1989). Influence of sparse vegetation cover on ersion and rill patterns: an experimental study. Department of Earth Resources. Fort Collins, Colorado State University: 65.
Roghair, C. N., et al. (2002). "Response of a Brook trout population and instream habitat to a catastrophic flood and debris flow." Transactions of the American Fisheries Society 131: 718-730.
Roghair, C. N., et al. (2002). "Response of a Brook Trouth Population and Instream Habitat to a Catastrophic Flood and Debris Flow." Transactions of the American Fisheries Society 131: 718-730.
Rollerson, T. P., et al. (1986). An approach to predicting post-logging slope stability for coastal British Columbia. NCASI West Coast Regional Meeting, Portland, Oregon.
Rollerson, T. P., et al. (2001). Predicting post-logging landslide activity using terrain attributes: Coast Mountains, British Columbia. Nanaimo, BC, Canada, British Columbia Ministry of Forests.
Rollerson, T. P., et al. (2002). Using terrain attributes to predict post-logging landslide likelihood on southwestern Vancouver Island. Nanaimo, BC, Canada, B.C. Ministry of Foresty: 15.
Romkens, M. J. M., et al. (2002). "Soil erosion under different rainfall intensities, surface roughness, and soil water regimes." Catena 46(2-3): 103-123.
Soil erosion is a complex phenomenon involving the detachment and transport of soil particles, storage and runoff of rainwater, and infiltration. The relative magnitude and importance of these processes depends on a host of factors, including climate, soil, topography, cropping and land management practices, control practices, the antecedent conditions, and the size of the area under consideration. In this study, the results of a series of experiments are reported, summarizing the soil loss and runoff response from a 0.6 x 3.75 in area to different rainstorm regimes, slope steepnesses, subsurface soil water pressures, and surface roughnesses under controlled laboratory conditions using a flume and rainfall simulator as water applicators, and a laser microreliefmeter and tensiometric system as soil response measuring devices. The soil chosen was a highly erodible Grenada loess (fine silty, mixed, thermic, Glossic Fragiudalf). The results showed: (1) a sequence of rainstorms of decreasing intensity on an initially air-dry smooth surface caused more soil loss than a sequence of similar storms of increasing intensity; (2) the surface roughness-sediment concentration relationship was not monotonic in nature; (3) subsurface soil water pressure substantially affected sediment concentration in runoff but only marginally impacted runoff amounts; (4) initially smooth, uniform surfaces may yield less soil loss than initially rough surfaces; (5) interrill runoff occurred as spatially varying flow in which flow patterns determine the locations of rills. Published by Elsevier Science B.V.
 
Romme, W. H. (1982). "Fire and Landscape Diversity in Subalpine Forests of Yellowstone National Park
doi:10.2307/1942611." Ecological Monographs 52(2): 199-221.
Romme, W. H., J. Clement, J. Hicke, D. Kulakowski, L.H. MacDonald, T.L. Schoennagel, and T.T. Veblen, . (2006). Recent forest insect outbreaks and fire risk in Colorado forests: a brief synthesis of relevant research. Colorado Forest Restoration Institute. CSU, Fort Collins, CO: 24 pp.
Roni, P., Beechie, T. J., Bilby, R. E., Leonetti, F. E., Pollock, M. M., and Pess, G. R. (2002). "A review of stream restoration techniques and a hierarchical strategy for prioritizing restoration in Pacific northwest watersheds." North American Journal of Fisheries Management 22(1): 1-20.
Roni, P., et al. (2002). "A review of stream restoration techniques and a hierarchical strategy for prioritizing restoration in Pacific Northwest watersheds." North American Journal of Fisheries Management 22: 1-20.
Roni, P. and T. Quinn (2001). "Effects of wood placement on movements of trout and juvenile coho salmon in natural and artificial stream channels." Transactions of the American Fisheries Society 130: 675-685.
We monitored the movements of marked juvenile coho salmon Oncorhynchus kisutch,
steelhead O. mykiss, and cutthroat trout O. clarki in a stream reach that had been ‘‘restored’’ with
placed wood and a reference reach with no wood placement and tracked the growth and movements
of individually marked coho salmon among habitats in artificial channels with and without woody
debris. Monthly surveys in Shuwah Creek, Washington, indicated that few (0–33%) of the marked
trout or coho salmon moved between the restored and reference reaches. However, a rapid decline
in both marked and unmarked fish in late fall and the increasing proportion of unmarked fish
indicated considerable migration to and from the study reaches. In the artificial channels, fewer
fish moved in the simple (with no wood) channel than in the complex (with wood) channel (22%
versus 37%), and the mean distance moved was shorter in the complex channel (4.4 versus 6.7
habitat units). In the simple channel, the fish that moved grew faster than those that did not.
Movement may facilitate increased growth in stream reaches with little woody debris, and the
placement of woody debris may lead to less frequent and shorter movements.
 
 
Rood, K. M. (1984). An Aerial Photograph Inventory of the Frequency and Yield of Mass Wasting on the Queen Charlotte Islands, British Columbia. Vancouver, British Columbia.
Rooney, J. J. (1995). Determination of sedimentation rates in Tomales Bay, California
    using a geographic information system. Honolulu, University of Hawaii at Manoa.
A geographic information system (GIS) was used to calculate sedimentation rates for Tomales Bay, California. Surfaces were constructed using digitized bathymetric data from National Ocean Service navigational survey charts published in 1861, 1931, 1957 and 1994 with a public-domain Unix-based software package, Generic Mapping Tool (GMT), that has GIS capabilities. The surfaces were subtracted from each other to determine the volume of sediment that accumulated in the Bay between navigational surveys. The assumption is made that sediment deposited in Tomales Bay remains there. Thus, sedimentation in the basin approximates sediment delivery to Tomales Bay. Bay-wide average sedimentation rates were found to be 1.8 kg m (super -2) yr (super -1) for the 1861-1931 interval, 5.8 kg m (super -2) yr (super -1) for the 1931-1957 interval, 1.0 kg m (super -2) yr (super -1) for 1957-1994, and 2.8 kg m (super -2) yr (super -1) for the entire period from 1861-1994. These rates were compared with results hindcast from existing models relating rainfall in the watershed to runoff and sediment yield. GIS results for the 1861-1931 and 1957-1994 intervals agree rather well with the sediment yield model, while that model apparently underestimates sedimentation for the 1931-1957 interval. Differences between the two sets of rates apparently reflect changes in land-usage that effected both erosion in the watershed and, after a lag time of several decades, sediment accumulation in the bay. Discrepancies between the GIS results and radiometrically based estimates of sedimentation may reflect limited coverage for the radiometrically dated cores, together with overestimated recent sedimentation because of sediment mixing from large scale bioturbation.
 
Root, T. L. and S. H. Schneider (1995). "Ecology and climate: research strategies and implications." Science 269: 334-341.
Roper, B. and D. L. Scarnecchia (1995). "Observer Variability in Classifying Habitat Types in Stream Surveys " North American Journal of Fisheries Management 15: 49-53.
Roper, B. B., et al. (2002). "An evaluation of physical stream habitat attributes used to monitor streams." Journal of the American Water Resources Association 38(6): 1637-1646.
The last few decades have seen an increased reliance
on the use of stream attributes to monitor stream conditions. The
use of stream attributes has been criticized because of variation in
how observers evaluate them, inconsistent protocol application,
lack of consistent training, and the difficulty in using them to
detect change caused by management activity. In this paper, we
evaluate the effect of environmental heterogeneity and observer
variation on the use of physical stream attributes as monitoring
tools. For most stream habitat attributes evaluated, difference
among streams accounted for greater than 80 percent of the total
survey variation. To minimize the effect that variation among
streams has on evaluating stream conditions, it may be necessary
to design survey protocols and analysis that include stratification,
permanent sites, and/or analysis of covariance. Although total variation
was primarily due to differences among streams, observers
also differed in their evaluation of stream attributes. This study
suggests that if trained observers conducting a study that is
designed to account for environmental heterogeneity can objectively
evaluate defined stream attributes, results should prove valuable
in monitoring differences in reach scale stream conditions. The failure
to address any of these factors will likely lead to the failure of
stream attributes as effective monitoring tools.
 
Rosenfeld, J. S. and L. Huato (2003). "Relationship between large woody debris characteristics and pool formation in small coastal British Columbia streams." North American Journal of Fisheries Management 23: 928-938.
The characteristics and function of large woody debris (LWD) were measured in 41
small (1.2–11.2-m bank-full channel width), fish-bearing streams in coastal British Columbia to
determine how total LWD abundance and the features of individual LWD pieces (diameter, length,
orientation, and presence of a rootwad) influenced the effectiveness of pool formation. Pool spacing
(the number of channel widths between channel-spanning pools) was a decreasing power function
of total LWD abundance, but the relationship was relatively weak. Stratification of sites by channel
gradient improved the model fit, steeper streams ($2% gradient) having a significantly lower pool
spacing than lower-gradient streams (,2%). The proportion of LWD that formed pools increased
from 6% for pieces with a diameter of 15–30 cm to 43% for pieces with a diameter of more than
60 cm. Large woody debris more than 60 cm in diameter formed a higher proportion of pools
across all channel widths. A simple, size-structured model of LWD abundance in small streams
suggests that loss of LWD larger than 60 cm in diameter will greatly decrease pool frequency
across all channel widths but have the greatest impact on large streams. Models that estimate pool
frequency based on total LWD abundance irrespective of size distribution may underestimate the
impact of riparian management that reduces the number of larger-diameter trees recruiting to the
stream channel.
 
Rosenfeld, J. S., et al. (2005). "Food abundance and fish density alters habitat selection, growth, and habitat suitability curves for juvenile coho salmon (Oncorhynchus kisutch)." Canadian Journal of Fisheries and Aquatic Science 62: 1691-1701.
To understand how fish density and food availability affect habitat selection and growth of juvenile coho
salmon (Oncorhynchus kisutch), we manipulated fish density (2–12 fish·m –2 ) and natural invertebrate drift (0.047–
0.99 mg·m –3 ) in 12 experimental stream channels constructed in a side-channel of Chapman Creek, British Columbia.
Increased food resulted in increased growth of both dominant and subdominant fish and a shift to higher average focal
velocities (from 6.5 to 8.4 cm·s –1 ) with maximum growth in the range of 10–12 cm·s –1 . Increased food appears to per-mit
juvenile coho to exploit higher velocity microhabitats that might otherwise be bioenergetically unsuitable at lower
food levels. Increased fish density resulted in lower growth of subdominant but not of dominant fish and a general dis-placement
of fish to both higher and lower focal velocities. The shapes of habitat suitability curves were sensitive to
food abundance, implying that differences in food availability may affect transferability of habitat suitability curves be-tween
streams of different productivity. While habitat suitability curves captured the change in extent of available habi-tat
following prey enrichment, actual increases in growth rate with enrichment (i.e., changes in habitat quality) were
poorly represented by habitat suitability values and better represented by bioenergetic model predictions.
 
Rosenfeld, J. S., et al. (2007). "Hydraulic geometry as a physical template for the River Continuum: application to optimal flows and longitudinal trends in salmonid habitat." Canadian Journal of Fisheries and Aquatic Science 64: 755-767.
Rosgen, D. (1996). Applied River Morphology. Pagosa Springs, Colorado, Wildland Hydrology.
Rosgen, D. and J. Kurz (?). Bankfull Discharge Determination Pacific Lumber Company and National Marine Fisheries Service. Pagosa Springs, Colorado.
Rosgen, D. L. (1985). A Stream Classification System. Riparian Ecosystems and their Management Reconciling Conflicting Uses. R. R. Johnson, C. D. Ziebell, D. R. Patton and P. F. Folliott, USDA Forest Service.
Rosgen, D. L. (1994). "A classification of natural rivers." Catena 22: 169-199.
A classification system for natural rivers is presented in which a morphological arrangement of stream characteristics is organized into relatively homogeneous stream types. This paper describes morphologically similar stream reaches that are divided into 7 major stream type categories that differ in intrenchment, gradient, width/depth ratio, and sinuosity in various landforms. Within each major category are six additional types delineated by dominate channel materials from bedrock to silt/clay along a continuum of gradient ranges. Recent stream type data used to further define classification interrelationships were derived from 450 rivers throughout the U.W., Canada, and New Zealand. Data used in the development of this calssification involved a great diversity of hydro-physiographic/geomorphic provinces from small to large rivers and in catchments from headwater streams in the mountains to the coastal plains. A stream hierarchical inventory system is presented which utilizes the stream classification system. Examples for use of this stream classification system for engineering, fish habitat enhancement, restoration and water resource management applications are presented. Specific examples of these applications include hydraulic geometry relations, sediment supply/availability, fish habitat structure evaluation, flow resistance, critical shear stress estimates, shear stress/velocity relations, streambank erodibility potential, management interpretations, sequences of morphological evolution, and river restoration principles.
 
Rosgen, D. L. (2004). Watershed assessment for river stability and sediment supply. WARSS. Fort Collins, CO.
Rot, B. W., et al. (2000). "Stream channel configuration, landform, and riparian forest structure in the Cascade Mountains, Washington." Canadian Journal of Fisheries and Aquatic Science 57: 699-707.
The hierarchical relationship of five key elements, valley constraint, riparian landform, riparian plant community,
channel type, and channel configuration, are described for 21 sites in mature to old-growth riparian forests of the
western Cascades Mountains, Washington, U.S.A. Channel type (bedrock, plane-bed, and forced pool–riffle) was
closely related to channel configuration (especially large woody debris (LWD) volume, density, and LWD-formed
pools) at the smallest spatial scale and valley constraint at the largest. Valley constraint significantly influenced
off-channel habitat (r 2 = 0.71) and LWD volume within forced pool–riffle channels (r 2 = 0.58). Riparian plant community
composition was differentiated by four landform classes: three alluvial landforms based on height above the channel
and one based on hillslope. Just above the active channel, floodplain landforms contained more deciduous stems
than conifer and greater conifer basal area than deciduous. Conifers dominated other landforms. The diameter of
in-channel LWD increased with the age of the riparian forest (r 2 = 0.34). In old-growth forests, LWD diameter was
equivalent to or greater than the average riparian tree diameter for all sites. In younger forests, the mixed relationship
between LWD and riparian tree diameter may reflect a combination of LWD input from the previous old-growth stand
and LWD input from the existing stand.
 
 
Roth, G. and F. Siccardi "Erosional development of drainage patterns. Network sensitivity to hillslope geometry, soil characteristics and rainfall intensity."
Drainage networks development at the hillslope scale is described by modeling two-dimensional flow of water and sediment. Deterministic equations of conservation are coupled with random fluctuations of the sediment transportation processes acting at the hydrodynamic scale. Controlled experiments have been carried out in order to analyze network sensitivity to hillslope geometry, soil characteristics, and rainfall intensity.
 
Roth, G. and F. Siccardi (1989). "Hydrodynamic Description of the Erosional Development of Drainage Patterns." Water Resources Research 25(2): 319-332.
Rothacher, J. (1970). "Increases in water yield following clear-cut logging in the Pacific Northwest." Water Resources Research 6(2): 653-658.
Rothermal, R. C. (1983). How to predict the spread and intensity of forest fires, USDA Forest Service.
Roy, A. G. and R. Roy (1988). "Changes in channel size at river confluences with coarse bed material." Earth Surface Processes and Landforms 13: 77-84.
Roy, A. G. and M. J. Woldenberg (1986). "A model for changes in channel form at a river confluence." Journa of Geology 94: 402-411.
Royall, D. (2001). "Use of mineral magnetic measurements to investigate soil erosion and sediment delivery in a small agricultural catchment in limestone terrain."
Understanding sediment delivery at the hillslope scale requires information on the spatial distributions and magnitudes of erosion and deposition. Empirical models such as the RUSLE may be useful for predicting erosion, but are poorly suited for quantifying deposition. Cesium-137 (Cs-137) is useful for quantifying both erosion and deposition, but is costly to inventory over a large area, and directly gauges soil redistribution only for recent decades. The use of rapidly acquired magnetic measurements represents a relatively new potential means of mapping and measuring soil redistribution. In this study, variations in surface magnetism are analyzed to determine patterns of erosion and sedimentation in a small agricultural catchment in northwestern Alabama (USA). Magnetic indicators of erosion are combined with published soil morphology, Cs-137 and short-term suspended sediment data from this former experimental watershed to evaluate long-term sediment delivery. All magnetic parameters measured could be related to soil erosion, although their patterns in space are not identical. Magnetic evidence suggests approximately 30 cm of soil loss on the steepest mid-slope portions of the catchment. Distributing soil loss across space according to magnetic patterns gives average long-term values lower than a prior estimate based on Cs-137 data. Long-term sediment delivery calculated using soil morphology to determine depositional volume ranges up to 45% depending on the magnetic parameter used to index soil loss, and assumptions regarding deposit geometry. These results suggest the need for continued refinement of magnetic techniques for purposes of erosion model validation and general sediment tracing applications.
 
Russell, W. O. I., et al. (in prep). "Fires, topography and gullies: quantifying spatial variation in long-term sediment yield among 2nd to 3rd order basins in the Blue Mountains of northeastern Oregon."
Rustomji, P. and I. Prosser "Spatial patterns of sediment delivery to valley floors: sensitivity to sediment transport capacity and hillslope hydrology relations."
There is considerable interest in large-scale spatial patterns of sediment transport in catchments, and this topic is often approached using terrain-based modelling. In such models topography influences the discharge of overland flow and its sediment transport capacity. The sediment transport capacity of overland flow is commonly expressed as a power function of slope and discharge (i.e. q sub(s) = k sub(1)q super( beta )S super( gamma )). The relationship between discharge and contributing area can also be expressed as a power function. Several reviews reveal a limited range of values for the two exponents beta and gamma . In this paper we examine the sensitivity of catchment-scale patterns of sediment delivery to valley floors to a range of sediment transport capacity and hillslope hydrology parameterizations, using two catchments on the southern tablelands of New South Wales. The results indicate that, over the limited range of beta and gamma identified within the literature, sediment deliveries to valley floors across the two catchments are similar for all but one of five sediment transport capacity relationships. The patterns are dominated by the trend in slope through each catchment. The sensitivity to hillslope hydrology of predicted sediment delivery patterns is strong in the catchment with systematic variation in unit hillslope area, and weak in the catchment for which there are no systematic trends in unit hillslope area. We believe there is less experimental evidence to restrict choice of hillslope hydrology parameters than there is for sediment transport capacity.
 
Rustomji, P. and I. Prosser (2001). "Spatial patterns of sediment delivery to valley floors: sensitivity to sediment transport capacity and hillslope hydrology relations." Hydrological Processes 15(6): 1003-1018.
Ryan, D. F. (2006). Introduction to synthesis of current science. CWE for fuels management in the Western U.S. W. J. Elliot, and L.J. Andrei.
Ryan, S. E. and Anonymous (2001). "The influence of sediment supply on rates of bedload transport; a case study of three streams on the San Juan National Forest." Proceedings - Federal Interagency Sedimentation Conference 7, Vol. 1: III48-III54.
This paper compares and contrasts differences in the rate and size of bedload moved over a range of flows in 3 streams on the San Juan National Forest that have different modes and supplies of sediment. The East Fork of the San Juan (EFSJ) drains an area with unstable volcanic bedrock and active mass wasting. The bed material is primarily loosely bound sand, gravel, and small cobbles that are readily mobilized, producing high rates of bedload transport; the estimate of the mean rate of bedload transport per basin area at the 1.5-year return interval flow is 0.029 kg s (super -1) km (super -2) . The grain size distribution of bedload observed at high flows approaches that of the bed, suggesting transport at EFSJ approximates an "equal mobility" mode. Silver Creek, a tributary to EFSJ, likewise has an abundant source of sediment from active avalanche chutes and debris slides that deliver material directly to the channel. This material is readily entrained, producing high rates of transport as it moves through the system. Consequently, transport rates were much lower on the recessional limb of the seasonal hydrograph due to diminishing sediment supply; mean rate of transport per basin area on the rising limb was approximately 0.031 kg s (super -1) km (super -2) and 0.002 kg s (super -1) km (super -2) on the falling limb, about an order of magnitude difference. Irregular transport rates in this channel reflect the episodic nature of the sediment supply. By contrast, observed rates of sediment transport at Florida River above Lemon Reservoir were very low, which is indicative of the stability of the landscape and lack of sediment supplied from upstream. The mean rate of transport per unit basin area measured at the 1.5-year return interval flow was 0.0003 kg s (super -1) km (super -2) . Sediment most likely originated from more mobile patches located between larger boulders and cobbles and does not reflect widespread entrainment over the channel surface. The presence of larger particles in the samples at increasingly greater discharge suggests that sediment is more selectively entrained.
 
Ryan, S. E. and G. E. Grant (1991). "Downstream Effects of Timber Harvesting on Channel Morphology in Elk River Basin, Oregon." Journal of Enivironmental Quality 20: 60-72.
Ryan, S. E., et al. (2002). "Defining phases of bedload transport using piecewise regression." Earth Surface Processes and Landforms 27(9): 971-990.
Differences in the transport rate and size of bedload exist for varying levels of flow in coarse-grained channels. For gravel-bed rivers, at least two phases of bedload transport, with notably differing qualities, have been described in the literature. Phase I consists primarily of sand and small gravel moving at relatively low rates over a stable channel surface. Transport rates during Phase II are considerably greater than Phase I and more coarse grains are moved, including material from both the channel surface and subsurface. Transition from Phase I to Phase II indicates initiation and transport of grains comprising the coarse surface layer common in steep mountain channels. While the existence of different phases of transport is generally acknowledged, the threshold between them is often poorly defined. We present the results of the application of a piecewise regression analysis to data on bedload transport collected at 12 gravel-bed channels in Colorado and Wyoming, USA. The piecewise regression recognizes the existence of different linear relationships over different ranges of discharge. The inflection, where the fitted functions intersect, is interpreted as the point of transition from Phase I to Phase II transport; this is termed breakpoint. A comparison of grain sizes moved during the two phases shows that coarse gravel is rarely trapped in the samplers during Phase I transport, indicating negligible movement of grains in this size range. Gravel larger than about D (sub 16) of the channel surface is more consistently trapped during Phase II transport. The persistence of coarse gravel in bedload samples provides good evidence that conditions suitable for coarse grain transport have been reached, even though the size of the sediment approaches the size limits of the sampler (76 mm in all cases). A relative breakpoint (R (sub br) ) was defined by the ratio between the discharge at the breakpoint and the 1.5-year flow (a surrogate for bankfull discharge) expressed as a percentage. The median value of R (sub br) was about 80 percent, suggesting that Phase II begins at about 80 percent of the bankfull discharge, though the observed values of R (sub br) ranged from about 60 to 100 percent. Variation in this value appears to be independent of drainage area, median grain size, sorting of bed materials, and channel gradient, at least for the range of parameters measured in 12 gravel-bed channels.
 
Ryder, J. M. and D. E. Howes (1939). A User's Guide to Terrain Maps in British Columbia, B. C. Ministry of Environment: 1-16.
Sable, K. A. and E. Wohl (2006). "The relationship of lithology and watershed characteristics to fine sediment deposition in streams of the Oregon Coast Range." Environmental Management 37(5): 659-670.
Lithology is one of many factors influencing the amount, grain size distribution, and location of fine sediment deposition on the bed of mountain stream channels. In the Oregon Coast Range, 18 pool-riffle stream reaches with similar slope and intact riparian area and relatively unaffected by logjams were surveyed for assessment of fine sediment deposition. Half of the streams were in watersheds underlain by relatively erodible sandstone. The other half were underlain by a more resistant basalt. Channel morphology, hydraulic variables, particle size, relative pool volume of fine sediment (V*), and wood characteristics were measured in the streams. A significantly higher amount of fine sediment was deposited in the sandstone channels than in the basalt channels, as indicated by V*. Grab samples of sediment from pools also were significantly finer grained in the sandstone channels. Geographic information systems (GIS) software was used to derive several variables that might correlate with fine sediment deposition. These variables were combined with those derived from field data to create multiple linear regression models to be used for further exploration of the type and relative influence of factors affecting fine sediment deposition. Lithology appeared to be significant in some of these models, but usually was not the primary driver. The results from these models indicate that V* at the reach scale is best explained by stream power per unit area and by the volume of wood perpendicular to the flow per channel area (R2 = 0.46). Findings show that V* is best explained using only watershed scale variables, including negative correlations with relief ratio and basin precipitation index, and positive correlations with maximum slope and circularity.
 
SAGE (2007). Scoping Document, Eastside Forest Hydrology: A proposed study for the Eastside Type N Characterization Project, CMER.
Sahu, B. K. (1964). "Transformation of weight frequency and number frequency data in size distribution studies of clastic sediments." Journal of Sedimentary Petrology 34(4): 768-773.
Sanborn, S. C., and B.P. Bledsoe (2006). "Predicting streamflow regime metrics for ungauged streams in Colorado, Washington, and Oregon." Journal of Hydrology 325: 241-261.
Sandberg, D. V., et al. (1999). National Strategic Plan: Modeling and Data Systems for Wildland Fire and Air Quality, USDA Forest Service: 60.
Sand-Jensen, K. (2007). "How to write consistently boring scientific literature." Oikos doi:10.1111/j.2007.0030-1299.15674.x.
Santacana, N., et al. (2003). "A GIS-based multivariate statistical analysis for shallow landslide susceptibility mapping in la Pobla de Lillet area (Eastern Pyrenees, Spain)." Natural Hazards 30: 281-295.
This paper presents a GIS-aided procedure for shallow landslide susceptibility mapping
at a regional scale. Most of the input data for the susceptibility assessment have been captured
automatically. A total of 13 parameters, related to the slope geometry, have been derived from the
digital elevation model (DEM) while vegetation cover and thickness of superficial formations have
been obtained from photointerpretation and field work. The susceptibility assessment is based on
multivariate statistical techniques (discriminant analysis), which has been tested in a pilot area in
La Pobla de Lillet (Eastern Pyreenes, Spain). The results obtained using a random sample show
that 82% of all the cells, and 90% of cells including slope failures, have been properly classified. A
susceptibility map based on the discriminant function has given consistent results. The susceptibility
assessment is very sensitive to the parameters selected. Compared to the traditional methods, the
main advantage of the GIS-aided procedure is the rapidity provided by the automatic capture of
parameters. It also has the capability of covering large areas, and the objectivity and reproducibility
of the results. The main drawback is that, at present, not all regions have DEM accurate enough to
cope with small landslides.
 
Santi, P. M. (1988). What Happens Between the Scar and the Fan? The Behavior of a Debris Flow in Motion. 24th Symposium on Engineering Geology and Soils Engineering.
Santi, P. M., et al. (2008). "Sources of debris flow material in burned areas." Geomorphology 96: 310-321.
The vulnerability of recently burned areas to debris flows has been well established. Likewise, it has been shown that many, if
not most, post-fire debris flows are initiated by runoff and erosion and grow in size through erosion and scour by the moving debris
flow, as opposed to landslide-initiated flows with little growth. To better understand the development and character of these flows,
a study has been completed encompassing 46 debris flows in California, Utah, and Colorado, in nine different recently burned
areas. For each debris flow, progressive debris production was measured at intervals along the length of the channel, and from these
measurements graphs were developed showing cumulative volume of debris as a function of channel length. All 46 debris flows
showed significant bulking by scour and erosion, with average yield rates for each channel ranging from 0.3 to 9.9 m3 of debris
produced for every meter of channel length, with an overall average value of 2.5 m3/m. Significant increases in yield rate partway
down the channel were identified in 87% of the channels, with an average of a three-fold increase in yield rate. Yield rates for short
reaches of channels (up to several hundred meters) ranged as high as 22.3 m3/m. Debris was contributed from side channels into the
main channels for 54% of the flows, with an average of 23% of the total debris coming from those side channels. Rill erosion was
identified for 30% of the flows, with rills contributing between 0.1 and 10.5% of the total debris, with an average of 3%. Debris
was deposited as levees in 87% of the flows, with most of the deposition occurring in the lower part of the basin. A median value of
10% of the total debris flow was deposited as levees for these cases, with a range from near zero to nearly 100%. These results
show that channel erosion and scour are the dominant sources of debris in burned areas, with yield rates increasing significantly
partway down the channel. Side channels are much more important sources of debris than rills. Levees are very common, but the
size and effect on the amount of debris that reaches a canyon mouth is highly variable.
 
Sarikhan, I. Y. and T. A. Contreras (2009). Landslide Field Trip to Morton, Glenoma, and Randle, Lewis County, Washington. Olympic, Washington, Washington Department of Natural Resources: 13.
Sarr, D. A. and D. E. Hibbs (2006). "Woody riparian plant distributions in western Oregon, USA: comparing landscape and local scale factors." Plant Ecology.
Sartz, R. S. (1953). "Soil Erosion on a fire-denuded forest areas in the Douglas-fir region." Journal of Soil and Water Conservation 8: 197-209.
SASInstitute (1999). The SAS System for Windows, v.9.1. Cary, North Carolina.
Sassa, K. (1984). The mechanism to initiate debris flows as undrained shear of loose sediments. Interpraevent 1984, Villach, Austria.
Sauder, E. A., et al. (1987). Logging and Mass Wasting in the Pacific Northwest with Application to the Queen Charlotte Islands, B. C.: A Literature Review. Vancouver, British Columbia, Forest Engineering Research Institute of Canada.
Saunders, J. W. and M. W. Smith (?). "Physical Alteration of Stream Habitat to Improve Brook Trout Production "? 18: 185-188.
Savage, M., et al. (1992). "Diversity and disturbance in a Colorado subalpine forest." Physical Geography 13(3): 240-249.
Savage, W. Z., et al. (1996). The Slumgullion earth flow, southwestern Colorado. Seventh International Symposium on Landslides, Balkema, Rotterdam.
Savic, D. A., G.A.Walters and J.W. Davidson (1999). "A genetic programming approach to rainfall-runoff modelling." Water Resources Management 13: 219-231.
Sawada, T., et al. (1983). "Relationship between channel pattern and sediment transport in a steep gravel bed river." Zietshrift fur Geomorphologie Suppl.-Bd. 46: 55-66.
Scanlon, T. M., et al. (2000). "Shallow subsurface storm flow in a forested headwater catchment: Observations and modeling using a modified TOPMODEL." Water Resources Research 36(9): 2575-2586.
Transient, perched water tables in the shallow subsurface are observed at the
South Fork Brokenback Run catchment in Shenandoah National Park, Virginia. Crest
piezometers installed along a hillslope transect show that the development of saturated
conditions in the upper 1.5 m of the subsurface is controlled by total precipitation and
antecedent conditions, not precipitation intensity, although soil heterogeneities strongly
influence local response. The macroporous subsurface storm flow zone provides a
hydrological pathway for rapid runoff generation apart from the underlying groundwater
zone, a conceptualization supported by the two-storage system exhibited by hydrograph
recession analysis. A modified version of TOPMODEL is used to simulate the observed
catchment dynamics. In this model, generalized topographic index theory is applied to the
subsurface storm flow zone to account for logarithmic storm flow recessions, indicative of
linearly decreasing transmissivity with depth. Vertical drainage to the groundwater zone is
required, and both subsurface reservoirs are considered to contribute to surface
saturation.
 
 
Scarlett, W. S. and C. J. Cederholm (1983). Juvenile coho salmon fall-winter utilization of two small tributaries in the Clearwater River, Washington. Olympic Wild Fish Conference, Port Angeles, Washington, Peninsula College, Fisheries Technology Program.
Scarnecchia, D. L. and B. B. Roper (2000). "Large scale differential summer habitat use of three anadromous salmonids in a large river basin in Oregon, U.S.A." Fisheries Management and Ecology 7: 197-209.
Scarnecchia, D. L. and B. B. Roper (2000). "Large-scale differntial summer habitat use of three anadromous salmonds in a large river basin in Oregon, USA." Fisheries Management and Ecology 7: 197-209.
Schaetzl, R. J. and L. R. Follmer (1990). "Longevity of tree throw microtopography: implications for mass wasting." Geomorphology 3: 113-123.
Schasse, H. W. (1987). Geologic Map of the Centralia Quadrangle, Washington. Olympia, Washington Division of Geology and Earth Resources.
Scheidt, N. E. (2006). Stream succession: channel changes after wildfire disturbance, University of Idaho. Master of Science: 60.
Scherer, R. (2004). Decomposition and longevity of in-stream woody debris: a review of literature from North America. Forest land-Fish Conference II. Edmonton, Alberta: 127-133.
This paper presents a brief summary of a literature review completed to document the
current state of knowledge regarding the decomposition and longevity of in-stream woody
debris in western Canada. The summary provides brief overviews of riparian management
implications, decomposition processes and rates, sensitivity of in-stream woody debris models
to decomposition parameters and a summary of the longevity of in-stream woody debris.
 
Schermer, E. R., et al. (1984). "The Orgin of Allochthonous Terranes: Perspectives on the Growth and Shaping of Continents." Annual Review of the Earth and Planet Science 12: 107-131.
Schick, A. P., et al. (1987). Bed load transport in desert floods: observations in the Negev. Sediment Transport in Gravel-bed Rivers. C. R. Thorne, J. C. Bathurst and R. D. Hey, Schick, A.P.
Lekach, J.
Hassan M.A. chapter 20: 617-642.
Schilling, S. P. (1998). LAHARZ: GIS programs for automated mapping of lahar-inundation hazard zones. Vancouver, Washington, U.S. Geological Survey.
Schlesinger, M. E. (?). "Insights on future climatic changes from model simulations."?
Schlosser, I. J. (1985). "Flow Regime, Juvenile Abundance, and the Assemblage Structure of Stream Fishes." Ecology 66: 1484-1490.
Schlosser, I. J. (1991). "Stream fish ecology: A landscape perspective." BioScience 41(10): 704-712.
Schmetterling, D. A., et al. (2001). "Effects of riprap bank reinforcement on stream salmonids in the western United States." Fisheries 26(7): 6-13.
Angular rock riprap is used to reduce riverbank erosion in developed riparian corridors.
We reviewed peer reviewed as well as non-refereed literature to determine the
effects of riprap on salmonid habitat and populations and to identify areas for future (applied) research. Although commonly used to armor banks, riprap affects salmonid
populations and stream function. Riprap may provide habitat for juvenile salmonids
and bolster densities on reaches of streams that have been severely degraded.
However, riprap does not provide the intricate habitat requirements for multiple age
classes or species provided by natural vegetated banks. Streambanks with riprap have
fewer undercut banks, less low-overhead cover and are less likely than natural stream
banks to contribute large woody debris to the stream. Lateral streambank erosion is
a natural process that occurs in many stream types. However, most valley-bottom
stream types, which have the greatest tendency to laterally migrate, lie within developed
corridors. Although permitting of individual projects may attenuate localized
negative effects to streambanks, it may not effectively curtail cumulative effects to a
watershed. Our review further demonstrated that the practice of riprapping banks
goes against current practices and philosophies of stream renaturalization and
impedes future restoration work. Future research should determine the true effects
of riprap banks on salmonid densities, the use of soft techniques using for stabilizing
banks on rivers, and the cumulative effects of riprap projects on watersheds and fluvial
processes. We foresee a continued struggle for resource managers trying to
maintain natural fluvial processes while protecting public infrastructure and private
property from those same processes.
 
Schmidt, J. C. and D. M. Rubin (1995). Regulated Streamflow, Fine-Grained Deposits, and Effective Discharge in Canyons with Abundant Debris Fans. Natural and Anthropogenic Influences in Fluvial Geomorphology. J. E. Costa, A. J. Miller, K. W. Potter and P. R. Wilcock, American Geophysical Union. 89: 177-195.
Schmidt, K.-H. (1989). "The significance of scarp retreat for cenozoic landform evolution on the Colorado Plateau, U.S.A." Earth Surface Processes and Landforms 14: 93--105.
Schmidt, K.-H. and P. Ergenzinger (1991). "Bedload entrainment, travel lengths, step lengths, rest periods - studied with passive (iron, magnetic) and active (radio) tracer techniques." ESPL 17: 147-165.
This paper reports results from bedload transport investigations with active (radio) and passive (iron, magnetic) tracers in the Lainbach, a step-pool mountain river, in Bavaria, Southern Germany. The spatial distributions of the iron tracers after flood events can be best described by exponential or Gamma distributions. There is some indication of a tendency of size-selective transport of the iron tracers, but there is also a considerable amount of scatter in the correlations between weight (size) and travel length owing to the masking influence of other variables, such as the shape of the particles and different positions in the river bed. The experiments with artificial magnetic tracers showed that elongated pebbles (rods) had the longest mean transport distance, platy ones (discs) remained relatively close to the starting points. The particles from the pool showed the greatest transport lengths and a 100% chance of being eroded.
  The Pebble Transmitter System (PETSY) consists of transmitters implanted into individual pebbles, a computerized receiver, a stationary antenna system with an antenna switchboard, and a data logging system. The particles do not move continuously but in a series of transport steps and non-movement intervals. A single value for a given size-class is not adequate to describe the critical conditions of entrainment under natural circumstances. A probability approach is much more suitable. The critical unit discharges (total discharge divided by active channel width) along the measuring reach are dependent on river bed morphology. In the steps bedload needs higher unit discharges to be entrained. Once entrained, the transport of bedload is stochastic in nature and the single particle transport is controlled by the step lengths and the duration of rest periods. The distributions of both parameters can be approximated by exponential functions. Applying the stochastic concept proposed by Einstein the mean values of the random variables 'step length' and 'duration of rest period' measured with the PETSY technique were used for the simulation of spatial distributions of bedload particles from point sources. More field and laboratory data are needed to include varying flow and roughness conditions with tracers representing different particle characteristics.
 
Schmidt, K. M. (1994). Mountain scale strength properties, deep-seated landsliding, and relief limits. Department of Geological Sciences. Seattle, WA, University of Washington: 166.
Schmidt, K. M. (1999). Root Strength, Colluvial Soil Depth, and Colluvial Transport on Landslide-Prone Hillslopes. Department of Geological Sciences. Seattle, University of Washington: 255.
Schmidt, K. M. and D. R. Montgomery (1993). A Rock Mass Strength Classification for Deep-Seated Landslide Susceptibility. 1993 Fall AGU Meeting.
Schmidt, K. M., 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.
Schnackenberg, E. and L. H. MacDonald (1998). "Detecting cumulative effects on headwater streams in the Routt Nation Forest, Colorado." Journal of American Water Resources Association 34(5): 1163-1177.
Schreider, S. Y., et al. (1996). "Modelling rainfall-runoff from large catchment to basin scale: the Goulburn Valley, Victoria." Hydrological Processes 10: 863-876.
Schreve, R. L. (1969). "Stream lengths and basin areas in topologically random channel networks." Journal of Geology 77: 397-414.
Schroeder, R. L. and G. W. Brown (1984). Debris torrents, precipitation, and roads in two coastal Oregon watersheds. Symposium on effects of forest land use on erosion and slope stability.
Schroeder, W. L. and J. V. Alto (1983). "Soil properties for slope stability analysis: Oregon and Washington Coastal Mountains." Forest Science 29(4): 823-833.
Schroeder, W. L. and J. V. Alto (1983). "Soil properties for slope stability analysis: Oregon and Washington Coastal Mountains. Forest Science." Forest Science 29(4): 823-833.
Schtickzelle, N. and T. P. Quinn (2007). "A metapopulation perspective for salmon and other anadromous fish." Fish and Fisheries 8: 297-314.
Salmonids are an important component of biodiversity, culture and economy in
several regions, particularly the North Pacific Rim. Given this importance, they have
been intensively studied for about a century, and the pioneering scientists recognized
the critical link between population structure and conservation. Spatial structure is
indeed of prime importance for salmon conservation and management. At first
glance, the essence of the metapopulation concept, i.e. a population of populations,
widely used on other organisms like butterflies, seems to be particularly relevant to
salmon, and more generally to anadromous fish. Nevertheless, the concept is rarely
used, and barely tested.
Here, we present a metapopulation perspective for anadromous fish, assessing in
terms of processes rather than of patterns the set of necessary conditions for
metapopulation dynamics to exist. Salmon, and particularly sockeye salmon in
Alaska, are used as an illustrative case study. A review of life history traits indicates
that the three basic conditions are likely to be fulfilled by anadromous salmon: (i) the
spawning habitat is discrete and populations are spatially separated by unsuitable
habitat; (ii) some asynchrony is present in the dynamics of more or less distant
populations and (iii) dispersal links populations because some salmon stray from their
natal population. The implications of some peculiarities of salmon life history traits,
unusual in classical metapopulations, are also discussed.
Deeper understanding of the population structure of anadromous fish will be
advanced by future studies on specific topics: (i) criteria must be defined for the
delineation of suitable habitats that are based on features of the biotope and not on
the presence of fish; (ii) the collection of long-term data and the development of
improved methods to determine age structure are essential for correctly estimating
levels of asynchrony between populations and (iii) several key aspects of dispersal are
still poorly understood and need to be examined in detail: the spatial and temporal
scales of dispersal movements, the origin and destination populations instead of
simple straying rates, and the relative reproductive success of immigrants and
residents.
 
Schuett-Hames, D., et al. (1992). T-F-W Ambient Monitoring Manual. Olympia, Washington, Northwest Indian Fisheries Commission.
Schuett-Hames, D. and B. Conrad (1995). Spawning Gravel Scour: Literature Review and Recommendations for a Watershed Analysis Monitoring Methdology (DRAFT), Northwest Indian Fisheries Commission: 30.
Schuett-Hames, D., et al. (1999, 1996). "Literature Review and Monitoring Recommendations for Salmonid Spawning Gravel Scour ". Retrieved 11/19/99, 1999, from http://199.245.112.101/TFW/reports/reports2.htm.
Schuett-Hames, D., et al. (1999). Method Manual for the Salmonid Spawning Gravel Scour Survey, NW Indian Fisheries Commission and Timber, Fish and Wildlife: 41.
Schuett-Hames, D., et al. (2000). "Patterns of gravel scour and fill after spawning by chum salmon in a western Washington stream." North American Journal of Fisheries Management 20: 610-617.
The patterns of gravel scour and fill during the incubation period of chum salmon
Oncorhynchus keta were studied in Kennedy Creek, a low-gradient alluvial channel draining into
southern Puget Sound, Washington. In 1991–1992, scour occurred during two storms having
estimated return intervals of 1.4 years and less than 1 year. Scour to median egg pocket depth
(0.2 m) occurred at 20% of the monitored locations during the greater event and at 1.3% during
the lesser event. Differences in depth of scour were observed between the two study reaches and
among habitat units (riffles, lower riffles, glides, pool tailouts, and pool lateral bars) within reaches.
Average depth of scour in a relatively simple, straight, and narrow reach of the creek was 0.075
m, whereas average scour was nearly twice as deep (0.140 m) at a more complex, sinuous, and
wide reach. Scour to median egg pocket depth occurred at 28% of the monitors in the complex
section compared with only 9% in the simpler section. Average scour depth and the percentage
of monitors scoured to median egg pocket depth were greater in pool-associated habitats (pool
lateral bars and pool tailouts) than in riffle-associated habitats (riffles, lower riffles, and glides).
Chum salmon often used sites such as pool tailouts that would be suitable under moderate flow
conditions but that would be prone to scour under higher flow events. The relatively widespread
scour to egg pocket depths that we observed during a bank-full event indicates that scour can be
a significant source of egg-to-fry mortality for salmonids spawning in low-gradient pool–riffle
channels.
 
 
Schuett-Hames, D., et al. (1994). Patterns of scour and fill in a low-gradient alluvial channel. Part C, Chapter 1. The Effect of Forest Practices on Fish Populations, Report TFW-F4-94-001. T. P. Quinn and N. P. Peterson. Olympia, Washington Dept. of Natural Resources: 42-55.
Schuett-Hames, D., et al. (?). Incubation Environment of Chum Salmon (Oncorhynchus keta) in Kennedy Creek: Part B. Patterns of Scour and Fill in a Low-Gradient Alluvial Channel 42-55.
Schuett-Hames, D., et al. (?). "Incubation Environment of Chum Salmon (Oncorhynchus keta) in Kennedy Creek: Part C. Patterns of Scour and Fill in a Low-Gradient Alluvial Channel "  Chapter I: 42-55.
Schuett-Hames, D. and A. Pleus (1996). Literature Review & Monitoring Recomomendations for Salmonid Spawning Habitat Availability.
Schuett-Hames, D., et al. (1999). Method Manual for the Large Woody Debris Survey. TFW Monitoring Program, Prepared for the Washington State Dept. of Natural Resources under the Timber, Fish, and Wildlife Agreement.
Schuett-Hames, D., et al. (1994). TFW Ambient Monitoring Program Manual. Olympia, Washington, Northwest Indian Fisheries Commision.
Schuldt, J. A. and A. E. Hershey (1995). "Effect of salmon carcass decomposition on Lake Superior tributary streams " J. N. Am. Benthol. Soc. 14(2): 259-268.
Schulz, W. H., et al. (2008). "Modeling the spatial distribution of landslide-prone colluvium and shallow groundwater on hillslopes of Seattle, WA." Earth Surface Processes and Landforms 33: 123-141.
Landslides in partially saturated colluvium on Seattle, WA, hillslopes have resulted in
property damage and human casualties. We developed statistical models of colluvium and
shallow-groundwater distributions to aid landslide hazard assessments. The models were
developed using a geographic information system, digital geologic maps, digital topography,
subsurface exploration results, the groundwater flow modeling software VS2DI and regression
analyses. Input to the colluvium model includes slope, distance to a hillslope–crest
escarpment, and escarpment slope and height. We developed different statistical relations
for thickness of colluvium on four landforms. Groundwater model input includes colluvium
basal slope and distance from the Fraser aquifer. This distance was used to estimate hydraulic
conductivity based on the assumption that addition of finer-grained material from downsection
would result in lower conductivity. Colluvial groundwater is perched so we estimated
its saturated thickness. We used VS2DI to establish relations between saturated thickness
and the hydraulic conductivity and basal slope of the colluvium. We developed different
statistical relations for three groundwater flow regimes. All model results were validated
using observational data that were excluded from calibration. Eighty percent of colluvium
thickness predictions were within 25% of observed values and 88% of saturated thickness
predictions were within 20% of observed values. The models are based on conditions common
to many areas, so our method can provide accurate results for similar regions; relations
in our statistical models require calibration for new regions. Our results suggest that Seattle
landslides occur in native deposits and colluvium, ultimately in response to surface-water
erosion of hillslope toes. Regional groundwater conditions do not appear to strongly affect
the general distribution of Seattle landslides; historical landslides were equally dispersed
within and outside of the area potentially affected by regional groundwater conditions.
 
 
Schumm, S. A. (1960). The Shape of Alluvial Channels in Relation to Sediment Type. Erosion and Sedimentation in a Semiarid Environment, US Geological Survey: 17-30.
Schumm, S. A. (1977). The Fluvial System. New York, John Wiley and Sons: 338.
Schumm, S. A. and R. W. Lighty (1965). "Time, space, and causality in Geomorphology." American Journal of Science 263: 110-119.
Schumm, S. A. and D. K. Rea (1995). "Sediment yield from disturbed earth systems." Geology 23(5): 391-394.
Schumm, S. A. and M. A. Stevens (1973). "Abrasion in place: a mechanism for rounding and size reduction of coarse sediments in rivers." Geology: 37-40.
Schuster, R. L., et al. (1987). Irrigation-induced landslides in fluvial-lacustrine sediments, south-central Washington state. Fifth International Conference and Field Workshop on Landslides, Christchurch, New Zealand.
Schuster, R. L., et al. (1996). "Mass wasting triggered by the 5 March 1987 Ecuador earthquakes." Engineering Geology 42: 1-23.
Schwab, J. W. (1998). Landslides on the Queen Charlotte Islands, rates and climatic events. Carnation Creek and Queen Charlotte Islands Fish/Forestry Workshop: Applying 20 Years of Coast Research to Management Solutions, Queen Charlotte City, B. C. Ministry of Forests, Research Branch, Victoria, B.C.
Schwarz, G. E., Smith, R.A., Alexander, R.B., and Gray, J.R., (2001). A spatially referenced regression model (SPARROW) for suspended sediment in streams of the conterminous U.S., in U.S. Subcommittee on Sedimentation, Proceedings of the Seventh Federal Interagency Sedimentation Conference,, Reno, Nevada USA.
Schwind, B. c. (2007). Monitoring Trends in Burn Severity: Report on the 2004 Mapped Fires.
Sciences, N. A. o. (1986). Ecological Knowledge and Enivronmental Problem Solving:Concepts and Case Studies. Washington D.C., National Academy Press.
Scott, K. M. (1971). Origin and Sedimentology of 1969 Debris Flows near Glendora, California, US Geological Survey: C242 - C246.
Scott, K. M. and G. C. Gravlee, Jr. (1968). Flood Surge on the Rubicon River, California - Hydrology, Hydraulics, and Boulder transport. Physiographic and Hydraulic Studies of Rivers, U.S. Geological Survey: M1-M38.
Scott, K. M. and R. P. Williams (1978). Erosion and sediment yields in the Transverse Ranges, Southern California, US Geological Survey: 38.
Scott Paper Company (1984). Application for License for Minor Water Power Projects, Scott Paper Company: ?
Scott Paper Company (1984). United States of America Before the Federal Energy Regulatory Commission: Application for License for Minor Water Power Project, Scott Paper Company.
Scrivener, J. C. and M. J. Brownlee (1989). "Effects of Forest Harvesting on Spawning Gravel and Incubation Survival of Chum (Oncorhynchus keta) and Coho Salmon (O. kistuch) in Carnation Creek, British Columbia " Canadian Journal of Fisheries and Aquatic Sciences 46: 681-696.
Scully, M. E., et al. (2002). "Application of an analytical model of critically stratified gravity-driven sediment transport and deposition to observations from the Eel River continental shelf, Northern California." Continental Shelf Research 22(14): 1951-1974.
Seaber, P. R., et al. (1987). Hydrologic Unit Maps, USGS Water-Supply Paper 2294, US Geological Survey: 66.
Seal, R. and C. Paola (1995). "Observations of downstream fining on the North Fork Toutle River near Mount St. Helens, Washington." Water Resources Research 31(5): 1409-1419.
Sear, D. A., et al. (1999). "Groundwater dominated rivers." Hydrological Processes 13: 255-276.
This paper explores the signi®cance of groundwater dominance in the surface water system through a com-
bination of review and an exposition of the general hydrology, ecology and geomorphology of rivers draining
the main UK aquifers. Groundwater dominance is shown to vary according to the nature of the aquifer
lithology, the mechanism of groundwater:surface water interaction and the scale at which one examines this
interaction. Using data derived from a range of studies including the UK Environment Agency River Habitat
Survey and the UK Institute of Freshwater Ecology RIVPACS invertebrate database it is shown that the nature
of the aquifer and mode of in¯uent discharge strongly control the hydrological and ecological characteristics of
the environment but that a speci®c groundwater ecology or hydrogeomorphology is masked by the overriding
controls exerted by aquifer geology and catchment topography. Despite this, it is clear that river systems
dominated by groundwater ¯ows have speci®c hydrological characteristics and management issues that require
holistic, multidisciplinary approaches that recognise the signi®cance of groundwater and the nature of the
interaction with the surface water environment.
 
Sedell, J. R., et al. (1988). What we know about large trees that fall into streams and rivers. From the Forest to the Sea: A Story of Fallen Trees. C. Maser, R. F. Tarrant, J. M. Trappe and J. F. Franklin. Portland, Oregon, USA, United States Forest Service General Technical Report, PNW-GTR-229.
Sedell, J. R. and C. N. Dahm (1984). Catastrophic disturbances to stream ecosystems: volcanism and clear-cut logging. Current perspectives in microbial ecology. M. J. Klug and C. A. Reddy. Washington, D.C., East Lansing and American Society of Microbiology: 531-539.
Sedell, J. R., et al. treamside vegetation management for aquatic habitat (draft), USDA Forest Service: 1-11.
Sedell, J. R., et al. (1982). Fish Habitat and Steamside Management: Past and Present. Annual Meeting of the Society of of American Foresters, Bethseda, Maryland.
Sedell, J. R. and J. L. Froggatt (1984). "Importance of streamside forests to large rivers: the isolation of the Willamette River, Oregon, U.S.A., from its floodplain by snagging and streamside forest removal." Verhandlungen-Internationale Vereinigung fuer Theorelifche und Angewandte Limnologie (International Association of Theoretical and Applied Limnology) 22: 1828-1834.
Sedell, J. R. and K. J. Luchessa (1981). Using the Historical Record as an Aid to Salmonid Habitat Enhancement. Symposium on Aquisition and Utilization of Aquatic Habitat Inventory Information, Portland, Oregon.
Sedell, J. R., et al. (1990). "Role of refugia in recovery from disturbances: modern fragmented and disconnected river systems." Environmental Management 14(5): 711-724.
Sedell, J. R. and F. J. Triska (1977). Biological consequences of large organic debris in northwest streams. Logging Debris in Streams Workshop 2, Corvallis, OR, Oregon State University.
Seed, H. B. (1979). "Soil Liquefaction and Cyclic Mobility Evaluation for Level Ground During Earthquakes." Journal of the Geotechnical Engineering Division 2: 201-255.
Seed, H. B., et al. (1971). "Simplified Procedure for Evaluating Soil Liquefaction Potential." Journal of the Soil Mechanics and Foundations Division 9: 1249-1273.
Seegrist, D. W. and R. Gard (1972). "Effects of Floods on Trout in Sagehen Creek, California " Transactions of the American Fisheries Society 3: 478-482.
Seibert, J. and B. L. McGlynn (2007). "A new triangular multiple flow direction algorithm for computing upslope areas from gridded digital elevation models." Water Resources Research 43: W04501.
Gridded digital elevation data, often referred to as DEMs, are one of the most widely
available forms of environmental data. Topographic analysis of DEMs can take many
forms, but in hydrologic and geomorphologic applications it is typically used as a
surrogate for the spatial variation of hydrological conditions (topographic indices) and
flow routing. Here we report on a new flow routing algorithm and compare it to three
common classes of algorithms currently in widespread use. The advantage of the new
algorithm is that unrealistic dispersion on planar or concave hillslopes is avoided, whereas
multiple flow directions are allowed on convex hillslopes. We suggest that this new
triangular multiple flow direction algorithm (MD1) is more appropriate for a range of
flow routing and topographic index applications
 
Selby, M. J. (1985). Earth's Changing Surface. Oxford, Clarendon Press.
Semmens, D. J., D.C.Goodrich, C.L. Unkrich,  R.E. Smith, D.A. Woolhiser,  and S.N. Miller "Kineros2 and the AGWA modeling framework."
Semmens, D. J., S.N. Miller, M. Hernandez,  I.S.Burns, W.P.Miller,  D.C.Goodrich, and W. G. Kepner (2004). Automated Geospatial Watershed Assessment (AGWA) - A GIS-Based Hydrologic Modeling Tool: Documentation and User Manual., U.S. Department of Agriculture, Agricultural Research Service, ARS-1446.
Sempeski, P. and P. Gaudin (1995). "Habitat selection by grayling - I. Spawning habitats " Journal of Fish Biology 47: 256-265.
Seo, J. I., et al. (2008). "Factors controlling the fluvial export of large woody debris, and its contribution to organic carbon budgets at watershed scales." Water Resources Research 44(W04428).
The fluvial export of large woody debris (LWD) was monitored in 131 reservoirs
throughout Japan. Published data on the fluvial export of dissolved and particulate
organic carbon were used to estimate the contributions of LWD in carbon budgets. Of all
variables tested, watershed area was most important in explaining LWD carbon
(LWDC) export, followed by annual precipitation. LWDC export per unit area was
relatively high in small watersheds, highest in intermediate-sized watersheds, and
decreased in large watersheds. In small watersheds, a large proportion of LWD retained on
narrow valley floors may fragment or decay and eventually be exported in forms other
than LWD. In intermediate-sized watersheds, LWD supplied from upstream and recruited
by bank erosion is consistently transported downstream. In large watersheds, LWD
recruitment is limited and LWD transported from upstream is stored on large floodplains.
These differences in LWD recruitment, retention and transport in watersheds of
different sizes lead to the proportion of LWDC in organic carbon exports to be maximum
in intermediate-sized watersheds and decline rapidly in large watersheds.
 
Service, U. F. (2004). Best Management Practices Evaluation Program: 1992-2002 monitoring results. R. USDA Forest Service, Vallejo, California.
Servizi, J. A. and D. W. Martens (1992). "Sublethal Responses of Coho Salmon (Oncorhynchus kisutch) to Suspended Sediments " Canadian Journal of Fisheries and Aquatic Sciences 49: 1389-1394.
Sessions, J., et al. (1987). "Road location and construction practices: effects on landslide fequency and size in the Oregon Coast Range." Western Journal of Applied Forestry 2(4): 119-124.
Shackley, S. and B. Wynne (1995). "Integrating knowledges for climate change: Pyramids, nets and uncertainties." Global Environmental Change 5(2): 113-126.
Shahlaee, A. K., et al. (1991). "Runoff and sediment production from burned forest sites in the Georgia Piedmont." Water Resources Bulletin 27(3): 485-493.
Shakesby, R. A., and S.H. Doerr (2006). "Wildfire as a hydrological amd geomorphological agent." Earth Science Reviews 74: 269-307.
Wildfire can lead to considerable hydrological and geomorphological change, both directly by weathering bedrock surfaces and changing soil structure and properties, and indirectly through the effects of changes to the soil and vegetation on hydrological and geomorphological processes. This review summarizes current knowledge and identifies research gaps focusing particularly on the contribution of research from the Mediterranean Basin, Australia and South Africa over the last two decades or so to the state of knowledge mostly built on research carried out in the USA.
 
Wildfire-induced weathering rates have been reported to be high relative to other weathering processes in fire-prone terrain, possibly as much as one or two magnitudes higher than frost action, with important implications for cosmogenic-isotope dating of the length of rock exposure. Wildfire impacts on soil properties have been a major focus of interest over the last two decades. Fire usually reduces soil aggregate stability and can induce, enhance or destroy soil water repellency depending on the temperature reached and its duration. These changes have implications for infiltration, overland flow and rainsplash detachment. A large proportion of publications concerned with fire impacts have focused on post-fire soil erosion by water, particularly at small scales. These have shown elevated, sometimes extremely large post-fire losses before geomorphological stability is re-established. Soil losses per unit area are generally negatively related to measurement scale reflecting increased opportunities for sediment storage at larger scales. Over the last 20 years, there has been much improvement in the understanding of the forms, causes and timing of debris flow and landslide activity on burnt terrain. Advances in previously largely unreported processes (e.g. bio-transfer of sediment and wind erosion) have also been made.
 
Post-fire hydrological effects have generally also been studied at small rather than large scales, with soil water repellency effects on infiltration and overland flow being a particular focus. At catchment scales, post-fire accentuated peakflow has received more attention than changes in total flow, reflecting easier measurement and the greater hazard posed by the former. Post-fire changes to stream channels occur over both short and long terms with complex feedback mechanisms, though research to date has been limited.
 
Research gaps identified include the need to: (1) develop a fire severity index relevant to soil changes rather than to degree of biomass destruction; (2) isolate the hydrological and geomorphological impacts of fire-induced soil water repellency changes from other important post-fire changes (e.g. litter and vegetation destruction); (3) improve knowledge of the hydrological and geomorphological impacts of wildfire in a wider range of fire-prone terrain types; (4) solve important problems in the determination and analysis of hillslope and catchment sediment yields including poor knowledge about soil losses other than at small spatial and short temporal scales, the lack of a clear measure of the degradational significance of post-fire soil losses, and confusion arising from errors in and lack of scale context for many quoted post-fire soil erosion rates; and (5) increase the research effort into past and potential future hydrological and geomorphological changes resulting from wildfire.
 
Shannon and Wilson, I. (1991). Tolt and Raging Rivers Channel Migration Study, King County, Washington. Seattle, WA, King County Surface Water Management Division: 80.
Sharma, A., et al. (1997). "Streamflow simulation: A nonparametric approach." Water Resources Research 33(2): 291-308.
Sharma, K. D. and J. F. Correia (1989). "Upland erosion in a mixed pine forest ecosystem." Proceedings of the Indian Academy of Sciences: Earth and Planetary Sciences 98(2): 133-145.
Sharma, K. D., et al. (1996). "Modeling spatial sediment delivery in an arid region using thematic mapper data and GIS." Transactions of the ASAE 39(2): 551-557.
Sharma, K. D., et al. (1996). "Modeling sediment delivery in arid upland basins." Transactions of the ASAE 39(2): 517-524.
Sharma, P. P., et al. (1995). "Raindrop-Induced Soil Detachment and Sediment Transport from Interrill Areas." Soil Science Society of America Journal 59(3): 727-734.
Sharma, R. and R. Hilborn (2001). "Empirical relationships between watershed characteristics and coho salmon (Oncorhynchus kisutch) smolt abundance in 14 western Washington streams." Canadian Journal of Fisheries and Aquatic Science 58: 1453-1463.
We assembled data on coho salmon (Oncorhynchus kisutch) from 14 streams in western Washington, including
annual smolt counts and annual escapement, either as absolute counts or as an index. We also compiled data on large
woody debris (number·km–1 of stream), road densities in the watersheds (km road·km–2), gradient of the streams (%), valley
slope adjacent to the stream (%), drainage area in the watershed (km2), and pool, pond, and lake areas (m2·km–1). We
explored the relationships between habitat variables and two measures of coho production, the maximum production of
smolts in the stream (capacity) and the maximum smolts/spawner (productivity). Using the 11 streams with pool and pond
counts, we found that pool and pond densities served as good predictors of smolt density (r 2 = 0.85 for pools and 0.68
for ponds, independently). Pools produced 0.39 smolts·m–2 and ponds produced 0.07 smolts·m–2 in the multiple regression
fit, accounting for 92% of the residual error. We also found that lower valley slopes, lower road densities, and lower
stream gradients were correlated with higher smolt density.
 
Sharma, S. (1990). XSTABL: an integrated slope stability analyusis program for personal computers. Moscow, ID, Interactive Software Designs.
Sharp, R. P. (1942). "Mudflow Levees." Journal of Geomorphology 5(3).
Sharp, R. P. and L. H. Nobles (1953). "Mudflow of 1941 at Wrightwood, Southern California." Bulletin of the Association of Engineering Geologists 64: 547-560.
Shaw, E. A., and J.S. Richardson (2001). "Direct and indirect effects of sediment pulse duration on stream invertbrate assemblages and rainbow trout (oncorhynchus mykiss) growth and survival." Canadian Journal of Fisheries and Aquatic Sciences 58: 2213-2221.
Shaw, S. C. and D. H. Johnson (1995). Slope morphology model derived from digital elevation data. Northwest ARC/INFO Users Conference. Coeur d'Alene, ID.
Shaw, S. C. and L. Vaugeois (1999). Comparison of GIS-based Models of Shallow Landsliding for Application to Watershed Management. Olympia, WA, Washington Department of Natural Resources, Forest Practices Division.
Shaw, S. C. and L. M. Vaugeois (1999). Comparison of GIS-based Models of Shallow Landsliding for Application to Watershed Management, Washington Department of National Resources: 1-103.
Sheer, M. B., et al. (2009). Development and management of fish intrinsic potential data and methods: state of the IP 2008 summary report: Pacific Northwest Aquatic Monitoring Partnership Series 2009-004: 56.
Sheer, M. B. and E. A. Steel (2006). "Lost watersheds: barriers, aquatic habitat connectivity, and salmon persistence in the Willamette and Lower Columbia river basins." Transactions of the American Fisheries Society 135: 1645-1669.
Abstract.—Large portions of watersheds and streams are lost to anadromous fishes because of
anthropogenic barriers to migration. The loss of these streams and rivers has shifted the distribution of
accessible habitat, often reducing the diversity of accessible habitat and the quantity of high-quality habitat.
We combined existing inventories of barriers to adult fish passage in the Willamette and Lower Columbia
River basins and identified 1,491 anthropogenic barriers to fish passage blocking 14,931 km of streams. We
quantified and compared the stream quality, land cover, and physical characteristics of lost versus currently
accessible habitat by watershed, assessed the effect of barriers on the variability of accessible habitats, and
investigated potential impacts of habitat reduction on endangered or threatened salmonid populations. The
majority of the study watersheds have lost more than 40% of total fish stream habitat. Overall, 40% of the
streams with spawning gradients suitable for steelhead (anadromous rainbow trout Oncorhynchus mykiss),
60% of streams with riparian habitat in good condition, and 30% of streams draining watersheds with all
coniferous land cover are no longer accessible to anadromous fish. Across watersheds, hydrologic and
topographic watershed characteristics were correlated with barrier location, barrier density, and the impacts of
barriers on habitat. Population-based abundance scores for spring Chinook salmon O. tshawytscha were
strongly correlated with the magnitude of habitat lost and the number of lowland fish passage barriers. The
characteristics of barrier and habitat distribution presented in this paper indicate that barrier removal projects
and mitigation for instream barriers should consider both the magnitude and quality of the lost habitat.
 
Shen, D. Y., A.N.Ma, H. Lin, X.H. Nie, S.J.Mao, B.Zhang, and J.J. Shi (2003). "A new approach for simulating water erosion on hillslopes." International Journal of Remote Sensing 24(14): 2819-2835.
Sheng, M. D., et al. (1990). Coho salmon enhancement in British Columbia using improved groundwater-fed side channels. Vancouver, British Columbia, Department of Fisheries and Oceans, Salmonid Enhancement Program: 81.
Shepard, B. B., et al. (1997). "Status and risk of extinction for westslope cutthroat trout in the Upper Missouri River Basin, Montana." North American Journal of Fisheries Management 17: 11588-11172.
Shepard, B. G., et al. (1986). "Relationships Between Stream and Intragravel Temperatures in Coastal Drainages, and Some Implications for Fisheries Workers." Canadian Journal of Fisheries and Aquatic Sciences 43: 1818-1822.
Sheridan, G. J., et al. (2000). "Estimation of erosion model erodibility parameters from media properties." Australian Journal of Soil Research 38(2): 265-284.
Sherriff, R., and  Veblen, T. (2007). "A Spatially-Explicit Reconstruction of Historical Fire Occurrence in the Ponderosa Pine Zone of the Colorado Front Range." Ecosystems 10(2): 311-323.
Shields, F. D. and C. Gippel (1995). "Prediction of effects of woody debris removal on flow resistance." Journal of Hydraulic Engineering 121(4): 341-353.
Shih, S.-M. and P. D. Komar (1990). "Differential bedload transport rates in a gravel-bed stream: a grain-size distribution approach." Earth Surface Processes and Landforms 15: 539-552.
Shindler, B. and L. A. Cramer (1999). "Shifting Public Values for Forest Management: Making Sense of Wicked Problems." Retrieved 8-18-99, 1999, from http://ww.fs.fed.us/eco/eco-watch/wickedpr.html.
Shirazi, M. A. and W. K. Seim (1981). "Stream System Evaluation With Emphasis on Spawning Habitat for Salmonids " Water Resources Research 17(3): 592-594.
Shirvell, C. S. (1989). "Ability of PHABSIM to Predict Chinook Salmon Spawning Habitat " Regulated Rivers: Research & Management 3: 277-289.
Shirvell, C. S. and R. G. Dungey (1983). "Microhabitats Chosen by Brown Trout for Feeding and Spawning in Rivers " Transactions of the American Fisheries Society 112(3): 355-367.
Shreve, R. L. (1966). "Statistical law of stream numbers." Journal of Geology 75: 178-186.
Shugart, H. H. (1987). "Dynamic ecosystem consequences of tree birth and death patterns: A set of computer models predicts long-term behavior of forests." BioScience 37(8): 596-601.
Shumm, S. A. (?). The Shape of Alluvial Channels in Relation to Sediment Type. Erosion and Sedimentation in a Semiarid Enivironment. ?: 17-29.
Sias, J. C. and et al. (1998). "A calibrated model of wood recruitment to channels from unmanged riparian zones."
Sickle, J. V. (1981). "Long-term distributions of annual sediment yields from small watersheds." Water Resources Research 17(3): 659-663.
Sickle, J. V. and R. L. Beschta (1983). "Supply-based models of suspended sediment transport in streams." Water Resources Research 19(3): 768-778.
Sickle, J. V. and S. V. Gregory (1990). "Modeling inputs of large woody debris to streams from falling trees." Canadian Journal of Forest Research 20: 1593-1601.
Sidle, R. C. (1985). Factors influencing the stability of slopes, USDA Forest Service: 17-25.
Sidle, R. C. (1987). "A dynamic model of slope stability in zero-order basins." IAHS Publ. 165: 101-110.
Sidle, R. C. (1992). "A theoretical model of the effects of timber harvesting on slope stability." Water Resources Research 28(7): 1897-1910.
Sidle, R. C., S. Sasaki, M. Otsuki, S. Noguchi, and A.R. Nik (2004). "Sediment pathways in a tropical forest: effects of logging roads and skid trails." Hydrological Processes 18: 703-720.
Sidle, R. C. and H. Ochiai (2006). Landslides Processes, Prediction, and Land Use, American Geophysical Union.
Sidle, R. C., et al. Effects of Land Management on Soil Mass Movement. Hillslope Stability and Land Use. Chapter 5: 73-88.
Sidle, R. C., et al. (1985). Hillslope Stability and Land Use. Washington D. C., American Geophysical Union.
Sidle, R. C. and D. N. Swanston (1982). "Analysis of a small debris slide in coastal Alaska." Canadian Geotechnical Journal 19(2): 167 - 174.
Sidle, R. C., et al. (2000). "Streamflow generation in steep headwaters: a linked hydro-geomorphic paradigm." Hydrological Processes 14: 369-385.
Sidle, R. C. and W. Wu (?). Evaluation of the Impacts of Timber Harvesting on Landslide Occurrence: ?
Siebert, J., and J.J. McDonnell (2002). "On the dialog between experimentalist and modeler in catchment hydrology: use of soft data for multicriteria model calibration." Water Resources Research 38(11): 1241, doi:1210.1029/2001WR000978.
Siepel, A. C., et al. (2002). "A simplified hillslope erosion model with vegetation elements for practical applications." Journal of Hydrology 258(1-4): 111-121.
Soil and water conservation practices are increasingly being considered for curbing non-point source pollution from agricultural land. Several studies have demonstrated that stream power is a simple and good predictor of soil detachment and transport and can be used to predict the effect of soil and water conservation practices on soil loss. Our objective was, therefore, to develop a simple water erosion simulation model that is physically based on stream power, handles vegetation in terms of contact cover, and considers the settling velocity characteristics of the eroding sediment. The model assumes that rill flow can occur on hillslope segments with net erosion, but on segments with net deposition sheet flow is assumed. Input parameters include the depositability of the soil, rill shape, rill density, net precipitation, and an empirical power function describing the decrease of sediment concentration with vegetative cover increase. The model was evaluated by comparison of predicted and observed relationships between sediment concentration, slope, and vegetative residue cover in two experimental studies using simulated rainfall: one that involved erosion plots with various uniform slopes and levels of vegetative cover, and another that involved the observation of soil movement on mechanically shaped concave, uniform, and convex slopes with negligible vegetation. Without calibration, the model appeared to represent soil erosion relationships observed in these studies and is simple enough to be included in grid-based variable source hydrology models. (C) 2002 Elsevier Science B.V. All rights reserved.
 
Sierra Pacific Industries (2003). Erosion Study: Judd Creek Basin, Southern Cascades, California. Redding, CA, By Lee Benda, Paul Bigelow, Kevin Andras; Lee Benda and Associates, Inc. Mt. Shasta, CA: 62.
Sierra Pacific Industries (2003). Terrain Mapping and Analysis: Supporting Timber Harvest Planning and Managing Research and Monitoring Programs. Redding, CA, By Benda, L., Andras, K., Bigelow, P, Lee Benda and Associates, Mt. Shasta, CA: 78.
Sierra Pacific Industries (2003). Wood Recruitment to Streams: Cascades and Klamath Mountains, Northern California. Redding, CA, By Lee Benda, Kevin Andras, Paul Bigelow, Lee Benda and Associates, Inc., Mt. Shasta, CA: 90.
Sigafoos, R. S. (1964). Botantical evidence of floods and flood-plain deposition, United States Geological Survey.
Sigler, J. W., et al. (1984). "Effects of Chronic Turbidity on Density and Growth of Steelheads and Coho Salmon " Transactions of the American Fisheries Society 113: 142-150.
Simberloff, D. (1998). "Flagships, umbrellas, and keystones: is single-species management pass'e in the landscape era? ." Biological Conservation 83(3): 247-257.
Simenstad, C. A., et al. (2002). Landscape structure and scale constraints on restoring estuarine wetlands. Concepts and Controversies in Tidal Marsh Ecology. M. P. Weinstein and D. A. Kreeger. New York, Kluwer Academic Publishers: 595-624.
Simon, A. (1989). "The discharge of sediment in channelized alluvial streams." Water Resources Bulletin 25(6): 1177-1188.
Simon, A. (1989). "A Model of Channel Response in Disturbed Alluvial Channels." Earth Surface Processes and Landforms 14: 11-26.
Simon, A. (1989). Sehar-strength determination and stream-bank instability in loess-derived alluvium, West Tennessee, USA. Applied Quaternary Research, Ottowa, Balkema, Rotterdam.
Simon, A., et al. (2007). "Critical evaluation of how the Rosgen classification and associated "natural channel design" methods fail to integrate and quantify fluvial processes and channel response." Journal of the American Water Resources Association 43(5): 1117-1131.
Over the past 10 years the Rosgen classification system and its associated methods of ‘‘natural
channel design’’ have become synonymous to some with the term ‘‘stream restoration’’ and the science of fluvial
geomorphology. Since the mid 1990s, this classification approach has become widely adopted by governmental
agencies, particularly those funding restoration projects. The purposes of this article are to present a critical
review, highlight inconsistencies and identify technical problems of Rosgen’s ‘‘natural channel design’’ approach
to stream restoration. This paper’s primary thesis is that alluvial streams are open systems that adjust to
altered inputs of energy and materials, and that a form-based system largely ignores this critical component.
Problems with the use of the classification are encountered with identifying bankfull dimensions, particularly in
incising channels and with the mixing of bed and bank sediment into a single population. Its use for engineering
design and restoration may be flawed by ignoring some processes governed by force and resistance, and the
imbalance between sediment supply and transporting power in unstable systems. An example of how C5 channels
composed of different bank sediments adjust differently and to different equilibrium morphologies in
response to an identical disturbance is shown. This contradicts the fundamental underpinning of ‘‘natural channel
design’’ and the ‘‘reference-reach approach.’’ The Rosgen classification is probably best applied as a communication
tool to describe channel form but, in combination with ‘‘natural channel design’’ techniques, are not
diagnostic of how to mitigate channel instability or predict equilibrium morphologies. For this, physically based,
mechanistic approaches that rely on quantifying the driving and resisting forces that control active processes
and ultimate channel morphology are better suited as the physics of erosion, transport, and deposition are the
same regardless of the hydro-physiographic province or stream type because of the uniformity of physical laws.
 
Simon, A. and C. R. Hupp (1987). Geomorphic and vegetative recovery processes along modified Tennessee streams: an interdisciplinary approach to disturbed fluvial systems. Vancouver Symposium: Forest Hydrology and Watershed Management, Building, Nashville.
Simon, A., et al. (1990). "The role of soil processes in determining mechanisms of slope failure and hillslope development in a humid-tropical forest, eastern Puerto Rico." Geomorphology 3: 263-286.
Sinclair, K. A. and C. F. Pitz (1999). Estimated Baseflow Characteristics of Selected Washington Rivers and Streams. Olympia, Washington State Department of Ecology.
Singer, M. B. and T. Dunne (2004). "Modeling decadal bed material sediment flux based on stochastic hydrology." Water Resources Research 40.
Estimates of decadal bed material sediment flux and net storage are derived by driving sediment transport calculations with a stochastic hydrology model. The resulting estimates represent the whole distribution of sediment flux based on natural variability in channel characteristics (gradient, width, and bed grain size) and the magnitude, duration, and interarrival time of flood events. A procedure for calibrating a fractional sediment transport equation of a commonly used form to bed material grain size distributions (BMGSDs) at cross sections is presented. The procedure was applied to the Sacramento River channel network to compute estimates of annual total and annual peak bed material discharges into and through the main stem over a 30-year period. Main stem bed material budgets were evaluated to identify reaches in states of net accumulation or scour. Simulations highlight large imbalances in sand and gravel storage throughout the Sacramento River, which can be explained by a combination of local hydraulics and BMGSDs and for which there is at least some empirical support.
 
Singh, H. (1981). Remote sensing for detecting riverbed sediments and computer mapping the time trends in the south fork of the Salmon River, University of Idaho: 9-59.
Singh, V. P. (2002). "Is hydrology kinematic." Hydrologic Processes 16: 667-716.
Singh, V. P., and D.A. Woolhiser (2002). "Mathematical modeling of watershed hydrology." Journal of Hydrological Engineering 7(4): 270-292.
Singh, V. P. (2003). "On the theories of hydraulic geometry." International Journal of Sediment Research 18(3): 196-218.
Sinowski, W., and K. Auerswald (1999). "Using relief parameters in a discriminant analysis to stratify geological areas with different spatial variability of soil parameters." Geoderma 89: 113-128.
Sinton, D. S. (1992). The Use of Ecological Principles for Riparian Zone Managment in Four Forest Planning Documents. Corvallis, Oregon, Oregon State University, Department of Geosciences: 57.
Sinton, D. S., et al. (2000). "Windthrow disturbance, forest composition, and structure in the Bull Run Basin, Oregon." Ecology 81(9): 2539-2556.
Siry, J., and F. Cubbage (2002). Clearcutting in the South: issues, status, and trends. Proceedings of the Fourth Annual Forest Inventory and Analysis Symposium.
Sitar, N., et al. Conditions for initiation of rainfall-induced debris flows. Slopes and Embankments: 834-849.
Sivakumar, B. (2003). "Forecasting monthly streamflow dynamics in the western United States: a nonlinear dynamical approach." Environmental Modelling & Software 18(8-9): 721-728.
Sivakumar, B. and A. W. Jayawardena (2003). "Sediment transport phenomenon in rivers: an alternative perspective." Environmental Modelling & Software 18(8-9): 831-838.
Sivapalan M, G. B., L Zhang, and R Vertessy (2003). "Downward approach to hydrological prediction." Hydrologic Processes 17: 2101-2111.
Skaugset, A. (?). Managing the risk of forest road-related landslides: the effects of road layout, construction and maintenance. ?, Oregon State University, Forest Engineering department: ?
Skinner, Q. (2003). "Rangeland monitoring: Water quality and riparian systems." Arid Land Research and Management 17(4): 407-428.
Ecological concepts serve as a foundation for developing a monitoring program to evaluate water quality and associated riparian systems. Ecological concepts used for developing a monitoring plan must be supported by scientific literature and related to streamflow dynamics and channel interactions. These interactions help determine natural or background habitat quality within and along river longitudinal and environmental gradients from mountains through basins in the western United States. In addition stream size, position in the watershed, and flow are related to sediment sorting, channel bank strength, and channel configuration. These relationships determine channel substrate habitat for aquatic organisms and population diversity. These habitat features may be modified by a channel's ability to store and transport sediment and associated pollutants within a watershed's drainage pattern. Sediment supply, delivery, and timing are altered by differences in snowmelt along elevation gradients, runoff from convective storms, water development history, and stream channel succession. Potential impairment of reference or background aquatic habitat in the western United States is generally sediment related and should be greater in basin river segments and during base flow conditions. Impairment sources can be shown to originate in the steep and first order tributaries of foothill and basin watersheds, and not from valley slopes where supply must cross established riparian zones. Water column, substrate disturbance, and channel bank disturbances may alter amount of sediment and bacteria pollution measured in basins and during base flow conditions.
 
Sklar, L. S. and W. E. Dietrich (2001). "Sediment and rock strength controls on river incision into bedrock." Geological Society of America 29(12): 1087-1090.
Sklar, L. S., et al. (2006). "Do gravel bed river size distributions record channel network structure?" Water Resources Research 42: W06D18.
Bed load sediment particles supplied to channels by hillslopes are reduced in size by
abrasion during downstream transport. The branching structure of the channel network
creates a distribution of downstream travel distances to a given reach of river and
thus may strongly influence the grain size distribution of the long-term bed load flux
through that reach. Here we investigate this hypothesis, using mass conservation and the
Sternberg exponential decay equation for particle abrasion, to predict bed material
variability at multiple scales for both natural and artificial drainage networks. We assume
that over a sufficiently long timescale, no net deposition occurs and that grains less than
2 mm are swept away in suspension. We find that abrasion during fluvial transport
has a surprisingly small effect on the bed load sediment grain size distribution, for the
simple case of spatially uniform supply of poorly sorted hillslope sediments. This occurs
because at any point in the channel network, local resupply offsets the size reduction
of material transported from upstream. Thus river bed material may essentially mirror the
coarse component of the size distribution of hillslope sediment supply. Furthermore,
there is a predictable distance downstream at which the bed load grain size distribution
reaches a steady state. In the absence of net deposition due to selective transport, large-scale
variability in bed material, such as downstream fining, must then be due primarily
to spatial gradients in hillslope sediment production and transport characteristics. A
second key finding is that average bed load flux will tend to stabilize at a constant value,
independent of upstream drainage area, once the rate of silt production by bed load
abrasion per unit travel distance is equal to the rate of coarse sediment supply per unit
channel length (q). Bed load flux equilibrates over a distance that scales with the
inverse of the fining coefficient in the abrasion rate law (摯瑬敳獩 ) and can be approximated
simply as q/3摯瑬敳獩 . Thus the efficiency of particle abrasion sets a fundamental length
scale, shorter for weaker rocks and longer for harder rocks, which controls the expression
in the river bed of variability in sediment supply. We explore the role of the abrasion
length scale in modulating the influence of sediment supply variability in a number of
channel network contexts, including individual tributary junctions, a sequence of
tributary inputs along a main stem channel, and variable basin shapes and network
architecture as expressed by the width function. These findings highlight the need
for both data and theory that can be used to predict the grain size distributions supplied to
channels by hillslopes.
 
Slattery, M. C. and R. B. Bryan (1992). "Hydraulic Conditions for Rill Incision under Simulated Rainfall - a Laboratory Experiment." Earth Surface Processes and Landforms 17(2): 127-146.
A series of controlled laboratory experiments were conducted in order to obtain precise data on the hydraulic and sediment transport conditions during rill formation. Tests were carried out using a crusting-prone binary mixed soil in a 15 m long flume at an average slope of 0.087 under simulated rainfall. Rainfall intensities varied from 30-35 mm h-1 and developed about 70 per cent of the kinetic energy of natural rainfall of similar intensity.Runoff and sediment discharge measured at the downstream weir were strongly influenced by rill forming processes. Essentially, rill incision reduced runoff discharge as a result of increased percolation in rill channels but increased sediment discharge. Secondary entrainment processes, such as bank collapse, also increased sediment discharge at the weir. Knickpoint bifurcation and colluvial deposition, however, decreased sediment discharge. Rills always developed through the formation of a knickpoint. The critical condition for knickpoint initiation was the development of supercritical flow and waves which mould and incise the bed. Prior smoothing of the soil surface by entrainment and redistribution of sediment facilitated supercritical flow. Statistical analysis showed that hydraulic and sediment transport conditions differed significantly in rilled and unrilled flows. The relationship between sediment discharge, rill erosion, and flow hydraulics was found to be nonlinear, conforming to a standard power function in the form y = ax(b). Rills were also associated with significantly increased sediment transport capacities. However, rill initiation was not clearly defined by any specific hydraulic threshold. Instead, rilled and unrilled flows were separated by zones of transition within which both types of flow occur.
 
Slattery, M. C. and T. P. Burt (1998). "Particle size characteristics of suspended sediment in hillslope runoff and stream flow." Earth Surface Processes and Landforms 72(8): 705-719.
This study examines the particle size characteristics of hillslope soils and fluvial suspended sediments in an agricultural catchment. Samples of surface runoff and stream flow were collected periodically and analysed for the size distributions of the effective (undispersed) sediment. This sediment was subsequently dispersed and the ultimate size distributions determined. The median effective particle size of stream suspended sediment was considerably coarser than the median ultimate particle size, indicating that most of the load included a substantial proportion of aggregates. Moreover, the proportion of fine material (i.e. silt and clay) increased, and the proportion of sand-sized material decreased, with increasing discharge. This decrease in sediment size with increased flow, which is contrary to the traditional assumption of a positive discharge/particle size relationship, is thought to reflect: (i) the influx of silt and clay, predominantly the former, originating on the catchment slopes and brought to the stream by overland flow along vehicle wheelings, roads and tracks; and (ii) erosion of fine material from the channel bed and banks. During large storms, however, the proportion of sand-sized sediment increased during the rising limb of the hydrograph, as a result of the entrainment of coarser source material from the valley floor during overbank flooding. The stream suspended sediment was finer than the catchment soils and considerably finer than material eroding from the catchment slopes during storms. The degree of clay and silt enrichment in the suspended sediments was largely the result of preferential deposition of the coarser fraction during the transport and delivery of sediment from its source to basin outlet. The data from this study confirm that a significant mode of sediment transport in fluvial systems is in the form of aggregates, and that the dispersed sediment size distribution is inappropriate for determining the transportability of sediment by flow.
 
Slaymaker, O. (2000). "Assessment of the geomorphic impacts of forestry in British Columbia." Ambio 29(7): 381-387.
Slaymaker, O. and M. Sonesson (2000). "Assessment of the geomorphic impacts of forestry in British Columbia." Ambio 29(7): 381-387.
Timber harvesting in British Columbia influences (a) forest hydrology; (b) fluvial geomorphology; (c) terrain stability; and (d) integrated watershed behavior. Impacts on forest hydrology are well understood and include increased average runoff, total water yield, increased storm runoff and advances in timing of floods. Stream channels and valley floors are impacted differently by fine sediment, coarse sediment and large woody debris transport. Terrain stability is influenced through gully and mass movement processes that are accelerated by timber harvesting. Impacts on integrated watershed behavior are assessed through disturbed sediment budgets and lake sediments. The Forest Practices Code (1995) is a significant step towards sustainable management of the land in so far as it attempts to minimize these geomorphic impacts of forest in B.C.
 
Smakhtin, V. U. (2001). "Low flow hydrology: a review." Journal of Hydrology 240: 147-186.
The paper intends to review the current status of low-flow hydrology — a discipline which deals with minimum flow in a
river during the dry periods of the year. The discussion starts with the analysis of low-flow generating mechanisms operating in
natural conditions and the description of anthropogenic factors which directly or indirectly affect low flows. This is followed by
the review of existing methods of low-flow estimation from streamflow time-series, which include flow duration curves,
frequency analysis of extreme low-flow events and continuous low-flow intervals, baseflow separation and characterisation
of streamflow recessions. The paper describes the variety of low-flow characteristics (indices) and their applications. A separate
section illustrates the relationships between low-flow characteristics. The paper further focuses on the techniques for low-flow
estimation in ungauged river catchments, which include a regional regression approach, graphical representation of low-flow
characteristics, construction of regional curves for low-flow prediction and application of time-series simulation methods. The
paper presents a summary of recent international low-flow related research initiatives. Specific applications of low-flow data in
river ecology studies and environmental flow management as well as the problem of changing minimum river flows as the result
of climate variability are also discussed. The review is largely based on the research results reported during the last twenty
years.
 
Small, E. E., et al. (1999). "Estimates of the rate of regolith production using 10Be and 26Al from an alpine hillslope." Geomorphology 27: 131-150.
Small, R. J. (1973). "Braiding Terraces in the Val D'Herens, Switzerland." Geography 58: 129-135.
Smith, D. G. (1973). Aggradation of the Alexandra-Saskatchewan River, Banff Park, Alberta. Fluvial Geomorphology. M. Morisawa. Binghamton Symposia in Geomorphology International Series 4:314: 201 - 209.
Smith, G. A. (?). "Coarse-grained nonmarine volcaniclastic sediment: Terminology and depositional process." Bulletin of the Geological Society of America 97: 1-10.
Smith, L. C., et al. (1998). "Stream flow characterization and feature detection using a discrete wavelet transform." Hydrological Processes 12: 233-249.
Smith, R. A., G. E. Schwarz, and R.B. Alexander (1997). "Regional interpretation of water-quality monitoring data." Water Resources Research 33(12): 2781-2798.
Smith, R. D. (1993). "Effects of experimental removal of woody debris on the channel morphology of a forest, gravel-bed stream." Journal of Hydrology 152: 153-178.
Smith, R. D., et al. (1993). "Effects on bedload transport of experimental removal of woody debris from a forest gravel-bed stream." Earth Surface Processes and Landforms 18: 455-468.
Smith, R. E., et al. (1999). "Simulation of selected events on the Catsop catchment by KINEROS2: A report for the GCTE conference on catchment scale erosion models."
Data from the Catsop catchment in South Limburg, Netherlands was simulated with the model KINEROS2. The results of calibration and validation on a split set of runoff and sediment data are reported and the variations in apparent parameters are analyzed. Calibration was performed with regard to the temporal distribution of runoff and sediment rather than single values such as total or peak rates. Based on the simulations, soil erodibility was considerably higher in 1993 than earlier years. Sediment discharge is quite sensitive to hydrologic simulation, as the amount and velocity of runoff affects sediment transport capacity which in turn determines the delivery of soil disturbed by rainsplash. Overall ability of the model to reproduce the measured data was considered relatively good.
 
Smith, R. K. (1999). Differential stability of spawning microhabitats of warmwater stream fishes Fisheries and Wildlife Sciences. Blacksburg, Virginia, Virginia Polytechnic Institute and State University.
Smith, S. M. and M. J. Brady (1997). "SUSAN - A new approach to low level image processing." International Journal of Computer Vision 23(1): 45-78.
This paper describes a new approach to low level image processing; in particular, edge and corner
detection and structure preserving noise reduction.
Non-linear filtering is used to define which parts of the image are closely related to each individual pixel; each
pixel has associated with it a local image region which is of similar brightness to that pixel. The new feature
detectors are based on the minimization of this local image region, and the noise reduction method uses this region
as the smoothing neighbourhood. The resulting methods are accurate, noise resistant and fast.
Details of the new feature detectors and of the new noise reduction method are described, along with test results
 
Snavely, P. D., Jr., et al. (1980). Makah Formation -- A Deep-Marginal-Basin Sequence of Late Eocene and Oligocene Age in the Northwestern Olympic Peninsula, Washington, Department of National Resources: 28.
Sobocinski, R. W., et al. (1990). "Sediment transport in a small stream based on 137Cs inventories of the bed load fraction." WRR 26(6): 1177-1187.
We report on sediment transport in White Oak Creek, a small third-order stream in eastern Tennessee, whose sediments have been contaminated by the irreversible adsorption of 137Cs. Sediment cores taken downstream from a point contamination source record te recent history of sedimentation and contamination in the creek.Sediment size distinguishes between different sidimentation events. The contamination profile developed since the last sedimentation event is used to estimate the time of the most recent sedimentation event which agrees with higher-than-normal periods of discharge measured by weirs on the creek. It is likely that at least seven major bed load ttransport events occurred diring 1985 and 1986. Total inventories of 137Cs in the bed load fraction were measured for six different size fractions for cores taken downstream from the point contamination source. The 137Cs concentration at any point in the creek is due to contaminated sediment transported to that location and subsequent contamination by 137Cs adsorbed directly from streamwater. The irreversible nature of 137Cs adsorption onto the sediments of White Oak Creek means that 137Cs can be used as a particle tracer in this system. Sediment transport in White Oak Creek was modeled by using a general transport model that has terms for dispersion, velocity, adsorption, and radioactive decay. For White Oak Creek, maximum velocities of 265, 215, 170, and 160 m/yr are estimated for the 1-2, 2-4, 4-8, and 8-16 mm size fractions, respectively.
 
Sobota, D. J. (2003). Fall directions and Breakage of Riparian Trees along Streams in the Pacific Northwest. Department of Fisheries and Wildlife. Corvallis, OR, Oregon State University. Masters of Science: 126.
Sobota, D. J., et al. (2006). "Riparian tree fall directionality and modeling large wood recruitment to streams." Canadian Journal of Forest Research 36: 1243-1254.
Directionality of tree fall in riparian forests can strongly influence predictions of large wood recruitment to
streams, yet accuracy of this model parameter has rarely been assessed with field data. We measured fall directions of
1202 riparian trees distributed among 21 stream sites across the Pacific Northwest, USA. Fall directions were oriented
towards the stream at 16 sites, upstream at four sites, and not distinguishable from random at one site. Average tree
fall direction across sites was correlated with valley constraint (Spearman r = –0.53; p = 0.02), but variability of fall
directions was not correlated with this variable. When grouped by species (six conifers and one deciduous), individual
trees exhibited stronger tendency to have fallen towards the channel on steep hillslopes (>40%) than on moderately
sloped landforms (<40%). Integration of field data into an established recruitment model indicated that 1.5 to 2.4 times
more large wood (by number of tree boles) would be recruited to stream reaches with steep hillslopes than to reaches
with moderate side slopes or flat banks, if riparian forest conditions are assumed to be constant. We conclude that
stream valley topography should be considered in models that use tree fall directions in predictions of large wood re-cruitment
to streams.
 
Soeters, R. and C. J. van Westen (1996). Slope instability recognition, analysis, and zonation. Landslides Investigation and Mitigation Special, Transportation Research Board Special Report 247. A. K. Turner and R. L. Schuster. Washington, D.C., National Academy Press: 129-177.
Soille, P., et al. (2003). "Carving and adaptive drainage enforcement of grid digital elevation models." Water Resources Research 39(12).
An effective and widely used method for removing spurious pits in digital elevation
models consists of filling them until they overflow. However, this method sometimes
creates large flat regions which in turn pose a problem for the determination of accurate
flow directions. In this study, we propose to suppress each pit by creating a descending
path from it to the nearest point having a lower elevation value. This is achieved by
carving, i.e., lowering, the terrain elevations along the detected path. Carving paths are
identified through a flooding simulation starting from the river outlets. The proposed
approach allows for adaptive drainage enforcement whereby river networks coming from
other data sources are imposed to the digital elevation model only in places where the
automatic river network extraction deviates substantially from the known networks. An
improvement to methods for routing flow over flat regions is also introduced. Detailed
results are presented over test areas of the Danube basin.
 
Soiseth, C. and A. M. Milner (1995). Predicting salmonid occurrence from physical characteristics of streams in Glacier Bay National Park and Preserve. Proceedings of the Third Annual Glacier Bay Science Symposium, National Park Service: Atlanta, Georgia.
Sommerfield, C. K. N., C. A. (1999). "Modern accumulation rates and a sediment budget for the Eel shelf: a flood dominated depositional environment." Marine Geology 154: 227-241.
Sorriso-Valvo, M. (1984). "Sackungen" and deep-seated rockslide/debris flows: evolution model. CNR-PAN Meeting on Progress in Mass Movement and Sediment Transport Studies: Problems in Recognition and Prediction, Torino.
Sowden, T. K. and G. Power (1985). "Prediction of Rainbow Trouth Embryo Survival in Relation to Groundwater Seepage and Particle Size of Spawning Substrates." Transactions of the American Fisheries Society 114: 804-812.
Sparks, R. E., et al. (1990). "Disturbance and recovery of large floodplain rivers." Environmental Management 14(5): 699-709.
Spence, B. C., et al. (1996). An Ecosystem Approach to Salmonid Conservation, Management Technology: 356.
Spencer, C. N. and F. R. Hauer (1991). "Phosphorous and nitrogen dynamics in streams during a wildfire." J. N. Am. Benthol. Soc. 10(1): 24-30.
Spicer, R. C. (1989). "Recent variations of Blue Glacier, Olympic Mountains, Washington, U.S.A." Arctic and Alpine Research 21(1): 1-21.
Spies, T. A., et al. (1988). "Coarse woody debris in Douglas-fir forests of western Oregon and Washington." Ecology 69(6): 1689-1702.
Spies, T. A. and K. N. Johnson (2007). "Projecting forest policy and management effects across ownerships in coastal Oregon." Ecological Applications 17(1): 3-4.
Spies, T. A., et al. (in press). "Assessing forest policies in the coastal province of Oregon: an overview of biophysical and socio-economic responses." Ecological Applications.
Spies, T. A., et al. Effects of Regional Landscape Dynamics and Landowner Activities on Watershed Condition and Biodiversity in the Oregon Coast Range, Oregon State University, University of Oregon: A-1 - D-11.
Spracklen, D. V., L. J. Mickley, J. A. Logan, R. C. Hudman, R. Yevich, M. D. Flannigan, and A. L. Westerling ((submitted)). "Impacts of climate change from 2000 to 2050 on wildfire activity and carbonaceous aerosol concentrations in the western United States." Geophysical Research Letters.
Sprugel, D. G. (1991). "Disturbance, equilibrium, and environmental variability: What is 'natural' vegetation in a changing environment?" Biological Conservation 58: 1-18.
 
(SREP) S. R. E. P. (2004). The State of the southern Rockies Ecoregion: A Report by the Southern Rockies Ecosystem Project. G. Colorado Mountain Club Press, CO: 138.
Staley, D. M., T.A. Wasklewicz, J.S. Blaszczynski (2006). "Surficial patterns of debris flow deposition on alluvial fans in Death Valley, CA using airborne laser swath mapping data." Geomorphology 74: 152-163.
Stallman, J. (1998). "Late Cenozoic Tectonism in the Battle Creek Structural Domain."?: 1-9.
Standford, J. A. and J. V. Ward (1988). "The hyporheic habitat of river ecosystems." Nature 335: 64-66.
Standford, J. A. and J. V. Ward (1993). "An ecosystem perspective of alluvial rivers: connectivity and the hyporheic corridor." J. N. Am. Benthl. Soc. 12: 48-62.
Stanford, J. A. and J. V. Ward (1991). Management of Aquatic Resources in Large Catchments: Recognizing Interactions Between Ecosystem Connectivity and Environmental Disturbance (draft). Watershed Management. N. R. J., Springer-Verlag.
Stanford, J. A. and J. V. Ward (1993). "An ecosystem perspective of alluvial rivers: Connectivity and the hyporheic corridor." Journal of the North American Benthological Society 12(1): 48-60.
Floodplains of large alluvial rivers are often expansive and characterized by high volume hyporheic flow through lattice-like substrata, probably formed by glacial outwash or lateral migration of the river channel over long time periods. River water downwells into the floodplain at the upstream end; and, depending on bedrock geomorphology and other factors, groundwater from the unconfined aquifer upwells directly into the channel or into floodplain springbrooks at rates determined by head pressure of the water mass moving through the floodplain hydrologic system. These large scale (km super(3)) hyporheic zones contain speciose food webs, including specialized insects with hypogean and epigean life history stages (amphibionts) and obligate groundwater species (stygobionts). Biogeochemical processes in the hyporheic zone may naturally load groundwaters with bioavailable solutes that appear to exert proximal controls on production and biodiversity of surface benthos and riparian vegetation.
 
Stark, C. P. and H. Hovius (2001). "The characterization of landslide size distributions." Geophysical Research Letters 28(6): 1091-1094.
Stark, P. L. (1989). A Cartiobibliography of Separately Published U.S. Geological Survey Special Maps and River Surveys, Western Association of Map Libraries.
Starkel, L. (?). Long-term and Short-term Rhythmicity in Terrestrial Landforms and Deposits. ? chapter 18: 323-332.
Statham, I. (1976). "Debris flows on vegetated screes in the Black Mountain, Carmarthenshire." Earth Surface Processes 1: 173-180.
Statzner, B. and B. Higler (1985). "Questions and comments on the river continuum concept." Canadian Journal of Fisheries and Aquatic Sciences 42(5): 1038-1044.
Stauffer, J. C., et al. Relationship of wooded riparian zones and runoff potential to fish community composition in agricultural streams. Canadian Journal of Fisheries and Aquatic Sciences 57 (2), 307-316 pp.;   35 ref.; 2000.
The relationship of fish community composition to riparian cover and runoff potential was investigated in 20 streams in the agricultural Minnesota River Basin in the USA during summer 1997. Analysis of variance indicated significant differences in fish community composition due to both riparian cover (wooded versus open) and runoff potential (high or low). Sites were classified as wooded if the amount of wooded riparian zone was >28% (the range on the study sites was 28-58%), and as open if the amount of wood was <10% (the rest of the area was row crops, pasture or native grasses); cover was determined from GIS data updated with aerial photographic data, and was confirmed on-site. The runoff classification factor was based on soil characteristics of the watersheds from GIS and soil characterization data. Streams with wooded riparian zones had fish communities with a higher index of biological integrity (IBI) scores, species richness, diversity and percentages of benthic insectivores and herbivores than streams with open riparian zones. Streams with low runoff potential had higher IBI scores and species richness than streams with high runoff potential. Fish were sampled in a representative reach of each stream by electrofishing. The riparian cover and runoff potential interaction was marginally significant with respect to IBI scores and species richness, suggesting a weak interaction between the two factors. Although both factors were important, riparian cover influenced fish community composition more than runoff potential in these streams, indicating that local factors (close to the stream) dominated landscape- or basin-level factors.
 
Stednick, J. D. (1996). "Monitoring the effects of timber harvest on annual water yield." Journal of Hydrology 176: 79-95.
Steel, E. A., et al. (2004). "Landscape models to understand steelhead (Oncorhynchusmykiss) distribution and help prioritize barrier removals in the Willamette basin, Oregon, USA." Canadian Journal of Fisheries and Aquatic Science 61: 999-1011.
We use linear mixed models to predict winter steelhead (Oncorhynchus mykiss) redd density from geology,
land use, and climate variables in the Willamette River basin, Oregon. Landscape variables included in the set of best
models were alluvium, hillslope < 6%, landslide-derived geology, young (<40 years) forest, shrub vegetation, agricul-tural
land use, and mafic volcanic geology. Our approach enables us to model the temporal correlation between annual
redd counts at the same site while extracting patterns of relative redd density across sites that are consistent even
among years with varying strengths of steelhead returns. We use our model to predict redd density (redds per kilo-metre)
upstream of 111 probable migration barriers as well as the 95% confidence interval around the redd density
prediction and the total number of potential redds behind each barrier. Using a metric that incorporates uncertainty, we
identified high-priority barriers that might have been overlooked using only stream length or mean predicted fish bene-fit
and we clearly differentiated between otherwise similar barriers. We show that landscape features can be used to de-scribe
and predict the distribution of winter steelhead redds and that these models can be used immediately to improve
decision-making for anadromous salmonids.
 
Steel, E. A., et al. (2008). "A spatially explicit decision support system for watershed-scale management of salmon." Ecology and Society 13(2): 50.
Effective management for wide-ranging species must be conducted over vast spatial extents,
such as whole watersheds and regions. Managers and decision makers must often consider results of multiple
quantitative and qualitative models in developing these large-scale multispecies management strategies.
We present a scenario-based decision support system to evaluate watershed-scale management plans for
multiple species of Pacific salmon in the Lewis River watershed in southwestern Washington, USA. We
identified six aquatic restoration management strategies either described in the literature or in common use
for watershed recovery planning. For each of the six strategies, actions were identified and their effect on
the landscape was estimated. In this way, we created six potential future landscapes, each estimating how
the watershed might look under one of the management strategies. We controlled for cost across the six
modeled strategies by creating simple economic estimates of the cost of each restoration or protection
action and fixing the total allowable cost under each strategy. We then applied a suite of evaluation models
to estimate watershed function and habitat condition and to predict biological response to those habitat
conditions. The concurrent use of many types of models and our spatially explicit approach enables analysis
of the trade-offs among various types of habitat improvements and also among improvements in different
areas within the watershed. We report predictions of the quantity, quality, and distribution of aquatic habitat
as well as predictions for multiple species of species-specific habitat capacity and survival rates that might
result from each of the six management strategies. We use our results to develop four on-the-ground
watershed management strategies given alternative social constraints and manager profiles. Our approach
provides technical guidance in the study watershed by predicting future impacts of potential strategies,
guidance on strategy selection in other watersheds where such detailed analyses have not been completed,
and a framework for organizing information and modeled predictions to best manage wide-ranging species.
 
 
Steel, E. A., et al. (2003). Managing uncertainty in habitat recovery planning. Ecosystem recovery planning for listed salmon: assessment approaches for salmon habitat. U.S. Dept. of Commerce, NOAA Tech. Memo NMFS-NWFSC-58. T. J. Beechie, P. Roni and E. A. Steel: 74-89.
Steel, E. A., et al. (2003). Habitat Analyses for the Lewis River Case Study: Interim Report. Seattle, NW Fisheries Science Center, NOAA Fisheries.
Steen, R. P. and T. P. Quinn (1999). "Egg burial depth by sockeye salmon (Oncorhynchus nerka): implications for survival of embryos and natural selection on female body size " Canadian Journal of Zoology 77: 836-841.
Stefferud, J. A. (1993). "Spawning season and microhabitat use by California Golden Trout (Oncorhynchus Mykiss Aguabonita) in the Southern Sierra Nevada " California Fish and Game 79: 133-144.
Steinblums, I. J. and J. K. Lyons (1984). "Designing Stable Buffer Strips for Stream Protection." Journal of Forestry 82(1): 49-52.
Steinbrecher, G. and W. T. Shaw (2007). "Quantile mechanics." European Journal of Applied Mathematics 19: 87-112.
In both modern stochastic analysis and more traditional probability and statistics, one way of characterizing a static or dynamic probability distribution is through its quantile function. This paper is focused on obtaining a direct understanding of this function via the classical approach of establishing and then solving differential equations for the function. We establish ordinary differential equations and power series for the quantile functions of several common distributions. We then develop the partial differential equation for the evolution of the quantile function associated with the solution of a class of stochastic differential equations, by a transformation of the Fokker–Planck equation. We are able to utilize the static formulation to provide elementary time-dependent and equilibrium solutions.
 
Such a direct understanding is important because quantile functions find important uses in the simulation of physical and financial systems. The simplest way of simulating any non-uniform random variable is by applying its quantile function to uniform deviates. Modern methods of Monte–Carlo simulation, techniques based on low-discrepancy sequences and copula methods all call for the use of quantile functions of marginal distributions. We provide web resources for prototype implementations in computer code. These implementations may variously be used directly in live sampling models or in a high-precision benchmarking mode for developing fast rational approximations also for use in simulation.
 
Stephens, R. (1984). Procedure to Estimate Sediment Yields for Watersheds on the Olympic National Forest, Olympic National Forest.
Stevens, D. L. (2002). Sampling design and statistical analysis methods for the integrated biological and physical monitoring of Oregon streams. Corvallis, OR, USA, Oregon Department of Fish and Wildlife, Corvallis Research Lab.
Stevens, D. L. and A. R. Olsen (2004). "Spatially balanced sampling of natural resources." Journal of the American Statistical Association 99(465).
The spatial distribution of a natural resource is an important consideration in designing an ef. cient survey or monitoring program for the
resource. Generally, sample sites that are spatially balanced, that is, more or less evenly dispersed over the extent of the resource, are more
ef. cient than simple random sampling.We review a uni. ed strategy for selecting spatially balanced probability samples of natural resources.
The technique is based on creating a function that maps two-dimensional space into one-dimensional space, thereby de. ning an ordered
spatial address.We use a restricted randomization to randomly order the addresses, so that systematic sampling along the randomly ordered
linear structure results in a spatially well-balanced random sample. Variable inclusion probability, proportional to an arbitrary positive
ancillary variable, is easily accommodated. The basic technique selects points in a two-dimensional continuum, but is also applicable to
sampling . nite populations or one-dimensional continua embedded in two-dimensional space. An extension of the basic technique gives
a way to order the sample points so that any set of consecutively numbered points is in itself a spatially well-balanced sample. This latter
property is extremely useful in adjusting the sample for the frame imperfections common in environmental sampling.
 
Stewart, J. H. and V. C. La Marche, Jr. (1967). Erosion and depostion produced by the flood of December 1964 on Coffee Creek, Trinity County, California, US Geological Survey: 22.
Stillwater Sciences (1999). South Fork Eel TMDL: Sediment Source Analysis Berkeley, California, prepared by Stillwater Sciences for the U.S. Environmental Protection Agency: 64.
Stock, G. M., et al. (2006). "Where does sediment come from? Quantifying catchment erosion with detrital apatite (U-Th)/He thermochronometry." Geology 34(9): 725-728.
We present a new method for tracing sediment using detrital apatite (U-Th)/He (AHe)
thermochronometry, and use this to quantify the spatial distribution of catchment erosion
in the eastern Sierra Nevada, California. Well-developed age-elevation relationships permit
detrital AHe ages to track the elevations where sediment grains were shed from bedrock.
We analyzed sediment exiting nonglaciated Inyo Creek and adjacent (formerly)
glaciated Lone Pine Creek. Statistical comparison of measured AHe age probability density
functions (PDFs) with predicted PDFs based on catchment hypsometries suggests that
Inyo Creek is eroding uniformly, consistent with field observations of weathered hillslopes
tightly coupled to the fluvial system. In contrast, significant mismatch between measured
and predicted PDFs from Lone Pine Creek reveals that sediment derives primarily from
the lower half of the catchment. The dearth of older ages is likely due to sediment storage
in cirques and moraines and/or focused erosion at intermediate elevations, both potential
consequences of glacial modification. Measured PDFs can also improve cosmogenic nuclide-
based erosion rates by more accurately scaling nuclide production rates. Our results
demonstrate the utility of detrital AHe thermochronometry for quantifying erosion in
fluvially and glacially sculpted catchments.
 
Stock, J. and D. R. Montgomery (1999). "Geologic constraints on bedrock river incision using the stream power law." Journal of Geophysical Research 104(B3): 4983-4993.
Stock, J. D. and W. E. Dietrich (2003). "Valley incision by debris flows: Evidence of a topographic signature." Water Resources Research 39(4): doi:10.1029/2001WR001057.
The sculpture of valleys by flowing water is widely recognized, and simplified
models of incision by this process (e.g., the stream power law) are the basis for most
recent landscape evolution models. Under steady state conditions a stream power law
predicts that channel slope varies as an inverse power law of drainage area. Using both
contour maps and laser altimetry, we find that this inverse power law rarely extends to
slopes greater than  0.03 to 0.10, values below which debris flows rarely travel.
Instead, with decreasing drainage area the rate of increase in slope declines, leading to a
curved relationship on a log-log plot of slope against drainage area. Fieldwork in the
western United States and Taiwan indicates that debris flow incision of bedrock
valley floors tends to terminate upstream of where strath terraces begin and where areaslope
data follow fluvial power laws. These observations lead us to propose that the
steeper portions of unglaciated valley networks of landscapes steep enough to produce
mass failures are predominately cut by debris flows, whose topographic signature is an
area-slope plot that curves in log-log space. This matters greatly as valleys with curved
area-slope plots are both extensive by length (>80% of large steepland basins) and
comprise large fractions of main stem valley relief (25–100%). As a consequence,
valleys carved by debris flows, not rivers, bound most hillslopes in unglaciated
steeplands. Debris flow scour of these valleys appears to limit the height of some mountains
to substantially lower elevations than river incision laws would predict, an effect absent in
current landscape evolution models. We anticipate that an understanding of debris flow
incision, for which we currently lack even an empirical expression, would substantially
change model results and inferences drawn about linkages between landscape morphology
and tectonics, climate, and geology.
 
Stock, J. D. and W. E. Dietrich (2006). "Erosion of steepland valleys by debris flows." Geological Society of America Bulletin 118: 1125-1148.
Episodic debris flows scour the rock beds
of many steepland valleys. Along recent
debris-flow runout paths in the western
United States, we have observed evidence for
bedrock lowering, primarily by the impact
of large particles entrained in debris flows.
This evidence may persist to the point at
which debris-flow deposition occurs, commonly
at slopes of less than ~0.03–0.10. We
find that debris-flow–scoured valleys have a
topographic signature that is fundamentally
different from that predicted by bedrock
river-incision models. Much of this difference
results from the fact that local valley
slope shows a tendency to decrease abruptly
downstream of tributaries that contribute
throughgoing debris flows. The degree of
weathering of valley floor bedrock may also
decrease abruptly downstream of such junctions.
On the basis of these observations, we
hypothesize that valley slope is adjusted to
the long-term frequency of debris flows, and
that valleys scoured by debris flows should
not be modeled using conventional bedrock
river-incision laws. We use field observations
to justify one possible debris-flow incision
model, whose lowering rate is proportional to
the integral of solid inertial normal stresses
from particle impacts along the flow and
the number of upvalley debris-flow sources.
The model predicts that increases in incision
rate caused by increases in flow event frequency
and length (as flows gain material)
downvalley are balanced by rate reductions
from reduced inertial normal stress at lower
slopes, and stronger, less weathered bedrock.
These adjustments lead to a spatially uniform
lowering rate. Although the proposed expression
leads to equilibrium long-profiles with
the correct topographic signature, the crudeness
with which the debris-flow dynamics are
parameterized reveals that we are far from a
validated debris-flow incision law. However,
the vast extent of steepland valley networks
above slopes of ~0.03–0.10 illustrates the need
to understand debris-flow incision if we hope
to understand the evolution of steep topography
around the world.
 
Stock, J. D. and D. R. Montgomery (1999). "Geologic constraints on bedrock river incision using the stream power law." Journal of Geophysical Research 104(B3): 4983-4993.
Stolum, H.-H. (1998). "Planform geometry and dynamics of meandering rivers." GSA Bulletin 110(11): 1485-1498.
Storck, P., et al. (1998). "Application of a GIS-based distributed hydrology model for prediction of forest harvest effects on peak stream flow in the Pacific Northwest." Hydrological Processes 12: 889-904.
Storey, H. C., et al. (1964). Hydrology of forest lands and range lands. Handbook of Applied Hydrology. V. T. Chow. New York, McGraw-Hill.
Stout, M. L. (1985). Relation Between Large Landslides and Debris Flows. Fourth International Conference and Field Workshop on Landslides, Tokyo, Japan.
Strahler, A. N. (1952). "Hypsometric (area-altitude) analysis of erosional topography." Bulletin of the Geological Society of America 63: 1117-1142.
Strategic Plan Core Team (1998). Strategic Plan for the Ecosystem Restoration Program Strategic Plan Core Team: 8.
Straub, J. W. (1998). Landslides on the Queen Charlotte Islands: Processes, Rates, and Climatic Events. Carnation Creek and Queen Charlotte Islands Fish/Forestry Workshop: Applying 20 Years of Coast Research to Management Solutions, B.C. Ministry of Forests.
Strauss, D., et al. (1989). "Do one percent of forest fires cause ninety-nine percent of the damage?" Forest Science 35(2): 319-328.
Strouss, D., et al. (1989). "Do one percent of forest fires cause ninety-nine percent of the damage?" Forest Science 35(2): 319-328.
Stubblefield, A. P. (2002). Spatial and temporal dynamics of watershed sediment delivery, Lake
    Tahoe, California. Davis, University of California.
Stuiver, M., et al. (1991). "Climatic, Solar, Oceanic, and Geomagnetic Influences on Late-glacial and Holocene Atmospheric ^14C/^12C Change." Quaternary Research 35: 1-24.
Styllas, M. N. (2001). Corvallis, Oregon State University.
Tillamook Bay is the second largest estuary on the Oregon coast, and concerns have been raised whether human induced impacts have been responsible for the perceived increase in sedimentation rates during the past century. Major land-use practices within the five watersheds of the Bay include logging, forest fires, the construction of forest roads, the placement of dikes along the channels of the main rivers and in the estuary, the removal of riparian vegetation, and the construction of jetties at the tidal inlet. Each of these practices has led to impacts on the entire ecosystem of the watersheds and the Bay, but this study focuses on the effects of human disturbances on the Bay's sediment accumulation. This study examines in detail the land-use practices that have occurred in the watersheds, on the beaches, and in the estuary, focusing on those that have had a direct impact on the sedimentation regime of the Bay. One goal of the study is to assess the relative roles of natural processes versus human impacts on the sedimentation. A general description of the physical characteristics of Tillamook Bay and its surroundings is included, and a brief discussion is provided about the tectonic setting of the Northwest Coast, including its history of subduction earthquakes and the associate sea-level changes. Also provided is a summary of the existing information concerning the arrival of Indians and their environmental impacts, followed by a more detailed account of the major impacts that have resulted from the settlement of the Euro-Americans in the Tillamook area, in the 1850's. The study then focuses on the description of the watersheds from a geomorphologic point of view, and the important land-use practices that may have affected sediment yields during the past century. Analyses of the hydrology of the Tillamook Bay watersheds are included, and the relations between annual water yields and total precipitation are examined in distinct time intervals, each corresponding to a different period with different amounts of land uses. The results of these hydrology analyses suggest that the Tillamook watershed gradually recovered from a period of major disturbances (from 1933 to 1955) to more normal conditions (from 1977 to 1998). In addition, this part of the study attempts to quantify the sediment transport regime of the rivers draining the watershed by using a hydraulic model that is based on the principle of stream power, and on considerations of availability of transported material. Application of this model during the 1933-1955 period for the major rivers suggests an average sediment yield on the order of 410,540 tons/year, but most important are the relative changes of the delivered sediment through time. The results of the model suggest a 1.6-factor decrease of the amount of river sediments from the Heavily Impacted Period (1933-1955) of major disturbances to the Normal Period (1977-1998). The spatial variations of beach and river derived sediments throughout the Bay are determined from textural and mineralogical analyses of surface sediment samples, with the beach sands dominating the area close to the inlet and the river derived sands being mainly deposited at the southeast and northeast parts of the Bay. The relative contributions of these two major sources of sediment were found to be 60% for the marine beach and 40% for the river sands. Further attempt is made to distinguish between the sand transported into the Bay from the individual rivers, and to determine the main processes that are responsible for the dispersion of sediments within the Bay. The attempts to distinguish sands contributed by the individual rivers involved modal analyses of the frequency curves of the surface sediment samples, and the results mainly suggest a grain-size increase away from the mouth of the rivers as a result of sediment reworking by estuarine processes following its initial deposition during episodic river flooding. The main processes that control the dispersion of sediments and their deposition within the Bay were identified by using factor analysis, the results of which suggest that various estuarine processes are responsible for the observed dispersal patterns. A brief review is provided of the study undertaken by Dr. James McManus for the collection and analyses of core samples from Tillamook Bay. Down-core geochemical analyses of major and minor elements indicate that there have been times of episodic input of marine sediment in the central and western portions of the Bay, which is a result of either periodic breaching or washover of Bayocean Spit, so that the beach sand source was more important in the past. This episodic input of marine sand as inferred from the down-core geochemical variations was related to the most recent subduction earthquake, which occurred on January 26th, 1700. Finally, a summary of the results and conclusions of different aspects of this study is presented, so that sedimentation in Tillamook Bay can be viewed as an integrated process involving the watersheds, the estuary, and the ocean beaches.
 
Styllas, M. N. (2001). "Sediment accumulation and human impacts in Tillamook Bay, Oregon."
Sueker, J. K., et al. Effect of basin physical characteristics on solute fluxes in nine alpine/subalpine basins, Colorado, USA. Hydrological Processes 15 (14), 2749-2769 pp.;   42 ref.; 2001.
Alpine/subalpine basins may exhibit substantial variability in solute fluxes despite many apparent similarities in basin characteristics. An evaluation of controls on spatial patterns in solute fluxes may allow development of predictive tools for assessing basin sensitivity to outside perturbations such as climate change or deposition of atmospheric pollutants. Relationships between basin physical characteristics, determined from geographical information system (GIS) tools, and solute fluxes and mineral weathering rates were explored for nine alpine/subalpine basins in Rocky Mountain National Park, Colorado, USA, using correlation analyses for 1993 and 1994 data. Stream-water nitrate fluxes were correlated positively with basin characteristics associated with the talus environment; i.e., the fractional amounts of steep slopes (more than or equal to 30 deg ), unvegetated terrain and young debris (primarily Holocene till) in the basins, and were correlated negatively with fractional amounts of subalpine meadow terrain. Correlations with nitrate indicate the importance of the talus environment in promoting nitrate flux and the mitigating effect of areas with established vegetation, such as subalpine meadows. Total mineral weathering rates for the basins ranged from about 300 to 600 mol ha-1 year-1. Oligoclase weathering accounted for 30 to 73% of the total mineral weathering flux, and was positively correlated with the amount of old debris (primarily Pleistocene glacial till) in the basins. Although calcite is found in trace amounts in bedrock, calcite weathering accounted for up to 44% of the total mineral weathering flux. Calcite was strongly correlated with steep slope, unvegetated terrain, and young debris - probably because physical weathering in steep-gradient areas exposes fresh mineral surfaces that contain calcite for chemical weathering. Oligoclase and calcite weathering are the dominant sources of alkalinity in the basins. However, atmospherically deposited acids consume much of the alkalinity generated by weathering of calcite and other minerals in the talus environment.
 
Sugden, B. D. (1989). The effect of fire suppression on forest stand dynamics: 19.
Sugden, D. E. and B. S. John (1976). Glacier Ice Glaciers and Landscape: A Geomorphical Approach, Edward Arnold. Chapter 2: 15-34.
Sullivan, D., et al. (2000). "Late Quaternary environmental change in western Colorado; results from close-interval sediment sampling." Program and Abstracts - American Quaternary Association. Conference 16: 102.
Sullivan, K., et al. (1987). A summary report of the Deschutes Basin: Sediment, flow temperature, and fish habitat, Weyerhaeuser Company.
Sullivan, K., et al. (1987). Stream channels: the link between forests and fishes. Steramside Management: Forestry and Fishery Interactions, University of Washington, Seattle, WA, University of Washington, Institute of Forest Resources.
Sumaryono, A. (?). Nuee Ardante Deposit and It's Subsequent Runoff Through Channel. Yogyakarta, Republic of Indonesia, Volcanic Sabo Technical Centre: 6.
Sumioka, S. S., et al. (1998). Magnitude and Frequency of Floods in Washington. Tacoma, Washington, U. S. Geological Survey: 91.
Summerfield, M. A. (1985). Plate tectonics and landscape development on the African continent. Tectonic Geomorphology. M. Morisawa and J. T. Hack. Boston and London, Allen and Unwin. Chapter 2: 27-51.
Sun, G., and S.G. McNulty (1998). Modeling soil erosion and transport on forest landscape. In: Winning solutions for risky problems: Proceedings of conference 29; 1998 February 16-20; Reno, NV. Steamboat Springs, CO: International Erosion Control Association: 189-198.
Century-long studies on the impacts of forest management in North America suggest sediment can cause major reduction on stream water quality. Soil erosion patterns in forest watersheds are patchy and heterogeneous. Therefore, patterns of soil erosion are difficult to model and predict. The objective of this study is to develop a user friendly management tool for land managers to design forest management activities (e.g., road building, prescribed burning) that may minimize water quality impacts. This system has the capability to predict long-term soil erosion and sediment transport from hillslopes to stream networks under different climate conditions and forest management scenarios. A Geographic Information System (GIS) coupled with the Universal Soil Loss Equation (USLE) model was used to facilitate database development, manipulation, and output display. The 1140 ha watershed was divided into 30 by 30 m grid cells and gross soil erosion was first predicted by the USLE model for each cell. The Arc/Info GIS utilities are employed to calculate the total mass of sediment moving from each cell to the nearest stream network. Field measurements were used to develop sediment movement routing functions. This study concluded that poorly managed roads are the main source of sediment in a forested watershed. The spatial location of forest roads affected sediment contribution to streams.
 
Sun, H., et al. (2002). "Contour-based digital elevation modeling of watershed erosion and sedimnetation: Erosion and sedimentation estimation tool (EROSET)." Water Resources Research 38(11): 10.1029/2001WR000960.
Sun, T., et al. (1996). "A simulation model for meandering rivers." Water Resources Research 32(9): 2937-2954.
SungRyong, H., et al. Effects of land use and municipal wastewater treatment changes on stream water quality. Environmental Monitoring and Assessment 70 (1/2), 135-151 pp.;   7 ref.; 2001. G. M. K. YoungJoon.
This study was conducted to analyse the quantitative impact of a municipal waste water treatment operation on the long term water quality changes in a tributary of the Han-river, Korea during 1994 to 1999. Changes of land use pattern in the study watershed are quantitatively analysed on the basis of land use maps that were created by classifying Landsat TM images acquired in April 1994 and March 1999. During this period, the average increase of land use area in terms of residence, cultivation, and barren was 5.89, 0.13, and 0.12%, respectively, and the corresponding decrease in water and forest area was 0.21 and 0.16%. The annual average reductions of biochemical oxygen demand, T-N, and T-P by the municipal waste water treatment operation were approximately 89, 11 and 27%, respectively. Spatial analysis of the pollution discharge from watershed was undertaken using a geographical information system (GIS) based model. A clear reciprocal relationship was found between the basin-wide self-purification coefficient and the watershed form ratio excepting a catchment area with water drain facilities. Due to land use changes over the five year study period, water quality change in terms of BOD, T-N, and T-P were (+)1.04 mg litre-1 (corresponding to a 13.7% increase of pollution), (+)0.58 mg litre-1 (10.0% increase), and (-)0.01 mg litre-1 (1.6% decrease). On the other hand, the effect of water quality restoration assessed by outward appearance during the same period was approximately 67.6, 39, and 36.5%, respectively. Consequently, it is understood that total stream water quality recovery in terms of BOD, T-N, and T-P were 81.3, 49.0, and 38.1% respectively, and that this included a negative contribution resulting from increased land use and a positive contribution due to the waste water treatment operation at Inchon.
 
Survey, U. F. S. T. N. F. S. (2002). Adaptive Management Services Enterprise Team, Nevada City, California.
Sutherland, D. G., et al. (2002). "Evolution of a landslide-induced sediment wave in the Navarro River, California." Geological Society of America Bulletin 114(8): 1036-1048.
ABSTRACT
A streamside landslide delivered 60 000
m 3 of mixed-size sediment to the Navarro
River, a sinuous gravel-bed channel (drain-age
area 5 535 km 2 ), at the end of the an-nual
high-runoff period in spring 1995. The
deposit formed a 9-m-high dam that par-tially
breached within several hours, but re-cessional
flows entrained little material un-til
the following high-runoff season. The
landslide afforded the opportunity to mea-sure
the evolution of a sediment wave from
its inception to near-obliteration and, par-ticularly,
to test relative tendencies for
translation and dispersion of a sediment
wave in a natural gravel-bed channel. This
study represents a simple case: The wave
originated from a single input, the preex-isting
channel was relatively uniform, and
resistant banks prevented adjustments in
width. We surveyed channel topography
over a 1.5–4.5 km reach centered on the
landslide dam each year from 1995 to 1999,
and we sampled bed material downstream
of the dam in 1995 and 1997. Landslide ma-terial
was coarser than ambient bed mate-rial,
but all sizes were mobilized by subse-quent
peak flows. Abrasion of weathered
and fractured graywacke sandstone land-slide
material was roughly an order of mag-nitude
greater than the ambient river
gravel.
The sediment wave dispersed and mostly
disappeared within a few years with no
measurable translation. Sediment filled the
reservoir created by the eroding landslide
dam until throughput of bed load was re-stored
in 1998. The stationary wave crest
eroded until in 1999 it was ,1 m higher
than the preslide elevation. As the wave
profile flattened, its detectable leading edge
extended downstream from 620 m in 1995
to ;1600 m in 1997. Downstream advance
of the wave was associated with coarsening
of bed material.
The sediment wave created a longitudi-nal
disturbance in sediment transport. By
using the dam as a reference datum of zero
bed-load transport, we computed longitu-dinal
variations in annual bed-load and
suspended-sediment transport rates in 100
m increments downstream of the dam.
These longitudinal variations were con-trolled
by scour and fill of the bed and by
abrasion of bed-load particles. Bed-load
transport rates in the first and second years
after the landslide increased in the land-slide
vicinity and then decreased down-stream
as sediment deposited behind the
advancing leading edge of the wave. The lo-cation
of peak bed-load transport rate ad-vanced
from the first year (400 m) to the
second (800 m).
We used a physically based, one-dimensional
model (Cui et al., 2002b) to
hindcast annual changes in transversely av-eraged
bed elevation over the study reach.
Agreement between measured and predict-ed
bed elevations was very good. This result
supports our conclusion that, once em-placed,
sediment waves in gravel-bed rivers
tend to disperse, with little or no
translation.
Keywords: bed-load waves, geomorpholo-gy,
modelling, sediment transport.
INTRODUCTION
Understanding the movement geomorphic problems. In the short term (10 0 –
10 1 yr), sediment commonly carries the sig-nature
of upstream disturbances in runoff and
erosion to downstream channels and valley
bottoms and their aquatic and riparian ecosys-tems.
The arrival, duration, and intensity of
sediment impacts depend strongly on how ex-cess
sediment is distributed and interacts with
stored sediment once it reaches the channel
system. At longer time scales (10 1 –10 3 yr),
sediment fluxes build and modify alluvial
landforms.
Sediment loads in channels are commonly
punctuated in time and space by large inputs
that produce sediment ‘‘pulses’’ or ‘‘waves.’’
Sediment waves are transient zones of sedi-ment
accumulation in channels that evolve by
interactions between flow and sediment trans-port
in the particular valley-bottom and chan-nel
topography through which they propagate.
We adopt the term ‘‘wave’’ to refer to a prop-agating
disturbance in bed elevation and sed-iment
properties that might translate and/or
disperse. Thus, a wave does not necessarily
consist of only the original sediment of the
input, but can also incorporate and mix with
preexisting sediment in the river. The problem
can be stated rhetorically: How do river sys-tems
digest large sediment inputs (Cui et al.,
2002a)? Understanding how single sediment
waves disperse and propagate should improve
understanding of the routing of sediment from
inputs distributed in a drainage network and
thereby aid analyses of cumulative watershed
effects. Such understanding should also im-prove
predictions of the fate of large volumes
of sediment released from decommissioned
dams. In this study, we focus on a wave of
bed material (sand and gravel) and neglect
 
Suwa, H. (1988). "Focusing Mechanism of Large Boulders to a Debris-Flow Front." Transactions of the Japanese Geomorphical Union 9(3): 151-178.
Suwa, H. and S. Okuda (1980). "Dissection of valleys by debris flows." Z. Geomorph. N.F. Suppl.-Bd. 35: 164-182.
Suwa, H. and S. Okuda (1985). Measurement of Debris Flows in Japan. IVth International Conference and Field Workshop on Landslides, Tokyo, Japan.
Suzuki, N. and W. C. McComb (1998). "Habitat classification models for beaver (Castor canadensis) in the streams of the central Oregon Coast Range." Northwest Science 72(2): 102-110.
Swales, S. and C. D. Levings (1989). "Role of off-channel ponds in the life cycle of juvenile coho salmon (Oncorhynchus kisutch) and other juvenile salmonids in the Coldwater River, British Columbia." Canadian Journal of Fisheries and Aquatic Sciences 46: 222-242.
Swansen, F. J. (1980). Geomorphology and Ecosystems (DRAFT). 40th Annual Biology Colloquim, Oregon State University.
Swansen, F. J. and C. T. Dyrness (1975). "Impact of clearcutting and road construction on soil erosion by landslides in the western Cascade Range, Oregon." Geology 3(7): 393-396.
Swansen, F. J., et al. (1985). Some effects of slope movements on river channels. Proceedings of the International Symposium on Erosion, Debris Flow and Disaster Prevention, Tokyo.
Swanson, D. N. (1969). Mass Wasting in Coastal Alaska, U.S.D.A. Forest Service: 15.
Swanson, F., et al. (1987). Mass erosion and other sediment sources. Streamside Management: Forestry and
Fishery Interaction, University of Washington, Institute of Forest Resources.
Swanson, F. J. (1981). "Fire and geomorphic processes." General Technical Report WO-26: 401-420.
Swanson, F. J. (1981). Fire and Geomorphic Processes. Fire Regimes and Ecosystem Properties, General Technical Report W0-26. Washington D.C., US Department of Agricuture Forest Service: 401-420.
Swanson, F. J., and R.L. Fredriksen (1982). Sediment routing and budgets: implications for judging impacts of forestry practices. In Sediment Budgets and Routing in Forested Drainage Basins: 129-137. USDA Forest Service, Pacific Northwest Forest and Range Experiment Station General Technical Report PNW-141: Portland, Oregon. In Sediment Budgets and Routing in Forested Drainage Basins: 129-137. USDA Forest Service, Pacific Northwest Forest and Range Experiment Station General Technical Report PNW-141: Portland, Oregon.
Swanson, F. J. (2003). Wood in rivers: a landscape perspective. The Ecology and Management of Wood in World Rivers. S. V. Gregory, K. L. Boyer and A. M. Gurnell. Bethesda, Maryland, American Fisheries Society: 299-313.
Swanson, F. J. (?). "Fire, Geomorphic Processes, and Landforms."
Swanson, F. J., et al. (1987). Mass Failures and Other Processes of Sediment Production in Pacific Northwest Forest Landscapes. Streamside Management: Forestry and Fishery Interactions. E. D. Salo and T. W. Conoy. Seattle, WA. Contribution #57: 9-37.
Swanson, F. J., et al. (1987). Mass failures and other processes of sediment production in Pacific Northwest forest landscapes. Streamside Management: Forestry and Fishery Interactions. E. O. Salo and T. W. Cundy. Seattle, WA, Institute of Forest Resources, University of Washington: 9-38.
Swanson, F. J. and C. T. Dyrness (1975). "Impact of clearcutting and road construction on soil erosion by landslides in the western Cascade Range, Oregon." Geology 3(7): 393-396.
Swanson, F. J., et al. (1990). Landscape Patterns Disturbance and Management in the Pacific Northwest, USA. Trends in Landscape: Ecology. I. S. Zonneveld and R. T. T. Forman. New York, Springer-Verhg. Chapter 11: 191-213.
Swanson, F. J., et al. Material Transfer in a Western Oregon Forested Watershed. ?
Swanson, F. J., et al. (1982). Material transfer in a western Oregon forested watershed. Analysis of Coniferous Forested Ecosystems in the western United States, Hutchinson, Dowden, Ross Publication Company, Stroudsburg, PA.
Swanson, F. J., et al. (1982). Material transfer in a western Oregon forested watershed. analysis of Coniferous Forest Ecosystems in the Western United States. R. L. Edmonds. Stroudsburg, Penn., Hutchinson Ross Publishing Co.: 233-266.
Swanson, F. J., et al. (1982). Land-Water Interactions: The Riparian Zone. Analysis of coniferous forest ecosystems in the western United States. R. L. Edmonds, Hutchinson Ross Publishing Company. Chapter 9: 267-291.
Swanson, F. J., et al. (1982). Introduction workshop on sidiment budgets and routing in forested drainage basins: 1-165.
Swanson, F. J., et al. (1982). Sediment budgets and routing in forested drainage basins, U.S. Department of Commerce National Technical Information Service
Pacific Northwest Research Station: 1-165.
Swanson, F. J., et al. (1998). "Flood disturbance in a forested mountain landscape." BioScience 48(9): 681-689.
Swanson, F. J., et al. (1998). "Flood Disturbance in a Forested Mountain Landscape:  Interactions of land use and floods."  48(9): 681-689.
Swanson, F. J., et al. (?). "The Physical Environment as a Basis for Managing Ecosystems "? Chapter 15: 229-238.
Swanson, F. J., et al. (1988). "Landform effects on ecosystem patterns and processes." BioScience 38(2): 92-98.
Swanson, F. J. and G. W. Lienkaemper (1978). Physical consequences of large organic debris in pacific northwest streams. Portland, Oregon, USDA Forest Service: 12.
Swanson, F. J., et al. (1976). History, physical effects, and management implications of large organic debris in western Oregon streams, USDA Forest Service.
Swanson, F. J., et al. (1986). Landslide Dams in Japan. Landslide Dams: Processes, Risk, and Mitigation, Seattle, Washington.
Swanson, F. J. and C. Roach (1987). Adminstrative Report Mapleton Leave Area Study. Corvallis, Oregon, USDA Forest Service: 139.
Swanson, F. J., et al. (1977). Inventory of mass erosion in the Mapleton Ranger District, Siuslaw National Forest, Siuslaw National Fores and the Pacific Northwest Forest and Range Experiment Station: 1-41.
Swanson, F. J., et al. (1981). Analysis of debris-avalanche erosion in steep forest lands: An example from Mapleton, Oregon, USA. Erosion and Sediment Transport in Pacific Rim Steeplands. Christchurch, I.A.H.S. Publ. No. 132: 67-75.
Swanson, F. J., et al. (1989). Landforms, disturbance, and ecotones. SCOPE (Scientific Committee on Problems of the Environment) Workshop, Paris.
Swanston, D. N. (1969). Mass Wasting in Coastal Alaska, USDA Forest Service: 15.
Swanston, D. N. (1970). Mechanics of debris avalanching in shallow till soils of Southeast Alaska. Juneau, Alaska, U.S. Forest Service: 2-17.
Swanston, D. N. (1974). Slope stability problems associated with timber harvesting in mountainous regions of the western United States. Portland, Oregon, U.S. Department of Agriculture: 9-14.
Swanston, D. N. (1980). Influence of forest and rangeland management on anadromous fish habitat in western North America, USDA Department of Agriculture
Forest Service
Pacific Northwest Forest and Range Experiment Station: 1-27.
Swanston, D. N. (1981). Creep and earthflow erosion from undisturbed and management impacted slopes in the Coast and Cascade Ranges of the Pacific Northwest, U.S.A. Erosion and Sediment Transport in Pacific Rim Steeplands. Christchurch, I. A. H. S.: 76-94.
Swanston, D. N. (?). "Effect of Geology on Soil Mass Movement Activity in the Pacific Northwest (DRAFT)."?: 89-115.
Swanston, D. N. (?). Landslide prediction and assessment: interpreting stability problems for the land manager.
Swanston, D. N., et al. (1988). "Timber harvest and progressive deformation of slopes in southwestern Oregon." Bulletin of the Association of Engineering Geologists 25(3): 371-381.
Swanston, D. N. and D. A. Marion (1991). Landslide response to timber harvest in Southeast Alaska. Fifth Federal Interagency Conference, Las Vegas, Nevada.
Swanston, D. N. and F. J. Swanson (1976). Timber Harvesting, Mass Erosion, and Steepland Forest Geomorophology in the Pacific Northwest. Geomorphology an engineering. D. R. Coates, Hutchinson and Ross, Inc.: 199-221.
Syverson, T. L. (1984). History and Origin of debris torrents in the Smith Creek Drainage, Whatcom County, Washington, Western Washington University: 1-81.
Tabor, R. W. and W. M. Cady (1978). Maps showing ground-failure hazards in the Columbia River Valley between Richland and Priest Rapids Dam, South-Central Washington. ?
Taft, C. E. (1961). "A Revised Key for the Field Idenfication of Some Genera of Algae." Tutox News 39(4): 99-103.
Tague, C., et al. (2008). "Deep groundwater mediates streamflow response to climate warming in the Oregon Cascades." Climate Change 86: 189-210.
Recent studies predict that projected climate change will lead to significant
reductions in summer streamflow in the mountainous regions of the Western US.
Hydrologic modeling directed at quantifying these potential changes has focused on the
magnitude and timing of spring snowmelt as the key control on the spatial–temporal pattern
of summer streamflow. We illustrate how spatial differences in groundwater dynamics can
also play a significant role in determining streamflow responses to warming. We examine
two contrasting watersheds, one located in the Western Cascades and the other in the High
Cascades mountains of Oregon. We use both empirical analysis of streamflow data and
physically based, spatially distributed modeling to disentangle the relative importance of
multiple and interacting controls. In particular, we explore the extent to which differences in
snow accumulation and melt and drainage characteristics (deep ground water vs. shallow
subsurface) mediate the effect of climate change. Results show that within the Cascade
Range, local variations in bedrock geology and concomitant differences in volume and
seasonal fluxes of subsurface water will likely result in significant spatial variability in
responses to climate forcing. Specifically, watersheds dominated by High Cascade geology
will show greater absolute reductions in summer streamflow with predicted temperature
increases.
 
Tague, C. and G. E. Grant (2004). "A geological framework for interpreting the low-flow regimes of Cascade streams, Willamette River Basin, Oregon." Water Resources Research 40(W04303).
In ungauged basins, predicting streamflows is a major challenge for hydrologists and
water managers, with approaches needed to systematically generalize hydrometric
properties from limited stream gauge data. Here we illustrate how a geologic/geomorphic
framework can provide a basis for describing summer base flow and recession behavior at
multiple scales for tributaries of the Willamette River in Oregon. We classified the basin
into High Cascade and Western Cascade provinces based on the age of the underlying
volcanic bedrock. Using long-term U.S. Geological Survey stream gauge records, we
show that summer streamflow volumes, recession characteristics, and timing of response
to winter recharge are all linearly related to the percent of High Cascade geology in
the contributing area. This analysis illustrates how geology exerts a dominant control on
flow regimes in this region and suggests that a geological framework provides a useful
basis for interpreting and extrapolating hydrologic behavior.
 
 
Takahashi, G. (1989). "Status of Charr and Masu salmon management in Japan; a call for conservation guidelines." Physiol. Ecol. Japan Spec. Vol. 1: 683-690.
Takahashi, G. (1990). "On the consistency of conservation of freshwater fish and 'Sabo' works in streams." Bulletin of the Institute of Zoology, Academia Sinica 29(3): 105-113.
Takahashi, G. (1990). "A Study in the Riffle Pool Concept." Transactions of the Japanese Geomorphical Union 11(4): 319-336.
Takahashi, G. (?). "Channel Course Fluctuation and Bed Load Movement in Stream Floodplain."  38(4): 3-9.
Takahashi, G. and S. Higashi (1984). "Effect of Channel Alteration on Fish Habitat." The Japanese Journal of Limnology 45(3): 178-186.
Takahashi, T. (1978). "Mechanical Characteristics of Debris Flow." Journal of Hydraulics Division 8: 1153-1169.
Takahashi, T. (1980). "Debris Flow on Prismatic Open Channel." Journal of the Hydraulics Division 106(HY3): 381-395.
Takahashi, T. (1981). "Debris Flow." Ann. Rev. Fluid Mech. 13: 57-77.
Takahashi, T., et al. (1981). Delineation of debris flow hazard areas. Erosion and sediment transport in the Pacific Rim steeplands, IAHS pub. 132., International Association of Hydrological Sciences: 589-603.
Takashi, G., et al. (2001). "The characteristics of woody debris and sediment distribution in headwater streams, southeastern Alaska." Can. J. For. Res. 31(8): 1386-1399.
Takken, I., et al. (1999). "Spatial evaluation of a physically-based distributed erosion model (LISEM)." Catena 37(3-4): 431-447.
Tappel, P. D. and T. C. Bjornn (1983). "A New Method of Relating Size of Spawning Gravel to Salmonid Embryo Survival " North American Journal of Fisheries Management 3: 123-135.
Tarboton, D. G., R.L. Bras, and I. Rodriguez-Iturbe (1991). "On the extraction of channel networks from digital elevation data." Hydrological Processes 5: 81-100.
Tarboton, D. G. (1997). "A new method for the determination of flow directions and upslope areas in grid digital elevation models." Water Resources Research 33(2): 309-319.
Tarboton, D. G. and D. P. Ames (2001). Advances in the mapping of flow networks from digital elevation data. World Water and Environmental Resources Congress, Orlando, Florida.
Tarboton, D. G., et al. (1991). "On the extraction of channel networks from digital elevation data." Hydrological Processes 5(1): 81-100.
Tarboton, D. G., et al. (1998). "Disaggregation procedures for stochastic hydrology based on nonparametric density estimation." Water Resources Research 34(1): 107-119.
Tarolli, P. and G. D. Fontana (2009). "Hillslope-to-valley transition morphology: New opportunities from high resolution DTMs." Geomorphology.
The search for the optimal spatial scale for observing landforms to understand physical processes is
3 a fundamental issue in geomorphology. Topographic attributes derived from Digital Terrain Models
4 (DTMs) such as slope, curvature and drainage area provide a basis for topographic analyses. The
5 slope–area relationship has been used to distinguish diffusive (hillslope) from linear (valley)
6 processes, and to infer dominant sediment transport processes. In addition, curvature is also useful
7 in distinguishing the dominant landform process. Recent topographic survey techniques such as
8 LiDAR have permitted detailed topographic analysis by providing high-quality DTMs. This study
9 uses LiDAR-derived DTMs with a spatial scale between 1 and 30 m in order to find the optimal
10 scale for observation of dominant landform processes in a headwater basin in the eastern Italian
11 Alps where shallow landsliding and debris flows are dominant. The analysis considered the scaling
12 regimes of local slope versus drainage area, the spatial distribution of curvature, and field
13 observations of channel head locations. The results indicate that: i) hillslope-to-valley transitions in
14 slope–area diagrams become clearer as the DTM grid size decreases due to the better representation
15 of hillslope morphology, and the topographic signature of valley incision by debris flows and
16 landslides is also best displayed with finer DTMs; ii) regarding the channel head distribution in the
17 slope–area diagrams, the scaling regimes of local slope versus drainage area obtained with grid
18 sizes of 1, 3, and 5 m are more consistent with field data; and iii) the use of thresholds of standard
19 deviation of curvature, particularly at the finest grid size, were proven as a useful and objective
20 methodology for recognizing hollows and related channel heads.
 
Tarolli, P. and D. G. Tarboton (2006). "A new method for determination of most likely landslide initiation points and the evaluation of digital terrain model scale in terrain stability mapping." Hydrology and Earth System Sciences 10: 663-677.
This paper introduces a new approach for determining
the most likely initiation points for landslides from
potential instability mapped using a terrain stability model.
This approach identifies the location with critical stability index
from a terrain stability model on each downslope path
from ridge to valley. Any measure of terrain stability may
be used with this approach, which here is illustrated using
results from SINMAP, and from simply taking slope as an
index of potential instability. The relative density of most
likely landslide initiation points within and outside mapped
landslide scars provides a way to evaluate the effectiveness
of a terrain stability measure, even when mapped landslide
scars include run out zones, rather than just initiation locations.
This relative density was used to evaluate the utility
of high resolution terrain data derived from airborne laser
altimetry (LIDAR) for a small basin located in the Northeastern
Region of Italy. Digital Terrain Models were derived
from the LIDAR data for a range of grid cell sizes (from 2 to
50 m). We found appreciable differences between the density
of most likely landslide initiation points within and outside
mapped landslides with ratios as large as three or more with
the highest ratios for a digital terrain model grid cell size of
10 m. This leads to two conclusions: (1) The relative density
from a most likely landslide initiation point approach is
useful for quantifying the effectiveness of a terrain stability
map when mapped landslides do not or can not differentiate
between initiation, runout, and depositional areas; and (2)
in this study area, where landslides occurred in complexes
that were sometimes more than 100 m wide, a digital terrain
model scale of 10m is optimal. Digital terrain model scales
larger than 10m result in loss of resolution that degrades the
results, while for digital terrain model scales smaller than 10 m the physical processes responsible for triggering landslides
are obscured by smaller scale terrain variability.
 
Tarutani, N., et al. (2000). "Spatial distribution pattern of rapid shallow landslides in Amakusa Island."?
Tate, N. J., et al. (2005). "Smoothing/filtering LiDAR digital surface models. Experiments with loess regression and discrete wavelets." Journal of Geographical Systems 7: 273-290.
This paper reports on the smoothing/filtering analysis of a digital
surface model (DSM) derived from LiDAR altimetry for part of the River
Coquet, Northumberland, UK using loess regression and the 2D discrete
wavelet transform (DWT) implemented in the S-PLUS and R statistical
packages. The chosen method of analysis employs a simple method to generate
noise’ which is then added to a smooth sample of LiDAR data; loess
regression and wavelet methods are then used to smooth/filter this data and
compare with the original smooth’ sample in terms of RMSE. Various
combinations of functions and parameters were chosen for both methods.
Although wavelet analysis was effective in filtering the noise from the data,
loess regression employing a quadratic parametric function produced the
lowest RMSE and was the most effective
 
Tayfur, G. (2002). "Applicability of sediment transport models for nonsteady state erosion from steep slopes." Journal of Hydrologic Engineering 7(3): 252-259.
Tayfur, G. (2002). "Artificial neural networks for sheet sediment transport." Hydrological Sciences Journal-Journal Des Sciences Hydrologiques 47(6): 879-892.
Taylor, G. (1997). Causes of the flood and a comparison to other climate events. The Pacific Northwest Floods of February 6-11, 1996: Proceedings of the Pacific Northwest Water Issues Conference. A. Laenen and J. D. Ruff. St. Paul, Minnesota, USA, American Institute of Hydrology: 3-7.
Taylor, G. H. (1997). "The Great Flood of 1996." from www.ocs.orst.edu/reports/flood96/Flood2.html.
Taylor, G. H. and C. Hannan (1999). The Climate of Oregon: From Rain Forest to Desert. Corvallis, OR, Oregon State University Press.
Taylor, J. R. (1997). An Introduction to Error Analysis: The Study of Uncertainties in Physical Measurements. Sausalito, California, University Science Books.
Taylor, S. B. and Anonymous (1998). "Geomorphic analysis of small-scale debris fans at three Central Appalachian watersheds; implications for controls on sediment transport efficiency." Abstracts with Programs - Geological Society of America 30(7): 141.
Comparative geomorphic analysis has been completed at three study sites in the central Appalachians: the Fernow Experimental Forest, Tucker County, WV; the North Fork basin, Pocahontas County, WV; and the Little River basin, Augusta County, VA. Each area is underlain by strata of the upper Devonian Acadian clastic wedge. Mapping (1:9600) of small-scale debris fans and morphometric analysis yield results that are useful in deciphering the local controls on sediment transport efficiency at the watershed scale. Debris fans are comprised primarily of gravel diamicton and are dominated by debris flow processes. Map units are delineated according to fan-surface morphology and height above channel grade. Debris fans are classified as either simple or compound. Simple fans are characterized by single map polygon types, with surfaces contoured to present channel grade or positioned at heights ranging from 2-15 m. Compound fans are characterized by multiple map polygon types and inset fan-terrace relationships. Composite fan surfaces range in area from 200 to 45,000 sq. m, averaging 3700 sq. m. Debris fans are preserved at tributary junctions, with greater than 75% occurring at the intersections of 1st- to 2nd-order channels and higher order trunk streams. Comparative geomorphic analysis of the three study sites suggests that the volume of fan deposits in storage vary according to local geologic factors. Areal analysis of map units combined with standard morphometric techniques provide a proxy for sediment storage volume. Critical parameters for the Fernow, North Fork and Little River sites respectively include: 1) Basin Area (sq. km) = 17.6, 49.3, 41.5; 2) Drainage Density (/km) = 4.2, 3.3, 4.7; 3) Strahler Order = 5, 5, 6; 4) Maximum Valley Width (m) = 120, 180, 290; 5) Tributary Junction Frequency (no./sq. km) = 9.0, 5.8, 9.1; 6) Fan Frequency (no./sq. km) = 2.1, 1.0, 2.8; and 7) Fan Area to Basin Area Ratio = 0.007, 0. 004, 0.012. These data document a greater volume of fan deposits in storage at the Little River site, suggesting that the Fernow and North Fork have been more effective at sediment transport during the Quaternary. Lithofacies variation in the Acadian bedrock sequence is interpreted as the primary factor controlling gross watershed morphology. The increased drainage density, valley width and tributary-junction frequency at the Little River site dramatically increase the degree of fan preservation. Assuming a constant climate variable, fan storage is therefore a function of bedrock geology, style of regolith production and transport mechanisms.
 
Teensma, P. D. A. (1987). Fire history and fire regimes of the central western Cascades of Oregon, University of Oregon: 117.
Teensma, P. D. A. (1987). Fire history and fire regimes of the central western Cascades of Oregon. Eugene, OR, University of Oregon: 188.
Teensma, P. D. A., et al. (1991). Preliminary Reconstruction and Analysis of Change in Forest Stand Age Classes of the Oregon Coast Range From 1850 to 1940. Portland, Oregon, USDI Bureau of Land Management: 9.
Terrapin Environmental (2004). Water typing model field performance assessment approach and procedures. Olympia, WA, Cooperative Monitoring Evaluation & Research: 30.
Terzagi, K. (1950). Mechanism of landslides. Application of geology to engineering practice. S. Paige, Berkeley Vol. Geological Society of America: 82-123.
Theobald, D. M. (2008). ICLUS SERGoM v3 User's manual (Preliminary draft). W. EPA, DC: 20.
Theobald, D. M., et al. (2005). Functional Linkage of Watersheds and Streams (FLoWS): Network-based ArcGIS tools to analyze freshwater ecosystems. ESRI User's Conference.
Thomas, A. L., et al. A spatial study of the relationships between streamwater acidity and geology, soils and relief (Vosges, northeastern France). Journal of Hydrology 217 (1/2), 35-45 pp.;   25 ref.; 1999.
A geographic information system (GIS) was used to study the relationships between streamwater pH and alkalinity at base flow with geology, soils and relief in 100 forested catchments located in the sandstone portion of the Vosges mountains, France. At base flow, streamwater acidity depends primarily on bedrock and soil content of weatherable minerals whose dissolution neutralizes acidity. Catchments are developed on three main stratigraphic levels, consisting of two sandstone layers rich in weatherable minerals, called 'rich', at the upper and lower extremity of the stratigraphic sequence, and a quartzitic sandstone, called 'poor', in between. Catchments were classified into three major groups in relation to the stratigraphic sequence: (1) only 'poor' bedrock, (2) 'rich' upstream and 'poor' downstream, (3) 'poor' upstream and 'rich' downstream. Results showed that streamwater pH differed depending on the group. Within each group, the pH was related to bedrock, soil and relief characteristics of the catchments. The relative surface covered by the bedrock located in the lower part of the catchment explained more than 50% of the variability of pH. More than 20% of variability could be explained by soil types in catchments composed of 'poor' bedrock only. Soil type apparently did not influence streamwater pH in the other groups. The catchment area, probably related to the increase of soil volume and water residence time, explained up to 20% of pH variability, depending on the catchment type.
 
Thomas, J. W. (2002). Dynamic vs. Static Management in a Fire-Induced Landscape - The Northwest Forest Plan. Missoula, Montana, University of Montana: 1-22.
Thomas, J. W. (2003). Application of the Northwest Forest Plan in National Forests in California, U.S.D.A., U.S. Forest Service, Pacific Southwest Region.
Thomas RB (1988). "Monitoring baseline suspended sediment in forested basin: the effects of sampling  on suspended sediment rating curves." Hydrological Sciences Journal/Journal des Sciences Hydrologiques 33(5): 499-514.
Thomas, R. B. and W. F. Megahan (1998). "Peak flow responses to clear cutting and roads in small and large basins, western Cascades, Oregon: A second opinion." Water Resources Research 34(12): 3393-3403.
Thomas, R. B. and W. F. Megahan (1998). "Peak flow responses to clear-cutting and roads in small and large basins, western Cascades, Oregon: A second opinion." Water Resources Research 34(2): 3393-3403.
In this paper, we conduct a reanalysis of methods and data used by Jones and
Grant [1996]. Data from three small watersheds (60–101 ha) and three pairs of large
basins (60–600 km2) in Oregon’s western Cascades were used to evaluate effects of timber
harvest and road construction on peak flows. We could not detect any effect of cutting on
peak flows in one of the large basin pairs, and results were inconclusive in the other two
large basin pairs. One small watershed was 100% clear-cut, a second was 31% patch-cut
with 6% of the area affected by road construction, and a third was held as a long-term
control. Peak flows were increased up to 90% for the smallest peak events on the clear-cut
watershed and up to 40% for the smallest peak flows on the patch-cut and roaded
watershed. Percentage treatment effects decreased as flow event size increased and were
not detectable for flows with 2-year return intervals or greater on either treated
watershed. Treatment effects decreased over time but were still found after 20 years on
the clear-cut watershed but for only 10 years on the patch-cut and roaded watershed.
 
Thomas, W. A. and A. L. Prasuhn (1977). "Mathematical Modeling of Scour and Deposition " Journal of the Hydraulics Division 103(8): 851-863.
Thompson, B. (1991). Annotated Bibliography of large organic debris (LOD) with regards to stream channels and fish habitat. Victoria, B.C.: 1-93.
Thompson, D. M., et al. (1999). "Velocity reversals and sediment sorting in pools and riffles controlled by channel constrictions." Geomorphology 27: 229-241.
Thompson, J. (2005). "Is it HIP? Identifying streams with high intrinsic potential to provide salmon and trout habitat." Science Findings 72(April).
Thompson, J. A., et al. (2001). "Digital elevation model resolution: effects on terrain attribute calculation and quantitative soil-landscape modeling." Geoderma 100: 67-89.
The accuracy of digital elevation models  DEM. and DEM-derived products depends on
several factors, including the horizontal resolution and vertical precision at which the elevation
data are represented, and the source of the elevation data. This accuracy becomes increasingly
important as we extend the use of DEM data for spatial prediction of soil attributes. Our objective
was to compare terrain attributes and quantitative soil-landscape models derived from grid-based
DEM represented at different horizontal resolutions  10 and 30 m., represented at different vertical
precisions  0.1 and 1 m., and acquired from different sources. Decreasing the horizontal resolution
of the field survey DEM produced lower slope gradients on steeper slopes, steeper slope gradients
on flatter slopes, narrower ranges in curvatures, larger specific catchment areas in upper landscape
positions, and lower specific catchment areas values in lower landscape positions. Overall, certain
landscape features were less discernible on the 30-m DEM than on the 10-m DEM. Decreased
vertical precision produced a large proportion of points with zero slope gradient and zero slope
curvature, and a large number of steeply sloping and more highly curved areas. Differences among
DEM from different sources were more significant, with less accurate representation of depressions
and drainage pathways with the USGS DEM as compared to the field survey DEM.
Empirical models developed from different DEM included similar predictive terrain attributes, and
were equally successful in predicting A-horizon depth  AHD. in the validation data set.
 
Thompson, J. R., et al. (2006). "Historical disturbance regimes as a reference for forest policy in a multowner province: a simulation experiment." Canadian Journal of Forest Research 36: 401-417.
Using a landscape simulation model, we examined ecological and economic implications of forest policies
designed to emulate the historical fire regime across the 2 × 106 ha Oregon Coast Range. Simulated policies included
two variants of the current policy and three policies reflecting aspects of the historical fire regime. Policy development
was guided by the management intentions of four owner groups: forest industry, nonindustrial private, state, and federal.
Fire severity was emulated with green-tree retention standards; fire frequency was emulated with annual harvestable
area restrictions; and fire extent was emulated with harvest-unit size regulations. Simulated disturbance-based policies
produced age-class distributions closer to the estimated historical range than those created by the current policy. Within
100 years, proportions of younger forests were within the historical range, while older forests moved closer to, but remained
below, historical conditions. In the near term, disturbance-based policies produced annual harvest volumes
20%–60% lower than those produced by the current policy. However, relative costs of disturbance-based policies diminished
over time. Our results suggest that if expediting a return to historical age-class distributions at a provincial-scale
was a goal, then public lands would be needed to provide large patches of old forest. In addition, this experiment illustrated
that distributing costs and benefits of conservation policies equitably across multiple private landowners is a significant
challenge.
 
Thompson, W. L. and D. C. Lee (2000). "Modeling relationships between landscape-level attributes and snorkel counts of chinook salmon and steelhead parr in Idaho." Canadian Journal of Fisheries and Aquatic Science 57: 1834-1842.
Knowledge of environmental factors impacting anadromous salmonids in their freshwater habitats, particularly
at large spatial scales, may be important for restoring them to previously recorded levels in the northwestern
United States. Consequently, we used existing data sets and an information-theoretic approach to model landscape-level
attributes and snorkel count categories of spring–summer chinook salmon (Oncorhynchus tshawytscha) and steelhead
(Oncorhynchus mykiss) parr within index areas in Idaho. Count categories of chinook salmon parr were negatively related
to geometric mean road density and positively related to mean annual precipitation, whereas those for steelhead
parr were negatively related to percent unconsolidated lithology. Our models predicted that chinook salmon parr would
be in low count categories within subwatersheds with >1 km·km –2 geometric mean road densities and (or) <700 mm
mean annual precipitation. Similarly, steelhead parr were predicted to be in low count categories in subwatersheds with
>30% unconsolidated lithology. These results provide a starting point for fish biologists and managers attempting to
map approximate status and quality of rearing habitats for chinook salmon and steelhead at large spatial scales.
 
Thoms, M. C. and M. Parsons (2002). Eco-geomorphology: an interdisciplinary approach to river science. The Structure, Function and Management Implications of Fluvial Sedimentary Systems, Alice Springs, Australia, IAHS.
Thorne, C. R., et al. Sediment Transport in Gravel-Bed Rivers, John Wiley and Sons.
Thorne, C. R. and L. W. Zevenbergen (1986). "Estimating mean velocity in mountain rivers." Am. Soc. Civ. Eng., J. of Hydraulic Engineering 111(4): 612-624.
Thorp, J. H., et al. (2006). "The riverine ecosystem sysnthesis: biocomplexity in river networks across space and time." River Research and Applications 22(123-147): 123-147.
We propose an integrated, heuristic model of lotic biocomplexity across spatiotemporal scales from headwaters to large rivers.
This riverine ecosystem synthesis (RES) provides a framework for understanding both broad, often discontinuous patterns along
longitudinal and lateral dimensions of river networks and local ecological patterns across various temporal and smaller spatial
scales. Rather than posing a completely new model, we arrange a conceptual marriage of eco-geomorphology (ecological
aspects of fluvial geomorphology) with a terrestrial landscape model describing hierarchical patch dynamics. We modify five
components of this terrestrial model for lotic ecosystems: (1) nested, discontinuous hierarchies of patch mosaics; (2) ecosystem
dynamics as a composite of intra- and inter-patch dynamics; (3) linked patterns and processes; (4) dominance of non-equilibrial
and stochastic processes; and (5) formation of a quasi-equilibrial, metastable state. Our conceptual model blends our perspectives
on biocomplexity with aspects of aquatic models proposed from 1980–2004.
Contrasting with a common view of rivers as continuous, longitudinal gradients in physical conditions, the RES portrays
rivers as downstream arrays of large hydrogeomorphic patches (e.g. constricted, braided and floodplain channel areas) formed
by catchment geomorphology and climate. The longitudinal distribution of these patches, which are identifiable using standard
geomorphic techniques, varies amongst rivers and is difficult to forecast above ecoregional scales. Some types of hydrogeomorphic
patches may reoccur along this downstream passage. Unique ecological ‘functional process zones’ are formed by individual
types of hydrogeomorphic patches because of physiochemical habitat differences which affect ecosystem structure and
function.
The RES currently includes 14 tenets predicting how patterns of individual species distributions, community regulation, lotic
ecosystem processes, and floodplain interactions will vary over spatiotemporal scales, especially as they relate to the functional
process zones formed by hydrogeomorphic differences in the river network.
 
Thorsen, G. W. (1978). "Landslide provinces in Washington." Engineering Geology in Washington I: 71-90.
Thorsen, G. W. (1989). Landslide Provinces in Washington. Engineering Geology in Washington. Bulletin 78. Olympia, Washington State Department of Natural Resources, Division of Geology and Earth Resources. 1: 71-89.
Thorsen, G. W. (1989). "Splitting and Sagging Mountains." Washington Geologic Newsletter 17(4): 3-13.
Thurow, R. F. and J. G. King (1994). "Attributes of Yellowstone Cutthroat Trout Redds in a Tributary of the Snake River, Idaho." Transactions of the American Fisheries Society 123: 37-50.
Tilman, D. (1996). "The benefits of natural disasters." Science 273.
Tiwari, A. K., L.M. Risse, and M.A. Nearing (2001). "Evaluation of WEPP and its comparison with USLE and RUSLE." Transactions of the ASAE 43(5): 1129-1135.
Toews, D. A. A. and D. R. Gluns (1986). Snow Accumulation and Ablation on Adjacent Forested and Clearcut Sites in Southeastern British Columbia. Western Snow Conference, Phoenix, Arizona.
Toews, D. A. A. and M. K. Moore (1982). The Effects of Three Streamside Logging Treatments on Organic Debris and Channel Morphology of Carnation Creek. Carnation Creek Workshop: A 10-year Review, Nanaimo, British Columbia.
Tonina, D., et al. (2008). "Hydrological response to timber harvest in northern Idaho: implications for channel scour and persistence of salmonids." Hydrological Processes.
The potential for forest harvest to increase snowmelt rates in maritime snow climates is well recognized. However, questions
still exist about the magnitude of peak flow increases in basins larger than 10 km2 and the geomorphic and biological
consequences of these changes. In this study, we used observations from two nearly adjacent small basins (13 and 30 km2) in
the Coeur d’Alene River basin, one with recent, relatively extensive, timber harvest, and the other with little disturbance in
the last 50 years to explore changes in peak flows due to timber harvest and their potential effects on fish. Peak discharge was
computed for a specific rain-on-snow event using a series of physical models that linked predicted values of snowmelt input to
a runoff-routing model. Predictions indicate that timber harvest caused a 25% increase in the peak flow of the modelled event
and increased the frequency of events of this magnitude from a 9-year recurrence interval to a 3Ð6-year event. These changes in
hydrologic regime, with larger discharges at shorter recurrence intervals, are predicted to increase the depth and frequency of
streambed scour, causing up to 15% added mortality of bull trout (Salvelinus confluentus) embryos. Mortality from increased
scour, although not catastrophic, may have contributed to the extirpation of this species from the Coeur d’Alene basin, given
the widespread timber harvest that occurred in this region.
 
Tooth, S. "Process, form and change in dryland rivers: a review of recent research."
Many of the world's extensive warm dryland regions support numerous, albeit often infrequently flowing, rivers. Dryland rivers are increasingly a focus of scientific and applied interest but empirical research and fluvial theory for drylands need to be strengthened. Recent research in arid central Australia indicates greater diversity in dryland river process, form and change than has hitherto been appreciated, and highlights the need for a global review assessing the present state of knowledge. This review outlines the distinctive characteristics of dryland fluvial environments (hillslope and channel hydrological and sediment transport processes, river pattern and geometry, temporal and spatial aspects of channel change, sedimentary structures and bedforms), many of which contrast with more humid fluvial environments. Although features common to many dryland fluvial environments can be identified (extreme temporal and spatial variability of rainfall, runoff and sediment transport, poor integration between tributary and trunk channels, importance of large floods as a control on channel morphology, lack of equilibrium between process and form), the fluvial diversity that exists within drylands requires recognition of the limitations to these generalisations. In particular, research in central Australia illustrates the need to understand the rivers of this region using empirical relationships, terms, and concepts additional to those defined by earlier work in drylands. Key deficiencies in dryland fluvial research are identified, and relate to three main areas: limited study of some aspects of modern dryland rivers (floodplain characteristics, influence of vegetation, downstream changes, importance of scale); limited understanding of dryland river behaviour over longer (Cenozoic) timescales; and lack of integration between the results from short-term, process-form studies and studies of the longer term histories of river behaviour. Linking knowledge of past hydrological and channel changes to present-day changes in dryland rivers is suggested as a key research priority. This will help develop a sound theoretical basis for the assessment of future developments in dryland river systems which will contribute to their improved scientific understanding and environmentally sensitive management.
 
Topping, B. J. D., et al. (2009). Oregon streamflow duration assessment method interim version.
Torgersen, C. E. and D. A. Close (2004). "Influence of habitat heterogeneity on the distribution of larval Pacific lamprey (Lampetra tridentata) at two spatial scales." Freshwater Biology 49: 614-630.
1. Spatial patterns in channel morphology and substratum composition at small
(1–10 metres) and large scales (1–10 kilometres) were analysed to determine the influence
of habitat heterogeneity on the distribution and abundance of larval lamprey.
2. We used a nested sampling design and multiple logistic regression to evaluate spatial
heterogeneity in the abundance of larval Pacific lamprey, Lampetra tridentata, and habitat in
30 sites (each composed of twelve 1-m2 quadrat samples) distributed throughout a 55-km
section of the Middle Fork John Day River, OR, U.SA. Statistical models predicting the
relative abundance of larvae both among sites (large scale) and among samples (small
scale) were ranked using Akaike’s Information Criterion (AIC) to identify the ‘best
approximating’ models from a set of a priori candidate models determined from the
literature on larval lamprey habitat associations.
3. Stream habitat variables predicted patterns in larval abundance but played different
roles at different spatial scales. The abundance of larvae at large scales was positively
associated with water depth and open riparian canopy, whereas patchiness in larval
occurrence at small scales was associated with low water velocity, channel-unit
morphology (pool habitats), and the availability of habitat suitable for burrowing.
4. Habitat variables explained variation in larval abundance at large and small scales, but
locational factors, such as longitudinal position (river km) and sample location within the
channel unit, explained additional variation in the logistic regression model. The results
emphasise the need for spatially explicit analysis, both in examining fish habitat
relationships and in developing conservation plans for declining fish populations.
 
 
Torgersen, C. E., et al. (2004). Pattern detection in stream networks: quantifying spatial variability in fish distribution. GIS/Spatial Analyses in Fishery and Aquatic Sciences. T. Nishida, P. J. Kailola and C. E. Hollingworth. Saitama, Japan, Fishery-Aquatic GIS Research Group. 2: 405-420.
Biological and physical properties of rivers and streams are inherently
difficult to sample and visualize at the resolution and extent necessary to
detect fine-scale distributional patterns over large areas. Satellite imagery
and broad-scale fish survey methods are effective for quantifying spatial
variability in biological and physical variables over a range of scales in
marine environments but are often too coarse in resolution to address
conservation needs in inland fisheries management. We present methods
for sampling and analyzing multiscale, spatially continuous patterns of
stream fishes and physical habitat in small- to medium-size watersheds
(500–1000 hectares). Geospatial tools, including geographic information
system (GIS) software such as ArcInfo dynamic segmentation and
ArcScene 3D analyst modules, were used to display complex biological
and physical datasets. These tools also provided spatial referencing
information (e.g. Cartesian and route-measure coordinates) necessary for conducting geostatistical analyses of spatial patterns (empirical
semivariograms and wavelet analysis) in linear stream networks.
Graphical depiction of fish distribution along a one-dimensional
longitudinal profile and throughout the stream network (superimposed on
a 10-metre digital elevation model) provided the spatial context necessary
for describing and interpreting the relationship between landscape pattern
and the distribution of coastal cutthroat trout (Oncorhynchus clarki clarki)
in western Oregon, U.S.A. The distribution of coastal cutthroat trout was
highly autocorrelated and exhibited a spherical semivariogram with a
defined nugget, sill, and range. Wavelet analysis of the main-stem
longitudinal profile revealed periodicity in trout distribution at three nested
spatial scales corresponding ostensibly to landscape disturbances and
the spacing of tributary junctions.
 
Torgersen, C. E., et al. (2008). Spatial identification of tributary impacts in river networks. River Confluences, Tributaries and the Fluvial Network. S. P. Rice, A. G. Roy and B. L. Rhoads, John Wiley & Sons, Ltd.: 159-181.
Torgersen, C. E., et al. (1999). "Multiscale thermal refugia and stream habitat associations of chinook salmon in northeastern Oregon." Ecological Applications 9(1): 301-319.
We quantified distribution and behavior of adult spring chinook salmon
(Oncorhynchus tshawytscha) related to patterns of stream temperature and physical habitat
at channel-unit, reach-, and section-level spatial scales in a wilderness stream and a disturbed
stream in the John Day River basin in northeastern Oregon. We investigated the effectiveness
of thermal remote sensing for analyzing spatial patterns of stream temperature and assessed
habitat selection by spring chinook salmon, evaluating whether thermal refugia might be
responsible for the persistence of these stocks in rivers where water temperatures frequently
exceed their upper tolerance levels (258C) during spawning migration. By presenting stream
temperature and the ecology of chinook salmon in a historical context, we could evaluate
how changes in riverine habitat and thermal spatial structure, which can be caused by land-use
practices, may influence distributional patterns of chinook salmon. Thermal remote
sensing provided spatially continuous maps of stream temperature for reaches used by
chinook salmon in the upper subbasins of the Middle Fork and North Fork John Day River.
Electivity analysis and logistic regression were used to test for associations between the
longitudinal distribution of salmon and cool-water areas and stream habitat characteristics.
Chinook salmon were distributed nonuniformly in reaches throughout each stream. Salmon
distribution and cool water temperature patterns were most strongly related at reach-level
spatial scales in the warm stream, the Middle Fork (maximum likelihood ratio: P , 0.01),
and most weakly related in the cold stream, the North Fork (P . 0.30). Pools were preferred
by adult chinook salmon in both subbasins (Bonferroni confidence interval: P # 0.05);
however, riffles were used proportionately more frequently in the North Fork than in the
Middle Fork. Our observations of thermal refugia and their use by chinook salmon at
multiple spatial scales reveal that, although heterogeneity in the longitudinal stream tem-perature
profile may be viewed as an ecological warning sign, thermal patchiness in streams
also should be recognized for its biological potential to provide habitat for species existing
at the margin of their environmental tolerances.
 
Torgersen, C. E., et al. (2001). "Airborne thermal remote sensing for water temperature assessment in rivers and streams." Remote Sensing and the Environment 76: 386-398.
Torgersen, C. R., et al. (1999). "Multiscale thermal refigua and stream habitat associations of chinook salmon in northeastern Oregon." Ecological Applications 9: 301-319.
Torres, R., et al. (1998). "Unsaturated zone processes and the hydrologic response of a steep, unchanneled catchment." Water Resources Research 34(8): 1865-1879.
Torri, D., J. Poesen, and L. Borselli (1997). "Predictability and uncertainty of the soil erodibility factor using a global dataset." Catena 31(1-22).
Toth, S. (1991). A Road Damage Inventory for the Upper Deschutes River Basin, submitted to the SHAW Committee: 20.
Toth, S. E. (1991). An Analysis of Hydrologic Cumulative Effects in a Large Watershed in Western Washington. Seattle, Washington, University of Washington: 84.
Townsend, C. R. (1989). "The patch dynamics concept of stream community ecology." J. North Am. Benthol. Society 8: 36-50.
Townsend, C. R. (1996). "Concepts in river ecology: pattern and process in the catchment hierarchy." Archiv für Hydrobiologie 113((Suppl)): 3-21.
Townsend, C. R. and A. G. Hildrew (1994). "Species traits in relation to a habitat templet for river systems." Freshwater Biology 31: 265-275.
Trainor, K. and M. Church (2003). "Quantifying variability in stream channel morphology." Water Resources Research 39(9): 1248.
Nine stream channel characteristics (channel unit frequency, channel unit length, pool spacing, depth variability, width variability, large woody debris jam spacing, large woody debris volume, relative roughness, and average bank-full width used as a scale) were measured in 12 reaches in old growth forests on Haida Gwaii and Vancouver Island. They are applied to calculate a Euclidean distance measure of dissimilarity between all possible reach pair combinations. Frequency distributions of the resulting dissimilarity values express the range of variability present in the streams analyzed and enable definition of ranges of favorable and unfavorable comparisons. Reach pairs exhibiting high dissimilarity values have significant differences in several key stream channel characteristics that vary between reach pairs. Those reaches consistently appearing in reach pairs with high dissimilarity values exhibit significant variance from the norm for the group. Dissimilarity distributions provide a basis for appraising the outcome of stream channel manipulation (for example, in channel “restoration” programs) and for selecting channel pairs that are sufficiently similar to act as treatment and control units in experimental manipulations.
 
Trent, R., A. Molinas and N. Gagarin (1993). An artificial neural network for computing sediment transport. Hydraulic Engineering 1993, San Francisco, CA, American Society of Civil Engineers.
Tribe, A. (1992). "Automated recognition of valley lines and drainage networks from grid digital elevation models: a review and a new method." Journal of Hydrology 139: 263-293.
Trillium (1993). Canyon Lake Creek and Kenney Creek Watershed Assessment. Bellingham, WA.
Trimble, S. W. (1977). "The fallacy of stream equilibrium in contemporary denudation studies." American Journal of Science 277: 876-887.
Trimble, S. W. (1983). "Changes in sediment storage in the Coon Creek basin, Driftless Area, Wisconsin, 1853 to 1975." Science 214: 181-183.
Trimble, S. W. (1983). "A sediment budget for Coon Creek Basin in the driftless area, Wisconsin, 1853-1977." American Journal of Science 283: 454-474.
Trimble, S. W. (1999). "Decreased Rates of Alluvial Sediment Storage in the Coon Creek Basin, Wisconsin, 1975-93." Science 285: 1244-1246.
Triska, F. J. (1984). "Role of wood debris in modifying channel geomoprhology and riparian areas of a large lowland river under pristine conditions: a historical case study." Verh. Internat.Verein. Limnol. 22: 1876-1892.
Troendle, C., and R.M. King (1985). "The effect of timber harvest on the Fool Creek Watershed." Water Resources Research 21(12): 1915-1922.
Troendle, C., M.S. Wilcox, G.S. Bevenger, and L.S. Porth (2001). "The Coon Creek Water Yield Augmentation Project: implementation of timber harvesting technology to increase streamflow." Forest Ecology and Management 143(179-189).
Troendle, C. A., and R.M.King (1986). "The effect of timber harvest on the Fool Creek watershed, 30 years layer." Water Resources Research 21(12): 1915-1922.
Troendle, C. A., and R.M.King, (1987). "The effect of partial and clearcutting on streamflow at Deadhorse Creek, Colorado." Journal of Hydrology 90: 145-157.
Troendle, C. A., and W.K. Olsen (1994). Potential effects of timber harvest and water management on streamflow dynamics and sediment transport. Sustainable Ecological Systems: Implementing an Ecological Approach to Land Management. Fort Collins, CO, USDA Forest Service, Rocky Mountain Forest and Range Experiment Station: p.34-40.
Troendle, C. A., L.H. MacDonald, and C.H. Luce (2006). Fuels management and water yield. CWE for fuels management in the Western U.S. W. J. Elliot, and L.J. Audin.
Truman, C. C. and J. M. Bradford (1993). "Relationships between Rainfall Intensity and the Interrill Soil Loss-Slope Steepness Ratio as Affected by Antecedent Water-Content." Soil Science 156(6): 405-413.
Trush, W. (?). "A step in the right direction "?
Trustrum, N. A. and R. C. DeRose (1988). "Soil depth - age relationship of landslides on deforested hillslopes, Taranaki, New Zealand." Geomorphology 1: 143-160.
Tschaplinski, P. J. (1999). The effects of forest harvesting, fishing, climate variation, and ocean conditions on salmonid populations of Carnation Creek, Vancouver Island, British Columbia. Sustainable fisheries management: Pacific salmon. E. E. a. f. c.-e. Knudsen. Boca Raton and New York, Lewis Publishers. Chapter 19: 297-313.
Tschaplinski, P. J. and G. F. Hartman (1983). "Winter Distribution of Juvenile Coho Salmon (Oncorhynchus kisutch) Before and After Logging in Carnation Creek, British Columbia, and Some Implications for Overwinter Surivival " Canadian Journal of Fisheries and Aquatic Sciences 40: 452-461.
Tsukamoto, Y. (1987). A Collection of Reprints. Fuchu, Tokyo, Tokyo University of Agriculture and Technology.
Tsukamoto, Y. and H. Minematsu (1987). Evaluation of the effect of lateral roots on slope stability. Symposium on Forest Hydrology and Watershed Management at XIX IUGG General Assembly, Vancouver, British Columbia.
Tsukamoto, Y., et al. (1982). Hydrological and geomorphological studies of debris slides on forested hillslopes in Japan. Recent Developments in the Explanation and Prediction of Erosion and Sediment Yield, I.A.H.S.
Tsukamoto, Y., et al. (1982). "Importance of pipeflow in Hillslope Hydrology."?: 281-282.
Tubbs, D. W. and T. Dunne (1977). Geologic Hazards in Seattle (DRAFT). A field guide for The Geologic Society American Meeting: 1-14.
Tucker, G., ST Lancaster, NM Gasparini, RL Bras, SM Rybarczyk (2001). "An object-oriented framework for distributed hydrologic and geomorphic modeling using triangulated irregular networks." Computers and Geosciences 27(2001): 959-973.
Tucker, G. E. and R. L. Bras (1998). "Hillslope processes, drainage density, and landscape morphology." Water Resources Research 34(10): 2751-2764.
Tucker, G. E., et al. (2001). "Statistical analysis of drainage density from digital terrain data." Geomorphology 36: 187-202.
Drainage density Dd., defined as the total length of channels per unit area, is a fundamental property of natural terrain
that reflects local climate, relief, geology, and other factors. Accurate measurement of Dd is important for numerous
geomorphic and hydrologic applications, yet it is a surprisingly difficult quantity to measure, particularly over large areas.
Here, we develop a consistent and efficient method for generating maps of Dd using digital terrain data. The method relies
on  i. measuring hillslope flow path distance at every unchanneled site within a basin, and  ii. analyzing this field as a
random space function. As a consequence, we measure not only its mean  which is half the inverse of the traditional
definition of drainage density. but also its variance, higher moments, and spatial correlation structure. This yields a
theoretically sound tool for estimating spatial variability of drainage density. Averaging length-to-channel over an
appropriate spatial scale also makes it possible to derive continuous maps of Dd and its spatial variations. We show that the
autocorrelation length scale provides a natural and objective choice for spatial averaging. This mapping technique is applied
to a region of highly variable Dd in the northern Apennines, Italy. We show that the method is capable of revealing
large-scale patterns of variation in Dd that are correlated with lithology and relief. The method provides a new and more
general way to quantitatively define and measure Dd, to test geomorphic models, and to incorporate Ddvariations into
regional-scale hydrologic models
 
Tucker, G. E. and R. Slingerland (1997). "Drainage basin responses to climate change." Water Resources Research 33(8): 2031-2047.
Tuolumne River Stakeholders Group and Tuolumne River Technical Advisory Committee (1997). Tuolumne River Floodway Emergency Repair and Long-term Habitat Restoration Project Proposal Arcata, California, Tuolumne River Stakeholders Group and Tuolumne River Technical Advisory Committee prepared for CALFED: 10.
Turcotte, D. L. (1992). Fractals and Chaos in Geology and Geophysics. Cambridge, Cambridge University Press.
Turner, K. A. and R. L. Schuster, Eds. (1996). Landslides Investigation and Mitigation, Transportation Research Board Special Report 247. Washington, D.C., National Academy Press.
Turner, M. G., et al. (1989). "Predicting across scales: theory development and testing." Landscape Ecology 3(3/4): 245-252.
Turner, M. G., et al. (1995). "Ecological Dynamics at Broad Scales: Ecosystems and landscapes." BioScience: S29-S35.
Turner, M. G., et al. (1993). "A revised concept of landscape equilibrium: Disturbance and stability on scaled landscapes." Landscape Ecology 8(3): 213-227.
Turner, T. R. (1995). Geomorphic Response of the Madison River to Point Sediment Loading at the Madison Slide, Southwest Montana. Earth Sciences. Bozeman, Montana State University: 99.
Turner, T. R. and W. W. Locke (1990). Spatial and temporal geomorphic response of the Madison River to point sediment loading; the Madison Slide, southwest Montana. Quaternary Geology of the Western Madison Range, Madison Valley, Tobacco Root Range and Jefferson Valley. R. D. Hall: 126-136.
Tuteja, N. K., Vaze, J., B. Murphy, and G. Beale (2004). CLASS: catchment scale multiple-landuse atmosphere soil water and solute transport. Technical Report 04/12. C. R. C. f. C. Hydrology. New South Wales, Australia: 53pp.
Twito, R. H., et al. (1987). The MAP Program, United States Department of Agriculture
Forest Service
Pacific Northwest
Research Station: 1-22.
U. S. D. A., U. S. F. S. (1994). Standards and Guidelines for management of habitat for late successional and old growth forest related species within the range of the northern spotted owl., U. S. Forest Service.
U. S. E. P. A. (2000). "Watershed Analysis and Management Guide (WAM)."
U. S. E. P. A. (2001). Trinity River total maximum daily load for sediment. Region IX, U. S. Environmental Protection Agency.
U. S. Environmental Protection Agency 1999 "Van Duzen River and Yager Creek Total Maximum Daily Load for Sediment, Region IX.".
U. S. Geological Survey (1997). Stream and Ground-Water Monitoring Program, Lake Tahoe Basin, Nevada and California U. S. Geological Survey: 1-6.
U. S. National Marine Fisheries Service (1998). A draft proposal concerning Oregon Forest Practices submitted to the Oregon Board of Forestry. Memorandum of Agreement advisory committee and the office of the Govenor. Portland, Oregon, National Marine Fisheries Service.
U.S. Army Corps of Engineers (1987). "Training at HEC." Advances in Hydrologic Engineering Information Exchange Bulletin 8(2): 9.
U.S. Bureau of Land Management (2000). Bureau of Land Management Science Strategy. Denver, Colorado, Bureau of Land Management: 1-19.
U.S. Geological Survey (1993). Digital Elevation Models: Data User's Guide. Reston, Virginia, U.S. Geological Survey: 50.
Underwood, J. and R. E. Crystal (2002). "Hydrologically enhanced, high-resolution DEMs." A Supplement to Geospatial Solutions and GPS World: 8-14.
Unknown "Sedmodl v.2 technical documentation."
Untersteiner, N. (?). "Artic sea ice and climate change (DRAFT)."
Updike, R. G., et al. (1988). "A model for earthquake-induced translatory landslides in Quaternary sediments." Geological Society of America Bulletin 100: 783-792.
UPSAG (2006) Landslide Hazard Zonatino Project Protocol. 
Urban, D. L., et al. (1987). "Landscape ecology: A hierarchical perspective can help scientists understand spatial patterns." BioScience 37(2): 119-127.
Ursic, S. J., and J. E. Douglass (1979). The effects of forestry practices on water resources. Proceedings of the W. Kelly Mosley environmental forum., Auburn University Press, Auburn, AL.
Reviews and discusses the effects of management of southern forests on annual water yield, distribution of yield both monthly and seasonally, storm runoff, sediment yield, and water quality.
 
USDA and USDI (1994). Standards and Guidelines for Management of Habitat for Late-Successional and Old-Growth Forest Related Species Within the Range of the Northern Spotted Owl, Attachment A to the Record of Decision for Amendments to Forest Service and Bureau of Land Management Planning Documents Within the Range of the Northern Spotted Owl. Portland, Oregon.
USDA Forest Service (1962). Windthrown Timber Survey, USDA Forest Serivice: 1-15.
USDA Forest Service (1983). Plumas National Forest Soil Survey.
USDA Forest Service (1985). Proceedings of a Workshop on Slope Stability: Problems and Solutions in Forest Management, USDA Forest Service: 122.
USDA Forest Service (1986). Eldorado National Forest Soil Survey. Refers to conditions in 1984.
USDA Forest Service (1998). Proceedings of the Conference on Coastal Watersheds: The Caspar Creek Story Conference on Coastal Watersheds: The Caspar Creek Story, Ukiah, California.
USDA Forest Service (1999). Recent Publications of the Pacific Northwest Research Station, Third Quarter 1999, USDA Forest Service: 1-18.
USDA Forest Service (2002). Landscape dynamics and forest management. Fort Collins, CO, USA, U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station: 1 CD-ROM.
USDA Forest Service (2002). Tahoe National Forest Soil Survey CD version 2.0.
USDA Forest Service (2003). Landscape dynamics and forest management. Fort Collins, CO, USA, U.S. Department of Agriculture, Forest Service, Rocky Mountain Research Station: 1 CD-ROM.
USDA Forest Service (2004). Best Management Practices Evaluation Program 1992-2002 Monitoring Results., USDA Forest Service, PSW.
Valdes, J. B., et al. (1979). "A rainfall-runoff analysis of the geomorphologic IUH." Water Resources Research 15(6): 1421-1434.
Valett, H. M., et al. (1994). "Vertical hydrologic exchange and ecological stability of a desert stream ecosystem." Ecology 75: 548-560.
van Asch, T. W. J. and J. T. Buma (1997). "Modelling groundwater fluctuations and the frequency of movement of a landslide in the Terres Moires region of Barcelonnette (France)." Earth Surface Processes and Landforms 22: 131-141.
van Asch, T. W. J. and J. T. Buma (1997). "Modelling groundwater fluctuations and the frequency of movement of a landslide in the Terres Noires region of Barcelonnette (France)." Earth Surface Processes and Landforms 22: 131-141.
Van den Berg, J. H. and B. P. Bledsoe (2003). "Comment on Lewin and Brewer (2001): "Predicting channel patterns", Geomorphology 40, 329-339." Geomorphology 53(3-4): 333-337.
Van Rampaey, A. J. J., et al. (2001). "Modelling mean annual sediment yield using a distributed approach." Earth Surface Processes and Landforms 26(11): 1121-1236.
In this paper a spatially distributed model for the calculation of sediment delivery to river channels is presented (SEDEM: SEdiment DElivery Model). The model consists of two components: (1) the calculation of a spatial pattern of mean annual soil erosion rates in the catchment using a RUSLE (Revised Soil Erosion Equation) approach; and (2) the routing of the eroded sediment to the river channel network taking into account the transport capacity of each spatial unit. If the amount of routed sediment exceeds the local transport capacity, sediment deposition occurs. An existing dataset on sediment yield for 24 catchments in central Belgium was used to calibrate the transport capacity parameters of the model. A validation of the model results shows that the sediment yield for small and medium sized catchments (10-5000 ha) can be predicted with an average accuracy of 41 per cent. The predicted sediment yield values with SEDEM are significantly more accurate than the predictions using a lumped regression model. Moreover a spatially distributed approach allows simulation of the effect of different land use scenarios and soil conservation techniques.
 
Van Rompaey, A. "Development and evaluation of spatially distributed sediment delivery models."
Van Rompaey, A., G. Covers, and C. Puttemans (2002). "Modelling land use changes and their impact on soil erosion and sediment supply to rivers." Earth Surface Processes and Landforms 27: 481-494.
Van Rompaey, A., et al. (2003). "Modelling sediment supply to rivers and reservoirs in Eastern Europe during and after the collectivisation period." Hydrobiologia 494(1-3): 169-176.
At present, the landscapes in the former communist countries of Eastern Europe are undergoing a second major change in 50 years. After World War II, collective agricultural systems, such as the United Agricultural Cooperatives and the State Farms, were established. Following the example of the Soviet Union, individual fields were put together in order to enablemass production. This, however, led to an enormous acceleration of erosion and sediment transport processes on arable land. Since the fall of the communist regimes in 1990, farmers, or their successors, can claim back their original property. The large collective fields are split up in smaller spatial units. The land that is not claimed back is abandoned. This new transition has again a huge impact on soil erosion and sediment export to rivers and reservoirs. In order to control and optimize the transition from collective to private farming systems, land arrangement administrations need modeling tools to simulate the geomorphic impact of possible future land use scenarios. In this paper, SEDEM, a spatially distributed sediment delivery model, is calibrated and validated with measured sediment yield data from Czech drainage basins. The results suggest that SEDEM is able to predict the impact of land use changes on the mean annual sediment supply to rivers. Next, SEDEM is applied to assess the impact of a range of possible future landscape scenarios such as the splitting up of large fields in smaller spatial units and conversions from arable land to pasture.
 
Van Rompaey, A. J. J., et al. (2001). "Modelling mean annual sediment yield using a distributed approach." Earth Surface Processes and Landforms 26(11): 1221-1236.
Van Sickle, J., et al. (2004). "Projecting the biological condition of streams under alternative scenarios of human land use." Ecological Applications 14(2): 368-380.
We present regression models for estimating the status of fish and aquatic
invertebrate communities in all second to fourth–order streams (1:100 000 scale; total stream
length 5 6476 km) throughout the Willamette River Basin, Oregon (USA). The models
project fish and invertebrate status as a function of physiographic, land-use/land-cover, and
stream flow variables, with the latter two sets of variables subject to change under historical
and alternative future scenarios of human development. Models are developed using sample
data collected between 1993 and 1997 from 149 wadeable streams in the basin. Model
uncertainties are propagated through model projections and into aggregated estimates of
regional status. The projections show no significant change in basin-wide status in year
2050, relative to Circa 1990, for scenarios either of increased human development or
continuation of current development trends, because landscape change under these scenarios
is dominated by conversion of agricultural land to rural residential and urban uses, and
because these changes affect only a small percentage of the basin. However, under a scenario
of increased conservation, regional medians of biotic status indicators are projected to
improve by 9–24% by year 2050. None of the changes projected between Circa 1990 and
year 2050 is as large in magnitude as the decline in status projected to have occurred
between the time of pre-European settlement and Circa 1990.
 
Van Sickle, J. and S. V. Gregory (1990). "Modeling inputs of large woody debris to streams from falling trees." Canadian Journal of Forest Research 20: 1593-1601.
Van Sickle, J. and C. B. Johnson (2008). "Parametric distance weighting of landscape influence on streams." Landscape Ecology 23: 427-438.
We present a parametric model for estimating
the areas within watersheds whose land use best
predicts indicators of stream ecological condition. We
regress a stream response variable on the distanceweighted
proportion of watershed area that has a
specific land use, such as agriculture. Distance weighting
functions model the declining influence of
landscape elements as a function of their flowpath
distances, first to the stream channel (to-stream
distance), and then down the channel to the location
at which stream condition was sampled (in-stream
distance). Model parameters specify different distance
scales over which to-stream and in-stream influences
decline. As an example, we predict an index of biotic
integrity (IBI) for the fish communities in 50 small
streams of the Willamette Basin of Oregon, USA, from
distance-weighted proportions of agricultural or urban
land use in their watersheds. The weighting functions
of best-fitting models (R2 = 0.57) represent landscape
influence on IBI as extending upstream tens of
kilometers along the stream channel network, while
declining nearly to zero beyond a distance of 30 m
from the channel. Our example shows how parametric
distance weighting can identify the distance scales, and
hence the approximate areas within watersheds, for
which land use is most strongly associated with a
stream response variable. In addition, distance-weighting
parameters offer a simple and direct language for
comparing the scales of landscape influence on streams
across different land uses and stream ecosystem
components.
 
Van Steeter, M. M. and j. Pitlick (1998). "Geomorphology and endangered fish habitats of the upper Colorado River. 1. Historic changes in streamflow, sediment load, and channel morphology." Water Resources Research 34(2): 287-302.
Van Wagner, C. E. (1978). "Age-class distribution and the forest fire cycle." Can. J. For. Res. 8: 220-227.
van Westen, C. J., et al. (2003). "Use of geomorphological information in indirect landslide susceptibility assessment." Natural Hazards 30: 399-419.
The objective of this paper is to evaluate the importance of geomorphological expert
knowledge in the generation of landslide susceptibility maps, using GIS supported indirect bivariate
statistical analysis. For a test area in the Alpago region in Italy a dataset was generated at
scale 1:5,000. Detailed geomorphological maps were generated, with legends at different levels
of complexity. Other factor maps, that were considered relevant for the assessment of landslide
susceptibility, were also collected, such as lithology, structural geology, surficial materials, slope
classes, land use, distance from streams, roads and houses. The weights of evidence method was
used to generate statistically derived weights for all classes of the factor maps. On the basis of these
weights, the most relevant maps were selected for the combination into landslide susceptibility maps.
Six different combinations of factor maps were evaluated, with varying geomorphological input.
Success rates were used to classify the weight maps into three qualitative landslide susceptibility
classes. The resulting six maps were compared with a direct susceptibility map, which was made by
direct assignment of susceptibility classes in the field. The analysis indicated that the use of detailed
geomorphological information in the bivariate statistical analysis raised the overall accuracy of the
final susceptibility map considerably. However, even with the use of a detailed geomorphological
factor map, the difference with the separately prepared direct susceptibility map is still significant,
due to the generalisations that are inherent to the bivariate statistical analysis technique.
 
Vandaele, K., J. Poesen, G. Govers, and B van Wesemael (1996). "Geomorphic threshold conditions for ephemeral gully incision." Geomorphology 16: 161-173.
Vanderheyden, J., et al. (1988). Watershed and Fisheries Cumulative Effect Assessment. Grants Pass, OR, U. S. Department of Agriculture, Forest Service: 59.
Vandine, D. F. (1984). Debris flows and debris torrents in the Southern Canadian Cordillera. Victoria, B. C., Vandine Geological Engineering Services: 44-68.
Vannote, R. L. and G. W. Minshall (1982). "Fluvial processes and local lithology controlling abundance, structure, and composition of mussel beds." Proc. Natl. Acad. Sci. USA 79: 4103-4107.
Vannote, R. L., et al. (1980). "The river continuum concept." Canadian Journal of Fisheries and Aquatic Science 37: 130-137.
Vanotte, R. L., et al. (1980). "The River Continuum Concept." Canadian Journal of Fisheries and Aquatic Sciences 37(1): 130-137.
VanSickle, J., and R.L. Beschta (1983). "Supply-based models of suspended sediment transport in streams." Water Resources Research 19(3): 768-778.
VanSickle, J. (?). Temporal Variability in Stream Ecosystems.
Varnes, D. J. (1978). Slope Movement Types and Processes. Landslides: Analysis and Control. R. L. Schuster and R. J. Krizek. Washinton D.C. Chapter 2: 11-31.
Varnes, D. J. (1984). Landslide Hazard Zonation: A Review of Principles and Practice.
Veblen, T. T., et al. (1994). "Disturbance regime and disturbance interactions in a Rocky Mountain subalpine forest." Journal of Ecology 82: 125-135.
Veldhuisen, C. (2000). Preliminary results and recommendations from the northwest Cascades Type 4/5 stream study, Skagit System Cooperative: 6.
Veldhuisen, C. (2004). Summary of headwater perennial stream surveys in the Skagit and neighboring basins: 2001-2003, Skagit River System Cooperative: 10.
Veldhuisen, C. and P. Russell (1999). Forest Road Drainage and Erosion Initiation in Four West-Cascade Watersheds, prepared for the Timber/fish/Wildlife Monitoring Advisory Group and the Northwest Indian Fisheries Commission: 37.
Veldhuisen, C. N. (1990). Coarse woody debris in streams of the Drift Creek Basin, Oregon. Department of Forest Science. Corvallis, OR, Oregon State University.
Ventura, E., et al. "The study of detachment and deposition on a hillslope using a magnetic tracer." Catena 48: 149-161.
Soil erosion by water involves the processes of detachment, transport and deposition of soil materials by the erosive forces of raindrops and surface flow of water. The redistribution of sediment within a field-sized area is important in estimating the effect of erosion and deposition on productivity, in helping the conservation planner to target efforts to reduce erosion, and to evaluate erosion models. The objective of this study was to use a magnetic tracer, with size and density similar to soil aggregates, to study detachment and deposition on a hillslope. Two interconnected plots were established on a hillslope. Two rainfall intensities (35 and 70 mm h super(-1)) combined with two different inflow rates (4 and 10 1 min super(-1)) were applied to the upper of the two plots. No rain or water was applied to the lower plot, which was used to study the deposition of eroded sediments from the upper plot. A 5% concentration of magnetic tracer was placed in the upper plot and mixed to depth of 3 cm. From this initial condition, areas of tracer detachment and deposition were identified using a magnetic sensor. Areas of detachment were associated with a decrease in magnetic signal, while areas of deposition were associated with an increase in the magnetic signal. In the lower plot, deposition of tracer correlated well with the magnetic susceptibility readings. Results indicated that the tracer was effective for identifying areas of net detachment and deposition, however, the tracer to soil ratio did not remain constant for all treatments. For this reason, a wider range of sizes and densities of the tracer should be tested if the method is to be useful to quantify erosion rates.
 
Ver Hoef, J. M., et al. (2006). "Spatial statistical models that use flow and stream distance." Environmental and Ecological Statistics 13: 449-464.
We develop spatial statistical models for stream networks that can estimate
relationships between a response variable and other covariates, make predictions at
unsampled locations, and predict an average or total for a stream or a stream segment.
There have been very few attempts to develop valid spatial covariance models
that incorporate flow, stream distance, or both. The application of typical spatial
autocovariance functions based on Euclidean distance, such as the spherical covariance
model, are not valid when using stream distance. In this paper we develop a large
class of valid models that incorporate flow and stream distance by using spatial moving
averages. These methods integrate a moving average function, or kernel, against
a white noise process. By running the moving average function upstream from a location,
we develop models that use flow, and by construction they are valid models based
on stream distance. We show that with proper weighting, many of the usual spatial
models based on Euclidean distance have a counterpart for stream networks. Using
sulfate concentrations from an example data set, the Maryland Biological Stream
Survey (MBSS), we show that models using flow may be more appropriate than models
that only use stream distance. For the MBSS data set, we use restricted maximum
likelihood to fit a valid covariance matrix that uses flow and stream distance, and then
we use this covariance matrix to estimate fixed effects and make kriging and block
kriging predictions.
 
Verstraeten, G. and J. Poesen (2000). "Estimating trap efficiency of small reservoirs and ponds: methods and implications for the assessment of sediment yield."
Throughout the world, several millions of small ponds exist for water supply, irrigation, flood control or to control water quality downstream. The reduced flow velocity in these ponds causes sedimentation of tranported particles. For most ponds this is a negative impact as their retention capacity decreases due to sedimentation processes. Sediment volumes in small ponds can be used to reconstruct sediment yield values and to study the spatial variation in sediment yield over large areas. Especially in developing countries, this technique can be very helpful in establishing large data sets on sediment delivery as there are often no resources for expensive monitoring programmes. However, when such studies are undertaken, one has to take into account the efficiency of the pond in trapping sediments. This trap efficiency is dependent on the characteristics of the inflowing sediment and the retention time of the water in the pond, which in turn are controlled by pond geometry and runoff characteristics. Because trap efficiency is one of the most important properties of a pond or reservoir, it has been studied for quite some time. This article provides an overview of the different methods available to estimate the trap efficiency of reservoirs and ponds. The first set of methods are empirical models that predict trap efficiency, mostly of normally ponded large reservoirs using data on a mid to long-term basis. These models relate trap efficiency to a capacity/watershed ratio, a capacity/annual inflow ratio or a sedimentation index. Today, these models are the most widely used models to predict trap efficiency, even for reservoirs or ponds that have totally different characteristics from the reservoirs used in these models. For small ponds, these models seem to be less appropriate. They also cannot be used for predicting trap efficiency for a single event. To overcome these restrictions, different theoretical models have been developed based on sedimentation principles. These can be very simple, such as the overflow rate method, but also very complex when runoff and sediment are routed through a pond with incremental time-steps. The theoretical-based models are probably more capable of predicting trap efficiency for small ponds with varying geometric characteristics, and some of them also provide data on effluent sediment concentrations and quality. However, when reconstructing sediment yield values using sedimentation rates over a period of a few months to a few year (the mid-term basis), one needs a trap efficiency value for the whole period, not for one single event. At present only limited research has been done on establishing mid-term trap efficiency models based on theoretical principles. This is probably the most important gap in trap efficiency research.
 
Verstraeten, G., et al. (2002). "Evaluating an integrated approach to catchment management to reduce soil loss and sediment pollution through modelling." Soil Use and Management 18(4): 386-394.
Verstraeten G, a. J. P. (2002). "Using sediment deposits in small ponds to quantify sediment yield from small catchements: possibilities and limitations." Earth Surface Processes and Landforms 27: 1425-1439.
Vianello, A., et al. (2009). "LiDAR-derived slopes for headwater channel network analysis." Catena 76: 97-106.
Slope is one of the most important distinguishing features for channel morphology. Variations in the
computation of slope from a digital elevation model can affect a wide range of hydrogeomorphically derived
applications. We compare different methods for computing channel slope using LiDAR-derived digital terrain
models (DTMs) with varying resolutions. We chose a headwater basin of the Eastern Italian Alps,
characterized by a dense ephemeral colluvial network and a main alluvial channel as our study area. The
identified alluvial morphologies are characteristic of steep mountain streams, namely, cascades and step
pools. Field surveys were carried out along the main channel and in some small tributaries. Results indicate
that a single method for slope calculation cannot estimate channel slope at the hydrographic network scale.
The differential geometry approach for slope calculation tends to overestimate field-surveyed channel slope
values for all the DTM resolutions (1, 2, 5 m). When a trigonometric approach for slope calculation is applied,
2 and 5 m DTM resolutions give more consistent results. Nevertheless, a reliable channel slope can be derived
from a DTM with an appropriate resolution by choosing a suitable method only after considering the channel
width
 
Viers, S. D. (1975). "Redwood vegetation dynamics." Bulletin of the Association of Engineering Geologists 56: 34-35.
Viney, N. R., and M. Sivapalan (2004). "A framework for scaling of hydrologic conceptualizations based on a disaggregation-aggregationb approach." Hydrological Processes 18: 1395-1408.
Vogel, R. M., et al. (1999). "Regional regression models of annual streamflow for the United States." Journal of Irrigation and Drainage Engineering(May/June): 148-157.
Estimates of annual streamflow volumes are needed in many different types of hydrologic studies.
Usually a streamgauge is unavailable at the location of interest, hence regional methods that relate streamflow
to readily measured geomorphic and climate characteristics provide a practical solution. Hydrologic, geomorphic,
and climatic characteristics of 1,553 undeveloped watersheds across the United States are used to develop
regional regression equations that relate the first two moments of annual streamflow to readily measured basin
and climate characteristics. These relations are summarized for each of 18 major U.S. water resource regions.
The relationships are remarkably precise, with adjusted R2 values ranging from 90.2–99.8% and an average
value of 96.2% across the continent. The usefulness of these relationships is evaluated by deriving their information
content in terms of equivalent record length. These results indicate that regional models of annual
streamflow, including runoff maps, are less accurate than suggested by traditional goodness-of-fit statistics. We
also provide estimates of precipitation and temperature elasticity of streamflow, by region.
 
Voinov, A. A., et al. (2010). "A community approach to earth systems modeling." EOS, Transaction American Geophysical Union 91(13): 117-124.
Vronskii, B. B. and V. N. Leman (1991). "Spawning Stations, Hydrological Regime and Survival of Progeny in Nests of Chinook Salmon, Oncorhynchus tshawytscha in the Kamchatka River Basin " Voprosy ikhtiologii 31(2): 282-291.
Waitt, R. B., Jr. and R. M. Thorson (1983). The Cordilleran Ice Sheet in Washington, Idaho, and Montana. Lake-Quaternary Environments of the United States. S. C. Porter, University of Minnesota Press. Vol I: The Late Pleistocene: 53-70.
Walder, J. P. and G. R. Willgoose (1999). "On the effect of digital elevation model accuracy on hydrology and geomorphology." Water Resources Research 35(7): 2259-2268.
Waldron, H. H. (?). "Debris flow and erosion control problems caused by the ash eruptions of Irazu volcano, Coasta Rica." U.S. Geological Survey: I1-I37.
Waldrop, M. M. (1992). Complexity: The Emerging Science at the Edge of Order and Chaos. New York, Simon and Schuster.
Walker, A. (1960). Water resources of the Nooksack River Basin and certain adjacent streams, Washington State Department of Ecology.
Walker, G. W. and N. S. MacLeod (1991). Oregon Geology. Reston, VA, U.S. Department of the Interior, Geological Survey.
Walker, J. P. and G. R. Willgoose (1999). "On the effect of digital elevation model accuracy on hydrology and geomorphology." Water Resources Research 35(7): 2259-2268.
This study compares published cartometric and photogrammetric digital
elevation models (DEMs) of various grid spacings with a ground truth data set, obtained
by ground survey, and studies the implications of these differences on key hydrologic
statistics. Inferred catchment sizes and stream networks from published DEMs were found
to be significantly different than those from the ground truth in most instances.
Furthermore, the width functions and cumulative area relationships determined from the
published DEMs were found to fall consistently outside the 90% confidence limits
determined from the ground truth for more than 60% of the relationship, suggesting that
these hydrologic properties are poorly estimated from published DEMs. However, the
slope-area relationships determined from published DEMs were found to be less sensitive
to catchment shape, size, and stream network, with the relationship falling outside the
90% confidence limits for less than 40% of the relationship for all catchments identified
from the published DEMs. A published relationship linking the horizontal resolution with
the vertical accuracy of the DEM was tested, predicting a horizontal resolution of about
10 m for the published DEMs tested.
 
Walker, L. R., et al. (1986). "The role of life history processses in primary succession on an Alaskan floodplain." Ecology 67: 1508-1523.
Wallace, J. B. and A. C. Benke (1984). "Quantification of Wood Habitat in Subtropical Coastal Plain Streams." Canadian Journal of Fisheries and Aquatic Sciences 41: 1643-1652.
Wallace, J. B., et al. (2001). "Small woody debris dynamics in a headwater stream." Verh. Internat. Verein. Limnol. 27: 1361-1365.
Wallbrink, P. J., et al. (2002). "A tracer budget quantifying soil redistribution on hillslopes after forest harvesting." Catena 47(3): 179-201.
Managing the impacts of erosion after forest harvesting requires knowledge of erosion sources; rates of sediment transport and storage; as well as losses from the system. We construct a tracer-based ( (super 137) Cs) sediment budget to quantify these parameters. The budget shows significant redistribution, storage and transport of sediment between landscape elements and identifies the snig tracks and log landings as the major impact sites in the catchment. Annual sediment losses from them were estimated to be 25+ or -11 and 101+ or -15 t ha (super -1) year (super -1) , respectively, however, it is probable that most of this is due to mechanical displacement of soil at the time of harvesting. The budget showed greatest net transport of material occurring from snig tracks; representing some 11+ or -4% of the (super 137) Cs budget. Of the latter amount, 18%, 28% and 43% was accounted for within the cross banks, filter strip and General Harvest Area (GHA), respectively. The (super 137) Cs budget also showed the GHA to be a significant sediment trap. The filter strip played a fundamental role in the trapping of material generated from the snig tracks, the mass delivery to them from this source was calculated to be 1.7+ or -0.6 kg m (super 2) year (super -1) . Careful management of these remains critical. Overall we could account for 97+ or -10% of (super 137) Cs. This retention suggests that (within errors) the overall runoff management system of dispersing flow (and sediment) from the highly compacted snig tracks, by cross banks, into the less compacted (and larger area) GHA and filter strips has effectively retained surface soil and sediment mobilised as a result of harvesting at this site.
 
Wallick, J. R., et al. (2006). "Determination of bank erodibility for natural and anthropogenic bank materials using a model of lateral migration and observed erosion along the Willamette River, Oregon, USA." River Research and Applications 22: 631-649.
Many large rivers flow through a variety of geologic materials.Within the span of several kilometres, bends may alternately flow
against recently reworked sediments, older, more indurated sediments or highly resistant materials. As sediment size,
cementation, and other properties strongly influence the erodibility of river banks, erosion rates and channel planform are
likely to vary significantly along the length of large rivers. In order to assess the role of bank materials on bank erosion rates, we
develop a method for detecting relative differences in erodibility between bank materials along large floodplains. By coupling
historic patterns of channel change with a simple model of bank erodibility we are able to track relative changes in bank
erodibility among time intervals and bank materials. We apply our analysis to the upper Willamette River, in northwestern
Oregon for three time periods: 1850–1895, 1895–1932 and 1972–1995 and compute relative differences in bank erodibility for
Holocene alluvium, partially cemented Pleistocene gravels, and revetments constructed in the 20th century. Although the
Willamette is fundamentally an anastomosing river, we apply the model to single-thread portions of the channel that evolved
through lateral migration. Our simple model of bank erodibility reveals that for all three-time periods, banks composed of
Holocene alluvium are at least 2–5 times more erodible than banks composed of Pleistocene gravels. Revetment installed in the
20th century is highly resistant to erosion and is at least 10 times less erodible than Pleistocene gravels.
 
Walling, D. (1999). "Linking land use, erosion, and sediment yields in river basins." Hydrobiologia 410: 223-240.
Abstract  Results obtained from erosion plots and catchment experiments provide clear evidence of the sensitivity of erosion rates to land use change and related human activity. Evidence for the impact of land use on the sediment yields of world rivers is less clear, although examples of rivers where sediment yields have both increased and decreased in recent decades can be identified. The apparent lack of sensitivity of river sediment loads to land use change reflects, at least in part, the buffering capacity associated with many river basins. This buffering capacity is closely related to the sediment delivery ratio of a river basin, in that basins with high sediment delivery ratios are likely to exhibit a reduced buffering capacity. Investigations of the impact of land use and related human activity on sediment yields should consider the overall sediment budget of a catchment rather than simply the sediment output. Information on the sediment budget of a drainage basin is difficult to assemble using traditional techniques, but recent developments in the application of fingerprinting techniques to establish sediment sources and in the use of environmental radionuclides, such as caesium-137 and lead-210, to document sediment storage offer considerable potential for providing such information. Sediment storage within a river basin can give rise to environmental problems where sediment-associated pollutants accumulate in sediment sinks. The accumulation of phosphorus on river floodplains as a result of overbank sedimentation can, for example, represent an important phosphorus sink.
 
Walling, D. E. (1983). "The sediment delivery problem." Journal of Hydrology 65(1983): 209-237.
Walling, D. E., et al. (2003). "Using (CS)-C-137 measurements to validate the application of the AGNPS and ANSWERS erosion and sediment yield models in two small Devon catchments." Soil & Tillage Research 69(1-2): 27-43.
Walling, D. E., et al. (2002). "Establishing sediment budgets for two small lowland agricultural catchments in the UK." Catena 47(4): 323-353.
Walsh SJ, D. B., and GP Malanson (1998). "An overview of scale, pattern, process relationships in geomorphology: a remote sensing and GIS perspective." Geomorphology 21: 183-205.
Walters, C. and J. Korman (1999). "Cross-scale modeling of riparian ecosystem responses to hydrologic management." Ecosystems 2: 411-421.
Wan, Y. and S. El-Swaify (1998). "Characterizing interrill sediment size by partitioning splash and wash processes." Soil Sci Soc Am J 62(2): 430-437.
Wan, Y., et al. (1996). "Partitioning interrill splash and wash dynamics: A novel laboratory approach." Soil Technology 9(1-2): 55-69.
Wang, X. and Z. Wang (1999). "Effect of land use change on runoff and sediment yield." International Journal of Sediment Research 14(4): 37-44.
Runoff and sediment yield are greatly affected by the type of land use. The paper studies the influence of agriculture, husbandry, mining and urban development, demonstrates the law of variation in sediment yield during the processes of land use changes. Land use changes from forest to agriculture, grassland to farm field and slope plough land to terrace field exhibit the same law influencing the sediment yield. During the transforming period the sediment yield is higher than those of either land used and is proved to be equal to he sum of the sediment yields of the two types of land use. This is named the superposition law of land use transformation. The correlation between the percentage of forest coverage with the runoff and sediment yields indicates that the runoff yield reduce to the minimum and the sediment yield is nearly zero if the coverage is higher than 60%. Whereas the runoff is doubled and sediment is tripled for 10% reduction in the forest coverage if the coverage is between 30-60%. Mining and urbanization increase the sediment yield in the development period.
 
Waples, R. S. (1991). "Genetic Interactions Between Hatchery and Wild salmonids: Lessons From the Pacific Northwest." Canadian Journal of Fisheries and Aquatic Sciences 48(Suppl. 1): 124-133.
Waples, R. S., et al. (2009). "Evolutionary history, habitat disturbance regimes, and anthropogenic changes: what do these mean for resilience of Pacific Salmon populations?" Ecology and Society 14(1).
Because resilience of a biological system is a product of its evolutionary history, the historical
template that describes the relationships between species and their dynamic habitats is an important point
of reference. Habitats used by Pacific salmon have been quite variable throughout their evolutionary history,
and these habitats can be characterized by four key attributes of disturbance regimes: frequency, magnitude,
duration, and predictability. Over the past two centuries, major anthropogenic changes to salmon ecosystems
have dramatically altered disturbance regimes that the species experience. To the extent that these
disturbance regimes assume characteristics outside the range of the historical template, resilience of salmon
populations might be compromised. We discuss anthropogenic changes that are particularly likely to
compromise resilience of Pacific salmon and management actions that could help bring the current patterns
of disturbance regimes more in line with the historical template
 
Ward, F. R. (1977). Prescribed underburning trials on second-growth Douglas-fir, USDA Forest Service.
Ward, J. a. C. J. (2004). "Sediment trapping within forestry streamside management zones: Georgia Piedmont, USA." Journal of the American Water Resources Association 40(6): 1421-1431.
Ward, J. V., et al. (2002). "Riverine landscape diversity " Freshwater Biology 47: 517-539.
Ward, T. J., et al. (1981). Use of a mathematical model for estimating potential landslide sites in steep forested drainage basins. Erosion  and Sediment Transport in Pacific Rim Steeplands. Christchurch, New Zealand, International Association of Hydrological Sciences. Publ. No. 132: 21-41.
Ward, T. J. and J. S. O'Brien (1981). Hydraulic and rheologic modelling of mud and grain flows. Erosion and Sediment Transport in Pacific Rim Steeplands. Christchurch, International Association of Hydrological Studies. Publ. No. 132: 269-279.
Waring, P. W. and W. H. Schlesinger (1985). Susceptibility and Response of Forests to Natural Agents of Disturbance. Forest Ecosystems and Management, Academic Press. Chapter 9: 211-235.
Warren, S. D., et al. (2004). "An evaluation of methods to determine slope using digital elevation data." Catena 58: 215-233.
Variation in the computation of slope from digital elevation data can result in significantly
different slope values and can, in turn, lead to widely varying estimates of environmental phenomena
such as soil erosion that are highly dependent on slope. Ten methods of computing slope from
distributed elevation data, utilizing capabilities inherent in five different geographic information
systems (GIS), were compared with field measurements of slope. The methods were compared based
on (1) overall estimation performance, (2) estimation accuracy, (3) estimation precision, and (4)
independence of estimation errors and the magnitude of field measured slopes. A method utilizing a
very high resolution digital elevation model (DEM) (1 m) produced slightly better estimates of slope
than approaches utilizing somewhat lower resolution DEMs (2–5.2 m), and significantly better
estimates than a method utilizing a 12.5 m DEM. The more accurate method was significantly
biased, however, frequently underestimating actual slope. Methods that averaged or smoothed high
resolution DEMs over larger areas also produced good estimates of slope, but these were somewhat
less accurate in areas of shallow slopes. Methods utilizing differential geometry to compute percent
slope from DEMs outperformed methods utilizing trigonometric functions. Errors in slope
computation are exaggerated in soil erosion prediction models because erosion typically increases
as a power function of slope
 
Washington Department of Natural Resources (2005). Chapter 222-10, Washington Administrative Code: State Environmental Policy Act Guidelines. Olympia, Washington, Washington Dep. of Natural Resources.
Washington Forest  Practices Board (1997). Board Manual: Standard Methodology for Conducting Watershed Analysis. Olympia, Washington, Washington Department of Natural Resources.
Washington Forest Practices Board (WFPB) (1994). "Standard Methodology for Conducting Watershed Analysis under Chapter 222-22 WAC. Version 2.1."
Washington State Forest Practices Board (2004). Section 2: Standard methods for identifying bankfull channel features and channel migration zones. Forest Practices Board Manual. Olympia, Washington, Washington State Department of Natural Resources.
Washington_Forest_Practices_Board (1997). Board Manual: Standard Methodology for Conducting Watershed Analysis. Olympia, Washington Department of Natural Resources.
Watanabe, M. and H. Ikeya (1981). Investigation and analysis of volcanic mud flows on Mt. Sakurajima, Japan. Erosion and Sediment Transport Measurement, Christchurch, IAHS.
Water Management Bureau (1999). Historic Channel Changes and Geomorphology of the upper Yellowstone River (DRAFT). Helena, Montana, Water Management Bureau: Montana Department of Natural Resources and Conservation: 45.
Wathen, S. J. and T. B. Hoey (1998). "Morphological controls on the downstream passage of a sediment wave in a gravel-bed stream." Earth Surface Processes and Landforms 23: 715-730.
Watson, C. C., et al. (2002). "Use of incised channel evolution models in understanding rehabilitation alternatives." Journal of the American Water Resources Association 38(1): 151-160.
Incised channels are caused by an imbalance between sediment transport capacity and sediment supply to the stream. The resulting bed and bank erosion alter channel morphology and stability. Geomorphological models of incised channel evolution can provide guidance in the selection of engineering design alternatives for incised channel rehabilitation. This paper describes how incised channel evolution models may be coupled with a dimensionless stability diagram to facilitate evaluation of rehabilitation alternatives. In combination, the models provide complementary views of channel processes from geomorphic and engineering perspectives.
 
Watters, R. J. (1983). "A Landslide Induced Waterflood-Debris Flow." Bulletin of the Association of Engineering Geology 28: 177-182.
Watters, R. J. and W. D. Delahaut (1995). Effect of argillic alteration on rock mass stability. Clay and Shale Slope Instability. W. C. hanebert and S. A. Anderson. Boulder, Colorado, Geological Society of America. 10: 139-150.
Waythomas, C. F., and G.P. Williams (1988). "Sediment yield and spurious correlation - toward a better portrayal of the annual suspended sediment load of rivers." Geomorphology 1: 309-316.
Weaver, W., et al. (1998). Aerial reconnaissance evaluation of recent storm effects on upland mountainous watersheds of Ohio: Wildland response to recent storms and floods in the Clearwater, Lochsa and Boise River Watersheds, Idaho. Arcata, California, Pacific Watershed Associates: 42.
Weaver, W. and D. K. Hagans (1996). Aerial reconnaissance evaluation of 1996 storm effects on upland mountainous watersheds of Oregon and southern Washington: wildland response to the February 1996 storm and flood in the Oregon and Washington Cascades and Oregon Coast Range Mountains. Arcata, California, prepared for The Pacific Rivers Council: 2-22.
Webb, R. H., et al. (1988). "Monument Creek debris flow, 1984: Implications for formation of rapids on the Colorado River in Grand Canyon National Park." Geology 16: 50-54.
Webb, T. I. (?). "What past climates can tell us about a future warmer earth."
Wechsler, S. P. (2007). "Uncertainties associated with digital elevation models for hydrologic applications: a review." Hydrology and Earth System Sciences 11: 1481-1500.
Digital elevation models (DEMs) represent the topography
that drives surface flow and are arguably one of
the more important data sources for deriving variables used
by numerous hydrologic models. A considerable amount of
research has been conducted to address uncertainty associated
with error in digital elevation models (DEMs) and the
propagation of error to derived terrain parameters. This review
brings together a discussion of research in fundamental
topical areas related to DEM uncertainty that affect the
use of DEMs for hydrologic applications. These areas include:
(a) DEM error; (b) topographic parameters frequently
derived from DEMs and the associated algorithms used to derive
these parameters; (c) the influence of DEM scale as imposed
by grid cell resolution; (d) DEM interpolation; and (e)
terrain surface modification used to generate hydrologicallyviable
DEM surfaces. Each of these topical areas contributes
to DEM uncertainty and may potentially influence results of
distributed parameter hydrologic models that rely on DEMs
for the derivation of input parameters. The current state of
research on methods developed to quantify DEM uncertainty
is reviewed. Based on this review, implications of DEM uncertainty
and suggestions for the GIS research and user communities
are offered.
 
Weddle, T. K., et al. (1993). "The Late-Glacial Marine Invasion of Coastal Central New england (Northeastern Massachusetts - Southwestern Maine): It's Ups and Downs "?: 1-1 - 1-28.
Weinberg, G. M. (1975). An Introduction to General Systems Thinking. New York, Wiley-Interscience.
Weins, J. A. (2002). "Riverine landscapes: taking landscape ecology into the water." Freshwater Biology 47: 501-515.
Weins, J. A., et al. (1993). "Ecological mechanisms and landscape ecology." Oikos 66: 369-380.
Weir, G. J. (1983). "One dimensional bed wave movement in lowland rivers." Water Resources Research 19(3): 627-631.
Weirich, F. H. (1987). Sediment transport and deposition by fire-related debris flows in Southern California. IAHS-AISH Publication, vol.165. R. L. Beschta, T. Blinn, G. E. Grant, G. G. Ice and F. J. Swanson: 283-284.
Weirich, F. H., et al. (1996). "The evaluation of peak flow effects on high-gradient, low-order mountain stream channels using a field-scale experimental facility." Abstracts with Programs - Geological Society of America 28(7): 261.
Our understanding of hydrologic and sediment transport processes in low order, high slope, gravel bedded mountain streams is quite limited. Remoteness of sites, the infrequent nature of events and related logistical difficulties have left us with little integrated, quantitative data on the basic hydraulic, hydrologic and sediment transport processes in such environments. A research facility has been established within the Ouachita National Forest of Arkansas at which an artificial stormflow generating system has been put into operation. The system consisting of a combined pump and reservoir system provided a controlled release of water in a manner which enabled replication of stormflow conditions in an adjacent test study reach. The results of initial runs are presented. A series of five releases were undertaken to evaluate questions of sediment supply depletion and basic transport dynamics. Instrumentation employed involved integrated arrays of electromagnetic flow meters, suspended and bedload samplers, tagged rocks, and pre- and post-event channel and bank surveys at selected cross sections in the test reach.
 
Wells, J. T. (1987). "Effects of sea-level rise on deltaic sedimentation in south-central Louisiana." Journal of Sedimentary Research: 157-165.
Wells, J. T. (1996). Subsidence, Sea Level Rise, and Wetland Loss in the Lower Mississippi River Delta. Sea  Level Rise and Coastal Subsidence: Causes, Consequences and strategies. Dordrecht, Kluwer Academic: 281-311.
Wells, N. A., et al. (1990). "Stonelines and landscape development on the laterized craton of Madagascar." Geological Society of America Bulletin 102: 615-627.
Welty, J. J., et al. (2002). "Riparian aquatic interaction simulator (RAIS): a model of riparian forest dynamics for the generation of large woody debris and shade." Forest Ecology and Management 162: 299-318.
Streams depend on riparian forests to supply many important functions such as delivery of large woody debris (LWD),
organic matter, and nutrients into the aquatic ecosystem and to provide shade to help maintain cool water. Both forests and
streams are dynamic, and inputs from the riparian forest are constantly replenished as organic matter and nutrients are
processed and transported. Forest stand dynamics have been intensively studied and foresters have developed a good
understanding of tree growth and mortality to support commercial forest management. In recent years, the relationships
between aquatic functions and adjacent forest stand characteristics have been increasingly quantified by ecologists.
Management of riparian forests to protect aquatic functions and water quality has received considerable attention in the Pacific
Northwest and elsewhere. For scientific knowledge to be useful to decision-makers, the complex set of riparian relationships,
increasingly well understood individually, need to be collectively and objectively linked in a form that can be easily used by
scientists and non-scientists alike to develop management strategies for riparian forests. In this paper we describe an analytical
system that quantitatively links widely used forest growth forecasting systems for coastal Pacific Northwest forest types to the
riparian ecological functions of large woody debris recruitment and shade. The riparian aquatic interaction simulator (RAIS),
with its user-friendly interface, allows managers to forecast aquatic functions for up to 300 years. RAIS provides these
forecasts over a range of critical input variables and produces realistic estimates of riparian functions when compared with
published research. RAIS is available at http://www.weyerhaeuser.com/rais.html # 2002 Elsevier Science B.V. All rights
reserved.
 
Welty, J. W., et al. (2002). "Riparian aquatic interaction simulator (RAIS): a model of riparian forest dynamics for the generation of large woody debris and shade." Forest Ecology and Management 162(2): 299-318.
Wemple, B. C., J.A. Jones, and G.E. Grant (1996). "Channel network extension by logging roads in two basins, Western Cascades, Oregon." Water Resources Bulletin 32(6): 1195-1207.
Wemple, B. C., et al. (1996). "Channel Network Extension by Logging Roads in Two Basins, Western Cascades, Oregon " Water Resources Bulletin 32(6): 1195-1207.
Wentworth, C. K. (1922). "A scale of grade and class terms for clastic sediments." Journal of Geology 30: 377-392.
Werner, B. T. (1999). "Complexity in natural landform patterns." Science 284: 102-104.
Westen, C. J. v. and M. T. J. Terlien (1996). "An approach towards deterministic landslide hazard analysis in GIS. A case study from Manizales (Colombia)." Earth Surface Processes and Landforms 21: 853-868.
Westerling, A. L., H.G. Hidalgo, D.R. Cayan, T.W. Swetnam (2006). "Warming and Earlier Spring Increases Western U.S. Forest Wildfire Activity." Science 313: 940-943.
Westerling, A. L., et al. (2003). "Climate and Wildfire in the Western United States." Bulletin of the American Meteorolgical Society 84(5): 595-604.
Western, A. W., et al. (1999). "Observed spatial organization of soil moisture and its relation to terrain indices." Water Resources Research 35(3): 797-810.
Western Water, A. (2008). Climate change in Colorado: a synthesis to support water resources management and adaptation. University of Colorado, Boulder Colorado.
Westman, W. E. and R. H. Whittaker (1975). "The pygmy forest region of northern California: studies on biomass and primary production." Ecological Monographs 63: 453-520.
Wharton, G. (1995). "The channel geometry method." Earth Surface Processes and Landforms 20(7): 649-660.
Wharton, G. (1995). Information from Channel Geometry-Discharge Relations. Changing River Channels. a. Gurnell and G. Petts, John Wiley & Sons Ltd.: 325-346.
Wheater, C. P. a. P. A. C. (2000). Using statistics to understand the environment. New York, Routledge.
Wheatley, J. M., et al. (2000). Automated Land Cover Mapping Using Landsat Thematic Mapper Images and Topographic Attributes. Terrain Analysis. Principles and Applications. J. P. Wilson and J. C. Gallant. New York, John Wiley & Sons, Inc.: 355-389.
Whipple, K. X. and T. Dunne (?). The influence of derbis-flow rheology on fan morphology, Owens Valley, California. Seattle, Washington, University of Washington: 1-28.
Whipple, K. X. and G. E. Tucker (2002). "Implications of sediment-flux-dependent river incision models for landscape evolution." Journal of  Geophysical Research 107(B2).
Whited, D. C., et al. (2011). "Estimation of juvenile salmon habitat in Pacfic Rim rivers using multiscalar remote sensing and geospatial analysis." River Research and Applications Wileyonlinelibrary.com, DOI: 10.1002/rra.1585.
Whitehouse, I. E., et al. (1986). Growth of Weathering Rinds on Torlesse Sandstone, Southern Alps, New Zealand. Rates of Chemical Weathering of rocks and minerals, Academic Press, Inc.: 419-435.
Whiting, P. J. (1998). "Floodplain Maintenance Flows " Rivers 6(3): 160-170.
Whiting, P. J., J.F.Stamm, D.B.Moog, and R.L. Orndorff (1999). "Sediment-transporting flows in headwater streams." GSA Bulletin 111(3): 450-466.
Whiting, P. J., and G. Matishoff (2005). "Suspended sediment transport distance in watersheds." Eos Trans. AGU,  86(52) Fall Meet. Suppl., Abstract #24B-03.
Whiting, P. J., et al. (1988). "Bedload sheets in heterogeneous sediment." Geology 16: 105-108.
Whiting, P. J., et al. (1999). "Sediment-transporting flows in headwater streams." Geological Society of America Bulletin 111(3): 450-466.
Whitlock, C., et al. (2003). "The role of climate and vegetation change in shaping past and future fire regimes in the northwestern U.S., and the implications for ecosystem management." Forest Ecology and Management 178: 5-21.
Whittaker, J. G. (1987). Sediment Transport in Step-pool Streams. Sediment transport in gravel-bed rivers. C. R. Thorne, J. C. Bathurst and R. D. Hey, John Wiley & Sons Ltd. Chapter 18: 545-579.
Wiberg, P. a. J. S. (1987). "Calculations of the critical shear stress for motion of uniform and heterogeneous sediments." Water Resources Research 23(8): 1471-1480.
Wiberg, P. L. and J. D. Smith (1987). "Calculations of the critical shear stress for motion of uniform and heterogeneous sediments." Water Resources Research 23(8): 1471-1480.
Wiberg, P. L. and J. D. Smith (1989). "Model for calculating bed load transport of sediment." J. Hydraul. Eng. 115(1): 101-123.
A bed load transport model based on the mechanics of sediment moving by saltation yields predicted values of bed load flux as a function of boundary shear stress, grain diameter and density. The parameters required to calculate bed toad transport (particle velocity, bed load sediment concentration, and the height of the bed load layer) can all be determined from our model, which computes sequences of trajectories of individual saltating grains as well as the concentration of moving grains that the flow can support. The latter is related to the momentum  the accelerating grains extract from the flow. Predicted curves of bed load transport vs. boundary shear stress agree well with transport data collected by Gilbert, Meyer-Peter et al., and the Waterways Experimental Station; measured shear stress was corrected for pressure drag when bed forms were present. Comparison of predicted bed load transport with common bed load equations reveals considerable similarity amoung the relationships. The best agreement with the data is produced by relationships in which the transport rate vanishes as the shear stress approaches the critical Shields' number, as in our model and Yalin's equation
 
Wicks, J. M., and J.C.Bathurst (1996). "SHESED: a physiclly based, distributed erosion and sediment yield component for the SHE hydrological modelling system." Journal of Hydrology 175: 213-238.
Wieczorek, G. F. (1984). "Preparing a detailed landslide-inventory map for hazard evaluation and reduction." Bulletin of the Association of Engineering Geologists 21: 337-342.
Wieczorek, G. F. (1987). Effect of rainfall intensity and duration on debris flows in central Santa Cruz Mountains, California. Debris Flows / Avalanches: Process, Recognition, and Mitigation. J. E. Costa and G. F. Wieczorek. Boulder, CO, The Geological Society of America. 7: 93-104.
Rainfall intensity and duration of storms has been shown to influence the triggering of debris flows. After examining storm records of the San Francisco Bay region, documenting when debris flows occurred, and measuring piezometric levels in shallow hillside soils, continuous high-intensity rainfall was found to play a key role in building pore-water pressures that trigger debris flows.
Debris flows in 10 storms between 1975 and 1984 in a 10-km2 area near La Honda, California, were examined, and their rainfall records compared to the records of other storms to determine the antecedent conditions and the levels of continuous, high-intensity rainfall necessary for triggering debris flows. no flows were triggered before 28 cm of rainfall had accumulated each season, which suggests that prestorm soil-moisture conditions are important. After this sufficient antecedent rainfall, a threshold of rainfall duration and intensity --- which accounted for triggering at least one debris flow per sotrm within the study area -- was identified. The number of debris flows increased in storms with intensity and duration characteristics significantly above this threshold.
By studying where debris flows initiated in storms of different intensity and duration, debris flow susceptibility was found to depend on soil thickness and hillside concavity and steepness. Moderate intensity storms of long duration triggered complex soil slump/debris flows in thick soils on concave slopes below large drainage areas, whereas high-intensity storms of short duration caused complex soil slide/debris flows in thinner soils without repsect ot size of drainage area. From these observations, an empirical model based on geology, hydrology, and topography is proposed to account for the triggering of debris flows at selective sites by storms with different combinations of intensity and duration once the antecedent and intensity-duration thresholds are exceeded.
 
Wieczorek, G. F., et al. (1983). Potential for debris flow and debris flood along the Wasatch front between Salt Lake City and Willard, Utah, and measures for their mitigation. Menlo Park, California, U.S. Geological Survey: 13.
Wieczorek, G. F., et al. (1983). Potential for debris flow and debris flood along the Wasatch Front between Salt Lake City and Willard, Utah, and measures for their mitigation. US Geological Survey Open-File Report 83-635., US Geological Survey: 45.
Wieczorek, G. F., et al. (1988). Debris flows and other landslides in San Mateo, Santa Cruz, Contra Costa, Alameda, Napa, Solano, Sonoma, Lake, and Yolo Counties, and factors influencing debris-flow distribution. Landslides, Floods, and Marine Effects of the Storm of January 3-5, 1982, in the San Francisco Bay Region, California: U.S. Geological Survey Professional Paper 1434. S. D. Ellen and G. F. Wieczorek, U.S. Geological Survey: 133-162.
Wieczorek, G. F., et al. (2000). "Debris-flow hazards in the Blue Ridge of Central Virginia." Environmental & Engineering Geoscience 6(1): 3-23.
Wieczorek, G. F., et al. (1995). Analysis of rock falls in the Yosemite Valley, California. Rock Mechanics. D. a. Schultz. Rotterdam, Balkema: 85-89.
Wieland, M. (1988). Effects of Floods of November 18-23, 1988 in Southern Thailand on Highway Bridges and Large Dams. Bangkok, Thailand, Asian Institute of Technology: 181.
Wiens, J. A. (2002). "Riverine landscapes: taking landscape ecology into the water." Freshwater Biology 47: 501-515.
Wigington, P. J. J., et al. (2006). "Coho salmon dependence on intermittent streams." Frontiers in Ecology and the Environment 4(10): 513-518.
In February 2006,the US Supreme Court heard cases that may affect whether intermittent streams are juris-
dictional waters under the Clean Water Act.In June 2006,however,the cases were remanded to the circuit
court,leaving the status of intermittent streams uncertain once again.The presence of commercial species,
such as coho salmon (Oncorhynchus kisutch ),can be an important consideration when determining jurisdic-
tion.These salmon spawn in the upper portions of Oregon coastal stream networks,where intermittent
streams are common.In our study of a coastal Oregon watershed,we found that intermittent streams were an
important source of coho salmon smolts.Residual pools in intermittent streams provided a means by which
juvenile coho could survive during dry periods;smolts that overwintered in intermittent streams were larger
than those from perennial streams.Movement of juvenile coho into intermittent tributaries from the main-
stem was another way in which the fish exploited the habitat and illustrates the importance of maintaining
accessibility for entire stream networks.Loss of intermittent stream habitat would have a negative effect on
coho salmon populations in coastal drainages,including downstream navigable waters.
 
Wigmosta, M. S., and W. A. Perkins (2001). Simulating the effects of forest roads on watershed hydrology. Land Use and Watersheds: Human Influence on Hydrology and Geomorphology in Urban and Forested Areas, American Geophysical Union. Vol. 2: 127-143.
Wigmosta, M. S. and D. P. Lettenmaier (1999). "A comparison of simplified methods for routing topographically driven subsurface flow." Water Resources Research 35(1): 255-264.
Wilcock, P. R. (1988). Methods for Estimating the Critical Shear Stress of Individual Fractions in Mixed-Size Sediment 1127-1135.
Wilcock, P. R. (1992). "Flow competence: A criticism of a classic concept." Earth Surface Processes and Landforms 17: 289-298.
Wilcock, P. R., and B. W. Ardell (1993). "Surface-based fractional transport rates: mobilization threshods and partial transport of a sand-gravel sediment." Water Resources Research 29(4): 1297-1312.
Wilcock, P. R. (1996). "Estimating local bed shear stress from velocity observations." Water Resources Research 32(11): 3361-3366.
Wilcock, P. R. (1997). "Entrainment, displacement and transport of tracer gravels." Earth Surface Processes and Landforms 22: 1125-1138.
Wilcock, P. R. (2001). "Comment on "Effects of hydraulic roughness on surface textures of gravel-bed rivers" and "Effects of sediment supply on surface textures of gravel-bed rivers" by John M. Buffington and David R. Montgomery." Water Resources Research 37(5): 1525-1526.
Wilcock, P. R., S.T. Kenworthy, and J.C.Crowe (2001). "Experimental study of the transport of mixed sand and gravel." Water Resources Research 37(12): 3349-3358.
Wilcock, P. R. (2001). "Toward a practical method for estimating sediment-transport rates in gravel-bed rivers." Earth Surface Processes and Landforms 26: 1395-1408.
Wilcock, P. R., and S.T. Kenworthy (2002). "A two-fraction model for the transport of sand/gravel mixtures." Water Resources Research 38(10): 1194-1206.
Wilcock, P. R., and R.M. Iverson, Ed. (2003). Prediction in Geomorphology, AGU, Washington, D.C.
Wilcock, P. R., and J.C.Crowe (2003). "Surface-based transport model for mixed-size sediment." Journal of Hydraulic Engineering-ASCE 129(2): 120-128.
Wilcock, P. R. and J. B. Southard (1988). "Experimental Study of Incipient Motion in Mixed-Sized Sediment " Water Resources Research 24(7): 1137-1151.
Wiley, T. J. (2000). "Relationship between rainfall and debris flows in western Oregon." Oregon Geology 62(2): 27-43.
Wilford, D. J. (1987). "Watershed Perspective in a Site Specific World." The Forestry Chronicle: 246-249.
Wilford, D. J. (1987). Watershed Workbook: Forest Hydrology Sensitivity Analysis For Coastal British Columbia Watersheds. Smithers, B.C., B.C. Ministry of Forests and Lands: 33.
Wilford, D. J., et al. (2004). "Recognition of debris flow, debris flood and flood hazard through watershed morphometrics." Landslides 1: 61-66.
Debris flows, debris floods and floods in mountainous
areas are responsible for loss of life and damage to infrastructure,
making it important to recognize these hazards in the early stage
of planning land developments. Detailed terrain information is
seldom available and basic watershed morphometrics must be
used for hazard identification. An existing model uses watershed
area and relief (the Melton ratio) to differentiate watersheds
prone to flooding from those subject to debris flows and debris
floods. However, the hazards related to debris flows and debris
floods are not the same, requiring further differentiation. Here,
we demonstrate that a model using watershed length combined
with the Melton ratio can be used to differentiate debris-flow and
debris-flood prone watersheds. This model was tested on 65 alluvial
and colluvial fans in west central British Columbia, Canada,
that were examined in the field. The model correctly identified
92% of the debris-flow, 83% of the debris-flood, and 88% of the
flood watersheds. With adaptation for different regional conditions,
the use of basic watershed morphometrics could assist land
managers, scientists, and engineers with the identification of
hydrogeomorphic hazards on fans elsewhere.
 
Wilkeson, R. (1999). Recovery of wild salmonids in western Oregon forests: Oregon forest practices act rules and measures in the Oregon plan for salmon and watersheds: 1-85.
Wilkinson, S. N., et al. (2009). "Predicting the distribution of bed material accumulation using river network sediment budgets." Water Resources Research 42(W10419).
Assessing the spatial distribution of bed material accumulation in river networks is
important for determining the impacts of erosion on downstream channel form and habitat
and for planning erosion and sediment management. A model that constructs spatially
distributed budgets of bed material sediment is developed to predict the locations of
accumulation following land use change. For each link in the river network, GIS
algorithms are used to predict bed material supply from gullies, river banks, and upstream
tributaries and to compare total supply with transport capacity. The model is tested in the
29,000 km2 Murrumbidgee River catchment in southeast Australia. It correctly predicts
the presence or absence of accumulation in 71% of river links, which is significantly
better performance than previous models, which do not account for spatial variability in
sediment supply and transport capacity. Representing transient sediment storage is
important for predicting smaller accumulations. Bed material accumulation is predicted in
25% of the river network, indicating its importance as an environmental problem in
Australia.
 
Willett, S. D. and M. T. Brandon (2002). "On steady states in mountain belts." Geology 30(2): 175-178.
The dynamic system of tectonics and erosion contains important feedback mechanisms
such that orogenic systems tend toward a steady state. This concept is often invoked,
but the nature of the steady state is commonly not specified. We identify four types of
steady state that characterize the orogenic system and illustrate these cases by using numerical-
model results and natural examples. These types are (1) flux steady state, (2)
topographic steady state, (3) thermal steady state, and (4) exhumational steady state: they
refer to the erosional flux, the topography, the subsurface temperature field, and the
spatial pattern of cooling ages, respectively. Models suggest that the topography will reach
a steady mean form at the scale of an orogenic belt, but perfect topographic steady state
is unlikely to be achieved at shorter length scales. Thermal steady state is a precondition
for exhumational steady state and in the case of temperature-dependent deformation,
topographic steady state. Exhumational steady state is characterized by reset age zones
spatially nested according to closure temperature, as illustrated in natural systems from
New Zealand, the Cascadia accretionary margin, and Taiwan.
 
Willgoose, G., et al. (1991). "A Coupled Channel Network Growth and Hillslope Evolution Model: 2. Nondimensionalization and Applications." Water  Resources Research 27: 1685-1696.
Willgoose, G., et al. (1991). "A Coupled Channel Network Growth and Hillslope Evolution Model: I. Theory." Water  Resources Research 27: 1671-1683.
Willgoose, G., et al. (1991). "Results from a new model of river basin evolution." Earth Surface Processes and Landforms 16: 237-254.
Williams, C. K. (1989). Major Forest Ecological Regions of Eastern Washington.
Williams, J. R., A.D. Nicks, and J.G. Arnold (1985). "Simulator for water resources in rural basins." Journal of Hydraulic Engineering-ASCE 111(6): 970-986.
Williams, P., Jr. (1948). "The Variation of the Time of Maximum Precipitation Along the West Coast of North America." The Bulletin of the American Meteorological Society 29(4): 143-145.
Williams, S. R. and S. W. Smith (1990). Magnitude Predictions for Debris Flows in Central Davis County.
Willis, K. J., et al. (1999). "The Role of Sub-Milankovitch Climatic Forcing in the Initiation of the Northern  Hemisphere Glaciation " Science 285: 568-571.
Wilmott, C. J. (1981). "On the validation of models." Physical Geography 2(2): 184-194.
Wilson, C. J. (1999). "Effects of logging and fire on runoff and erosion of highly erodible granitic soils in Tasmania." Water Resources Research 35(11): 3531-3546.
Wilson, C. J., J.W. Carey, P.C. Beeson, M.O. Gard, and L.J. Lane (2001). "A GIS-based hillslope erosion and sediment delivery model and its application in the Cerro Grande burn area." Hydrological Processes 15(15): 2995-3010.
Wilson, C. J. and W. E. Dietrich (1987). The contribution of bedrock groundwater flow to storm runoff and high pore pressure development in hollows. Erosion and Sedimentation in the Pacific Rim. R. L. e. a. Beschta. Corvallis, Oregon, USA, International Associateon of Hydrological Science. IAHS Publication 165: 49-59.
Wilson, E. O. (?). The Natural Sciences Chapter 4: 45-65.
Wilson, J. P. and J. C. Gallant, Eds. (2000). Terrain Analysis. Principles and Applications. New York, John Wiley & Sons.
Wimberly, M. C. (2002). "Spatial simulation of historical landscape patterns in coastal forests of the Pacific Northwest." Canadian Journal of Forest Resources 32(8): 1316-1328.
Wimberly, M. C. and R. S. H. Kennedy (2008). "Spatially explicit modeling of mixed-severity fire regimes and landscape dynamics." Forest Ecology and Management 254: 511-523.
Simulation models of disturbance and succession are being increasingly applied to characterize landscape composition and dynamics under
natural fire regimes, and to evaluate alternative management strategies for ecological restoration and fire hazard reduction. However, we have a
limited understanding of how landscapes respond to changes in fire frequency, and about the sensitivity of model predictions to assumptions about
successional pathways and fire behavior.We updated an existing landscape dynamics model (LADS) to simulate the complex interactions between
forest dynamics, fire spread, and fire effects in dry forests of the interior Pacific Northwest. Experimental model runs were conducted on a
hypothetical landscape at fire rotations ranging from 5 to 50 years. Three sensitivity analyses were carried out to explore the responses of landscape
composition to (1) parameters characterizing succession and fire effects on vegetation, (2) the probability of fire spread into different successional
stages, and (3) the size and spatial pattern of static fire refugia. The area of old open-canopy forests was highest at the shortest fire rotations, and was
particularly sensitive to the probability of stand-replacement fire in open-canopy forests and to the fire-free period required for ingrowth to occur in
open-canopy forests. The area of old closed-canopy forests increased with lengthening fire rotation, but always comprised a relatively small portion
of the landscape (<10%). The area of old closed-canopy forests increased when fire spread was more rapid in open-canopy forests than in closedcanopy
forests, and when the physical landscape incorporated large ‘‘fire refugia’’ with low fire spread rates. Old closed-canopy forests appear to
comprise a relatively minor landscape component in mixed-severity fire regimes with fire rotations of 50 years or less. However, these results are
sensitive to assumptions about the spatial interactions between fire spread, landscape vegetation patterns, and the underlying physical landscape.
 
Wimberly, M. C. and T. A. Spies (2001). "Influences of environment and disturbance on forest patterns in coastal Oregon watersheds." Ecology 82(5): 1443-1459.
Modern ecology often emphasizes the distinction between traditional theories
of stable, environmentally structured communities and a new paradigm of disturbancedriven,
nonequilibrium dynamics. However, multiple hypotheses for observed vegetation
patterns have seldom been explicitly tested. We used multivariate statistics and variation
partitioning methods to assess the relative influences of environmental factors and disturbance
history on riparian and hillslope forests. Our study area was the Cummins Creek
Wilderness, located in the Oregon Coast Range. Most of the wilderness burned at least
once between the mid-19th and early 20th centuries, creating a mosaic of younger forests
with a few old-growth patches. Species composition on hillslopes varied primarily along
a climatic gradient from moist maritime environments to drier inland climates but was
relatively insensitive to forest age structure. The abundance of Tsuga heterophylla, a firesensitive,
late-successional tree species, decreased with distance from old-growth patches,
suggesting possible seed-source limitations following the historical fires. In contrast to
species composition, hillslope forest structure was primarily related to fire history but was
largely independent of environmental gradients. Old-growth structure characteristics such
as large dominant trees, large snags, high down-wood volumes, and high tree size diversity
increased with stand age and with the presence of remnant trees that survived the fires.
Riparian forests had high shrub cover, abundant hardwoods, and high down-wood volumes,
while the conifer-dominated hillslopes had high overstory density and basal area. Maritime
climates and their associated plant species extended further inland in riparian areas than
on hillslopes. Advance regeneration densities were higher in riparian forests within 5 km
of the coast than in any other portion of the study area. Riparian forest structure and
composition were related to both environmental and disturbance variables, with stream
gradient and size integrating much of the fine-scale variability in disturbance regimes. No
single theoretical framework was sufficient to explain the vegetation patterns observed in
these forested watersheds. Our findings suggest a conceptual model of forest landscapes
in which the relative influences of environment and disturbance on vegetation patterns are
contingent on the facet of vegetation considered (composition vs. structure) and the portion
of the landscape examined (hillslope vs. riparian).
 
Wimberly, M. C., et al. (2000). "Simulating historical variability in the amount of old forests in the Oregon Coast Range." Conservation Biology 14: 167-180.
Wing, M. G. and A. E. Skaugset (2002). "Relationships of channel characteristics, land ownership, and land use patterns to large woody debris in western Oregon streams." Canadian Journal of Fisheries and Aquatic Science 59: 796-807.
Regression tree analysis was used to test the relationship of channel and aquatic habitat characteristics from
3793 stream reaches in western Oregon State to the abundance of large woody debris (LWD). Stream reaches were
drawn from diverse ownerships and land uses – land cover types over a broad geographic extent. When all land uses –
land covers were considered, ownership and land use patterns were related to LWD abundance. When nonforested land
uses were excluded, however, these factors became less important. In forested streams, LWD abundance was predicted
primarily by stream gradient and bankfull channel width, with the volume, frequency, and size of LWD pieces decreasing
as channel size increased. Within forested lands, stand age and forest distribution were related to LWD size but had
small correlations with LWD volume and abundance. The strong relationship of stream gradient and bankfull channel
width with LWD suggests that in forested areas, the most significant factor related to LWD counts is the geomorphology
of stream reaches and their surrounding areas. Land managers in western Oregon who want to improve aquatic
habitat quality may want to direct their efforts to increasing LWD in larger streams, which typically include smaller
quantities of LWD.
 
Winkler, R., D. Spittlehouse, T. Giles, B. Heise, G. Hope, and M. Schnorbus (2004). "Upper Penticton Creek: how forest harvesting affects water quantity and quality." Streamline: Water Management Bulletin 8(1): 18-20.
Winograd, I. J., et al. (1997). "Duration and Structure of the Past Four Interglaciations " Quaternary Research 48: 141-154.
Winter, T. C. (2007). "The role of ground water in generating streamflow in headwater areas and in maintaining base flow." Journal of the American Water Resources Association 43(1): 15-25.
The volume and sustainability of streamflow from headwaters to downstream reaches commonly depend on contributions from ground water. Streams that begin in extensive aquifers generally have a stable point of origin and substantial discharge in their headwaters. In contrast, streams that begin as discharge from rocks or sediments having low permeability have a point of origin that moves up and down the channel seasonally, have small incipient discharge, and commonly go dry. Nearly all streams need to have some contribution from ground water in order to provide reliable habitat for aquatic organisms. Natural processes and human activities can have a substantial effect on the flow of streams between their headwaters and downstream reaches. Streams lose water to ground water when and where their head is higher than the contiguous water table. Although very common in arid regions, loss of stream water to ground water also is relatively common in humid regions. Evaporation, as well as transpiration from riparian vegetation, causing ground-water levels to decline also can cause loss of stream water. Human withdrawal of ground water commonly causes streamflow to decline, and in some regions has caused streams to cease flowing.
 
Winterbottom, S. J. and D. J. Gilvear (1997). "Quantification of channel bed morphology in gravel-bed rivers using airborne multispectral imagery and aerial photography." Regulated Rivers: Research & Management 13: 489-499.
Winterbourn, M. J., et al. (1981). "Are New Zealand stream ecosystems really different?" New Zealand Journal of Marine and Freshwater Research 15: 321-328.
Winzler and Kelly (1980). Yager Creek Drainage Sediment Investigation 1980. Scotia, California, prepared by Winzler and Kelly for The Pacific Lumber Company: 1-35.
Wipfli, M. S. and D. P. Gregovich (2002). "Export of invertbrates and detritus from fishless headwater streams in southeastern alaska: implications for downstream salmonid production." Freshwater Biology 47(47): 957-969.
1. We examined the export of invertebrates (aquatic and terrestrial) and coarse organic
detritus from forested headwaters to aquatic habitats downstream in the coastal mountains
of southeast Alaska, U.S.A. Fifty-two small streams (mean discharge range: 1.2–3.6 L s)1),
representing a geographic range throughout southeast Alaska, were sampled with 250-lm
nets either seasonally (April, July, September) or every 2 weeks throughout the year.
Samples were used to assess the potential subsidy of energy from fishless headwaters to
downstream systems containing fish.
2. Invertebrates of aquatic and terrestrial origin were both captured, with aquatic taxa
making up 65–92% of the total. Baetidae, Chironomidae and Ostracoda were most
numerous of the aquatic taxa (34, 16 and 8%, respectively), although Coleoptera (mostly
Amphizoidae) contributed the greatest biomass (30%). Mites (Acarina) were the most
numerous terrestrial taxon, while terrestrial Coleoptera accounted for most of the
terrestrial invertebrate biomass.
3. Invertebrates and detritus were exported from headwaters throughout the year,
averaging 163 mg invertebrate dry mass stream)1 day)1 and 10.4 g detritus stream)1
day)1, respectively. The amount of export was highly variable among streams and seasons
(5–6000 individuals stream)1 day)1 and <1–22 individuals m)3 water; <1–286 g detritus
stream)1 day)1 and <0.1–1.7 g detritus m)3 water). Delivery of invertebrates from
headwaters to habitats with fish was estimated at 0.44 g dry mass m)2 year)1. We estimate
that every kilometre of salmonid-bearing stream could receive enough energy (prey and
detritus) from fishless headwaters to support 100–2000 young-of-the-year (YOY) salmonids.
These results illustrate that headwaters are source areas of aquatic and terrestrial
invertebrates and detritus, linking upland ecosystems with habitats lower in the
catchment.
 
Wipfli, M. S. and D. P. Gregovich (2002). "Export of invertebrates and detritus from fishless headwater streams in south-eastern Alaska: implications for downstream salmonid production." Freshwater Biology 47: 1-13.
Wipfli, M. S. and D. P. Gregovich (2002). "Export of invertebrates and detritus from fishless headwater streams in southeastern Alaska: implications for downstream salmonid production." Freshwater Biology 47: 957-969.
Wipfli, M. S., et al. (1998). "Influence of salmon carcasses on stream productivity: response of biofilm and benthic macroinvertebrates in southeastern Alaska, U.S.A." Canadian Journal of Fisheries and Aquatic Sciences 55(6): 1503-1511.
Wissmar, R., et al. (2009). "Radar-derived digital elevation models and field-surveyed variables to predict distributions of juvenile coho salmon and Dolly Varden in remote streams in Alaska." Transactions of the American Fisheries Society 139: 288-302.
Wissmar, R. C. (1993). "The need for long-term stream monitoring programs in forest ecosystems of the Pacific Northwest." Environmental Monitoring and Assessment 26: 219-234.
Wissmar, R. C. (1994). Influences of Disturbance Regimes on the Spatial and Long-term Development of Riparian and River Landscapes (From RFF proposal 92-93), Fisheries Research Institute.
Wissmar, R. C., et al. (1990). "Contribution of organic acids to alkalinity in lakes within the Mount St. Helens blast zone." Limnol. Oceanogr. 35(2): 535-542.
Wissmar, R. C. and F. J. Swanson (1990). "Landscape disturbances and lotic ecotones." The Ecology and Management of Aquatic-Terrestrial Ecotones.
Wittenberg, h. (1999). "Baseflow recession and recharge as nonlinear storage processes." Hydrological Processes 13: 715-726.
Witzel, L. D. and H. R. MacCrimmon (1983). "Embryo survival and alevin emergence of brook charr, Salvelinus fontinalis, and brown trouth, Salmo trutta, relative to redd gravel composition." Can. J. Zool. 61: 1783-1791.
Witzel, L. D. and H. R. MacCrimmon (1983). "Redd-Site Selection by Brook Trout and Brown Trout in Southwestern Ontario Streams " Transactions of the American Fisheries Society 112: 760-771.
Wohl, E. (2000). Mountain Rivers. Washington, D.C., American Geophysical Union.
Wohl, E., et al. (2005). "River restoration." Water Resources Research 41: 12.
River restoration is at the forefront of applied hydrologic science. However, many
river restoration projects are conducted with minimal scientific context. We propose two
themes around which a research agenda to advance the scientific basis for river
restoration can be built. First, because natural variability is an inherent feature of all river
systems, we hypothesize that restoration of process is more likely to succeed than
restoration aimed at a fixed end point. Second, because physical, chemical, and biological
processes are interconnected in complex ways across watersheds and across timescales,
we hypothesize that restoration projects are more likely to be successful in achieving goals
if undertaken in the context of entire watersheds. To achieve restoration objectives, the
science of river restoration must include (1) an explicit recognition of the known
complexities and uncertainties, (2) continued development of a theoretical framework that
enables us to identify generalities among river systems and to ask relevant questions,
(3) enhancing the science and use of restoration monitoring by measuring the most
effective set of variables at the correct scales of measurement, (4) linking science and
implementation, and (5) developing methods of restoration that are effective within
existing constraints. Key limitations to river restoration include a lack of scientific
knowledge of watershed-scale process dynamics, institutional structures that are poorly
suited to large-scale adaptive management, and a lack of political support to reestablish
delivery of the ecosystem amenities lost through river degradation. This paper outlines an
approach for addressing these shortcomings.
 
Wohl, E. and D. Merrit (2005). "Prediction of mountain stream morphology." Water Resources Research 41: 10.
We use a large and diverse data set from mountain streams around the world to
explore relationships between reach-scale channel morphology and control variables. The
data set includes 177 step-pool reaches, 44 plane-bed reaches, and 114 pool-riffle reaches
from the western United States, Panama, and New Zealand. We performed several
iterations of stepwise discriminant analysis on these data. A three-variable discriminant
function using slope (S), D84, and channel width (w) produced an error rate of 24% for the
entire data set. Seventy percent of plane-bed reaches were correctly classified (16%
incorrectly classified as pool-riffle and 14% incorrectly classified as step-pool). Sixtyseven
percent of pool-riffle channels were correctly classified (31% incorrectly
classified as plane-bed and 2% as step-pool). Eighty-nine percent of step-pool reaches
were correctly classified (9% incorrectly classified as plane-bed and 2% as pool-riffle).
The partial R2 values and F tests indicate that S is by far the most significant single
explanatory variable. Comparison of the eight discriminant functions developed using
different data sets indicates that no single variable is present in all functions, suggesting
that the discriminant functions are sensitive to the specific stream reaches being analyzed.
However, the three-variable discriminant function developed from the entire data set
correctly classified 69% of the 159 channels included in an independent validation data
set. The ability to accurately classify channel type in other regions using the three-variable
discriminant function developed from the entire data set has important implications for
water resources management, such as facilitating prediction of channel morphology using
regional S-w-D84 relations calibrated with minimal field work.
 
Wohl, E. E. (2004). "Limits of downstream hydraulic geometry." Geology 32(10): 897-900.
Wohl, E. E. (2005). "Compromised rivers: understanding historical human impacts on rivers in the context of restoration." Ecology and Society 10(2).
Wohl, E. E., P.L. Angermeier, B. Bledsoe, G.M. Kondolf, L. MacDonnell, D.M. Merritt, M.A. Palmer, N.L. Poff, and D. Tarboton (2005). "River restoration." Water Resources Research 41(W10301): 12p.
Wohl, E. E., et al. (1996). "A comparison of surface sampling methods for coarse fluvial sediments." Water Resources Research 32(10): 3219-3226.
Wohl, E. E. and D. M. Merritt (2001). "Bedrock channel morphology." Geological Society of America Bulletin 113(9): 1205-1212.
Analyses of 41 bedrock channel reaches
indicate quantifiable relationships between
bedrock channel morphology and reachscale
hydraulic and substrate variables.
Discriminant analysis was used to develop
a discriminant criterion based on reach-averaged
channel gradient, substrate heterogeneity,
and Selby rock-mass strength. This
criterion correctly classified 70% of the observations
into one of five channel morphologic
types. Channels formed at higher gradients
have a morphology that effectively
maximizes the erosional force, whereas a
morphology that evenly distributes flow energy
or dissipates flow energy internally is
associated with lower gradients. These results
suggest that bedrock channel morphology,
like alluvial channel morphology,
reflects a quantifiable balance between hydraulic
driving and substrate resisting
forces.
 
Wohl, E. E. and P. P. Pearthree (1991). "Debris flows as geomorphic agents in the Huachuca Mountains of southeastern Arizona." Geomorphology 4: 273-292.
Wohlgemuth, P. M., et al. (2001). "The effects of log erosion barriers on post-fire hydrologic response and sediment yield in small forested watersheds, southern California." Hydrological Processes 15(15): 3053-3066.
Wildfire usually promotes Hooding and accelerated erosion in upland watersheds. In the summer of 1999, a high-severity wildfire burned a series of mixed pine/oak headwater catchments in the San Jacinto Mountains of southern California. Log erosion barriers (LEBs) were constructed across much of the burned area as an erosion control measure. We built debris basins in two watersheds, each about I ha in area, one with LEBs, the other without, to measure post-fire hydrologic response and sediment yield and to evaluate the effectiveness of the LEBs. The watersheds are underlain by granitic bedrock, producing a loamy sand soil above large extents of weathered bedrock and exposed core stones (tors) on the surface. Measured soil water-repellency was similar over the two catchments. Rain gauges measured 348 mm of precipitation in the first post-fire year. The ephemeral stream channels experienced surface flow after major rainstorms. and the source of the water was throughflow exfiltration at the slope/channel interface. Post-fire overland flow produced some rilling, but hillslope erosion measured in silt fences away from any LEBs was minor, as was sediment accumulation behind the LEBs. Stream channels in the catchments exhibited minor net scour. Water yield was much greater in the LEB-treated watershed. This resulted in 14 times more sediment yield by weight than the untreated watershed. Average soil depths determined by augering were nearly double in the catchment without the LEBs compared with the treated watershed. This suggests that differences in water and sediment yield between the two catchments are due to the twofold difference in the estimated soil water-holding capacity in the untreated watershed. It appears that the deeper soils in the untreated watershed were able to retain most of the precipitation, releasing less Water to the channels and thereby reducing erosion and sediment yield. Thus, the test of LEB effectiveness was inconclusive in this study, because soil depth and soil water-holding capacity may have masked their performance. Published in 2001 by John Wiley & Sons, Ltd.
 
Woillez, M., et al. (2007). "Indices for capturing spatial patterns and their evolution in time, with application to European hake (Merluccius merluccius) in the Bay of Biscay." ICES Journal of Marine Science 64(3): 537-550.
A series of candidate statistical indices is used in an attempt to capture spatial patterns of fish populations from research survey data.
To handle diffuse population limits, indices are designed not to depend on arbitrary delineation of the domain. They characterize the
location (centre of gravity and spatial patches), the occupation of space (inertia, isotropy, positive area, spreading area, and equivalent
area), statistical dispersion (Gini index and coefficient of variation of strictly positive densities), and microstructure. Collocation
between different ages and years is summarized by a global index of collocation. Indices are estimated for hake from a bottomtrawl
data series in the Bay of Biscay in autumn of 1987–2004. The study provides a detailed description of the spatial patterns
of different hake age groups, age 3 appearing to be a turning point in these dynamics. Capturing spatial patterns through indices
allows the comparison of surveyed populations and identification of trends and outliers in the time-series. Spatial indices are used
in a multivariate approach to obtain an overview of the relationships between the different spatial indices characterizing the
spatial behaviour of six age groups of hake, and to assess their persistence through time.
 
Wolfe, M. D. and J. W. Williams (1986). "Rates of landsliding as impacted by timber management activities in northwestern California." Bulletin of the Association of Engineering Geologists 232(1): 53-60.
The impacts on the rates of landsliding by forest management activities (road construction and timber harvesting) in northwestern California are evaluated. The effects of these land disturbing activities were studied for various geomorphic zones of a watershed, valley inner gorges, land with slopes greater than 80 percent, and ' other ' watershed lands. An analysis of historical aerial photography and computer generated digital terrain slope maps was used to determine changes, over time, in the distribution of active landslides in third order watersheds, where the present watershed conditions vary from pristine to highly disturbed. In all of the disturbed watersheds, forest management activities increased the landslide rates (landslides per square mile) of each geomorphic zone. The valley inner gorges, and land with slopes greater than 80 percent had the greatest landslide rate increases. Managed valley inner gorges and managed slopes greater than 80 percent had 11 to 26 and 3 to 26 times more landslides per square mile than managed ' other lands ', respectively. These data suggest that the valley inner gorges and land with slopes greater than 80 percent are the most landslide prone geomorphic zones in a watershed, and the most sensitive to forest management activities.
 
Wolff, N., et al. (1989). The effects of partial forest-stand removal on the availability of water for groundwater recharge (DRAFT), Department of Natural Resources: 12.
Wollheim, W. M., et al. (1999). "A coupled field and modeling approach for the analysis of nitrogen cycling in streams." Journal of the North American Benthological Society 18(2): 199-221.
Wolman, G. M. and J. P. Miller (1960). "Magnitude and frequency of forces in Geomorphic processes." The Journal of Geology 68: 54-74.
Wolman, M. G. (1954). "A method of sampling coarse bed material." Transactions, American Geophysical Union 35: 951-956.
Wolman, M. G. (1954). "A method of sampling coarse river-bed material." Transactions of the American Geophysical Union 35(6): 951-956.
Wolman, M. G. (1959). "Factors influencing erosion of a cohesive river bank." American Journal of Science 257: 204-216.
Wolman, M. G., and J.P. Miller (1960). "Magnitude and frequency of forces in geomorphic processes." Journal of Geology 68(1): 54-74.
Conclusions: effectiveness of processes which control many landforms depend upon their distribution in time as well as their magnitude. Rare and infrequent events may not be the most significant in terms of forming the landscape. Much of the "work" is performed by events of moderate magnitude and frequency eg. bankfull discharge, not floods. May be different for different landforms - requires observations of landforms and processes.
 
Wolman, M. G. (1977). "Changing needs and opportunities in the sediment field." Water  Resources Research 13(1): 50-54.
Wolman, M. G. and R. Gerson (1978). "Relative scales of time and effectiveness of climate in watershed geomorphology." Earth Surface Processes 3: 189-208.
Wolman, M. G. and L. B. Leopold (1957). River Flood Plains: Some Observations On Their Formation US Geologic Survey: 107.
Wolman, M. G. and L. B. Leopold (1957). River Flood Plains: Some observations on their formation: 107.
Wolman, M. G. and J. P. Miller (1960). "Magnitude and frequency of forces in geomorphic processes." Journal of Geology 68: 54-74.
Wondzell, S. M. (1994). Flux of ground water and nitrogen through the floodplain of a fourth-order stream. Department of Forest Science, Oregon State University: 1-113.
Wondzell, S. M. (2006). "Effect of morphology and discharge on hyporheic exchange flows in two small streams in the Cascade Mountains of Oregon, USA." Hydrological Processes 20: 267-287.
Stream-tracer injections were used to examine the effect of channel morphology and changing stream discharge on
hyporheic exchange flows. Direct observations were made from well networks to follow tracer movement through the
hyporheic zone. The reach-integrated influence of hyporheic exchange was evaluated using the transient storage model
(TSM) OTIS-P. Transient storage modelling results were compared with direct observations to evaluate the reliability
of the TSM. Results from the tracer injection in the bedrock reach supported the assumption that most transient
storage in headwater mountain streams results from hyporheic exchange. Direct observations from the well networks
in colluvial reaches showed that subsurface flow paths tended to parallel the valley axis. Cross-valley gradients were
weak except near steps, where vertical and cross-valley hydraulic gradients indicated a strong potential for stream
water to downwell into the hyporheic zone. The TSM parameters showed that both size and residence time of transient
storage were greater in reaches with a few large log-jam-formed steps than in reaches with more frequent, but smaller
steps. Direct observations showed that residence times in the unconstrained stream were longer than in the constrained
stream and that little change occurred in the location and extent of the hyporheic zone between low- and high-baseflow
discharges in any of the colluvial reaches. The transient storage modelling results did not agree with these observations,
suggesting that the TSM was insensitive to long residence-time exchange flows and was very sensitive to changes in
discharge. Disagreements between direct observations and the transient storage modelling results highlight fundamental
problems with the TSM that confound comparisons between the transient storage modelling results for tracer injections
conducted under differing flow conditions. Overall, the results showed that hyporheic exchange was little affected by
stream discharge (at least over the range of baseflow discharges examined in this study). The results did show that
channel morphology controlled development of the hyporheic zone in these steep mountain stream channels.
 
Wondzell, S. M., et al. (2007). "Flow velocity and the hydrologic behavior of streams during baseflow." Geophysical Research Letters 34.
Diel variations in stream discharge have long been
recognized, but are relatively little studied. Here we
demonstrate that these diel fluctuations can be used to
investigate both streamflow generation and network routing.
We treat evapo-transpiration (ET) as a distributed impulse
function in an advection model and analyze the effect of
ET on diel fluctuations in discharge. We show that when
flow velocity is high during high baseflow, discharge
fluctuations tend to be in phase and constructive
interference reinforces ET-generated signals resulting in
strong diel fluctuations measured at a gauging station at the
mouth of the watershed. As flow velocity slows with
baseflow recession, ET-generated signals are increasingly
out of phase so that fluctuations in discharge are masked by
destructive interference. These results demonstrate that
naturally produced fluctuations in discharge constitute
discrete impulse functions that can be used to analyze
eco-hydrologic behavior of whole-watersheds during
baseflow periods.
 
Wondzell, S. M., et al. (2006). "Simulating riparian vegetation and aquatic habitat dynamics in response to natural and anthropogenic disturbance regimes in the Upper Grande Ronde River, Oregon, USA." Landscape and Urban Planning 80(3): 249-267.
We developed state and transition models (STMs) to evaluate the effects of natural disturbances and land-use practices on aquatic and riparian habitats. The STMs consisted of discrete states defined by channel morphology and riparian vegetation. Transitions between states resulted from plant succession and from natural and anthropogenic disturbances. Channel conditions and habitat suitability for anadromous salmonids were ranked by using a qualitative four-factor scale for each state in the STMs. Disturbance probabilities were varied to define both historical and current disturbance regimes. Models were run for 120 years with the current disturbance regime to illustrate changes associated with Euro-American settlement, and then run for an additional 50 years under the historical disturbance regime to illustrate the potential for passive recovery. Results suggested that Euro-American settlement dramatically changed riparian vegetation and channel conditions, which resulted in substantial declines in habitat quality. Passive recovery of channel conditions and habitat suitability was quick in some stream types, but slow in others. Overall, our results underestimate the effects of human land uses on streams and overestimate the rate of recovery under passive restoration because the models do not yet include the effects of many management activities, especially those resulting from forest harvest and roads.
 
Wondzell, S. M. and J. King (2003). "Fire and hillslope erosion, transport, and sedimentation." Forest Ecology and Management 178(1-2): 75-88.
Wondzell, S. M. and J. G. King (2003). "Post-fire erosional processes in the Pacific Northwest and Rocky Mountain regions." Forest Ecology and Management 178: 75-87.
The objective of this paper is to provide a general overview of the influence of wildland fires on the erosional processes common to the forested landscapes of the western United States. Wildfire can accelerate erosion rates because vegetation is an important factor controlling erosion. There can be great local and regional differences, however, in the relative importance of different erosional processes because of differences in prevailing climate, geology and topography; because of differences in the degree to which vegetation regulates erosional processes; and because of differences in the types of fire regimes that disrupt vegetative cover. Surface erosion, caused by overland flow, is a dominant response to wildfire in the Interior Northwest and Northern Rocky Mountains (Interior region). A comparison of measured post-fire infiltration rates and long-term records of precipitation intensity suggest that surface runoff from infiltration-excess overland flow should also occur in the Coastal and Cascade Mountains of the Pacific Northwest after fires, but this has not been documented in the literature. Debris slides and debris flows occur more frequently after wildfire in the Interior region and in the Coastal and Cascade Mountains of the Pacific Northwest (Pacific Northwest region). Debris flows can be initiated from either surface runoff or from soil-saturation-caused debris slides. In the Pacific Northwest region, debris flows are typically initiated as debris slides, caused by soil saturation and loss of soil cohesion as roots decay following fire. In the Interior region, both overland-flow-caused and debris-slide-caused debris flows occur after wildfire. Surface erosion, debris slides, and debris flows all occur during intense storms. Thus, their probability of occurrence depends upon the probability of intense storms occurring during a window of increased susceptibility to surface erosion and mass wasting following intense wildfire.
 
Wondzell, S. M. and J. G. King (?). Post-Fire Erosional Processes in the Pacific Northwest and Rocky Mountain Regions. Forest Ecology and Management: 1-12.
Wondzell, S. M. and F. J. Swanson (1996). "Seasonal and storm dynamics of the hyporheic zone of a 4th-order mountain stream. I: Hydrologic processes." J. N. Am. Benthol. Soc. 15(1): 3-19.
advected channel water
ground water
groundwater flow models
streams
aquifer
water budget
flow path
geomorphology
 
Wondzell, S. M. and F. J. Swanson (1996). "Seasonal and storm dynamics of the hyporheic zone of a 4th-order mountain stream. II: Nitrogen cycling." J. N. Am. Benthol. Soc. 15(1): 20-34.
Wondzell, S. M. and F. J. Swanson (1999). "Floods, channel change, and the hyporheic zone." Water Resources Research 35(2): 555-567.
Wood, A. W., L.R.Leung, V. Sridhar, and D. P. Lettenmaier (2004). "Hydrologic implications of dynamical and statistical approaches to downscaling climate model outputs." Climatic Change 62: 189-216.
Wood, P. J., and P.D. Armitage, (1997). "Biological effects of fine sediment on the lotic environment." Environmental Management 21(2): 203-217.
Wood, S. H., et al. (1999). Contrasting sediment transport events generated by rain on snow and convective storms on burned basins in the Idaho Batholith. Abstracts with Programs - Geological Society of America. 31: 440-441.
Recent stand-replacing fires in the ponderosa pine-dominated forests of the southern Idaho batholith have been followed by catastrophic sediment transport events with very different generative mechanisms. On January 1, 1997, approximately 7 cm of intense rain fell on a melting snowpack of 13-23 cm water equivalent in the lower South Fork Payette River basin and adjacent areas. Although many debris flows, hyperconcentrated flows, and sediment-charged floods issued from steep low-order basins burned in the 1989 Lowman fire, others initiated on unburned forested slopes and south-facing grass- and forb-covered slopes. Failures of grussy colluvium 2-6 m deep occurred in slope hollows >20 degrees , with up to 4-12 failures per 0.5-1 km (super 2) basin. We found no clear differences in basin morphometry or total volume of failures to explain why the grassy Hopkins Creek basin produced dominantly sheetflooding on its fan, while a burned forest basin east of Archie Creek produced a large, rapid debris flow. Deposition at Hopkins Creek took place over several hours, however, implying continuing slope failures, whereas a large single failure may have generated the debris flow near Archie Creek, which crossed the South Fork Payette channel and expanded over an arc of 110 degrees as it flowed 90 m across 5-7 m high terraces on the opposite bank. Charred logs are common in this boulder-poor deposit, especially around its outer margin. In contrast, an intense thunderstorm in August 1995 caused sediment transport primarily by widespread runoff from slopes burned in a 1995 fire in the North Fork Boise River drainage. Rills several cm deep integrated downslope into gullies >1 m deep, then fed highly erosive flows that scoured stored alluvium from trunk channels. Although colluvium adjacent to channels was incorporated through undercutting and minor failures, en masse failure of colluvial slope hollows was rare. Again, a variety of transport and depositional processes operated to produce diverse facies on tributary fans along the North Fork Boise River. We continue to document differences in deposit character and inferred transport processes between these events, between different basins in each event, and within individual basins.
 
Wooding, R. A. (1991). "Growth on Natural Dams by Deposition from Steady Supersaturated Shallow Flow." Journal of Geophysical Research 96(B1): 667-682.
Woods, R. A. and M. Sivapalan (1997). "A connection between topographically driven runoff generation and channel network structure." Water Resources Research 33(12): 2939-2950.
Woods, R. A., et al. (1997). "Modeling the spatial variability of subsurface runoff using a topographic index." Water Resources Research 33(5): 1061-1073.
Woodsmith, R. D. and J. M. Buffington (1996). "Multivariate geomorphic analysis of forest streams: implications for assessment of land use impacts on channel condition." Earth Surface Processes and Landforms 21: 377-393.
Wood-Smith, R. D. and J. M. Buffington (1996). "Multivariate Geomorphic Analysis of Forest Streams: Implications for Assessment of Land Use Impacts on Channel Condition " Earth Surface Process and Landforms 21: 377-393.
Wooster, J. and S. Hilton (2004). Large woody debris volumes and accumulation rates in cleaned streams in redwood forests in  southern Humbolt County, California, United States Department of Agriculture, Forest Service: 1-8.
Wooten, R. M. (1988). Level I Stability Analysis Validation Report: 1-47.
Worona, M. and C. Whitlock (1995). "Late Quaternary vegetation and climate history near Little Lake, central Coast Range, Oregon." Geological Society of America Bulletin 107(7): 867-876.
Wright, H. E. J. (1974). "Landscape development, forest fires, and wilderness management." Science 186: 487-495.
Wright, J., et al. (2008). GIS-Based Atlantic Salmon Habitat Model, Draft, Appendix C. Atlantic Salmon Recovery Program, NOAA Fisheries Service.
Wright, K., et al. (1990). "Logging Effects on Streamflow: Storm Runoff at Caspar Creek in Northwestern California " Water Resources Research 26(7): 1657-1667.
Wright, K. K., et al. (2005). "Restricted hyporheic exchange in an alluvial river system: implications for theory and management." Journal of the North American Benthological Society 24(3): 447-460.
Large-scale patterns of hyporheic exchange are predictable within some river systems, but our understanding of the factors driving hyporheic processes and the magnitude of hyporheic exchange needed to influence biophysical patterns at larger scales remains limited. We investigated the patterns, magnitude, and potential effects on biota of reach-scale hyporheic exchange in an alluvial river of the Pacific Northwest. The river was topographically similar to and in the same geographic region as other systems where large-scale hyporheic exchange and associated biological responses have been observed. We hypothesized that predictable reach-scale patterns of hyporheic exchange would occur in alluvial valley segments of the river and that hyporheic upwelling would be associated with reach-scale patterns of physical and biological characteristics. We used in-channel piezometers and synoptic stream flow measurements to quantify hyporheic exchange. We measured temperature, dissolved O2, pH, specific conductivity, chlorophyll a biomass, primary production, and benthic macroinvertebrates as indicators of physical and biological responses. Contrary to our expectations, we found no evidence, physical or biological, of reach-scale hyporheic exchange. Hyporheic connectivity in this river system probably is constrained by geologic and geomorphic characters as well as the legacy of human land use in the basin. Thus, our results illustrate the variability of hyporheic processes that can occur among alluvial river systems and may have implications for watershed management.
 
Wu, J. and O. L. Loucks (1995). "From balance-of-nature to hierarchicalpatch dynamics:A paradigm shift in ecology." Quarterly Review of Biology 70: 439-466.
Wu, J. Q., A.C. Xu, and W.J. Elliott (2002). Adapting WEPP for forest watershed erosion modeling. 12th ISCO Conference, Beijing.
Wu, T. H. (1996). Soil strength properties and their measurement. Landslides, Investigation and Mitigation, Transportion Research Board, National Research Council, Special Report 247. A. K. Turner and R. L. Schuster. Washington, D. C., National Academy Press.
Wu, T. H. and A. Abdel-Latif (1996). Prediction and mapping of landslide hazard (FINAL DRAFT). Columbus, OH, Ohio State University: 41.
Wu, T. H. and D. N. Swanson (1980). "Risk of Landslides in Shallow Soils and Its Relation to Clearcutting in Southeastern Alaska." Forest Sci. 26(3): 495-510.
Wu, W. and R. C. Sidle (1995). "A distributed slope stability model for steep forested basins." Water Resources Research 31(8): 2097-2110.
Xu, J. (2004). "Comparison of hydraulic geometry between sand and gravel bed rivers in relation to channel pattern discrimination." Earth Surface Processes and Landforms 29: 645-657.
Yang, C. (1977). "The movement of sediment in rivers." Geophysical Surveys 3(1977): 39-68.
Yang, C. T. (1979). "Unit stream power equations for total load." Journal of Hydrology 40: 123-138.
Yang, C. T., and C. Huang (2001). "Applicability of sediment transport formulas." International Journal of Sediment Research 16(3): 335-353.
Yang, S. J. a. C. (1970). Hydraulic geometry of 12 selected stream systems of the United States. Urbana, Ill., Illinois State Water Survey.
Yang, S. L., J. Zhang, and X.J. Xu (2007). "Influence of the Three Gorges Dam on downstream delivery of sediment andf its environmental implications, Yangtze River." Geophysical Research Letters 34, L10401, doi10.1029/2007GL029472.
Yano, K. (1984). Development of a warning and evacuation system against debris flow. Internationales Symposion Interpraevent 1984, Villach.
Ye, M., et al. (2008). "On model selection criteria in multimodel analysis." Water Resources Research 44(W03428).
Hydrologic systems are open and complex, rendering them prone to multiple
conceptualizations and mathematical descriptions. There has been a growing tendency to
postulate several alternative hydrologic models for a site and use model selection criteria
to (1) rank these models, (2) eliminate some of them, and/or (3) weigh and average
predictions and statistics generated by multiple models. This has led to some debate
among hydrogeologists about the merits and demerits of common model selection (also
known as model discrimination or information) criteria such as AIC, AICc, BIC, and KIC
and some lack of clarity about the proper interpretation and mathematical representation of
each criterion. We examine the model selection literature to find that (1) all published
rigorous derivations of AIC and AICc require that the (true) model having generated the
observational data be in the set of candidate models; (2) though BIC and KIC were
originally derived by assuming that such a model is in the set, BIC has been rederived by
Cavanaugh and Neath (1999) without the need for such an assumption; and (3) KIC
reduces to BIC as the number of observations becomes large relative to the number of
adjustable model parameters, implying that it likewise does not require the existence of a
true model in the set of alternatives. We explain why KIC is the only criterion accounting
validly for the likelihood of prior parameter estimates, elucidate the unique role that the
Fisher information matrix plays in KIC, and demonstrate through an example that it
imbues KIC with desirable model selection properties not shared by AIC, AICc, or BIC.
Our example appears to provide the first comprehensive test of how AIC, AICc, BIC, and
KIC weigh and rank alternative models in light of the models’ predictive performance
under cross validation with real hydrologic data.
 
Yee, C. S. (?). Scour and Fill of Spawning Gravels in a Small Coastal Stream of Northwestern California funded jointly by US Forest Service and Humboldt State University.
Yee, C. S. and R. D. Harr (1977). Effect of wetting mode on shear strength of two aggregated soils. Portland, Oregon, USDA Forest Service: 9.
Yee, C. S. and R. D. harr (1977). "Influence of soil aggregation on slope stability in the Oregon Coast Ranges." Environmental Geology 1: 367-377.
Yitian, L. a. R. R. G. (2003). "Modeling flow and sediment transport in a river system using an artificial neural network." Environmental Management 31(1): 122-134.
Young, K. A. (2000). "Riparian Zone Management in the Pacific Northwest: Who's Cutting What?" Environmental Management 26(2): 131-144.
Young, M. K. (1991). "Selection of Measures of Substrate Composition to Estimate Survival to Emergence of Salmonids and to Detect Changes in Stream Substrates " North American Journal of Fisheries Management 11: 339-346.
Young, M. K. (1994). "Movement and characteristics of stream-borne coarse woody debris in adjacent burned and undistrubed watersheds in Wyoming." Canadian Journal of Forest Research 24: 1933-1938.
Following fire, changes in streamflow and bank stability in burned watersheds can mobilize coarse woody debris. In 1990 and 1991, I measured characteristics of coarse woody debris and standing riparian trees and snags in Jones Creek, a watershed burned in 1988, and in Crow Creek, an unburned watershed. The mean diameter of riparian trees along Jones Creek was less than that of trees along Crow Creek, but the coarse woody debris in Jones Creek was greater in mean diameter. Tagged debris in Jones Creek was three times as likely to move, and moved over four times as far as such debris in Crow Creek. In Jones Creek, the probability of movement was higher for tagged pieces that were in contact with the stream surface. Larger pieces tended to be more stable in both streams. It appears that increased flows and decreased bank stability following fire increased the transport of coarse woody debris in the burned watershed. Overall, debris transport in Rocky Mountain streams may be of greater significance than previously recognized.
 
Young, M. K. (1994). "Movement and characteristics of stream-borne coarse woody debris in adjacent burned and undisturbed watersheds in Wyoming." Canadian Journal of Fisheries and Aquatic Sciences 24: 1933-1938.
Young, M. K., et al. (2005). "Predicting cutthroat trout (Oncorhynchus clarkii) abundance in high-elevation streams: revisiting a model of translocation success." Canadian Journal of Fisheries and Aquatic Science 62: 2399-2408.
Assessing viability of stream populations of cutthroat trout (Oncorhynchus clarkii) and identifying streams
suitable for establishing populations are priorities in the western United States, and a model was recently developed to
predict translocation success (as defined by an index of population size) of two subspecies based on mean July water
temperature, pool bankfull width, and deep pools counts. To determine whether the translocation model applied to
streams elsewhere with more precise abundance estimates, we examined the relation between electrofishing-based esti-mates
of cutthroat trout abundance and these habitat variables plus occupied stream length. The preferred model was
(population size)1/2 = 0.00508(stream length (m)) + 5.148 (N = 31). In contrast, a model based on data from the origi-nal
translocation model included stream temperature and deep pool counts as variables. Differences in models appear to
largely have a methodological rather than biological basis. Additional habitat coupled with increased habitat complexity
may account for the form of the abundance – stream length relation in the electrofishing-based model. Model-derived
estimates imply that many cutthroat trout populations are below thresholds associated with reduced risk of extinction.
We believe that this model can reduce uncertainty about projected population sizes when selecting streams for
reintroductions or evaluating unsampled streams.
 
Young, M. K. and W. A. Hubert (1990). "Fines in Redds of Large Salmonids " Transactions of the American Fisheries Society 119: 156-162.
Young, M. K., et al. (2006). "Characterizing and contrasting instream and riparian coarse wood in western Montana basins." Forest Ecology and Management 226: 26-40.
The importance of coarse wood to aquatic biota and stream channel structure is widely recognized, yet characterizations of large-scale patterns in
coarse wood dimensions and loads are rare. To address these issues, we censused instream coarse wood ( 2 mlong and 10 cm minimum diameter)
and sampled riparian coarse wood and channel characteristics in and along 13 streams in westernMontana. Instream coarse wood tended to be shorter
but of larger diameter than riparian pieces, presumably because of fluvial processing. Instream coarse wood also displayed highly variable spatial
patterns.Most segments lacked significant spatial correlation in coarsewood abundance in adjacent 50 mreaches andwhen present, coarsewood patch
sizes (100–1200 m) were specific to particular streams. Estimation of instreamand riparian piece dimensionswithin 25%of themean required samples
of 13–314 pieces, whereas estimation of wood loads instream segments required samples of 8–210 reaches (400–10 500 m). If these results are
representative of other systems, few previous studies have used sample sizes adequate to characterize instream coarse wood loads.
 
Young, W. J., et al. (2002). Predicting channel type from catchment and hydrological variables. The Structure, Function and Management Implications of Fluvial Sedimentary Systems. F. J. Dyer, M. C. Thoms and J. M. Olley, Interenational Association of Hydrological Sciences. 276: 53-60.
Yu, B. and M. G. Wolman (1987). "Some Dynamic Aspects of River Geometry." Water Resources Research 23(3): 501-509.
Yue, J. and J. L. Anderson (1990). Numerical Modeling of a Trench Fill and Scour Hydraulic Engineering, American Society of Civil Engineers.
Yulianti, J. S., et al. (1999). "Non-point source water quality management under input information uncertainty."
Non-point source pollution in the form of erosion can be controlled with watershed management schemes that identify the optimum set of farming practices for given conditions of topography, climate, crop yield, crop prices and allowable sediment load. However, since input conditions may be difficult to estimate or may vary naturally, it is important to develop approaches for determining good farming practices while accounting for the uncertainty of the input conditions. This work identifies: (1) the sensitivity of different watershed management policies to uncertain input information; (2) the input information that is significant for modelling and managing erosion and sedimentation; and (3) the farming practices that are least sensitive to uncertainty in the input information. Monte Carlo Simulation, and Modified Generalized Sensitivity and Regret Analyses are applied in combination to address these issues and this approach is demonstrated for the control of erosion and sedimentation in the Highland Silver Lake Watershed in Illinois. Two different watershed management policies, the Erosion Standard and Erosion Tax Policies, and a benchmark representing a least cost bound, the Least Opportunity Cost, are examined. Results show that for the Highland Silver Lake Watershed the Least Opportunity Cost approach yields the least variable solution with respect to opportunity cost and that the Erosion Standard Policy produces the most variable "optimal" opportunity cost values under different watershed conditions, crop yields and crop prices. The rainfall erosivity factor and the prices of corn and soybeans are the most important parameters in the linked sediment delivery and economic model used in this work and the importance of the crop yield parameters depends on the management policy chosen. The Regret Analysis shows that the most often selected "optimal" sets of farming practices under many scenarios of uncertain input data and parameters vary depending on how regret is expressed, but in general, only a small selection of farming practices are identified to be relatively less sensitive to uncertain inputs.
 
Yuyi, W., et al. (1992). Shearing stress, strain, and debris flow --taking the erosion and deposition of debris flow in Jiangjia gully as an example. Interpraevent 1992, Bern.
Zecharias, Y. B. and W. Brutsaert (1988). "The influence of basin morphology on groundwater outflow." Water Resources Research 24(10): 1645-1650.
Zecharias, Y. B. and W. Brutsaert (1988). "recession characteristics of groundwater outflow and base flow from mountainous watersheds." Water Resources Research 24(10): 1651-1658.
Zevenbergen, L. W. and C. R. Thorne (1987). "Quantitative analysis of land surface topography." Earth Surface Processes and Landforms 12: 47-56.
Zhang, L., et al. (1999). "Estimation of soil moisture and groundwater recharge using the TOPOG_IRG model." Water Resources Research 35(1): 149-161.
Zhang, W., and D.R. Montgomery (1994). "Digital elevation model grid size, landscape representation, and hydrologic simulations." Water Resources Research 30(4): 1019-1028.
Zhang, W. and T. W. Cundy (1987). "Laminar Einstein bed load transport equation for overland sheet flow." Journal of Hydraulic Engineering 113(12): 1525-1538.
Zhang, W. and T. W. Cundy (1989). "Modeling of two-dimensional overland flow." Water Resources Research 25(9): 2019-2035.
Zhang, W. and D. R. Montgomery (1994). "Digital elevation model grid size, landscape representation, and hydrologic simulations." Water Resources Research 30(4): 1019-1028.
Zhang, x., et al. (1999). "Comparison of slope estimates from low resolution DEMs: Scaling issues and a fractal method for their solution." Earth Surface Processes and Landforms 24: 763-779.
Zhang, X. C., M.A. Nearing,W.P. Miller, L.D. Norton, and L.T. West (1998). "Modeling interrill sediment delivery." Soil Science Society of America Journal 62(2): 438-444.
Zheng, F. L., et al. (2000). "Vertical hydraulic gradient and run-on water and sediment effects on erosion processes and sediment regimes." Soil Science Society of America Journal 64(1): 4-11.
Zhou, Q. and X. Liu (2004). "Error analysis on grid-based slope and aspect algorithms." Photogrammetric Engineering and Remote Sensing 70(8): 957-962.
Zhu, A. X., and D.S. Mackay (2001). "Effects of spatial scale of soil information on watershed modeling." Journal of Hydrology 248: 54-77.
Ziegler, A. D., R. A. Sutherland, and T. W. Giambelluca (2000). "Runoff generation and sediment production on unpaved roads, footpaths and agricultural land surfaces in northern Thailand." Earth Surface Processes and Landforms 25(5): 519-534.
Rainfall simulation was used to examine runoff generation and sediment transport on roads, paths and three types of agricultural fields in Pang Khum Experimental Watershed (PKEW), in mountainous northern Thailand. Because interception of subsurface flow by the road prism is rare in PKEW, work focused on Horton overland flow (HOF). Under dry antecedent soil moisture conditions, roads generated HOF in c. 1 min and have event runoff coefficients (ROCs) of 80 per cent, during 45 min, c. 105 mm h-1 simulations. Runoff generation on agricultural fields required greater rainfall depths to initiate HOF; these surfaces had total ROCs ranging from 0 to 20 per cent. Footpaths are capable of generating erosion-producing overland flow within agricultural surfaces where HOF generation is otherwise rare. Paths had saturated hydraulic conductivity (Ks) values 80-120 mm h-1 lower than those of adjacent agricultural surfaces. Sediment production on roads exceeded that of footpaths and agricultural lands by more than eight times (1·23 versus &lt; 0·15 g J-1). Typically, high road runoff volumes (owing to low Ks, c. 15 mm h-1) transported relatively high sediment loads. Initial road sediment concentrations exceeded 100 g l-1, but decayed with time as loose surface material was removed. Compared with the loose surface layer, the compacted, underlying road surface was resistant to detachment forces. Sediment concentration values for the road simulations were slightly higher than data obtained from a 165 m road section during a comparable natural event. Initial simulation concentration values were substantially higher, but were nearly equivalent to those of the natural event after 20 min simulation time. Higher sediment concentration in the simulations was related to differences in the availability of loose surface material, which was more abundant during the dry-season simulations than during the rainy season natural event. Sediment production on PKEW roads is sensitive to surface preparation processes affecting the supply of surface sediment, including vehicle detachment, maintenance activities, and mass wasting. The simulation data represent a foundation from which to begin parameterizing a physically based runoff/erosion model to study erosional impacts of roads in the study area. Copyright © 2000 John Wiley & Sons, Ltd.
 
Ziemer, R. R. (1981). "Roots and the stability of forested slopes." I.A.H.S. Publ. 132: 343-357.
Ziemer, R. R. (1981). Roots and the Stability of Forested Slopes. Erosion and Sediment Transport in Pacific Rim Steeplands. T. R. H. Davies and A. J. Pearce, International Association of Hydrological Sciences. 132: 343-361.
Ziemer, R. R. (1981). "Storm Flow Response to Road Building and Partial Cutting In Small Streams of Northern California." Water Resources Research 17(4): 907-917.
Ziemer, R. R., T.E. Lisle, R.M. Rice, and J. Lewis (1991). "Modeling the cumulative watershed effects of forest management strategies." Journal of Environmental Quality 20(1): 36-42.
Ziemer, R. R. (1998). Flooding and Stormflows US Forest Service: 15-24.
Ziemer, R. R., et al. (1991). "Modeling the Cumulative Watershed Effects of Forest Management Strategies " J. Environ. Qual. 20: 36-42.
Zienkiewicz, O. C. and R. L. Taylor (1988). The Finite Element Method. London, McGraw-Hill.
Zimmerman, R. C., et al. (1967). The influence of vegetation on channel form in small streams. Sympoiusm on River Morphology, Chirstchurch, New Zealand, International Association of Hydrological Sciences.
Zybach, B. (1988). The great fires of the Oregon Coast range: 1770-1933 and their basic effects upon current vegetation and property ownership patterns, Oregon State University.
Zybach, B. (1993). Forest History and Femat Assumptions: A Critical Review of President Clinton's 1993 Northwest Forest Plan. Corvallis, Oregon, prepared for American Forest and Paper Association and the Northwest Forest Resource Council: 101.
Zybach, B. (1994). Siletz Gorge Forest History: Methods and Recommendations for Using Northwest Forest History Information with GIS Applications. Corvallis, Oregon, prepared for the Center for Environmental Studies: 81.
 
Built with help of Dr.Explain