Você está na página 1de 25

Catena 64 (2005) 247 271 www.elsevier.

com/locate/catena

A review of hillslope and watershed scale erosion and sediment transport models
Hafzullah Aksoy a,*, M. Levent Kavvas b
a

Istanbul Technical University, Department of Civil Engineering, Hydraulics Division 34469 Maslak, Istanbul, Turkey b University of California, Department of Civil and Environmental Engineering, Davis, CA 95616, USA

Abstract This study reviews the existing erosion and sediment transport models developed at hillslope and watershed scales. The method followed in this review is to summarize the models with a focus on the physically based modeling technique as well as with a brief discussion about empirical and conceptual models. Approaches for determining the sediment transport capacity of flow are explained. The extension of a sediment transport model to a nutrient transport model is then discussed. Finally, the future of erosion and sediment transport models are projected to include the probabilistic description of hydrology, the physical characteristics of the watershed, and the stochastic structure of soil properties. The review is expected to be of interest to researchers, watershed managers and decision-makers while searching for models to study erosion and sediment transport phenomena and related processes such as pollutant and nutrient transport. D 2005 Elsevier B.V. All rights reserved.
Keywords: Conceptual models; Empirical models; Erosion; Physically based models; Sediment transport; Upland erosion

1. Definitions and basic concepts Defined by ASCE Task Committee (1970) as the loosening or dissolving and removal of earthy or rock materials from any part of the earths surface, erosion is a process of detachment and transportation of soil materials by erosive agents (Foster and Meyer,
* Corresponding author. Fax: +90 212 2856587. E-mail address: haksoy@itu.edu.tr (H. Aksoy). 0341-8162/$ - see front matter D 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.catena.2005.08.008

248

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

1972). Erosion can be caused by wind (wind erosion), by rainfall (rainfall erosion), or by runoff (runoff erosion). Runoff erosion can happen in unconcentrated flow (sheet erosion), in rills (rill erosion), or gullies (gully erosion). Rills are such small concentrations of running water that they can be completely removed by normal cultivation methods, whereas gullies cannot be. Erosion in the channel is called channel erosion. Soil eroded from a given area is defined in terms of rate of erosion. Total sediment outflow from a watershed per unit time is called sediment yield. It is obtained by multiplying the sediment loss by a delivery ratio (Novotny and Chesters, 1989). The transported portion of the eroded sediment (ratio of yield to the total eroded material) is called sediment delivery or sediment delivery ratio. Sediment delivery decreases with increasing basin size as large basins have more sediment storage sites where eroded sediment is kept. Sediment delivery can be limited by reducing either the detachment rate or the transport capacity depending on which has a lower value. Sediment particles are detached from their current places if the sediment load in the flow is smaller than the sediment transport capacity of flow. For this, the shear stress exerted by flow should be greater than the critical shear stress that is required for sediment particles to be removed from their current locations. Otherwise deposition is observed whenever the flow has a sediment load exceeding its transport capacity. This is illustrated in Table 1 where three movement modes of sediment are given. Entrainment is the actual rate of mass temporarily leaving the overland surface, which may or may not be transported. Deposition is the actual rate of mass temporarily reaching the overland surface. If deposition exceeds entrainment, then net deposition takes place; if entrainment exceeds deposition, then net erosion takes place. When deposition is equal to entrainment, then the sedimentation process is said to be in equilibrium (Croley, 1982). Annual amount of sediment eroded from a watershed is called annual gross erosion. A small portion (less than one-fourth) of the eroded sediment is delivered to the receiving bodies (e.g. sea, inland lakes, streams, etc.) while the remaining is deposited in their way. The major part of the annual sediment discharge is transported in a short period of time by a few storms during which the discharge of the stream is continuously changing. Therefore temporal variation of the stream discharge is very important (Bennett, 1974). Sediment concentration is higher at higher rainfall intensity due to its higher detachability capacity. With increasing flow discharge sediment concentration decreases. Sediment concentration is the highest at the beginning of the rainfall and decreases until the steady state is reached. Detachability or re-detachability, hence soil loss, is expected to decrease as the water depth in the flow increases (Proffitt et al., 1991). The decrease in soil loss at the greater water depth can be attributed to the increased protection of the soil
Table 1 Movement mode of sediment for different cases of transport capacity (Tc) of flow and sediment load ( Q s) in the flow Case I II III Tc b Q s Tc = Q s Tc N Q s Deposition Transport Erosion

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

249

surface from the raindrop impact as the mean water depth of surface water increases (Singh and Prasad, 1982). Factors affecting water erosion are climate, topography, soil, vegetation and antropogenic activities such as tillage systems and soil conservation measures (Kuznetsov et al., 1998). Natural vegetative cover is sparse in arid regions and, therefore, runoff events cause entraining very high concentrations of sediment resulting in a limitless sediment supply. However, in the more humid regions, the vegetation usually prevents runoff from entraining soil. Therefore, the antecedent soil conditions become important although they are region dependent. Winter storms result in less sediment transport than summer storms. However, large amounts of precipitation and runoff occurring during winter could cause high erosion rates if the soil cover is minimal (ASCE Task Committee, 1970). This fact was observed by Emmett (1970) who concluded, based on nearly 10-year experimental data, that sediment concentration in overland flow is negatively correlated with vegetation cover of the region. Sediment transport is sensitive to the flow hydraulics. Therefore, when applying physically based models, detailed and accurate simulation of the velocity and depth of flow is more important than an accurate calibration of rainfall erosion and runoff erosion parameters although, in some cases, erosion can be sensitive to those parameters (Smith et al., 1999). An bupland areaQ in a watershed is where surface runoff can be considered as overland flow in hydrological analysis. Upland erosion is affected by hydrology, topography, soil erodibility and transportability, vegetation cover, land use, subsurface effects, tillage roughness and tillage marks (Foster, 1982). Sediment discharge becomes a function of hydraulic properties of flow, physical properties of soil, and surface characteristics. On the upland area, four different processes accomplish sediment removal and transport: detachment by raindrop impact, detachment by runoff, transport by raindrop splash, and transport by runoff. Flow and raindrop detachment rates are not simple, and they are, therefore, given by empirical formulas (Bennett, 1974). The inception of sediment motion on a slope was found to be dependent upon the slope plus the Reynolds number of the flow over the slope (Lau and Engel, 1999). Increasing bed slope reduces the critical shear stress that should be exceeded for a sediment particle to start its motion. The upland area is usually considered to have both interrill areas and rills. Consequently, erosion in such areas is divided into rill erosion and interrill erosion. The upland sediment load is given by the detachment (or deposition) rate in rills plus delivery rate of particles detached by interrill erosion to rill flow (Foster and Meyer, 1975). Interrill erosion process is rainfall dominated, whereas rill erosion is mostly defined by runoff. It is important to note that rainfall erosion controls the rising limb of the sediment graph while runoff (flow) erosion rates are closely related to the recession (Park et al., 1982). Erosion in rills is due mainly to flow detachment. An equation such as that of Yalin (1963), developed for streamflow, seems to be applicable for determining the transport capacity of flow in the rill. Gully erosion is, in general, similar to rill erosion. Gully erosion and rill erosion differ from rainfall erosion in that raindrop impact is not important (Foster and Meyer, 1975). The initiation and development of gully erosion is a response to widespread clearing and

250

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

intensive land use. Landslides are other kinds of erosion that can occur in a basin. Different kinds of assessment techniques are available. Montgomery and Dietrich (1994) proposed a slope stability model for predicting the location of shallow landsliding. Increasing rainfall rate makes the catchment unstable. Steep channels in the basin headwaters transport huge debris flow, while the lower-gradient channels in the major valleys are within the depositional zone. Lowland streams in which the flow is usually perennial carry coarser bed material. It is usually assumed that the bed material transport velocity is the same as the mean stream velocity and that the longitudinal dispersion is negligible (Bennett, 1974). Lowland (channel) erosion consists of streambed and stream bank erosions. A simple way to predict erosion in the channel is to use the formula QS~Gs ds 1

in which Q is stream discharge, S slope of stream channel, G s bed load sediment discharge and d s particle diameter of bed load sediment (ASCE Task Committee, 1970). Flow velocity near the banks of the channel is much smaller than that in the center. This results in relatively coarse materials being moved down while fine material being deposited near the banks, and suggests the use of two-dimensional models for stream channel processes. However, practical considerations force one to use one-dimensional models. This effect can, for instance, be considered by subdividing the cross-sectional area of the channel into mid-channel and side-channel subsections. Alternatively, an effective depth can be defined in order to take the variability along the channel cross-section into account (Johanson and Leytham, 1977). The following section summarises the existing watershed and hillslope scale erosion and sediment transport models with a focus on those that are physically based. It is then explained how the transport capacity of flow is determined. Data requirements of the models are discussed. Extension of a sediment transport model to a nutrient transport model is given in a later section together with some existing nutrient transport models. Finally the review is summarised and projections for the future are drawn.

2. Existing erosion and sediment transport models A review of watershed hydrology models was recently given by Singh and Woolhiser (2002). Three review papers were observed in literature for erosion and sediment transport models developed at hillslope and watershed scales. Bryan (2000) performed a review on the water erosion on hillslopes while Zhang et al. (1996a) reviewed modelling approaches used for the prediction of soil erosion in catchments. Models mentioned in that review were limited to a number of very well known models only. The third one (Merritt et al., 2003) gave the most recent review that analyzes specific models based on model input output, model structure, runoff, erosion/transport and water quality modeling, and accuracy and limitation of the model. Some models are similar, because they are based on the same assumptions and some are distinctly different. Models so far developed can be categorized according to different

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

251

criteria that may encompass process description, scale, and technique of solution (Singh, 1995). A model may be based on a conceptual or an empirical framework. Such models are called conceptual and empirical models, respectively. In conceptual models, a watershed is represented by storage systems. Empirical models are limited to conditions for which they have been developed. If a model is constructed by using mass conservation equation of sediment, it is called a physically based erosion and sediment transport model. For instance, the USLE (Wischmeier and Smith, 1978) is an empirical model which is based on a large amount of data from the United States. AGNPS (Young et al., 1989) uses a modified form of USLE. The hydrological part of ANSWERS (Beasley et al., 1980) is a conceptual process. KINEROS (Smith, 1981), WESP (Lopes, 1987), SEM (Storm et al., 1987), SHESED (Wicks, 1988) and EUROSEM (Morgan et al., 1998) are some examples for the physically based erosion and sediment transport models. Table 2 categorises the erosion and sediment transport models depending upon the process description used in their formulation. These models will be discussed into detail in the following sections. Most of the models in Table 2 are physically based although the first attempts in the modeling of erosion and sediment transport started with empirically based approaches of the USLE. Also some conceptual models such as LASCAM (Viney and Sivapalan, 1999) were developed. 2.1. Empirical and conceptual models Given below are summaries of the empirical and conceptual erosion and sediment transport models listed in Table 2.
Table 2 Erosion and sediment transport models Model USLE MUSLE RUSLE SEDD AGNPS LASCAM ANSWERS LISEM CREAMS WEPP EUROSEM KINEROS KINEROS2 RUNOFF WESP CASC2D-SED SEM SHESED Empirical Conceptual Physically based

252

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

2.1.1. USLE The Universal Soil Loss Equation (USLE) (Wischmeier and Smith, 1978) is given by E RKSLCP 2

where E is average annual soil loss in tons/acres, R rainfall erosivity index, K soil erodibility index, S slope, L length of the slope, C cropping management factor and P supporting conservation practice factor. The equation is based on a huge amount of data from the United States. The USLE computes annual soil loss. Its modified version (MUSLE) has been an attempt to compute soil loss for a single storm event. The USLE was revised (RUSLE) (Renard et al., 1991) and revisited (Renard et al., 1994) for improvement. 2.1.2. SEDD The SEdiment Delivery Distributed (SEDD) model, which is based on the empirical USLE model was proposed by Ferro and Porto (2000). A Monte Carlo technique was used to test the effect of uncertainty in the model parameters on sediment yield computations, similar to the study by Biesemans et al. (2000) on the RUSLE. 2.1.3. AGNPS An event-based model, AGNPS (AGricultural NonPoint Source) simulates runoff, sediment and nutrient transport from agricultural watersheds. The model divides the watershed into square cells uniformly distributed over the watershed. The erosion and sediment transport component is based on estimating the upland erosion by the USLE and routing it by the steady-state continuity equation of sediment. Eroded soil is subdivided into five size classes: Clay, silt, small aggregates, large aggregates and sand. The model produced favourably comparable results for runoff and sediment (Young et al., 1989). AGNPS was applied to two medium size (80130 km2) watersheds in Central Europe (Rode and Fredo, 1999; Pekarova et al., 1999). Although total runoff volume was simulated with small deviations from observations in both watersheds, sediment was computed satisfactorily for only one of the two watersheds. 2.1.4. LASCAM A continuous (daily time interval), conceptual sediment generation and transport algorithm was coupled to an existing water and salt balance model, LASCAM (Viney and Sivapalan, 1999). LASCAM was originally developed to predict the effect of land use and climate change on the daily trends of water yield and quality in forested catchments in Western Australia. The developed sediment transport algorithm does not discriminate between sediment size classes. It was found that the amount of runoff and sediment produced by the model was matched well in monthly and daily time intervals. Viney et al. (2000) later coupled a conceptual model of nutrient mobilisation and transport to the LASCAM. 2.2. Physically based models Table 3 focuses on watershed scale physically based erosion and sediment transport models for which detailed information is provided below.

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

Table 3 Physically based erosion and sediment transport models Model ANSWERS LISEM CREAMS WEPP EUROSEM KINEROS KINEROS2 RUNOFF WESP CASC2D-SED SEM SHESED Lumped Distributed Stochastic Deterministic 1-D 2-D Steady Unsteady Event-based Continuous Rilled No rilled Single-size Multi-size state state structure structure

253

254

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

Process based classification divides this type of models into lumped and distributed. A lumped model uses single values of input parameters with no spatial variability and results in single outputs. A distributed model, however, uses spatially distributed parameters and provides spatially distributed outputs by taking explicit account of spatial variability of the process. A model can be considered deterministic or stochastic depending upon the way the process is described. In physically based models, the model is called one- or twodimensional depending upon the number of dimensions of the mass conservation equation used in the model. Erosion and sediment transport models generally take the nonstationarity in the erosion process into account although a number of them are interested only in the steady state case. A model is called an event-based model if it is used for the simulation of sediment produced by one single rainfallrunoff event. A continuous model is used for the simulation of sediment due to many consecutive rainfallrunoff events occurring during a season or longer time period. Single-size erosion and sediment transport models can only predict sediment transport for a mean grain size and can give the total sediment mass leaving the catchment. The sediment size distribution is very important in sediment quality since pollutants are usually sorbed to finest particles. This is achieved in multi-size models. In a similar manner, models with rilled structure perform better in the simulation of the natural topography in the watershed. 2.2.1. ANSWERS The ANSWERS (Areal Nonpoint Source Watershed Response Simulation) model (Beasley et al., 1980) includes a conceptual hydrological process and a physically based erosion process. The erosion process assumes that sediment can be detached by both rainfall and runoff but can only be transported by runoff. ANSWERS model divides a watershed into small, independent elements. Within each element the runoff and erosion processes are treated as independent functions of the hydrological and erosion parameters of that element. In the model, surface conditions and overland flow depth in each element are considered uniform. No rilling is considered. The effect of rills is assumed to be described by the roughness coefficient of the Manning equation used in the model. According to ANSWERS subsurface return flow and tile drainage are assumed to produce no sediment. A detached sediment particle is reattached to the soil, if it deposits. Detachment of such a particle requires the same amount of energy as required for the original detachment. Channel erosion is negligible. In the erosion part, the differential equation given by Foster and Meyer (1972) is used. Preparing input data file for ANSWERS is rather complex (Norman, 1989) as it is the case for many physically based hydrology and erosion and sediment transport models. The model can be considered a tool for comparative results for various treatment and management strategies (Beasley et al., 1980). Park et al. (1982) added a sediment transport component to the previously developed ANSWERS hydrological model. In this new form ANSWERS is a single event model, and it uses distributed parameters. Also a channel erosion component is included in the new version. 2.2.2. LISEM Because of spatial and temporal variation in runoff and soil erosion processes, GIS has been a very useful tool to use in hydrological applications. The LImburg Soil Erosion

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

255

Model (LISEM) (De Roo et al., 1996) is one of the first models that use GIS. Although it is physically based, LISEM mostly uses empirically derived equations. The model, in the soil erosion part, accounts also for roads, wheel tracks and channels. There are some indices used for prediction of the soil erosion (De Roo, 1998). For example, wetness index, defined as the natural logarithm of A s / S where A s = contributing area and S = slope gradient, can be used to identify possible stream paths. Wetness index can also be used for indicating wet and dry areas, and thus possible source areas for saturation overland flow, which is one of the main causes of erosion. The wetness index originally comes from the TOPMODEL of Beven and Kirkby (1979). Flow detachment risk is related to the stream power index, which is the product of the unit contributing area and slope gradient. There is also a sediment transporting capacity index, which is again a function of contributing unit area and slope. The above given indices are used to give the soil erosion hazard index. GIS can be used in obtaining the index maps of the basin. CATSOP experimental watershed located in the Limburg area in the Netherlands was used for calibration and validation of the LISEM model parameters. EROSION 2D/3D (Schmidt et al., 1999) is another model, with similar structure to LISEM, for which an application based on the same data set from the CATSOP watershed was performed. 2.2.3. CREAMS The sediment transport component of CREAMS (Chemicals, Runoff, and Erosion from Agricultural Management Systems) analyzes the interrill area and rill separately. Detachment on both rill and interrill area is determined by the modified USLE. The procedure allows parameters to change along the overland flow profile and along waterways to describe spatial variability (Foster et al., 1981). 2.2.4. WEPP WEPP (Water Erosion Prediction Project) (Nearing et al., 1989) is a model to predict soil erosion and sediment delivery from fields, farms, forests, rangelands, construction sites and urban areas (Laflen et al., 1997). It is a daily continuous model. WEPP divides runoff between rills and interrill areas. Consequently, it calculates erosion in the rills and interrill areas separately. The steady-state sediment continuity equation is used to predict rill and interrill processes (Nearing et al., 1989). Rill erosion occurs if the shear stress exerted by flow exceeds the critical shear stress while sediment load in the flow is smaller than the transport capacity of flow. Interrill erosion is considered to be proportional to the square of the rainfall intensity. Interrill area delivers sediment to rills. The model solves the non-dimensional (normalized) detachment and deposition equations. The normalized load is calculated and then is converted to the actual load. It was found by Zhang et al. (1996b) that the model was reliable in predicting long term averages of soil loss under cropped conditions. 2.2.5. EUROSEM The EUROpean Soil Erosion Model, EUROSEM, (Morgan et al., 1998) is a model for predicting soil erosion by water from fields and small catchments. The model was designed as an event-based model, since it was thought that erosion was dominated by only a few events per year. Moreover, continuous models require substantial amount of

256

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

data, as many of them are not physically measurable at field or in the laboratory. EUROSEM is a dynamic erosion model and is able to simulate sediment transport, erosion and deposition by rill and interrill processes over the hillslope. The model provides total runoff, total soil loss, storm hydrograph and storm sediment graph. In EUROSEM, soil detachment by raindrop impact is the sum of the direct throughfall and leaf drainage and it depends on the kinetic energy of the rainfall. Splash erosion takes place before runoff begins. Therefore, initial sediment concentration should be taken as a non-zero value. Three cases were considered for hillslope erosion: hillslopes without rills, hillslopes with rills and interrill areas, and hillslopes with very dense rill structure. The channel erosion process is treated similar to the rill erosion with the exception that the raindrop impact is neglected and lateral inflow of sediment to the channel from the hillslope becomes important. Bank collapse is not simulated (Morgan et al., 1998, 1999; Kinnell, 1999). EUROSEM was applied to the CATSOP experimental watershed in the Netherlands. This is a watershed on which LISEM (De Roo et al., 1996) and EROSION 2D/3D (Schmidt et al., 1999) were also applied. The model was found to be useful for the short duration storms, which were characterized by a single pulse of rainfall (Folly et al., 1999). Veihe and Quinton (2000) and Veihe et al. (2000) used Monte Carlo simulations for the sensitivity analysis of hydrological, soil and vegetation parameters of EUROSEM as well as the effect of rills and rock fragments. Hydrological parameters were found to be the most important parameters. Detachability and cohesion of the soil were also found to be important but vegetation parameters were found to have insignificant effect. 2.2.6. KINEROS KINEROS (KINematic EROsion Simulation) (Smith, 1981; Woolhiser et al., 1990) is composed of elements of a network, such as planes, channels or conduits, and ponds or detention storages, connected to each other. KINEROS is an extension of KINGEN, a model developed by Rovey et al. (1977), with incorporation of erosion and sediment transport components. The sediment component of the model is based upon the onedimensional unsteady state continuity equation. Erosion/deposition rate is the combination of raindrop splash erosion and hydraulic erosion/deposition rates. Splash erosion rate is given by an empirical equation in which the rate is proportional to the second power of the rainfall. Hydraulic (runoff) erosion rate is estimated to be proportional to the transport capacity deficit, which is the difference between the current sediment concentration in the flow and steady state maximum concentration. Hydraulic erosion may be positive or negative depending upon the local transport capacity. A modified form of the equation of Engelund and Hansen (1967) was used for determining the steady state flow concentration. A single-mean sediment particle size was used in the formulation. KINEROS does not explicitly separate rill and interrill erosion. Channel erosion is taken the same as the upland erosion except for the omission of the splash erosion as it is no longer effective on erosion in the channel phase. Soil and sediment are characterised by a distribution of up to five size class intervals in the new version of the model, KINEROS2 (Smith et al., 1995a,b). Smith et al. (1999) applied the model to a catchment in the Netherlands. It was also applied to a catchment in Northern Thailand to see its applicability for unpaved mountain roads (Ziegler et al., 2001).

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

257

2.2.7. RUNOFF The sediment transport component of RUNOFF (Borah, 1989) computes soil erosion and routes the sediment to the downstream end of the slope on which flow occurs. The model has two parameters: flow detachment coefficient that should be calibrated by the observed data, and the raindrop detachment coefficient that is fixed. Although the model simulated the sediment discharge reasonably, there are, however, some discrepancies due to parameters fixed in time. 2.2.8. WESP Using the one-dimensional continuity equation for sediment transport Lopes (1987) and Lopes and Lane (1988) developed a physically based, event-oriented mathematical model for sedimentation in small watersheds. The sediment-flux term of the model for overland flow area is an outcome of the sediment entrainment by overland flow shear stress, the rate of sediment entrainment by rainfall impact and the rate of sediment deposition. In the model, it is assumed that both erosion and deposition occur simultaneously. Erosion by overland flow shear stress is proportional to a power of the average shear stress acting on the soil surface. ER KR sk : 3

In Eq. (3), K R is soil detachability factor, s average effective shear stress, and k an exponent equal to 1.5 in Lopes (1987). Note that in this formulation there is no critical shear stress that should be exceeded for the initiation of sediment particles. WESP assumes there are always fine particles of sediment, detached by the action of wind or other elements between storm events which will be available for transport by sheet flow as soon as rainfall exceeds infiltration rate, without any resistance to removal (Lopes, 1987). Sediment deposition rate, simultaneously occurring together with erosion, is given by d aVs Cs 4

where a is dimensionless coefficient depending upon the soil and fluid properties, V s particle fall velocity, and C s sediment concentration. In the case of uniform rainfall intensity (r), detachment by raindrop impact (E I) is given as EI ar2 5

where a is a coefficient to be calibrated. The channel erosion part of the model uses a standard continuity equation with a source term composed of the rate of entrainment by channel flow, rate of sediment deposition and lateral sediment inflow from surrounding overland flow areas. A critical shear stress needs to be exceeded for initiation of sediment in the channel bed. There will be no sediment as long as the flow shear stress is smaller than the critical shear stress. The model was applied to rainfall simulator and natural data sets. Hydrographs were simulated well, whereas sedigraphs were not. However, the total sediment load was reproduced quite well. Using the same method as in WESP, Santos et al. (1998) found that the initial moisture content of the soil influenced the runoff hydrograph and hence the sedigraph. Hydrograph and sedigraph of a rainfall event, occurring very shortly after the previous one, were

258

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

simulated well. WESP was modified by Santos et al. (2000) for large watersheds. Instead of using the simultaneous erosion and deposition concept of the original WESP model, flux of sediment transport by overland flow was calculated as the difference between the erosion and deposition. 2.2.9. CASC2D-SED The upland erosion routine of the physically based hydrological model CASC2D was introduced by Johnson et al. (2000). The hydrological model uses two-dimensional continuity and momentum equations for runoff and adopts the diffusion wave approximation. In the upland erosion part of the model, transport capacity of flow is determined by a modified version of the regression equation given by Kilinc and Richardson (1973). Although runoff hydrographs were computed reasonably well, sedigraphs could not be simulated adequately. The sediment yield was found within a range of 50% to 200%. 2.2.10. SEM A distributed soil erosion and sediment transport model (SEM) (Storm et al., 1987) was incorporated into the SHE hydrological modeling system (Abbott et al., 1986a,b). SEM simulates the spatial and temporal variation of soil erosion in catchments. The splashdetached soil particles are transported by overland flow. Overland flow itself has a detachment potential, which was called flow entrainment, and was taken equal to the transport capacity of flow. The net erosion or deposition is calculated as the difference between the sediment load entering and leaving each grid in the catchment. The model has two parameters to be calibrated from the available data that are related to the soil erodibility and flow entrainment. 2.2.11. SHESED SHESED (Wicks, 1988) is the sediment transport component of the SHE hydrological model (Abbott et al., 1986a,b). SHESED considers erosion as the sum of erosion by raindrop and leaf drip impacts and that by overland flow. Erosion takes place in the channel bed too. The eroded sediment is transported by overland flow to channels. Once the eroded sediment gets to the channel, it is further transported downstream. Soil erosion by raindrop and leaf drip impacts is given by an equation based on the theoretical work of Storm et al. (1987). The overland flow soil detachment is given by an equation accounting for interrill areas and rills together. Therefore rills are not accounted for explicitly in the model. Ground cover, given in the raindrop detachment equation, is low-lying cover, which shields the soil from raindrop impact erosion. Canopy cover refers to taller vegetation, which shields the soil from the direct impact of the raindrops but allows the rainwater to coalesce on its surface and fall to the ground as large leaf drips (Wicks and Bathurst, 1996). In SHESED, overland flow and sediment transport are based upon the two-dimensional mass conservation equations. Either the Ackers and White (1973) equation or the Engelund and Hansen (1967) equation is used in determining the transport capacity of flow. Selection of the transport capacity equation in SHESED is based upon a trial and error technique, and is chosen in the calibration stage of the model. Also the raindrop and overland flow erodibility coefficients are calibrated. The sediment yield

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

259

simulations showed sensitivity to the erodibility coefficient. Therefore, accurate calibration is needed (Wicks et al., 1992). Particle size distribution is not considered. The equation is solved by an explicit finite difference method (Bathurst et al., 1995). Channel erosion in SHESED includes local bed erosion (bed load plus suspended load) in the channel, sediment inflow from upstream, and sediment flow from overland flow. A one-dimensional transport equation is used. Inputs of the channel component are overland flow and rainfall conditions, supplied by either SHE or taken directly from measurements. Gullying, mass movement, channel bank erosion, or erosion of frozen soil are not considered in the SHESED. It does not feedback to SHE, meaning that change at the channel bed elevation due to erosion is not given as input to SHE, as the change is very small. 2.3. Hillslope scale erosion and sediment transport models Detachment of sediment by rainfall and entrainment of sediment by overland flow on a plane slope without rills were studied by Rose et al. (1983a). Rainfall detachment was taken proportional to a power of rainfall rate, whereas sediment entrainment by overland flow was given by an equation based on the stream power concept for bed-load sediment (Bagnold, 1977). Deposition rate was considered to depend on the settling velocity of sediment particles. The analysis resulted in an ordinary differential equation (ODE) expressing the conservation of mass of sediment. Application of the theory to data from a plane slope in an arid region in Arizona (Rose et al., 1983b) showed good agreement between measured and calculated sediment concentrations and fluxes over time. Hairsine and Rose (1991) developed a formula describing rainfall detachment in the absence of overland flow driven erosion. Rate of rainfall detachment per unit area of soil (E I) was assumed to be dependent upon rainfall rate (r) as EI arb 6

where a is detachability coefficient of the original soil, b an exponent that can be equated to unity (Proffitt et al., 1991). Detachment of the original soil and re-detachment of the deposited sediment were considered separately. Sediment was classified based upon the particle settling velocity. In the study by Hairsine and Rose (1992a), a new model for erosion on a plane surface was developed using physical principles. The model uses the stream power approach (Bagnold, 1966) and considers the raindrop impact and overland flow removal. The model also contains the deposition and re-entrainment of the deposited sediment. By considering the rill formations (Hairsine and Rose, 1992b) the model was further developed. In the stochastic model of Lisle et al. (1998), trajectories of sediment particles were represented by alternating periods of rest (stationary phase) and motion (mobile phase). A sediment particle was supposed to have only two velocity states: resting on the bed (zero velocity), and moving in the water at the water velocity u. Parlange et al. (1999) gave an analytical approximation with similar mathematical structure of studies mentioned above, for rainfall induced erosion on a hillslope. Sediment sorting was investigated by Hairsine et al. (1999), and theoretical results were compared to experimental results provided by Proffitt et al. (1991). A simple experimental set-up

260

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

was developed for rainfall induced erosion on hillslope (Heilig et al., 2001). Another study performed by Siepel et al. (2002) took the effect of vegetation elements on rainfall induced erosion at hillslope scale. Govindaraju and Kavvas (1991) coupled analytical solutions developed by Govindaraju et al. (1990) for overland flow to the erosion model of Foster and Meyer (1972). The overland flow component used the diffusion wave approximation. The overland flow depth was approximated by a sinusoidal expression. Analytical solutions, obtained after coupling, were compared to the experimental results of Kilinc and Richardson (1973) and Singer and Walker (1983). The influence of micro-topography on overland flow and erosion and sediment transport is of great importance. Flow discharge and sediment discharge in rills are greater than those on interrill areas. Knowing the importance of the micro topography Kavvas and Govindaraju (1992) and Govindaraju and Kavvas (1992) modeled the rill structure of a hillslope. A non-dimensional erosion formulation was developed by Govindaraju (1995) who reduced the number of calibration parameters, to be used in the formulation, to only two. The first gave the erodibility characteristics of the soil, whereas the second defined the stage of the erosion process. Soil erosion is sensitive to both the critical shear stress and the soil erodibility. Govindaraju (1998) explained the stochastic structure of these two factors by taking the flow dynamics into account. The critical shear stress, assumed to be exponentially distributed along the slope length, and soil erodibility were treated as homogeneous uncorrelated random variables. Laguna and Giraldez (1993) aimed to explore the fit of kinematic wave simplification to the sediment transport processes. A sensitivity analysis was performed. The analysis showed the sensitivity to interrill erosion parameters to be high at the rising stage during which erosion processes are controlled by the rainfall impact detachment. Erosion was, however, found to be sensitive to the rill erosion component of the model at the recession stage during which soil loss is primarily due to rill erosion. This means, as stated before, that erosion on the interrill area controls the rising stage of erosion while the rill erosion controls the recession stage. From the study, it was also seen that there was no clear relationship between sediment yield and peak runoff rate. However, important relations between sediment yield and runoff volume and rainfall rate were found. The kinematic wave approximation was found to be an applicable approach for modelling sediment transport. Tayfur (2001) used the two-dimensional flow and sediment continuity equations with the kinematic wave approximation. The erosion term in the sediment mass conservation equation was considered as the sum of raindrop induced (rainfall) erosion and sheet flow generated (runoff) erosion. Soil detachment due to raindrop was related to the rainfall intensity and overland flow depth. Erosion by overland flow is a linear function of the difference between the transport capacity of the flow and the sediment load in the flow. Based upon the analysis in the study, the most sensitive parameters were found to be the soil erodibility coefficient (g) and the exponent (k) in T c g s s c k 7

where Tc is transport capacity of flow, s and s c flow shear stress and its critical value, respectively. A recent study was performed by Aksoy and Kavvas (2001) for erosion and

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

261

sediment transport at hillslope scale. Rill and interrill interaction over the hillslope, which was not taken into account in many of the existing models, was considered. Formulation for the interrill area used the two-dimensional continuity equation plus momentum equation simplified with kinematic wave approximation. Erosion was divided into rainfall erosion for which a simple formula depending upon the rainfall intensity was used, and runoff erosion for which the transport capacity was determined by Yalin (1963) equation. Non-physical erosion parameters included in the rainfall and runoff erosion equations were calibrated by using a field experimental data set. Aksoy and Kavvas (2001), in order to simplify the modelling technique, reduced the two-dimensional partial differential equation (PDE) governing the erosion and sediment transport process over interrill area into a one-dimensional form. The original PDE was even reduced to an ODE at hillslope scale by means of areal averaging of the original conservation equation. 2.4. Transport capacity of overland flow Sediment transport capacity of overland flow is the maximum flux of sediment that flow is capable to transport. All physically based soil erosion models contain a sediment transport equation. Many of the existing models use either a bed load or a total load formula originally developed for rivers. Other soil erosion models use simple empirical formulas. Sediment transport capacity can be formulated by either sediment concentration or sediment load. Concentration is a more fundamental variable than the sediment load. Early approaches to the sediment transport capacity have used the shear stress (Yalin, 1963), stream power (Bagnold, 1966), or unit stream power (Yang, 1972). Alonso et al. (1981), after comparison of nine sediment transport formulas, suggested the use of Yalins (1963) equation in computing the sediment transport capacity for overland flow. Nearing et al. (1989) used a simplified function of the hydraulic shear stress acting on the soil for calculating the sediment transport capacity of flow. Tayfur (2002) analysed those approaches and concluded that the unit stream power could be selected for the simulation of unsteady state erosion and sediment transport from very mild bare slopes and, under low rainfall intensities, it could also be employed to simulate loads from mild and steep slopes. For the very steep slopes, the shear stress and stream power models could be used. The stream power and the shear stress models could also be employed in order to simulate sediment load from mild and steep slopes under high rainfall intensity. Based upon data of nearly 10 years in a semi-arid area of New Mexico, USA, Emmett (1970) concluded that sediment and organic content in overland flow was positively correlated with ground slope and negatively correlated with vegetation. A general relationship between variables that affect the sediment transport capacity was developed by Julien and Simons (1985) as  sc e qs aS b qc rd 1 s 8

where q s is sediment discharge, S slope, q discharge, r rainfall intensity, s c critical shear stress, s actual shear stress, a a coefficient and b, c, d, e exponents to be determined from laboratory or field experiments. When s c remains very small compared to s and when it is

262

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

considered that the sediment transport capacity of turbulent flow in deep channels is not a function of rainfall then Eq. (8) reduces to qs aS b qc : 9

Prosser and Rustomji (2000) addressed the same equation for the sediment transport capacity. As q is a function of the upslope contributing area, sediment discharge is evaluated completely by topographic factors. From examination of many studies based upon flumes, laboratory and field plots and rivers, b and c, exponents of S and q in Eq. (9), were found to be bounded by 0.5 and 2.0, as lower and upper limits, respectively. When one single combination is desired, a median value of 1.4 can be used for both exponents. The sediment transport capacity (Tc) of overland flow was also found to be proportional to the overland flow discharge ( q) only, as Tc ~ q c , where c ranged between 1.2 and 1.5. Then the sediment concentration (C s) in the runoff becomes C s ~ q c 1 (Novotny and Chesters, 1989). Abrahams et al. (1998) obtained a regression equation for the transport capacity of overland flow by combining results of laboratory experiments. 2.5. Data requirement The data requirements of any model dramatically increase with the complexity included in the model. Distributed models, in particular, need more data than other models. Erosion and sediment transport models contain non-physical parameters in formulating the rainfall and runoff erosion. This requires data for calibrating and then validating those parameters. However, it is known that it is a difficult task to collect erosion and sediment data from a watershed or from a specific hillslope in a watershed. Data collection is much harder for detailed models. For example, collecting data for a model with no rill and interrill area distinction will be much easier than doing it for a model with this distinction. Input data for erosion and sediment transport models include outputs from hydrological models. Therefore, in order to be able to run any erosion model it is first required to run a hydrological model so that the hydrological outputs can be supplied as input for the erosion model. An erosion modeller should either run the hydrological part of the model or the modeller should be supplied with the hydrological inputs. GIS has been a very important tool in developing data files required for the models. It helps modellers to use more complicated models, as preparing data with GIS is easier than doing it by traditional ways. By using GIS, a parameter can be distributed not only in time but also in space.

3. Extension of a sediment transport model to a sediment-bound pollutant transport model Pollutant or nutrient yield can be simply calculated by multiplying the sediment yield by a potency factor, which is pollutant content of the sediment. This content is usually given in grams of pollutant in grams of soil. Non-point pollution is caused by humankinds activities on the land and differs from the natural erosion and sediment movement. For example; erosion and sediment transport caused by cutting a forest down is considered

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

263

pollution, while a mudslide, caused by an earthquake, is not. Sediment concentrations two orders of magnitude lower than the natural erosion are not tolerable if they are caused by non-point pollution. Use of lumped models is avoided in water quality studies as the delivery process and related parameters represent a hydrologic stochastic process. It is therefore suggested to take the stochastic structure of the nonpoint pollution processes into account and also to establish the statistical characteristics of the processes (Novotny and Chesters, 1989). Sediment yield of a stream is strongly related to the flow. Flow is monitored in streams more frequently than the sediment concentration or phosphorus loads. Therefore, relations between flow and sediment or phosphorus are usually based on some regression equations. For example, sedimentturbidity relationship is used to convert a time-series of turbidity to suspended sediment concentration. If turbidity is missing for a period but flow has been measured at that period, the suspended sedimentflow relationship can be used to fill the gaps in the data (Green et al., 1999). Phosphorus transported by the flow is much more than that associated with the soil since phosphorus is mainly associated with finer particles (Quinton, 1999). It is known that phosphorus mainly moves with sediment by being attached to the surface of sediment particles. Therefore, it is reasonable to assume the sediment transport process as an indicator of phosphorus transport. Chemical properties are other factors that should be taken into account in the soil detachment processes, yet none of the existing models do so. Akan (1987) studied pollutant washoff by overland flow on impervious surfaces. Ashraf and Borah (1992) worked on the modeling of pollutant transport in runoff and sediment. Yan and Kahawita (1997, 2000) and Wallach et al. (2001) studied modeling pollutants in the overland flow at the hillslope scale. A model called SPNM (SedimentPhosphorusNitrogen Model) was developed by Williams (1980) for simulating contribution of agriculture to water pollution. SPNM was designed to predict sediment, P, and N yields for individual storms and to route these yields through streams. The model computes the total sediment yield predicted by the Modified Universal Soil Loss Equation (MUSLE). The P model predicts average annual P yields. The N model simulates both organic and inorganic N yields associated with the sediment and runoff. The organic N model has the same structure as the P model because both N and P are transported with sediment. The organic N tends to associate with fine clay, whereas P tends to associate with coarse clay and silt as well as fine clay. The nitrate concentration in surface and subsurface flow are modeled separately. SPNM gave good results for sediment yield. Results for nutrients were found realistic. AGNPS (Young et al., 1989) has a subcomponent for estimating P, N and COD (chemical oxygen demand). Chemical transport calculations are divided into soluble and sediment adsorbed phases. Nutrient yield in the sediment-adsorbed phase is obtained by multiplying the total sediment yield in a cell by the nutrient content in the field soil and the enrichment ratio, which is a function of sediment yield. Soluble nutrient yield is estimated by multiplying total runoff by the mean concentration of the nutrient at the soil surface during runoff and an extraction coefficient of nutrient for movement into runoff. SHETRAN (Ewen et al., 2000) is a reactive solute transport model. Three main components in SHETRAN are water flow, sediment transport and solute transport. Flow is assumed not to be affected by sediment transport and sediment transport not to be affected

264

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

by solute transport. Therefore, the three components are independent of each other. SHETRAN models a single complete river basin. It has a stream link and column structure. River network is modeled as stream links and the rest of the basin is modeled as a set of columns. Transport along the links and vertical transport in the columns are the two main movements. There is also lateral movement between cells in neighboring columns. Later, Birkinshaw and Ewen (2000) developed a nitrogen transformation component and integrated it into the SHETRAN.

4. Summary of review and projections for the future Erosion is a very important natural phenomenon ending with soil loss. It causes also loss of storage volume in river reservoirs where eroded sediment deposits. The USLE was designed as a tool to be used in the management practices of agricultural lands. It is the first attempt in computing the sediment yield of a catchment. Although its development is based on data from the United States, it has been used widely all over the world. The USLE with some modifications and revisions is still a useful tool in watershed management. A large number of the existing erosion and sediment transport models are based on the USLE. Their applications are, however, limited to the environmental circumstances from which the USLE was generated. Limitations of the USLE and its modified or revised versions (MUSLE and RUSLE) forced modelers to use distinctly different alternatives. WEPP in the United States and SHESED and EUROSEM in Europe were derived based on physical description of the erosion and sediment transport processes. Although preparation of data for physically based models is a hard task, they have been used extensively. It is obvious that a physically based model has much more detail than USLE or its derivatives have. Therefore, there has been a big effort in developing physically based erosion and sediment transport models. A physically based model may use lumped or distributed inputs to generate lumped or distributed outputs. A distributed model is constructed by using partial differential equations, whereas a lumped model is expressed by ordinary differential equations (Singh, 1995). A physically based model may be a semi-distributed model as well. This means that not all model parameters need to be of distributed type. Some parameters, especially those that cannot be collected easily in the field, can be used as lumped. It may also be noted that some physically based models may include non-physical descriptions in their formulation. For example, LISEM contains many empirically derived equations although it is presented as a physically based model. A differential equation, used for constructing a physically based model, may be deterministic or stochastic. All the existing models are deterministic models where the erosion and sediment transport processes are formulated by deterministic differential equations. None of the models can yet consider the stochasticity included in the erosion and sediment transport processes. Thus, models in Table 3 are categorised as distributed type deterministic physically based erosion and sediment transport models. Physically based models use different approximations by which they simplify the system (nature) in formulating the erosion and sediment transport processes. One

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

265

simplification used is to reduce the number of dimensions of the governing equations. For example, the three dimensional topographical terrain is reduced into a two-dimensional form. SHESED is one of those models that use erosion and sediment transport in the twodimensional form. Increasing number of dimensions results in more intensive computations by a model. A hillslope can be thought of as a sheet where processes take place in one dimension. Even in two-dimensional models, the number of dimensions may be reduced to one by performing local scale averaging. However, in such a simplification the effect of the second dimension is not neglected but indirectly incorporated into the model. The partial differential equation governing the process may even be reduced to an ordinary partial differential equation (Aksoy and Kavvas, 2001). Some physically based models do not consider the time derivative term. WEPP is such a model. Also Foster and Meyer (1972) used the steady state continuity equation of mass transport, which is the basis for the ANSWERS model. The case for most of the existing physically based erosion and sediment transport models is the unsteady state where the time derivative of sediment concentration is taken into consideration. Initial and boundary conditions become very important in cases where the model simulates erosion and sediment transport continuously. Continuous simulation models require large quantities of data for weather and land use. They generate a large number of small events that may not cause significant runoff or soil loss. Some physically based models were, therefore, designed as event based models that can be run for each specific event. This indicates that erosion is dominated by only a few events per year. EUROSEM is such a model. Only SEM, SHESED and WEPP can simulate the erosion and sediment transport continuously. Smoothing irregularities (rills and interrill areas) over a hillslope is another simplification although it has been shown experimentally by Govindaraju et al. (1992) that, erosion in rills is, at least, one order of magnitude greater than erosion on interrill areas. Therefore, modellers should be aware of the rillinterrill interaction on a hillslope although it is not easy to incorporate it into a model. Some models are capable of distinguishing among the sediment sizes. Modellers aimed to construct models that are less complex than the physically based models but that yield simulations more precise than those obtained by the USLE or its derivatives. This directed modellers to build conceptual models where the erosion process is conceptualised. In such a model there is a non-physical but conceptually meaningful relation between the elements of the process. Geographical Information Systems (GIS) have been a very useful tool for hydrologists, in particular for physically based modellers in providing the spatially distributed data. GIS can also supply the time distribution of the hydrological data. GIS uses the digital elevation model (DEM) that can provide information on elevation, slope and aspect of the catchment. By using the 10-m DEM, one can delineate the rill and gully structure of the watershed. GIS seems, at least for now, as the only way for supplying the necessary data for the physically based models. It is possible to incorporate the physical heterogeneity in a catchment by using GIS. Heterogeneity in hydrological variables can, however, be obtained by performing hydrological data measurements not only at the outlet of a catchment but also at least at the outlet of each subcatchment. The measurements help in assessing the hydrological behaviour of each subcatchment.

266

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

Erosion and sediment transport models are extensions of hydrological models. Therefore, erosion and sediment transport equations are coupled to existing hydrological algorithms. In such a coupling, output of the hydrological model becomes input for the erosion part of the model. In the same sense, an erosion and sediment transport model can be extended easily to a nutrient transport model, as it is known that nutrients are mainly transported by sediment particles. It is much easier to extend a multi-size erosion and sediment transport model to a nutrient transport model since nutrient transport is a size selective process. Current physically based models are deterministic where rainfallrunoff, erosion and sediment transport are thought of as deterministic processes. Probability based stochastic modelling techniques can be derived in the future for the erosion and sediment transport modelling. Such a technique should include the probability distribution of rainfall. Spatial and temporal distribution of rainfall as a random input to the rainfallrunoff part of the model will result in randomly simulated runoff. In such a modelling technique heterogeneity in the physical structure of watershed (Kavvas, 1999) can be given by probability distribution functions. Rill occurrence probability over an interrill area (Govindaraju and Kavvas, 1992) is an example of this. Also non-physical erosion parameters have probability distributions (Govindaraju, 1998). This is because of the critical soil properties, such as aggregation or soil resistance to erosion, which are random processes due to heterogeneity of soils.

Acknowledgements The first author (H. Aksoy) was a post-doctoral researcher with the second author (M.L. Kavvas) when this study was performed at the Hydrologic Research Laboratory, Department of Civil and Environmental Engineering of University of California at Davis (UCDavis). This stay was supported by Istanbul Technical University (ITU) through a postdoctoral research scholarship, by Scientific and Technical Research Council of Turkey (TUBITAK) through NATO B-1 postdoctoral scholarship, and by UCDavis through CA State EPA 205J Grant. The second authors work was supported partially by US EPA Center for Ecological Health Research (Grant No. R819658) at University of California, Davis, and partially by US EPA Grant No. R826282010. However, the views expressed in this paper do not necessarily reflect those of the funding agencies, and no official endorsement should be inferred. The authors also would like to acknowledge constructive comments by two anonymous reviewers, guest editors (Dr. W. Cornelis and Dr. D. Gabriels of Ghent University, Belgium), and editor-in-chief at CATENA at different stages of this manuscript.

References
Abbott, M.B., Bathurst, J.C., Cunge, J.A., OConnell, P.E., Rasmussen, J., 1986a. An introduction to the European hydrological systemsysteme hydrologique Europeen, bSHEQ: 1. History and philosophy of a physically-based, distributed modeling system. Journal of Hydrology 87, 45 59.

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

267

Abbott, M.B., Bathurst, J.C., Cunge, J.A., OConnell, P.E., Rasmussen, J., 1986b. An introduction to the European hydrological systemsysteme hydrologique Europeen, bSHEQ: 2. Structure of a physically-based, distributed modeling system. Journal of Hydrology 87, 61 77. Abrahams, A.D., Li, G., Krishnan, C., Atkinson, J.F., 1998. Predicting sediment transport by interrill overland flow on rough surfaces. Earth Surface Processes and Landforms 23, 1087 1099. Ackers, P., White, W.R., 1973. Sediment transport: new approach and analysis. ASCE Journal of the Hydraulics Division 99 (HY11), 2041 2060. Akan, A.O., 1987. Pollutant washoff by overland flow. ASCE, Journal Of Environmental Engineering 113 (4), 811 823. Aksoy, H., Kavvas, M.L., 2001. A physically based erosion and sediment transport component for watershed hydrologic models. Hydrologic Research Laboratory, Department of Civil and Environmental Engineering, University of California, Davis, California. Alonso, C.V., Neibling, W.H., Foster, G.R., 1981. Estimating sediment transport capacity in watershed modelling. Transactions of the ASAE 12111220, 1226. ASCE Task Committee, 1970. Sediment sources and sediment yields. ASCE, Journal of the Hydraulics Division 96 (HY6), 1283 1329. Ashraf, M.S., Borah, D.K., 1992. Modeling pollutant transport in runoff and sediment. Transactions of the ASAE 35 (6), 1789 1797. Bagnold, R.A., 1966. An approach to the sediment transport problem for general physics. Geological Survey Professional Paper (U.S.) 442-I. Bagnold, R.A., 1977. Bed load transport by natural rivers. Water Resources Research 13, 303 311. Bathurst, J.C., Wicks, J.M., OConnell, P.E., 1995. The SHE/SHESED basin scale water flow and sediment transport modeling system. In: Singh, V.P. (Ed.), Computer Models of Watershed Hydrology. Water Resources Publications, Littleton, CO, pp. 563 594. Beasley, D.B., Huggins, L.F., Monke, E.J., 1980. ANSWERS: a model for watershed planning. Transactions of the ASAE, 938 944. Bennett, J.P., 1974. Concepts of mathematical modeling of sediment yield. Water Resources Research 10 (3), 485 492. Beven, K.J., Kirkby, M.J., 1979. A physically-based variable contributing area model of basin hydrology. Hydrological Sciences Bulletin 24 (1), 43 69. Biesemans, J., Van Meirvenne, M., Gabriels, D., 2000. Extending the RUSLE with the Monte Carlo err or propagation technique to predict long term average off-site sediment accumulation. Journal of Soil and Water Conservation 55 (1), 35 42. Birkinshaw, S.J., Ewen, J., 2000. Nitrogen transformation component for SHETRAN catchment nitrate transport modelling. Journal of Hydrology 230, 1 17. Borah, D.K., 1989. Sediment discharge model for small watersheds. Transactions of the ASAE 32 (3), 874 880. Bryan, R.B., 2000. Soil erodibility and processes of water erosion on hillslope. Geomorphology 32, 385 415. Croley II, T.E., 1982. Unsteady overland sedimentation. Journal of Hydrology 56, 325 346. De Roo, A.P.J., 1998. Modelling runoff and sediment transport in catchments using GIS. Hydrological Processes 12, 905 922. De Roo, A.P.J., Wesseling, C.G., Ritsema, C.J., 1996. LISEM: a single-event physically based hydrological and soil erosion model for drainage basins: I. Theory, input and output. Hydrological Processes 10, 1107 1117. Emmett, W.W., 1970. The hydraulics of overland flow on hillslopes. Prof. Pap.-Geol. Surv. (U. S.), vol. 662A. USGS (United States Geological Survey), Washington, DC. 68 pp. Engelund, F., Hansen, E., 1967. A Monograph on Sediment Transport in Alluvial Streams. Teknish Vorlag, Copenhagen. Ewen, J., Parkin, G., OConnell, P.E., 2000. SHETRAN: distributed river basin flow and transport modeling system. ASCE, Journal of Hydrologic Engineering 5 (3), 250 258. Ferro, V., Porto, P., 2000. Sediment delivery distributed (SEDD) model. ASCE, Journal of Hydraulic Engineering 5 (4), 411 422. Folly, A., Quinton, J.N., Smith, R.E., 1999. Evaluation of the EUROSEM model using data from the Catsop watershed, the Netherlands. Catena 37, 457 475.

268

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

Foster, G.R., 1982. Modeling the erosion process. In: Haan, C.T., Johnson, H.P., Brakensiek, D.L. (Eds.), Hydrologic Modeling of Small Watersheds, ASAE Monograph, vol. 5. ASAE (American Society of Agricultural Engineers), St. Joseph, MI, pp. 297 380. Foster, G.R., Meyer, L.D., 1972. A closed-form soil erosion equation for upland areas. In: Shen, H.W. (Ed.), Sedimentation Symposium in Honor Prof. H.A. Einstein. Colorado State University, Fort Collins, CO, pp. 12.1 12.19. Foster, G.R., Meyer, L.D., 1975. Mathematical simulation of upland erosion by fundamental erosion mechanics. Present and Prospective Technology for Predicting Sediment Yields and Sources. USDA Agricultural Service, Washington, DC, pp. 190 207. Foster, G.R., Lane, L.J., Nowlin, J.D., Laflen, J.M., Young, R.A., 1981. Estimating erosion and sediment yield on field-sized areas. Transactions of the ASAE, pp. 1253 1262. Govindaraju, R.S., 1995. Non-dimensional analysis of a physically based rainfallrunoff-erosion model over steep slopes. Journal of Hydrology 173, 327 341. Govindaraju, R.S., 1998. Effective erosion parameters for slopes with spatially varying properties. ASCE, Journal of Irrigation and Drainage Engineering 124 (2), 81 88. Govindaraju, R.S., Kavvas, M.L., 1991. Modeling the erosion process over steep slopes: approximate analytical solutions. Journal of Hydrology 127, 279 305. Govindaraju, R.S., Kavvas, M.L., 1992. Characterization of the rill geometry over straight hillslopes through spatial scales. Journal of Hydrology 130, 339 365. Govindaraju, R.S., Kavvas, M.L., Jones, S.E., 1990. Approximate analytical solutions for overland flows. Water Resources Research 26 (12), 2903 2912. Govindaraju, R.S., Kavvas, M.L., Tayfur, G., Krone, R., 1992. Erosion control of decomposed granite at Buckhorn Summit. Final Project Report. University of California, Davis, CA. Green, T.R., Beavis, S.G., Dietrich, C.R., Jakeman, A.J., 1999. Relating stream bank erosion to in-stream transport of suspended sediment. Hydrological Processes 13, 777 787. Hairsine, P.B., Rose, C.W., 1991. Rainfall detachment and deposition: sediment transport in the absence of flowdriven processes. Soil Science Society of America Journal 55, 320 324. Hairsine, P.B., Rose, C.W., 1992a. Modeling water erosion due to overland flow using physical principles: 1. Sheet flow. Water Resources Research 28 (1), 237 243. Hairsine, P.B., Rose, C.W., 1992b. Modeling water erosion due to overland flow using physical principles: 2. Rill flow. Water Resources Research 28 (1), 245 250. Hairsine, P.B., Sander, G.C., Rose, C.W., Parlange, J.-Y., Hogarth, W.L., Lisle, I., Rouhipour, H., 1999. Unsteady soil erosion due to rainfall impact: a model of sediment sorting on the hillslope. Journal of Hydrology 220, 115 128. Heilig, A., DeBruyn, D., Walter, M.T., Rose, C.W., Parlange, J.-Y., Steenhuis, T.S., Sander, G.C., Hairsine, P.B., Hogarth, W.L., Walker, L.P., 2001. Testing a mechanistic soil erosion model with simple experiment. Journal of Hydrology 244, 9 16. Johanson, R.C., Leytham, K.M., 1977. Modeling sediment transport in natural channels. In: Corell, D.L. (Ed.), Watershed Research in Eastern North America: A Workshop to Compare Results. Chesapeake Bay Center for Environmental Studies, Smithsonian Institution, Edgewater, MD. Johnson, B.E., Julien, P.Y., Molnar, D.K., Watson, C.C., 2000. The two-dimensional upland erosion model CASC2D-SED. Journal of the American Water Resources Association 36 (1), 31 42. Julien, P.Y., Simons, D.B., 1985. Sediment transport capacity of overland flow. Transactions of the ASAE 28 (3), 755 762. Kavvas, M.L., 1999. On the coarse-graining of hydrologic processes with increasing scales. Journal of Hydrology 217, 191 202. Kavvas, M.L., Govindaraju, R.S., 1992. Hydrodynamic averaging of overland flow and soil erosion over rilled hillslopes. IAHS Publications 209, 101 111. Kilinc, M., Richardson, E.V., 1973. Mechanics of soil erosion from overland flow generated by simulated rainfall. Hydrology Papers, vol. 63. Colorado State University, Fort Collins, CO. Kinnell, P.I.A., 1999. Discussion on bThe European soil erosion model (EUROSEM): a dynamic approach for predicting sediment transport from fields and small catchmentsQ. Earth Surface Processes and Landforms 24, 563 565.

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

269

Kuznetsov, M.S., Gendugov, V.M., Khalilov, M.S., Ivanuta, A.A., 1998. An equation of soil detachment by flow. Soil & Tillage Research 46, 97 102. Laflen, J.M., Elliot, W.J., Flanagan, D.C., Meyer, C.R., Nearing, M.A., 1997. WEPP-predicting water erosion using a process-based model. Journal of Soil and Water Conservation, 96 102 (MarchApril). Laguna, A., Giraldez, J.V., 1993. The description of soil erosion through kinematic wave model. Journal of Hydrology 145, 65 82. Lau, Y.L., Engel, P., 1999. Inception of sediment transport on steep slopes. ASCE, Journal of Hydraulic Engineering 125 (5), 544 547. Lisle, I.G., Rose, C.W., Hogarth, W.L., Hairsine, P.B., Sander, G.C., Parlange, J.Y., 1998. Stochastic sediment transport in soil erosion. Journal of Hydrology 204, 217 230. Lopes, V.L., 1987. A numerical model of watershed erosion and sediment yield. PhD thesis, The University of Arizona. Lopes, V.L., Lane, L.J., 1988. Modeling sedimentation processes in small watersheds. IAHS Publications 174, 497 508. Merritt, W.S., Latcher, R.A., Jakeman, A.J., 2003. A review of erosion and sediment transport models. Environmental Modelling & Software 18, 761 799. Montgomery, D.R., Dietrich, W.E., 1994. A physically based model for the topographic control on shallow landsliding. Water Resources Research 30 (4), 1153 1171. Morgan, R.P.C., Quinton, J.N., Smith, R.E., Govers, G., Poesen, J.W.A., Auerswald, K., Chisci, G., Torri, D., Styczen, M.E., 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. Morgan, R.P.C., Quinton, J.N., Smith, R.E., Govers, G., Poesen, J.W.A., Auerswald, K., Chisci, G., Torri, D., Styczen, M.E., 1999. Reply to discussion on bThe European soil erosion model (EUROSEM): a dynamic approach for predicting sediment transport from fields and small catchments. Earth Surface Processes and Landforms 24, 567 568. Nearing, M.A., Foster, G.R., Lane, L.J., Finkner, S.C., 1989. A process-based soil erosion model for USDA-water erosion prediction project technology. Transactions of the ASAE 32 (5), 1587 1593. Norman, S.E., 1989. An evaluation of ANSWERS, a distributed parameter watershed model. Thesis submitted in partial satisfaction of the requirements for the degree of master of science in Water Science in the Graduate Division of the University of California, Davis, California. Novotny, V., Chesters, G., 1989. Delivery of sediment and pollutants from nonpoint sources: a water quality perspective. Journal of Soil and Water Conservation, 568 576 (NovemberDecember). Park, S.W., Mitchell, J.K., Scarborough, J.N., 1982. Soil erosion simulation on small watersheds: a modified ANSWERS model. Transactions of the ASAE, 1581 1588. Parlange, J.-Y., Hogarth, W.L., Rose, C.W., Sander, G.C., Hairsine, P., Lisle, I., 1999. Addendum to unsteady soil erosion model. Journal of Hydrology 217, 149 156. Pekarova, P., Konicek, A., Miklanek, P., 1999. Testing of AGNPS model application in Slovak microbasins. Physics and Chemistry of the Earth (B) 24 (4), 303 305. Proffitt, A.P.B., Rose, C.W., Hairsine, P.B., 1991. Rainfall detachment and deposition: experiments with low slopes and significant water depths. Soil Science Society of America Journal 55, 325 332. Prosser, I.P., Rustomji, P., 2000. Sediment transport capacity relations for overland flow. Progress in Physical Geography 24 (2), 179 193. Quinton, J.N., 1999. Detachment and transport of particle-bound P: process and prospects for modelling. Paper Presented at the COST Action on Phosphorus Meeting, Cordoba, Spain, 1315 May 1999. Renard, K.G., Foster, G.R., Weesies, G.A., Porter, J.P., 1991. RUSLE: revised universal soil loss equation. Journal of Soil and Water Conservation, 30 33 (JanuaryFebruary). Renard, K.G., Foster, G.R., Yoder, D.C., McCool, D.K., 1994. RUSLE revisited: status, questions, answers, and the future. Journal of Soil and Water Conservation, 213 220 (MayJune). Rode, M., Fredo, H.G., 1999. Testing AGNPS for soil erosion and water quality modelling in agricultural catchments in Hesse (Germany). Physics and Chemistry of the Earth (B) 24 (4), 297 301. Rose, C.W., Williams, J.R., Sander, G.C., Barry, D.A., 1983a. A mathematical model of soil erosion and deposition processes: I. Theory for a plane land element. Soil Sci. Soc.Am. J. 47, 991 995.

270

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

Rose, C.W., Williams, J.R., Sander, G.C., Barry, D.A., 1983b. A mathematical model of soil erosion and deposition processes: II. Application to data from an arid-zone catchment. Soil Science Society of America Journal 47, 996 1000. Rovey, E.W., Woolhiser, D.A., Smith, R.E., 1977. A distributed kinematic model of upland watersheds. Hydrology Papers, vol. 93. Colorado State University, Fort Collins, CO. Santos, C.A.G., Watanabe, M., Suzuki, K., Srinivasan, V.S., 1998. Influence of the moisture-tension parameter on sedigraphs and hydrographs from a semiarid region in Brazil. IAHS Publications 249, 231 240. Santos, C.A.G., Watanabe, M., Suzuki, K., 2000. Application of a physically-based erosion model for a largeriver basin in Japan. Paper Presented in the Symposium on Integrated Water Resources Management, Davis, CA, April 2000. Schmidt, J., Werner, M.V., Michael, A., 1999. Application of the EROSION 3D model to the CATSOP watershed, The Netherlands. Catena 37, 449 456. Siepel, A.C., Steenhuis, T.S., Rose, C.W., Rose, J.-Y., McIsaac, G.F., 2002. A simplified hillslope erosion model with vegetation elements for practical applications. Journal of Hydrology 258, 111 121. Singer, M.J., Walker, P.H., 1983. Rainfall-runoff in soil erosion with simulated rainfall, overland flow and cover. Australian Journal of Soil Research 21, 109 122. Singh, V.P., 1995. Watershed modelling. In: Singh, V.P. (Ed.), Computer Models of Watershed Hydrology. Water Resources Publ., Highlands Ranch, CO, pp. 1 22. Singh, V.P., Prasad, S.N., 1982. Explicit solution to kinematic equations for erosion on an infiltrating plane. In: Singh, V.P. (Ed.), Modeling Components of Hydrologic Cycle. Water Resources Research, Littleton, CO, pp. 515 538. Singh, V.P., Woolhiser, D.A., 2002. Mathematical modeling of watershed hydrology. ASCE, Journal of Hydrologic Engineering 7 (4), 270 292. Smith, R.E., 1981. A kinematic model for surface mine sediment yield. Transactions of the ASAE, 1508 1514. Smith, R.E., Goodrich, D.C., Quinton, J.N., 1995a. Dynamic, distributed simulation of watershed erosion: the KINEROS2 and EUROSEM models. Journal of Soil and Water Conservation 50 (5), 517 520. Smith, R.E., Goodrich, D.C., Woolhiser, D.A., Unkrich, C.L., 1995b. KINEROSa kinematic runoff and erosion model. In: Singh, V.P. (Ed.), Computer Models of Watershed Hydrology. Water Resources Publications, Littleton, CO, pp. 697 732. Smith, R.E., Goodrich, D.C., Unkrich, C.L., 1999. Simulation of selected events on the Catsop catchment by KINEROS2, a report for the GCTE conference on catchment scale erosion models. Catena 37, 457 475. Storm, B., Jorgensen, G.H., Styczen, M., 1987. Simulation of water flow and soil erosion processes with a distributed physically-based modeling system. IAHS Publications 167, 595 608. Tayfur, G., 2001. Modeling two-dimensional erosion process over infiltrating surfaces. ASCE, Journal of Hydrologic Engineering 6 (3), 259 262. Tayfur, G., 2002. Applicability of sediment transport capacity models for nonsteady state erosion from steep slopes. ASCE, Journal of Hydrologic Engineering 7 (3), 252 259. Veihe, A., Quinton, J., 2000. Sensitivity analysis of EUROSEM using Monte Carlo simulation I: hydrological, soil and vegetation parameters. Hydrological Processes 14, 915 926. Veihe, A., Quinton, J., Poesen, J., 2000. Sensitivity analysis of EUROSEM using Monte Carlo simulation II: the effect of rills and rock fragments. Hydrological Processes 14, 927 939. Viney, N.R., Sivapalan, M., 1999. A conceptual model of sediment transport: application to the Avon River Basin in Western Australia. Hydrological Processes 13, 727 743. Viney, N.R., Sivapalan, M., Deeley, D., 2000. A conceptual model of nutrient mobilisation and transport applicable at large catchment scales. Journal of Hydrology 240, 23 44. Wallach, R., Grigorin, G., Rivlin (Byk), J., 2001. A comprehensive mathematical model for transport of soildissolved chemicals by overland flow. Journal of Hydrology 247, 85 99. Wicks, J.M., 1988. Physically-based mathematical modelling of catchment sediment yield. Thesis submitted for the degree of doctor of philosophy, Department of Civil Engineering, University of Newcastle Upon Tyne. Wicks, J.M., Bathurst, J.C., 1996. SHESED: a physically based, distributed erosion and sediment yield component for the SHE hydrological modeling system. Journal of Hydrology 175, 213 238. Wicks, J.M., Bathurst, J.C., Johnson, C.W., 1992. Calibrating SHE soil-erosion model for different land covers. ASCE, Journal of Irrigation and Drainage Engineering 118 (5), 708 723.

H. Aksoy, M.L. Kavvas / Catena 64 (2005) 247271

271

Williams, J.R., 1980. SPNM, a model for predicting sediment, phosphorus, and nitrogen yields from agricultural basins. AWRA, Water Resources Bulletin 16 (5), 843 848. Wischmeier, H., Smith, D.D., 1978. Predicting rainfall erosion losses. Agriculture Handbook no 537, USDA Science and Education Administration. Woolhiser, D.A., Smith, R.E., Goodrich, D.C., 1990. KINEROS, a kinematic runoff and erosion model. Documentation and User Manual, USDA, Agricultural Research Service, ARS-77. 130 pp. Yalin, M.S., 1963. An expression for bed-load transportation. ASCE, Journal of the Hydraulics Division 89 (HY3), 221 250. Yan, M., Kahawita, R., 1997. Modeling pollutant transport in overland flow with infiltration. Proceedings of the 27th Congress of the IAHR, Volume 1, August 1015, San Francisco, CA, pp. 347 351. Yan, M., Kahawita, R., 2000. Modelling the fate of pollutant in overland flow. Water Research 34 (13), 3335 3344. Yang, C.T., 1972. Unit stream power and sediment transport. ASCE Journal of the Hydraulics Division 98 (HY10), 1805 1826. Young, R.A., Onstad, C.A., Bosch, D.D., Anderson, W.P., 1989. AGNPS: a nonpoint-source pollution model for evaluating agricultural watersheds. Journal of Soil and Water Conservation, 168 173 (MarchApril). Zhang, L., ONeill, A.L., Lacey, S., 1996a. Modelling approaches to the prediction of soil erosion in catchments. Environmental Software 11 (13), 123 133. Zhang, X.C., Nearing, M.A., Risse, L.M., McGregor, K.C., 1996b. Evaluation of WEPP runoff and soil loss predictions using natural runoff plot data. Transactions of the ASAE 39 (3), 855 863. Ziegler, A.D., Giambelluca, T.W., Sutherland, R.A., 2001. Erosion prediction on unpaved mountain roads in northern Thailand: validation of dynamic erodibility modeling using KINEROS2. Hydrological Processes 15, 337 358.

Você também pode gostar