You are on page 1of 11

EARTH SURFACE PROCESSES AND LANDFORMS Earth Surf. Process.

Landforms 36, 805815 (2011) Copyright 2010 John Wiley & Sons, Ltd. Published online 7 December 2010 in Wiley Online Library (wileyonlinelibrary.com) DOI: 10.1002/esp.2108

Back-calculation of bedload transport in steep channels with a numerical model


1

Michael Chiari1* and Dieter Rickenmann1,2 University of Natural Resources and Life Sciences, Institute of Mountain Risk Engineering, Vienna, Austria 2 Swiss Federal Research Institute WSL, Birmensdorf, Switzerland
Received 5 May 2010; Revised 11 October 2010; Accepted 18 October 2010 *Correspondence to: Michael Chiari, University of Natural Resources and Life Sciences, Institute of Mountain Risk Engineering, Peter Jordanstrasse 82, 1190 Vienna, Austria. E-mail: michael.chiari@boku.ac.at

ABSTRACT: In August 2005 severe ood events occurred in the Alps. A sediment routing model for steep torrent channel networks called SETRAC has been applied to six well-documented case study streams with substantial sediment transport in Austria and Switzerland. For these streams information on the sediment budget along the main channel is available. Flood hydrographs were reconstructed based on precipitation data and stream gauges in neighbouring catchments. Different scenarios are modelled and discussed regarding sediment availability and the effect of armouring and macro-roughness on sediment transport calculations. The simulation results show the importance of considering increased ow resistance for small relative ow depth when modelling bedload transport during high-intensity ood events in torrents and mountain rivers. Without any correction of increased ow resistance using a reduced energy slope, the predicted bedload volumes are about a factor of 10 higher on average than the observed values. Simulation results were also used for a back-calculation of macro-roughness effects from bedload transport data, and compared with an independent estimate of ow resistance partitioning based on ow resistance data. Copyright 2010 John Wiley & Sons, Ltd.
KEYWORDS: bedload transport, numerical modelling, macro roughness, stress partitioning, steep gradient

Introduction
Sediment transport dynamics in steep channels may be quite different from low-gradient channels. Sediment storage in torrents and mountain rivers is often limited, whereas the transport capacity during extreme ood events is very high, resulting in rapid morphological changes. Bathurst et al. (1987) dened a mountain stream as a river in an upland area or an area with steep topography. Channel gradients are steep (ranging from 001 to 01 or steeper) and bed sediment contains a high proportion of gravel, cobbles and boulders. For the most upstream reaches in mountain catchments the term torrent is used in Europe. In the following the term mountain stream includes both torrents and mountain rivers. To distinguish between the latter two stream types, we follow a denition proposed by Rickenmann and Koschni (2010): a torrent is a stream with a catchment area less than 25 km2 and a mean channel gradient larger than 005, and it is prone to debris ow occurrence; a mountain river is less steep, but still substantial bedload transport may occur during ood events. There is often a strong interaction between hillslope processes and the channel network. Sediment pulses and sediment supply disturbances inuence the transport of bedload (Cui and Parker, 2005). Sediment transport may be supply limited (Bathurst,

2007) rather than controlled by the sediment transport capacity under given discharge and channel conditions. Steep headwater streams are characterized by a wide range of sediment sizes. The presence of large boulders, woody debris and bedrock constrictions inuences the bed morphology, resulting in large variations in channel geometry, stream ow velocity and roughness (Hassan et al., 2005). Flow resistance in steep mountain streams with irregular bed topography includes both grain friction and important form drag. Flows are typically characterized by low relative ow depths. Recent investigations and discussions of such ow conditions have been made by Bathurst (2002), Lee and Ferguson (2002), Smart et al. (2002), Aberle and Smart (2003) and Church and Zimmerman (2007). Limited sediment supply and increased ow resistance are the main reasons that observed bedload transport at steep channel gradients may be considerably smaller than predicted by conventional bedload transport equations (Rickenmann et al., 2006; Bathurst, 2007; Yager et al., 2007; Chiari et al., 2010). Under such conditions, the application of theoretical sediment transport equations derived from laboratory experiments may be problematic (Gomi and Sidle, 2003). Quantitative measurements of sediment and bedload transport in steep streams are very limited (Rickenmann, 2001). Measured

806

M. CHIARI AND D. RICKENMANN

transport rates in mountain streams may differ substantially from values predicted with such formulae (e.g. Bathurst et al., 1987; Gomez and Church, 1989; Rickenmann, 2001). Yager et al. (2007) demonstrated that accurately predicting bedload transport in steep, rough streams requires accounting for the effects of local sediment availability and drag due to rarely mobile particles. Also, measured transport rates in gravel-bed rivers and boulder-bed streams may vary by several orders of magnitude under similar (mean) ow conditions (e.g. Bathurst et al., 1987; Gomez, 1987; Reid and Laronne, 1995; Hegg and Rickenmann, 1999; Rickenmann, 2001). The importance of accounting for additional energy losses (or increased total ow resistance) in steep streams in the context of bedload transport calculations has been pointed out in several studies (e.g. Govers and Rauws, 1986; Palt, 2001; Rickenmann, 2001, 2005; Yager et al., 2007; Rickenmann and Koschni, 2010; Chiari et al., 2010; Zimmermann, 2010). In some of these studies, similar concepts to the grain/form resistance partitioning in lowland rivers were applied in steep streams with some success. Bedload transport equations for steep channels were typically developed and calibrated with ume experiments, where grain ow resistance was prevalent and macro-roughness bed elements were absent. Therefore sediment transport rates in steep streams may be overestimated using traditional sediment transport formulae because most of the ow energy in steep streams is lost to form and spill drag (Zimmermann, 2010). However, from his ume experiments on ow resistance in steep channels including a wide range of ow conditions, Zimmermann (2010) concluded that the traditional concept of separating grain and form resistance does not make sense for shallow ows in steep streams. Indeed, in steep and rough streams it appears to be difcult to distinguish between grain and form resistance (the latter term sometimes including also other sources of resistance such as spill drag, variation in cross-section geometry and ow width or woody debris), since larger particles are also part of the bed, of bedforms and of the bank, and may contribute in varying degrees to different types of resistance. Based on a recent analysis of more than 3500 items of eld data on ow resistance including many steep and rough streams, it is proposed to abandon the terminology of grain/form resistance partitioning as it was used in several studies on steep streams (Rickenmann and Recking, 2010). In fact, some of the previous
Table I. Formulae used for the simulations Type Flow velocity Flow resistance Reference Strickler (1923) Rickenmann (1996) Rickenmann (1996) Flow resistance partitioning Energy gradient associated with base-level resistance only Bedload transport Incipient motion Armour layer

approaches to correct for overestimation of bedload transport in steeper streams (e.g. Palt, 2001; Rickenmann, 2005), including the one implemented in the SETRAC model (Chiari et al., 2010), can also be considered as an attempt to separate between a base-level resistance and macro-roughness (sum of additional) ow resistance, the latter type normally not being accounted for with traditional bedload transport formulae. Increased ow resistance (due to macro-roughness) is important for intermediate- and large-scale roughness conditions sensu Bathurst et al. (1981). The objective of this study is to indirectly investigate the role of increased ow resistance on bedload transport by analysing a number of ood events which occurred in August 2005 in the Alps. To this end, we performed back-calculations of six extreme bedload transporting events with the recently developed numerical bedload transport model SETRAC for steep streams (Rickenmann et al., 2006; Chiari et al., 2010). SETRAC has been designed to reproduce sediment transfer at the catchment scale in steep streams during ood events for engineering applications and hazard assessment.

Simulations
The SETRAC model
The one-dimensional sediment routing model for steep torrent channel networks called SETRAC has been developed at the University of Natural Resources and Applied Life Sciences, Vienna. SETRAC is the acronym for Sediment Transport Model in Alpine Catchments. The model is described in Chiari et al. (2010). The model equations used in this study are presented in Table I. Different sediment transport formulae (Rickenmann, 1990, 1991, 2001; Smart and Jggi, 1983) and ow resistance approaches (Rickenmann, 1994; Smart and Jggi, 1983; Strickler, 1923) can be selected in SETRAC for steep channel gradients. An important new element is the possibility of taking increased ow resistance into account, which reduces the energy level available for bedload transport. Based on similar approaches of Meyer-Peter and Mueller (1948) and Palt (2001), Rickenmann (2005) proposed a rough and general empirical function to estimate increased ow (form) resistance in mountain rivers and torrents. The main idea is to calculate the

Equation

v=

1 067 05 R S ntot

(1) (2) (3) (4) (5)


15

097 g 041Q 019 1 = for S 0008 064 ntot S 019d90 436 g 049Q 002 1 = for S 0008 023 ntot S 003d90 nr h = 0092S 035 ntot d90
a

after Rickenmann (2005) Rickenmann et al. (2006) Rickenmann (2001) after Bathurst et al. (1987) after Jggi (1992) and Badoux and Rickenmann (2008)

( nn ) with 1 a 2 d q = 31( q q S s 1 d )
Sred = S
r tot 90 30 02 b

( )

033

c)

15

(6) (7) (8)

15 qc = 0065 ( s 1)167 g 05d50 S 112

qc ,D = qc

( dd )
90 m

10/9

Copyright 2010 John Wiley & Sons, Ltd.

Earth Surf. Process. Landforms, Vol. 36, 805815 (2011)

BACK-CALCULATION OF BEDLOAD TRANSPORT IN STEEP CHANNELS Table II. Characterization of the catchments used for the back-calculation of bedload transport during the August 2005 ood events Simulated case study Sessladbach Schnannerbach Suggadinbach Chirelbach Chiene Ltschine Catchment area (km2) 99 63 750 1305 905 1800 Mobilized bedload (m3) 16 33 50 240 120 150 000 000 000 000 000 000 Mean channel gradient 025 024 008 006 005 003 Sim. river length (m) 3 250 3 000 8 800 7 800 8 250 12 800 Discretisation length (m) 25 25 50 50 50 50 Peak discharge (m3 s-1) 25 24 100 100 90 140 Event duration (h) 47 35 43 59 58 59

807

a (best t) 15 10 15 (10) 15 15

ratio of grain to total ow resistance, and determine a reduced energy gradient which is related to grain resistance only (Chiari et al., 2010), and the approach is implemented in SETRAC (compare Equations (4) and (5) in Table I). Using this approach, much better agreement was obtained between observed and predicted bedload volumes for a ood event in 2005 in an Austrian torrent (Rickenmann et al., 2006), and for several ood events in 2000 and 1993 in mountain rivers in southwestern Switzerland (Badoux and Rickenmann, 2008). Armouring effects can also be considered in SETRAC. In addition, it is possible to calculate fractional bedload transport and consider grain-sorting effects by a hiding function (e.g. Parker, 2008) in combination with mobile bed conditions (Chiari, 2008; Chiari and Rickenmann, 2009). In SETRAC the channel network is represented by nodes, cross-sections and reaches. The sediment is routed through the channel network considering erosion and deposition via the sediment continuity condition. An initial, maximum erosion depth can be assigned for each channel reach; this feature is important in taking into account limited sediment availability. Morphological changes due to erosion and deposition can be calculated. For the calculations the cross-section is divided into strips to get a representative discretization of the prole. The number of strips will typically depend on the complexity of the cross-section. The stream width remains constant during the simulations. Flow hydrographs are routed as kinematic waves through the channel system. Sediment input as sedigraphs is also possible. The successor of the SETRAC model (TomSed) as well as a user guide, a technical manual and tutorials will be made available as a free download at www.bedload.at.

Case studies
Flood event data Extreme ood events occurred in August 2005 in Austria and in Switzerland. A general event documentation was prepared for Austria (Hochwasser, 2006). Field investigations of ood marks and geomorphological changes were made in several torrents and documented by the Institute of Mountain Risk Engineering at the University of Natural Resources and Life Sciences, Vienna (Hbl et al., 2005). Detailed investigations including a reconstruction of the ood hydrograph and the sediment budget along the main channel were made for the Sessladbach and Schnannerbach in Tyrol and for the Suggadinbach in Vorarlberg. The ood event in these torrents resulted in severe economic damage, due to massive bedload transport. In Switzerland detailed investigations for the Chirel mountain river (LLEDiemtigtal, 2006), the Chiene mountain river (LLEReichenbach, 2006) and the Schwarze Ltschine mountain river (LLELtschine, 2007) are available, and a general documentation of the ood events was prepared
Copyright 2010 John Wiley & Sons, Ltd.

(Bezzola and Hegg, 2007, 2008). This work also includes a more detailed analysis of bedload volumes transported by 33 debris ows and 39 events with uvial sediment transport (Rickenmann and Koschni, 2010). A characterization of the investigated catchments is given in Table II. Detailed information on the event reconstructions and back-calculations of the associated oods in the considered streams can be found in Chiari (2008). For all catchments, a sediment budget was established after the event. For the steep torrents Sessladbach and Schnannerbach the sediment erosion and deposition along the channel were mapped in the eld shortly after the ood events. Erosion and deposition volumes were measured (whenever possible) or estimated in the eld for (quasi-)homogeneous reaches with similar bed, erosion/ deposition and channel gradient. Separate estimates of morphological changes were made for the main channel, the banks and sediment input from the banks or hill slopes. For all other catchments morphological changes caused by the ood events were quantied with airborne LiDAR data, using highresolution digital elevation models (DEM) (Chiari et al., 2009). Two LiDAR-based DEMs are available for the catchments, representing the pre- and post-ood situation. For both methods, the erosion and deposition volumes are then accumulated for the entire channel starting from the most upstream point. The difference between accumulated erosion and deposition is the transported sediment volume. In most cases, massive deposition of sediments occurred on the alluvial fan outside of the stream channel during an early phase of the event, as reported by eyewitness accounts. This implies that most sediment transported as bedload was retained in the measurement area and did not leave the system at the channel outlet. These volumes were corrected for the pore volumes and the amount of ne sediment transported as suspended load further downstream. For this study it is assumed that the pore volume and the content of ne sediment make up about 50% of the erosion volumes. For depositional situations 30% of pore volume is considered. It is assumed that the suspended sediment and washload are transported farther downstream. This estimate is based on an earlier study assessing sediment transport conditions in the mountain river Reuss in central Switzerland during the ood event of 1987, where 45% of the total mobilized sediment was determined to have been ne material transported in suspension (Bezzola et al., 1991), and on measured fractions of 39% and 56% suspended load in two glacierized Swiss catchments (Gurnell et al., 1988). The sum of mobilized (eroded) bedload volumes are given in Table II. In the Chirel catchment sediment input by debris ows from some tributaries was estimated from geomorphological eld evidence, and the corresponding sediment volumes were represented as triangular sedigraphs in the simulation. For the eld-based volume estimates, depositions on the fan could be be assessed more accurately than channel or lateral erosion in steep reaches. These estimates of sediment volumes may be
Earth Surf. Process. Landforms, Vol. 36, 805815 (2011)

808

M. CHIARI AND D. RICKENMANN

associated with an uncertainty range of up to a factor of 2 depending on the experience of the expert in the eld (Rickenmann and Koschni, 2010). The accuracy of the LiDAR-based volume estimates is slope dependent and is higher for milder reaches (Scheidl et al., 2008). The mean volume error has been determined as 03 m3 m-2 stream bed area for the analysed catchments (Chiari et al., 2009), which results in an uncertainty of 35% on average regarding the maximum accumulated bedload transport volumes calculated with LiDAR data. The grain size distribution was estimated with a pebble count and evaluated after Fehr (1987) on a channel reach basis. The input hydrographs were generated with the Hydrologic Modeling System (HEC-HMS) (Scharffenberg and Flemming, 2005) and calibrated according the ood reconstructions. In three Swiss case studies (Chirel, Chiene, Ltschine) an independent estimate of the streamow runoff is based on stream gauge measurements upstream and downstream of the conuence with the main river. Agreement of these estimates with HEC-HMS simulations is discussed in Rickenmann and Koschni (2010). These reconstructions are described for each case study in this paper seperately. The formulae used in SETRAC for the systematic backcalculation of the presented extreme ood events are listed in Table I. To account for limited sediment availability, all simulations consider supply-limited conditions. Possible erosion depths were estimated in the eld and range from 005 m in channel reaches mainly in bedrock to 100 m for reaches in alluvial bed sediment. A constant uid density of 1100 kg m-3 is assumed for all simulations. Sediment transport is calculated using the one-grain model. Bed-level changes due to erosion and deposition are considered along the longitudinal prole. For each catchment several simulations are performed. The rst simulation considers the full transport capacity, the second an armour layer criterion, and one simulation is performed accounting for increased ow resistance (i.e., macro-roughness), with the exponent a in Equation (5) serving as a calibration parameter. Sessladbach The torrent Sessladbach is a tributary of the Trisanna River in the Paznauntal valley in Tyrol (Austria). The catchment and the simulated channel are characterized in Table II. The steepest parts are the middle reaches, with channel gradients up to 04. The gradient of the fan is 018. No streamow measurements are available for this channel. Therefore the input hydrographs needed for SETRAC were generated with HEC-HMS (Scharffenberg and Flemming, 2005). Data from a precipitation gauge situated in the catchment were used and the simulated peak discharge was calibrated using eld-based estimates of the peak discharge at several cross-sections. About 16 000 m3 of bedload were mobilized during the event. The sediment budget is based on eld estimates of erosion and deposition along the channel and the fan (Hochwasser, 2006). On the fan 8000 m3 of bedload sediment were deposited. Despite the steep gradient of the channel no evidence of debris ow occurrence was found in the catchment. Nevertheless, high sediment concentrations comparable to a debris ood (Hungr, 2005) may have developed during the event in the steep middle reaches. Figure 1 shows the longitudinal prole, the local channel gradient in the sections and the assumed storage layer depth (DSL) according to the eld survey. A comparison of the reconstructed transported bedload volume (ABT) for the August 2005 event and the SETRAC simulation results is shown. The simulation with the full transport capacity overestimates the recalculated bedload transport. All the sediment stock is depleted. Considering an armour layer criterion overestimates the total bedload transport to the same degree.
Copyright 2010 John Wiley & Sons, Ltd.

40000

ABT [m]

30000 20000 10000 0


longitudinal profile channel gradient depth storage layer

reconstructed bedload transport Simulation transport capacity Simulation armour layer (ident. with t. c.) Simulation macro-roughness a=1.5

1800 1600 1400 1200 1000 0 1 2 3

0.6 0.4 0.2 0

distance from outlet [km]

Figure 1. Longitudinal prole of Sessladbach, showing channel gradient, depth of storage layer, reconstructed ABT and comparison with SETRAC simulation results. This gure is available in colour online at wileyonlinelibrary.com/journal/espl

As in the rst simulation, all sediment stock is depleted. With a reduction of the energy gradient due to macro-roughness the modelled bedload transport is comparable with the observed volumes of transported bedload. The exponent a was optimized to nd best overall agreement between observed and reconstructed bedload volumes, resulting in a = 15. Schnannerbach The second well-documented torrential event of this study is the Schnannerbach in Tyrol. The Schnannerbach is a tributary of the Rosanna River in the Stanzertal valley (Tyrol, Austria). The catchment and the simulated channel are characterized in Table II. The steepest reaches have a channel gradient of 037 and the gradient of the fan is 010. No streamow measurements are available for Schnannerbach. The input hydrographs needed for SETRAC were generated with the hydrological model HEC-HMS. Data from precipitation gauges situated around the catchment were used. The simulated peak discharge was calibrated using eld-based estimates of the peak discharge at several cross-sections. About 33 000 m3 of bedload sediment were mobilized during the extreme event. In the upstream catchment sediment supply from the hillslopes is nearly unlimited. The middle reaches are mainly in bedrock, with some sediment supply from lateral erosion. The reach 200 m upstream of the fan is formed by a gorge. About 18 000 m3 of bedload were deposited on the inhabited fan (Hochwasser, 2006). Figure 2 shows the longitudinal prole, the local channel gradient in the sections and the assumed storage layer depth according to the eld survey. A comparison of the reconstructed transported bedload volume for the August 2005 event and the SETRAC simulation results is shown. The simulation with the full transport capacity overestimates the recalculated bedload transport to the same degree. All the sediment stock is depleted. Considering an armour layer criterion overestimates the total bedload transport. All sediment stock is depleted during the simulation. With a reduction of the energy gradient due to macro-roughness the modelled bedload transport is comparable with the observed amount of transported bedload. The exponent a was optimized to nd the best agreement between observed and reconstructed bedload volumes, resulting in a = 10. Suggadinbach The catchment of the Suggadin mountain river is situated in the western part of Vorarlberg (Austria). The catchment and the
Earth Surf. Process. Landforms, Vol. 36, 805815 (2011)

S [-] or DSL/10 [m]

altitude [m a.s.l.]

2000

0.8

BACK-CALCULATION OF BEDLOAD TRANSPORT IN STEEP CHANNELS

809

80000

ABT [m]

40000 20000 0
longitudinal profile channel gradient depth storage layer

ABT [m]

60000

reconstructed bedload transport Simulation transport capacity Simulation armour layer (ident. with t. c.) Simulation macro-roughness a=1.0

200000 150000 100000 50000 0


longitudinal profile channel gradient depth storage layer

reconstructed bedload transport Simulation transport capacity Simulation armour layer Simulation macro-roughness a=1.0

S [-] or DSL/10 [m]

1800 1600 1400 1200 1000 0 1 2 3 0 0.4 0.2

0.2 1000 800 600 0 2 4 6 8 0.15 0.1 0.05 0

distance from outlet [km]

distance from outlet [km]

Figure 2. Longitudinal prole of Schnannerbach, showing channel gradient, depth of storage layer, reconstructed ABT and comparison with SETRAC simulation results. This gure is available in colour online at wileyonlinelibrary.com/journal/espl

Figure 4. Longitudinal prole of Chirel, showing channel gradient, depth of storage layer, reconstructed ABT and comparison with SETRAC simulation results. This gure is available in colour online at wileyonlinelibrary.com/journal/espl

400000

ABT [m]

300000 200000 100000 0


longitudinal profile channel gradient depth storage layer

reconstructed bedload transport Simulation transport capacity Simulation armour layer Simulation macro-roughness a=1.5

1400 1200 1000 800 600 0 2 4 6 8

0.3 0.2 0.1 0

S [-] or DSL/40 [m]

altitude [m a.s.l.]

1600

overestimates the recalculated bedload transport. Considering an armour layer criterion slightly reduces the accumulated bedload transport but still overestimates the total transport. With a reduction of the energy gradient due to macroroughness the modelled bedload transport is in better agreement with the observed volumes of transported bedload. The exponent a was optimized to nd best agreement between observed and reconstructed bedload volumes, resulting in a = 15. Chirel The catchment of the Chirel mountain river is situated in the Canton Berne (Switzerland). The catchment and the simulated channel are characterized in Table II. The channel gradient varies from 0015 in the attest reaches up to 014 in the steepest reaches. There are no streamow measurements for the Chirel mountain river, but the discharge could be reconstructed with streamow measurements upstream and downstream of the conuence with the Simme River (LLEDiemtigtal, 2006). There is one gauging station 125 km upstream (Oberwil) and one 03 km downstream (Latterbach) in the Simme River. The catchment is surrounded by nine precipitation gauges, but none is situated in the catchment. Therefore radar precipitation data (calibrated with groundbased precipitation gauges; MeteoSchweiz, 2006) were used in an HEC-HMS simulation to generate the input hydrographs, which were calibrated with the reconstructed hydrograph at the conuence. Airborne LiDAR data were used to determine the morphological changes during the extreme ood event that occurred in August 2005. Two high-resolution elevation models for the Chirel river watershed were available for the pre- and post-event situation. Areas of erosion and deposition were veried using aerial photographs. About 240 000 m3 of bedload sediment were mobilized during the ood event. Several debris ow events supplied material from tributaries to the Chirel mountain river. The volumes of the debris ows were determined by analysis of the morphological changes in the tributaries. They were modelled as triangular-shaped sedigraphs at the respective conuence nodes with an assumed duration of 15 min, triggered at the highest rainfall intensities in the catchment. Figure 4 shows the longitudinal prole, the local channel gradient in the sections and the assumed storage layer depth according to the eld survey. A comparison of the reconstructed transported bedload volume for the August 2005 event and the SETRAC simulation results is shown. The
Earth Surf. Process. Landforms, Vol. 36, 805815 (2011)

distance from outlet [km]

Figure 3. Longitudinal prole of Suggadinbach, showing channel gradient, depth of storage layer, reconstructed ABT and comparison with SETRAC simulation results. This gure is available in colour online at wileyonlinelibrary.com/journal/espl

simulated channel are characterized in Table II. The channel gradient varies from 001 to 020. In the Suggadin mountain river there are two stream ow gauges, but the measurements are not reliable for the August 2005 event. Therefore a backcalculation of the discharge was made (Chiari et al., 2008). The peak discharge was recalculated for cross-sections with minor morphological changes and the time evolution of the ood event was calibrated with the stages of the stream gauge measurements. The precipitation gauges in the catchment failed during the August 2005 ood; therefore two other gauges close to the catchment were used for the rainfallrunoff simulation using the HEC-HMS model. The reconstructed bedload transport is based on pre- and post-event airborne LiDAR data to determine the morphological changes along the main channel. Areas of erosion and deposition were veried with aerial photographs. For each channel reach the amount of sediment eroded or deposited was calculated. These data were completed with records from sediment dredging after the ood. About 50 000 m3 of bedload sediment were mobilized during the ood event. Figure 3 shows the longitudinal prole, the local channel gradient in the sections and the assumed storage layer depth according to the eld survey. A comparison of the reconstructed transported bedload volume for the August 2005 event and the SETRAC simulation results is shown. The simulation with the full transport capacity clearly
Copyright 2010 John Wiley & Sons, Ltd.

S [-] or DSL/40 [m]

altitude [m a.s.l.]

altitude [m a.s.l.]

2000

0.6

1200

0.25

810

M. CHIARI AND D. RICKENMANN

200000

ABT [m]

ABT [m]

150000 100000 50000 0


longitudinal profile channel gradient depth storage layer

reconstructed bedload transport Simulation transport capacity Simulation armour layer Simulation macro-roughness a=1.5

250000 200000 150000 100000 50000 0


longitudinal profile channel gradient depth storage layer

reconstructed bedload transport Simulation transport capacity Simulation armour layer Simulation macro-roughness a=1.5

S [-] or DSL/20 [m]

1100 1000 900 800 700 0 2 4 6 8

0.2 0.15 0.1 0.05 0

0.2 0.15 0.1 0.05

1000 800 600 4 8 12

distance from outlet [km]

distance from outlet [km]

Figure 5. Longitudinal prole of Chiene, showing channel gradient, depth of storage layer, reconstructed ABT and comparison with SETRAC simulation results. This gure is available in colour online at wileyonlinelibrary.com/journal/espl

Figure 6. Longitudinal prole of Schwarze Ltschine, showing channel gradient, depth of storage layer, reconstructed ABT and comparison with SETRAC simulation results. This gure is available in colour online at wileyonlinelibrary.com/journal/espl

simulations with the full transport capacity overestimates the recalculated bedload transport. The simulation considering an armour layer criterion is in good agreement with the reconstructed bedload transport and can be regarded as a best-t simulation for the Chirel mountain river. Considering a moderate reduction of the energy gradient due to macro-roughness (with a = 10) underestimates the total amount of transported bedload material. Chiene The catchment of the Chiene mountain river is situated in the Canton Berne (Switzerland). The catchment and the simulated channel are characterized in Table II. The channel gradient varies from 0004 in the at middle reaches up to 017 in the steepest reaches. No streamow measurements for the Chiene mountain river are available, but the discharge could be reconstructed with streamow measurements upstream and downstream of the conuence with the Kander River (LLEReichenbach, 2006). There is one gauging station 6 km upstream (Hondrich) and one 6 km downstream (Rybrgg Frutigen) in the River Kander. Radar precipitation data (MeteoSchweiz, 2006) were used in an HEC-HMS simulation to generate the runoff hydrographs, which were calibrated with the reconstructed hydrograph at the conuence. Airborne LiDAR data of the pre- and post-ood event situation are available for the catchment and were used for the reconstruction of the morphological changes. During the event about 120 000 m3 of bedload sediment were mobilized. Most of the material was deposited in the at middle reaches and in the village of Kien, situated on the alluvial fan. Figure 5 shows the longitudinal prole, the local channel gradient in the sections and the assumed storage layer depth according to the eld survey. A comparison of the reconstructed transported bedload volume for the August 2005 event and the SETRAC simulation results is shown. The simulation with the full transport capacity clearly overestimates the recalculated bedload transport (Figure 5). Considering an armour layer criterion overestimates the total bedload transport to a similar degree. With a reduction of the energy gradient due to macroroughness, the modelled bedload transport is comparable with the observed transport. For the best-t simulation the calibration parameter was found to be a = 15. Schwarze Ltschine The catchment of the Schwarze Ltschine mountain river is situated in the Canton Berne (Switzerland) (LLELtschine,
Copyright 2010 John Wiley & Sons, Ltd.

2007). The catchment and the simulated channel are characterized in Table II. The Schwarze Ltschine has a channel gradient ranging from 0003 in the at lower reaches up to 019 in the steepest reaches. No streamow measurements are available for the Schwarze Ltschine mountain river, but further downstream streamow measurements are available for the rising and the falling limb of the ood hydrograph for the locations Gsteig (35 km downstream of the conuence with the Weisse Ltschine River) and Zweiltschinen (05 km upstream of the conuence at the Weisse Ltschine River). During the peak ow the stream gauges failed. Therefore the event hydrograph was generated with an HEC-HMS rainfall runoff simulation, using calibrated values of the radar precipitation data (MeteoSchweiz, 2006). The whole catchment of the Ltschine has been modelled in order to achieve a more reliable calibration for the August 2005 ood event. The simulated peak discharge was calibrated with reconstructed values of the peak discharge (LLELtschine, 2007) and the rising and falling limbs of the stream gauges. Airborne LiDAR data before and after the ood event are available for the catchment and were used for the reconstruction of the morphological changes. Figure 6 shows the longitudinal prole, the local channel gradient in the sections and the assumed storage layer depth (DSL) according to the eld survey. A comparison of the reconstructed transported bedload volume for the August 2005 event and the SETRAC simulation results is shown. During the event about 150 000 m3 of bedload sediment were mobilized. Most of the material was deposited in the at lower reaches. The simulation with full transport capacity overestimates the recalculated bedload transport. Considering an armour layer criterion overestimates the total bedload transport particularly in the steep channel reaches. For the atter reaches this simulation is in good agreement with the accumulated bedload transport. The simulation considering a reduction of the energy gradient due to macro-roughness losses (with the calibrated parameter a = 15) is in good agreement with the reconstructed bedload transport in the atter reaches, but the channel erosion in the steep part from 46 to 62 km is still overestimated.

Discussion
Simulation results
The SETRAC simulation results conrm the difculty of accurate predictions of bedload transport in steep mountain rivers
Earth Surf. Process. Landforms, Vol. 36, 805815 (2011)

S [-] or DSL/20 [m]

altitude [m a.s.l.]

altitude [m a.s.l.]

1200

0.25

1200

0.25

BACK-CALCULATION OF BEDLOAD TRANSPORT IN STEEP CHANNELS

811

30

20

10

0 0 10 20 30 40 50

simulation time [h]

Figure 7. Hydrograph and sedigraphs for the channel outlet of the Sessladbach.

and torrents. Generally the application of theoretical sediment transport equations derived from laboratory experiments results in an overestimation of the bedload transport in steep rough channels of up to three orders of magnitude (e.g. Bathurst et al., 1987; Gomez and Church, 1989; Reid and Laronne, 1995; Hegg and Rickenmann, 1999; Rickenmann, 2001; Gomi and Sidle, 2003; Rickenmann and Koschni, 2010). To model supply-limited conditions the maximum erosion depth can be limited. Performing the SETRAC calculations for such conditions does not reproduce the time evolution of observed bedload transport for several of the streams investigated in this study. The hydrograph as well as the sedigraphs of the simulation with full transport capacity (case A) are shown along with a simulation considering an armour criterion (case B) and the best-t simulation (case C with the exponent a = 15) in Figure 7 for the Sessladbach eld study. For case A the sedigraph peaks before the hydrograph, because sediment delivery is initially more rapid and then limited from the upstream reaches. This temporal evolution of the simulated bedload transport is not in agreement with observations of the inhabitants on the fan, indicating that most of the sediment was deposited in the village during times with higher discharges. After 17 h simulation time all the sediment stock is depleted. Conversely, for the cases B and C, considering armouring or macro-roughness, the temporal sequence of the event shows a simultaneously peaking hydrograph and sedigraph. For case C the total transported volume is in better agreement with the event reconstruction. For sediment routing calculations in gravel-bed rivers armouring may be an important element in reducing bedload transport (Porto and Gessler, 1999; Hunziker and Jggi, 2002). An armour criterion as implemented in SETRAC (Table I, Equation (8)) can be regarded as upper boundary (resulting in a lower limit for bedload transport volume estimates), because the incipient motion criterion is increased for the whole simulation time. In natural rivers, once the armour layer is broken the incipient motion may be reduced because of the destroyed armour layer, or effectively a mobile armour layer may form. The simulation results of the mountain streams modelled within this study show that, considering armouring, the total bedload transport is reduced. Overall about 1020% less bedload is transported compared to simulations neglecting armouring in mountain streams (considering full transport capacity). For the Chirel mountain river this reduction of the transport capacity is sufcient to obtain a resonable agreement
Copyright 2010 John Wiley & Sons, Ltd.

bedload discharge [m/s]

water discharge [m/s]

Q Qb with full transport capacity (case A) Qb with armour criteria (case B) Qb with macro-roughness (case C)

between simulated and observed bedload transport. For the other mountain rivers (Suggadinbach, Chiene and Schwarze Ltschine) the total bedload transport is overestimated. Thus macro-roughness has to be considered for the recalculation of the August 2005 events in these mountain streams. For steep torrents, the application of the incipient motion criterion due to armouring also results in clear overestimation of observed bedload transport volumes. For the Sessladbach and the Schnannerbach no reduction of the accumulated bedload transport is achieved, because all the sediment stock is depleted. Therefore the consideration of a simple armour criterion in combination with a one-grain-size model cannot reproduce the recalculated bedload transport. Macro-roughness energy losses appear to be more important at very steep channel gradients. For torrent channels, where also colluvial sediment makes up the channel bed, the concept of armouring is questionable. Stable bed-forms like step-pool systems are typically an important morphological element and may only be partially destroyed during extreme ood conditions. The quantication of macro-roughness ow resistance is difcult. Approaches derived from natural data are rarely available, because ow resistance measurements are particularly rare in torrent channels. The dataset compiled by Rickenmann (1996) for the estimation of total roughness has been used for the quantication of the contribution of grain roughness to total roughness. Macro-roughness ow resistance calculated in SETRAC can be regarded as upper boundary because, if considered, it is calculated for the whole simulation time. Steep torrents with stable bed structures may show a different behaviour concerning form roughness, once the stable structures are partly or completely destroyed. Macro-roughness ow resistance may then become less important, but may still be present due to big boulders, other roughness structures and a highly variable channel width. On the other hand, some increase in ow resistance may be expected due to bedload transport (e.g. Song et al., 1998; Gao and Abrahams, 2004; Smart, 2006; Recking et al., 2008). In this study the exponent a in Equation (5) has been used as a calibration parameter in combination with Equation (4), accounting for macroroughness energy losses. In SETRAC possible values for a are in the range 10 < a < 20. For this study a best t was found for the calibrated parameter a ranging from 10 to 15.

Comparison between transport capacity and observed transport on a channel reach basis
For further analysis of the presented case studies, the full transport capacity for each cross-section was also calculated (i.e., not considering a macro-roughness ow resistance correction). For this additional set of simulations, we did not consider morphological changes of the cross-section geometry and of the channel gradient between the cross-sections. To model unlimited supply conditions, the possible erosion depth was set to the high value of 100 m. For comparison with the reconstructed bedload volumes, time-integrated bedload volumes were calculated. For further analysis the ratio of the simulated to the reconstructed (observed) bedload transport, r = Vb,sim/Vb,obs, is plotted in Figure 8. For steeper torrents (Sessladbach and Schnannerbach) limited sediment supply may be more important than for mountain rivers, where generally more sediment is stored in the channel bed. Figure 8 shows different patterns for the torrents (lled symbols) and mountain rivers (open symbols). No clear trend can be found for the torrents, but the mountain rivers indicate an increasing overestimation with increasing channel reach gradient. With the exception of some reaches, the observed bedload transport is
Earth Surf. Process. Landforms, Vol. 36, 805815 (2011)

812
1000.0

M. CHIARI AND D. RICKENMANN


1 eq. (10) eq. (4) eq. ( 9) eq. (11) eq. (12)

100.0

0.8

r=Vb,sim/Vb,obs

10.0

0.6

nr /ntot
1.0 Sessladbach Schnannerbach Suggadinbach Chirel Chiene Ltschine 0.01 0.1 1 0.1

0.4

0.2

0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7

0.0 0.001

Figure 8. Ratio of calculated transport capacity to reconstructed bedload transport versus channel gradient for all channel reaches. This gure is available in colour online at wileyonlinelibrary.com/journal/ espl

Figure 9. Comparison of two approaches to estimate the relative proportion of macro and grain roughness (ow resistance partitioning) in steep streams, shown as a function of channel gradient.

overestimated by one to three orders of magnitude. In several reaches with low to moderate channel gradients the actual bedload transport is underestimated. This may be an effect of the non-mobile bed approach used in the SETRAC calculations for this comparison and may not be representative, because changes of the channel gradient along the longitudinal prole can increase the transport capacity, particularly in atter reaches.

nr = 007S 043 ntot nr = 007S 047 ntot

(based on Eqn ( 4)) (based on Eqn (10))

(11)

(12)

Back-calculated ow resistance partitioning on a channel reach basis


Palt (2001) derived an approach to estimate the contribution of grain roughness to total roughness from eld data of Himalayan mountain rivers for the channel gradient range 0002 < S < 012:

nr = 01S 036 ntot

(9)

where nr is Mannings roughness coefcient associated with grain friction only (or base-level resistance), ntot is Mannings roughness coefcient of total roughness and S is the channel gradient. Rickenmann et al. (2006) proposed an alternative equation to the one given in Table I to estimate ow resistance partitioning as a function of discharge, a characteristic grain size and the channel gradient, based on the same eld data (Rickenmann, 1994) and accounting for grain roughness after Wong and Parker (2006):

Equations (10) and (4) are compared in Figure 9, where they are shown together with the dataset of ow velocity measurements used by Rickenmann (1994, 1996) and with the trend lines for nr/ntot as a function of channel gradient S. A large variation in the nr/ntot ratio is observed for channel gradients up to about 01. For steeper channels grain roughness tends to be only a small fraction of total roughness. The main reason to deduce trend lines in the form of Equations (11) and (12) is to allow a comparison with similar trend functions as estimated from bedload data. A more recent analysis of ow resistance data (Rickenmann and Recking, 2010) suggests that nr/ntot is mainly a function of relative ow depth. A comparable expression can be derived from the presented eld and bedload transport simulation data assuming that nr/ntot only depends on the channel gradient. Considering only data where the ratio of simulated transport and reconstructed bedload transport is r > 1, it is possible to express the contribution of grain (or base-level) roughness to total roughness (nr/ntot) as a function of the channel gradient by considering the bedload transport Equation (6) integrated over the channel width:
5 1 Vb ,sim A (Q Qc ) S15 S 15 Sred Vb ,obs A (Q Qc ) Sred

r=

(13)

0133Q 019 nr = 0096 047 019 ntot g d90 S

(10)

where Vb,sim is the simulated bedload volume, Vb,obs is the observed bedload Volume, A is a constant, Qc is the critical discharge for the beginning of motion and Sred is the energy gradient associated with grain friction only. Combining Equations (13) and (5) gives

where Q is the discharge, g is the acceleration due to gravity and d90 the characteristic grain size for which 90% of the material is ner by weight. Fitting a power law regression for nr/ntot as a function of channel gradient to the eld data on ow resistance (Figure 9) calculated using Equations (4) and (10), respectively, results in
Copyright 2010 John Wiley & Sons, Ltd.

nr ( 1a ) = r 15 ntot

(14)

According to the ManningStrickler equation an appropriate value of the exponent a in Equation (5) should be a = 2. Using
Earth Surf. Process. Landforms, Vol. 36, 805815 (2011)

BACK-CALCULATION OF BEDLOAD TRANSPORT IN STEEP CHANNELS


1

813

0.8

0.8

0.6

eq. (9)

0.6

eq. (9)

0.4

y = 0.21x-0.28 R2 = 0.43

0.4

y = 0.07x-0.47 R2 = 0.43

eq. (12)

0.2 eq. (11)

eq. (12)

0.2 eq. (11)

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7

Figure 10. Contribution of grain roughness to total roughness as a function of channel gradient using one single a = 20 in Equation (14) for all stream reaches.

Figure 11. Contribution of grain roughness to total roughness as a function of channel gradient using one single a = 12 in Equation (14) for all stream reaches.

this value and Equation (14) in combination with the bedload data (r-values) presented in Figure 8 results in nr/ntot estimates as shown in Figure 10. Fitting a power law regression to the data in Figure 10 gives

nr = 021S 028 ntot

(15)

In the case study simulations a correction procedure to account for macro-roughness after Equation (4) has been applied and the exponent a in Equation (5) has been used as calibration parameter. For the best-t simulations the calibrated exponent a was in the range 1015 for the different case study catchments; in this case, a represents a mean value for a given stream along the entire channel.

Conclusions
Equation (15) is based on bedload data for the channel gradient range 0007 < S < 052 and the coefcient of correlation is R2 = 043. A comparison between Equations (9), (11), (12) and (15) is also presented in Figure 10. With a = 2 macroroughness effects appear to be less important than predicted with Equations (9), (11) and (12). Meyer-Peter and Mueller (1948) showed theoretically that the exponent a in Equation (5) may vary between 20 and 133, and from their experiments they empirically determined a value of 15. Therefore the exponent a has been varied to bring the eld data from this study into better agreement with the trend of macro-roughness ow resistance (Figure 9; Equations (9), (11) and (12)). The best t was found for a = 120 (Figure 11). For this condition, the trend line is represented by a function identical to Equation (12): The SETRAC model has been applied successfully to bedloadtransporting ood events in the Austrian and Swiss Alps in August 2005. The back-calculations of well-documented extreme events stress the importance of considering macroroughness or increased ow resistance. Neglecting the effect of macro-roughness in steep channels results in an overestimation of the observed bedload transport by about a factor of 10 on average for our channel reach back-calculations. Modelling supply-limited conditions by limiting the sediment stock in reaches cannot reproduce the time evolution of some extreme events. All the sediment stock would be depleted within the rising limb of the hydrograph, which is contrary to observations. A simple armour criterion in combination with a one-grain model appears not to be sufcient in many case studies to better match calculated and observed bedload transport, particularly at steep channel slopes. Therefore macro-roughness is likely to be important for a better prediction of bedload transport during ood events in steep channel systems. Further investigations are required to develop more reliable and site-specic approaches to take into account increased ow resistance for the calculation of bedload transport in steep channels.

nr = 007S 047 ntot

(16)

Data shown in Figures 10 and 11 include all channel reaches with r > 1, assuming full transport capacity and neglecting sediment availability. Using an average exponent a = 12, the trend line of the bedload data (Equation (16)) in Figure 11 is in good agreement with the trend lines based on the ow resistance data (Equations (9), (11) and (12)). It is important to note that the derivation of the two groups of trends are based on completely independent datasets, but the different datasets show the same general trend for the contribution of grain (or base-level) roughness to total roughness. The eld data on ow resistance (Figure 9) show a comparable scatter for milder channel gradients (S smaller than about 01) to the back-calculated form roughness based on the bedload data (Figures 10 and 11).
Copyright 2010 John Wiley & Sons, Ltd.

AcknowledgementsWe would like to thank the Austrian Science Fund (FWF):[L147] for nancial support through the Translational Research Program on Sediment routing model for steep torrent channels, which enabled this work to be undertaken. The Swiss Federal Ofce for the Environment (BAFU) supported the case studies in the canton of Berne and provided the LiDAR DTM for the situation after the 2005 event. The LiDAR DTM before the 2005 event was provided by the Swiss Federal Ofce of Topography (swisstopo). LiDAR data for the Suggadin case study were provided by the Landesvermessungsamt Feldkirch. We thank the editors and the two referees for their constructive comments, which helped to improve the manuscript.
Earth Surf. Process. Landforms, Vol. 36, 805815 (2011)

814

M. CHIARI AND D. RICKENMANN dHrens, Switzerland. In Sediment Budgets. IAHS Publication no. 174; 431441. Hassan MA, Church M, Lisle TE, Brardinoni F, Benda L, Grant GE. 2005. Sediment transport and channel morphology of small, forested streams. Journal of the American Water Resources Association 41: 853876. Hegg C, Rickenmann D. 1999. Comparison of bedload transport in a steep mountain torrent with a bedload transport formula. In Hydraulic Engineering for Sustainable Water Resources Management at the Turn of the Millennium: Proceedings 28th IAHR Congress, Graz, Austria (CD-ROM). Hochwasser. 2006. Ereignisdokumentation. Technical report, der Bundeswasserbauverwaltung, des Forsttechnischen Dienstes fr Wildbach- und Lawinenverbauung und des Hydrographischen Dienstes. Bundesministerium fr Land- und Forstwirtschaft, Umwelt- und Wasserwirtschaft. Hbl J, Ganahl E, Bacher M, Chiari M, Holub M, Kaitna R, Prokop A, Dunwoody G, Forster A, Kerschbaumer M, Schneiderbauer S. 2005. Ereignisdokumentation. Tirol, Band 1: Generelle Aufnahme (5W-Standard). IAN Report 109, Institut fr Alpine Naturgefahren, BMLFUW. Hungr O. 2005. Classication and terminology. In Debris-Flow Hazards and Related Phenomena, Jakob M, Hungr O (eds). Springer: Berlin; 923. Hunziker R, Jggi MNR. 2002. Grain sorting processes. Journal of Hydraulic Engineering 128: 10601068. Jggi M. 1992. Sedimenthaushalt und Stabilitt von Flussbauten. Mitteilungen der Versuchsanstalt fr Wasserbau, Hydrologie und Glaziologie der ETH Zrich Nr. 119. Lee A, Ferguson R. 2002. Velocity and ow resistance in step-pool streams. Geomorphology 46: 5971. LLEDiemtigtal. 2006. Lokale Lsungsorientierte Ereignisanalyse (LLE) Hochwasser 2005 im Diemtigtal. Technischer Bericht, Auftrag des Tiefbauamtes des Kantons Bern. LLELtschine. 2007. Lokale Lsungsorientierte Ereignisanalyse (LLE) Ltschine. Technischer Bericht, Auftrag des Tiefbauamtes des Kantons Bern. LLEReichenbach. 2006. Lokale Lsungsorientierte Ereignisanalyse (LLE) Reichenbach. Technischer Bericht, Auftrag des Tiefbauamtes des Kantons Bern. MeteoSchweiz. 2006. Starkniederschlagsereignis. Arbeitsberichte der MeteoSchweiz 211, Bundesamt fr Meteorologie und Klimatologie. Meyer-Peter E, Mueller R. 1948. Formulas for bedload transport. In Proceedings of the 2nd meeting of the International Association of Hydraulic Structures Research, Stockholm, Sweden; 3964. Palt S. 2001. Sedimenttransporte im Himalaya-Karakorum und ihre Bedeutung fr Wasserkraftanlagen. Mitteilung 209 des Instituts fr Wasserwirtschaft und Kulturtechnik, Universitt Karlsruhe. Parker G. 2008. Transport of gravel and sediment mixtures. In Sedimentation Engineering: Theories, Measurements, Modeling, and Practice. ASCE Manuals and Reports on Engineering Practice No. 110, Reston, VA; 165252. Porto P, Gessler J. 1999. Ultimate bed slope in Calabrian streams upstream of check dams: eld study. Journal of Hydraulic Engineering 125: 12311242. Recking A, Frey P, Paquier A, Belleudy P, Champagne JY. 2008. Bed-load transport ume experiments on steep slopes. Journal of Hydraulic Engineering 134: 13021310. Reid I, Laronne J. 1995. Bedload sediment transport in an ephemeral stream and comparison with seasonal and perennial counterparts. Water Resources Research 31: 773781. Rickenmann D. 1990. Bedload transport capacity of slurry ows at steep slopes. Mitteilung 103 der Versuchsanstalt fr Wasserbau, Hydrologie Glaziologie, ETH Zrich. Rickenmann D. 1991. Hyperconcentrated ow and sediment transport at steep slopes. Journal of Hydraulic Engineering 117: 1419 1439. Rickenmann D. 1994. An alternative equation for the mean velocity in gravel-bed rivers and mountain torrents. In Proceedings of the ASCE 1994 National Conference on Hydraulic Engineering, Vol. 1, Cotroneo G, Rumer RR (eds), Buffalo, NY; 672676. Rickenmann D. 1996. Fliessgeschwindigkeit in Wildbchen und Gebirgsssen. Wasser Energie Luft 88: 298304.
Earth Surf. Process. Landforms, Vol. 36, 805815 (2011)

References
Aberle J, Smart G. 2003. The inuence of roughness structure on ow resistance on steep slopes. Journal of Hydraulic Research 41: 259 269. Badoux A, Rickenmann D. 2008. Berechnung zum Geschiebetransport whrend der Hochwasser 1993 und 2000 im Wallis. Wasser Energie Luft 3: 217226. Bathurst J. 2002. At-a-site variation and minimum ow resistance for mountain rivers. Journal of Hydrology 269: 1126. Bathurst J. 2007. Effect of coarse surface layer on bed-load transport. Journal of Hydraulic Engineering 133: 11921205. Bathurst J, Li R, Simons D. 1981. Resistance equation for large scale roughness. Journal of the Hydraulics Division 107: 15931613. Bathurst J, Graf W, Cao H. 1987. Bed load discharge equations for steep mountain rivers. In Sediment Transport in Gravel-Bed Rivers, Thorne CR, Bathurst JC, Hey RD (eds). Wiley: New York; 453477. Bezzola GR, Hegg C. 2007. Ereignisanalyse Hochwasser 2005, Teil 1: Prozesse, Schden und erste Einordnung. Umwelt-Wissen 0707, Bundesamt fr Umwelt BAFU, Eidgenssische Forschungsanstalt WSL. Bezzola GR, Hegg C. 2008. Ereignisanalyse Hochwasser 2005, Teil 2: Analyse von Prozessen, Massnahmen und Gefahrengrundlagen. Umwelt-Wissen 0825, Bundesamt fr Umwelt BAFU, Eidgenssische Forschungsanstalt WSL. Bezzola G, Hunziker R, Jggi M. (1991). Flussmorphologie und Geschiebehaushalt whrend des Ereignisses vom 24/25 August 1987 im Reusstal. Mitteilung Nr. 4 des Bundesamtes fur Wasserwirtschaft sowie Mitteilung Nr. 14 der Landeshydrologie und -geologie, Bern; 101104. Chiari M. 2008. Numerical modelling of bedload transport in torrents and mountain streams. PhD thesis, Dissertation an der Universitt fr Bodenkultur Wien, Institut fr Alpine Naturgefahren. Chiari M, Rickenmann D. 2009. Modellierung des Geschiebetransportes mit dem Modell SETRAC fr das Hochwasser im August 2005 in Schweizer Gebirgsssen. Wasser Energie Luft 101: 319327. Chiari M, Mair E, Rickenmann D. 2008. Geschiebetransportmodellierung in Wildbchen und Vergleich der morphologischen Vernderung mit LiDAR Daten. In Schutz des Lebensraumes vor Hochwasser, Muren, Massenbewegungen und Lawinen. Interpraevent, Dornbirn, Vorarlberg, Austria. Conference Proceedings, Vol. 1; 295306. Chiari M, Scheidl C, Rickenmann D. 2009. Assessing morphologic changes caused by torrential ood events with airborne lidar data. In Land Conservation (LANDCON 0905) (CD-ROM). Chiari M, Friedl K, Rickenmann D. 2010. A one dimensional bedload transport model for steep slopes. Journal of Hydraulic Research 48: 152160. Church M, Zimmerman A. 2007. Form stability of step-pool channels: research progress. Water Resources Research 43: W03415. Cui Y, Parker G. 2005. Numerical model of sediment pulses and sedimentsupply disturbances in mountain rivers. Journal of Hydraulic Engineering 131: 646656. Fehr R. 1987. Geschiebeanalysen in Gebirgsssen. Mitteilung 92 der Versuchsanstalt fr Wasserbau, Hydrologie und Glaziologie, ETH Zurich. Gao P, Abrahams AD. 2004. Bedload transport resistance in rough open-channel ows. Earth Surface Processes and Landforms 29: 423435. Gomez B. 1987. Bedload. In Glacio-uvial Sediment Transfer: An Alpine Perspective, Gurnell AM, Clark MJ (eds). Wiley: New York; 355376. Gomez B, Church M. 1989. An assessment of bedload sediment transport formulae for gravel bed rivers. Journal of Water Resources Research 25: 11611186. Gomi T, Sidle R. 2003. Bed load transport in managed steep-gradient headwater streams of southeastern Alaska. Water Resources Research 39: 1336. Govers G, Rauws G. 1986. Transporting capacity of overland ow on plane and on irregular beds. Earth Surface Processes and Landforms 11: 515524. Gurnell A, Warburton J, Clark M. 1988. A comparison of the sediment transport and yield characteristics of two adjacent glacier basins, Val

Copyright 2010 John Wiley & Sons, Ltd.

BACK-CALCULATION OF BEDLOAD TRANSPORT IN STEEP CHANNELS Rickenmann D. 2001. Comparison of bed load transport in torrents and gravel bed streams. Water Resources Research 37: 32953305. Rickenmann D. 2005. Geschiebetransport bei steilen Gefllen. Mitteilung 190 der Versuchsanstalt fr Wasserbau, Hydrologie und Glaziologie, ETH Zurich; 107119. Rickenmann D, Koschni A. 2010. Sediment loads due to uvial transport and debris ows during the 2005 ood events in switzerland. Hydrological Processes 24: 9931007. Rickenmann D, Recking A. 2010. Evaluation of ow resistance in gravel-bed rivers through a large eld dataset. Water Resources Research (submitted). Rickenmann D, Chiari M, Friedl K. 2006. SETRAC: a sediment routing model for steep torrent channels. In River Flow 2006, Vol. 1, Ferreira R, Leal EAJ, Cardoso A (eds). Taylor & Francis: London; 843 852. Scharffenberg W, Flemming M. 2005. Hydrologic modeling system HEC-HMS. Users Manual CPD-74A, Version 3.0.0. US Army Corps of Engineers, Davis, CA. Scheidl C, Rickenmann D, Chiari M. 2008. The use of airborne LiDAR data for the analysis of debris ow events in Switzerland. Natural Hazards and Earth System Science 8: 11131127.

815

Smart G. 2006. The resistance of transported sediment. In River Flow 2006, Vol. 1, Ferreira RML, Alves ECTL, Leal JGAB, Cardoso AH (eds). Taylor & Francis: London; pp. 867875. Smart G, Jggi M. 1983. Sedimenttransport in steilen Gerinnen. Mitteilung 64 der Versuchsanstalt fr Wasserbau, Hydrologie und Glaziologie, ETH Zurich. Smart G, Duncan M, Walsh M. 2002. Relatively rough ow resistance equations. Journal of Hydraulic Engineering 128: 568578. Song T, Chiew Y, Chin C. 1998. Effect of bedload movement on ow friction factor. Journal of Hydraulic Engineering 124: 165175. Strickler A. 1923. Beitrge zur Frage der Geschwindigkeitsformel und der Rauhigkeitszahlen fr Strme, Kanle und geschlossene Leitungen. Mitteilungen des Amtes fr Wasserwirtschaft 16. Wong M, Parker G. 2006. Re-analysis and correction of bedload relation of Meyer-Peter and Mller using their own database. Journal of Hydraulic Engineering 132: 11591168. Yager EM, Kirchner JW, Dietrich WE. 2007. Calculating bed load transport in steep boulder bed channels. Water Resources Research 43: W07418. Zimmermann A. 2010. Flow resistance in steep streams: an experimental study. Water Resources Research 46: W09536.

Copyright 2010 John Wiley & Sons, Ltd.

Earth Surf. Process. Landforms, Vol. 36, 805815 (2011)

You might also like