the role of sediment in controlling steady-state bedrock...

26
The role of sediment in controlling steady-state bedrock channel slope: Implications of the saltationabrasion incision model Leonard S. Sklar a, , William E. Dietrich b a Department of Geosciences, San Francisco State University, 1600 Holloway Avenue, San Francisco, CA 94132, USA b Department of Earth and Planetary Science, University of California, Berkeley, CA 94720, USA Received 4 April 2004; received in revised form 30 August 2005; accepted 30 August 2005 Available online 13 July 2006 Abstract The saltationabrasion model is a mechanistic model for river incision into bedrock by saltating bedload, which we have previously derived and used experimental data to constrain all parameter values. Here we develop a method for applying the saltationabrasion model at a landscape scale, and use the model as a reference for evaluating the behavior of a wide range of alternative incision models, in order to consider the implications of the saltationabrasion model, as well as other models, for predicting topographic steady-state channel slope. To determine the single-valued discharge that best represents the effects of the full discharge distribution in transporting sediment and wearing bedrock, we assume all runoff can be partitioned between a low-flow and a high-flow discharge, in which all bedload sediment transport occurs during high flow. We then use the gauged discharge record and measurements of channel characteristics at a reference field site and find that the optimum discharge has a moderate magnitude and frequency, due to the constraints of the threshold of grain motion and bed alluviation by high relative sediment supply. Incision models can be classified according to which of the effects of sediment on bedrock incision are accounted for. Using the predictions of the saltationabrasion model as a reference, we find that the threshold of motion is the most important effect that should be represented explicitly, followed in order of decreasing importance by the cover effect, the tools effect and the threshold of suspension effect. Models that lack the threshold of motion over-predict incision rate for low shear stresses and under-predict the steady-state channel slope for low to moderate rock uplift rates and rock strengths. Models that lack the cover effect over-predict incision rate for high sediment supply rates, and fail to represent the degree of freedom in slope adjustment provided by partial bed coverage. Models that lack the tools effect over-predict incision rate for low sediment supply rates, and do not allow for the possibility that incision rate can decline for increases in shear stress above a peak value. Overall, the saltationabrasion model predicts that steady-state channel slope is most sensitive to changes in grain size, such that the effect of variations in rock uplift rate and rock strength may affect slope indirectly through their possible, but as yet poorly understood, influence on the size distribution of sediments delivered to channel networks by hillslopes. © 2006 Elsevier B.V. All rights reserved. Keywords: Rivers; Erosion; Sediment supply; Grain size; Magnitudefrequency; Landscape evolution 1. Introduction 1.1. Motivation One of the central problems in understanding the relief structure and topographic texture of tectonically Geomorphology 82 (2006) 58 83 www.elsevier.com/locate/geomorph Corresponding author. Tel.: +1 415 338 1204; fax: +1 415 338 7705. E-mail address: [email protected] (L.S. Sklar). 0169-555X/$ - see front matter © 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.geomorph.2005.08.019

Upload: others

Post on 28-Mar-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

(2006) 58–83www.elsevier.com/locate/geomorph

Geomorphology 82

The role of sediment in controlling steady-state bedrock channelslope: Implications of the saltation–abrasion incision model

Leonard S. Sklar a,⁎, William E. Dietrich b

a Department of Geosciences, San Francisco State University, 1600 Holloway Avenue, San Francisco, CA 94132, USAb Department of Earth and Planetary Science, University of California, Berkeley, CA 94720, USA

Received 4 April 2004; received in revised form 30 August 2005; accepted 30 August 2005Available online 13 July 2006

Abstract

The saltation–abrasion model is a mechanistic model for river incision into bedrock by saltating bedload, which we havepreviously derived and used experimental data to constrain all parameter values. Here we develop amethod for applying the saltation–abrasionmodel at a landscape scale, and use themodel as a reference for evaluating the behavior of a wide range of alternative incisionmodels, in order to consider the implications of the saltation–abrasion model, as well as other models, for predicting topographicsteady-state channel slope. To determine the single-valued discharge that best represents the effects of the full discharge distribution intransporting sediment and wearing bedrock, we assume all runoff can be partitioned between a low-flow and a high-flow discharge, inwhich all bedload sediment transport occurs during high flow.We then use the gauged discharge record and measurements of channelcharacteristics at a reference field site and find that the optimum discharge has a moderate magnitude and frequency, due to theconstraints of the threshold of grain motion and bed alluviation by high relative sediment supply. Incision models can be classifiedaccording to which of the effects of sediment on bedrock incision are accounted for. Using the predictions of the saltation–abrasionmodel as a reference, we find that the threshold of motion is themost important effect that should be represented explicitly, followed inorder of decreasing importance by the cover effect, the tools effect and the threshold of suspension effect. Models that lack thethreshold of motion over-predict incision rate for low shear stresses and under-predict the steady-state channel slope for low tomoderate rock uplift rates and rock strengths. Models that lack the cover effect over-predict incision rate for high sediment supplyrates, and fail to represent the degree of freedom in slope adjustment provided by partial bed coverage.Models that lack the tools effectover-predict incision rate for low sediment supply rates, and do not allow for the possibility that incision rate can decline for increasesin shear stress above a peak value. Overall, the saltation–abrasion model predicts that steady-state channel slope is most sensitive tochanges in grain size, such that the effect of variations in rock uplift rate and rock strength may affect slope indirectly through theirpossible, but as yet poorly understood, influence on the size distribution of sediments delivered to channel networks by hillslopes.© 2006 Elsevier B.V. All rights reserved.

Keywords: Rivers; Erosion; Sediment supply; Grain size; Magnitude–frequency; Landscape evolution

⁎ Corresponding author. Tel.: +1 415 338 1204; fax: +1 415 3387705.

E-mail address: [email protected] (L.S. Sklar).

0169-555X/$ - see front matter © 2006 Elsevier B.V. All rights reserved.doi:10.1016/j.geomorph.2005.08.019

1. Introduction

1.1. Motivation

One of the central problems in understanding therelief structure and topographic texture of tectonically

Page 2: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

59L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

active landscapes is the question of what controls theslope of an actively incising bedrock channel. Previouswork over several decades has led to an understandingof the dynamics of slope adjustment in alluvial channels(e.g. Mackin, 1948; Schumm, 1973; Bull, 1979; Snowand Slingerland, 1987), in which slope is controlled bythe long-term balance between sediment supply andsediment transport capacity. Howard (1980) identifiedtwo distinct alluvial regimes based on the grain sizedistribution of the sediment supply: an ‘active’ fine-bedregime in which the slope is set by the total sedimentsupply, and a ‘threshold’ coarse-bed regime in which theslope is controlled by the threshold of motion for thecoarsest grain size supplied in amounts sufficient toform a channel-spanning deposit. Early efforts to modellandscape evolution (e.g. Willgoose et al., 1991;Beaumont et al., 1992) treated channel incision andslope adjustment as an essentially alluvial process. Incontrast, fluvial incision and slope adjustment arerepresented in the current generation of landscapeevolution models (e.g. Howard, 1994; Moglen andBras, 1995; Tucker and Slingerland, 1996; Willett,1999; Bogaart et al., 2003) as controlled primarily by therate of bedrock detachment, rather than by thedivergence of sediment transport. The rate of riverincision into bedrock is commonly assumed to scalewith a flow intensity parameter, such as unit streampower, average boundary shear stress, or the boundaryshear stress in excess of a detachment threshold. Forthese ‘detachment limited’ models, the dominantinfluences on predicted channel slope are rock resis-tance, drainage area (a surrogate for discharge), andrates and spatial patterns of rock uplift, rather thansediment supply or grain size characteristics.

Sediment supply and grain size should stronglyinfluence rates of bedrock incision, however, as we haverecently shown through both experimental (Sklar andDietrich, 2001) and theoretical (Sklar and Dietrich,1998; Sklar and Dietrich, 2004) studies. In particular,sediment supplies tools that abrade exposed bedrock,but which can also inhibit incision by forming transientdeposits that limit the extent of bedrock exposure in thechannel bed. Moreover, grain size strongly influencesthe erosional efficiency of individual abrasive impactsas well as the minimum shear stress needed to preventcomplete alluviation of the channel bed. Thus, thecontrols on the slope of a channel actively incising intobedrock are likely to include elements drawn fromalluvial models (sediment supply, grain size andtransport capacity) as well as from detachment-limitedmodels (rock strength, flow intensity, tectonic boundaryconditions).

Here we explore the controls on bedrock channelslope using a mechanistic model for river incision intobedrock by saltating bedload (Sklar and Dietrich, 2004),in which the dynamics of bedload sediment transportand bedrock detachment by abrasion are directlycoupled. We address three distinct issues that arise inconsidering the implications of the ‘saltation–abrasion’model for river longitudinal form.

We begin with the challenge of scaling up a modelthat was derived at the temporal and spatial scale ofindividual bedload saltation events to the scale atwhich landscapes evolve. In particular, we develop anew method for determining the magnitude andfrequency of the discharge and sediment supply thatcan represent the integrated effect of the full distribu-tion of flow events that are responsible for erodingbedrock over geomorphic time. Next we compare thesaltation–abrasion model with other proposed bedrockincision models, to illustrate the importance ofrepresenting the effects of sediment supply and grainsize explicitly, and to evaluate the extent to which othermodels capture those effects. Finally, we consider theimplications of the saltation–abrasion model forunderstanding what controls the slope of a riverchannel for the limited but illustrative case when thelandscape is in topographic steady state. In particular,we contrast the predictions of several models for thevariation in steady-state channel slope with rock upliftrate, rock strength, and the magnitude and duration ofthe representative erosive flow event. Because profilerelief and concavity are directly related to the integraland derivative of local channel slope with downstreamdistance, this analysis provides a foundation forunderstanding the implications of the model forlongitudinal profile form.

1.2. Model overview

The saltation–abrasion model is based on the ideathat incision rate can be treated as the product of threeterms: the average volume of rock removed by anindividual bedload impact, the impact rate per unit timeper unit bed area, and the fraction of the bed not armoredby transient deposits of alluvium (Sklar and Dietrich,1998). We used several fundamental theoretical andexperimental findings to derive expressions for each ofthese terms, including: (1) wear is proportional to theflux of impact kinetic energy normal to the bed; (2)bedload saltation trajectories can be represented aspower functions of non-dimensional excess shear stress;(3) rock resistance to bedload abrasion scales with thesquare of rock tensile strength; and (4) the assumption

Page 3: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

60 L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

that bed cover depends linearly on the ratio of sedimentsupply to sediment transport capacity (Sklar andDietrich, 2004). The predicted instantaneous rate ofbedrock incision Ei can be expressed as

Ei ¼ 0:08RbgY

kvr2TQs

s4sc4

� 1

� ��1=2

1� Qs

Qc

� �

� 1� u4wf

� �2 !3=2

ð1Þ

where Rb is the non-dimensional buoyant density[Rb= (ρs−ρw) /ρw], g is the gravitational acceleration,Y is the rock modulus of elasticity, σT is the rocktensile strength, kv is a non-dimensional rock resistanceparameter (experiments indicate that kv∼106, which isequivalent to 1012 J m− 3 MPa− 1, Sklar and Dietrich,2004), u⁎ is the shear velocity [u⁎=(gRhS)

1/2], wf isthe grain fall velocity in still water (calculated hereusing the algorithm of Dietrich, 1982), and Qs is themass sediment supply rate (all variables in S.I. units).In Eq. (1), τ⁎ is the non-dimensional shear stress

s4 ¼ RhS=RbDs ð2Þwhere Rh is the hydraulic radius, S is the channel slope,Ds is the representative sediment grain diameter, and τc⁎is the value of τ⁎ at the threshold of sediment motion.Hydraulic radius is solved for iteratively, assuming arectangular channel cross-section [Rh=HwW / (2Hw+W)],using the continuity equation [Qw=UwHwW] and theManning roughness relation

Uw ¼ R2=3h S1=2

nmð3Þ

where Hw is flow depth, Uw is mean flow velocity, W ischannel width and nm is the Mannings roughnessparameter. Bedload sediment transport capacity Qc iscalculated here using the Fernandez-Luque and van Beek(1976) expression

Qc ¼ 5:7qsW ðRbgD3s Þ1=2ðs4� sc4Þ3=2: ð4Þ

Note that our choice of bedload transport relation issomewhat arbitrary as many other equivalent expressionsare available, however, the essential outcome of theanalysis below would not be changed so long as theexpression forQc included the nearly universal non-lineardependence of transport rate on shear stress in excess of athreshold of motion stress.

The model is unique in three important respects.First, it predicts the rate of bedrock incision by a singleprocess, abrasion by saltating bedload, rather than

lumping the effects of many possible erosional mechan-isms together in a single expression. Second, because allthe model parameters are independent of the modelformulation, it can be calibrated using laboratoryexperiments so that there are no freely adjustableparameters. And third, because it is derived at thetemporal and spatial scale at which the erosional processoccurs, it can be scaled up in space and time explicitly,so that, for example, the role of the tradeoffs betweenmagnitude and frequency in discharge and sedimentsupply can be explored systematically.

2. Model application at the landscape scale

The saltation–abrasion model (Sklar and Dietrich,2004) was derived at the temporal and spatial scale ofindividual bedload impacts, and predicts an instanta-neous bedrock erosion rate (Ei) for specific values of theseven input variables: channel width (W), slope (S),discharge (Qw), coarse sediment supply rate (Qs),representative grain size (Ds), rock strength (σT) andchannel roughness (nm). These values can be treated asconstant at a spatial scale of a channel reach, ignoringlateral variations in shear stress, and at the temporalscale of several hours to a day. To apply the model in thefield and to explore the model implications for riverlongitudinal profile form and dynamics, the model mustbe scaled up, in both time and space, by many orders ofmagnitude. The spatial variables can be scaled usingsimple relationships of downstream hydraulic geometry,while temporal scaling requires a method for integratingover the full range of discharges and sediment supplyevents that the channel will experience.

Although bridging the gap in scale between themechanistic derivation and the landscape applicationrequires the introduction of several additional para-meters, the physical meaning of these parameters canbe explicit. Maintaining a quantitative link to the scaleof individual erosive events allows us to be mecha-nistically consistent and observe the limitationsimposed by the fundamental scales of the problem(Dietrich and Montgomery, 1998). For example, themodel assumptions are violated if the grain diameter isgreater than the flow depth, or if the flow is nothydraulically rough (i.e. particle Reynolds numberRep=ρsu⁎Ds /μ<100).

To help ground the scaling up of the model inphysical reality, we use the device of a reference site, anactively incising mixed bedrock–alluvial bedded reachof the South Fork Eel River in Northern California(Seidl and Dietrich, 1992; Howard, 1998; Sklar andDietrich, 2004). We use gauge records of discharge,

Page 4: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

Table 1Reference site model input values

Drainage area (A) 112 km2

Channel slope (S) 0.0053Median grain diameter (Ds) 0.060 mChannel width (W) 18.0 mRock tensile strength (σT) 7.0 MPaAssumed long-term incision rate (Elt) 0.9 mm/yearPredicted instantaneous incision rate (Ei) 20 mm/year

These values approximate conditions in a gauged reach of the SouthFork Eel River, Mendocino County, California. Other model parametervalues held constant are listed in Sklar and Dietrich (2004), Table 3.

61L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

knowledge of rates of tectonic forcing, and measure-ments of channel geometry, grain size, rock strength andextent of bedrock exposure, to evaluate whether themodel predictions are consistent with the field setting.The reference site is also useful as an anchor point inappreciating how widely we vary the value of inputvariables in the sections below where we comparevarious bedrock incision models and explore thecontrols on channel slope. Table 1 lists some of therelevant characteristics of the South Fork Eel Riverreference site.

2.1. Spatial scaling from reach to longitudinal profile

For modeling purposes, a river profile can berepresented as a series of reaches of uniform width,slope and rock strength, each with a correspondingupstream drainage area (A). For simplicity, we treatchannel roughness as constant throughout the profile,although roughness may in fact vary systematically withchannel slope and grain size (e.g. Bathurst, 1993).Discharge for a given recurrence interval (Qw) isassumed to vary with drainage area as a power function

Qw ¼ bAp ð5Þwhere the exponent p captures the spatial variation inprecipitation due to orographic or other effects. Channelwidth is assumed to vary as a power function of somegeomorphically significant discharge

W ¼ cQ fw ð6Þ

such that width can be parameterized in terms ofdrainage area

W ¼ cb f Apf : ð7ÞIn all the calculations to follow we treat width asconstant for a given drainage area, consistent with theassumption of a confined bedrock gorge with arectangular channel cross-section, lacking any flood-

plain. Drainage area is parameterized as a function ofdistance downstream of the channel head (x)

A ¼ Ao þ dxh ð8Þwhere Ao is the drainage area upstream of the channelhead, d is a dimensional coefficient and the exponenth captures the variation in drainage area shape. Note thatthis formulation ignores the abrupt changes in drainagearea that occur at tributary junctions.

The commonly observed downstream fining of grainsize is typically treated as an exponential function (e.g.Morris andWilliams, 1999), however, many field studieshave found good agreement with a power law finingrelation (e.g. Brush, 1961; Brierley and Hickin, 1985;Ohmori, 1991). For consistency with the other spatialscaling relations we therefore model the downstreamvariation in representative grain diameter of the bedloadsize class as a power function of downstream distance

Ds ¼ Doxþ xoxo

� ��a

ð9Þ

whereDo is the grain size at the channel head and xo is theunchanneled distance from the drainage divide to thechannel head. Downstream changes in lithology can berepresented by continuous or abrupt variation in rocktensile strength; in all calculations below we treat rockstrength as uniform along the length of the profile.

Assuming that hillslope erosion rates closely trackchanges in rates of incision of neighboring channels, wecan express the time-averaged, or mean-annual bedloadsediment supply rate (QsM) as the integral of theupstream rates of river incision

QsM ¼ qsFb

ZEltðAÞdA ð10aÞ

which for the simple case of uniform uplift becomes

QsM ¼ qsFbEltA ð10bÞ

where Elt is the long-term bedrock incision rateaveraged over all discharge and sediment supply events

Elt ¼REiðtÞdtRdt

ð11Þ

and Fb is the fraction of the total sediment supply that iscoarse enough tomove as bedload. If we assume the riverprofile is in topographic steady state, such that the long-term bedrock incision rate is equal to the rock uplift rate(Elt =Ur), then, for the case of uniform rock uplift

QsM ¼ qsFbUrA: ð12Þ

Page 5: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

62 L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

The bedload fraction of the total load (Fb) can becalculated for a given boundary shear stress, if the fullgrain size distribution is specified, using a criterion forthe threshold of suspension, such as when the shearvelocity (u⁎) is equal to the fall velocity in still water(wf) (Rouse, 1937). However, because we use a singlegrain diameter (Ds) to represent the size of sedimentmoving as bedload, Fb becomes a parameter that mustbe estimated, rather than calculated. Measured values ofFb, from simultaneous sampling of both bedload andsuspended load, range from 0.01 to 0.5 (e.g. Leopold,1994), although field estimates of Fb in coarse-beddedmountain channels range as high as 0.82 (Bezinge,1987).

2.2. Representative discharge and sediment supply

Landscape evolution models typically treat bedrockincision as a continuous process, and do not accountexplicitly for the differing effects of individualdischarge and sediment supply events. This approxi-mation may be valid for erosional processes that do nothave thresholds, but because the saltation–abrasionmodel is based on the dynamics of bedload sedimenttransport we must consider the constraint imposed bythe threshold of sediment motion. For example, for theSouth Fork Eel River reference site, the mean-annualdischarge of 4 m3/s does not produce sufficient shearstress to mobilize the representative grain size of0.06 m, and thus would not cause bedrock wear by themechanism of bedload abrasion. We note that otherincision mechanisms, for example hydration-fracturingof clay-rich sedimentary rocks (Stock et al., 2005), maycause bedrock detachment at flows below the thresholdof sediment motion.

Ideally, to calculate a long-term incision rate, wewould model a time-series of discharges or integrateover a frequency distribution of discharges so that wecould account for the weighted contributions ofdischarges of all magnitudes. For example, Tucker andBras (2000) used the Poisson pulse rainfall model toderive a frequency distribution of flood discharges,which they integrated to explore (separately) thesensitivity of long-term rates of sediment transport andbedrock erosion to variability in runoff. Tucker and Bras(2000) did not, however, couple sediment transport andbedrock erosion, as we do here, as thus did notencounter the problem of specifying the instantaneoussediment supply rate for conditions of excess sedimenttransport capacity. In actively incising, mixed bedrock–alluvial channels, sediment flux will commonly be lessthan the discharge-dependent potential transport capa-

city. The extent of under-supply, or over-capacity,should vary considerably from discharge to discharge.This is because sediment supply from hillslopes to thechannel network is inherently episodic, and thefrequency distribution of sediment supply events, suchas landslides, debris-flows and bank failures, should beto some extent decoupled from the distribution ofrainfall and runoff events.

Here we use the representative discharge concept todevelop a rational method for treating incision as a semi-continuous process while accounting for the effect of thethreshold of sediment motion. The representativedischarge is a single-value flow rate of limited duration,which we assume to represent the integrated effect of thefull distribution of discharges. For consideration of theeffects of sediment on bedrock incision, the method alsomust specify a corresponding representative sedimentsupply rate, which we also assume represents theintegrated effect of the full distribution of sedimentsupply events. Here we seek to develop a simplemethod, with a limited number of parameters, that: (1)can be constrained by measured discharge distributionsto reflect differing rainfall and runoff variability acrossclimatic regimes; (2) accounts for the total volume ofrunoff, not just flood flows; (3) accounts for the totalcoarse sediment supply that will move as bedload ratherthan as suspended load; and (4) accounts for theinfluence of grain size on the threshold of sedimentmotion and thus on the bedload sediment transportcapacity of the representative discharge.

In the analysis that follows, we begin by arguing forthe use of two representative discharges, a high (flood)flow, and a low (base) flow, each with a specific partialannual duration, rather than a single intermittentrepresentative discharge. We then show that a cumula-tive discharge distribution can be used to relate therelative magnitudes and durations of the high and lowrepresentative discharges. Next we consider the effectsof the threshold of sediment motion and the non-linearity of the dependence of bedload sedimenttransport capacity on the representative discharge, tofind an optimal partitioning of cumulative runoffbetween the high and low flow representative dis-charges. The result is a determination of the meanannual duration and magnitude of the representativehigh flow discharge, and the associated sedimentsupply rate, which are responsible for bedrock incisionby bedload abrasion. Importantly, the representativesediment supply rate will be less than the sedimenttransport capacity of the representative high flow,consistent with the assumption of the saltation–abrasionmodel that excess transport capacity is required for

Page 6: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

Fig. 1. Cumulative distribution of mean daily discharges from 24 yearrecord of USGS gauge at South Fork Eel River reference site, (USGSgauge 11475500) ‘near Branscom’, California.

63L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

bedrock to be exposed in the channel bed and thussubject to abrasion.

2.3. Magnitude–frequency model development

A simple approach to constraining the magnitude ofthe representative discharge would be to assume that thetotal mean annual runoff was carried by a singlerepresentative discharge (QwT), such that

QwT ¼ QwM=Ft ð13Þwhere QwM is the mean annual discharge and Ft is anintermittency parameter equal to the fraction of totaltime during which the representative discharge isassumed to occur. In the simplest case, Ft =1 and themean annual discharge is assumed to represent the fulldischarge distribution (e.g. Tomkin et al., 2003). Asnoted above, in many channels the mean annualdischarge will be insufficient to mobilize the bedsediments, and will therefore be a poor choice for arepresentative discharge for estimating long-term bed-rock incision rates. For the general case of Ft <1 (e.g.Tucker and Slingerland, 1997), the use of Eq. (13) mayexaggerate the effects of extreme events because itneglects the fraction of annual runoff that occurs in lowmagnitude, high frequency discharges. Such low flowsmay account for a significant fraction of the annualrunoff in coarse-bedded channels with a high thresholdof motion.

To account for the potentially important contributionof low-magnitude flows, we define two characteristicdischarges, a representative high flow (QwH) andresidual low flow (QwL), which together account forall the annual runoff. We define a new parameter, thefraction of the total runoff carried by the high flow (Fw),such that

QwH ¼ QwMFw=Ft ð14Þand

QwL ¼ QwMð1� FwÞ=ð1� FtÞ: ð15ÞWe assume that by definition bedload transport onlyoccurs during the high-flow discharge so that thesediment supply during high flow is

Qs ¼ QsM=Ft: ð16Þsuch that no sediment is supplied during low flow, be-cause the shear stress is below the threshold of motion.

The two magnitude–frequency parameters, Fw andFt, are not independent, but can be related byconsidering a cumulative discharge distribution. Fig. 1

shows the cumulative distribution of mean dailydischarge for the South Fork Eel reference site. In Fig.1, the probability of non-exceedence (Pne) for a givendischarge is computed by ranking the entire dischargerecord in ascending order, such that

Pne ¼ m=ðN þ 1Þ ð17Þwhere m is the rank of an individual discharge and N isthe total number of measured discharges.

We can define the intermittency parameter Ft as

Ft ¼ 1� Pne ð18Þso that the fraction of the total volume of runoff carriedby the high flow Fw is

Fw ¼

Z11�Ft

QwðPneÞPnedPne

Z 1

0QwðPneÞPnedPne

: ð19Þ

Fig. 2a shows the relationship between Fw and Ft for thereference site gauge record. If the representative highflow is assumed to occur relatively frequently (Ft∼1) itwould carry most of the annual runoff, while a lessfrequent, but higher magnitude representative high flow(Ft≪1) would carry only a small fraction of the annualrunoff. This relationship is a way of characterizing the‘storminess’ of the discharge distribution, for examplethe curve would shift to the left for storm dominatedrivers, and to the right for spring-fed or snow-meltdominated rivers. The relationship between Fw and Ft isunlikely to be uniform within a drainage basin, forexample shifting to the right with increasing distancedownstream due to the integrating effect of flow routingthrough the channel network (Bras, 1990). Note thatalthough we have used empirical data, this approach

Page 7: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

64 L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

could be generalized for use with synthetic dischargefrequency distributions, such as the log-normal or log-Pearson distributions.

Fig. 2b shows the variation in the representativehigh-flow discharge (QwH) and low-flow discharge(QwL) with the high-flow time fraction (Ft), for thereference site discharge record, using Eqs. (14), (15),and (19). Also plotted in Fig. 2b is the representativedischarge that carries the total runoff (QwT, Eq. (13)). AsFt declines from 1.0 (where QwH=QwT=QwM, andQwL=0), the magnitude of the high-flow discharge

increases, but the rate of increase declines because of theincreasing fraction of the total runoff that is carried bythe low flow. In contrast, the rate of increase in totalrunoff representative discharge (QwT) is constant, so thatfor a 1 year recurrence interval (i.e. Ft =1 /365) QwT isabout an order of magnitude greater than QwH. Note thatfor very low values of the high flow time fraction Ft, therepresentative low flow is approximately equal to themean annual discharge (QwL∼QwM) and the high flowcarries a negligible fraction of the annual runoff.

The difference between the high flow (QwH) and totalflow (QwT) representative discharges is most significantwhen we consider sediment transport. For eachdischarge, we can calculate a corresponding sedimenttransport capacity (Qc, Eq. (4)). Fig. 2c shows thebedload sediment transport capacity at the South ForkEel reference site (Table 1) calculated for the high-flow(QcH) and total-runoff (QcT) representative dischargemethods. Also shown is the representative sedimentsupply rate (Qs), which is inversely proportional to thehigh flow time fraction (Ft) because we assume that allcoarse sediment is moved only during high flows (Eq.(16)). For Ft∼1, the average boundary shear stress isbelow the threshold of motion and the transport capacityis zero. With decreasing Ft (i.e. higher magnitude, lowerfrequency), once the threshold of motion is exceeded thenon-linearity of the sediment transport relation leads to arapid increase in transport capacity for both the high-flow and total-runoff methods. Transport capacity (Eq.(4)) exceeds sediment supply (Eq. (16)) when Ft <0.2(i.e. <70 days/year) for the total-runoff method andwhen Ft <0.08 (i.e. <29 days/year) for the high-flowmethod. For further decreases in Ft, the extent of excess

Fig. 2. Determination of a model high-flow discharge that bestrepresents the integrated effect of the full distribution of discharge andsediment supply events on bedload transport and river incision intobedrock. (a) Fraction of total runoff carried by representative high flowdischarge as a function of the fraction of total time that high flowoccurs. (b) Variation in representative discharge with high flow timefraction for two formulations, a single discharge that carries the totalannual runoff (QwT), and a pair of low-flow (QwL) and high-flow(QwH) discharges between which total annual runoff is partitioned. (c)Variation with high flow time fraction, for the total runoff formulation,of sediment transport capacity (QcT), and for the high flowformulation, sediment transport capacity (QcH) and sediment supply(Qs), assuming all bedload transport occurs during the representativehigh-flow discharge. (d) Variation with high flow time fraction of: theratio of sediment supply (Qs) to high flow transport capacity (QcH),assumed equivalent to the extent of bedrock exposure; the ratio of thenet annual high flow transport capacity to the net annual transportcapacity integrated over the full gauged discharge distribution, whichpeaks closest to unity at Ft =0.0437; and the ratio of sediment supply tothe transport capacity of the representative discharge using the totalrunoff formulation (QcT).

Page 8: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

65L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

sediment transport capacities predicted by the twomethods diverge significantly. The high flow transportcapacity (QcH) only exceeds the sediment supply for anarrow range of Ft (0.016≤Ft≤0.08), while the totalrunoff transport capacity exceeds sediment supply for allFt <0.2.

Fig. 2d shows the ratio of sediment supply totransport capacity (Qs /Qc), which in the saltation–abrasion model corresponds to the fraction of the bedcovered with alluvium, for the high-flow discharge (Qs/QcH) and the total-discharge (Qs/QcT), for reference siteconditions. The high-flow method predicts only alimited extent of bedrock exposure, at an optimum Ft,while the total-runoff method predicts very largebedrock exposure over a wide range of Ft.

Because the representative discharge approach isintended to approximate the long-term net effect of thefull distribution of discharges, we can use the timeintegral of the sediment transport capacity of the gaugeddischarges at the reference site as a metric to evaluate theefficacy of the two representative discharge methods.Over the 24-year period of record, the annual meanintegrated bedload transport capacity (∑QcA), forreference conditions, is 7.26×107 kg/year. Plotted inFig. 2d is the ratio of the high flow annual transportcapacity to the annual capacity of the gauged discharges(∑QcH/∑QcA), calculated over the range of Ft.

The high flow representative discharge methodapproximates the net sediment transport capacity ofthe actual discharge record best at the optimum Ft valueof 0.0437 (i.e. ∼16 days/year), differing only by about5%. In contrast, the total runoff method annual netsediment transport capacity (not plotted) exceeds theactual integrated mean capacity by a factor of 2 atFt∼0.2, and grows rapidly to a ratio of greater than 95 atFt =0.0437, and greater than 5000 by Ft =0.001.

The preceding analysis provides a rational procedurefor assigning values to the magnitude–frequency scalingparameters; for the reference site we find an optimalvalue for the high flow time fraction of Ft =0.0437,which corresponds to a high flow runoff fraction ofFw=0.41. The result that nearly 60% of the annualdischarge is carried by the representative low-flowdischarge shows the importance of using a flowpartitioning approach in generating a representativedischarge for simulating bedload sediment transport as asemi-continuous process.

2.4. Field reference site calibration

We have used the discharge record and channelconditions of the South Fork Eel River reference site to

describe a rational method for constraining the magni-tude–frequency scaling parameter Ft. We next use thereference site to constrain the value of the remaining freescaling parameter, the fraction of total load in thebedload size class Fb, so that the predicted incision ratesof the calibrated saltation–abrasion model are consistentwith the field conditions.

For simplicity, we assume that the reference sitereach and upstream watershed are in approximatetopographic steady state, with bedrock incision andlandscape lowering rates equal to the local rate of rockuplift. From analysis of marine terraces, Merritts andBull (1989) mapped the variation in rock uplift rate overa 130 km transect parallel to the Northern Californiacoastline. For the South Fork Eel River reference site,located at about km 80 of the transect, Merritts andBull's (1989) data (their Fig. 5) suggest a long-termaverage rock uplift rate of 0.9 mm/year. Assuming thatbedrock incision is confined to high flows that occur16 days/year on average (Ft =0.0437), the corres-ponding instantaneous incision rate (Ei =Elt /Ft) is20 mm/year. The saltation–abrasion model predictsthis incision rate if the bedload fraction Fb is set to 0.22,for the reference site drainage area, rock strength, andchannel geometry (Table 1). The corresponding pre-dicted fraction of bed cover is 0.85, roughly consistentwith the mapped extent of bedrock exposure undercontemporary conditions.

These calculations illustrate how the saltation–abrasion model, which was derived at the scale ofindividual particle impacts can be scaled up in time andspace to apply at the scale of landscape adjustment torates of rock uplift. The scaling parameters (Ft, Fw, andFb) are constrained by the gauged discharge record andby estimates of the regional rock uplift field, to occupy alimited range of physically reasonable values. Incombination with the experimental calibration of therock strength parameter kv, the reference site calibrationdemonstrates that the saltation–abrasion model isphysically consistent with both the dynamics ofindividual bedload grain impacts on an exposed bedrockbed, and the long-term evolution of river channels inresponse to climatic and tectonic forcing.

3. Contrasts with other bedrock incision models

Although many other bedrock incision models havebeen proposed, none account fully for the effects ofsediment that emerge from a mechanistic considerationof the coupling of bedload sediment transport andbedrock wear by bedload particle impacts. In thissection we use the saltation–abrasion model as a

Page 9: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

66 L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

reference model to evaluate how other bedrock incisionmodels treat the effects of sediment, and explore therelative importance of explicitly representing each ofthese effects. Despite the importance of bedrock incisionrules for the outcomes of landscape evolution simula-tions, few studies have systematically compared thebehavior of multiple incision models. Tucker andWhipple (2002) contrasted the topographic outcomesof several incision models, some of which explicitlyrepresented sediment supply, however, none of themodels included the effects of grain size. Similarly, vander Beek and Bishop (2003) and Tomkin et al. (2003)tested the ability of various incision models to simulatethe evolution of specific river longitudinal profiles,however, the effects of grain size were also not includedin any of the models investigated.

Sediment supply and grain size introduce fourdistinct effects: the tools and cover effects, and theconstraints imposed by the thresholds of motion andsuspension. The tools effect arises because there must besome coarse sediment supplied for there to be toolsavailable to abrade the bed. The cover effect refers to theassumption that the sediment supply rate must be lessthan the sediment transport capacity for some bedrock tobe exposed in the channel bed, and thus subject toimpacts by saltating bedload. The threshold of motion,which is determined by grain size for a given shearstress, must be exceeded for abrasion by mobilesediment to occur. Finally, particle motion must bebelow the threshold of suspension, also determined bygrain size for a given flow intensity, for abrasion bybedload motion to occur. Note that the saltation–abrasion model only applies to bedload; abrasion bysuspended grains and all other incision mechanisms arenot considered here. Note, however, that the cover effectshould inhibit all bedrock detachment mechanisms, withthe possible exception of dissolution.

3.1. Model classification

Bedrock incision models can be grouped accordingto which, if any, of the effects of sediment are explicitlyrepresented. Three of the four sediment effects havebeen recognized to some extent in previous models, thetools effect, the cover effect, and the threshold ofmotion. The fourth effect, the threshold of suspension, isintroduced in the saltation–abrasion model becausebedload impacts are the only erosional mechanismconsidered. Below we review model formulations thatencompass the range of possible combinations ofsediment effects, including some that have not beenpreviously proposed, in order to clearly demonstrate the

implications of each of the sediment effects forpredictions of bedrock incision rates.

We begin by defining a generic bedrock incisionexpression that can be used as a common basis forcomparing different models. Each of the publishedbedrock incision models that we are aware of can beexpressed in a form based on Eq. (1)

Ei ¼ KggaQb

s 1� Qs

Qt

� �c

1� u4wf

� �2" #d

ð20Þ

where γ is a placeholder variable for the various modelmeasures of flow intensity (e.g. stream power), and thecoefficient Kγ is a placeholder erosional efficiencycoefficient that represents rock strength and otherfactors. The tools effect, cover effect and threshold ofsuspension are controlled by the exponents b, c, and drespectively, and the threshold of motion effect isexpressed in the choice of γ and in the choice ofsediment transport capacity relation for Qc. Table 2summarizes the parameterization of each model formu-lation in terms of the variables, coefficient andexponents of Eq. (20). Note that for each modelformulation the value of Kγ is tuned to match theassumed representative conditions and incision rate ofthe S. Fork Eel River reference site.

3.1.1. Class I: no sediment effectsThe most widely used bedrock incision models do

not explicitly account for any of the possible effects ofsediment. These include the stream power model (Seidland Dietrich, 1992)

E ¼ KXXn ð21Þ

the unit stream power model (Howard et al., 1994)

E ¼ Kxxn ð22Þ

and the shear stress model (Howard and Kerby, 1983)

E ¼ Kssn ð23Þ

where Ω=ρwgQwS is the stream power per unitdownstream length, ω=ρwgQwS /W is the streampower per unit bed area, τ=ρwgRhS is the force perunit area exerted by the flow averaged over the channelboundary, and KΩ, Kω and Kτ are proportionalityconstants intended to represent implicitly the effects ofrock resistance and many other factors, includingsediment (Howard et al., 1994; Sklar and Dietrich,1998; Whipple and Tucker, 1999). Eqs. (21), (22) and(23) are equivalent to Eq. (20) with a=n and b=c=d=0.

Page 10: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

Table 2Incision model comparison: classification, Eq. (19) model parameters and reference site parameterization

Class Equation Model γ a b c d Kγ (for reference site) Original citation

I 21 Unit StreamPower [SP]

ω 1 0 0 0 0.18 (mm/year)/(W/m2)

Seidl andDietrich, 1992

I 22 Shear Stress [SS] τ 1 0 0 0 0.41 (mm/year)/Pa Howard andKerby, 1983

I 23 Excess Shear Stress(rock) [ESS-rock]

τ−τc (rock) 1 0 0 0 0.20 (mm/year)/Pa Howard, 1994

I DimensionlessShear Stress

τ⁎ 1 0 0 0 400 mm/year Lavé andAvouac, 2000

II 24 Excess Shear Stress(sediment) [ESS-sed]

τ−τc (sed) 1 0 0 0 1.0 (mm/year)/Pa This study

II Dimensionless ExcessShear Stress

τ⁎ /τc⁎ 1 0 0 0 962 mm/year This study

II Alt DimensionlessExcess Shear Stress

τ⁎ /τc⁎−1 1 0 0 0 28.9 mm/year This study

III 26 Modified-Alluvial(n=1) [MA(1)]

ω 1 0 1 0 2.7 (mm/year)/(W/m2)

Beaumont et al.,1992

III 26 Modified-Alluvial(n=2) [MA(2)]

ω 2 0 1 0 0.0106 (mm/year)/(W2/m4)

Whipple andTucker, 2002

IV 27 Modified-Alluvial(Bedload) [MA(bdld)]

τ⁎ /τc⁎−1 1.5 0 1 0 225 (mm/year)/(kg/s)

This study

V 28 Tools [Tools] τ⁎ /τc⁎−1 −0.5 1 0 0 0.405 (mm/year)/(kg/s)

Foley, 1980;this study

VI 30 Parabolic Stream Power(n=1) [SPP(1)]

ω 0 1 1 0 0.353 (mm/year)/(W/m2)

Whipple andTucker, 2002

VI 30 Parabolic Stream Power(n=2) [SPP(2)]

ω 1 1 1 0 0.003 (mm/year)/(w2/m4)

Whipple andTucker, 2002

VII 1 Saltation–abrasion(no suspension)

τ⁎ /τc⁎−1 −0.5 1 1 0 2.70 (mm/year)/(kg/s)

Sklar andDietrich, 2004

VIII 1 Saltation–abrasion[SA]

τ⁎ /τc⁎−1 −0.5 1 1 1 2.70 (mm/year)/(kg/s)

Sklar andDietrich, 2004

67L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

Considerable debate has surrounded the choice of n andγ, with n=1 most commonly used (e.g. Tucker andSlingerland, 1994; Willett, 1999) although Whipple etal. (2000) argue for values of n ranging between 2/3 and5/3 depending on the dominant incision mechanism.

The excess shear stress model (Howard et al., 1994;Tucker and Slingerland, 1997)

E ¼ Ks�crðs� sc−rockÞn ð24Þaccounts for the possibility that incision only occurswhen shear stress exceeds a critical value τc-rock belowwhich no bedrock wear occurs. A similar model couldbe formulated with incision proportional to streampower in excess of a threshold. The threshold shearstress τc-rock is assumed to be a property of bedrockstrength, with higher values for more resistant litholo-gies. No data are available to guide parameterization ofthis threshold, although Whipple et al. (2000) discussthe threshold behavior of rock erosion by the mecha-nism of plucking. It is important to note that nodetachment threshold was observed in the bedrockabrasion mill experiments of Sklar and Dietrich (2001)

in which bedload abrasion was the dominant erosionalmechanism. A similar detachment threshold for cohe-sive hillslope materials is a key parameter in mechanis-tic theories for channel initiation (e.g. Montgomery andDietrich, 1988, 1992), and has been used in numericalsimulations of landscape evolution focused on drainagedensity (Tucker and Bras, 1998).

3.1.2. Class II: grain size effects onlyThe excess shear stress model can alternatively be

defined to incorporate the threshold of sediment motion

E ¼ Ks�csðs� sc−sedÞn ð25Þwhere τc-sed is the shear stress required to mobilize non-cohesive bed sediments and is a generally a linearfunction of grain size (Buffington and Montgomery,1997). This represents a second class of incision modelsin which grain size must be specified but the tools, coverand suspension effects are neglected. The magnitude ofτc-rock should be significantly greater than τc-sed becauseof the cohesive nature of rock; for purposes of modelcomparison we arbitrarily set τc-rock to be five times

Page 11: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

68 L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

greater than τc-sed for the reference site grain diameter of0.06 m. The sediment-based excess shear stress modelcan equivalently be expressed in non-dimensional formwith γ=τ⁎−τc⁎ or γ=τ⁎ /τc⁎−1. Like the stream powerand shear stress models, the various formulations of theexcess shear stress model are represented by Eq. (20)with a=n and b=c=d=0.

Another model that requires specification of thesediment grain size was proposed by Lave and Avouac(2001). It is a modified form of the shear stress model(Eq. (23)) in which the flow intensity variable γ=τ⁎ andn=1. Although this approach does explicitly includegrain size, it does not account for the non-linear effectintroduced by the size-dependent threshold of grainmotion.

3.1.3. Class III: sediment supply effects onlyThe simplest approach to including the effect of

sediment supply is to treat bedrock incision rate asproportional to the sediment transport capacity in excessof the supply

E ¼ KmaðQc � QsÞ ¼ KmaQcð1� Qs=QcÞ ð26Þwhere Qc= f(γ) and Kma is an erosional efficiencyconstant, as first proposed by Beaumont et al. (1992).Eq. (26) represents a third class of incision models,which capture the coverage effect but not the tools effectnor the grain size dependent effects of the thresholds ofmotion and suspension. Although the term ‘under-capacity’ is often used, we refer to models of this type as‘modified alluvial’ because they differ from alluvialincision models only in the value of the parameter Kma.If Kma=1 /ρs(1−λ)WLr, where λ is bed porosity and Lris a characteristic reach length, then Eq. (26) is simplythe conservation of mass equation for a channel with aninfinitely thick alluvial mantle. In landscape evolutionsimulations Kma is typically set to a much lower value,however, to account for the cohesive properties ofbedrock (e.g. Kooi and Beaumont, 1996). In theframework of Eq. (20), c=1, b=d=0, and a and γdepend on the choice of sediment transport relation.Two variations of the modified-alluvial model havebeen proposed, which assuming that sediment transportcapacity is proportional to stream power (i.e.Qc=Kqaω),can be written as

E ¼ Knxnð1� Qs=KqaxÞ ð27Þ

where n=1 (Beaumont et al., 1992) or n=2 (Whippleand Tucker, 2002), and a=n in Eq. (20).

Note that the modified-alluvial model is commonlymotivated by the argument that sediment supply

inhibits incision by extracting energy from the flowfor sediment transport, energy that would otherwise beavailable for incision of bedrock (Beaumont et al.,1992). However, as we argued in developing thesaltation–abrasion model (Sklar and Dietrich, 2004), itis precisely because energy is transferred from theflow to the sediment that bedrock erosion occurs, atleast for the mechanism of abrasion. The coverageeffect arises because grain interactions inhibit energytransfer from the mobile sediment to the bedrock.Although this is a subtle distinction, it may accountfor the omission of a tools effect term in landscapeevolution models using the modified-alluvial incisionrule.

3.1.4. Class IV: sediment supply and grain size effectsonly

If we use a bedload sediment transport relation in themodified alluvial model, we obtain a fourth class ofincision model, which, in addition to accounting for thecoverage effect, explicitly includes the threshold ofmotion and scales flow intensity by grain size.Combining Eqs. (4) and (26) we obtain

E ¼ Kma−b

�5:7qsW ðRbgD

3s Þ1=2ðsc4Þ3=2

� s4sc4

� 1

� �3=2

�Qs

�ð28Þ

which can be parameterized in Eq. (20) with γ=τ⁎ /τc⁎−1, a=1.5, and b=1.0.

3.1.5. Class V: grain size and tools effects onlyFoley's (1980) pioneering mechanistic incision

model represents a fifth class of possible incisionmodels, those which include the tools effect andthreshold of motion, but not the coverage effect andthreshold of suspension. This model, which appliesBitter's (1963) impact wear theory to the saltatingmotion of an individual grain without accounting forgrain interaction, can be expressed as

E ¼ Kfw2siQs

LsW: ð29Þ

Eq. (29) is equivalent to Eq. (20) with b=1 and c=d=0.Unfortunately, Foley (1980) did not specify thefunctional relationships Ls= f(γ) and wsi = f(γ), however,we can use the empirical saltation trajectory relationsdeveloped previously (Sklar and Dietrich, 2004; Eqs.(27) and (34)) to obtain γ=τ⁎ /τc⁎−1 and a=−0.5.Another version of the tools-based model was proposedby Howard et al. (1994) who used data from Head and

Page 12: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

69L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

Harr (1970) and Wiberg and Smith (1985) to parame-terize Foley's (1980) model as

E ¼ KhðQs=W ÞðQw=W Þ0:5S0:4D�0:1s ð30Þ

for τ⁎ /τc⁎>1.3; this is approximately equivalent toγ=ω and a=0.45 in Eq. (20).

3.1.6. Class VI: cover and tools effects onlyA sixth class of bedrock incision models account for

both the tools and coverage effects but neglect thethresholds of motion and suspension. For example,Whipple and Tucker (2002), responding to our earlierfindings (Sklar and Dietrich, 1998), proposed aparabolic modification of the unit stream power model(Eq. (22)) that can be written as

E ¼ Kxxn½1� 4ðQs=Qt � 0:5Þ2�

¼ Kpxn�1Qsð1� Qs=QtÞ ð31Þ

for Qc=Kqaω and Kp=4Kω /Kqa. The parabolic model ofEq. (31) is equivalent to Eq. (20) with b=c=1, d=0 anda=n−1. Whipple and Tucker (2002) discuss the caseswhere n=1 and n=2.

Fig. 3. Instantaneous bedrock erosion rate as a function of sedimentsupply for four classes of incision models that each account fordifferent sets of effects of sediment, calculated holding all othervariables constant at South Fork Eel River references site conditions.Open circle indicates the instantaneous incision rate that correspondsto the assumed long term bedrock erosion rate at the reference site, forFt =0.0437.

3.1.7. Class VII: tools, cover and threshold of motiononly

A seventh class of incision models account for thethreshold of motion along with the tools and covereffects, neglecting only the effect of the threshold ofsuspension. This is what we would obtain with thesaltation–abrasion model of Eq. (1) if we omit thesaltation length extrapolation term [1− (u⁎ /wf)

2], so thata=−0.5, b=c=1 and d=0 in Eq. (20). Note that this isequivalent to the linear parabolic model (Eq. (31) withn=1) if γ=τ⁎ /τc⁎−1 rather than ω (i.e. if a bedloadsediment transport relation is used).

3.1.8. Class VIII: tools, cover, threshold of motion andthreshold of suspension

Finally, the full saltation–abrasion model (Eq. (1))constitutes an eighth model class, in which each of thefour principal effects of sediment are explicitlyaccounted for.

3.2. Model comparisons

In this section we compare the incision ratespredicted by the various models defined above, forvariations in sediment supply rate and transport stage(τ⁎ /τc⁎). As noted above, the coefficients of each of themodels are tuned to predict the assumed long-term

incision rate at the South Fork Eel reference site. Thiscreates a common benchmark erosion rate, so that wecan compare predicted deviations from that erosion rateas we vary sediment supply and transport stage.

Fig. 3 shows incision rate as a function of sedimentsupply, holding all other variables, such as shear stress,grain size, and rock strength, constant at reference sitevalues. The reference site is indicated by the open circle.Four types of behavior can be seen. In the trivial case,models with no sediment supply effects (Classes I andII) plot as a horizontal line because predicted incisionrate is independent of sediment supply. Models thataccount for the cover effect, but not the tools effect(Classes III and IV) predict a linear decrease in incisionrate with increasing sediment supply, becoming zerowhen the sediment supply equals the transport capacity;maximum incision rates occur when sediment supply isminimized. Conversely, tools-based models (Class V)predict a linear increase in incision rate with increasingsediment supply; incision rate is maximized whensediment supply is greatest. The saltation–abrasionmodel, and the other models that account for both thecover and tools effects (Classes VI, VII and VIII) predicta parabolic dependence of incision rate on sedimentsupply, in which the peak incision rate occurs at amoderate sediment supply rate.

To compare model predictions of incision rate as afunction of transport stage, it is convenient to use athree-dimensional depiction of the incision model

Page 13: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

70 L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

functional surface, as we have done previously (Sklarand Dietrich, 2004). As shown in Fig. 4, each incisionmodel can be represented fully by plotting contours ofconstant incision rate (axis out of the page) as functionsof transport stage and the ratio of sediment supply tosediment transport capacity (Qs/Qc; also referred to asrelative sediment supply). Note that the contour spacingis logarithmic to encompass incision rates ranging overmany orders of magnitude. Note also that transport stage

Fig. 4. Bedrock incision rate (mm/year) predicted by various models (contoustage and relative sediment supply (Qs/Qc), for reference site values of grain

occurs on the relative sediment supply axis as well,because sediment transport capacity is a non-linearfunction of transport stage (Eq. (4)).

Depicting the incision functions in this format allowsthe principal effects of sediment to occur along discreteaxes. The thresholds of motion and suspension areexpressed along the transport stage axis, while the coverand tools effect are expressed along the relative supplyaxis. As we have shown previously (Sklar and Dietrich,

rs, axis out of the page, logarithmic spacing) as a function of transportsize (Ds=0.06 m) and rock tensile strength (σT=7.0 MPa).

Page 14: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

Fig. 5. Bedrock erosion rate as a function of transport stage for variousincision models, for constant relative sediment supply (Qs/Qc=0.5).Model abbreviations listed in Table 2.

71L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

2004), when incision rate is non-dimensionalized byrock strength and grain size, the saltation–abrasionmodel plots in this parameter space as a unique surface,valid for all physically reasonable combinations ofdischarge, slope, width, sediment supply, grain size, rockstrength and channel roughness. For model comparisonpurposes we have used the dimensional instantaneousincision rate Ei (mm/year), with model parameters tunedto match the reference site incision rate (Table 1).

The fundamental differences between the incisionmodel classes are clearly apparent in Fig. 4. Models withno sediment effects plot as simple inclined planes (Fig.4a and b), with the sensitivity to shear stress expressedby the density of contour lines. Models that require ashear stress in excess of a threshold for incision to occur,either a rock strength-dependent detachment threshold(Fig. 4c) or the threshold of motion (Fig. 4d, g, j, k andl), have a zone of ‘no erosion’ for stresses below thethreshold. Where the cover effect is accounted for (Fig.4e, f, g, i, j, k, l), incision rate contours curve to the rightas the threshold of alluviation is approached (Qs/Qc∼1)and incision rate goes to zero. Similarly, where the toolsaffect is accounted for (Fig. 4h, i, j, k and l), incision ratecontours curve to the right as sediment supply, and thusthe incision rate, approach zero. Because the saltation–abrasion model also accounts for the threshold ofsuspension, there is also a zone of ‘no erosion’ for shearstresses above this threshold (Fig. 4l). Due to thesuspension effect, the saltation–abrasion model predictsthat there is a limit to the potential rate of bedrock wearby bedload abrasion, for all possible values of transportstage and sediment supply, as indicated by the peak ofthe incision function surface (Fig. 4l). All other modelsconsidered here, which in principle are intended torepresent incision by multiple erosional mechanisms,predict continuously increasing incision rates withincreasing transport stage.

Comparison of Figs. 3 and 4 shows that the curves ofincision rate as a function of sediment supply, plotted inFig. 3, are simply slices through the functional surfacesplotted in Fig. 4, parallel to the relative sediment supplyaxis. A similar slice through the functional surfaces canbe made parallel to the transport stage axis, in order todepict the dependence of incision rate on transport stage,independent of the influence of transport stage on thesediment transport capacity. Fig. 5 shows slices througheach of the functional surfaces of Fig. 4, along a line ofconstant relative sediment supply (Qs/Qc=0.5).

Three fundamental differences between incisionmodel classes are clearly illustrated in Fig. 5. First,models that do not account for the threshold of motionpredict significant rates of incision when the flow

intensity is too weak to move bedload. These include thestream power [SP, SP(2)], shear stress (SS), modified-alluvial [MA, MA(2)], and parabolic stream power[SPP, SPP(2)] models. Second, with the exception ofsaltation–abrasion [SA], all models neglect the thresh-old of suspension, and thus predict extremely rapid ratesof incision when sediment transport is dominated bysuspended load (τ⁎ /τc⁎>30), despite the probability thatsuspended grains collide very infrequently with thechannel bed. Third, most models tend to over-predict therate of growth of incision rate, relative to the saltation–abrasion model, even at moderate levels of transportstage. This is particularly true of the non-linear (n=2)versions of the modified alluvial and parabolic streampower models. The bedload version of the modifiedalluvial model [MA(bdld)] also predicts very rapidgrowth in incision rate as transport stage increasesbeyond the threshold of motion. Note that the excessshear stress model with a rock strength-dependentdetachment threshold [ESS(R)] plots apart from theother models in Fig. 5 because we have set the thresholdshear stress to a value five times larger than thethreshold of sediment motion shear stress, to account forthe cohesive nature of rock.

Each incision model considered here can be tuned tobe approximately equivalent to the saltation–abrasionmodel, over a limited range of variation in sedimentsupply and transport stage. For example, as shown in Fig.

Page 15: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

72 L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

3, the modified alluvial model predicts a similar variationincision rate for relatively high sediment supply rates andfixed transport stage, but vastly over predicts incisionrate for low sediment supply rates, compared to thesaltation–abrasion model. Similarly, each of the modelsthat account for the threshold of motion predict similarincision rates for relatively low transport stages (forconstant Qs/Qc; Fig. 5), but diverge rapidly as τ⁎ /τc⁎increases above ∼2.0. The bedload version of themodified alluvial model perhaps best approximates thesaltation–abrasion model, but only for the narrow rangeof low transport stage (τ⁎ /τc⁎<3) and high relativesediment supply (Qs/Qc>0.8). These conditions may,however, be relatively common in many mountain riverswhen bedrock incision is taking place.

The fundamentally different model behaviors illus-trated in Fig. 4 can be expected to correspond to largedifferences in predicted landscape response to changesboundary conditions, such as rock uplift rate, rockstrength and discharge regime. In the following sectionwe explore the landscape implications of including orneglecting the various effects of sediment in bedrockincision models, by considering the controls on channelslope for the case of topographic steady state.

4. Controls on steady-state channel slope

The complex role of sediment in influencing bedrockincision rates has important implications for understand-ing the controls on the slope of a given reach of riverchannel. And because river longitudinal profile relief andconcavity are directly related to the integral andderivative of channel slope along the length of the profilerespectively, understanding the controls on local channelslope is essential to explaining river profile form ingeneral. For simplicity, we focus here on steady-statechannel slopes. Landscapes subject to prolonged steadyrock uplift may approach an equilibrium condition whereerosion rate everywhere balances rock uplift rate andtopography reaches a steady state (Hack, 1960). Atsteady state, erosion rate is determined simply by the rateof rock uplift, and channel slope becomes the dependentvariable. The equilibrium slope of an incising river is thena function of the rate of rock uplift, as well as the othervariables influencing the instantaneous incision rate, inparticular sediment supply, grain size, discharge, androck strength. In this section we explore the implicationsof the cover and tools effects and the thresholds of motionand suspension in controlling channel slope at a reachscale, and contrast the predictions of the saltation–abrasion model with the other types of models used inlandscape evolution simulations.

4.1. Components of steady-state channel slope

Recognition of the constraints on bedrock incisionrates imposed by the threshold of grain motion and thethreshold of alluviation, suggests that a hierarchy ofchannel slopes exists. The minimum slope required forbedrock incision to occur is set by the threshold ofmotion. For a given discharge and channel geometry, ifslope is not steep enough to generate shear stresses inexcess of the threshold of motion, no sedimenttransport and no incision occur. If the slope is steepenough to mobilize bed sediments, but not steepenough to transport bedload out of the reach at the ratethat coarse sediments are supplied to the reach fromupstream and from adjacent hillslopes, then we assumethat no bedrock will be exposed in the channel bed,incision rate will be zero, and bed aggradation willoccur. For incision to occur, we assume that the slopemust be steeper than the slope required to transport theimposed sediment load, in order to provide the excesstransport capacity to generate sufficient bedrockexposure and bedload particle impact energy andfrequency.

Therefore, steady-state total channel slope (S) can beconsidered to be a sum of component terms

S ¼ SDs þ DSQs þ DSE ð32Þwhere SDs

is the threshold of motion slope, ΔSQsis the

increment of slope above SDsrequired to transport

sediment at the rate of supply, and ΔSE is an additionalslope increment required to erode bedrock at the rate ofrock uplift. Using the bedload sediment transportrelation of Eq. (4) (Fernandez-Luque and van Beek,1976), we obtain

SD ¼ sc4RbDs=Rh ð33Þand

DSQs ¼RbDs

Rh

Qs

5:7qsW ðRbgD3s Þ1=2

!2=3

ð34Þ

For the saltation abrasion model, because slope occursin both the flow intensity term and the cover term(Eq. (20); Table 2), ΔSE cannot be solved foranalytically, and instead must be solved for numeri-cally. Below, in comparing the steady-state channelslopes predicted by the various classes of incisionmodels, we will use the slope component frameworkto help understand the behavior of the saltation–abrasion model, and the deviations from this behaviorof the other models.

Page 16: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

73L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

4.2. Analytical solutions for steady-state channelslope

Analytical solutions for steady-state channel slopecan be derived for the other bedrock incision modelsdiscussed in the preceding section. In order toevaluate the predicted influence of rock strength onsteady-state channel slope (below), we modify theincision efficiency coefficients of each model expres-sion to explicitly include the square of rock tensilestrength (σT

2) as a proxy for rock resistance to wear.We have demonstrated previously (Sklar and Dietrich,2001) that rock resistance to wear by bedloadabrasion scales with σT

2. Here we assume forsimplicity that the same scaling relationship appliesto rock resistance to wear by all incisional mechan-isms represented by the other model expressions.Using the unit stream power model as an example(n=1), we now write Eq. (22) as

E ¼ KxrqwgQwS=r2TW ð35Þ

where Kωσ=Kω/σT2.

Rearranging Eq. (35) to solve for slope we obtain

S ¼ Er2TW=KxrqwgQw: ð36ÞSimilarly, for the shear stress model (Eq. (23))

S ¼ Er2TKsrqwg

� �10=7WQwn

� �6=7

ð37Þ

the excess shear stress model, with a rock detachmentthreshold (Eq. (24)

S ¼ Er2TKs�crr

þ sc−rock

� �W 3=5

qg Qwnð Þ3=5" #10=7

ð38Þ

the excess shear stress model, with a sediment motionthreshold (Eq. (25))

S ¼ Er2TKscsr

þ sc−sed

� �W 3=5

qwg Qwnð Þ3=5" #10=7

ð39Þ

the modified-alluvial model (Eq. (27)), with n=1

S ¼ Er2TKar

þ Qs

� �W

KblqwgQwð40Þ

and with n=2

S ¼ QsW2KblqwgQw

1þ 1þ Er4T4K2bl

Q2sKa2r

� �1=2" #

ð41Þ

the bedload modified alluvial model (Eq. (28))

S ¼ f Er2T þ QsKma−br

5:7Kma− brqsW RbgD3s

� �1:2ðsc4Þ3=2 !2=3

þ1

24

35

� RbDssc4QwnW

� �3=5g10=7 ð42Þ

the tools model (Eq. (29) parameterized as Eq. (20) withb=1 and c=d=0)

S ¼ KfrQs

Er2T

� �2

þ1

" #RbDssc4QwnW

� �3=58<:

9=;

10=7

ð43Þ

the parabolic stream power model (Eq. (31)), with n=1

S ¼ QsW

KblqwgQwð1� Er2TKbl=4KxrQsÞ ð44Þ

and with n=2

S ¼ WqwgQw

Kblr2TE4Kx2rQs

þ Qs

Kbl

� �ð45Þ

4.3. Comparisons of incision model predictions ofsteady-state channel slope

In this section we compare the steady-state channelslope predicted by the saltation–abrasion model with theslopes predicted by the other incision models discussedabove. Our analysis focuses on the sensitivity of steady-state slope to variations in rock uplift rate and rockresistance to erosion, in part because the two mostcommon incision models in landscape evolution studies,the stream power and shear stress models, predict linearvariation in slope with both variables. As we demon-strate below, the saltation–abrasion model predicts ahighly non-linear relationship between slope and rockuplift rate and very little dependence of slope on rockstrength. To the extent that other models incorporate thevarious effects of sediment on incision rate, they toodeviate from the linear relationships of the stream powerand shear stress models.

We also evaluate the sensitivity of slope to themagnitude–frequency parameter Ft, which, as definedabove, represents the fraction of time during which thehigh flows that transport sediment and incise bedrockoccur. Note that because topographic steady-state occurswhen rock uplift rate is equal to the long-term incisionrate Elt, the absolute values of rock uplift rate and

Page 17: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

Fig. 6. Influence of variable rock uplift rate on: (a) steady-state channelslope for various models; (b) fraction of bed covered for models thataccount for cover effect; and (c) fraction of total slope due to the slopecomponents responsible for grain motion (SDs

), bedload sedimenttransport (ΔSQs

) and bedrock wear (ΔSE) as predicted by the saltation–abrasion model. Reference site indicated by open circle. Maximumpossible incision rate for saltation–abrasion model, for reference siteconditions, indicated by asterisk.

74 L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

steady-state instantaneous erosion rate differ by a factorof Ft. As before, to help ground the analysis in the rangeof physically reasonable values, we hold all othervariables constant at South Fork Eel River reference siteconditions (Table 1). Note also that we consider 11 ofthe 12 incision models discussed in the previous section;for clarity, we have omitted the version of the saltation–abrasion model that lacks the threshold of suspensionterm.

4.3.1. The influence rock uplift rate on steady-statechannel slope

Fig. 6a shows the predicted steady-state channelslope as a function of rock uplift rate for the variousincision models. As before, the reference site is indicatedby an open circle. Over a range of rock uplift ratespanning more than four orders of magnitude, the slopepredicted by the saltation–abrasion model varies by onlyslightly more than one order of magnitude. As rockuplift rate declines from the reference site value, thesteady-state saltation–abrasion model slope declinesslowly until reaching a nearly constant value, set by SDs

,the threshold of motion slope. As rock uplift rateincreases from the reference site value, the predictedslope increases rapidly, until reaching a maximum,denoted by an asterisk. The saltation–abrasion modelpredicts that steady-state cannot be achieved for rates ofrock uplift any more rapid, given the reference sitevalues of grain size, rock strength, and other variablesheld constant. This maximum possible incision rateoccurs because of the tradeoff between the increase inparticle impact energy with the decrease in impactfrequency, with increasing transport stage. As describedin detail in Sklar and Dietrich (2004), for sufficientlyhigh shear stress, the derivative of the saltation–abrasionincision function with transport stage becomes negative.Thus, further increases in slope beyond the peak valuecause incision rate to decline, rather than increase.

The relatively small change in steady-state channelslope predicted by the saltation–abrasion model over avery large range of rock uplift occurs in part because ofthe adjustment in the fraction of the bed exposed toerosive wear. This is shown in Fig. 6b, where thefraction of the bed covered, assumed to correspond tothe relative sediment supply (Qs/Qc), is plotted againstrock uplift rate. For low rock uplift rates, only smallchanges in bed cover are required to accommodate largechanges in the steady-state incision rate. As rock upliftrate becomes relatively large, and begins to approach themaximum possible steady-state value, the extent ofbedrock exposed on the channel bed increases morerapidly.

The non-linear increase in steady-state channel slopewith increasing rock uplift rate is best understood byconsidering the relative contributions of each of theslope components defined above to the total channelslope. Fig. 6c shows each of the slope components,normalized by the total channel slope, for the range ofrock uplift rates. The threshold of motion componentdominates the total slope for low values of rock upliftrate, but declines in importance as rock uplift becomesrelatively rapid. Because increases in steady-state rock

Page 18: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

75L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

uplift rate result in accelerated rates of erosion andsediment production in the upstream landscape, sedi-ment supply increases linearly with rock uplift rate (for aconstant bedload fraction of the total load Fb). Thiseffect influences the increase in channel slope above SDs

much more significantly than does the requirement thatthe channel incise at the rate of rock uplift. For relativelyhigh rock uplift rates, ΔSQs

is the dominant control onchannel slope. The slope increment responsible forbedrock incision ΔSE grows rapidly as the peak erosionrate is approached, but is never the dominant slopecomponent term. Overall, the saltation–abrasion modelpredicts that the effects of grain size and sedimentsupply primarily control steady-state channel slope,over the full range of rock uplift rates for which steady-state can be achieved by bedload abrasion.

The behavior of the saltation–abrasion model isfundamentally different from the stream power andshear stress models, as well as the linear forms (n=1) ofthe modified-alluvial and parabolic stream powerincision models. Each of these models predicts a rapidand continuous decline in slope with reduced uplift rate,without any limit to how low the slope can become (Fig.6a). Orders of magnitude variation in uplift rate areaccommodated by orders of magnitude change inchannel slope, and thus predicted topographic relief.This power law behavior is expected from the streampower and shear stress models. In contrast, the modifiedalluvial and parabolic stream power models explicitlyinclude the cover effect and thus might be expected topredict changes in the extent of bed cover with changingerosion rate and sediment supply. However, as shown inFig. 6b, the linear form of these models predicts aconstant extent of bed cover because the reduction insediment supply with reduced rock uplift rate is exactlymatched by a reduction in sediment transport capacity.The bedload version of the modified alluvial model alsopredicts constant bed exposure despite large changes inrock uplift rate and sediment supply (Fig. 6a), however,this model closely follows the slope variation predictedby the saltation–abrasion model (Fig. 6a) because, asdefined above, they share the same form of sedimenttransport relation (Eq. (4)).

The non-linear (n=2) forms of the modified-alluvialand parabolic stream power models do predict changesin bed cover with changing uplift rate, however theeffect is opposite to that predicted by the saltation–abrasion model. As shown in Fig. 6b, these modelspredict that the fraction of the bed covered becomesincreasingly small as uplift rate is reduced, approachingbare bedrock as erosion rate becomes negligibly small.This occurs in the case of the non-linear modified

alluvial model because steady-state slope depends onerosion rate to a power less than unity in Eq. (41) (notethat Qs

2 appears in the denominator of the last term), sotransport capacity declines more slowly than sedimentsupply as uplift rate is reduced. In the case of the non-linear parabolic stream power model, as uplift rate isreduced the predicted steady-state slope declines to aconstant value, much like the saltation–abrasion model.However, unlike the threshold of motion slope, thisconstant slope value has no physical meaning, but ismerely a mathematical artifact. At steady-state, erosionand sediment supply cancel in the first term within theparentheses of Eq. (45), and the resulting constant termis independent of grain size, sediment supply rate androck uplift rate. Because slope remains large, sedimenttransport capacity remains high as uplift rate andsediment supply decline, and the fraction of the bedcovered goes to zero.

The two versions of the excess shear stress model,with the critical shear stress term set alternatively by arock detachment threshold or the threshold of sedimentmotion, both predict a non-linear variation in channelslope similar to the saltation–abrasion model. Thesediment mobilization version closely tracks the salta-tion–abrasion model for low rock uplift rates wheregrain size is the dominant control on channel slope, butpredicts a more rapid increase in slope for relatively highuplift rates as it approaches the power law behavior ofthe simple shear stress model. The rock detachmentversion of the excess shear stress model is exactlyparallel, only shifted up by a factor set by the differencein the threshold value. Finally, the tools model predictsno dependence of steady-state channel slope on rockuplift rate. This surprising result occurs because theincrease in the incision rate required to maintain steadystate with increasing rock uplift rate is exactly balancedby the increase in tools provided by the increasedsediment supply.

4.3.2. The influence of rock strength on steady-statechannel slope

Fig. 7a shows the variation in predicted steady-statechannel slope as a function of rock tensile strength forthe various incision models. In this calculation, rockstrength varies by a factor of 20, which corresponds to a400-fold difference in rock resistance to abrasive wear(Sklar and Dietrich, 2001). For this range of rockresistance, which covers most competent lithologiesencountered in tectonically active landscapes, thesaltation–abrasion model predicts only subtle changesin steady-state channel slope. Unlike rock uplift rate,variations in rock strength do not change the steady-

Page 19: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

Fig. 7. Influence of variable rock tensile strength on: (a) steady-statechannel slope for various models; (b) fraction of bed covered formodels that account for cover effect; and (c) fraction of total slope dueto the slope components responsible for grain motion (SDs

), bedloadsediment transport (ΔSQs

) and bedrock wear (ΔSE) as predicted by thesaltation–abrasion model. Reference site indicated by open circle.Maximum possible incision rate for saltation–abrasion model, forreference site conditions, indicated by asterisk.

76 L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

state sediment supply rate, and thus only affect slope byrequiring changes in the efficiency of the incisionprocess. As shown in Fig. 7b, the efficiency of thebedload abrasion mechanism is increased in response tomore resistant rock by increases in the extent of bedrockexposed in the channel bed. Because sediment supplyand grain size are constant, only small increments inslope are required to reduce bed cover substantially. Forthe reference site values of grain size and sedimentsupply, the slope components due to these variables are

coincidentally equal (Fig. 7c), and dominate the totalslope except for the highly resistant rocks. Thus, forweak and moderate strength rocks, the channel slope isessentially independent of rock strength and is con-trolled by the requirement to mobilize and transport thesupplied coarse sediment load. However, at the high endof the range of rock tensile strengths, the saltation–abrasion model predicts that bedload abrasion cannoterode efficiently enough to match the rate of rock uplift(for reference site conditions). Increases in channelslope above the peak value denoted by the asterisk inFig. 7a would increase the extent of bedrock exposure,but not enough to compensate for the accompanyingreduction in the frequency of particle impacts.

In contrast, models that do not include sedimenteffects, such as the stream power and shear stressmodels predict a simple power law dependence ofsteady-state channel slope on rock tensile strength (Fig.7a). Models that include the cover effect, such as thelinear and non-linear versions of the modified alluvialand parabolic stream power models, predict a narrowrange of slopes over a wide range of rock strengths,because the constant sediment supply rate imposes aminimum slope which must be exceeded for incision tooccur. As with the saltation–abrasion model, thereduced erosional efficiency as rock strength increasesis offset by a corresponding increase in the fraction ofthe bed exposed (Fig. 7b), which is accomplished withsmall changes in channel slope. Counter-intuitively, thetools model predicts a decrease in slope with increasingrock strength. This occurs because for this modelerosional efficiency is increased by increasing thefrequency of particle impacts, which in the absence ofa change in sediment supply rate, can only beaccomplished by a reduction in shear stress, whichreduces the saltation hop length.

4.3.3. The influence of the magnitude–frequency trade-off on steady-state channel slopes

In this calculation we vary the magnitude–frequen-cy parameter Ft to evaluate how the steady-statechannel slopes predicted by the various models varywith storminess. As described in detail in Section 2.2,above, when Ft =1.0, discharge and coarse sedimentoccur continuously at the mean annual rates of waterand sediment delivery from the upstream watershed.As Ft decreases, the magnitudes of both the high flowdischarge and coarse sediment supply increase, whilethe duration decreases. Because in this approachdischarge is partitioned into high and low flowcomponents, but all coarse sediment is assumed tobe transport only during the high flows, sediment

Page 20: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

77L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

supply grows more rapidly than discharge as high-flowevents become increasingly infrequent (Ft≪1.0). Asbefore, all variables not affected by changes in Ft areheld constant at South Fork Eel River reference sitevalues.

Fig. 8a shows the variation in channel slope as afunction of the high flow time fraction Ft for the variousincision models. The slope predicted by the saltation–abrasion model is highest at the two extremes of the

Fig. 8. Influence of variable high flow time fraction parameter on: (a)steady-state channel slope for various models; (b) fraction of bedcovered for models that account for cover effect; and (c) fraction oftotal slope due to the slope components responsible for grain motion(SDs

), bedload sediment transport (ΔSQs) and bedrock wear (ΔSE) as

predicted by the saltation–abrasion model. Reference site indicated byopen circle. Maximum possible incision rate for linear parabolicstream power model, for reference site conditions, indicated byasterisk.

range of Ft, and reaches a minimum at the reference sitevalue of Ft. The range of predicted slopes varies byabout a factor of two. The slope is steepest for Ft =1.0because high flow discharge, and thus flow depth, arelowest, and a high slope value is needed to exceed thethreshold of motion. As shown in Fig. 8b and c, atFt =1.0 the bed is nearly completely covered, and thetotal slope is dominated by the threshold of motioncomponent SDs

. Because the instantaneous erosion rateEi =Elt /Ft (=Ur /Ft for steady-state), Ei is minimized atFt =1.0 and the slope increment due to incision ΔSE isalso minimized.

The saltation–abrasion slope is also relatively steepfor Ft≪1.0. For this highly ‘stormy’ representation ofthe distribution of discharges, the low frequency of highflow events results in very high coarse sediment supplyand a high instantaneous incision rate. In order totransport this high load, the slope must be considerablysteeper than the threshold of motion slope, and to inciseat this rapid rate, exposed bedrock must occupy a largefraction of the channel bed (Fig. 8b). For these reasons,the total slope is dominated by ΔSQs

and ΔSE (Fig. 8c).The predicted slope reaches a minimum at the referencesite value of Ft =0.0437 for the same reason that thisvalue was selected in Section 2.2, because the product ofdischarge magnitude and duration is maximum at thisvalue, resulting in the maximum net annual sedimenttransport capacity for a given slope. (Note that in theprevious calculation slope was held constant, here slopeis allowed to vary while long-term incision rate is heldconstant.)

Because the stream power and shear stress modelsdo not account for the effects of sediment, thosemodels predict a simple continuous increase in steady-state channel slope with increasing high flow timefraction (Ft). This results from a tradeoff betweenslope and discharge; as high flow discharge declineswith increasing Ft, slope increases to compensate. Therate of slope increase is tempered by the accompa-nying decline in instantaneous incision rate withincreasing Ft. The excess shear stress models followthe same pattern as the shear stress model, with therock detachment version offset again by the assumeddifference in threshold value. Most of the models thataccount for one or more of the effects of sediment onbedrock incision generally follow the pattern of slopevariation predicted by the saltation–abrasion model,with slope maxima at either end of the range of Ft witha minimum at a moderate value of Ft. However, aswith the previous calculation of steady-state slope as afunction of rock uplift rate, the pattern of bed coveragepredicted by the various versions of the modified

Page 21: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

78 L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

alluvial and parabolic stream power models is oppositethat predicted by the saltation–abrasion model. Asshown in Fig. 8b, those models predict large extents ofbedrock cover during high magnitude, low frequencyevents (Ft≪1), and little alluvial cover for lowmagnitude, high frequency events (Ft∼1). The linearparabolic stream power model cannot maintain steadystate for low bed coverage; the maximum possiblesteady-state value of Ft is indicated with an asterisk.This occurs due to a reversal in the sign of thederivative of the incision function with respect to flowintensity, as the incision function moves down into thetools-deprived region of the parabolic sediment supplyeffect (Fig. 3). Finally, the tools-based model shows anexaggerated version of the slope variation pattern ofmost of the other sediment-influenced models, but forthe opposite reason. The slope predicted by the toolsmodel is very high for Ft≪1 in order to offset thelarge increase in the frequency of bedload impacts, andconsequent increase in incisional efficiency, caused bythe high sediment supply rate.

4.3.4. Non-dimensional representation of saltation–abrasion model slope variation

An additional way of understanding the pattern ofvariation in steady-state channel slope predicted by thesaltation–abrasion model in the three preceding calcula-tions, is by plotting the associated path across the non-dimensional representation of the incision functionsurface. This is shown in Fig. 9, where, unlike thecontour plots of Fig. 4, we plot contours of incision rate

Fig. 9. Non-dimensional bedrock erosion rate (E⁎=EiσT2 /ρsY(gDs)

3/2×10−15; contours, axis out of the page) as a function of transport stageand relative sediment supply, showing plotting position path for thecalculation of steady-state channel slope predicted by the saltation–abrasion model shown in Figs. 6–8.

non-dimensionalized by grain diameter and rock tensilestrength

E4 ¼ Er2TqsY ðgDsÞ3=2

ð46Þ

as derived in Sklar and Dietrich (2004). The South ForkEel River reference site is again indicated with an opencircle. With this concise representation of the saltation–abrasion incision function, the calculated variation inslope with rock uplift rate and with the magnitude–frequency parameter Ft traverse the identical path. Forslow rock uplift rates (Ur), or large high flow timefractions (Ft), the steady-state channel reach is justabove the threshold of motion (S∼SD; Figs. 6c and 8c),with nearly complete bed coverage (Qs/Qc∼1; Figs. 6band 8b), and relatively low instantaneous incision rateΔSE≪S (Figs. 6c and 8c). As Ur increases, or Ft

decreases, the channel plotting position in Fig. 9 climbstoward the peak of the functional surface, reflecting theincrease in the instantaneous incision rate (Ei). At themaximum possible incision rate (asterisks), it is evidentthat extending the curve further by increasing transportstage would cause the plotting position to descend downthe flank of the functional surface, and result in areduced incision rate and a failure to maintain steady-state.

For the calculation of slope as a function of rockstrength, the plotting position in Fig. 9 for weak rocks isQs/Qc∼1 (Fig. 7b), but in this case τ⁎ /τc⁎>1 becausesediment supply is constant, and SDs

∼ΔSQs(Fig. 7c).

As rock strength increases, the plotting position climbsto higher values of E⁎, following a path of moresubdued change in τ⁎ /τc⁎ that reflects the small changein predicted steady-state slope (Fig. 7a). Fig. 9 showsagain that the peak possible steady-state slope occurs asthe plotting position path is about to begin descendingdown the low relative supply flank of the functionalsurface. The difference in the plotting position paths,between the S= f(Ur) or S= f(Ft) cases and the S= f(σT)case, is due to the fact that sediment supply varies withUr and Ft but not with σT. The increase in sedimentsupply pushes the path for the Ur and Ft cases to a muchhigher transport stage to achieve the approximatelysame extent of bed exposure at the peak slope, than isrequired in the σT case where sediment supply isconstant.

5. Discussion

The preceding analysis suggests that there is ahierarchy of factors controlling channel slope in

Page 22: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

79L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

tectonically active landscapes that are in approximatelytopographic steady-state. The grain size of the coarsefraction of the total load is the dominant factor over awide range of conditions, including weak to moderatestrength rocks, low to moderate rock uplift rates, andrelatively uniform to moderately ‘stormy’ dischargedistributions. The supply rate of coarse sediments is thesecond most important factor in setting steady-stateslopes, while bedrock wear itself is only a significantcontrol for very resistant rocks and extremely rapid ratesof rock uplift.

Given the importance of grain size in setting channelslopes, an obvious weakness of the saltation–abrasionmodel, as applied here, is the use of a single graindiameter to represent the effect of the wide range ofgrain sizes that may move as bedload through a reach ofan actively incising channel. Explicitly accounting forgrain size distributions would require use of bedloadsediment transport relations that predict sediment fluxfor each grain size class (e.g. Parker, 1990). Also neededare new observations and insights into the role of mixedgrain sizes in controlling the thresholds of alluviationand the relationship between sediment supply, transportcapacity, and the partial coverage of bedrock. Prelim-inary wear rate measurements (unpublished) usingmixed grain sizes in a bedrock abrasion mill suggest,however, that for relatively low sediment supply rates,the presence of a grain mixture has little effect. Totalwear rates are similar to what would be obtained by amass-weighted sum using data from uniform grain sizeruns.

Assessment of the morphologic implications of thesaltation–abrasion model is limited by our lack ofunderstanding of the controls on the grain sizedistribution of sediment supplied to the channel networkby hillslopes. Here we have treated grain size asindependent of all other variables, however, it is likelythat grain size will covary with rock uplift rate, rockstrength, and other variables such as discharge. Steady-state landscapes underlain by more resistant rocks,eroding more rapidly, or in more arid climates, are likelyto produce and deliver to channels relatively coarsergrain size distributions. Not only will grain size beaffected, but the fraction of total load in the bedload sizeclass Fb should be larger as well.

Further research is also needed to improve ourunderstanding of the effects of episodic sediment supplyon the extent of partial alluviation in bedrock channels.Although we have assumed here that stochastic varia-tions in sediment supply can be effectively averagedover time, the bed condition of actively eroding riverssuch as the South Fork Eel may be highly variable over

an annual to decadal time scale (Howard, 1998). Thesensitivity of the extent of alluviation to episodicsediment supply may be enhanced by non-linear effectsin the relationship between relative sediment supply andthe percentage of bed cover. As discussed in more detailin the development of the saltation–abrasion model(Sklar and Dietrich, 2004), there may be a thresholdsediment supply required before any transient depositsof alluvium can form, and a threshold extent of bedcover above which a run away alluviation occurs, due tochanges in bed roughness and bedload sedimenttransport efficiency.

The hierarchy of slope controlling factors identifiedby the saltation–abrasion model provides a usefulframework for evaluating the utility of the other incisionmodels discussed above. The most important sedimenteffect is the threshold of motion, which surprisingly isnot explicitly included in any other published models,although versions of the excess shear stress model havebeen used without specifically linking the value of τc tosediment grain size (Tucker and Slingerland, 1997;Tucker and Bras, 1998). Treating τc as a rock strengthdependent parameter potentially has the same effect insetting a minimum slope (Fig. 6a), however we know ofno direct evidence for the existence of a generaldetachment threshold. As we noted previously (Sklarand Dietrich, 2004), bedrock abrasion mill experimentsfailed to detect a detachment threshold for the erosionalmechanism of abrasion by sediment impacts.

The influence of sediment supply on the extent ofbedrock exposure in the channel bed is the second-mostimportant effect that consideration of sediment intro-duces to the bedrock incision problem. As we haveargued previously (Sklar and Dietrich, 2001), partial bedcoverage introduces an important degree of freedom forchannel adjustment. As illustrated above (Figs. 6b, 7band 8b), small changes in channel slope can result inlarge changes in the efficiency of bedrock incision, dueto shifts in the percentage of bed alluviation. Interest-ingly, models that are intended to represent the covereffect, such as the modified alluvial and parabolic streampower models, can predict no adjustment in partial bedcoverage with variable rock uplift rate, and can evenpredict patterns of bed cover adjustment opposite tothose predicted by saltation–abrasion model (Fig. 6b).

The case of the non-linear parabolic stream powermodel illustrates a potential pitfall in developinggeomorphic process models without a strong mechanis-tic basis. That model predicts an apparent thresholdslope for low uplift rates, which, as described above, hasno physical meaning. The non-linear version (n=2) wasproposed to eliminate the computationally troublesome

Page 23: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

80 L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

upper limit to the range of possible erosion ratespredicted by the linear version (n=1) of the parabolicstream power model (Fig. 8b). Unfortunately, choosingparameter values on the basis of mathematical conve-nience can lead to profoundly counter-intuitive modelbehavior as well as reasonable yet wholly artifactualresults.

The tools effect is third in order of importance amongsediment effects on bedrock incision rates. For example,the tools-based model, which includes the tools effect aswell as the threshold of motion, predicts stronglycounter-intuitive and highly divergent dependencies ofslope on rock strength and uplift rate. On the other hand,the bedload version of the modified alluvial model,which includes the threshold of motion and cover effectsbut not the tools effect, is arguably the most likelyamong the other models discussed above to produce theresults similar to those predicted by saltation–abrasionmodel. And only minor differences exist between thepredictions of the published versions of the modifiedalluvial and parabolic stream power models, both ofwhich include the cover effect while only the latteraccounts for the tools effect.

Finally, the suspension effect is the least important ofthe four sediment effects in influencing predictedsteady-state channel slopes. Although the version ofthe saltation–abrasion model lacking the suspension–extrapolation term was not tested here, the fact that themaximum possible bedload abrasion rate was encoun-tered at low values of transport stage relative to thethreshold of suspension (Fig. 9) indicates that thesuspension effect is not responsible for the existence ofthat instability. As discussed in detail by Sklar andDietrich (2004), the decline in incision rate withincreasing shear stress is fundamentally due to theelongation of saltation trajectories, even without anyshift in transport mode. Rather than being a computationannoyance, this result may be evidence of a real effect inriver profile evolution. Over-steepening of tributarychannels upstream of junctions could lead to hangingvalleys, or formation of migrating knickpoints. Thesehypotheses illustrate the benefits of developing geo-morphic process models beginning with the detailedmechanics, and remaining faithful to the originalassumptions through the subsequent steps of scalingup and exploring morphologic implications.

6. Conclusion

We have demonstrated how the saltation–abrasionmodel, which was derived at the spatial and temporalscale of individual bedload sediment saltation impacts,

can be scaled up for application at the scale oflongitudinal profile evolution. Because the core modelhas no freely adjustable parameters, the additionalparameters introduced in scaling up have explicitphysical significance. As a result, the sensitivity ofspecific morphologic implications, such as predictedsteady-state channel slope, to those parameter values canbe meaningfully evaluated. Two important ideas areintroduced in our treatment of the problem of balancingthe tradeoffs between the magnitude and frequency ofhigh flow events that transport bedload sediment andincise bedrock. First is the use of a real discharge record,and estimates of long-term mean denudation rates, froma reference field site, to make sure that the methodrespects the constraints that the total average annualrunoff and sediment yield must be conveyed through agiven reach of river. Second, because we recognize thecontribution of low flow discharges to the annual waterbudget, we can rationally select the representative highflow magnitude and duration that most closely appro-ximates the net effect of the full distribution ofdischarges in transporting bedload. We find thatmoderate magnitude discharges best represent the fulldistribution because low magnitude discharges fallbelow the threshold of motion, and high magnitudedischarges occur so infrequently that the associatedbedload sediment supply is too great to allow bedrockexposure in the channel bed.

Comparison of the behavior of the saltation–abrasionmodel with a range of other bedrock incision models,each of which fails to capture at least one of the foureffects of sediment, reveals important differences withsignificant morphologic implications. Models that lackthe threshold of motion over-predict incision rate forlow shear stresses and under-predict the steady-statechannel slope for low to moderate rock uplift rates androck strengths. Models that lack the cover effect over-predict incision rate for high sediment supply rates, andfail to represent the degree of freedom in channeladjustment provided by partial bed coverage. Modelsthat lack the tools effect over-predict incision rate forlow sediment supply rates, and do not allow for thepossibility that incision rate can decline for increases inshear stress above a peak value.

Two conceptual tools help clarify the controls onsteady-state channel slope predicted by the saltation–abrasion model. The first is the idea that total slope canbe considered a sum of components, each responsible forone of three tasks of a river channel at steady state, listedhere in order of descending importance overall:mobilization of bed sediments, transport of suppliedcoarse sediment load, and incision into bedrock at the

Page 24: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

81L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

rate of rock uplift. The second is the non-dimensionalrepresentation of the incision model as a function oftransport stage and relative sediment supply. Using thesetools we have shown that the influence of rock uplift rate,rock strength and the ‘storminess’ of the dischargedistribution on steady-state slope are expressed primarilythrough their effect on relative sediment supply, whichcontrols the extent of bed exposure in the channel bed.

NotationA upstream drainage area (km2)Ao drainage area at channel head (km2)b coefficient for change in Qw with A (m3 s−1

km−2p)c coefficient for change in W with Qw (m1–3f/s f )d coefficient for change in A with x (km2−h)Do grain diameter at channel head (m)Ds grain diameter (m)Ei instantaneous bedrock erosion rate (m/s)Elt long-term bedrock erosion rate (mm/year)E⁎ non-dimensional erosion rate per unit rock

resistancef exponent for change in W with Qw

Fb fraction of total load in coarse bedload sizeclass

Ft fraction of total time when high flow occursFw fraction of total runoff carried by high flowsg gravitational acceleration (m/s2)h exponent for change in Awith xKf tools (Foley, 1980) model efficiency (m s2/kg)Kh tools (Howard et al., 1994) model efficiency

(m0.9 s1.5 kg−1)Kma modified alluvial model efficiency (m/kg)Kma-b bedload modified alluvial model efficiency

(m/kg)Kn stream power modified alluvial model effi-

ciency (m2n+1 s−1 W−n)Kp parabolic stream power model efficiency

(m2n− 1 W1− n kg−1)Kqa stream power sediment transport coefficient

(kg m2 s−1 W−1)kv saltation–abrasion rock resistance coefficientKγ placeholder erosional efficiency coefficientKΩ stream power model efficiency (m1+ n s−1

W−n)Kω unit stream power model efficiency (m2n+1 s−1

W−n)Kτ shear stress model efficiency (m s−1 Pa−1)Kτ-cr excess shear stress (rock) model efficiency

(m s−1 Pa−1)Kτ-cs excess shear stress (sediment) model efficiency

(m s−1 Pa−1)

Ir impact rate per unit area per unit time (1/m2 s)Lr representative reach length (m)N number of discharges in flow recordm rank of individual dischargen exponent in stream power and shear stress

modelsnm mannings channel roughness coefficient (s/m1/3)p exponent for change in Qw with APne probability of non-exceedanceQs sediment supply (kg/s)QsM mean annual sediment supply (kg/s)Qc sediment transport capacity (kg/s)QcT total-runoffmethod sediment transport capacity

(kg/s)QcH high-flow sediment transport capacity (kg/s)Qw discharge (m3/s)QwT representative discharge, total-runoff method

(m3/s)QwH representative discharge, high flow method

(m3/s)QwL residual low flow discharge, high flow method

(m3/s)Rb non-dimensional bouyant densityRep particle Reynolds numberRh hydraulic radius (m)S channel slopeSD slope component for threshold of sediment

motionΔSQs

slope increment for transport of suppliedsediment

ΔSE slope increment for erosion of bedrockUr rock uplift rate (mm/year)u⁎ flow shear velocity (m/s)W channel width (m)wf particle fall velocity (m/s)x distance downstream of channel head (km)xo unchanneled distance (km)Y rock modulus of elasticity (MPa)α downstream fining exponent∑QcA annual transport capacity summed over flow

record (kg/year)∑QcH annual transport capacity of representative

high flow (kg/year)γ placehold flow intensity variableλ sediment porosityω unit stream power (W/m2)Ω stream power (W/m)μ viscosity of water (kg m−1 s−1)ρs sediment density (kg/m3)ρw water density (kg/m3)σT rock tensile strength (MPa)τb mean boundary shear stress (Pa)

Page 25: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

82 L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

τc-rock rock strength-determined critical shear stress(Pa)

τc-sed sediment grain size-determined critical shearstress (Pa)

τ⁎ non-dimensional shear stressτc⁎ non-dimensional critical shear stressτ⁎ /τc⁎ transport stage

Acknowledgements

We thank G. Hauer, J. Kirchner, A. Luers, J. Roering,M. Stacey, J. Stock, and K. Whipple for stimulating andinsightful discussions. Thoughtful comments by twoanonymous reviewers helped to improve the manu-script. This work was supported by the National Centerfor Earth-Surface Dynamics, by NSF grant EAR970608, and by a Switzer Environmental Fellowshipto L.S.S.

References

Bathurst, J.C., 1993. Flow resistance through the channel network. In:Beven, K., Kirkby, M.J. (Eds.), Channel Network Hydrology. JohnWiley & Sons, Chichester.

Beaumont, C., Fullsack, P., Hamilton, J., 1992. Erosional control ofactive compressional orogens. In: McClay, K.R. (Ed.), ThrustTectonics. Chapman and Hall, New York, pp. 1–18.

Bezinge, A., 1987. Glacial meltwater streams, hydrology and sedimenttransport: the case of the Grande Dixence Hydroelectricity Scheme.In: Gardiner, V. (Ed.), International Geomorphology 1986, Part I,pp. 473–498.

Bitter, J.G.A., 1963. A study of erosion phenomena, part I. Wear 6,5–21.

Bogaart, P.W., Tucker, G.E., de Vries, J.J., 2003. Channel networkmorphology and sediment dynamics under alternating periglacialand temperate regimes: a numerical simulation study. Geomor-phology 54, 257–277.

Bras, R.L., 1990. Hydrology: An Introduction to Hydrologic Science.Addison Wesley, Reading, MA.

Brierley, G.J., Hickin, E.J., 1985. The downstream gradation ofparticle sizes in the Squamish River, British Columbia. EarthSurface Processes and Landforms 10, 597–606.

Brush, L.M., 1961. Drainage basins, channels, and flow characteristicsof selected streams in Central Pennsylvania. Professional Paper-Geological Survey (U.S.) 282-F.

Buffington, J.M., Montgomery, D.R., 1997. A systematic analysis ofeight decades of incipient motion studies, with special reference togravel-bedded rivers. Water Resources Research 33, 1993–2029.

Bull, W.B., 1979. Threshold of critical power in streams. GeologicalSociety of America Bulletin 90, 453–464.

Dietrich, W.E., 1982. Settling velocity of natural particles. WaterResources Research 18, 1615–1626.

Dietrich, W.E., Montgomery, D.R., 1998. Hillslopes, channels, andlandscape scale. In: Sposito, G. (Ed.), Scale Dependence and ScaleInvariance in Hydrology. Cambridge University Press, Cambridge,pp. 30–60.

Fernandez-Luque, R., van Beek, R., 1976. Erosion and transport ofbed-load sediment. Journal of Hydraulic Research 14, 127–144.

Foley, M.G., 1980. Bedrock incision by streams. Geological Society ofAmerica Bulletin Part II, 91, 2189–2213.

Hack, J.T., 1960. Interpretation of erosional topography in humidtemperate regions. American Journal of Science 258A, 80–97.

Head, W.J., Harr, M.E., 1970. The development of a model to predictthe erosion of materials by natural contaminants. Wear 15, 1–46.

Howard, A.D., 1980. Thresholds in river regime. In: Coates, D.R.,Vitek, J.D. (Eds.), The Concept of Geomorphic Thresholds. Allenand Unwin, Boston, pp. 227–258.

Howard, A.D., 1994. A detachment-limited model of drainage basinevolution. Water Resources Research 30, 2261–2285.

Howard, A.D., 1998. Long profile development of bedrock channels:interaction of weathering, mass wasting, bed erosion, and sedimenttransport. In: Tinkler, K., Wohl, E.E (Eds.), Rivers Over Rock:Fluvial Processes in Bedrock Channels. Geophysical MonographSeries, vol. 107. American Geophysical Union, Washington, DC,pp. 237–260.

Howard, A.D., Kerby, G., 1983. Channel changes in badlands.Geological Society of America Bulletin 94, 739–752.

Howard, A.D., Dietrich, W.E., Seidl, M.A., 1994. Modeling fluvialerosion on regional to continental scales. Journal of GeophysicalResearch 99, 13971–13986.

Kooi, H., Beaumont, C., 1996. Large-scale geomorphology: classicalconcepts reconciled and integrated with contemporary ideas via asurface processes model. Journal of Geophysical Research 101,3361–3386.

Lavé, J., Avouac, J.P., 2000. Active folding of fluvial terraces across theSiwaliks Hills, Himalayas of central Nepal. Journal of GeophysicalResearch 105 (B3), 5735–5770. doi:10.1029/1999JB900292.

Lave, J., Avouac, J.P., 2001. Fluvial incision and tectonic uplift acrossthe Himalayas of central Nepal. Journal of Geophysical Research106 (B11), 26561–26591.

Leopold, L.B., 1994. AView of the River. Harvard University Press,Cambridge. 298 pp.

Mackin, J.H., 1948. Concept of the graded river. Geological Society ofAmerica Bulletin 59, 463–512.

Merritts, D., Bull, W.B., 1989. Interpreting Quaternary uplift rates atthe Mendocino triple junction, northern California, from upliftedmarine terraces. Geology 17, 1020–1024.

Moglen, G.E., Bras, R.L., 1995. The effect of spatial heterogeneitieson geomorphic expression in a model of basin evolution. WaterResources Research 31, 2613–2623.

Montgomery, D.R., Dietrich, W.E., 1988. Where do channels begin?Nature 336, 232–234.

Montgomery, D.R., Dietrich, W.E., 1992. Channel initiation and theproblem of landscape scale. Science 255, 826–830.

Morris, P.H., Williams, D.J., 1999. A worldwide correlation forexponential bed particle size variation in subaerial aqueous flows.Earth Surface Processes and Landforms 24, 835–847.

Ohmori, H., 1991. Change in the mathematical function typedescribing the longitudinal profile of a river through anevolutionary process. Journal of Geology 99, 97–110.

Parker, G., 1990. Surface-based bedload transport relation for gravelrivers. Journal of Hydraulic Research 28 (4), 417–436.

Rouse, H., 1937. Modern conceptions of the mechanics of turbulence.Transactions of the ASCE 90, 463–543.

Schumm, S.A., 1973. Geomorphic thresholds and complex response ofdrainage systems. In: Morisawa,M. (Ed.), Fluvial Geomorphology.Publications in Geomorphology, Binghamton University (SUNY),pp. 299–310.

Page 26: The role of sediment in controlling steady-state bedrock ...angelo.berkeley.edu/wp-content/uploads/Sklar_2006_Geomorpho.pdfThe role of sediment in controlling steady-state bedrock

83L.S. Sklar, W.E. Dietrich / Geomorphology 82 (2006) 58–83

Seidl, M.A., Dietrich, W.E., 1992. The problem of channel erosion intobedrock. In: Schmidt, K.H., de Ploey, J. (Eds.), FunctionalGeomorphology. Catena Suppl., vol. 23, pp. 101–124.

Sklar, L.S., Dietrich, W.E., 1998. River longitudinal profiles andbedrock incision models: stream power and the influence ofsediment supply. In: Tinkler, K., Wohl, E.E. (Eds.), Rivers OverRock: Fluvial Processes in Bedrock Channels. GeophysicalMonograph Series, vol. 107. American Geophysical Union,Washington, DC, pp. 237–260.

Sklar, L.S., Dietrich, W.E., 2001. Sediment and rock strength controlson river incision into bedrock. Geology 29, 1087–1090.

Sklar, L.S., Dietrich, W.E., 2004. A mechanistic model for riverincision into bedrock by saltating bedload. Water ResourcesResearch 40, W06301. doi:10.1029/2003WR002496.

Snow, R.S., Slingerland, R.L., 1987. Mathematical modeling of gradedriver profiles. Journal of Geology 95, 15–33.

Stock, J.D., Montgomery, D.R., Collins, B.D., Dietrich, W.E., Sklar, L.S., 2005. Field measurements of incision rates following bedrockexposure: implications for process controls on the long profiles ofvalleys cut by rivers and debris flows. Geological Society ofAmerica Bulletin 117, 174–194. doi:10.1130/B25560.1.

Tomkin, J.H., Brandon, M.T., Pazzaglia, F.J., Barbour, J., Willett, S.D.,2003. Quantitative testing of bedrock incision models, ClearwaterRiver, WA. Journal of Geophysical Research 108 (B6), 2308.doi:10.1029/2001JB000862.

Tucker, G.E., Bras, R.L., 1998. Hillslope processes, drainage density,and landscape morphology. Water Resources Research 34,2751–2764.

Tucker, G.E., Bras, R.L., 2000. A stochastic approach to modeling therole of rainfall variability in drainage basin evolution. WaterResources Research 36, 1953–1964.

Tucker, G., Slingerland, R., 1994. Erosional dynamics, flexuralisostasy, and long-lived escarpments: a numerical modelingstudy. Journal of Geophysical Research 99, 12229–12243.

Tucker, G.E., Slingerland, R.L., 1996. Predicting sediment flux fromfold and thrust belts. Basin Research 8, 329–349.

Tucker, G.E., Slingerland, R., 1997. Drainage basin responses toclimate change. Water Resources Research 33 (8), 2031–2047.

Tucker, G.E., Whipple, K.X., 2002. Topographic outcomes predictedby stream erosion models: sensitivity analysis and intermodelcomparison. Journal of Geophysical Research 107 (B9), 2179.

van der Beek, P., Bishop, P., 2003. Cenozoic river profile developmentin the Upper Lachlan catchment (SE Australia) as a test ofquantitative fluvial incision models. Journal of GeophysicalResearch 108 (B6), 2309.

Whipple, K.X., Tucker, G.E., 1999. Dynamics of the stream-powerriver incision model: implications for height limits of mountainranges, landscape response timescales, and research needs. Journalof Geophysical Research 104B, 17661–17674.

Whipple, K.X., Tucker, G.E., 2002. Implications of sediment-flux-dependent river incision models for landscape evolution. Journalof Geophysical Research, [Solid Earth] 107, 2039.

Whipple, K.X., Hancock, G.S., Anderson, R.S., 2000. River incisioninto bedrock: mechanics and relative efficacy of plucking,abrasion, and cavitation. Geological Society of America Bulletin112, 490–503.

Wiberg, P.L., Smith, J.D., 1985. A theoretical model for saltatinggrains in water. Journal of Geophysical Research 90 (C4),7341–7354.

Willett, S.D., 1999. Orogeny and orography: the effects of erosion onthe structure of mountain belts. Journal of Geophysical Research104, 28957–28981.

Willgoose, G., Bras, R.L., Rodriguez-Iturbe, I., 1991. A coupledchannel network growth and hillslope evolution model 1. Theory.Water Resources Research 27, 1671–1684.