earth surface processes and landforms conceptual model for ... · a physically based conceptual...

17
Copyright © 2006 John Wiley & Sons, Ltd. Earth Surface Processes and Landforms Earth Surf. Process. Landforms 31, 875–891 (2006) Published online 29 March 2006 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/esp.1297 A conceptual model for meander initiation in bedload-dominated streams Brett C. Eaton, 1 * Michael Church 1 and Tim R. H. Davies 2 1 Department of Geography, The University of British Columbia, Vancouver, British Columbia, Canada, V6T 1Z2 2 Department of Geological Sciences, University of Canterbury, Canterbury, New Zealand Abstract A simple analytic model is presented relating local sediment transport capacity to variance in the transverse shear stress distribution in a stream channel. The model is used to develop a physically based conceptual model for the initiation of meandering in straight, bedload- dominated streams as a result of a feedback mechanism. The feedback maximizes the cross- sectional shear stress variance and – in order to achieve stability – ultimately minimizes the energy slope at repeated locations along the channel, subject to steady-state mass flux and the stability of the channel boundary. These locations develop into pools in a fully developed meandering channel; they represent attractor states wherein sediment continuity is satisfied using the least possible energy expenditure per unit length of channel. However, since the cross-sectional geometry of a pool (and the adjacent bar) is asymmetric, these attractor states are only conditionally stable, requiring strong, curvature-induced secondary circula- tion to maintain their asymmetry. Between two successive pools, a stream occupies a metastable, higher energy state (corresponding to a riffle) that requires greater energy ex- penditure per unit length of channel to transport the same volume of sediment. The model we present links processes at the scale of a channel width to adjustments of the channel sinuosity and slope at the scale of a channel reach. We argue that the reach-scale extremal hypotheses employed by rational regime models are mathematical formalisms that permit a one-dimensional theory to describe the three-dimensional dynamics producing stream mor- phology. Our model is consistent with the results from stream table experiments, with respect to both the rate of development of meandering and the characteristics of the equilib- rium channel morphology. Copyright © 2006 John Wiley & Sons, Ltd. Keywords: sediment transport; meander development; bank stability; fluvial system; extremal hypotheses Received 13 October 2004; Revised 12 July 2005; Accepted 6 September 2005 Introduction Experimental studies of bedload-transporting streams with erodible banks demonstrate that straight, rectangular chan- nels are unstable once stream power exceeds a critical competence limit (e.g. Schumm and Khan, 1972; Eaton and Church, 2004). The dominant form of channel adjustment is the development of a sinuous, single-thread channel with pools on alternate sides of the channel. This adjustment occurs even in experiments when the ultimate channel pattern is braided (e.g. Ashmore, 1991). Where an equilibrium single-thread channel planform is reached, the degree of sinuo- sity is functionally determined by the sediment supply (Q b ), valley slope (S v ) and discharge (Q), and does not seem to depend on the time-to-equilibrium or the sequence of vertical and lateral adjustments that produce the equilibrium form (Eaton and Church, 2004). This is described by Eaton and Church (2004) as slope-minimizing behaviour (SMB), which denotes the tendency for streams to bypass slopes at which the sediment supply could theoretically be trans- ported by the available discharge – according to one-dimensional numerical models – in favour of lower slopes. SMB is physical evidence of the sorts of processes implicit in extremal hypotheses describing reach-scale river behaviour (Yang, 1976; Kirkby, 1977; Chang, 1979; White et al., 1982; Davies and Sutherland, 1983). Using the conceptual model developed herein, we attempt to demonstrate that, in mechanical terms, SMB is a manifestation of the linkages between the reach-average geometric properties of an alluvial system – in particular the sinuosity ( L*) and average slope (S) – and the three-dimensional, time-varying stress/transport field that deforms the channel boundaries *Correspondence to: B. C. Eaton, Department of Geography, The University of British Columbia, Vancouver, British Columbia, Canada. E-mail: [email protected]

Upload: others

Post on 08-Jul-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

Conceptual model for meander initiation 875

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

Earth Surface Processes and LandformsEarth Surf. Process. Landforms 31, 875–891 (2006)Published online 29 March 2006 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/esp.1297

A conceptual model for meander initiationin bedload-dominated streamsBrett C. Eaton,1* Michael Church1 and Tim R. H. Davies2

1 Department of Geography, The University of British Columbia, Vancouver, British Columbia, Canada, V6T 1Z22 Department of Geological Sciences, University of Canterbury, Canterbury, New Zealand

AbstractA simple analytic model is presented relating local sediment transport capacity to variancein the transverse shear stress distribution in a stream channel. The model is used to developa physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a result of a feedback mechanism. The feedback maximizes the cross-sectional shear stress variance and – in order to achieve stability – ultimately minimizes theenergy slope at repeated locations along the channel, subject to steady-state mass flux andthe stability of the channel boundary. These locations develop into pools in a fully developedmeandering channel; they represent attractor states wherein sediment continuity is satisfiedusing the least possible energy expenditure per unit length of channel. However, since thecross-sectional geometry of a pool (and the adjacent bar) is asymmetric, these attractorstates are only conditionally stable, requiring strong, curvature-induced secondary circula-tion to maintain their asymmetry. Between two successive pools, a stream occupies ametastable, higher energy state (corresponding to a riffle) that requires greater energy ex-penditure per unit length of channel to transport the same volume of sediment. The modelwe present links processes at the scale of a channel width to adjustments of the channelsinuosity and slope at the scale of a channel reach. We argue that the reach-scale extremalhypotheses employed by rational regime models are mathematical formalisms that permit aone-dimensional theory to describe the three-dimensional dynamics producing stream mor-phology. Our model is consistent with the results from stream table experiments, withrespect to both the rate of development of meandering and the characteristics of the equilib-rium channel morphology. Copyright © 2006 John Wiley & Sons, Ltd.

Keywords: sediment transport; meander development; bank stability; fluvial system; extremalhypotheses

Received 13 October 2004;Revised 12 July 2005;Accepted 6 September 2005

Introduction

Experimental studies of bedload-transporting streams with erodible banks demonstrate that straight, rectangular chan-nels are unstable once stream power exceeds a critical competence limit (e.g. Schumm and Khan, 1972; Eaton andChurch, 2004). The dominant form of channel adjustment is the development of a sinuous, single-thread channel withpools on alternate sides of the channel. This adjustment occurs even in experiments when the ultimate channel patternis braided (e.g. Ashmore, 1991). Where an equilibrium single-thread channel planform is reached, the degree of sinuo-sity is functionally determined by the sediment supply (Qb), valley slope (Sv) and discharge (Q), and does not seem todepend on the time-to-equilibrium or the sequence of vertical and lateral adjustments that produce the equilibriumform (Eaton and Church, 2004). This is described by Eaton and Church (2004) as slope-minimizing behaviour (SMB),which denotes the tendency for streams to bypass slopes at which the sediment supply could theoretically be trans-ported by the available discharge – according to one-dimensional numerical models – in favour of lower slopes.

SMB is physical evidence of the sorts of processes implicit in extremal hypotheses describing reach-scale riverbehaviour (Yang, 1976; Kirkby, 1977; Chang, 1979; White et al., 1982; Davies and Sutherland, 1983). Using theconceptual model developed herein, we attempt to demonstrate that, in mechanical terms, SMB is a manifestation ofthe linkages between the reach-average geometric properties of an alluvial system – in particular the sinuosity (L*)and average slope (S) – and the three-dimensional, time-varying stress/transport field that deforms the channel boundaries

*Correspondence to:B. C. Eaton, Department ofGeography, The University ofBritish Columbia, Vancouver,British Columbia, Canada.E-mail: [email protected]

Page 2: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

876 B. C. Eaton, M. Church and T. R. H. Davies

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

according to well-known Newtonian principles and gives rise to these geometric properties. While this argument is notmade by way of analogy with thermodynamics, the approach is broadly equivalent: our conceptual model is anattempt to describe the large-scale manifestation of Newtonian physics in much the same way as thermodynamicdiffusion describes the outcome of many individual molecular collisions.

Following on from the pioneering work on rational regime theory (Yang, 1976; Kirkby, 1977; Chang, 1979; Whiteet al., 1982; Davies and Sutherland, 1983; Millar and Quick, 1993), Eaton et al. (2004) proposed that stable channelmorphologies arise from the maximization of flow resistance within what they define as the fluvial system ( fsys),subject to the limits of bank stability and the ability to pass the imposed sediment supply with the available discharge.This conjecture can be shown to be similar to the minimum energy argument advanced by Huang et al. (2004), thoughthey pay no attention to the important constraint of bank erodibility. However, both propositions are essentiallyprobabilistic considerations of the problem constructed at the reach scale. Nothing is said about the process–forminteractions occurring at the sub-reach scale, at the level of individual bars, riffles and pools. The striking regularity ofthe meanders produced in the laboratory, and the organized way in which equilibrium evidently is achieved (Eaton andChurch, 2004), does not seem to support the position that channel change is simply a random process, guided in onlya general way by a drift towards stability: rather, fluvial systems appear to be self-organized, exhibiting emergentbehaviour as a result of interactions among random processes occurring at the grain scale.

Linear stability models (LSMs) of alternate bar development (e.g. Engelund and Skovgaard, 1973; Parker, 1976;Fredsoe, 1978) represent an attempt to model the apparent regularity exhibited by meandering channels. The form ofthe bed adjustment (the perturbation, in the language of stability analysis) is specified at the outset: usually a doubleharmonic perturbation of diminishingly small amplitude is imposed. These models, and the more recent LSMs thatcouple the alternate bar-scale analysis with an analysis of the channel alignment using some form of bank erosionrelation (Blondeaux and Seminara, 1985; Johannesson and Parker, 1989; Sun et al., 2001), consistently predict thatuniform straight channels are unstable, once perturbed, and produce reasonable predictions of stream channel charac-teristics. But, in order to achieve solutions, these models must constrain the nature of the initial perturbation andsimplify the relation between transport dynamics and channel geometry. As a result, while they are very informativewith respect to the existence and persistence of instabilities producing meandering stream morphology, they are notwell suited to developing the mechanistic links between reach-scale properties and the patterns of sediment transport.Our conceptual model, based on the feedback process described in this paper between the asymmetry of the cross-sectional shear stress distribution and the local transport capacity, attempts to explain why uniform shear stressdistributions are nearly always unstable, once perturbed.

The regime argument put forward by Eaton et al. (2004) recognizes three separate scales of flow resistance (grainscale, bedform scale and reach scale). It is hypothesized that, as a system attains stability, it will minimize the energyavailable to deform the system by adjusting one or more of these flow resistance components so as to optimize thedissipation of energy as frictional resistance to flow. The adjustment of each component must be the result of someprocess–form interaction. For example, the relation between sediment supply and surface texture and structure, andthereby the grain-scale flow resistance, is well known (Dietrich et al., 1989; Church et al., 1998). When stream banksare relatively erodible, it appears that the reach-scale flow resistance (i.e. the sinuosity of the thalweg) is preferentiallyadjusted (Eaton and Church, 2004), and the flow resistance maximization is manifested by minimizing the channelslope. The conceptual model presented herein primarily describes this reach-scale adjustment, but it must be remem-bered that adjustments of the grain-scale and the bedform-scale flow resistance components will modulate the reach-scale response. This is perhaps the defining difference between our model of meander initiation and a more completemodel of stream channel dynamics.

We do not, then, claim that the model described here incorporates all necessary and sufficient conditions to describealluvial channel behaviour. We also wish to emphasize that the model is primarily concerned with initiation ofmeandering in single-thread channels, since the added degree of freedom introduced by channel bifurcations andconfluences requires a more sophisticated modelling approach than is consistent with the use of regime theory in thispaper. We will: (i) present a theoretical basis for the hypothesized feedback founded on a simplified representation ofthe shear stress distribution to estimate the sediment transport capacity; and (ii) present evidence from the experimentsdiscussed in Eaton and Church (2004) that demonstrates the action of the hypothesized feedback mechanism toproduce equilibrium channels.

Transport Capacity for Asymmetric Shear Stress Distributions

The feedback initiating meandering and slope reduction (SMB) operates locally, at the level of the channel cross-section. Once initiated, it propagates downstream. The feedback is described using a simplified shear stress

Page 3: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

Conceptual model for meander initiation 877

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

Figure 1. Assumed shear stress distributions used to evaluate the transport capacity. For all three configurations, W and τ- areconstant; only the variance of the distribution changes.

distribution characterized by a width (W), mean shear stress ( 1 ) and a parameter (β) representing a measure ofthe variance of the shear stress distribution. This is shown in Figure 1, where the magnitude of the shear stressincreases downward, and the horizontal coordinate, x, represents a distance in the distribution space. The coordinate xis defined as the proportion of the distribution (P) less than a given shear stress value (P varies from 0 to 1), scaled bythe channel width (x varies from 0 to W). The result is that the shear stress distributions in Figure 1 can be interpretedas graphical representations of channel cross-sections under conditions where one accepts the uniform flow approxi-mations, and as simple shear stress distributions in their own right when one does not accept them.

In the limit case when β = 0, the distribution is uniform, corresponding to a transport capacity TC . Once thecharacteristic grain size (D) is known, the transport capacity is estimated using a simplified form of the Meyer-Peterand Müller equation (Meyer-Peter and Müller, 1948):

TC = Wξ[τ − τc]3/2 (1)

wherein ξ is a constant, τc is the critical shear stress for entrainment, given by τc = 0·045(γs − γ )D (cf. Meyer-Peterand Müller, 1948), γ is the unit weight of water and γs is the unit weight of the sediment.

For non-uniform distributions (β > 0) the sediment transport capacity increases. For example, the effect of longitu-dinal shear stress variations associated with ripples is reportedly sufficient to maintain sediment transport and ripplemigration at shear stresses less than that required to initiate transport from a flat bed (Rathbun and Guy, 1967).Similarly, Schumm and Khan (1972) observed that meanders formed at lower stream powers if the inlet was angled soas to induce an initial bend than if the inlet was straight: the initial bend presumably created an asymmetric shearstress distribution with sufficient transport capacity that the riffle–pool sequence could be propagated downstream, aprocess that they did not observe in the absence of significant cross-sectional shear stress variance. The transportcapacity (TC) for a non-uniform distribution is given by:

TC dx

x

W

cc

[ ] /= −ξ τ τ� 3 2 (2)

where xc is the point at which τ(x) = τc and W corresponds to the position in the distribution space (defined above)of the maximum shear stress (τo). Geometry dictates that there is a maximum value of β that can be imposed whilestill maintaining the same W and 1 as the equivalent uniform distribution. This limit, expressed as an angle basedon the approximation ∂ = 1/γS, is given by:

βγmax tan=

−1 21

SW(3)

Ferguson (2003) presents a more complete treatment of the difference between the quantities defined by Equations 1and 2 in which he employs a shear stress distribution represented by two planar surfaces. A shape parameter, b, isvaried between the limits 0 and 1, representing distribution shapes ranging from uniform to a convex dogleg, respec-tively. A distribution with a triangular geometry, similar to those in Figure 1, is described by b = 0·5 in Ferguson’sframework. Ferguson observes that the distribution of shear stress in one particular test case (Fraser River, BritishColumbia) is consistent with a simple distribution (b ≅ 0·4 to 0·7). It may be inferred, then, that the simple shear stressdistributions assumed in this analysis are adequately realistic.

Page 4: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

878 B. C. Eaton, M. Church and T. R. H. Davies

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

The nature of the feedback process depends on how the ratio TC/ TC behaves. When TC/ TC > 1, a local excesstransport capacity results, which will necessarily deform the channel through net erosion until the disparity is reduced.For all β in the range βmax > β > 0, this ratio becomes infinite in the limit (1 → τc), since TC approaches 0 but TCremains finite, if small. At the other extreme (1 → ∞), the ratio approaches 1 (no effect). The behaviour between theselimits depends on the channel width (W) and the distribution variance (β).

If feedback strength is expressed as (TC – TC )/ TC, excess shear stress is expressed as (1 – τc)τc, and shear stressdistribution variance is expressed as a proportion of the theoretical maximum (β/βmax), then the dependence on Wbecomes implicit. The strength of the feedback tendency is functionally related to excess shear stress for given valuesof β/βmax, as shown in Figure 2.

Conceptual Model of Meander Initiation

The feedback which is the basis for the meander initiation model is based on the idea that: (i) a disparity between TCand TC produces net erosion centred around the position of maximum shear stress; and (ii) this local erosion willincrease the variance of the shear stress distribution, thus increasing the disparity between TC and TC, and producinga positive feedback. This ultimately necessitates a change in the mean shear stress via a channel slope adjustment inorder to establish a stable, meandering channel pattern.

Characteristics of the positive feedback mechanismTwo key points must be made before describing in detail the feedback mechanism. First, it is evident in Figure 2 that,as the mean shear stress approaches τc, the effect of the distribution variance (β/βmax) on the local transport capacityis proportionally greater. For example, a mean shear stress 50 per cent greater than τc has a transport capacity only30 per cent greater than TC, assuming a relatively high variance (β/βmax = 0·5). For the same distribution variance, ashear stress only 10 per cent larger than τc has a transport capacity that is fully 300 per cent greater than TC. Second,

Figure 2. Proportional increase in TC relative to TC plotted against excess shear stress for a range of relative variances (β/βmax)

Page 5: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

Conceptual model for meander initiation 879

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

provided the mean shear stress (1) does not change (which amounts to holding the channel slope constant), thefeedback intensifies as β/βmax increases. Once the feedback starts, it quickly increases the local disparity between thereach-average sediment supply and the local transport capacity, TC. For a constant excess shear stress, TC increasesquickly with increasing β/βmax (Figure 2). For example, if the excess shear stress is 25 per cent, then TC is 5 per centgreater than TC when β/βmax is about 1/8, and 80 per cent greater when β/βmax is about 1/2.

Both of these effects are likely to be generally true of any reasonable shear stress distribution model. They certainlyapply to the shear stress distribution models introduced by Ferguson (2003). However, it is appropriate to emphasizethat the driving variable is (τ − τc) not simply τ, and that sorting of bed sediment under partial transport conditions caninhibit the sort of feedback that we invoke. Sorting amounts to a form of grain-scale flow resistance response in theframework described by Eaton et al. (2004), and the topic is relatively well studied (Parker et al., 1982; Parker andKlingeman, 1982; Dietrich et al., 1989; Wilcock and McArdell, 1993; Paola and Seal, 1995). So, the formative flowsresponsible for the construction of riffle and pool sequences must be capable of moving nearly all of the grains foundon the bed (locally, if not everywhere) in something approaching full mobility (Wilcock and McArdell, 1993) in orderto prevent the development of an armour layer sufficient to suppress the feedback. The area subject to fully mobiletransport need not be extensive to initiate meander development; it may be sufficient to have very localized fullymobile transport, which can occur in laboratory experiments, at least, under flow conditions not very different fromthreshold.

Linking the positive feedback to planform adjustment and stabilitySimplifying assumptions. Since we are concerned primarily with the relative (in)stability of a flat-bottomed, straightchannel we can usefully apply the uniform flow approximations, wherein the distribution of shear stress (τ) has adirect correspondence to depth distribution (after Lane, 1955). Visual inspection during stream table experiments(Eaton and Church, 2004) suggests that the distribution of sediment transport is highly correlated with the local waterdepth, particularly during the phase of meander initiation. Once a channel develops significant cross-sectional asym-metry, the role of three-dimensional flow structure becomes more important, but by that time the initial morphologicstructure has been set (i.e. the meanders have been initiated), which conditions the further development of the channelmorphology.

In addition to disregarding the three-dimensional flow structure, we ignore the accelerations and decelerations of theflow that may produce spatial variations in velocity head (and thereby the energy gradient), and assume that the meanvelocity is similar everywhere. We justify this assumption on the observation that the velocity for riffle and pool cross-sections in laboratory experiments (Eaton and Church, 2004) were not systematically different, and on the reportedconvergence of velocities at near-bankfull conditions in the field (Carling, 1991; Clifford and Richards, 1992).

We also note that the maximum possible reach average sediment throughput (Qb) for the system is equal to TCsince sections with uniform shear stress distributions act as chokes within the system. It must also be equal to thesediment supply rate if a system is in equilibrium.

The feedback process. Assuming an initially uniform shear stress distribution, a small local perturbation is imposedin the form of bed erosion (see Figure 3 (first channel configuration) and Figure 4). At this stage, the thalweg remainsstraight, and the water surface slope remains unchanged. However, for flow conditions that are nearly uniform, arandom local increase in depth will increase β, thereby increasing TC at the site of the perturbation. We assume thatthis process does not modify ∂ and W appreciably. Nevertheless, a perturbation that increases the variance, β, willquickly amplify through localized net erosion at an increasingly rapid rate. The system is no longer stable. Excesstransport capacity is localized around the initial perturbation and net scour continues, increasing the maximum depth(Yo) and β. This net change is primarily vertical, the banks remaining stable.

Downstream, where the shear stress distribution remains uniform, deposition necessarily occurs (see Figure 3,second channel configuration). The deposition will affect β there and initiate another perturbation. Since entrainedsediment will be deposited on the same side of the channel, the new perturbation will occur on the opposite side of thechannel. That is, divergence of the sediment flux modulates the shear stress distribution and replicates the initialperturbation. The longitudinal scale of the flux divergence, which presumably is related to the step length for sedimenttransport at formative discharge (Pyrce and Ashmore, 2003), sets the wavelength for the emergent channel bars. Thecharacteristic wavelength is considered in more detail below.

The rate of morphologic change accelerates rapidly, since the initial increase in TC will cause further net erosion,around the deepest part of the channel, increasing β and TC in the positive feedback loop shown at the top of Figure4. At some point, the vertical degradation will cause the adjacent stream bank to fail by over-steepening it, introducinga lateral component to the adjustment and producing the third configuration in Figure 3. The onset of bank erosion iscritical to the development of meanders (Friedkin, 1945), at which time the flux divergence increases rapidly (Dietrich

Page 6: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

880 B. C. Eaton, M. Church and T. R. H. Davies

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

and Smith, 1983), and lateral adjustment comes to dominate the system. Channel slope begins to adjust and a negativefeedback is initiated (bottom of Figure 4), through which channel sinuosity increases and slope decreases. Stabilityresults from the reduction in channel slope (fourth configuration, Figure 3). At this point, the assumptions of our initialanalysis clearly are no longer valid, but the meander form has been initiated and further development is stronglyconditioned by the initial morphologic change.

Longitudinal variations in energy gradient must develop within the system in order to maintain a constant TC in achannel where β varies. The slope over the apex of the developing bend (i.e. the zone of most intense positivefeedback – which will be referred to as an attractor) is quickly reduced, and the slope over the crossover (where β isnearly 0) will increase due to the drawdown effect of pool deepening (downstream of the crossover) and the deposi-tion of sediment eroded from the pool (on the upstream side of the crossover). At equilibrium, the shallower, moreuniform crossover sections cannot have the same local slope as the apex sections if they are to continue to transport,on average, the same throughput sediment load.

Meander wavelength. There remains the issue of the wavelength for the adjustment. In the field, meander wave-length is remarkably consistent across a range of channel scales, ranging from about 5W to 15W (Leopold et al., 1964;Yalin, 1971; Rhoads and Welford, 1991). In laboratory experiments (e.g. Garcia and Nino, 1993), the wavelength ofalternate bars developed between fixed, straight banks is nearly constant, despite a range of flow depths and channelgradients. On all evidence, the critical length scale appears to be the width of the channel. Yalin (1971) argues that thisis the result of some spatial sinusoidal autocorrelation with a characteristic length scale equal to the channel width. Heattributes the autocorrelation to channel width-scale turbulent eddies, but this seems an unlikely control given thecirculation cell flow structure typical of meandering rivers (Hey, 1976; Dietrich and Smith, 1983).

We relate the characteristic meandering wavelength to the diffusion of sediment, eroded at the initial perturbations,across the channel. Experimental evidence reported by Nikora et al. (2002) suggests that the rate of lateral sedimentdiffusion is consistent with the lower range of observed meander wavelengths (c. 8W). This sediment diffusionprocess has also been observed in experiments studying bar development, wherein bedload sheets have been observedto migrate downstream and expand across the channel (Fujita, 1989; Fukuoka, 1989; Pyrce and Ashmore, 2003).

Figure 3. Schematic diagram of cross-sectional and planimetric adjustments resulting from the action of the positive and negativefeedback described in Figure 4. Slope adjustments are initially associated with changing flow path length, but are reinforced by thepattern of erosion and sedimentation, indicated by the paired zones of erosion (E) and deposition (D).

Page 7: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

Conceptual model for meander initiation 881

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

This sediment transport-based view of meander initiation and propagation is consistent with Parker’s (1976) analysis,which concludes that sediment transport is the primary cause of alternate bar-type instability, and with Smith’s (1998)observation that channel width-scale bedload deposition is the key mechanism giving rise to meandering in his streamtray experiments.

Though experimental evidence provides some sense for the rate of diffusion, we do not fully understand thephysical processes controlling the diffusion (Nikora et al., 2002). Given that cross-stream velocity components instraight channels are typically on the order of a few per cent of the downstream flow velocity (Tracy, 1965), it seemsunlikely that lateral momentum diffusion is responsible for sediment wave diffusion. We note, however, that, in afluid, a change in the flow structure (and hence the sediment flux) can be transmitted at a maximum speed equal to thecelerity of a gravity wave (vg). Accordingly, we suppose that the sediment wave emanating from the perturbation canpropagate laterally by influencing the flow structure at a speed (vp) that is proportional to vg (cf. Werner, 1951;Anderson, 1967), thus:

v gdg = and vp = αvg (4)

Figure 4. Flow chart representing the feedback responsible for the initiation of meander development and the feedback haltingmeander development at the equilibrium slope and sinuosity.

Page 8: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

882 B. C. Eaton, M. Church and T. R. H. Davies

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

where α is a parameter between zero and unity. The angle at which the disturbance propagates across the channel isgiven by the ratio (vp/v), where v is the average stream velocity, or, equivalently, by (α/Fr), where Fr is the Froudenumber, since Fr = v/vg.

Figure 5 shows a schematic representation of our conceptual view of the diffusion process. As the edge of thesediment wave approaches the opposite bank, flow diverging over the edge of the wave front encounters the bank, andstrong turbulent circulation cells develop. This stalls the advance of the wave front. Although sediment can betransported across the front of the sediment wave, it is no longer accreted to the front of the bedform. Downstream,the development of strong secondary circulation adjacent to the bank forces the flow to converge and limits the lateralbedform growth, thereby distorting the wave, increasing β and triggering another local scour and a new sedimentwave that diverges across the channel in the opposite direction. This view assumes that sediment is transported in aseries of nearly closed erosion and deposition cells with length scales of 0·5λ, which Pyrce and Ashmore (2003)demonstrate to be the case.

The point at which the proximity of the bank first produces an effect sufficient to stall the advance of the wave frontcorresponds to the location of the root of what becomes the crossover riffle (position ‘b’ in Figure 5). The interplaybetween the tendency for the sediment wave to diffuse laterally and increasingly strong secondary circulation betweenthe wave front and the channel bank results in the equilibrium wave front configuration and the associated alternatebar deposit (Figure 5B). If we assume that the root of the riffle is approximately midway between the upstream anddownstream pools ( ab ≈ bc ), then the pool-to-pool spacing is ac, and therefore the wavelength for the alternate barsis 2( ac ). Thus, we can combine the divergence angle (α/Fr) with the channel geometry parameters (W, λ) such that,from geometrical similarity:

λα

=

4FrW (5)

W is the obvious natural scale that constrains wave diffusion and Fr reflects the rate at which a disturbance in the fluidcan be transmitted laterally. The details of the physical processes governing the lateral propagation of a bedload sheetare parameterized in the coefficient, α, which can be expected to vary with channel shape, the degree of divergencebetween the sediment flux and the velocity field and/or relative roughness. An LSM-derived equation for predictingwavelength, presented by Parker and Anderson (1975) and modified by Ikeda (1984), can be expressed in a similarform:

λκ

κ , =

=

5FrW S

W

dwhere (6)

Equation 6, then, can be interpreted as an expression of the dynamics shown in Figure 5, implying that our conceptualview of the processes controlling meander wavelength is consistent with the scaling predicted by linear stability analysis.

Figure 5. Diffusion of a sediment wave originating at an initial perturbation (A), and the resultant morphology associated with anumber of such equilibrium, morphologically expressed wave fronts (B).

Page 9: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

Conceptual model for meander initiation 883

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

While it is difficult (perhaps impossible) to derive α theoretically, we can at least estimate α based on data from thefield and from laboratory experiments. Lewin (1976) provides a field example of meander initiation in a gravel-bedriver that was artificially straightened (Fr ≈ 0·85). The initial characteristic wavelength for the bars that developed wasabout 6·1W, while the ultimate stable wavelength, once cut-banks and point bars developed, was about 6·75W, suggest-ing α ≈ 0·5. Data from laboratory experiments generally support this estimate. The wavelengths for the constantsinuosity experiments reported by Eaton and Church (2004) – for which we know the effective width tolerably well –are on average about 7·3W. The Froude number is about 0·83, implying α ≈ 0·46. However, data from experimentsreported by Garcia and Nino (1993) indicate that α is not constant; rather, it varies with Fr (Figure 6). The data fromall of the laboratory experiments exhibit a consistent general trend, despite the fact that the wavelengths for Garciaand Nino’s (1993) experiments are significantly higher (λ ≈ 9W, on average) than those reported by Eaton and Church(2004). A similar behaviour that may be analogous is the development of dune fields downstream of initial perturbations,as reported by Venditti et al. (2005) (Figure 7). The angle of cross-channel propagation at Fr ≈ 0·33 for the bedformsimplies α ≈ 0·25 (and from Equation 5, λ ≈ 5·5W).

The results are generally consistent with the λ ≈ 2πW scaling suggested by Yalin (1971), and conform to the lowerrange of the reported wavelengths for fully developed meanders in natural channels. Since the argument above is

Figure 6. Relation between α and Fr for laboratory experiments. The black circles correspond to the channels described byEaton and Church (2004) and the grey circles indicate the results reported by Garcia and Nino (1993). An approximate curvedescribing the general trend of the data was fitted by eye.

Figure 7. Bedforms developing downstream of both positive and negative defects in the experiments reported by Venditti et al.(2005). The dashed lines approximately indicate the rate of lateral propagation of the bedforms across the channel.

Page 10: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

884 B. C. Eaton, M. Church and T. R. H. Davies

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

based on the maximum possible rate of cross-channel diffusion, it is expected that the analysis should represent thelower bound, and that other factors, such as the role of secondary circulation and bank erosion, could easily distort thealternate bar wavelengths predicted by Equation 5. According to Parker (1976), the development of significant second-ary circulation is a trailing phenomenon that enhances the development of meandering, and informs the ultimate stablechannel morphology. That secondary circulation is not incorporated into our conceptual model is not a seriouslimitation, then, since we are concerned with understanding the initiation of the meander structure, not with predictingthe ultimate meander geometry.

On the persistence of the instabilityThe LSMs presented by Engelund and Skovgaard (1973) and by Parker (1976) show that, provided the bed is activelytransporting sediment, some sort of instability always grows. Fredsoe (1978) predicts general instability for W/dgreater than 8. However, since they assume that the perturbation, though infinitesimal, is ubiquitous, and that bars willemerge spontaneously everywhere, they cannot be used to investigate the processes that actually produce the instabil-ity. We have suggested a mechanism that produces a perturbation of similar geometry with an appropriate length scale,which diffuses downstream from an initially localized perturbation via sediment flux divergence.

An important question remains: what is the initial cause of the local perturbation? One plausible cause is the effectof a channel wall on the flow structure and turbulence intensity. It has long been known that, in a straight channel, theturbulence intensity is greatest adjacent to the side-wall (Gessner and Jones, 1965), and Tominaga et al. (1989) relatethe production of vorticity there to secondary circulation cells adjacent to the bank. These cells scale with the waterdepth and the relative roughness of the banks, consistent with results reported by Einstein and Shen (1964), whoobserved alternate bar formation only in the presence of rough channel banks. The cells are not large enough to affectthe entire channel cross-section, but they can generate an initial perturbation of the sort we invoke above. Allen (1994)and Smith (1996) summarize the turbulent structure along the banks as a series of corkscrew vortices that areperiodically disrupted by a horseshoe vortex, which may coalesce and grow in scale. These two scales of periodicbehaviour are implicit in the time-averaged secondary flow cell, and may be responsible for the initiation and mainte-nance of a perturbation for long enough that the feedback between cross-sectional shear stress variance, flow structureand transport capacity can begin. However, the feedback will amplify only at flows that can produce full mobility inthe developing pools.

The cause and persistence of a local perturbation must fundamentally lie in the interaction between the probabilitythat a flow excursion will produce local erosion, and the probability that an altered sediment supply from upstreamdue to some other perturbation will interact with it, thereby suppressing or enhancing the net erosion. That is, randomprocesses interact with each other, producing the self-organization (the alternating series of riffles, pools and bars)characteristic of most bedload-dominated streams. The outcome is related to the step length for sediment transport.While step length may be set, essentially, by the spatiotemporal structure of the flow over a nearly uniform bed or,later, by the larger scale structure of pools and bars (Pyrce and Ashmore, 2003), it is appropriate to view it as a pro-babilistic statement about mean distances of movement, and thus related primarily to the mean boundary shear stress.

Scales of stability for developed meandersIn this section we relate the processes and feedbacks in our model of meander initiation to the structure and functionof fully developed meanders. At the reach scale, stability has been argued to result from the maximization of thesystem-scale flow resistance and the consequent minimization of the energy available to deform the system (Eatonet al., 2004). However, at the scale of morphologic units, stability is conditional, and there is a continual oscillationbetween various states along the stream channel and over time (Figure 8). The minimum variance sections (i.e. theriffles) are metastable, at best, as a result of the susceptibility of nearly uniform shear stress distributions to perturbations.The maximum variance sections (i.e. the pools, or attractors) are also unstable. Transitions to maximum variancestates are achieved by changing the local channel path length and thereby lowering the local channel gradient. Theresultant channel shape becomes intimately linked with the centripetal acceleration of flow around a bend, and theresultant secondary circulation currents (Thompson, 1986). When the centripetal acceleration ceases and the second-ary currents weaken, sediment that is tracking along the bar face descends and deposits into the thalweg, reducing thecross-channel shear stress variance and necessitating an increased slope to transport the incoming sediment (cf.scouring and filling sections; Andrews, 1979; Lisle, 1979). The maximum variance solutions are stable, conditionalon the existence of sufficient centripetal acceleration to maintain their asymmetric cross-sectional form.

The pools can be considered to be attractor states since, due to their greater shear stress variance, they can do thesame geomorphic work while expending less energy, as described above. The spatiotemporal pattern of channel

Page 11: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

Conceptual model for meander initiation 885

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

Figure 8. Meandering as a product of oscillation between maximum variance attractors on opposite sides of the channel. Theassociated patterns of change for the local slope and relative cross-sectional shear stress variance are shown on the right(0 = uniform shear stress distribution, 1 = maximum variance of the shear stress distribution).

adjustment involves the attainment of a maximum variance state on one side of the channel (a triangular shear stressdistribution with the maximum shear stress, τo. along the right bank, for example), followed by the abandonment ofthat state for one with a nearly uniform shear stress distribution, and ultimately followed by the development of amaximum variance state on the opposite side of the channel (τo along the left bank) (Figure 8).

For meandering streams, then, there is no single, stable state. They maintain relative pattern stability by oscillatingbetween two attractors having τo on opposite sides of the channel which are associated with curvature-inducedsecondary circulation in opposite directions. In the transition between the two attractors, the stream occupies ametastable state. Thus, meandering can be viewed as a dynamic process, produced by the attraction toward condition-ally stable maximum variance conditions on alternate sides of the channel. The oscillation is responsible for thecharacteristic pattern of alternating pools with asymmetric cross-sectional shear stress distributions having a largevariance, separated by riffles where the shear stress distribution is generally more uniform. This can be interpreted asa spatial oscillation along the path of sediment and fluid flux, or alternatively as a temporal oscillation at a fixed point,produced by the downstream migration of the meander form.

Page 12: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

886 B. C. Eaton, M. Church and T. R. H. Davies

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

Comparisons with Experimental Data

The argument presented above suggests that the feedback mechanism should produce some characteristic features inalluvial systems. Experiments on a stream table where Q, Qb and Sv are treated as independent variables demonstratemany of these expected features (Eaton and Church, 2004).

Observed slope reductionMost obviously, Eaton and Church’s (2004) experiments demonstrate that initially straight channels with a mobile bedincrease their sinuosity until they reach equilibrium configurations at slopes less than the initial slope. Holding Qconstant, the reported equilibrium slope depends on the sediment supply to the system, and the documented trajecto-ries toward equilibrium show sinuosity increasing monotonically – the so-called slope minimizing behaviour. Eatonand Church (2004) also demonstrate that the response of systems at equilibrium to subsequent changes in the govern-ing conditions is consistent with the hypothesized feedback mechanism, since when the sediment supply was lowered,the channel – initially at equilibrium with the higher feed rate – responded by instigating another phase of bankerosion at the apices, producing an increase in sinuosity until equilibrium was re-established at a lower slope. Whenthe sediment supply to an equilibrium channel was increased, the reverse seemed to occur, though no equilibriumchannels were established since the single-thread channels tended to give way to multiple-thread ones.

Time-to-equilibriumThe rate of adjustment for these experiments was also consistent with our model, which predicts that when the shearstress acting on the bed is much larger than τc, the amplification of the transport capacity due to non-uniformity of thetransverse shear stress distribution is small, and thus the feedback is relatively weak (Figure 2). As the average shearstress approaches the critical value for entrainment, the strength of the feedback increases rapidly. Figure 9 plots theinitial specific discharge (q) against the time it took for the stream table experiments to reach equilibrium, which mustbe related in some general way to the strength of the feedback mechanism. Since the initial channel slope (i.e. Sv) andthe bed material grain size distribution were kept constant, q is a surrogate for shear stress, with the threshold for bedentrainment from a flat bed estimated to be about 0·6 litres per second per decimeter (L/s/dm). There is a strongpositive correlation between the time required to establish equilibrium and the initial q. For the longest runningexperiment (1–3), increases in channel sinuosity did not occur until around 10 hours, when the sediment transport rateand, by inference, average shear stress were substantially reduced, and the strength of the feedback increased. It tooka further 14 hours to reach equilibrium. In contrast, when the initial q was near threshold, equilibrium configurationswere reached after one or two hours.

Another perspective supposes that the higher rates of sediment throughput associated with higher mean shear stressshould result in a more rapid approach to equilibrium. But it is clear that the lateral change in channel morphology implicitin our feedback mechanism requires divergence of the sediment flux – deposition must be possible within the channel.When the mean shear stress is high and the mean step length is long, relative to channel width, then there can be rela-tively little divergence of the sediment flux. As a result, construction of an alluvial morphology by deposition is verylimited, even though substantial morphologic modification may occur due to net degradation. It is only when the meanshear stress approaches the value for entrainment that divergence (and deposition) readily occurs. Thus, while highermean shear stress results in a higher sediment throughput, it produces a lower rate of constructional morphologic change.

Figure 9. Time to equilibrium versus initial specific discharge for stream table experiments.

Page 13: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

Conceptual model for meander initiation 887

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

Longitudinal slope variationsAnother testable prediction of our model pertains to the distribution of local energy slope along the channel. Theinferred difference in β between bend apices and crossovers implies that a higher energy slope must exist at thecrossovers if they are to transport the same sediment load. The water surface slopes for three experiments reported byEaton and Church (2004) are presented in Figure 10. Water surface elevations are used instead of total hydraulic head

Figure 10. Water surface slopes along the thalweg for three equilibrium channel configurations. The minimum slope in each bendis shown in bold italics, and the maximum is shown in bold. Sv is 1·09 for these experiments. The direction of flow is from right toleft.

Page 14: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

888 B. C. Eaton, M. Church and T. R. H. Davies

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

because there is a greater density of these data since the total hydraulic head can only be estimated at the crossoversand the apices (where there are velocity estimates). However, as mentioned above, the mean velocities at the crosso-vers and apices for these experiments were not significantly different, suggesting that the approximation of energygradient using water surface slope is appropriate.

The experiments represent a range of thalweg sinuosities and initial transport intensities. For the higher thalwegsinuosity channels (experiments 1–1 and 1–3, Figure 10), the pattern is relatively clear: the lowest slopes are associatedwith the thalweg apices and the highest slopes are associated with the entrances to and exits from the bends. Thispattern is also evident for a lower sinuosity equilibrium channel (experiment 1–6), but there are notable exceptions tothe general rule. This most likely implies failure of the simple reasoning that relates β to position along the thalwegtrace. The details associated with the individual bedforms that undoubtedly affect the transverse distribution of shearstress seem to become relatively more important as the sinuosity decreases.

Interestingly, the typical pattern of slope variation exhibited in these experiments is also evident in fixed bank,mobile bed experiments. Hooke (1975) discusses an experiment conducted in a channel bend based on a sine-generated curve with a mobile sand bed. In the pool downstream of the bend apex (where helicoidal flow strength isgreatest) the energy slope at all of his experimental discharge values is less than over the riffles upstream anddownstream of the pool. The magnitude of difference in surface slopes is similar to the differences between themaximum and minimum slopes shown in Figure 10 for our experiments.

Hooke (1975) presents maps of shear stress variance which, for a discharge of 50 L/s, show that over the pooland adjacent bar, 33 per cent of the bed experiences flow less than 50 per cent of the average shear stress (1), while29 per cent of the bed experiences shear stresses greater than 150 per cent of the average. In contrast, only 34 percent of the bed at the riffles experience shear stresses outside the range 0·5(1) ≤ τ ≤ 1·5(1), demonstrating that thereis indeed a systematic difference in the shear stress variance with channel morphology. This pattern holds over a rangeof discharge values: for a discharge of 20 L/s, Hooke’s maps indicate that only 40 per cent of the riffle experiencedshear stresses outside the range defined above, while 51 per cent of the pool–bar had shear stresses outside therange.

Discussion and Conclusions

By changing the scale of inquiry from the reach scale to the scale of individual morphologic units, one can gainan understanding of the optimality criterion necessary to complete the regime formulation presented by Eatonet al. (2004), and evident in the graded channel response documented by Eaton and Church (2004). Eaton et al.(2004) interpreted the optimality criterion as a maximization of the system-scale flow resistance, which is conceptu-ally equivalent to a potential energy well (cf. Huang et al., 2004). When the reach-scale resistance component isdominant, system-scale resistance is maximized by a feedback between the transport capacity for a section andthe variance of the transverse shear stress distribution that reduces the channel slope. Slight initial variations in theshear stress distribution result in local net scour, typically along the banks, which increases the local transport capa-city by increasing the variance of the shear stress distribution. This process leads to further local net scour in a posi-tive feedback (Figure 4). The local net scour may be primarily vertical until the channel banks begin to fail, at whichpoint bank erosion initiates a negative feedback that increases channel sinuosity (Figure 4). If the reach-averagesediment throughput is to remain constant (that is, if equilibrium is to be regained), then the local channel slope atthe locus of most intense feedback (the attractor) must decrease since more material is leaving the area than is enter-ing it.

Once one initial bend develops, others necessarily follow, because of the non-uniform cross-sectional distribution ofsediment transport. That is, the local increase in sediment transport resulting from the net scour at an asymmetriccross-section will induce significant divergence in the sediment flux field, expressed as periodic deposition down-stream on alternate sides of the channel. Order is generated in the system by alternating loci of erosion and depositionon either side of the channel. The longitudinal scale for this cycle can be approximated in a physically meaningfulway as the length scale for the diffusion of a sediment wave across the channel, moving laterally at a speed propor-tional to that for a gravity wave. The initiation of meandering, then, is proposed to be related to the characteristics ofthe sediment transport field. The secondary flow circulation that is closely related to the meandering pattern is thoughtto be a trailing phenomenon that develops as a consequence of the sediment transport field dynamics, but whichultimately influences the equilibrium channel form (after Parker, 1976).

Based on the feedback process, we can identify two features of an equilibrium produced this way, beyond thedefinitive equality of local transport capacity and sediment supply. First, the cut-bank at the bend apex must becritically stable at formative discharge. In fact, this may be an appropriate way of defining the formative discharge, at

Page 15: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

Conceptual model for meander initiation 889

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

least with respect to the channel planform. A reasonable (and testable) hypothesis claims that cut-banks are near thethreshold of motion for bankfull flows in alluvial channels, which may be the most effective means of estimating bankstrength for structurally complex stream banks. In any case, we argue that bank stability is a key feature of equilib-rium, as have many previous researchers (e.g. Millar and Quick, 1993; Griffiths and Carson, 2000).

In comparison, the discharge responsible for bed surface texture or in-channel bedforms is probably different fromthat controlling the channel pattern. That is, channel morphology is best thought of as the result of a range of flows,each effecting different adjustments: we have been concerned herein only with the flows that initiate and maintain thechannel planform.

The second point relates to the shear stress distribution. A channel formed in erodible material has some maximumlimit to the variance in the shear stress that can be sustained without the bed failing, just as banks can withstandonly a certain shear stress before bank retreat occurs. That is, there is some configuration at which the variance ismaximized and beyond which the boundary will fail, thereby reducing the variance. In our framework, this is equiva-lent to maximizing β by maximizing Yo, subject to the strength of the channel boundary. However, since – for aspecified slope – an increase in variance will produce an increase in TC, the system will tend to exhibit a maximumvariance consistent with cross-sectional stability. If the variance is not maximized, then there is still scope forthe positive feedback shown in Figure 4 to act on the system, causing further deformation and increasing β. Thereforethe equilibrium channel configuration will be characterized by cut-banks at the apices that are at the threshold ofmovement and by a cross-sectional shape that exhibits the maximum variance that can be supported by the bedmaterial.

However, since crossovers between two apices will have a lower β, the feedback mechanism described aboveimplies that water surface slope must vary systematically downstream in order to maintain a steady-state mass fluxthroughout the system. Experimental observations from stream table experiments and from fixed bank physical modelssupport this deduction.

The theory also predicts that the strength of the feedback mechanism falls off rapidly as excess shear stressincreases. This is consistent with the observed time-to-equilibrium for laboratory experiments for a range of initialshear stresses, which proceed more quickly at lower shear stresses. The relation between relative variance, transportcapacity and shear stress also has implications for the magnitude of the variation in the energy gradient for differentsystems. In streams that are always near threshold, which is true of most gravel-bed streams, the effect of variance onthe transport capacity is relatively large, and there ought to be relatively large differences in the energy gradient overthe pools and the riffles. As discussed above, we certainly see such differences in laboratory models of this kind ofstream. However, in stream channels with very high dimensionless shear stress, such as sand-bed channels withcohesive banks, the relative effect of shear stress variance on the transport capacity will be much smaller and,presumably, so too will be the relative change in energy slope along the stream channel.

In all of the experiments reported by Eaton and Church (2004), the equilibrium channel form was the result of areduction of the channel slope following the development of alternate bars. The equilibrium slope achieved is func-tionally determined by the sediment concentration, while other possible adjustments, such as changes in the resistancecoefficients (e.g. Manning’s n), could not be identified. This slope minimizing behaviour is the product of the initia-tion and development of pools on alternating sides of the channel. In this way, a three-dimensional interactionamongst channel morphology, local hydraulics and sediment transport responsible for pool growth informs the one-dimensional properties of the fluvial system, and produces the observed SMB. SMB is consistent with the cross-scalelinkage described by the feedback mechanism that occurs between the width-scale variables Yo and β and the length-scale variables L* and S.

Since SMB is similar to the behaviour predicted by the minimum slope hypothesis (and by extension, all analogousoptimality criteria), the process–form interactions embodied in the feedback mechanisms suggest a physically based,rational interpretation of what such hypotheses actually represent. Regime models are one-dimensional, and since theycannot explicitly consider the transverse distribution of shear stress, they cannot distinguish which solution among themany possible solutions admitted by the model is actually stable in a three-dimensional fluvial system. That is, theycannot explicitly describe the self-forming action of the fluvial system. The selection of a single, stable solution isaccomplished by application of an optimality criterion that facilitates the description of a three-dimensional reality bya one-dimensional model.

Acknowledgements

This work has benefited from comments made by Trevor Hoey, Gerald Nanson, Jim Pizzuto, Peter Wilcock and, especially, RobFerguson. We thankfully acknowledge funding provided by the Natural Sciences and Engineering Research Council of Canadathrough M. Church’s discovery grant and B. Eaton’s post-graduate scholarship.

Page 16: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

890 B. C. Eaton, M. Church and T. R. H. Davies

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

References

Allen JRL. 1994. Fundamental properties of fluids and their relation to sediment transport processes. In Sediment Transport and DepositionalProcesses, Pye K (ed.). Blackwell: Oxford: 25–60.

Anderson AG. 1967. On the development of stream meanders. Proceedings of the 12th Congress. International Association of HydraulicResearch: Fort Collins; 370–378.

Andrews ED. 1979. Scour and fill in a stream channel, East Fork River, Western Wyoming. Professional Paper 1117. United StatesGeological Survey: Washington.

Ashmore P. 1991. Channel morphology and bed load pulses in braided, gravel-bed streams. Geografiska Annaler 73: 37–52.Blondeaux P, Seminara G. 1985. A unified bar-bed theory of river meanders. Journal of Fluid Mechanics 157: 449–470.Carling PA. 1991. An appraisal of the velocity-reversal hypothesis for stable pool-rifle sequences in the River Severn, England. Earth

Surface Processes and Landforms 16: 19–31.Chang HH. 1979. Minimum stream power and river channel patterns. Journal of Hydrology 41: 303–327.Church M, Hassan MA, Wolcott JF. 1998. Stabilizing self-organized structures in gravel-bed stream channels: Field and experimental

observations. Water Resources Research 34(11): 3169–3179.Clifford NJ, Richards KS. 1992. The reversal hypothesis and the maintenance of riffle-pool sequences: a review and field appraisal. In

Lowland Floodplain Rivers: Geomorphological Perspectives, Carling PA, Petts GE (eds). John Wiley & Sons: Chichester; 43–70.Davies TRH, Sutherland AJ. 1983. Extremal Hypotheses for River Behavior. Water Resources Research 19(1): 141–148.Dietrich WE, Smith JD. 1983. Processes controlling the equilibrium bed morphology in river meanders. In River Meandering, Elliott CM

(ed). ASCE: New York; 759–769.Dietrich WE, Kirchner JW, Ikeda H, Iseya F. 1989. Sediment supply and the development of the coarse surface layer in gravel–bedded

rivers. Nature 340: 215–217.Eaton BC, Church M. 2004. A graded stream response relation for bedload dominated streams. Journal of Geophysical Research – Earth

Surface 109(F03011). DOI: 10.1029/2003JF000062.Eaton BC, Church M, Millar RG. 2004. Rational regime model of alluvial channel morphology and response. Earth Surface Processes and

Landforms 29(4): 511–529.Einstein HA, Shen HW. 1964. A study on meandering in straight alluvial channels. Journal of Geophysical Research 69(24): 5239–5247.Engelund F, Skovgaard O. 1973. On the origin of meandering and braiding in alluvial streams. Journal of Fluid Mechanics 57(2): 289–302.Ferguson RI. 2003. The missing dimension: effects of lateral variation on 1-D calculations of fluvial bedload transport. Geomorphology

56(1–2): 1–14.Fredsoe J. 1978. Meandering and braiding of rivers. Journal of Fluid Mechanics 84(4): 609–624.Friedkin JF. 1945. A laboratory study of the meandering of alluvial rivers. US Waterways Engineering Experimental Station: Vicksberg.Fujita Y. 1989. Bar and channel formation in braided streams. In River Meandering, Ikdea S, Parker G (eds). American Geophysical Union:

417–462.Fukuoka S. 1989. Finite amplitude development of alternate bars. In River Meandering, Ikdea S, Parker G (eds). American Geophysical

Union: 237–265.Garcia M, Nino Y. 1993. Dynamics of sediment bars in straight and meandering channels: experiments on the resonance phenomenon.

Journal of Hydraulic Research 31(6): 739–761.Gessner FB, Jones JB. 1965. On some aspects of fully-developed turbulent flow in rectangular channels. Journal of Fluid Mechanics 23(4):

689–713.Griffiths GA, Carson MA. 2000. Channel width for maximum bedload transport capacity in gravel-bed rivers, South island, New Zealand.

Journal of Hydrology (NZ) 39(2): 107–126.Hey RD. 1976. Geometry of river meanders. Nature 262: 482–484.Hooke R. 1975. Distribution of sediment transport and shear stress in a meander bend. Journal of Geology 83(5): 543–565.Huang HQ, Chang HH, Nanson GC. 2004. Minimum energy as the general form of critical flow and maximum flow efficiency and for

explaining variations in river channel pattern. Water Resources Research 40(W04502). DOI: 10.1029/2003WR002539.Ikeda S. 1984. Prediction of alternate bar wavelength and height. Journal of Hydraulic Engineering 110(14): 371–386.Johannesson H, Parker G. 1989. Linear theory of river meanders. In River Meandering, Ikeda S, Parker G (eds). Water Resources Mono-

graph. American Geophysical Union: 181–213.Kirkby MJ. 1977. Maximum sediment efficiency as a criterion for alluvial channels. In River Channel Changes, Gregory KJ (ed.). John

Wiley & Sons: Chichester; 429–442.Lane EW. 1955. Design of stable channels. American Society of Civil Engineers, Transactions 120: 1234–1260.Leopold LP, Wolman MG, Miller JP. 1964. Fluvial Processes in Geomorphology. WH Freeman: San Fransico.Lewin J. 1976. Initiation of bed forms and meanders in coarse-grained sediment. Geological Society of America Bulletin 87: 281–285.Lisle T. 1979. A sorting mechanism for a riffle-pool sequence: summary. Geological Society of America Bulletin 90: 616–617.Meyer-Peter E, Müller R. 1948. Formulas for bed-load transport. International Association for Hydraulic Structures Research, 2nd Meeting,

Proceedings. Stockholm.Millar RG, Quick MC. 1993. Effect of bank stability on geometry of gravel rivers. Journal of Hydraulic Engineering – ASCE 119(12):

1343–1363.Nikora V, Habersack H, Huber T, McEwan I. 2002. On bed particle diffusion in gravel bed flows under weak bed load transport. Water

Resources Research 38(6). DOI: 1081, 1010.1029/2001WR000513.

Page 17: Earth Surface Processes and Landforms Conceptual model for ... · a physically based conceptual model for the initiation of meandering in straight, bedload-dominated streams as a

Conceptual model for meander initiation 891

Copyright © 2006 John Wiley & Sons, Ltd. Earth Surf. Process. Landforms 31, 875–891 (2006)

Paola C, Seal R. 1995. Grain-size patchiness as a cause of selective deposition and downstream fining. Water Resources Research 31(5):1395–1407.

Parker G. 1976. On the cause and characteristic scales of meandering and braiding in rivers. Journal of Fluid Mechanics 76(3): 457–480.Parker G, Anderson AG. 1975. Modelling of meandering and braiding in rivers. Proceedings of the ASCE Modelling Symposium. American

Society of Civil Engineers: San Francisco; 575–591.Parker G, Klingeman PC. 1982. On why gravel bed rivers are paved. Water Resources Research 18(5): 1409–1423.Parker G, Dhamotharan S, Stefan H. 1982. Model experiments on mobile, paved gravel bed streams. Water Resources Research 18(5):

1395–1408.Pyrce RS, Ashmore PE. 2003. Particle path length distributions in meandering gravel-bed streams: results from physical models. Earth

Surface Processes and Landforms 28: 951–966.Rathbun RE, Guy HP. 1967. Measurement of hydraulic and sediment transport variables in a small recirculating flume. Water Resources

Research 3(1): 107–122.Rhoads BL, Welford MR. 1991. Initiation of river meandering. Progress in Physical Geography 15(2): 127–156.Schumm SA, Khan HR. 1972. Experimental study of channel patterns. Geological Society of America Bulletin 83: 1755–1770.Smith CR. 1996. Coherent flow structures in smooth-wall turbulent boundary layers: facts mechanisms and speculation. In Coherent Flow

Structures in Open Channels, Ashworth PJ, Bennett SJ, Best JL, McLelland SJ (eds). John Wiley & Sons: Chichester: 1–39.Smith CE. 1998. Modeling high sinuosity meanders in a small flume. Geomorphology 25: 19–30.Sun T, Meakin P, Jossang T. 2001. A computer model for meandering rivers with multiple bed load sediment sizes 1. Theory. Water

Resources Research 37(8): 2227–2241.Thompson A. 1986. Secondary flows and the pool–riffle unit – a case-study of the processes of meander development. Earth Surface

Processes and Landforms 11(6): 631–641.Tominaga A, Ezaki K, Nakagawa H. 1989. Three-dimensional turbulent structure in straight open channel flows. Journal of Hydraulic

Research 27(1): 149–173.Tracy HJ. 1965. Turbulent flow in a three-dimensional channel. Journal of the Hydraulics Division – ASCE 9(HY6): 9–35.Venditti JG, Church MA, Bennett SJ. 2005. Bed form initiation from a flat sand bed. Journal of Geophysical Research – Earth Surface

110(F1). DOI: 10.1029/2004JF000149.Werner PW. 1951. On the origin of river meanders. Transactions, American Geophysical Union 32(6): 898–902.White WR, Bettess R, Paris E. 1982. Analytical approach to river regime. Journal of the Hydraulics Division – ASCE 108(HY10):

1179– 1193.Wilcock PR, McArdell BW. 1993. Surface-based fractional transport rates: mobilization thresholds and partial transport of a sand-gravel

sediment. Water Resources Research 29(4): 1297–1312.Yalin SM. 1971. On the formation of dunes and meanders. Proceedings of the 14th Congress of the International Association of Hydraulic

Research: 1–8.Yang CT. 1976. Minimum unit stream power and fluvial hydraulics. Journal of the Hydraulics Division – ASCE 102(HY7): 769–784.