Sie sind auf Seite 1von 8

River, Coastal and Estuarine Morphodynamics: RCEM 2007 Dohmen-Janssen & Hulscher (eds)

2008 Taylor & Francis Group, London, ISBN 978-0-415-45363-9

Effects of particle exposure, near-bed velocity and pressure fluctuations


on incipient motion of particle-size mixtures
Stefan Vollmer
Federal Institute of Hydrology, Department River Morphology, Koblenz, Germany

Maarten G. Kleinhans
Universiteit Utrecht, Fac. Geosciences, Dept. Physical Geography, Utrecht, The Netherlands

ABSTRACT: For the prediction of river bed destabilisation and fractional sediment transport of mixtures, we
aim to solve two problems that are poorly understood. First, the flow and pressure fluctuations surrounding both
the embedded and exposed particles must be parameterised for hydraulically smooth to rough flow. Second, an
adequate relation between particle size and particle exposure should be based on the particle size distribution and
the (water-worked) bed structure. We use a recently developed force balance model for the threshold of motion
of uniform sediments incorporating the effects of particle exposure, pressure fluctuations into the bed, very
shallow flow and bed slope. The flow module is extended to non-uniform roughness of sediment mixtures. Our
extended model predicts the critical Shields values of arbitrary mixtures directly as function of exposure and no
longer needs empirical hiding-exposure relations. Several empirical and geometrical relations between particle
size and exposure were tested. The results are compared to extensive datasets from the literature of incipient
fractional transport rates. The modelled hiding-exposure relations are very sensitive to the relation between
particle size and exposure, which differ for unimodal, skewed and bimodal mixtures. This is explained by the
pore structure of these sediments. The existing relations fail particularly for the smaller particles in bimodal and
skewed distributions. These small particles percolate through the pores so their exposure or embedding strongly
depends on the fractional content and pore structure, in agreement with empirical data. We are working on a
universal relation for exposure containing particle size distribution, pore structure and water-working. The model
reproduces data of uniform sediments well for the entire physically possible range of particle exposures and for
hydraulically rough to nearly smooth conditions. Trends in existing data for mixtures are also reproduced but
depend strongly on exposures that were not measured.

INTRODUCTION

parameters is limited, after which conclusions are


drawn.

The aim of this paper is to develop a model for the


prediction of the threshold of motion of particle-size
mixtures. Prediction of incipient motion of sediment
mixtures is presently limited to empirically derived
equations. First we review the relevant parameters
to predict incipient motion of uniform sediments by
physics-based modelling. Then we review the current
empirical knowledge of incipient motion of natural
mixtures of various particle sizes. We then extend a
recently developed force balance model for the threshold of motion of uniform sediments to sediment mixtures. Our extended model predicts the critical Shields
values of arbitrary mixtures directly as function of
exposure and no longer needs empirical relations. The
discussion is used to validate and verify the model to
demonstrate where current knowledge of the relevant

REVIEW

The classical empirical study by A.F. Shields demonstrated that incipient motion of sands and gravels of
all densities occurs in a narrow range of a nondimensional ratio of the sediment entraining and detraining
forces:

where c = critical shear stress, s and are the densities of the sediment and water, g = gravitational acceleration and D = particle size. This critical Shields

541

Shields number *cr

10

angles and particle exposures to the flow. As a result,


the model predicts nearly equal mobility of all sizes;
that is, all particles are entrained at the same shear
stress. The reason is that small particles are hidden in
the lee of large particles, which are more exposed to
the flow. This agrees with data and a model for bed
surface armouring by Parker and Klingeman (1982).
However, later investigations showed that incipient motion of bimodal sediments differs from equal
mobility (Wathen et al., 1995; Wilcock 1993; Wilcock
2003; wu and Yang 2004). The sand mode behaves
equally mobile but incipient gravel motion is slightly
size-dependent. A simple model by Egiazaroff (1965)
explained this based on a logarithmical velocity profile and a simple geometrical assumption for the hiding
and exposure of different particle sizes:

Soulsby
Zanke
Wiberg Smith
this paper
data

10

10

10

10
10
grain size D (mm)

10

*cr model

10

10

10

*cr data

Figure 1. A. Comparison between the Soulsby (1997) fit


and the models by Zanke (2003), Wiberg and Smith (1987)
and Vollmer and Kleinhans (2007) models. The dataset was
compiled by Buffington and Montgomery (1997). The lowest points are by Coleman (1967, in vollmer and Kleinhans
2007) for maximum grain, exposure. B. Model predictions
compared to data for individual data points including slope
and depth effects. The lines are perfect agreement and a factor
of 1.5 deviation.

number has a value of about 0.057 for fine sand and


gravel, a somewhat smaller value for medium sand and
an increasing value for finer sediment (Fig. 1).
Wiberg and Smith (1987) reduced the Shields curve
partly to physics of flow around spherical particles.
The elements of their model are pivoting angles of
naturally packed particles, drag and lift forces with
independently derived coefficients, and a universal
velocity profile for the entire range of hydraulically
smooth to rough flow. The dip in the Shields curve
coincides with the transition between smooth and
rough 3.5 < Re < 30, where

where u = shear velocity and = fluid viscosity. To


extend the model to sediment mixtures wiberg and
Smith (1987) employed empirical relations of pivot

for Di /Dm > 0.3; for Di /Dm < 0.3Ashida and Michiue
(1971) proposed a correction: ci = 0.07, wherein
Di = fractional and Dm = geometric mean particle
size. The Egiazaroff model cannot account for variations in the particle size distribution, however, which
has large effects on the effective hiding function (Van
der Zwaard 1974; Shvidchenko et al., 2001). In fact,
the exposure of the fine mode of bimodal sediment
strongly depends on its relative abundance. Starting
from a gravel bed with few fines deeply embedded, adding fine sediment will increasingly fill the
pores of the coarse sediment until the coarse sediment is entirely covered while the fine sediment is fully
exposed. At the same time the decreasing roughness
of the bed affects the mean shear stress and the turbulence characteristics (Van der zwaard 1974; Sambrook
smith and Nichols, 2005). So far no model is available
incorporating these effects, but the problem has relevance for downstream fining and for aquatic ecology,
e.g. spawning of salmon.
Kleinhans and Van Rijn (2002); Zanke (2003);
Hofland (2005) and many earlier workers showed that
near-bed turbulence increases the mobility. The largest
shear stresses in the turbulent flow field entrain the
sediment. Hence the original Shields curve refers to
generalised sediment motion while individual particle motion is attained at smaller Shields values.
vollmer et al., (2002) measured and parameterised
turbulence-induced pressure fluctuations in the porous
bed. vollmer and Kleinhans (2007) showed that the
pressure fluctuations cause lift forces on the particles
that contribute to initial motion even if particles are
flush embedded and not exposed to the attacking nearbed flow velocity. Their model performed well to the
entire range of possible particle exposure: from deeply
embedded to nearly fully exposed. Verification on a
dataset for a large range of particle sizes (Buffington

542

and Montgomery, 1997) and a large range of exposures


gave good results (Fig. 1). The Vollmer and Kleinhans
(2007) model is the only available model that accounts
for this range of exposures and for in-bed pressure fluctuations and is therefore the most suitable to extend to
sediment mixtures.
Shvidchenko et al. (2001) systematically experimented with mixtures of varying standard deviation
and skewness and measured sediment transport rates
for systematic combinations of depth, slope and discharge. They analysed the data by fitting an empirical
transport function to the nondimensional transport rate
against nondimensional shear stress to interpolate or
extrapolate to a reference transport rate. The resulting
reference shear stress indicates incipient motion in a
consistent manner between datasets and has similar
values as the critical shear stress for the beginning of
motion. However, Shvidchenko et al (2001) include
an additional empirical slope factor in the fitted transport function so that it becomes difficult to distinguish
between the effects on reference shear stress of this
slope factor, the bed slope and the water depth.
Therefore we reanalysed the whole dataset entirely
according to the procedure described by Parker and
Klingeman (1982) with empirical transport functions
without a slope effect. The model also will be compared to the well-known hiding functions of Egiazaroff
(1965) (adjusted) and Parker and Klingeman (1982)
(equal mobility).
3
3.1

MODEL DESCRIPTION
Model for uniform sediment

The model equation, based on the momentum balance, is:

where = bed slope, = pivot angle of the centre


of the particle, / = pivot angle corrected for the
point of direct flow attack (see Vollmer and Kleinhans 2007, for details), keff = 1.4 is the effective
instantaneous shear stress associated to the average
shear stress (eff = keff avg ) that entrains the particles
(Zanke 2003), cD and cL are drag and lift coefficients
(see vollmer and Kleinhans, 2007, for constitutive
equations) kturb accounts for the turbulence-induced
pressure fluctuations in the bed (discussed later), kstd
accounts for the direct, steady flow attack.
The model equation is solved by integrating the drag
and lift forces of the direct flow over the height of the
particle that is exposed to the flow, integrating the lift
forces due to pressure fluctuations over the entire particle (also the embedded ones) and calculating the pivot
angle related to the exposure and size of the particle.

Of additional importance for model verification is the


combined effect of bed surface slope (e.g. wiberg and
Smith 1987) and shallow flow, which leads to larger
critical Shields numbers in steep and shallow mountain gravel bed streams and sandy laboratory models,
fans and deltas. This effect is incorporated in the model
through the first (slope) term and kstd (see vollmer and
Kleinhans, 2007, for details).
We will assume that the model described above is
suitable for beds of a range of grain sizes, which leaves
the problems of 1) finding an appropriate velocity profile for beds of variable, natural roughness for the
entire range of hydraulically smooth to rough flow, 2)
parameterising the lift forces on embedded grains
induced by pressure fluctuations that are caused by turbulence, and 3) specifying the exposure of the grains
of different sizes depending on the size distribution.
The exposure is defined as E = P/Dm where the protrusion is the height of the top of the grain above the
surrounding bed, which is defined at 0.25Dm below
their tops.
3.2 Velocity profile for steady flow attack
The attacking forces, calculated from integration of
the flow velocity profiles over the particle height,
have been included in several Shields curve models
for uniform and heterogeneous sediments. For example, wiberg and Smith (1987) extended the logarithmic
velocity profile for rough flow down to z = z0 = ks /30.
For the hydraulically smooth flow they corrected the
velocity profile near the bed and used this formulation
also for the transitional flows. The resulting Shields
curves show the well known dip at the 10 < Re < 100
and D 0.7 mm. However, the hydraulic resistance
parameters (Chzy coefficient C or Darcy-Weissbach
friction factor f ), do not show a dip for natural bed
roughness such as for heterogeneous sediments (Zanke
2006, in Zanke 2003). Several authors proposed formulations of the hydraulic resistance parameters C or
f , which represent the ratio of depth-averaged flow
velocity (averaged over time and water depth) to shear
velocity um /u. These formulations hold for the entire
Re range and show a more gradual transition between
smooth and rough flow for heterogeneous bed roughness. We use a vertical velocity profile for heterogeneous sediments that corresponds to these formulations and is similar to the transition between smooth
and rough flow for technical (non-uniform) roughness
in pipe flow after colebrook and white (1937):

where the von Krmn coefficient = 0.42. The


hydraulic roughness of the bed, necessary for the
velocity profile and the length scale of turbulent
fluctuations, is here taken constant at ks = 1.5Dm .

543

(Nikora et al., 2002; wu and yang, 2004) and others


showed that spatially averaged velocity distributions
deviate from the logarithmical function near the bed,
but concentrate solely on hydraulic rough flow. We will
extend our model with future work on this effect.
3.3

Particle lift by pressure fluctuations

The model for the critical Shields parameter of uniform


sediment is now adapted to include the effect of nearbed pressure fluctuations for sediment mixtures. This
has an advantage over existing models, in that grains
can be mobilized that are only partly, or not at all,
exposed to the flow (vollmer and Kleinhans 2007).
The decay of pressure fluctuations into the porous bed
could be described with an exponential decay function,
which is in agreement with a solution of the Laplace
equation for describing the groundwater flow regime
Vollmer et al., 2002).
Pressure variations at the riverbed are caused by
coherent eddy structures or wave-like structures that
scale with the roughness length, or the size of flow
obstacles, respectively. In open channel flow over natural beds of heterogeneous sediments exist a wide range
of sizes of coherent turbulent structures and pressure
disturbances at the bed (Hofland 2005). The dominant
wave length Ld due to the larger roughness elements is
likely to be decisive for the pressure forces on the larger
exposed particles, while other length scales (generated
closer to the bed and acting in a smaller region) are critical for the incipient motion of the smaller embedded
particles. The critical wave length of pressure disturbances is related to its grain size and has an order of
magnitude of 2 D or a multiple of D (Hofland 2005).
We propose to calculate the pressure propagation
as the exponential decay with wavelength Ld = 9.5ks
for E > 0 and Ld = 9.5Di for E < 0. This guarantees a
good transition at Di = Dm at which E 0 and agrees
with the model for uniform sediments (Vollmer and
Kleinhans, 2007) where ks = Dm .
3.4 Size- and frequency-dependent exposure
of grains
For uniform sediments there is a direct relation
between the exposure and the pivot angle (Luckner 2002 in vollmer and Kleinhans, 2007). wiberg
and Smith (1987) and Zanke (2003) simply assumed
E = 0.3 for all grain size fractions i, where E is
the absolute exposure above the average bed surface,
which is defined at level 0.25Dm below the tops of Dm .
For mixtures there are only two empirical relations available (Kirchner et al., 1990; Buffington
et al., 1992), hereafter called Kirchner or Kir and
Buffington or Buf . The Kir relations are:

Figure 2. Bed structures and the related functions for


pivot angle and protrusion p. The exposure E = p/Dm .
Buffington refers to eq. 8. and Kirchner refers to eq. 6.

for exposure, where E = P/Dm with P the protrusion


of a particle above (positive) or below (negative) the
Dm . The pivot angle is:

Buf use the same and their exposure relation is


written as:

where n = percentile of size fraction. For better understanding of these empirical relations we also study
three relations based on simplified bed structures
(Fig. 2).
The pivot angles for unimodal and bimodal sediment demonstrate the effect of the presence of fines
in the bed (Figs 3, 4). The exposures of large grains
Di > 2Dm (Figs 3, 4) hardly differ between the various
models, except for the same centre model (see Fig. 2)
which predicts half the typical numbers. The modelled
exposures for the smaller grains Di < 2Dm , on the other
hand, are very different (see inset in Figs). The exposure is the largest for small grains on Dm and the
smallest for eq. 8. There is a considerable difference
between the Kir and Buf equations. For large grains
both resemble the same centre result, illustrating that
the large grains sink about halfway into the bed rather
than be fully exposed (the latter would lead to overpassing with extremely small critical shear stresses).
For the small grains, Kir sets the grains only slightly
deeper than the same centre, while Buf puts the grains
much deeper: down to the same floor function. The
result will be a much decreased mobility for the fines

544

70

70
Buffington
Kirchner
same floor
same centre
on Dm

50 1
40
30

60
Exposure E = p/Dm ()

Exposure E=p/Dm ()

60

20 1
10

10

-4

10

10
D (m)

10

10

pivot angle i (o), p i (arbitrary)

pivot angle i (o), pi (arbitrary)

150

100

50

10

10
D (m)

10

10

20 1
10

10

-4

10

10
D (m)

10

10

RESULTS
Exposure models

The model was applied to a unimodal mixture with


standard deviation 1 and Dm = 1 mm. All model results
show a kink near E = 0 where the direct drag force
of the flow disappears but the turbulence-induced lift
emerges. In reality this transition would be more gradual so the kink can be ignored. The effects of the various exposure and pivot angle models are dramatically
different (Fig. 5), while the difference between unimodal and bimodal sediment (not shown) is negligible
compared to this.
The higher levels of fine grains in Kir (compared
to Buf) exposes these grains to the turbulence-induced
pressure fluctuations propagating into the bed. Hence,
the mobility of fines with Kir is much larger than
with Buf where grains Di < 0.2Dm become immobile. The Kir and Buf functions give critical shear
stresses in between the Egiazaroff curve and equal

Buffington/Kirchner
same floor
same centre
on Dm
size distrib pi

150

100

50

0
5
10

according to Buf. For the bimodal sediment the results


are nearly the same except that the large peak of fines
leads to slightly larger exposures of the fines in Buf.

4.1

30

200

Buffington/Kirchner
same floor
same centre
on Dm
size distrib p

Figure 3. Exposure E (A) and pivot angle (B) of various


functions for a unimodal grain size-distribution. Cross-hairs
in inset are at E = 0 and D = Dm .

40

-10
5
10

200

0
5
10

50 1

0
-10
5
10

Buffington
Kirchner
same floor
same centre
on Dm

10

10
D (m)

10

10

Figure 4. Exposure E (A) and pivot angle (B) of various


functions for a bimodal grain size-distribution.

mobility, which is realistic, but do not clearly exhibit


a dependence on the grain size-distribution.
Grains lying on Dm are much more exposed with
smaller pivot angles than other scenarios and are much
more mobile. The grains having the same centre on
the other hand are much less exposed and mobile. In
between these two scenarios the Kir and Buf results
and the Egiazaroff curve are found. The scenario with
grains lying on the same floor hardly has any mobile
grains Di < Dm , while coarser grains are more mobile
than the Dm , which is unrealistic.
4.2 Comparison to Shvidchenko et al. (2001) data
The model is tested on the empirical hiding functions
derived from the Shvidchenko et al. data. The data are
averaged for each sediment mixture (Fig. 6), which
removes effects of slope and water depth (not shown
here). The experimental results for the unimodal sediments demonstrate no significant effect of the width
of the grain size-distribution (Fig. 7). The shape of
the distribution on the other hand has a large effect,
which is partly due to the differences between the Dm of
the mixtures and probably partly due to the bed structure. The coarse skew sediment with a fine tail has
a decreased mobility for the larger grain whereas the

545

0.5

10

narrow
wider
wide

0.4
1

10

pi

cr,i

0.3

0.2

10

0.1
10

10

10

10

10

10

0
3
10

10

cr,i

10

fine skew
coarse skew
bimodal

0.4

0.3
pi

10

0.5

Buffington
Kirchner
same floor
same centre
on Dm
Egiazaroff
Ashida Michue

10

10

0.2

0.1
3

10

10

10

10
D (m)

10

10

0
3
10

10

D (m)

Figure 5. Hiding functions for unimodal sediment predicted


by the model with various pivot angle and exposure functions
(see Fig. 3). A. Dimensional shear stress (c for equal mobility
would plot as horizontal line). B. Nondimensional shear stress
(c for uniform sediment would plot as horizontal line).

fine skewsediment with a coarse tail has an increased


mobility for the larger grains.
The model captures the general trend seen in the
data. The predictions for the different skewed and
bimodal mixtures lie at the same relative heights in
the graphs (Fig. 7b). The fact that the modelled values
are larger than observed in all cases is due to the different definition of critical and reference shear stress
and can be ignored. The model is rather sensitive to
the presence of a sharp peak in the grain size distribution (see narrow mixture) which seems unrealistical.
This effect is entirely due to the Kir exposure and pivot
angle predictors.

DISCUSSION

Our model is more based on physics of near-bed flow


and grain pivoting than earlier models, and predicts
differences in ci in general agreement with a detailed
dataset (Shvidchenko et al., 2001) and with the two
paradigms of hiding and exposure: the Egiazaroff
function and equal mobility (Parker and Klingeman,

Figure 6. Grain size-distributions of the Shvidchenko et al.


(2001) mixtures and the approximations for the model by
addition of two lognormal distributions.

1982). The reason is that the new model can predict incipient motion for various exposures above and
below the bed with appropriate, independently derived
drag and lift coefficients. For individual grains this
approach is well verified (Fig.1) (Vollmer et al., 2002;
vollmer and Kleinhans, 2007). Semi-empirical relations for the velocity profile over rough beds, for the
effect of low submergence (h/Dm ) in shallow flow
and for turbulent pressure fluctuations over and in the
bed can be improved or replaced in the future pending further work on double-averaging of the Reynolds
equations (Nikora et al., 2002).
For mixtures, a description of the exposures and
their related pivot angles of individual grain sizes is
necessary. The effects of this bed structure on the
beginning of motion are very large. The advance of
insight by the model presented in this paper is therefore not that the empirical hiding functions can now
be replaced, but in pointing out the missing of knowledge on the bed structure of mixtures. In the future
the model may predict hiding functions for different
mixtures on a physical basis rather than an empirical
as has been common practice for half a century.
Simple bed structures were shown to be inadequate
as empirical results clearly transitioned between two

546

of such a bed will change the size-distribution of the


surface sediment depending on the composition of the
sediment supplied upstream (Parker and Klingeman,
1982; Kleinhans, 2005).
Type III. When much fine sediment deposits on a
gravel bed, the pores may be entirely filled so that
the bed becomes matrix-supported. The gravel may
then either become so embedded, while turbulence is
no longer generated over their tops, that their mobility dramatically decreases. On the other hand, once
entrained, the gravel may pass over the smooth sand
bed rapidly with a negligible critical shear stress.
Type III occurs in pools (Lisle and Hilton, 1999)
and highly bimodal sediments with a separate source
for fines (e.g. experiments of Van der Zwaard 1974)
and greatly affects the presence and morphology of
bedforms (Kleinhans et al., 2002); a complexity not
present in the shallow flow data of Wilcock and Crowe
(2003). Moreover, when the fines are present on the
bed they may fill up the pores of the coarse sediment to
such extent that the turbulence associated to the coarse
sediment is no longer generated (Sambrook smith and
Nicholas 2005). It is not known whether the entire filling of pore space by fines will affect the propagation of
pressure fluctuations into the bed, which we neglected
so far.

10

narrow
wider
wide
Egiazaroff
1

cr,i

10

10

10

10

10

fine skew
coarse skew
bimodal
Egiazaroff
1

cr,i

10

10

10

10

D (m)

6
Figure 7. Nondimensional reference shear stress of the
Shvidchenko et al. (2001) data and predicted by the model
with Kir.

types of simple structures. Two empirical bed structures were shown to give very different results while
their authors already warned that these functions may
only be valid for the specific beds from which they
were derived (Buffington et al., 1992).
This dependence of exposure functions on the grain
size-distribution and water-workedness of the bed is
entirely natural and must be investigated in experiments and nature. The physical basis for this dependence is illustrated by reference to three hypothetical
types of sediment for which the model will be suitable
after an appropriate exposure function has been constructed based on grain size-distribution and porosity
considerations. We are presently attempting to find an
exposure function that incorporates these three types:
Type I. A few fines settling from suspension on a
gravel bed will percolate deeply into the bed, which is
then defined as clast-supported. The negative exposure
of these grains will be too low for entrainment even by
in-bed pressure fluctuations.
Type II. Fine sediment larger than the pore size of the
clast-supported bed will not percolate into the bed but
remain near the surface. The exposures of these grains
will be large enough for entrainment. Water-working

CONCLUSIONS

We present a model for the beginning of motion of


all grain sizes in a mixture, based on the (parameterised) physics of near-bed flow, turbulence and
in-bed pressure fluctuations. The model results generally behave between the commonly used hiding
functions of Egiazaroff (1965) and equal mobility of
Parker and Klingeman (1982). Trends in the data of
Shvidchenko et al. (2001) for mixtures of various distribution shapes are moderately well predicted with the
main cause for deviations being the empirical exposure
function. We demonstrate that the model results, particularly for the fines, depend strongly on the exposure
of individual grain sizes above the bed and argue that
the physical basis for this dependence is the porosity
and pore-filling of fines depending on their proportion
and the grain size-distribution shape.

ACKNOWLEDGEMENTS
MGK is supported by the Netherlands Earth and Life
sciences Foundation (ALW) with financial aid from
the Netherlands Organisation for Scientific Research
(grant ALW-VENI-863.04.016). A. Shvidchenko is
gratefully acknowledged for discussion and providing
his data and MGK thanks Gary Parker for discussion.

547

REFERENCES
Ashida, K. and M. Michiue (1971). An investigation of river
bed degradation downstream of a dam. In Proc. of 14th
IAHR Congress, Volume C30, Paris, pp. 19. IAHR.
Buffington, J., W. Dietrich, and J. Kirchner (1992). Friction angle measurements on a naturally formed gravel
streambed: implications for critical boundary shear stress.
Water Resources Research 28(2), 411425.
Buffington, J. and D. Montgomery (1997). A systematic analysis of eight decades of incipient motion studies, with
special reference to gravel-bedded rivers. Water Resources
Research 33(8), 19932029.
Colebrook, C. F. and C. M. White (1937). Turbulent flow
in pipes with particular reference to the transition regio
between the smooth and rough pipe laws. J. Inst. Civ. Eng.
7, 9, 99118, 318400.
Egiazaroff, I. (1965). Calculation of nonuniform sediment
concentrations. J. of the Hydraulics Division, ASCE
91(HY4), 225248.
Hofland, B. (2005). Rock and Roll turbulence-induced damage to granular bed protections, Published PhDthesis.
Delft, The Netherlands: Department of Civil Engineering
and Geosciences, TU-Delft.
Kirchner, J., W. Dietrich, F. Iseya, and H. Ikeda (1990). The
variability of critical shear stress, friction angle and grain
protrusion in water-worked sediments. Sedimentology 37,
647672.
Kleinhans, M. (2005). Upstream sediment input effects on
experimental dune trough scour in sediment mixtures. J.
of Geophysical Research 110, F04S06.
Kleinhans, M., A. Wilbers, A. D. Swaaf, and J. Van den Berg
(2002). Sediment supply-limited bedforms in sand gravel
bed rivers. J. of Sedimentary Research 72(5), 629640.
Kleinhans, M. G. and L. C. Van Rijn (2002). Stochastic prediction of sediment transport in sand-gravel bed rivers. J.
of Hydraulic Engineering 128(4), 412425.
Lisle, T. and S. Hilton (1999). Fine bed material in pools of
natural gravel bed channels. Water Resources Research
35(4), 12911304.
Nikora, V., D. Goring, and A. Ross (2002). The structure and
dynamics of the thin near-bed layer in a complex marine
environment: a case study in Beatrix Bay, New Zealand.
Estuarine, Coastal and Shelf Science 54, 915926.

Parker, G. and P. Klingeman (1982). On why gravel bed


streams are paved. Water Resources Research 18(5),
14091423.
Sambrook Smith, G. and A. Nicholas (2005). Effect on flow
structure of sand deposition on a gravel bed: results from
a two-dimensional flume experiment. Water Resources
Research 41, W10405.
Shvidchenko, A., G. Pender, and T. Hoey (2001). Critical shear stress for incipient motion of sand-gravel
streambeds. Water Resources Research 37, 22732283.
Soulsby, R. (1997). Dynamics of marine sands. London, UK:
Thomas Telford Publications.
Van der Zwaard, J. (1974). Ruwheidsonderzoek Waal: invloed
van bodemtransport op de ruwheid van een vaste laag,
Volume M988-II. WLDelft Hydraulics.
Vollmer, S., F. de los Santos Ramos, H. Daebel, and G. Khn
(2002). Micro scale exchange processes between surface
and subsurface water. J. of Hydrology 269, 310.
Vollmer, S. and M. G. Kleinhans (2007). Predicting incipient
motion including the effect of turbulent pressure fluctuations in the bed. Water Resources Research accepted.
Wathen, S., R. Ferguson, T. Hoey, and A. Werritty
(1995). Unequal mobility of gravel and sand in weakly
bimodal river sediments.Water Resources Research 31(8),
20872096.
Wiberg, P. and J. Smith (1987). Calculations of the critical shear stress for motion of uniform and heterogeneous
sediments. Water Resources Research 23, 14711480.
Wilcock, P. (1993). Critical shear stress of natural sediments.
J. of Hydraulic Engineering 119(4), 491505.
Wilcock, P. and J. Crowe (2003). Surface-based transport
model for mixed-size sediment. J. of Hydraulic Engineering 129(2), 120128.
Wu, W. and S. Yang (2004). A stochastic partial transport
model for mixed-size sediment: Application to assessment of fractional mobility. Water Resources Research 40,
W04501.
Zanke, U. C. E. (2003). On the influence of turbulence on the
initiation of sediment motion. International J. of Sediment
Research 18(1), 115.

548

Das könnte Ihnen auch gefallen