ArticlePDF Available

Height and wavelength of alternate bars in rivers: Modelling vs. laboratory experiments

Authors:

Abstract and Figures

Alternate bars are large wave patterns in sandy beds of rivers and channels. The crests and troughs alternate between the banks of the channel. These bars, which move downstream several meters per day, reduce the navigability of the river. Recent modelling of alternate bars has focused on stability analysis techniques. We think, that the resulting models can predict large rhythmic patterns in sandy beds, especially if the models can be combined with data-assimilation techniques. The results presented in this paper confirm this thought. We compared the wavelength and height of alternate bars as predicted by the model of Schielen et al. [14], with the values measured in several flume experiments. Given realistic hydraulic conditions R root Re > 2*10(3), (R the width-to-depth ratio and R, the Reynolds number), the predictions are in good agreement with the measurements. In addition, the model predicts the bars measured in experiments with graded sediment. If R root Re < 2*103, the agreement between model results and measurements is lost. The wave height is clearly underestimated, and the standard deviation of the differences between predictions and measurements increases. This questions the usefulness of small flume experiments for morphodynamic problems.
Content may be subject to copyright.
JOURNAL OF HYDRAULIC RESEARCH, VOL. 39, 2001, NO. 2 147
Fig. 1. Alternate bars in a channel as measured in a straight laboratory
flume Lanzoni [9]. The crests alternate between the side banks.
(The flow is from right to left.)
Height and wavelength of alternate bars in rivers: modelling vs. laboratory experi-
ments
Hauteur et longueur d’onde des bancs alternés en rivières : la modélisation faceaux
expériences de laboratoire
M.A.F. KNAAPEN, Dep. of Civil Engineering, University of Twente, P.O. Box 217, 7500 AE Enschede, The Netherlands.
S.J.M.H. HULSCHER, Dep. of Civil Engineering, University of Twente, P.O. Box 217, 7500 AE Enschede, The Netherlands.
H.J. DE VRIEND, Dep. of Civil Engineering, University of Twente, P.O. Box 217, 7500 AE Enschede, The Netherlands.
A. VAN HARTEN, Dep. of Management Studies, University of Twente, P.O. Box 217, 7500 AE Enschede, The Netherlands.
ABSTRACT
Alternate bars are large wave patterns in sandy beds of rivers and channels. The crests and troughs alternate between the banks of the channel. These
bars, which move downstream several meters per day, reduce the navigability of the river. Recent modelling of alternate bars has focused on stability
analysis techniques. We think, that the resulting models can predict large rhythmic patterns in sandy beds, especially if the models can be combined
with data-assimilation techniques. The results presented in this paper confirm this thought.
We compared the wavelength and height of alternate bars as predicted by the model of Schielen et al. [14], with the values measured in several flume
experiments. Given realistic hydraulic conditions > 2*10³, (Rthe width-to-depth ratio and Rethe Reynolds number), the predictions are in good
R Re
agreement with the measurements. In addition, the model predicts the bars measured in experiments with graded sediment. If < 2*10³, the agree-
R Re
ment between model results and measurements is lost. The wave height is clearly underestimated, and the standard deviation of the differences between
predictions and measurements increases. This questions the usefulness of small flume experiments for morphodynamic problems.
RÉSUMÉ
Les bancs alternés sont des configurations de grandes ondulations dans les fonds sablonneux des rivières et des chenaux. Les crêtes et les creux alternent
entre les rives du chenal. Ces bancs, qui se déplacent vers l’aval de plusieurs mètres par jour, réduisent la navigabilité de la rivière. La modélisation
récente des bancs alternés s’est concentrées sur les techniques d’analyse de stabilité. Nous pensons que les modèles résultant peuvent prédire les grandes
configurations régulières dans les lits sablonneux, notamment si les modèles peuvent être combinés à des techniques intégrant des données. Les résultats
présentés dans l’article confirment cette idée. Nous avons comparé la longueur d’onde et la hauteur des bancs alternés prédites par le modèle de Schielen
et al., avec les valeurs mesurées dans plusieurs expériences en canal. Des conditions hydrauliques réalistes étant données, > 2.10³ (Rétant le rapport
R Re
de la largeur à la profondeur, et Rele nombre de Reynolds), les prédictions sont en bon accord avec les mesures. En outre, le modèle prévoit les bancs
mesurés dans les expériences avec des sédiments calibrés. Si < 2.10³ , on perd l’accord entre les résultats du modèle et les mesures. La hauteur
R Re
des ondulations est nettement sous-estimée et l’écart-type des différences entre prévisions et mesures augmente. Ceci remet en cause l’utilité des
expériences à petites échelles pour les problèmes de morphodynamique.
1 Introduction
The interaction between a non-cohesive bed and the water flow-
ing over it results in interesting phenomena. Several types of
wave patterns can be seen on the bed, each caused by a different
process. The largest bedform observed in fixed-bank rivers and
channels is termed alternate bar. Alternate bars are wave patterns,
of which the crest and trough alternate between the banks of the
channel (see Figure 1). These bars move downstream at a speed
of several meters per day. Their existence reduces the navigability
and influences the water capacity of the channel. Therefore, it is
important to predict the behaviour of these bars.
The existence of alternate bars is thought to be an inherent insta-
bility of the bed-flow system. Therefore, state of the art research
into modelling of alternate bars has focused on stability analysis
techniques [2][3][4][14][16][18]. Both Colombini et al. [4] and
Schielen et al. [14] formulate a weakly non-linear model, which
allows for small temporal variations of the amplitudes of the bars.
Only the latter model also allows for slow spatial variation.
Colombini and Tubino [3] developed a fully non-linear model.
The model of Schielen et al. [14] describes behaviour, similar to
the behaviour of alternate bars in rivers. However, to predict us-
ing their model, more lengthy and error-prone mathematical deri-
vations are necessary. Data-assimilation techniques [5] may be
the solution to avoid these derivations. Further, these techniques
can generalise the results to other large rhythmic bed waves, like
sand waves and shore parallel bars in the sea. As a start, such an
approach requires the model to estimate the characteristics of the
patterns correctly.
So far, the models have hardly been tested against either field or
laboratory data. Schielen et al. [14] used their model for a qualita-
tive analysis of the phenomenon only. Colombini et al. [4] com-
pared the results of their model with data from small-scale experi-
ments, in which the width of the flume never exceeded 0.7 m.
148 JOURNAL OF HYDRAULIC RESEARCH, VOL. 39, 2001, NO. 2
run T Q i
b
h
uC
z
F
r
L
b
H
b
[h] [m
3
/s] [%] [cm] [m/s] [m
1/2
/s] [-] [m] [cm]
uniform
sediment
P1801 816 0.03 0.16 7.3 0.27 25.2 0.32 11.3 8.5
P2403 260 0.047 0.21 8.3 0.38 28.9 0.42 4.5-7.5 5.0-6.0
P0404 192 0.04 0.2 7.7 0.35 27.8 0.4 4.3-8.0 6.0
P1505 28 0.03 0.45 4.4 0.45 32.2 0.69 10 7.0
P1605 24 0.02 0.5 3.3 0.4 31.6 0.71 11 7.7
P2709 24 0.045 0.51 5.7 0.53 30.7 0.7 9.7 4.5
P2809 24 0.04 0.52 5.3 0.5 30.4 0.7 10.6 4.7
P2909 24 0.045 0.52 5.6 0.53 31.3 0.72 9.5 4.4
graded
sediment
P0807 73 0.03 0.42 0.043 0.47 34.7 0.71 10.4 0.042
P0109 51 0.04 0.51 0.047 0.57 36.1 0.83 11.7 0.04
P1309 29 0.045 0.53 0.05 0.6 36.7 0.85 10.3 0.034
P2009 3 0.045 0.53 0.058 0.6 36.6 0.85 10.2 0.034
Table 1 Conditions and results of experiments on alternate bars by Lanzoni [9][10]. Here Tgives the duration of
the experiment, Qthe average water flux, ibthe longitudinal water surface slope, h*the average water
depth, uthe average flow velocity, Czthe Chezy coefficient and Fr the Froude number. Lband Hbare the
length and the height of the bars respectively
b
zi
h
u
C
=(1)
The differences between their model results and the measure-
ments were imputed on the scale effects due to this small channel
width. Recently, Lanzoni [9][10] performed morphological labo-
ratory experiments in a flume of 1.5 m wide, 1 m deep and 50 m
long. The experiments in this flume, and therefore the results are
assumed more realistic.
In this paper, the wavelength and height of alternate bars as pre-
dicted by the model of Schielen et al. [14] are compared with the
length and height of alternate bars measured in several flume ex-
periments. The large flume experiments of Lanzoni [9][10] are
included in this comparison. The comparison leads to an assess-
ment of the validity of using a weakly non-linear stability analysis
to predict alternate bar behaviour.
2 Available results of laboratory experiments
The available data of experiments can be divided in 2 groups. The
first group consists of experiments carried out by Lanzoni [9][10].
These experiments are well documented. The second group con-
sists of experiments by several researchers [1][6][7][12][17].
Only little information on these experiments is available to the
authors. However, this group is much larger then the first one.
2.1 Lanzoni’s experiments
In the large straight sand flume of Delft Hydraulics, Lanzoni
[9][10] generated alternate bars, using steady flow conditions.
The flume is 1.5 m wide, 1 m deep and 50 m long. The
bathymetry can be measured over 43.8 m. During the experi-
ments, all flow characteristics were controlled. Both the water
depth and the flow velocity were held constant at the required
values. The sand leaving the flume was weighed and subse-
quently fed back at the upstream end of the flume, evenly spread
over the width. The experiments were divided into two series. In
the first series the sediment was uniform (d50 = 481 µm, d90 = 710
µm, ρs= 2.65 g/cm³). In the second series graded sediment was
used (d50 = 262 µm, d90 = 3210 µm, ρs= 2.65 g/cm³). Table 1
summarises the conditions of the experiments using both uniform
and graded sediment.
A water-level indicator and a profile indicator measure the water
level and the bed profile respectively. In bursts of 4 to 6 minutes,
these measurements were taken over three longitudinal sections,
one in the middle and the other two at 0.20 m from each bank.
The interval between two successive measurements depended on
the bed form celerity. At the end of an experiment, more mea-
surements of the bathymetry were performed at 0.40 and 0.60
meter from each wall. The sediment transport rate was deter-
mined from the immersed weight of the sediment collected at the
end of the flume. The Chezy coefficient was estimated from the
measured water depth using:
where h*is the average bottom depth, ibis the bed slope, and uis
the average flow velocity.
The measurements of the water level and the bed level can be
used to determine the water depth: h*=ζ–z
b, in which zband ζ
are the bed-level and the water level, respectively. Figure 2 shows
an example of three resulting longitudinal bed profiles.
For each section, the average longitudinal bed slope (ib1,ib2 and
ib3) is calculated using linear regression. The noise, caused by
JOURNAL OF HYDRAULIC RESEARCH, VOL. 39, 2001, NO. 2 149
Fig. 2. Longitudinal bed profile measured during experiment P1801.
Top down is subsequently shown: the profiles along a section
near the left bank, in the middle and near the right bank of the
flume.
Name y
#d
sρsuh
[m] [-] [mm] [g/cm3] [m/s] [c m]
Ashida 0.5 13 1 2.65 0.30-0.50 2.0-3.5
Jaeggi 0.3 13 0.52 2.65 0.25-0.50 1.6-2.8
0.3 23 3 1.45 0.15-0.40 1.6-4.1
0.3 11 1.8 2.65 0.35-0.50 1.4-4.1
0.3 1 4 2.65 0.37 2.44
Kinoshita 0.132 4 0.38 2.65 ±0.30 0.8-1.1
0.132 3 0.76 2.65 0.35-0.45 0.6-1.2
0.132 14 1.24 2.65 0.30-0.50 0.4-1.2
0.132 5 1.7 2.65 0.30-0.50 0.7-1.7
Muramoto 0.55 8 0.99 2.65 0.35-0.75 2.1-4.5
0.25 3 0.99 2.65 0.35-0.60 1.3-2.0
Sukegawa 0.15 5 2.3 2.65 0.40-0.60 0.6-1.9
0.15 6 0.45 2.65 0.20-0.50 0.6-1.5
0.31 8 2.3 2.65 0.40-0.65 1.7-3.7
0.31 9 0.45 2.65 0.30-0.60 1.2-2.5
0.3 9 3.55 2.65 0.60-0.90 1.0-4.2
Table 2 Conditions during several small-scale flume experiments. Here
y*is the width of the flume, #is the number of experiments,
with the range of the flow velocity and the water depth given by
uand h*, respectively. dsand ρsare the size and the density of
the sediment, respectively.
Fig. 3. Sketch explaining the definitions in the model (after Schielen et
al. [14]).
(2)
() ( )
()
0z-hUz-
tbb =++
ζζ rr
(3)
gUU
tψζ r
rrrr =++
U
(4)
hz
vuv
C,gi)}
hz
vuv
Cb
22
d
b
22
d
+
+
+
+
+
=ζζ
ψb
r
small-scale ripples and dunes, is filtered out of the bottom pro-
files, using the moving-frame averaging method [13]. (The size
and shape of the windows is unknown.) Finally, the large-scale
bar characteristics are estimated. The bar height is defined as the
difference between the maximum and the minimum bed eleva-
tion, within a bar unit. Note that the filtering of the measurements
results in low estimates of the bar height, since it reduces the ex-
tremes. The bar length is estimated using a spectral analysis [13]
of the filtered bottom profile Figure 2 shows that only a few
large-scale bars exist in the flume. Therefore, the accuracy of the
height and length estimates of the bars is limited. Table 1 presents
the characteristics of the observed alternate bars.
2.2 Small scale experiments by others
In the past, several small-scale flume experiments were carried
out. Here the experimental results of Ashida [1], Jaeggi [6],
Kinoshita [7], Muramoto [12], Sukegawa [17] are relevant. In
these experiments, the width of their flumes varied from 13 to 55
cm. In the experiments, the water depth, the surface slope, and the
flow velocity were measured. The information of the sediment
used is limited: only the mean diameter and the density
characterise the sediment. Table 2 summarises the conditions un-
der which these experiments were run.
3 The amplitude evolution model of Schielen et al. [14]
3.1 Basic equations
To describe the behaviour of alternate bars in rivers, the
morphodynamic model of Schielen et al. [14], which is an exten-
sion of Colombini et al. [4], will be used. The model of Schielen
et al. [14] is discussed briefly below.
Schielen et al. [14] considered a uniform, shallow-water flow in
a straight, infinitely long channel, with a uniform, mild slope: ib
<< 1. The banks are assumed to be non-erodible and the bottom
sediment to be non-cohesive. (See the sketch in Figure 3.) The
flow is described using the mass balance and the depth-averaged
St. Venant equations:
in which represents the forcing and friction mechanism:
ψ
r
Here gis the acceleration due to gravity and zbis the elevation of
the disturbed bed relative to the undisturbed bed. ζis the eleva-
tion of the disturbed free surface, with respect to the undisturbed
150 JOURNAL OF HYDRAULIC RESEARCH, VOL. 39, 2001, NO. 2
x,
y
(5)
2
z
d
C
Cg
=
(6)
.z-
|U|U
|U|S
b
b
= r
r
r
rγσ
(7)
0
S
t
b
z=+
(8)
0|S0,|v0,|S0,|v
yyy0yy0y
====
==
(9)
,0,0,0)(u)z,,v,(u b0000
=
ζ
(10)
()
dkyf
ek
tikx
0
0ω
εφφ +
+=
R=y*
h*
Fig. 4. Example of a neutral curve, dividing the region with a stable
basic state from the region with an unstable basic state. R is the
width-to-depth ratio; k gives the wave number of the perturba-
tions.
(11)
()
r1RR
2
c
ε+=
(12)
ε+=
1c
kkk
(13)
c.c.h.o.t)y'cos(),A(ht)y,(x,z
t'x'ki
cc
++πτχε=
ω+
b
e
water level h*.=(u,v) is the flow velocity vector in xand y-
U
r
direction and the operator is defined as
r
.
Finally,
is the drag coefficient, in which Czis the Chezy coefficient.
The sediment is transported as bed-load, which is modelled as:
The sediment mass balance yields:
where = (Sx,Sy) is the sediment transport, in volume per unit of
S
time, in xand y-direction. Furthermore, the non-linearity of the
sediment transport with respect to the flow bis limited to 2 < b<
7. The downhill preference of the sediment transport is accounted
for by γ> 0. Both parameters are dimensionless. The sediment
transport proportionality parameter σdepends on the sediment
properties and includes the effect of the porosity of the bed.
To close the model at both side walls (y=0,y=y
*) boundary
conditions have to be defined:
The model defined by the Equations (2),(3),(6),(7) and the bound-
ary conditions (8) allows for a solution, describing uniform flow
over a plane sloping bed:
(Note that u*is the uniform flow velocity, not the shear velocity.)
3.2 Stability analysis
The starting point of every stability analysis should be a physi-
callyrelevant,exactsolutiontothe mathematical model. Equation
(9) represents such a basic solution. This basic state is perturbed
by small-amplitude, periodic bed waves:
in which = (u,v,ζ,zb) and denotes the basic state given by
φ
φ0
Equation (9). The complex morphological wave frequency
ω(k,R,b,γ,Cd) is related to the morphological wave number k, the
width-to-depth ratio
and the model parameters Cd,band γ. If the growth rate (given by
the real part of ω) is negative, the basic state is stable: the pertur-
bationsdecrease inamplitude anddisappear.Under otherphysical
circumstances, the basic state is unstable: the disturbances start to
grow, forming a rhythmic pattern.
Schielen et al. [1] substituted the perturbation (10) into the model
Equations (3)-(7), linearised for small perturbations. Thus, they
found a relationship between the complex wave frequency and the
wave number. Using this relationship, one can determine whether
the basic state is stable or unstable (see Figure 4). As can be seen
in Figure 4, a whole range of waves has positive growth rates.
However, for each condition the alternate bars have only one
wavelength. More information is needed to find this wavelength.
To derive information about the shape and the behaviour of the
bars, Schielen et al. [14] had to consider the non-linear terms. In
agreement with Colombini et al. [4], they assumed a weakly non-
linear regime:
with ε<< 1 and some unknown –1 < k1<1.(Rc,kc) is the first
combination of the width-to-depth ratio Rand the wave number
kto give an unstable basic state for increasing R(the lowest point
above the neutral curve in Figure 4). Assumption (11) means, that
the width-to-depth ratio is just large enough to give growing per-
turbations.
This approach led to the following wave [14]:
JOURNAL OF HYDRAULIC RESEARCH, VOL. 39, 2001, NO. 2 151
(14)
0A|A|
A
A
A
2
3
2
2
21
=+
++
α
ξ
αα
τ
(15)
t'
2
ετ=
(16)
t'x'
k
νεχ+=
(17)
u
hig
C
2
b
d
=
(18)
50
s
z
2
d1-C
u
=
ρ
ρ
θ
(19)
()
3
50
s
2
3
cdg1--13.3S
=ρ
ρ
θθµ
(20)
θµ
θc
1
3
b
(20)
4
1
c
0.75
=θµ
θ
γ
Fig. 5. The predicted wave length plotted against the measured values
(a) for all experiments and (b) for the experiments with >
R Re
2000. The dotted lines show the 20% error boundaries.
in which εis defined by Equation (11). kcand ωcare the critical
wave number and the critical frequency, respectively. Further-
more, (x’,y’,t’) are the dimensionless co-ordinates
.
y*h*
σu
b
,t
y*
y
,
y*
x
(
(
The cosine in Equation (13)models the alternation of the crest,
h.o.t. denotes the higher order terms (O (ε2)) and c.c. means the
complex conjugates. The amplitude A(ξ,τ) follows from:
This equation is known as the Ginzburg-Landau equation. Here,
τis the morphological time and ξis the morphological co-ordi-
nate in a frame moving with the group velocity νbof the bars:
The parameters αi(i=1,2,3) in Equation (14) are complex func-
tions of the drag coefficient Cdand the transport parameters band
γand σ.
3.3 Model parameters
To calculate the wave length and wave height using Equation
(13), we need values for Cd,band γ. Unlike small bedforms, al-
ternate bars do not influence the bed roughness. (During the ex-
periments the depth h*, slope iband flow velocity u*do not change
significantly, while the alternate bars develop.) Therefore, the
drag coefficient Cdcan be estimated using Equations (1) and (5):
Accurate values of exponent bare unknown. However, using
Chezy’s law:
(where θis the shields stress) and the commonly used formula by
Meyer-Peter and Müller [11], describing transport of sediment:
one can derive that:
where θcis the critical Shields parameter and µthe bed-form or
efficiency factor.
Finally, the slope coefficient γcan be estimated using the relation
derived by Sekine and Parker [15]:
4 Comparison between theoretical and experimental results
Given the conditions of the experiments described in section 2,
the model of Schielen et al. [14] predicts the wave length and
wave height of alternate bars. These predictions are compared
with the measured values. Figure 5(a) and Figure 6(a) show the
results of the wavelength and wave height predictions respec-
tively.
The model predicts the characteristics of the larger bars accu-
rately, but the predictions of the smaller sized bars are not satis-
factory. These smaller bars are measured in the small-scale flume
experiments. Figure 7 shows, that the errors in the predictions are
related to the parameter . Here is the Reynolds num-
R Re
Re=h u
**
v
ber, with νthe kinematic viscosity. More generally, used parame-
ters, like the Reynolds number or the Froude number, give no
relation with the errors.
The relation to the parameter can be explained from the as-
R Re
sumptions in the model of Schielen et al. [14]. They assume that,
in zeroth order (O(1)), the dissipation is small compared to the
advection: Re>> 1. In first order i.e. O(ε) they assume that the
total dissipation is dominated by the vertical dissipation: the
width-to-depth ratio R>1. In the model, the effect of the horizon-
tal dissipation is neglected. This assumption is valid if >> 1
R Re
Horizontal dissipation distributes the friction from the boundaries
to the inner domain. The horizontal dissipation distributes the
friction generated at the side-banks, while the vertical dissipation
distributes the friction generated at the bed. Using the reasoning
in the above, Schielen et al. [14] neglect the boundary effects near
the banks.
Due to the horizontal dissipation of the friction with the side
152 JOURNAL OF HYDRAULIC RESEARCH, VOL. 39, 2001, NO. 2
Fig. 6. The predicted wave height plotted against the measured values
(a) for all experiments and (b) for the experiments with >
R Re
2000. The dotted lines show the 20% error boundaries.
Fig. 7. Mean of the absolute errors in the wave height predictions as a
function of the dimensionless parameter , exp denotes the
R Re
measured values.
walls, the flow velocity in the boundary layer near these walls is
smaller thanthe mean velocity. The model of Schielen et al. [14]
assumes uniform flow velocities over the channel width. Conse-
quently, the model locally underestimates the drag coefficient.
Finally, in the model the bar height is proportional to the drag
coefficient. Thus, the bar height, which is measured near the
walls, is underestimated. Apparently, the effect of the boundary
layer is significant to the bar height if < 2*10³ (Figure 7)
R Re
This value has to be regarded as an empirical result. So far, we
can not derive this number from theoretical arguments.
In agreement with Colombini et al. [4], the authors assume that
this is an artefact of the small channel width in the experiments.
The dynamics in the small flume experiments differ significantly
from the dynamics in the field, where in general > 2*10³.
R Re
Therefore, one has to consider, whether the results of small-scale
flume experiments are useful in case of morphodynamic prob-
lems.
If > 2*10³, the model estimates both wave length and wave
R Re
height accurately, as can be seen in Figure 5(b) and Figure 6(b).
(The standard deviation is of the same order as the measurement
noise, which is merely caused by small bed forms. Even in the
strong non-linear cases, in which the width-to-depth ratios are
large compared to the critical values, the results are good. This is
striking, since the model is based on the assumption that the
width-to-depth ratio is close to its critical value. This means that
the model is useful outside the theoretical restrictions.
In the Lanzoni experiments, the predicted wave heights are high
compared to the measurements. However, the errors are small and
may be caused by the filter method used by Lanzoni. As stated in
section 2.1 the moving-average method results in low estimates
of the wave height.
In the experiments with larger flumes, the predicted wavelengths
coincide with the measured values. Most errors are less than 10%
of the values, which is small compared to the measurement noise.
The two outliers in the Lanzoni experiments (P2403,P0404) can
be explained (see [8]). In one case, the power spectrum is double
peaked. The wavelength of the peak that is disregarded by
Lanzoni [9] coincides with the predicted values. In the other case,
there is no clear alternate bar pattern. The outliers in the small-
scale flume experiments are related to a high Froude number (Fr
> 1). Schielen et al. [14] assume a Froude number < 1.
The model predicts the wavelength and height of the alternate
bars equally well in experiments with graded sediment as in ex-
periments with uniform sediment. Apparently, the grading of sed-
iment has no significant influence on the behaviour of the alter-
nate bars. Therefore, it can be concluded, that the sediment trans-
port parameters capture the effect of the sediment composition
sufficiently.
5 Conclusions
The model of Schielen et al. [14] predicts the wavelength and
height of alternate bars accurately in flume experiments if >
R Re
2*10³. For these experiments, the errors can be considered as
white noise. The standard deviation is of the same order as the
measurement noise, which is merely caused by small-scale bed
forms, such as dunes and ripples.
There are no differences between the results in the experiments
using uniform sediment [9] and the experiments using graded
sediment [10]. The model parameters capture the effects of the
sediment composition sufficiently.
However, if < 2*10³, the measured heights are significantly
R Re
larger than predicted. The model used neglects the dissipation
near the side banks. This results in an underestimation of the al-
ternate bar heights. The dynamics in the small-scale flume experi-
ments differ significantly from the dynamics in the field. This
questions the usefulness of small flume experiments for
morphodynamic problems.
We conclude, that the model predicts the basic characteristics of
alternate bars well, even far from critical conditions. This justifies
further research into predicting alternate bar behaviour, using the
model of Schielen et al. [14] in combination with data-assimila-
tion.
6 Acknowledgement
The authors thank the Dutch Organisation for Scientific Research
JOURNAL OF HYDRAULIC RESEARCH, VOL. 39, 2001, NO. 2 153
symbol description unity
A scaled alternate bar amplitude [-]
Cddrag coefficient [-]
CzChezy coefficient [m½/s]
FrFroude number [-]
Hbalternate bar height [m]
Lbalternate bar length [m]
Q water flux [m3/s]
R width-to-depth ratio [-]
ReReynolds number [-]
S sediment transport vector [m3/s]
T duration of experiment [s]
U flow velocity vector [m/s]
b on linearity of sediment transport [-]
d grain diameter [µm]
g gravity constant [m/s2]
h water depth [m]
iblongitudinal free surface slope [-]
k morphological wave number [-]
t time [s]
u,v longitudinal and transverse flow velocity [m/s]
x,y,z longitudinal, transverse and vertical position [m]
y
channel width [m]
zbelevation disturbed bed [m]
φState vector of the morphological problem [-]
ΨForcing and friction term in the flow equations [-]
α1exponential amplitude growth coefficient [-]
α2horizontal amplitude variation coefficient [-]
α3non-linear amplitude decay coefficient [-]
εsmall parameter [-]
γdownhill preference of sediment transport [-]
µbed form factor [-]
νbgroup velocity of the alternate bars [-]
ψforcing of the flow [m/s2]
ρdensity [g/cm3]
σsediment transport proportionality [-]
τmorphological time [-]
θShields parameter [-]
ξmorphological length [-]
ζelevation disturbed free surface [m]
ωmorphological wave frequency [m/s2]
subscripts
50 median
90 90\% is smaller
ccritical value
sof sediment
xlongitudinal component
ytransversal component
average value
0basic solution
bf the river bed
superscript
' scaled value
Table 3. List of symbols, subscripts and superscripts
(NWO) for funding this work through the ’Non Linear Systems’
priority program (under project number 620-61-349). It is linked
to the PACE-project in the framework of the EU-sponsored Ma-
rine Science and Technology Program (MAST III), under con-
tract number MAS3-CT95-0002.
7 Symbols
8 References
1. Ashida, K. and Shiomi, Y., Study on the hydraulics behavi-
our of meander in channels, in Disaster Prevention Research
Institute Annuals, pp. 457-477, Kyoto University, 1966.
2. Blondeaux, P. and Seminara, G., A unified bar-bend the-
ory of river meanders, Journal of Fluid Mechanics, 157, pp.
449-470, 1985.
3. Colombini, M. and Tubino, M., Finite-amplitude free bars:
a fully non-linear spectral solution, in Euromech 262, edited
by Soulsby and Bettess, pp. 163-169, Balkema, Rotterdam,
1991.
4. Colombini, M., Seminara, G. and Tubino, M., Finite am-
plitude alternate bars, , Journal of Fluid Mechanics, 181, pp.
213-232, 1987
5. Fletcher, R., Practical Methods of Optimization, Wiley,
1987.
6. Jaeggi, M., Formation and effects of alternate bars, Journal
Hydraulic Division, pp. 1103-1122, ACSE, 110, 1984.
7. Kinoshita, R. Investigation of channel deformation in
Ishikari river, Tech. report, Bureau of Resources, Department
of Science and Technology, Japan, 1961.
8. Knaapen, M.A.F., Hulscher, S.J.M.H., de Vriend, H.J.,
and van Harten, A., Alternate bar patterns in rivers: Mod-
elling vs. laboratory experiments, in Advances in Hydro-Sci-
ence and –Engineering, edited by K. Holtz, W. Bechteler, S.
Wang and M. Kawahara, ICHE’98, CD-ROM, 1998.
9. Lanzoni, S. Experiments on free and forced bar formation
in a straight flume. Uniform sediment, research report Q1774,
Delft Hydraulics, 1995.
10. Lanzoni, S. Experiments on free and forced bar formation
in a straight flume. Graded sediment, research report Q1774,
Delft Hydraulics, 1997.
11. Meyer-Peter, E. and Müller, R., Formulas for bed-load
transport, IAHR, Stockholm, 1948.
12. Muramoto, Y. and Fujita, Y., The classification of meso-
scale river bed configuration and the criterion of its forma-
tion, Proceedings 22nd Japanese Conference on Hydraulics,
pp. 275-282, ISCE, 1978.
13. Oppenheim, A. V. and Schafer, R. W., Discrete-Time Sig-
nal Processing, Prentice-Hall, Englewood Cliffs, NJ, 1989.
14. Schielen, R. M. J., Doelman, A., and de Swart, H. E., On the
non-linear dynamics of free bars in straight channels, Journal
of Fluid Mechanics, 252, pp. 325-356, 1993.
15. Sekine, M. and Parker, G., Bed-load transport on trans-
verse slope, Journal of Hydraulic Engineering, 118, 513-535,
1992.
16. Seminara, G. andTubino, M., Alternate bars and Meander-
ing: Free, Forced and Mixed Interactions, American Geo-
physical Union, 1989.
17. Sukegawa, N., Study on meandering of streams in straight
channels, Tech. report, Bureau of Resources, Department of
Science and Technology, Japan, 1971.
18. Tubino, M., Growth of alternate bars in unsteady flow, Wa-
ter Resource Research, 27(1), pp. 37-52, 1991.
... In contrast, bed material of sand bed rivers like the Loire River in France or the lower Elbe River in Germany is transported at nearly all discharges, so that the grain size distribution of the bed material affects bar characteristics less. This is especially the case in sand bed rivers that are characterized by a uniform grain size distribution and transport of all fractions (Knaapen et al., 2001;Venditti et al., 2012), and bars in these environments may thus also form during low discharges (Rodrigues et al., 2012(Rodrigues et al., , 2015. Moreover, sand bed rivers transport an increasing part of bed material in suspension with increasing discharge. ...
... The bar heights on the left and right side of the channel were characterized using the standard deviation (SD) of the longitudinal outer 20 right and left elevation profiles as a surrogate measure. We refrained from calculating the absolute bar height defined as the distance between the lowest and the highest elevation of one bar complex (e.g., Ikeda, 1984;Knaapen et al., 2001;Adami et al., 2016) as the identification of a single bar is rather subjective and can be easily biased by single extreme values (e.g., Redolfi et al., 2020). It is worth noting that restricting the analysis to the 100 m wide riverbed may underestimate bar height, as bars could be even higher in the omitted areas. ...
Article
Full-text available
This study investigates the occurrence and dynamics of single row alternate bars forming in a particular reach of the sand bed Elbe River in Germany. Although the formation and dynamics of alternate bars have been intensively investigated in the literature, there exists only a limited number of studies focusing on the characteristics of alternate bars forming under complex field conditions. This is particular the case for bars forming in trained sand bed rivers, as most previous field studies have focused on gravel bed rivers. Moreover, little is known on the impact of river training structures on bar characteristics in anthropogenic rivers. To close this gap, we present a comprehensive bed elevation data set that was collected over a period of 10 years within a 30 km long reach of the lower Elbe River in Germany by the Federal Waterways and Shipping Administration (WSV). The reach is characterized by a sand bed, has curved as well as straight parts, and exhibits a section that is less trained by groins than the neighboring sections. For our analyses, we propose a novel approach to estimate bar characteristics based on statistically derived geometrical parameters. The outcomes of the approach are used to show that bars in Elbe River belong to the free bar type and that their origin of formation and characteristics depend on hydrological and geometrical boundary conditions. The results reveal that the active width of the river bed, defined as the distance between the groin heads, is a crucial parameter for the occurrence of alternate in the reach. We further highlight the impact of river bends on bar characteristics, as bars in the outer bend were longer and higher than their inner bend counterparts. Finally, we show that simple predictors for bar formation can be successfully applied to predict bar formation in sand bed rivers but that care needs to be taken when applying such approaches to more complex boundary conditions.
... Such works tended to include a large experimental component, with the existing alternate bar data to a large extent resulting from them. More recent years have been marked by substantial analytical and numerical modeling efforts, primarily involving stability analyses of the phenomenon, and aiming at developing a better understanding of the full dynamics of alternate bars including their origin and subsequent development (e.g., Colombini et al. 1987;Tubino and Seminara 1990;Schielen et al. 1993;Tubino et al. 1999;Knaapen et al. 2001;Defina 2003;Crosato et al. 2012;Bertagni and Camporeale 2018;Redolfi 2021). Classically, the acquisition of data on alternate bars has been accomplished solely through laboratory experiments. ...
... These alternate bars are forced in that they are not freely moving, and they have wavelengths scaled to 3-12 channel widths (Ikeda, 1984;Jaeggi, 1984;Colombini et al., 1987;Seminara and Tubino, 1989;Nelson, 1990;Schielen et al., 1993;Knaapen et al., 2001), features broadly shared by the point assemblages in the Missouri River (Figs. 2 and 4). Traditionally, these forced alternate bars would grow as bank-attached compound point bars into a roughly uniformwidth single-thread channel (e.g., Blondeaux and Seminara, 1985). ...
Article
Full-text available
This paper offers a mechanism for meandering in an otherwise braided river and then discusses its general implications for river processes and fluvial deposits. Braided rivers manage to meander without the paired point bars and single-thread channels that are instrumental in developing bends in other meandering rivers. The driving processes for meandering in these braided systems remain enigmatic. The unchannelized and prechannelized Missouri River is an example of a braided meandering river, and it provides an opportunity to gain insight into these processes. This study utilized historical maps, sequential air photos, and surficial geologic maps both to define the processes by which this braided river meanders, and to characterize the deposits produced by these processes. These data show that the Missouri River meanders by building point assemblages instead of point bars. Repeated accretion of midchannel and lateral bars to a common point on the bank forces development of a meander bend around a point assemblage comprising multiple amalgamated compound bars. This differs from single-thread systems, which expand and translate bends around a single compound point bar. Alternating development of point assemblages forces meandering over successions of meander bends. Braided meander loops grow by expansion and translation like single-thread rivers, but they also may contract to produce counterpoint assemblages. Contraction appears to be the more common means of loop abandonment compared to loop cutoff for the braided Missouri River. This differs from single-thread meandering rivers, where contraction is limited, and loop cutoff is consistently the dominant abandonment process. Deposits of the braided meandering Missouri River differ from deposits of single-thread rivers in the rarity of both meander scrolls and single-thread channel fills. Instead, point and counterpoint assemblages comprise fusiform bar elements bound by small filled remnants of anabranch channels. These assemblages are commonly bound by meander cutbank scars. Cutbank scars associated with contraction, however, tend to be composite rather than discrete erosional surfaces, and they do not tend to bind river-scale abandoned channel fills. The braided meandering Missouri River also differs from wandering rivers because wandering rivers meander by building compound bars instead of assemblages, are more gravelly, have less pervasive and much less mobile midchannel bars, and appear to reflect a transitional intermediate pattern instead of a stable hybrid pattern. Braiding and meandering both expend stream power, and both are mechanisms for achieving channel equilibrium. The Missouri River exhibits both of these processes in tandem; thus, meandering and braiding are not mutually exclusive processes. Braided meandering rivers like the Missouri River are less common than either straight-braided or single-thread-meandering rivers, but they are not unique. The long-held distinction of braided versus meandering patterns for rivers thus may be practical but is not definitive.
... Secondary flow exerts an important role in several processes involving in a natural stream, such as the dispersion of nutrients and/or of chemical substances as well as the trapping or transport of sediments. In particular, the cross-sectional motion contributes both to shift the core of the maximum velocity (among others Leschziner and Rodi, 1979 ;Johannesson and Parker, 1989 ;Wilson et al., 2003 ) and of the boundary shear stress along the bend, thus modifying and reshaping the channel's bed and platform (see as an example de Vriend and Gendolf, 1983 ;Whiting and Dietrich, 1993 ;Knaapen et al., 2001 ;Blanckaert and Graf, 2004 ;Roca et al., 2009 ;Termini and Piraino, 2011 ), and to enhance the mass mixing and the energy losses, thus affecting the channel's conveying capacity ( Camporeale et al., 2005 ;Nepf, 2012 ;Gurnell, 2014 ;De Serio et al., 2018 ). ...
Article
Literature shows the application of the entropy concept in open channels flow. No existing research has been addressed to the application of the entropy concept for evaluating the effect of secondary flow which especially develops in curved channels. The purpose of this paper is to explore the possibility to evaluate the effect of the secondary flow on the stream-wise velocity distribution starting only from the knowledge of the surface velocity and the entropic concept application. The analysis is conducted with the aid of data collected in a large-amplitude meandering laboratory flume for two different values of the width-to-depth ratio. Results show that the location of the maximum velocity varies not only along the bend but also along the water depth, depending on the relative entity of the cross-circulation and the convective flow acceleration. It has been also verified that the entropic parameter varies along the bend as a function of the width-to-depth ratio. The application of the entropic model to the examined cases has allowed the verification that not only the shape of the velocity profile but also the bed shear stress evolution along the bend could be estimated by the entropic model starting from the knowledge of the surface velocity only.
... The earlier works (in the 1970s and 1980s) largely focused on the establishment of threshold conditions of alternate bar formation (for reviews, see da Silva 1991 ;Yalin 1992;Katolikov 2000;and Ahmari and da Silva 2011). More recently, substantial analytical and numerical modeling efforts have been made to penetrate the dynamics of the origin and subsequent development of alternate bars, primarily through stability analysis (e.g., Colombini et al. 1987;Tubino and Seminara 1990;Schielen et al. 1993;Tubino et al. 1999;Knaapen et al. 2001;Defina 2003;Crosato et al. 2012;Duró et al. 2016). ...
Article
A new empirical equation for alternate bar height is introduced. It is assumed that the bars are formed under a steady and uniform flow; the stage of interest is that where bars have grown to their fully developed state. The equation is developed on the basis of dimensional considerations and all data available to the authors; the formulation also incorporates findings by stability analysis. The data result from a total of 191 flume experiments reported in 16 different works carried out from 1961 to 2014 using either sand or gravel as bed material. Both fully rough and transitionally rough flows were used in the experiments. Overall, width-to-depth ratios and relative depths varied from 3.5 to 54.4 and 3.8 to 191, respectively; the ratios of bed shear stress to critical bed shear stress ranged from 1.1 to 14.7. Data from the Naka River, Japan, are also used. The equation correlates bar height, normalized by flow depth, with the excess width-to-depth ratio (B=h) with regard to the smallest, or critical, value of B=h at which alternate bars occur as well as with the relative depth and, in the case of transitionally rough flows, also the grain-size Reynolds number. It is shown that when compared with previous empirical equations, the present equation produces a considerably improved overall alignment of the data with the perfect agreement line and a significant larger (nearly double) percentage of data falling within the 20% error range. The results by the present equation are also substantially more congruent with those derived from existing theoretical and numerical analyses of the development of alternate bars, and more specifically, those resulting from a stability analysis of the phenomenon. As a by-product of this work, a first comparative evaluation of existing equations for alternate bar height is also presented.
... They are characterized by a bar mode equal to one, in contrast to the higher bar modes of multiple bars (Ikeda, 1984;Jaeggi, 1984). Alternate bars have been studied extensively in laboratory experiments (Fujita and Muramoto, 1985;Struiksma and Crosato, 1989;Tubino, 1991;Lisle et al., 1993;Lanzoni and Tubino, 1999;Lanzoni, 2000;Knaapen et al., 2001;Crosato et al., 2011) and theoretical analyses (reviewed later), but studies of their formation and further development in real rivers have remained limited (Welford, 1994;Eekhout et al., 2013). Theoretical analyses of the mathematical equations for flow, sediment transport and morphological bed evolution show that alternate bars occur only within a specific range of the width-to-depth ratio of an alluvial channel. ...
... They are characterized by a bar mode equal to one, in contrast to the higher bar modes of multiple bars (Ikeda, 1984;Jaeggi, 1984). Alternate bars have been studied extensively in laboratory experiments (Fujita and Muramoto, 1985;Struiksma and Crosato, 1989;Tubino, 1991;Lisle et al., 1993;Lanzoni and Tubino, 1999;Lanzoni, 2000;Knaapen et al., 2001;Crosato et al., 2011) and theoretical analyses (reviewed later), but studies of their formation and further development in real rivers have remained limited (Welford, 1994;Eekhout et al., 2013). Theoretical analyses of the mathematical equations for flow, sediment transport and morphological bed evolution show that alternate bars occur only within a specific range of the width-to-depth ratio of an alluvial channel. ...
Article
A field study was carried out to investigate the development of alternate bars in a secondary channel of the Loire River (France) as a function of discharge variations. We combined frequent bathymetric surveys, scour chains and stratigraphical analysis of deposits with measurements and modelling of flow dynamics. The channel exhibited migrating bars, non-migrating bars and superimposed dunes. Possible mechanisms of bar initiation were found to be chutes associated with changes of bank direction and instability resulting from interactions between existing bars during the fall in water level after floods. We propose that the reworking of bar sediments during low flows (high width-to-depth ratio β), reinforced by high values of the Shields mobility parameter, can explain the formation or re-generation of new alternate migrating bars during a subsequent flood. The migration pattern of the bars was found to be cyclic and to depend mainly on (i) channel layout and (ii) the dynamics of superimposed dunes with heights and lengths depending on location and discharge value. For instance, the hysteresis affecting the steepness of dunes influences the flow resistance of the dunes as well as the celerity of migrating bars during flood events.We compare the findings from the field with results from theoretical studies on alternate bars. This gives insight in the phenomena occurring in the complex setting of real rivers, but it also sheds light on the extent to which bar theories based on idealized cases can predict those phenomena.
Chapter
This article gives and overview of the large spectrum of models based on first principles (mass and momentum conservation) developed in the field of fluvial morphodynamics. After describing the main morphological features of alluvial rivers and their floodplains, the chapter gives a general classification of the modeling approaches available in literature. Next, it delineates the morphodynamic problem and briefly summarizes the procedures used to solve the governing equations, depending on the spatial and temporal scales resolved. The problem of the closures arising from depth-averaging is then discussed. The various models are grouped with reference to the morphological features to be simulated. The dynamics of small-scale and large-scale bedforms in single-thread straight rivers and the modeling of braided, anastomosing and meandering rivers are then considered. Finally, the simulation of the coupled river-floodplain evolution is tackled. The chapter ends with a delineation of the modeling perspectives.
Book
Cambridge Core - Geomorphology and Physical Geography - River Dynamics - by Bruce L. Rhoads
Article
We extend the class of simple offshore models that describe large-scale bed evolution in shallow shelf seas. In such seas, shallow water flow interacts with the seabed through bed load and suspended load transport. For arbitrary topographies of small amplitude we derive general bed evolution equations. The initial topographic impulse response (initial sedimentation and erosion patterns around an isolated feature on a flat seabed) provides analytical expressions that provide insight into the inherent instability of the flat seabed, the Coriolis-induced preference for cyclonically oriented features, and bed load transport being a limiting case of suspended load transport. The general evolution equation can be used to describe sandbank formation, known as the result of self-organization. Examples of human intervention at the seabed include applications to a dredged channel and an offshore sandpit. An outlook toward future research is also presented.
Article
Full-text available
A two-dimensional model for predicting finite-amplitude characteristics of migrating bars in straight alluvial channels is presented. A fully nonlinear numerical solution is proposed whereby flow equations and sediment continuity equations are solved by using spectral techniques. Theoretical findings on alternating bars derived by means of a weakly nonlinear analytical solution (Colombini et al. 1987) are recovered when the width to depth ratio is relatively close to its threshold value for bar formation. For larger values of the above ratio the numerical model provides a more refined representation of actual bar topography. The model is used for investigating the process of nonlinear competition between different transverse modes (single-alternating and multiple-row bars) as the width to depth ratio of the channel increases. A tendency toward the selection of higher-order transverse modes is predicted by the numerical solution. (A)
Article
Full-text available
The mean current velocity profile in the combined wave-current motion is calculated by use of the depth-integrated momentum equation. The velocity distribution is assumed to be logarithmic inside as well as outside the wave boundary layer, but with different slopes. Hereby, it is possible to describe the flow in the whole range from the pure wave motion (without any mean current) to the pure current motion (without any waves). The theory covers an arbitrary angle between wave propagation direction and mean current direction.
Article
Full-text available
A simple morphological model is considered which describes the interaction between a unidirectional flow and an erodible bed in a straight channel. For sufficiently large values of the width-depth ratio of the channel the basic state, i.e. a uniform current over a flat bottom, is unstable. At near-critical conditions growing perturbations are confined to a narrow spectrum and the bed profile has an alternate bar structure propagating in the downstream direction. The timescale associated with the amplitude growth is large compared to the characteristic period of the bars. Based on these observations a weakly nonlinear analysis is presented which results in a Ginzburg-Landau equation. It describes the nonlinear evolution of the envelope amplitude of the group of marginally unstable alternate bars. Asymptotic results of its coefficients are presented as perturbation series in the small drag coefficient of the channel. In contrast to the Landau equation, described by Colombini et al. (1987), this amplitude equation also allows for spatial modulations due to the dispersive properties of the wave packet. It is demonstrated rigorously that the periodic bar pattern can become unstable through this effect, provided the bed is dune covered, and for realistic values of the other physical parameters. Otherwise, it is found that the periodic bar pattern found by Colombini et al. (1987) is stable. Assuming periodic behaviour of the envelope wave in a frame moving with the group velocity, simulations of the dynamics of the Ginzburg-Landau equation using spectral models are carried out, and it is shown that quasi-periodic behaviour of the bar pattern appears.
Article
Bed-load particles tend to move downslope in response to gravity and downstream in response to fluid drag. The effect of gravity becomes particularly important when the downslope gravitational force has a component normal to the streamwise direction. Such phenomena as bank erosion, bed topography in river bends, sediment sorting, and stream braiding are strongly influenced by transverse bed-load transport under the influence of gravity. Here, the problem is analyzed from a consideration of the dynamics of saltating grains. Simultaneous solution of the streamwise and transverse equations of motion for a grain saltating on a transversely tilting bed of moderate slope allows a determination of mean particle trajectory. This mean trajectory, while predominately directed in the streamwise direction, includes a transverse downslope component as well. The results are used to determine a relation for the ratio of transverse to streamwise bed-load transport. This ratio increases with increasing transverse slope and decreases with increasing streamwise shear stress. The relation is compared with observations as well as existing relations obtained from simpler considerations.
Article
Major river regulation works have often resulted in alternate bar formation. Experimental work is described which provides more information on this phenomenon. Alternate bar formation is closely associated with meandering, inferring that alternate bar formation is caused by flow induced periodic disturbances of the horizontal velocity profile. Based on these considerations a criterion has been developed which defines the upper limit of alternate bar formation, and this has been satisfactorily compared to other formation criteria and experimental data. The lower limit is given by the initiation of bedload transport. A simple relation between slope, channel width and grain size is then deduced to indicate whether for a given channel, alternate bars can develop for any discharge. This relation takes into account armoring effects. Another equation is proposed to predict the scour depth resulting from alternate bar formation. Such scour often induces costly river maintenance works. Finally, some interesting results on flow resistance hav been found, leading to a tentative equation for the intermediate roughness range.
Article
In this paper, the wave length of alternate bars as predicted by linear and weakly non-linear stability theory is compared to wave lengths measured in laboratory experiments in a large sand flume. Linear theory states that the wavelength of the bars coincides with the fastest growing waves. Nonlinear theory expects those waves to prevail that are the first to start growing when the width to depth ratio exceeds its critical value. The wave lengths found in the experiments coincide surprisingly well with these critical wave lengths predicted by weakly nonlinear theory, but little agreement is found with linear theory.
Article
A theoretical model is formulated to investigate the development of the amplitude of alternate bars in unsteady flows. The problem is tackled by means of a weakly nonlinear analysis developed in a neighborhood of the threshold conditions for bar formation. Bar response to unsteady flow is found to depend on a parameter U ^ that is a measure of the ratio between the time scale of the basic flow and the time scale of bar growth. The present theory shows that if U ^ is O(1), as often occurs in nature, flow unsteadiness affects the instantaneous growth rate and phase of bar perturbations and controls the final amplitude reached by the bed configuration. A procedure for determining the final amplitude for a given flood event is proposed. Flume experiments were performed to test the main theoretical results. The bed response to unsteady flow was measured for different values of the period of the flood. The observed temporal behavior of the bar amplitude proves to be strongly affected by the unsteady character of the flow for U ^ of O(1), as predicted by the theory.