Sei sulla pagina 1di 487

Water Resources Monograph 12

River
Meandering
SyunsukeIkeda
Gary Parker
Editors

American Geophysical Union

Published under the aegis of the AGU Water Resources Monograph Board.

Library of Congress Cataloging-in-Publication Data


River meandering Syunsuke Ikeda and Gary Parker, editors.
p. cm. - (Water resources monograph: 12)
ISBN 0-87590-316-9
1. Meandering rivers. I. Ikeda, Syunsuke. II. Parker, Gary. III. Series.
GB1205.R55
1989
551.48'3-dc20
89-6873

Copyright 1989 by the American Geophysical Union, 2000 Florida Avenue,


NW, Washington, ,DC 20009
Figures, tables, and short excerpts may be reprinted in scientific books and
journals if the source is properly cited.
Authorization to photocopy items for internal or personal use, or the
internal or personal use of specific clients, is granted by the American
Geophysical Union for libraries and other users registered with the Copyright
Clearance Center (CCC) Transactional Reporting Service, provided that the
base fee of $1.00 per copy, plus $0.10 per page is paid directly to CCC, 21
Congress Street, Salem, MA 01970. 0065-8448/89/$01. + .10.
This consent does not extend to other kinds of copying, such as copying for
creating new collective works or for resale. The reproduction of multiple
copies and the use of full articles or the use of extracts, including figures and
tables, for commercial purposes requires permission from AGU.
Printed in the United States of America.

CONTENTS

1.

Boundary Shear Stress and Sediment Transport in River Meanders of


Sand and Gravel, by Willianl E. Dietrich and Peter Whiting 1

2.

Sedimentary Controls on Channel Migration and Origin of Points Bars in


Sand-Bedded Meandering Rivers, by Hiroshi Ikeda 51

3.

Flow in Meandering Channels with Natural Topography, by Jonathan 1v1.


Nelson and J. Dungan Smith 69

4.

Sedinlent Transport and Sorting at Bends, by Syunsuke Ikeda 103

5.

Sediment Control by Submerged Vanes.


Odgaard and Anita Spoljaric 127

6.

Analysis ofN-D Bed Topography Model for Rivers, by Nico Struiks111a


and AI~ssandra Crosato 153

7.

Linear Theory of River 1\1eanders, by I-Ielgi Johannesson and Gary


Parker 181

8.

Studies on Qualitative and Quantitative Prediction of Meander Channel


Shift, by I(azuyoshi Hasegawa 215

9.

Finite Amplitude Development of Alternate Bars, by Shoji Fukuoka 237

10.

Alternate Bars and 1\1eandering;


Free, Forced and 1\.fixed Interactions,
by G. Seminara and 1\1. Tubino 261

11.

Evolution and Stability of Erodible Channel Beds, by Jonathan 1\/1.


Nelson and J. Dungan Slnith 321

12.

Observations on Several Recent Theories of Resonance and Overdeepening


in 1\1eandering Channels, by Gary Parker and Helgi Johannesson 379

13.

Bar and Channel Forluation in Braided Streanls, by Yuichiro Fujita 417

14.

Topographic Response of a Bar in the Green River, Utah to Variation in


Discharge, by E.D. Andrews and J. 1\1. Nelson 463

Design Basis, by A. Jacob

Q..,

iii

PREFACE

River channel pattern comes in three flavors: straight, meandering, and


braided. Of these, meandering is perhaps the most common, but at the same
time the most mysterious: it is strikingly rich in pattern, yet is encumbered
with neither the sterile order of its straight cousin, nor the undecipherable
disorder of its braided relative.
Consider Figures 1 and 2 herein.
The former is of the artificially
straightened Naka River, Shikoku, Japan.
Thwarted in its quest for a
meandering planform, the river has, nevertheless, generated of its own the
precursor known as alternate bars. In the latter photograph, a pair of images
of the East Nishnabotna River, Iowa, USA, illustrate the elegance of the
sinuous river, and the inexorable migration and deformation of bends so
characteristic of freely meandering strealns.
The inception and growth of meander bends, and their eventual demise
through the mechanism of cut-off formation, have been treated by many, and
the literature on the subject lies scattered about the world in such fields as
hydraulics, geology, and geography.
This volume represents an attempt to
dra\v together an international group of theoretical researchers at the cutting
edge, supplemented with several leading field-oriented researchers.
(Indeed,
some among our numbers comfortably occupy both camps.) The underlying
philosophy \vas an attempt to extract from the various schools a more unified
understanding of the mechanics of meandering, as a base from which to build
for the future.
The concept \vas originally born in 1981, in a letter communicated fron1
S. Ikeda to G. Parker. It developed rapidly in the aftermath of the Specialty
Conference on River Meandering, held in Ne\v Orleans in 1983 under the
sponsorship of the American Society of Civil Engineers.
Eventually, our
proposal, "Development and Application of the Theory of River Meandering",
was adopted in 1985 by the U.S. National Science Foundation and the Japan
Society for the Promotion of Science, as a bilateral research project.
This monograph is not simply a collection of papers from a conference,
but rather the product of a joint research effort of unique and unusual
character. In order to appreciate the mood of the volunle, it is of use to
describe hovv the project proceeded.
The core group of participants \\Tas chosen to consist of five Japanese
and five Americans, all relatively young. (These \vords will surely come back
to haunt us.) The research was organized about three workshops. The first
of these, held in Tokyo in July of 1985, was en1bedded in the midst of a
t,vo-week tour of rivers and research facilities, one that ,vas to take us to
three of the four lnajor islands of Japan. Naturally, the Naka River, sho,vn in
Figure 1, was on the itinerary. The second \vorkshop was held in September
of 1986, at the Brinkerhoff House, a lodge on Jenny Lake, Grand Teton
v

PREFACE

vi

Figure 1.

Alternate bars in the Naka River, Tokushima Prefecture, Japan


(courtesy Tokushima Construction Office, Ministry of
Construction, Japan).

National Park, Wyoming (Figure 3). It was part of a similar two-week series
of site visits which eventually included nine rivers in California, Wyoming,
Iowa (Le. the East Nishnabotna River), and Minnesota. The third and final
workshop was held at Poipu Beach, Kauai, Hawaii, in October of 1987 (Figure

4).

The researchers were in contact not only during the workshops, but
through the length of the site visits. We travelled together, ate together, and
often jointly occupied hotel rooms. During our stay at Jenny Lake, we even
washed dishes and swept the floor together, and in a cooperative effort of
special import, chased a bear off the veranda.
The workshops were all of
relatively free format, with the emphasis on active, and often heated
The' extended period together helped greatly to reduce the
discussion.
language barrier, and facilitated a seemingly unending interplay of ideas. A

Ikeda and Parker

Figure 2.

vii

The meandering East Nishnabotna River just south of Red Oak,


Iowa: (left) October 5, 1973; (right) May 25, 1979. Channel
migration can be clearly discerned (courtesy A. J. Odgaard).

spin-off of this atmosphere is an exchange of researchers that continues until


this day.
In order to provide a more complete perspective on the problem, two
leading European researchers were invited to participate in the final workshop.
The monograph thus includes fourteen papers authored by: the Japanese,

PREFACE

viii

Figure 3.

The second workshop, held at Brinkerhoff House, Jenny Lake,


Wyoming, on September 8, 1986.

American, and European participants and their graduate students. It is fair to


say that no individual came out of the workshops without a significantly
changed perspective on the problem of river meandering.
The participants thus included eight civil engineers, three geologists, and
a geographer. The tone of the monograph is mechanistic and interdisciplinary.
The interplay between theory, computation, experiment, and field observation

Figure 4.

The third workshop, held at the Poipu Beach Sheraton Hotel,


Kauai, Hawaii, on October 21, 1987.

Ikeda and Parker

ix

is fully exploited.
One sign evidencing the heightened spirit of cooperation
and understanding is the prodigious cross-referencing among the papers.
The list of people and organizations deserving thanks for making this
effort possible is long.
First and foremost are the U.S. National Science
Foundation and the Japan Society for the Promotion of Research, who
provided the funding.
In Japan, the Ministry of Construction and the
I-Iokkaido Development Bureau deserve special thanks. In the United States,
the Department of the Interior and the Army Corps of Engineers greatly
facilitated the program.
Various graduate students from the University of
California, Berkeley and the University of Minnesota drove vehicles for us.
Professor J. F. I(ennedy kindly made a personal trip to the East Nishnabotna
River to show us the "Iowa Vanes" being tested there as a means of
preventing bank erosion.
Patricia Swanson, Diana Dalbotten, and Donna
Elftmann spent long hours typing and editing the papers to a common format,
enabling us to meet the many externally-imposed deadlines.
It is our hope that the volume represents a \vatershed in progress on
river meandering, and a stimulus to younger researchers who might be inclined
to en1bark upon the study of this most elegant and beautiful of natural
phenomena.
Syunsuke Ikeda
Gary Parker

Water Resources Monograph

River Meandering

Copyright 1989 by the American Geophysical Union.

Boundary Shear Stress and Sediment Transport


In River Meanders of Sand and Gravel
William E. Dietrich and Peter Whiting
Department of Geology and Geophysics
University of California, Berkeley, 9-17!O

Field measurements in a sand-bedded river and in two gravel-bedded ones


are compared to examine controls on boundary shear stress fields, sediment
transport processes, and sorting in meanders. Analysis of detailed flow field
measurements in the sand-bedded river meander and over a gravel-bedded
alternate bar reveals a well-defined spatial structure to the magnitude and sign
of forces controlling boundary shear stress that arise from topographicallyinduced spatial accelerations. The relationship between bedload transport and
boundary shear stress fields in river meanders varies with size and
heterogeneity of bed material. In bends of moderately to well sorted sand in
flows generating boundary shear stresses well above critical (such as in large
sandy rivers), downstream varying boundary shear stress is matched by
topographically-induced cross-stream transport of sediment. In meanders with
high excess shear stress but poorly sorted coarse sand and fine gravel,
boundary shear stress variation downstream is partially matched by surface
grain size adjustments and by net cross-stream sediment flux.
Maxima of
bedload transport rate and boundary shear stress do not correspond in some
areas. In gravel-bedded meanders with low excess boundary shear stress and
low sediment supply, bedload may be much finer than the bed surface, and
significant areas of bar surface are covered with grain sizes that constitute a
very small portion of the bedload. Substantial bedload transport may only
occur over a narrow portion of the bed width where boundary shear stress
relative to critical stress of the surface is highest and where the sediment flux
from upstream is locally concentrated. In this case, grain size adjustments
dominate over topographically-induced cross-stream sediment transport in
controlling the relationship between boundary shear stress and bedload
transport fields.

Introduction
Morphologic adjustments occur in rivers when the divergence of the
boundary shear stress field causes sediment flux divergence leading to either
net scour or deposition.
In . exploring the mechanisms of morphologic
adjustments in rivers, then, two key questions can be posed: 1) what is the
relationship between the boundary shear stress field and channel topography,
and 2) what is the relationship between the boundary shear stress field and
the sediment transport field. Theories for channel morphology must partially
assume the answer to these questions in order to solve the complex coupled

Copyright American Geophysical Union

Vol. 12

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

flow and sediment transport equations.


A goal of empirical field studies
should be to test these assumptions. Both of these questions are addressed
here.
The first question is motivated in part by debate over the role of
topographically-induced convective accelerations in the overall downstream and
cross-stream force balance controlling flow through river meanders. Scaling
arguments by Smith and McLean [1984] and Nelson [1988] showed that such
convective accelerations were large and properly belonged in a "zero-order"
force balance when the equations of motion are solved by perturbation
expansion.
Although Dietrich and Smith [1983] reported some data on
convective accelerations. from a field study, it was recently concluded by
Odgaard and Bergs [1988, p. 43] that in particular "they were not able to
provide clear cut evidence from their field data... " for the inclusion of
convective acceleration terms neglected in the seminal theory by Engelund
[1974]. The first part of our paper endeavors to show "clear cut" evidence
that convective accelerations associated with downstream changing topography
are large and systematic in their contribution to the total force balance in a
natural meandering river with well developed point bars. These accelerations
are computed from the spatial derivatives of large terms; consequently, they
are prone to large error even in laboratory settings, but we argue here that the
observed general spatial pattern of these terms in a bend and the consistent
magnitude of these terms are well defined with our field data. We also report
data on forces due to convective accelerations obtained from detailed
measurements over a bar in a nearly straight reach to demonstrate the
significance of these terms in the alternate bar case.
Although the issue
regarding convective accelerations may seem rather narrow, it has broad
implications.
The key effect of downstream changes in bed topography on flow is not so
much on the magnitude of the boundary shear stress at any point on the
stream bed, but rather on the direction of the boundary shear stress vector,
particularly over the top of the point bar. In a sequence of bends, the effects
of changing curvature alone on the growth and decline of superelevation will
cause a zone of maximum boundary shear stress to shift from near the
upstream inside bank to the downstream outside bank [see, for example,
Dietrich, 1987, p. 181].
Dietrich [1987] has shown that, even using a
downstream force balance just between the pressure gradient and boundary
shear stress (Le. no downstream convective accelerations included), the
vertically-averaged downstream velocity field is fairly closely predicted at his
Muddy Creek study site. Not surprisingly then, others, Le. Odgaard [1988]
and Bridge [19831, have also modeled relatively accurately aspects of Muddy
Creek flow and geometry with equations which did not include all
similarly-scaled convective accelerations.
But neglecting to include all
appropriate topographically-induced convective accelerations prevents fully
quantifying the most important effect of the point bar: point bars force the
flow over the bar toward the opposite bank, even at the bed. This effect,
called quite graphically "topographic steering" by Nelson [1988], is linked to
the convective acceleration terms, as illustrated in Figure 1.
In essence,
shoaling of the flow over the point bar generates convective accelerations that
cause a pressure rise over the bar and drop over the pool such that, in the
cross-stream direction, centrifugal force exceeds the opposing pressure gradient
force and net outward flow occurs. Put more" simply, the flow goes around
the bar.
This "shoaling" or "steering" effect is particularly important because it
provides a mechanism for topographic adjustments due to stage change and for
the development of an equilibrium bar topography, especially in channels with
relatively flat bar tops. The instability that leads to point bar growth in

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Dietrich and Whiting


reduced cross-stream slope

due to convec:~\:rations

--

Fig. 1.

Changes in the water surface topography due to shoaling of flow over the
point bar. Solid lines delineate channel outline and water surface in a curved
channel with downstream varying bed topography but neglecting effects of
spatial accelerations.
Dashed lines delineate the water surface in the same
channel including the effects of convective accelerations in the downstream
direction. Elevation changes are exaggerated for illustration [from Dietrich and
Smith, 1983].

curved channels is usually thought to reach equilibrium when the outward


component of the weight of the bed particle on the outward sloping bar
surface is exactly balanced by inward component of drag caused by
This
curvature-induced secondary circulation [see review in Hill, 1987].
hypothesis, which implies sediment travels through bends along lines of equal
depth, is widely used despite the lack of published measurements on transport
direction which demonstrate the validity of the hypothesis.
In contrast,
outward flow over the bar creates a dynamic balance rather than purely the
assumed static force balance. Growth of the bar occurs because sediment flux
to the bar exceeds removal rate. But the growth of the bar creates convective
accelerations that force the flow toward the opposite bank, directing sediment
away from the bar top into the deepening pool, resulting in equilibrium of bar
top and pool depth. This mechanism was proposed by Dietrich and Smith
[1984a,b] and since has been shown quantitatively to be essential in explaining
equilibrium point bars in river bends and alternate bars in straight sections
[Nelson, 1988]. Odgaard [1986] has also recognized the need to include the
effect of outward directed flow over bars in his model to predict equilibrium
bed topography in river meanders.
This new mechanism for equilibrium bed topography in rivers with bars
motivated in part the second question addressed here. In order to predict
channel deformation from a flow model, the boundary shear stress responsible
for sediment transport must be calculated and the relationship between the
direction of the local boundary shear stress and the sediment transport vector
on inclined surfaces must be predicted.
The task in field studies is to
quantify each of these two parts, and this task is perhaps the most difficult
aspect of field investigations of river mechanics.
Although important
experimental work on initiation and transport on an inclined plane has been
accomplished [Ikeda, 1982], the effects of sediment sorting on rates of
downstream transport have not been included in available sorting models. For
a sand-bedded river, Dietrich and Smith [19841 have demonstrated that the
outward shifting zone of maximum boundary shear stress is tracked by the
zone of maximum bedload transport due to a small, but significant, net
outward bedload transport caused by topographic effects. To our knowledge,

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

no detailed bedload transport data from a gravel-bedded meander, however, is


available from previous studies to compare with these results.
The second part of this paper is devoted to field observations and methods
to calculate boundary shear stress and sediment transport in curved channels.
Both a sand-bedded and a gravel-bedded river are examined and, in contrast
to the sand-bedded site, in gravel-bedded rivers it is proposed that spatial
grain size adjustments rather than net cross-stream bedload transport dominate
response to skewed and shifting boundary shear stress field caused by a curved
channel with a well-developed point bar. It is not clear, however, what the
relationship is between bedload and bed surface grain size in a gravel bend.

Boundary Shear Stress and Topography

Theory
Smith and McLean [1984] showed that, if shear stresses due to lateral
boundary layers are neglected, the vertically averaged equation of motion
expressed in a curvilinear coordinate system can be written as

- p ~ <usun>h
Ull

BE
( Tzn ) b -_ - pgh on

- p

on

2p<usu n>h
( I-N) R

<us>h
1
p (I-N)R - P I-N

a
us

(1)

<UsU n>

<un>h

<un>h
P (I-N)R

(2)

where (Tzs)b and (Tzn)b are the downstream and cross-stream components of
boundary shear stress, hand E are depth of flow and elevation of the water
surface with respect to an arbitrary datum, and Us and Un are the downstream
and cross-stream components of the velocity. The angle brackets indicate that
the enclosed quantity has been vertically averaged.
The fluid density and
gravitational acceleration are p and g, respectively.
The coordinate system
consists of an s-axis which points downstream parallel to the centerline, a
z-axis that is nearly vertical, and a cross-stream n-axis that is positive
toward the left bank. The metrical coefficient in the downstream direction
that accounts for the differing lengths along the s-axis between the inside and
outside of the channel is (l-n) = 1 - n/R where R is the radius of curvature
of the centerline and the sign is given by N. The last three terms in (1) and
(2) arise from spatial acceleration of the fluid that can be generated by
downstream changes in channel topography.
Scaling arguments for single
perturbation expansion by Smith and McLean [1984] and Nelson [1988] show
that in (2), these convective acceleration terms, as well as the cross-stream
boundary shear stress, should be small, reducing (2) to
2

on = - f<us>h
I-N) R

pgh OE

whereas in (1) all terms were of comparable magnitude.

Copyright American Geophysical Union

(3)

Water Resources Monograph

River Meandering

Vol. 12

Dietrich and 'Vhiting

The vertically-averaged continuity equation for steady flow, according to

Smith and McLean [1984], is


1

r=N

8<us>h

as

- <un>h
(1-N )R +

B<un>h

lhi

(4)

which can be solved for <un>h,

=-

<un>h

1-N

8<us>h

as

(5)

On

-w/2

where -w 12 is the right-bank position of the channel with a width "w". In


the case where <usun>h ~ <us><un>h and <u~> ~ <u~> , substitution of
(5) into (1) and (2), and solving for S, the downs.tream water surface slope
(S = (-l/{l-N))(aE/Os)) and Sn, the cross-stream water surface slope (Sn =
BEl On), yields

s = pgn
(TzS)h

<us>

(l-N)g

S - -(' ZS)h _
n -

B<us> + <un> [8<u s>

OS

<Us>

1-N ) Rg

<us>]
-g- ( J I l - (l-N)R

__1_ <us> 8<un>


1-N

--os

<un>

2 ( 1-N) Rg

(6)
(7)

Equations (6) and (7) have been written to compare with Equations (7)
and (8) of Odgaard and Bergs [1988], which appeared to have been derived in
the manner described by Smith and McLean l1984]. Following Odgaard and
Bergs' format, which is that originally used by Yen and Yen [1971], Equation
(6) can be written as

(8)
and (7) as
Sn

Snl

Sn2

Sn3

Sn4

(9)

This makes apparent several differences between the equations reported by


Odgaard and Bergs and that given here. The most important difference is
that they incorrectly write
1 8<us>2h
S2 -- 2gli

as

This clearly cannot be correct, as depth, h, should not be inside the


derivative, because, with the substitution of continuity, the term
1 8<us>h
1-N
03

was replaced with the two other terms in (4). This error, if not corrected
during their subsequent calculation, may substantially overestimate the
In (9), Odgaard and Bergs exclude Sn4, although
magnitude of S2.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

substitutions using the continuity equation in (2) (their Equation (5)) must
produce this term. Nonetheless, this term is quite small.
As Dietrich and Smith [1983] pointed out, deletion of S3 in (6) gives the
equivalent to equation (49) of Engelund [1974]. They argued, however, the S3
should not be left out, both because of scaling arguments that led to Equation
(1) and because their field data showed that .all terms in (1) were of
comparable magnitude. Odgaard and Bergs, on the other hand, present data
from a laboratory flume that purported to show that 83 was insignificant.
They used this result to justify ignoring this same term in Odgaard's f1986]
model for flow meanders. They leave it as a mystery as to why Dietrich and
Smith f1983] argue for the significance of S3 in general, and instead attribute
the difterence in results to a possible effect of radius of curvature to width
ratio.
It is probable that the small curvature to width ratio at Muddy Creek, the
site of Dietrich and Smith's observations, contributed to enhanced bar
development and large topographic effects on the flow, but there are other
First, as mentioned
reasons for the discrepancy in the significance of 83.
above, Odgaard and Bergs' 82 appears to be incorrect and systematically too
large relative to S3. Second, according to their Figure 3, at the entrance to
their laboratory flume bend, the maximum velocity and flow depth were
already near the outside bank. Not surprisingly, this will minimize S3. Third,
in the critical reach where S3 would be largest (between their section 20 and
43), their sections were farthest apart. Consequently, locally large values of
<Un> and S3 may have been averaged with smaller values for the reach.
Finally, the Muddy Creek bend is downstream of a sharp bend which
introduces strongly skewed flow that must go around the point bar in the
study reach, whereas Odgaard and Bergs' bend was preceded by a long
straight reach. Although the width changes considerably through the Muddy
Creek site, the cross-sectional area does not; hence the magnitude of 83 is
largely controlled by downstream varying bed topography.
Dietrich and Smith were not proposing that in all channels S2 and S3 are
of equal magnitude, but rather, in most natural rivers where bars are present
and radius of curvature is continuously changing, these terms are of a similar
order of magnitude. To make this point clearer, we present analyses for two
study sites where measurements were of sufficient density and accuracy to
warrant at least drawing some semi-quantitative conclusions regarding the
magnitude, sign and spatial distribution of the convective acceleration and
centrifugal force terms in (1) and (2) or (8) and (9).

Analysis of Field Data


Dietrich and Smith [1983] showed that, to analyze cross-stream velocities
correctly, field data on horizontal velocity vector fields taken at successive
sections through a reach of channel needed to be oriented with respect to
section orientations established according to continuity.
This requirement
arises because of the difficulty of recognizing proper section orientation by
visual inspection on natural rivers. Using the coordinate system defined above,
they showed that the cross-stream discharge Qnw of water required by
continuity is

w/2

Qnw =

<un>h dn

-w/2

-1
[ 1-N

fn
-w/2

ikus>h dnJ dn

as

Copyright American Geophysical Union

(10)

Water Resources Monograph

River Meandering

Vol. 12

Dietrich and \Vhitillg

and reasoned that correct orientation of a section was one that gave the
cross-stream discharge determined in (10).
Observed Qnw can then be
corrected to that computed from (10) by section rotation. There are many
subtleties in applying (10) to field that, in general, will not alter the gross
cross-stream field predicted, but can have significant effects on terms in (1),
(2), (8), and (9). A common problem is the effect of discharge variation
between sections.
Variation in discharge arises in two manners. Real variation in discharge
may occur as a result of small stage fluctuations that are almost unavoidable
in most natural rivers. Apparent variation in discharge may occur as a result
of the integration of measurements across the sections and uncertainty in how
to properly project near-boundary measurements to the no--slip boundary,
especially at the banks.
If Qsw is the downstream discharge, then

w/2
Qsw

(11 )

<us>h dn

-w/2

w/2

J <us>hdn

-w/2

and, according to Leibniz rule

w/2

-w/2

<us>hdn

which, using (5), gives

(l~N) ~ =

<un>h at w/2

that is, the downstream changing discharge will equal a calculated influx or
outflow at the channel bank. If this term is small relative to calculated Qnw,
then this error can be ignored. Otherwise, all sections should be normalized
to the same discharge. The simplest normalization is to multiply each local
<us> h value by a correction factor equal to the ratio of constant downstream
discharge divided by the individual section discharge.
Equation (10) requires computing the difference in local downstream
discharge between sections. . Ideally, the difference would be calculated between
the first and third section for checking the orientation of an intermediate one.
Regrettably, this requires very close spacing of sections, which is not practical
in most field settings. One approach employed by Dietrich and Smith [1984]
is to do a forward difference between two successive sections, but write
8<us>h _ h 8<us>

iJd

<us>

8h
os

(12)

and use h in the first term and <us> in the second at the lower of the two
sections for which the cross-stream discharge is calculated. This has the effect

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

of biasing the results toward conditions at the section for which the
orientation is to. be determined. We have compared this procedure with a
standard forward difference using the left-hand side of (12) only and found the
two procedures give results that differ by usually less than 1 or 2 degrees.
We feel the procedure using (12) yields a more accurate result.
Another problem in using (10) is accounting properly for effects of width
change between sections. Although uncertainty of Qnw is relatively small, near
bank values of <Un> are greatly affected by this problem and, consequently,
so are terms in the force balance. We dealt with this problem by defining a
regularly spaced interval of calculation points either side of the centerline
between successive sections. If a large change occurred in width, then local
near bank values of <Un> were relatively large. In effect, this assumes that,
in the short distance between sections, most of the cross-stream flux to
accommodate width changes is concentrated near the bank.

A problem that has a moderate effect on (10), but a large effect on force
balance calculations, is the location of the centerline of the channel. Near
bank eddy zones, partial overflow of grassy partitions, and similar problerl1s
can introduce important uncertainty in location of the centerline. We do not
know a simple quantitative procedure to deal with this problem.

24m
............_ ...'
distance scale

25

flow velocity scale

~1r~ij%;~

near-bed flow toward right bank

~
o 40 80 cm/s

Fig. 2.

Planform map of study bend at Muddy Creek showing vertically-averaged


downstream velocities at individual sections. Areas with near-bed cross-stream
velocities oriented toward the right bank (inner bank) are stippled; otherwise
flow is toward the left bank.
The thick dashed line traces the left bank
position in 1976, while the thick solid line traces the bank in 1978.

Force Balance Analysis


The site and methods of data collection at Muddy Creek are thoroughly
described elsewhere [Dietrich and Smith, 1983, 1984; Dietrich et al., 1979,
1984]. Figure 2 shows the section location, and vertically-averaged velocity
field and near-bed flow direction through the bend. The Muddy Creek study

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Dietrich and \iVhiting

bend during most years of measurement had a median grain size of 0.7 mm, a
water surface slope of 0.0014, a mean depth and velocity of 40 em and 55
em/sec. During the critical year of measurement, 1978, the channel had a
minimum radius of curvature to width ratio of 1.5. Because of the upstream
bend, flow enters the study reach with the highest velocity near the inside, or
right bank (Figures 2, 3).
The flow slowly decelerates over the bar and
progressively accelerates over the pool, with the centerline area experiencing
the least velocity variation.
The data presented on the Muddy Creek convective acceleration terms were
computed in two ways. For terms in (1) and (2), with derivatives in the
downstream direction, calculated values of verticalfy integrated velocities at
each measurement point were interpolated to values at a regular interval
(every 50 cm) across the channel and differences were then determined. For
analysis of (8), interpolated values of <us> were used to compute <Un> via
~~~: and these <Us> and <Un> values were used to determine all terms in
Comparison
acceleration,

of

Figure

with

Figure

shows

that

the

convective

is high over the bar due to shoaling (80./ Os) and large in the downstream pool
2

region due to flow acceleration (8<u s> / Os).


location of nlaximum centrifugal force,

Figure 4 also indicates that the

<us>h

p (I-NJR

shifts across the channel, causing the cross-stream water surface to be convex
up in the upstream part of the bend and concave up in the downstream
region. Figure 5 shows the observed and predicted (Equation (3)) cross-stream
water surface profile at Muddy Creek, generally, supporting the use of (3)
rather than (2). Data for water surface topograJ?hy and channel topography
were collected in 1979 [Dietrich and Smith, 1983J, a year after the velocity
measurements, so some differences between predicted and observed are
attributed to small changes in flow and bed topography, although the
discharges were the same. Also, as will be shown next, some of the terms
neglected in (3) are locally significant.
The four terms that sum to the total downstream boundary shear stress in
(1) are portrayed in contour maps in Figure 6. As Dietrich and Smith [1983]
discussed, it is difficult to evaluate terms with large spatial derivatives
accurately from field measurements. Nonetheless, Figure 6 shows clearly two
features of these terms: the first two spatial acceleration terms are as large or
larger than the local pressure gradient term, and the contribution in magnitude
and sign of these terms to the overall force balance in (1) varies systematically
through the bend.
These two convective acceleration terms most directly
controlled by topography (the second and third maps of Figure 6) are of
comparable magnitude but opposite sign throughout the bend. Generally, the
last term, which is controlled primarily by channel curvature, is only
important in the zone of curvature minimum, where the cross-stream velocity,
<un>, is also relatively high.
Because of the tendency for the spatial
acceleration terms to balance each other, the local boundary shear stress is
approximately equal to the local pressure gradient throughout the bend. In a

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

10

Boundary Shear Stress

~!

51

~I

221

~
o

~!~~21
o

~!---.-----.-----~191
o

~!~~181
o

~!-----.-----.---~141
o

:f---~----'----'
~12l
2:t

~1~101
o

300

200

100

-100

-200

300

Distance from centerline (cm)


Fig. 3.

Vertically-averaged downstream flow velocity at Muddy Creek sections.

sequence of river meanders with well-developed bar topography, this tendency


for the dominant convective acceleration terms to be counterbalancing is
probably common. Although the gross downstream vertically-averaged velocity
field can be estimated without including the convective acceleration terms, the
local boundary shear stress and consequent bedload transport may not be
accurately determined [Dietrich, 1987].

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

11

Dietrich and Whiting


500000
----0---

.-

400000

-.3

..c

N"::J
v

----.--

--0--

~
..........

Me

19

---0--

-...-

--0--

300000

20
22
24
25

200000
100000
0

S
-.3
..c

0
40

Po.
Q)

80

300

-300

300

-300

Distance from centerline (cm)


Fig. 4.

<U s>h at Muddy Creek sections and the associated cross-stream depth profile
illustrating the large convective acceleration over the point bar is due to
shoaling, while over the downstream pool the convective acceleration is due to
the flow acceleration.
The figure indicates the position of the maximum
centrifugal force shifts across the channel causing the cross-stream water
surface profile to be convex up in the upstream portion of the bend and
concave up downstream.

In order to define the relative importance of the cross-stream convective


acceleration terms, each value was divided by the magnitude of the
corresponding centrifugal force term at each location. The resulting contour
maps (Figure 7) indicate that, like the convective acceleration terms in the
downstream direction, these two terms tend to be of opposite sign and similar
magnitude through the bend, reducing the error caused by neglecting these two
terms in (3). Over the upstream part of the bar (sections 14-19), however, it
appears that the negative terms exceed the positive ones.
This imbalance
would reduce the cross-stream pressure gradient, and cause the centrifugal
force to exceed the cross-stream pressure gradient, resulting in outward flow
over the bar. Hence, these generally small terms may be locally important in
defining the effects of channel topography on flow.
The conversion of (1) to (8) highlights the role of <Un> in advecting
momentum across the channel by showing its contribution to the two parts of
the S3 term. Figure 8 shows the cross-stream structure of <Un> calculated
two ways: from use of the downstream velocity, <us>, in (5), and from
measured cross-stream velocities oriented to a section direction defined by
continuity constraints. In general, the predicted and observed <Un> are close.
In the upstream part of the bend (12 to 19, 24 to 25), <un> values reach 20
to 35% of the average <us> for the channel of 55 em/sec. This suggests S3
should be significant.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

12

::1 1~ ' 3
J
057.

I:::
~------:

0.54

Q~

Q~

I[

::

124:=:~

O~$

1:

I:;
I~

14
0.58

t---------~_-----.....;

0.57~

0~6

0.54

O:~
::112:: : : :j 120~~_1:::
::[
I::
j

0.57

.... 54

300

200

100

100

-200

Distance from centerline (cm)

Fig. 5.

-3:J0

300

200

100

-100

200

-300

Distance from centerline (cm)

Predicted and observed cr~tream water surface profiles at Muddy Creek.


Water surface and bed topography data were collected in 1979, while velocity
used in prediction of water surface profile was measured in 1978.

To calculate terms in (8), we have rewritten (6) as


S

CD <u s >2

2gh

2(1-N)g

O<us>2 + <un> o<us>

as

-g-

-----an- -

and (8)

Copyright American Geophysical Union

<un> <u s >


- g - (l-N)R

Water Resources Monograph

River Meandering

Vol. 12

Dietrich and Whiting

13

25

Fig. 6.

Contour maps of the value of the components of the downstream force balance
at Muddy Creek:
a) downstream pressure gradient force, b) convective
acceleration associated with the change in momentum of the downstream flow
in the downstream dir~ction, c) convective acceleration associated with the
change in momentum of the downstream flow in the cr~tream direction,
and d) force associated with the channel curvature.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

14

25

4m
t:=e::::c::::r::::::::t'

Fig. 7.

Contour maps of the logarithm of the ratio of terms in the cr~tream force
balance for Muddy Creek: a) ratio of force due to the change in momentum
of the cross-stream flow in the cross-stream direction to the centrifugal force,
and b) ratio of the force due to the change in momentum of the cro~tream
flow in the downstream direction to the centrifugal force.
Areas of negative
sign are shaded.

(13)
replacing Tb with p(C D /2)( <us>2), where CD is held constant at 3.61x10- 2
[Dietrich, 1987, p. 202] and S3 has been broken into its two components, S3A
and S3B. By holding CD constant, we ignore spatial structure in boundary
shear stress due to differing sizes of bedforms through the bend, local effects of
lateral boundary layers, spatial variation in bed surface grain size and other
such features that impose local structure on CD. Given the uncertainty in
this analysis, correction of CD due to these features seemed an unnecessary
exercise. The terms in (13) were calculated for the interval between sections
and the values were labelled with the downstream section number.
In Figure 9, each of the S components and the sum are plotted, and the
sum is compared with observed slope in Figure 10. The important result
depicted in Figure 9 is that in the upstream parts of the bends where
shoaling-induced outward flow is occurring (section 14-18, 25), S3A and S3B
are large but of opposite sign, having the effect of making S3 relatively small

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Dietrich and Whiting

15

:~125:~:1

-40

:~I~.~~,~~"""
~m : 1
:~I~~~.:.:

. .I

bU

00
S

20 1 .. ..

10'

..

..

=====::::::::::::::::::::=.. .. . .. ...j.

.. .. :1"

60

.,20

40

1----------,ii~+J=..-1

/-,

20

..-

""~..

'161

1
1

20
40
100
80

~
0

,.-..4

r.n.

19

60
40
20

-20

~11~:'~:m:1
-40

:~ 1------'
. 14 -::---.-------..........................
-----'
....... ~.....
...-1
o ---- .

-20

300

200

100

-100

-200

-300

Distance from centerline (cm)

------I..---'--........--""""'-~....&.......--

:~l1~ ~
Jor

20C

10r

-, 00

200

j
-300

Distance from centerline (cm)

Fig. 8.

Comparison
sections.
continuity.
downstream

Fig. 9.

Components of 8 from Equation (8) plotted at each section.


is the sum of 81, 82, 83a, and 83b'

of measured and predicted cross-stream velocities at Muddy Creek


Observed velocities oriented with respect to section satisfying
Predicted velocities calculated from the downstream changes in the
velocities.
The total slope

but not zero. This result is similar to that found in Figure 6. Overall, 8g or
8gA + 8gB is significant in this bend. 82, on the other hand, is largest at
the crossing between the bends, where rapid changes in channel curvature and
bed topography occur.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

16

~1- -,- - -,2:'~- . . . . -" :" " "':': : : " :"~: :" " "

: '""'"--'------::
..1

:~I?~_~~m1

20

:~~~

100 19
80..._
~

601---------- ---

40

20

r::n

20

o"___----'-_.. . . . . . ._--'----.::........L..-_....a....--.I
300

200

100

100

200

300

IJistance from centerline (cm)


Fig. 10.

Comparison at sections of the calculated S from Equation


measured S from the observed water surface topography.

(8)

and

the

In general, predicted and observed slopes (Figure 10) are similar in


magnitude and cross-stream structure. This is particularly encouraging because
elevation differences between sections were generally less than 1 cm.
The
largest disagreement, at the inner bank of 12 and 19, are probably not due to
calculation or measurement error. At section 12, there is a local effect of a
lateral wake formed by the indentation of the bank just above section 10
(Figure 2) that appeared to deflect the high velocity toward the centerline
(Figure 2) and locally increase the downstream slop between 10 and 12. This
effect was also unsteady, causing considerable error in attempting to surve~ the
local surface elevation of the water. Measurements in a subsequent year l1981]
did not detect such a large water surface gradient; instead values much closer
to that predicted were found. At section 19, the difference between observed

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

17

Dietrich and Whiting

and predicted slope may be related to local differences in channel topography


between 1978 and 1979 [see Figures 2--6 and 2-7, Dietrich, 1982J.
The
disagreement at section 19 is probably not due to error in field measurements
because in 1983 water surface measurements also revealed a very low
downstream gradient between sections 14 and 18 near the right bank.
In order to examine the effects of bed topography more directly, detailed
field measurements were also collected in Solfatara Creek (Figure 11), a 5.1 m
wide, sinuous gravel-bed channel near Norris Junction, Yellowstone National
Park, Wyoming.
The reach of channel studied consists of straight reach
Flow exits the bend and is rapidly
downstream of a bend (Figure 11).
shoaled by the downstream bar, causing acceleration over and diversion around
the bar. The data reported here were made at a discharge of 1.08 m 3 /s,
which is 45% of the estimated bankfull discharge. The average depth was
0.5 m and the water surface slope was .00095. These conditions correspond to
a total boundary shear stress of 50 dynes/cm 2, a value below the critical
boundary shear stress for entrainment of the 9 mm median grain-size bed.
Locally, sand was transported as bedload. At this moderate flow over a bar
formed at higher discharges, topographically-induced convective accelerations
are expected to be pronounced.

I
Fig. 11.

contour interval

0.2 m

Topographic map of the .study reach at Solfatara Creek.


shaded.

The pools are

The discharge, near depth and total boundary shear stress, were nearly the
same as that for the Muddy Creek data set. Despite very large differences in
curvature and bed grain size between Muddy Creek and Solfatara Creek, the
spatial pattern of the magnitude and sign of the convective acceleration terms,
S2 and 83, were found to be quite similar, demonstrating the dominant control
of bed topography on these terms r Whiting, in preparation].
The magnitude of 8 , normafized by the absolute value of the other
3
convective acceleration term, S2' is shown in a map of the data for Solfatara
Creek in Figure 12. Because the. channel is essentially straight, 8 3B is zero
is used. This map clearly shows that, at least in this moderate
and only 8
3A
flow case, the magnitude of 83 is greater than 0.1 82" The most striking

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Boundary Shear Stress

18

Fig. 12.

Vol. 12

Contour map of the logarithm of the ratio of cross-stream (S3a) and


downstream (S2) transport of downstream momentum at Solfatara Creek.
Positive areas are shaded.

feature is the diagonal shift of the positive zone across the bar. Positive S3
results in cross-stream advection of high momentum fluid toward the right
bank and into the pool. The greater sampling density at Solfatara than at
Muddy Creek appears to have resulted in better defined structure to the
convective acceleration terms. Attempts to predict boundary shear stress in
(1), however, still yielded unrealistic results in parts of the channel; as in the
Muddy Creek case, there were large uncertainties in the convective acceleration
terms.
Our analysis of the terms in (1), (2), and (8) for our two study sites
shows that all the convective acceleration terms are important to the fluid
force balances. The scaling by Smith and McLean [1984] and Nelson r19881 of
(1) applies equally to (8), so it should not be surprising that S3 was round to
be a large term. In a sense, it is generally assumed that, if a theory predicts
reasonably well the vertically averaged downstream velocity or the gross bed
topography of a bend, then the model is correct physically. We feel, however,
that the most direct test of a theory is to examine whether all the forces
In this way, counteracting effects
driving the fluid are correctly modeled.
which otherwise would not be discovered can be examined for their
significance.
Boundary Shear Stress and Sediment Transport

Methods
Many scales of resistance contribute to the total boundary shear stress
determined in ( 1) (Figure 13).
In order to apply the results of ( 1) to
predicting sediment transport, however, the appropriate boundary shear stress
is just that balanced by the combination of static grain resistance and
momentum extraction due to bedload transport [e.g., Bagnold, 1973; Smith and
McLean, 1978].
In a theoretical analysis, values from (1) must be

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

19

Dietrich and Whiting

Boundary shear stress components


1) Grain resistance

2) Bedload resistance

3) Bedform resis tance


4} Bar resis tance

5) Bank and planfor m


resistance

Fig. 13.

~------

Scales of resistance to flow in natural channels.

systematically reduced, accounting for effects of topography and bedform


resistance, whereas, in the field, the challenge is to evaluate the individual
components illustrated in Figure 13.
A large number of theoretical and
empirical formulae exist to account for form drag due to bedforms, although
few field studies have fully evaluated these formulae [Le., Smith and McLean,
1978].
Table 1 lists nine methods to estimate from field measurements the local
boundary shear stress. The first three only give the boundary shear stress
causing sediment transport if the channel is straight and free of bedforms
(making (1) collapse to method (2), anyway). Method (4) is the only direct
measurement of shear stress; all others are derived from theoretical models.
After (2), the most widely used method in the analysis of field data is (5),
but this strictly can only be employed for velocity data close to the bed
0.2 flow depth) and even then unsteady wakes from migrating bedforms on
sand beds and from the stationary or moving coarser fraction of gravel rivers
makes data from (5) regrettably crude. Method (3) is the most common one
used to relate mean velocity to boundary shear stress in theoretical predictions.
Method (8) was, in effect, proposed by Middleton [1976] although he didn't use
it to map boundary shear stress, and we are not aware of examples where (7)
and (9) have been employed to map the spatial distribution of boundary shear
stress in rivers. Method (6) is similar to the Preston tube method [e.g., Nece
and Smith, 1970], in which a single, near-bed velocity measurement, coupled
with an argument for boundary roughness, is used to calculate from (5), the
local boundary shear stress. This method has been employed by Dietrich and
Smith [1984] and Dietrich et ale [1984] to predict bedload transport fields in
Muddy Creek at high and low f1ow. Considerable field measurements have
been collected to explore this approach and these data and a theoretical
analysis will be presented elsewhere [Dietrich and Whiting, in preparation1.
Because we make use of this approach here, we will briefly review the method.

Single - velocity method.


A simple approach to estimate zo, the roughness term in method (5), for a
mobile bed is to hypothesize either that Zo is proportional to the saltation

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

20
Table 1.

Methods to estimate local boundary shear stress.

Equation

Comment

Eq (1) of this paper

Difficult to define all terms


accurately in field studies.

Tb

~1ethod

pghS

Only approximately correct in


short reaches because of
convective accelerations.

Cn is not a constant [see


Nelson, 1988, for discussion].

Turbulence measurements have


proven difficult to obtain in
natural rivers with mobile beds
[see McLean and Smith, 1979].

Tb

zo

p(UK)2
(In(z/zo))2

f(D) and (5)

Difficult to make reliable


velocity profile measurements
over mobile bed and beds with
large grain sizes; need to make
profile measurements very close
to bed to avoid form drag of
bedforms.
Need data on bed surface grain
sizes; zo/D 84 may not be a
constant for all problems.
Results depend on bedload
equation used and, if necessary,
accurate estimate of critical
boundary shear stress.

Suspended load may be highly


variable; not clear which
representative settling veloci ty
to use.

Only applies to gravel bed


rivers with Tb/ Tc close to 1.0
at stage of interest.

height of the moving grains due to momentum extraction required to cause


bedload transport [Le. Smith and McLean, 1978], or that zo is controlled by a
representative coarser fraction of the moving or static bed surface.
In
expanding upon the saltation approach formulated by Smith and McLean,
Dietrich lI982], in collaboration with Smith, included a drag force neglected by

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

21

Dietrich and Whiting

the Smith and McLean theory which is exerted on the ascending particle, and
formulated a semi-theoretical model to predict the height of saltating grains,
CB,
(14)
where J is the friction angle of the grains resting in the pockets. Wiberg
and Smith [1985] have subsequently derived from fully theoretical grounds a
saltation model which confirms the general form of (14), although an
additional grain size dependency was included. Equation (14) does predict the
very limited available data on saltation heights of grains quite well. Dietrich
proposed that
(15)
such that
(16)
where at was an empirical constant. The resistance due to static grains at
the bed surface, Zn, is calculated from the Nikuradse diagram [Le., Smith,
1977], and is generally much smaller than at c5B .
To evaluate at, Dietrich made simultaneous near-bed velocity and bedload
transport measurements, then used these data to solve iteratively for the 0'1
that .gave boundary shear stresses that, when used in the Valin bedload
equation, gave the least variance between predicted and observed bedload
transport rates. He used velocity data from a low flow in 1977, collected at
2 cm above the crest of migrating bedforms, and calculated bedload flux based
on simultaneous observations on bedform migration rate and from a few
samples collected with a 7.62 cm Helley-Smith sample. In using the Valin
bedload equation, he asSUllled that the median grain size of sediment collected
in the bedload sample, 0.6 cm, applied to all samples. The least-squares fit
to the velocity and transport data (Figure 14) yielded an 0'1 of 0.077. This
value is based on Von Karman's constant, k, equal to 0.43, a value thought
appropriate at the time of his analysis. The computed bedload transport fields
closely matched the observed for the low flow data [Dietrich et al., 1984] . In
addition, Dietrich and Smith [1984, p. 1369] have shown that the bedload
transport was on average predicted to within 5% of observed for Muddy Creek
in the high flow year of 1978, even when the near-bed velocity data used in
the prediction were collected two years previously from a similar stage.
Because of the imprecise nature of the data in Figure 14, 0'1 is not welldefined by this test, but as Figure 15 shows, the estimated boundary shear
stress is only weakly dependent on at within a reasonable range of error. The
method has the great .advantage over velocity profile data of relying only on
velocities at a single position above the bed; hence, effects of wakes of
upstream bedforms can be avoided. In addition, when one is constrained to
work with only one current meter, a single near-bed velocity measurement is
considerably faster and probably more reliable than velocity profiles.
Dietrich [1982, p. 105-108] showed that Hooke's [1974, 1975] experimental
data could be reinterpreted using the above method. Hooke used a Preston

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

22

E
u

BEDFORMS NORMAL TO FLOW

en

o:
o
c..
en
z

o~

OCP
00

7t/t
/+ 08

~
0

~o ~8

0 0
IOd'

.1

8~

at- '"

00

t-

eel'
g -I9t-

++ +.D"

"'-

",

o /.

+ HELLEY-SMITH SAMPLES

Q)

07

BE OFORMS OBLIQUE TO FLOW

.0

cd:

a::
tO

<X

.01

C
W
CD

.00 I

......................-.._.-....-__

----..II.....-""""-

I 00

I000

10,000

U AT 2cm ABOVE BED, (cm/sec)2


Fig. 14.

Measured bedload transport as a function of average flow velocity at 2 cm


above the bed. The curve is the least squares fit of the combined hopping
height model for roughness and the Valin bedload equation in which 0'1 was
found to be 0.077.

tube to measure velocities at approximately 3 mm above the bed to estimate


boundary shear stress and plotted measured sediment transport rate against
this calculated boundary shear stress (Figure 16). The dashed line and the
corrected axis are the results predicted using Hooke's velocities converted from
boundary shear stress using his drag coefficient, a constant grain size of
0.03 cm in the single velocity method outlined above, and the Valin bedload
equation. The difference in predicted boundary shear stress is due to the
inclusion of bedform resistance in the drag coefficient calculation by Hooke.
Based on reported bedform geometry and the Smith and McLean [1978] form
drag correction, this drag coefficient was probably about 2.4 times too large, a
value consistent with the difference between abcissa values in Figure 16.
Since these initial applications, a large data set on paired near-bed velocity
and bedload transport rate have been collected for sand and gravel [Dietrich

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

23

Dietrich and Whiting


-

0.25

~urved ..tiffed ~ ~~predicfed

.~.
.

by eye by
Hooke

~ 0.20

by model

........

rtl

0.15

t-

Il::

oQ.

25
N

.........
fJ)

Q,)

c:

20

tt-

15

___- - - - - - - 46.6

10

"'C

h. '.

~ 0.10

I '.
Ii'

h" .

<t
Il::

Z
LaJ

/> :

:l~::

0.05

:IE

,'.2. ".

en o.a

LaJ

o
.05

.10

..

-I.

L.....-......---..... -.-L.........-_""""""--_
o
10
20
30
2
SHEAR STRESS (dynes/em [HOOKE]

4. I

8 .9

17.0

SHEAR STRESS PREDICTED


2
BY MODEL ( dynes /em )

ROUGHNESS PARAMETER
COEFFICIENT (eL)
Fig. 15.

Variation in predicted boundary shear stress with at for two values of velocity
at 2 cm above the bed.

Fig. 16.

Variation in sediment transport rate with boundary shear stress in the


laboratory meander studied by Hooke [1974]. The solid line is the curve fit
by eye by Hooke to the data. Lower scale on the abscissa is the boundary
shear stress predicted with the roughness scaled by the saltation height and
the velocities measured by Hooke at 0.3 cm above the bed composed of 0.03
cm grains. The dashed line is predicted by the model (Equation 16 and the
Yalin bedload equation).

and Whiting, in preparationJ. Analysis of these data with a more appropriate


von Karman constant of 0.40 yielded an at of 0.109 and of 0.121, using the
Valin bedload equation and the Luque and van Beek [1976] equation,
respectively. The higher values of at largely result from the reassessment of
von Karman's constant to 0.40; the computed boundary shear stress is nearly
the same (within about 5%) whether at is 0.077 and k is 0.43 or at is closer
to 0.11 and k is 0.40.
It is fair to question precision here and a first
approximation of at is 0.1.
An alternative approach was taken when it was realized that in
poorly-sorted sediment, the largest grains commonly only roll and the saltating
grains do not hop above these rolling ones. Another formulation, even more
empirical than the first, but steeped in a long tradition, is to propose
_
ZQ -

AD x

f( R*J

(17)

where A is an empirical constant, Dx is a representative length scale of the


grains controlling boundary resistance (equivalent to k in Wiber.Q and Smith

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

24

[1987]) in which x is the percent finer than that size fraction, and R. is

u.Dxlv, the Reynolds roughness number equivalent to that in the Nikuradse

diagram, which ranges from 45 to 30 for sand to gravel. Following the logic
of others that one standard deviation from the median, D84 , is a good measure
of boundary length scale [Leopold et al., 1964], we used a least squares
procedure involving the Luque and van Beek [1976] bedload equation to
evaluate A, and found the best fit value to be about 3.5. Note that, if f(R.)
is 30, then (17) becomes
Zo ~ 0.1 D84

(18)

a result similar to (15), where O't = 0.1, and t5 , the hopping height of the
B
median grain size, is approximately equal to D84. The best fit value using the
Valin bedload equation is not substantially different.
When grain size
distributions are only approximately known, equation (18) is relatively easy to
use and appears to give a reasonable estimate for sand and fine gravels. The
details of this analysis are given elsewhere [Dietrich and Whiting, in
preparation] .
Comparison of Boundary Shear Stress Estimatel
In order to show the effects of various scales of resistance, we will first
show a com:parison of methods (2) and (3), and then compare results of (3)
with (5), (6), and (7). The analysis presented earlier in this section would
suggest that, due to large convective accelerations, boundary shear stress
calculated from method (2) would differ substantially from that of method (3)
in some parts of the bend. Figure 17 shows that the two differ most over the
downstream pool (sections 19 and 20) and at the entrance to the central bend
(section 12). The steep slope along the right bank at 12 was discussed earlier.
The general magnitude and structure of the boundary shear stress fields are
similar, although method (3) tends to predict higher values at several sections.
To use method (6), we computed hopping heights and Zo values from
equations (14) and (15), respectively, as described above for single-velocity
measurements at dune crests. Individual data points are shown in Figure 4-11
of Dietrich JI982].
Predicted bedload transport values for each observation
were then ivided by two to estimate the average bedload transport rate
rDietrich and Smith, 1984], and, as mentioned above, the observed bedload
field was predicted quite accurately by this method. The pattern of boundary
shear stress shown in Figure 18 is similar to that of Figure 3 for the
vertically-averaged velocity:
the boundary shear stress declines downstream
over the bar, increases over the pool and varies the least along the channel
centerline.
Fi~ure 19 shows the boundary shear stress field calculated from methods
(3), (5), (6), and (8). The methods for collecting data for (5) and (8) are
described in Dietrich et ale [1979] and Dietrich and Smith [1984J, respectively.
In general, all four methods show the same cross-stream structure and differ
primarily in magnitude.
As expected, method (3), in which the drag
coefficient, CD' was held constant and include~ all length scales of resistance,
was systematically highest. The velocity profile data were initially thought to
be close enough to the bed to exclude the effects of drag from bedforms
coverin~ the stream bed rDietrich et al., 1979, p. 310].
However, as Dietrich
et ale l1984] reported, and clearly showed in Figure 19, these data lie entirely

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

25

Dietrich and Whiting

----0-

100

from drag coefficient


from pressure gradient

- . - - - -..

O~----&.--------.....a...._...a....--..L-IoI-~

t:~I~~I
i':I~~1

_._...

-100

1:1~.~4_~~
_~.m_1
~
-100 ----"-----'-

--1---....._

::11~.:~
, - :
---'--.....0....-....1.....-"""'---

1.

~..

100

0----'-----'-.:..............a....-...J...--..L--~----'--

300

200

100

-100

200

300

Distance from centerline (cm)


Fig. 17.

Comparison of the total boundary shear stress calculated with a drag


coefficient equal to 0.0361 (method 3) and the total boundary shear stress
calculated with the downstream pressure gradient force (method 2).

above the single velocity data which accurately predict bedload transport.
Hence, the velocity profile data include form drag effects. The boundary shear
stress estimates from suspended load (method 7) were computed as the square
of the median settling velocity from samples collected using a U.S. DH-48
sampler. The relatively large sampling error, the approximate nature of the
suspension criterion of Ws = u*, and uncertainty about the appropriate settling
velocity class to use, makes this estimate fairly crude. Also, diffusion-based
theories for suspended sediment profiles [e.g., Hunt, 1954] predict a declining
settling velocity with distance above the bed. Because the suspended load was

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

26

fP1jj..
8

19

20 22

14

12

en
en

24 25

40

a:
...-en
30
N

e
a: u

<t .......
~ : 20

en ~

"'0

10

<t

o
Z

::>
~

0 """--"""---'
200
+300

100

100

200

-300

DISTANCE FROM CHANNEL


CENTERLINE (em)
( TOWARD LEFT BANK)

Fig. 18.

(TOWARD RIGHT BANK)

Downstream variation in the structure and magnitude of the boundary shear


stress field.
Maximum boundary shear stress shifts from near the right bank
to near the outer left bank through the bend.

low and concentrated near the bed [Dietrich and Whiting, in preparation], we
think the calculated settling velocities are biased toward the near-bed values.
Despite the uncertainty in this method, the structure and magnitude of the
estimates are generally similar to that of the single-velocity method. We also
calculated the boundary shear stress field with the coarsest reliable settling
velocity value, the 95 percent size, for each site and found that the estimated
boundary shear stress was closest to that of method (3), that is the local total
boundary shear stress.
For both high and low flow at Muddy Creek, Dietrich et al. [1984l
computed the average total boundary shear stress (method 2), the average 0
that estimated from velocity profile measurement (method 5), and the average
from the single-velocity calculations (method 6).
These represent the
boundary shear stress associated with all roughness scales, that associated with
bedform drag and skin resistance, and that associated with skin resistance.
Figure 20A shows the results of their analysis. To test the validity of these
results, they compared predicted ratios of method 2 to 5 and 5 to 6 for each
stage with observed values, using the Smith and McLean [1978] method. In
Figure 20B, predicted and observed values are compared, and the agreement is
fairly good. In that analysis, the flow over the bar was assumed to separate,
which isnIt strictly correct. However, application of the simple drag equation
proposed by Nelson [1988, p. 43] yields similar results when a drag coefficient
for unseparated flow is used. The systematic nature of the data shown in
Figure 20A and the general agreement between predicted and observed values
gives further support to the conclusion that these calculated boundary shear
stress values are accurate.
Figure 20A also illustrates one other important point, stressed by Dietrich
et al. [1984]. In sand-bedded rivers, the bed generally remains mobile through
a large range of stages. Point bar crest height or amplitude, which largely
determines the form drag associated with bar resistance, should increase with
higher stage as the bar top builds and pool deepens.
Consequently, the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

27

Dietrich and Whiting


100

25
80
60
40
20

120

----tr---

friction coefficient

_0 _ _ ___ - _ _ . __.

----...-

vel~_t!. ..~~file

80 -,
60
40
20

100

22

80
60
40

20

as
(,,)

Q,)

s::

~
en

,.
en
,.

00

40

20

Q,)

CO

Q,)

..c:en

19

,.

"'0

s::

::s0

~I-

-I

100

~;=

~I

300

~~=
200

100

-100

~I

200

300

Distance from centerline (cm)

Fig. 19.

Comparison of methods 3, 5, 6, and 8 for computing the local boundary shear


strees.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

28
N'

UJ
UJ

~UJ
~

as
Q,)
~
UJ

3.0

60
50
0

40

:0

30

2.5

as
~

Q,)

i:

20

2.0

UJ

.a
0

10

s:::

:s0

r:Q

s:::

~:B

UJ;E

Fig. 20.

@~

cS~

Q,)

1.5

as

form drag/skin friction -- high


form drag/skin friction -- low
totaVform drag -- high
totaVform drag -- low

[]

1.0
..0

1.5

2.0

2.5

3.0

Predicted ratio

Variation with stage in the boundary shear stress due to skin friction, and
bedform drag, and their total (a) ,and the proportion of the total boundary
shear stress taken up with each component (b).

portion of the total boundary shear stress controlled by bar form drag
resistance should increase with rising stage, as the bed topography adjusts.
Figure 20A clearly shows this effect. Although the total boundary shear stress
doubled between the low and high flow, the sediment transport controlling
boundary shear stress increased by only 25 percent and, as a result, the
bedload transport rate only increased by 2.2 times. This illustrates the need
to account for change in form-drag resistance with stage when computing
bedload transport rate in sand-bedded rivers.

Sediment Transport in a Sand-Bedded Meander


The data collected at Muddy Creek [Dietrich et al., 1979, 1984; and
Dietrich and Smith, 1983, 1984] provide a unique opportunity to examine the

relationship between boundary shear stress, bedload, suspended load, and grain
size sorting processes through a bend. Figures 21 and 22 show that the rone
of maximum bedload transport, the position of the maximum transport rate of
the coarsest size fraction in the bedload, and the median grain size of the bed
closely track the zone of maximum boundary shear stress as it shifts across
the channel into the pool through the bend.
The position of maximum
transport rate of sizes finer than the median grain diameter of o. 7 mm,
however, lie in lower shear stress areas progressively across the channel, except
where the sediment crosses the channel (sections 19/20-21).
As discussed
extensively elsewhere [Dietrich et al., 1979; Dietrich and Smith, 1984; Dietrich,
1987], the shoaling of the flow over the bar forces near-bed flow and,
consequently, bedload transport across the top of the bar (sections 18-19/20).
In the pool and along the steeper sloping surface of the point bar, the near
bed flow direction is strongly inward. At the edge of the bar top, the coarse
sediment travels against this inward component of the secondary circulation by
rolling, avalanching obliquely down from the crests of migrating dunes on the
side or face of the point bar, and by being transported by trough-wise
currents of obliquely oriented dunes. The finer sediment crosses the coarse
sediment as it is carried inward from the deeper water and up onto the
downstream end of the bar by the inward directed boundary shear stress
associated with the secondary circulation and flow in the lee of obliquely
oriented bedforms. The mechanics of this sorting process is discussed further
by Dietrich [1987]. Sorting occurs because grain weight, which opposes the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

29

Dietrich and Whiting

25
0.5

~
24

0.5

------

---+-- Boundary shear me.

----0--

S
E

22

Median bedloed Ii.


Bedload tnnaport

Q,)

0.5

.;j

s::

60

0
1

s::
as

:.a

::s
Q,)

0.5

"'0

fa
a;

"'-

co.

CD

ern
m

0.5

oS
as
~

t=0

as

Q.

Q,)

..d

UJ

s::as
~

as

"'0

as

]
:s0

=a
0

0.5

Q,)

14
0.5

12
0.5

o'--_......

....1001.-........_--'-_-oA.

Joo

200

100

100

"---.I

-200

-300

Distance from centerline (em)


Fig. 21.

Bedload transport, boundary shear stress, and median grain size (DSO) at
Muddy Creek sections in 1978.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

30

LOG
--1.6-1.4i 1.4-1.2
- 6 - 1.2-1.0
- 0 - 1.0-0.8
,
- e - 0.8-0.6
- + - 0.4-0.2 i 0.6-0.4
CONVERSION SCALE:

.3 .4 .5 .6 .7 .8 .9 1.0 1.1 1.2 1.3 1.4 1.5

.'13 ' .I'~ .~5 Ii .~O" ~:O " '2:~ " 5~7
.35

Fig. 22.

.71

1.41

4.0

meters

LOG Ws
O(mm)

Map of Muddy Creek bend showing the trace of the locus of maximum
bedload transport for different size classes, and the region of maximum
boundary shear stress (stippled pattern).

cross--stream near-bed flow, depends on the cube of the gra~n diameter,


whereas the inward drag varies with the square of the graIn diameter.
Consequently, the inward directed drag on the sloping point bar surface forms
an ideal environment for grain sorting, and, in Muddy Creek, fine and coarse
sediment trade sides in a downstream distance of about one channel width.
The net cross-stream transport sediment in Muddy Creek results in a
dynamic balance controlling the equilibrium bed topography.
The channel
curvature and bar-pool topography causes a zone of maximum boundary shear
stress to develop and to shift across the channel through the bend. Along the
inside bank and onto the bar top, the magnitude of the boundary shear stress
declines, whereas it increases downstream through the pool near the outside
bank (Figure 21). At equilibrium, this spatial variation in boundary shear
stress can be balanced either through corresponding changes in bed grain size,
coarsening where boundary shear stress increases, or by cross-stream bedload
transport, Le. convergent transport into areas of increasing boundary shear
stress and divergent transport from decreasing boundary shear stress areas. In
Muddy Creek, this balance is achieved through both cross-stream transport
and grain size adjustments.
Figure 23A and B illustrates the relationships between bed grain size,
near-bed velocity, shear velocity and bedload transport rate predicted from
(15) and the Valin bedload equation.
This shows that, for a constant
near-bed velocity, increasing grain size of the bed causes the shear velocity
and, therefore, the boundary shear stress to increase. But the increasing grain
size also increases the critical boundary shear stress, and this leads to a
maximum transport rate where the difference between the critical and applied
Downstream increase in
boundary shear stress is greatest (Figure 23B).
boundary shear stress can, therefore, be accommodated by bed coarsening
without net convergence of bedload transport. Closer inspection of Figure 21
shows local difference between cross-stream structure of boundary shear stress,

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

31

Dietrich and 'Vhiting

a
Velocity at 2 em
above bed
(em/sec)
~_ _----~ 60.0

CD

en

......

eu

!::
u

~_-------50.0

_ _- - - - - - - - 40.0

L&J

>
a:

L&J

eX, = .077

OL.......&........,j&........,jI--l---lI--l-.L-.L-.L--L---&.--L.--L.--L.--L.---I..---I..--L.--L.--J

0.0

.05

.10

.15

PARTICLE SIZE (em)

b
e

1.5

Z = 2.0

I
U
CD

0(.=.077

en

......

e01.0

a:

Q.
(/)

a:
~

0.5

0
-oJ
0
L&J

CD

0.0
0.0

0.5

.10

.15

.20

PARTICLE SIZE (em)

Fig. 23.

a) Variation in predicted shear velocity with median particle size of the


bedload for a given flow velocity at 2 cm above the bed according to
Equation (16).
Above a particle size of 0.07 cm, the shear velocity for a
given velocity above the bed is only weakly dependent on particle size.
b) Variation in bedload transport with median particle size of the bedload
for given average flow velocities at 2 cm above the bed according to Equation
(16) and the Valin bedload equation.

bedload transport rate, and median surface grain size. Between sections 19
and 20, the maximum transport rate shifts rapidly toward the centerline, not
because the boundary shear stress maximum is there, but because the
boundary shear stress increased rapidly there without bed coarsening causing

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

32

the excess shear stress and bedload transport to be greatly increased [see
Figure 5-28, Dietrich, 1982]. Between stations 19 and 22, the bed samples in
the deepest water were distinctly bimodal [Figures 5-18, Dietrich, 19821, unlike
any other samples in the bend. This bimodality apparently resulted trom the
arrival of grains, coarser than 2 mm, that had rolled down the point bar slope
into the pool. Comparison of sections 19 through 24 shows that the pool
coarsens due to influx of large grains from the point bar slope before the
boundary shear stress maximum shifts completely into the pool, and that the
zone of maximum bedload transport does not shift completely across the
channel into the pool. Although the bedload transport maximum entered the
central bend near the inside bank, it did not enter the next bend downstream
against the inner bank.
These observations suggest that bed inclination
toward the outside bank and consequent net cross-stream transport of coarse
particles is primarily responsible for the coarsening of the bedload in the pool.
The relatively close tracking of the coarsest bed particles by the zone of
maximum boundary shear stress is perhaps more coincidental than causal.
In the upstream part of the bend and at slightly higher stages, a different
effect, but one that probablf is common in other rivers, becomes important.
preViOUSlY~ Dietrich et ale l1979, Figure 61 and Dietrich and Smith 11984,
Figure 10 reported finding immobile grave in the pool.
This grave was
derived rom erosion of the adjacent terrace bank.
Comparison of
cross-sectional profiles at stations 18, 19, and 20 show that the pool was
about 10 em deeper near the outer bank in 1976 when higher flows had
occurred and, correspondingly, the area of immobile gravel covered a much
larger area in 1976. The gravel was a lag due to the divergence of sediment
transport caused by strong inward flow resulting from curvature-induced and
point-bar-eonstrained secondary circulation. We hypothesize that, if the gravel
had been absent, then the pool would have continued to deepen, steepening
the point bar slope until grains could not travel out of the pool by the inward
near bed flow component, or reducing the boundary shear stress to the critical
value of the fine sand reaching the upstream part of the pool.
The cross-stream variation in median grain size of the suspended load and
bedload were nearly the same (Figure 24), with the suspended load
systematically finer than the bedload. In areas of boils, the suspended load
size temporarily approached that of the bedload. Boils regularly ejected from
the lees of the three-dimensional dunes.
For example, the average period
between boils of 124 observations in the upstream part of the bend was 3
seconds with a standard deviation of 0.9 seconds. The boils made sampling
the time-averaged suspended sediment difficult, but despite the crude sampling
method and this inherent variance, the concentration varied systematically
across the channel (Figure 25). The average suspended load was 14 gm/sec,
with a coefficient of variation between the eight sections of measurement of
48% or nearly twice the variation in the measured bedload rate. At sections
12, 14, 18, 19, and 22, there appears to be a concentration minimum near the
centerline. Observations at the time of sampling suggest that this variation
may be due to a lack of influence of boils in the centerline area. Maximum
concentrations at each section are generally located outside the maximum
boundary shear stress zone, but where bed grain size is considerably smaller.
The primary exception is at section 20, where maximum transport rate of each
bedload size, the maximum suspended load concentration and transport rate,
and the zone of maximum boundary shear stress coincide.
The displaced
concentration maximum, coupled with rapid outward shift of the position of
maximum unit water discharge, causes the maximum suspended load to shift
guickly to the side of the bar or pool well before the bedload maximum
(Figure 26).

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Dietrich and Whiting

33

1.:)~-""""""-T'"'""""""'P'"_~...............----r-...............

1.0

'I---tr--1___

0.5

bedload
suspended load

20

--------

l=:

0.0 L--..o....----L-------...J~..........L.._.-.

j:

:~9J

18

! :2
::1 '

l=:

~ J~~
4------.----.---~:1
14

;~[2J
200

100

-100

200

'300

Distance from centerline (cm)

200

19

160
120

Q)

~
ot,)
~

l=:

0.0

300

.9
~
as$.l

80
40

t ~I=:~===181
! =[
o

14

'/\::

~t---------:~---:
:--:- :1
12

300

200

100

100

-200

300

Distance from centerline (cm)

Fig. 24.

Variation in median particle size of bedload and suspended load through the
successive sections through the Muddy Creek bend.

Fig. 25.

Suspended sediment concentration at Muddy Creek sections.


ordinate scale for concentration.

Note the varying

Sediment Transport in Gravel-Bedded Meander


In order to compare boundary shear stress and sediment transport
relationships between sand- and gravel-bedded meanders, data were obtained
from a tributary of the Rio Grande del Ranchos River near Talpo, New
Mexico, upstream of a reach investigated by Leopold and Emmett [19831. The

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

34

20

200

100

100

-200

300

Distance from centerline (cm)


Fig. 26.

Comparison of bedload and suspended load transport fields.


Each value was
divided by the section-averaged value to yield the normalized transport.

field work was accomplished during spring snowmelt when a small portion of
the flow crossed the partially vegetated flood plain on the inside of the bend
(Figure 27). Mean width and depth were 4.0 and 0.4 m, respectively, and
discharge was about 2 m 3 /sec. The bend was preceded by one less strongly

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Dietrich and Whiting

35

II

Fig. 27.

Contour map of depth and section locations at Rio Grande de los Ranchos.
Flow depths greater than 60 cm are shaded.

curved and discharged into a bend partially blocked by a beaver dam.


Average water surface slope through the bend was about 0.0066. The study
reach was strongly curved with a radius of curvature to width ratio of almost
1.8. Pebble counts of the bed surface between sections I and II yielded a
median grain size of 32 mm and DS 4 of about 60 mm. During this high
runoff event, the flow was relatively clear, although the bed surface could not
be seen in the deeper water over the pool.
Our goal was to investigate whether the zone of the maximum bedload
transport tracked the shifting zone of maximum boundary shear stress into the
pool through the bend as occurs in sand-bedded channels. For this purpose,
and in order to make measurements at the same stage, only three
measurement sections were used (Figure 27). Flow velocity at 5 cm above the
bed was measured with a Smith pulse-type current meter [Dietrich and Smith,
1983] mounted to a rod with a vane so that it could freely rotate into the
flow direction. Velocity was recorded for 20 second periods, with typically
three but up to 14 separate 20 second readings being used at each position
across the channel. Simultaneous with the velocity measurements, bedload was
collected with the Helley-Smith sampler [Emmett, 1980] for 5 minutes and
then an attempt was made to characterize the bed surface size by scraping the
sample location. This procedure unfortunately mixes surface grains with an
unknown
amount
of
subsurface
sediment,
probably
systematically
underestimating the size of the surface grains.
Each sample was dried,
weighed, and sieved to determine transport rate and size distribution.
Analysis of the bedload and bed surface samples required consideration of
what the Helley-Smith sampler actually collects and whether the sampling
period was sufficient. Because of the height of the bedload sampler, not all of
the sediment caught moved as bedload. If we use the standard suspension
criteria that grains with settling velocity smaller than the shear velocity travel

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

36

in suspension and use curves from Dietrich [1982] to estimate settling velocity
from sieve analysis, then about 60 percent of the sediment was bedload. At
the two downstream sections, the bedload to total load ratio increased
systematically across the channel approaching unity on top of the bar. In
order to calculate grain sizes of the bed surface and bedload, all sediment
expected to travel in suspension based on calculated local boundary shear
stresses was subtracted from the total sample mass.
Sampling period in the study was chosen to complete the three sections
within the nearly constant flow period during the day. Analysis of grain size
distributions, however, suggests that a procedure, biased by expected grain size
distribution in the load, should be used to insure that the rarer, coarsest
fraction is appropriately sampled. Ideally, the sampling period during which a
Helley-Smith sampler rests on the bed should be long enough such that
individual coarse grains should contribute a small percentage of the total
weight. If M equals the mass of a grain with nominal diameter, Do, and p
equals the percent of the total load that is in a size class represented by Do,
the minimum sampling period, T, such that the mass of a single large grain is
equal to its proportion in the expected size distribution, is given by

(19)
where qb is the measured bedload transport rate (mass per time per width)
and w is the width of the sampler. Solving for the minimum sampling period,
T gives

T=

P qb w

(20)

and, using the nominal diameter, Do to calculate mass, (20) becomes


1r

Ps 0 Do
P qb w

(21)

For example, if the bedload size distribution was the same as the bed surface
size distribution, as represented by our pebble count in the study reach, then
22 percent of the bedload should have a size between 45 and 64 mm. A
representative grain size of this size class is 5.4 cm and, assuming Ps = 2.0
gm/cm 3 , the minimum sampling period for transport collected in a 7.6 cm
wide bedload sampler is

T = 98
qb
For the low (.02 gm/sec--em), average (.184 gm/sec--em), and high (1.45
gm/sec--em) measured bedload transport rate in our study reach at individual
points, T is 82, 9, and 1 minutes, respectively. To state this another way, in
order for only one grain with a diameter of 5.4 cm to be collected and not be
more than 22 percent of the total weight of the bedload sample, the sampling
period calculated from (21) must be used. Clearly T is a minimum. We will
return to this issue later.
Boundary shear stress was calculated using the single-velocity method
proposed above, with the observed velocities and the estimated DS 4 (Figure 28)
from surface samples collected by scraping the bed.
Figure 29 shows the
cross-stream structure of the estimated boundary shear stress at the three
sections. The cross-sectional average boundary shear stress calculated by this

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

37

Dietrich and Whiting

III
-G

-.-

o ---.----0----0--

II

l:l..
Q)

Cl 100
-6

250

-250

Distance from centerline (cm)


Fig. 28.

Grain size variation of the bed surface and the bedload at Rio Grande de los
Ranchos and the cross-stream depth profile.

method at I, II, and III yielded 195, 176, and 190 dynes/cm 2 , respectively.
The average of 187 dynes/cm2 compared to the reach-averaged boundary shear
stress calculated by method (2) of 246 dynes/cm2 yields a form drag ratio of
1.3. The Nelson form drag corrections for point bars [Nelson, 1988] yields, for
a bar height of 48 em, Zo of 0.2 and wavelength of 4500 em, a predicted total
boundary shear stress to local ratio of 1.2, in close agreement to that
estimated from the field data. The consistency of the data suggest that the
estimated local boundary shear stresses are reasonable.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

38

Although these data are sparse, they reveal some well-defined patterns that
raise important questions about sediment transport in gravel-bedded meanders.
Figure 29 strongly suggests that most of the sediment in transport near the
bed occurs in a narrow band along the centerline through the bend. The
sediment load was considerably finer than the bed surface in the pool and at
the entrance to the bend, but over the shallow bar top where the surface was
sandy, the load and surface sediment size were comparable (Figure 28). With
the exception of two of the 22 samples, maximum bedload grains size collected
was less than 22 mm. All grains collected were finer than 45 rom. Because
the bedload transport rate was small, it is possible that the five minute
sampling period was insufficient to collect the less frequently moved large
rocks, as discussed above. Two observations, however, suggest that the coarser
size on the bed should have at least been detected if they were moving. A
total of 65 minutes of sample collection time was used to obtain samples from
the 13 sites where the bed surface DS 4 was greater than 22 mm. If, in fact,
the bedload should have been 22 percent by weight sediment between 45 and
64 mm, as suggested above, then, according to (21), within 16 minutes for the
observed transport rate of 0.1 gm/sec~m, 5.4 em grain could have been
caught.
This is, in a sense, a crude statement of probability of catching
grains coarser than 45 mm if they were moving. None were found in the
sampler. Also, in the narrow band of high transport along the centerline, our
sampling period relative to grain sizes was much more favorable, yet no
large grains were collected there either.
We conclude that, at this stage,
grains coarser than 45 em were not moving, or moving very rarely, and that,
except over the point bar top, the bedload was finer than the bed surface.
As expected, the position of maximum boundary shear stress shifted from
the center of the channel to the pool through the bend (Figure 29). Although
the coarsest bedload and bed surface grains coincide with the maximum
boundary shear stress position across the channel, the maximum bedload
transport does not track the outward shifting maximum boundary shear stress
zone. Instead, it is apparent that the bed surface fines over the shallower
portion of the bar, reducing the critical boundary shear stress there, and the
maximum transport occurs where the boundary shear stress relative to the
reduced critical boundary shear stress is greatest. Many bedload equations
state that the transport rate increases with the difference between the
dimensionless shear stress, Tb((Ps - p)DSO)-I, and the critical dimensionless
shear Tc((Ps - p)gDso)- 1. This difference at our site was smallest in the pool
where the transport was least and was over an order of magnitude higher at
the centerline maximum bedload transport band.
A second maximum
difference occurred near the inside bank where the transport was low, but here
the actual coarseness of the surface layer was probably underestimated because
of the large amount of sand collected from the subsurface in the scoop
sampler.
Efforts to use excess shear stress bedload equations proved
unsuccessful, as the bedload transport rate was greatly over-predicted in the
pool.
The calculated boundary shear stress, as argued above, is probably
approximately correct, but our method of grain size characterization was crude.
Perhaps more important is the theoretical problem of how to calculate
transport rate when the grain size of the moving load is smaller than that
resting on the bed. This is an area of active debate [e.g., Thorne et al., 19871
and, at present, we cannot predict from estimated boundary shear stress and
grain size the observed bedload transport field.
One other result of our field study that is well expressed with these limited
data is a large spatial variation in the size of material that should be carried
in suspension, and this has important effects on the sorting of sediment. Over
the downstream part of the pool at sections II and III, the estimated
boundary shear stress would cause all sand to be carried in suspension. llere

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Dietrich and Whiting

39
1.0

600

ill

soo
400

0.5

300
200
100

0.0

-e

600

rJJ

II

500
400

.,;l
~

300

~
rJJ

~Q,)
t::

~
rJJ

rn

Q,)

~
.,;l

rn
~

t::

as
~

200

"'0

100

as
Q,)

rn

.,;l

as0

::aC>

r:Q

~
as

"'0

t::

::s

600

0.5

I
0.4

500
400

0.3
300
0.2
200
0.1

100

100

-100

0
-200

Distance from centerline (em)


Fig. 29.

Total boundary shear stress and measured bedload transport at Rio Grande de
Note the variation in the ordinate scale for transport at
los Ranchos.
sections.

the flow is also strongly toward the inside bank; based on vane orientation
measurements made with an indicator on the holding rod, flow direction at 5
cm above the bed was typically between 30 and 40 degrees toward the inside
bank relative to an orientation perpendicular to the cross-sectional direction.
At section I, the entrance to the bend, all sand across the active bed surface
in the middle of the section should also travel in suspension, and here the
flow direction near the bed was about 22 degrees toward the inside bank.
The combined effects of downstream decreasing boundary shear stress along the
inside bank over the bar, increasing boundary shear stress into the pool, and
flow toward the inside bank will cause sand to be transported inward and
become bedload and bed surface material over the bar top where the boundary
shear stresses are less.
The convergent transport and increasing bedload
transport of sand over the bar top appears to explain the abrupt increase in
sand in the bed material, as measured with the scoop samples.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

40
Di8CU8Sion

Accurate prediction of observed boundary shear stress fields in natural


rivers is a crucial and sensitive test of theories concerned with channel form
and dynamics. But, to provide such a test, reliable, internally consistent field
measurements must be made. This leads to the dilemma that virtually all
methods of estimating boundary shear stress that can be successfully employed
in the field rely on the use of theory.
In applying these methods, it is
essential to examine whether the assumptions used to arrive at the method are
satisfied in the field setting. The most obvious example of this is equating S
with 81 in (8) to estimate local boundary shear stress. Clearly, if substantial
spatial variation in bed topography occurs, this approximation is very crude.
Commonly, 8 is set equal to 81 + 82 and is then called the "energy slope" or
"energy grade line" [e.g. Petit, 1987].
This approximation should be
reasonably accurate for prediction of mean longitudinal profile of the water
surface in flood routing, but exclusion of 83 degrades the accuracy of predicted
cross-stream variation in total boundary shear stress over bar and pool
topography. The analysis of the nearly straight channel flow over an alternate
bar on 80lfatara Creek suggests that 83A, which, because of the insignificant
channel curvature, is not balanced by S3B' plays an essential role in flow and
bed morphology adjustments in alternate bar and possibly braid-bar channels.
Our empirical analysis of boundary shear stress suggests that, despite the
large potential error in calculations, consistent and apparently accurate data
can be obtained. Although Dietrich and Smith [1983] correctly point out the
difficulty of using (1) to calculate boundary shear stress, analysis of all the
terms in (1) still provides insight about the relative magnitude of terms.
Results shown in Figure 9 indicate that quantitative analysis is, in fact,
possible. The estimate of local boundary shear stress responsible for sediment
transport seems less error-prone and can be highly constrained if other
measurements, particularly of the sediment transport field, are made. 'fhe
proposed near-bed velocity method gave results that seem quite accurate. The
use of this method in the Rio Grande del Ranchos tributary made it possible
to define the spatial variation in boundary shear stress, where, because of
limited time and the structure of the interior flow field (which caused the
velocity maximum to be near the bed over the pool) conventional velocity
profile measurements were unproductive.
Application of this method to
gravel-bedded rivers is not without its problems. Characterization of the bed
surface is required and, as extensively discussed elsewhere [Church et al., 1987;
Diplas and Sutherland, 1988], scoop samples may not yield accurate estimates
of the surface roughness.
Comparison of boundary shear stress and sediment transport relationships
observed in Muddy Creek and Rio Grande del Ranchos point to important
differences between sand and gravel-bedded river meanders. In either case, the
cross-stream shift of the zone of maximum boundary shear stress into the pool
through a bend can either be accommodated by net scour in the pool,
coarsening of the bed surface in the pool, or convergent sediment transport to
the pool.
With stage change in sand-bedded rivers net scour in pools is
common, but at equilibrium, one or both of the latter responses must occur.
In the equilibrium case at the sandy-bedded Muddy Creek, both coarsening of
the pool bed and net bedload transport across the channel into the pool
achieves the balance with the downstream increasing boundary shear stress.
The coarsening in the pool results from rolling and avalanching of coarse sand
and granules from the point bar top after these particles have been carried to
the outer edge of the top by shoaling-induced outward flow. The cross-stream
bed slope must effect the rate of cross-stream transport of sediment such that,

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

DieLrich and Whiting

Vol. 12

41

for a given opposing inward near-bed flow, the steeper the slope (and,
consequently, the deeper the pool), the larger the cross-stream transport.
Moreover, the steeper the slope, the finer the sediment size that could shif~
outward against the opposing inward near-bed flow.
Normally, this
cross-stream slope is assumed to be that required to cause a balance between
inward drag and outward gravitational pull on the particles [see review,
Dietrich, 1987; Hills, 1987] such that there is no net cross-stream transport of
sediment. When there is net cross-stream transport of sediment in the bends,
this requirement only applies once the boundary shear stress maximum is in
the pool and grain sorting is completed.
The small imbalance causing net outward sediment transport may only
require small steepening of slope relative to that with no net transport. For
example, Ikeda '8 [1987] Equation 15 for cross-stream sediment transport
suggests that, if the deviation angle of the near-bed flow from the downstream
direction and the ratio of critical shear stress to boundary shear stress are held
constant, a cross-stream point bar slope of only about two degrees steeper
than the slope for no net cross-stream transport will cause the cross-stream
transport into the pool to be 10 percent of the downstream transport. When
this small imbalance acts over a large portion of the bend, however, it shifts
the zone of maximum bedload transport into the pool. Dietrich and Smith
[1984, Figure 121 found the cross-stream bedload transport to be on average 10
percent of the downstream in Muddy Creek. Because the cross-stream slope
of the point bar surface in Muddy Creek is about 15 degrees, if Ikeda's
equation is applicable, an approximation that assumes no net cross-stream
transport will only err in estimating the mean slope by about 2/15, or 10
percent. Not surprisingly then, predictions of the Muddy Creek cross-stream
point bar slope that assume equilibrium of cross-stream forces on the resting
bed surface particles [Odgaard, 1986] or moving ones [Bridge, 1984] are found
to be approximately correct. It should be pointed out, however, that the
assumpt10n used in Ikeda '5 [1987] model and in other sorting models is that
the grains are always in contact with the bed. Once the grains leave the bed
during saltation, however, the outward gravitational pull is zero and the grains
should travel in the direction of near-bed flow [see Dietrich and Smith, 1984,
p. 1375-1376].
The quantitative significance of this effect on sorting of
sediment in sand-bedded rivers is not known. No model as yet has predicted
the complete spatial variation in grain size distribution through a bend.
Because of the apparent relatively small contribution of net cross-stream
transport to the equilibrium cross-stream bed slope in sand-bedded rivers, it
could be argued that it is insignificant and can be ignored. Exploration of
three important features of bends, however, appear to require net cross-stream
transport. First, Struiksma et al. [1985] argued that the "overshoot" tendency
in bends in which the maximum pool depth tends to be at the upstream part
of a bend of constant radius is due to an oversteepened cross-stream bar slope
that causes net sediment transport into the pool. This induces a periodic
response in pool depth through the bend.
Second, the point bar top or
platform [Bluck, 1971] is often nearly flat [Dietrich, 1987, p. 1821, and when
bends have a high width to depth ratio, there is a tendency toward flow depth
minimum to develop toward the center of the channel, rather than near the
inside bank.
Both of these features are probably controlled by the
shoaling-induced outward flow over the bar, which can move sediment across a
surface not steep enough to cause cross-stream rolling and which causes
convergence of sediment transport with that coming up from the pool along
the edge of the bar top. This latter effect and its contribution to flow depth
minimum away from the inner bank has been explored by Nelson and Smith
[1985]. Finally, the topographic changes in sand-bedded rivers associated with
stage change appear to be due to divergence in the boundary shear stress field

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

42

not being balanced by net cross-stream transport of sediment. In sand-bedded


rivers in which all bed material sizes remain mobile over a wide range of
stage, the pool scours and the point bar and crossing aggrades during stage
rise. The stage changes in bed topography have been attributed to slope and
consequently boundary shear stress changes in pools versus the crossings or
riffles [Le., Leopold and Wolman, 1960]. Shoaling-induced outward transport
over the bar top also contributes to topographic adjustments. During stage
rise, the outward transport should decrease, causing net deposition and growth
of the bar into the flow. This net deposition prevents cross-stream transport
to the pool so that scour there is more likely.
The crossings also may
aggrade during stage rise because net transport across the channel there, due
to shoaling, is also reduced. Equilibrium at the increased stage height can
occur if aggradation of the crossing and bar top causes sufficient shoaling to
force outward transport toward the pool, and if scour has steepened the
cross-stream bar slope to cause sufficient net transport to the pool.
The
relative role of changes in the divergence of boundary shear stress field and
net cross-stream sediment in causing bed topography changes with stage has
not, to our knowledge, been explored theoretically.
Gravel-bedded bends differ from sand-bedded ones in several ways. The
most obvious is that stage decline below over-bank flow often causes relatively
minor topographic adjustments in gravel rivers. The large reduction in water
surface slope and boundary shear stress over the pool at low stage tends to
cause the pool area to collect some fine bedload or coarse suspended load
carried in from the crossing and edge of the bar where the surface slope tends
to increase with stage decline. Although Keller [1971] is often cited for his
"velocity reversal" hypothesis which argues near-bed velocity should become
greater in the pool than in the riffle at high flow, the idea was clearly stated
in terms of boundary shear stress changes by Leopold and Wolman [1960, p.
777] many years before him. Further, Keller presented no data that In gravel
bed rivers the riffles are, in fact, coarser than pools at high flow, which is
what he was trying to explain. At the high flow observed in Rio Grande del
Ranchos, for example, the average local boundary shear stress and the grain
size distribution at the riffle and in the pools were nearly the same.
Observation in gravel-bedded meanders in New Zealand indicate that the pool
gravel is coarsest [Carson and Griffith, 1987].
This controversy, nonetheless, points to the general problem that, unlike in
sand-bedded meanders, it is not obvious from examining the bed surface of
bends in gravel what is the bedload size distribution.
From a theoretical
perspective, one would like to know what the size distribution of gravel is that
enters a bend from upstream and from the banks over a period of sufficient
time that all grains on the surface have been replaced. This is needed to
predict bed morphology, bedload transport fields, and grain size variation in a
given bend.
In straight laboratory flumes without well-developed alternate
bars, Parker et ale [1982] have argued that the subsurface sediment grain size
distribution closely approximates that of the bedload. The Rio Grande del
Ranchos data, although crude, suggest the cross-stream variation in bedload
and bed surface size is very large, as it,,- is in the well-documented
sand-bedded Muddy Creek. Hence, although a coarse surface layer commonly
mantles the bed in gravel bends, the subsurface may, at best, approximate
only the size distribution of the local load that passes over a particular point
on the bed. But even with data in hand about local subsurface grain size
distribution through a bend, it is not obvious how these data would be
combined to define the average size distribution entering the bend.
In
practice, unless a straight reach without bars can be found such that lateral
sorting and lateral variations in boundary shear stress and the bedload
transport rate are minimized, the bedload size distribution may _only be

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Dietrich and vVhiting

Vol. 12

43

determined through extensive direct measurement, and from long-term


deposition of sediment in reservoirs.
The sorting in bends has the effect of laterally distributing the
heterogeneous mixture of grain sizes that travel as bedload. Because many
gravel-bedded rivers experience only low excess shear stress even at bankfull
flow, cross-stream variation in grain size may have a larger control on bedload
transport and equilibrium bed topography development than in sand-bedded
bends. One effect is that the zone of significant bedload transport may be
confined to a narrow band along the channel centerline, as in Rio Grande del
Ranchos. Unpublished results [E. Larson, personal communication] on bedload
transport in the 22 m wide Pole creek in Wyoming for a discharge slightly
above bankfull discharge show a similar pattern of bedload transport near the
centerline through the bend.
Another effect is that the bed surface may
become covered with gravel whose size is only a small fraction of the total
bedload. The lack of grains larger than 50 mm in the bedload at Rio Grande
del Ranchos suggests that, although these grains are common in the riffle and
pools, it is probably a very small fraction of the total bedload that passes the
bend over many years. Comparison of bed surface grain size at section I and
II indicate that the coarse sediment toward the inside bank of the centerline
at I must travel outward and into the pool by II or that the transport rate is
so low that when the grains go into the bar top they are mixed with a very
large quantity of much finer sediment. The first interpretation seems more
probable.
The relationship between surface grain size variation in a bend and bedload
transport rate probably depends not just on average size distribution of
bedload that enters the bend, and the geometry of the bend, but also on the
total amount of bedload that is imposed from upstream and bank erosion.
Experiments by Dietrich et ale [1987] in a straight flume in which bedload feed
rate was varied while water discharge, grain size distribution of the load, and
mean boundary shear stress stayed nearly constant showed that, with
decreasing feed rate, the bed surface coarsened and the area of active transport
became a narrow band, bounded by nearly static bed surfaces. Based on these
experiences, we can hypothesize that the widths of significant bedload transport
and the grain sizes found on the bed surface in a gravel-bedded meander are
controlled in part by the amount of sediment entering the bend.
Hence,
unlike what is normally assumed for sand-bedded bends, the effects of
upstream sediment supply and sediment eroded from banks appear to be
important to predicting spatial variation in sediment size and transport rates
in bends.
Unlike in sand-bedded rivers, flood events in gravel-bedded rivers may
cause bed surface grain size and morphologic adjustments that may persist for
years after the flood. It is possible that areas of coarser grains may only
move in rare events. The controversy remains about whether high discharge
events disrupt significantly the coarse surface layer in gravel rivers [see papers
and discussions in Thorne et al., 1987].
One effect that we inferred from our Rio Grande del Ranchos data, which
may be common in other gravel-bedded rivers, is that much of the sand was
probably thrown into suspension at the upstream part of the bend where the
maximum boundary shear stress is near the centerline. Rapid decline of the
boundary shear stress in the downstream direction onto the bar caused the
sand to come out of suspension and travel as bedload. This contributes to the
very rapid decline in bed surface size in the bend and the development of a
fine grained "bartail" like that described by Bluck [1971]. At low discharge,
the sand will also travel through the entire bend as bedload and infiltrate the
pores of the coarse gravel in the pools. The variation in transport mode is
most pronounced in the coarse sand range because boundary shear stresses

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

44

Vol. 12

Boundary Shear Stress

required for suspension increase rapidly in sediment coarser than sand.


In
effect, the strong cross-stream variation in local boundary shear stress caused
by channel curvature and bar-pool topography causes portions of the bed
surface and perhaps significant portions of the bedload to be sand in a stream
that at high flow would otherwise carry this sediment in suspension. A model
for sorting of sediment in gravel-bedded meanders will need to consider the
spatial variation in sand suspension through a bend and variation with stage.
Conclusion
Methods described here for analyzing flow data, calculating local boundary
shear stress, and measuring bedload transport indicate that accurate .data,
sufficient to test assumptions in theories, can be obtained from detailed field
measurements. There is no absolute methodology for analyzing flow field data
for convection acceleration forces because of the inevitable effects of width and
discharge changes and limited number of cross-sections at which data are
usually available. It is fair to say that you cannot have too many crosssections in a reach of channel.
The method we have proposed for mapping boundary shear stress in rivers
based on near-bed velocity and grain size measurements appears to have much
promise in sand-bedded rivers and possibly in gravel-bedded ones as well. For
Muddy Creek the method gave accurate results which were highly constrained
by several other estimates of boundary shear stress and from sediment
transport measurement.
Its use in the gravel-bedded case was much less
constrained, but the magnitude and cross stream of the calculated boundary
shear stress seemed consistent with other observation. One encouraging result
at our Muddy Creek site was the strong correspondence in cross-stream
structure between boundary shear stress calculated from the vertically-averaged
velocity and a constant drag coefficient and that calculated from the near-bed
velocity data.
The difference in magnitude of these two stresses can be
successfully predicted from the form-drag correction equation proposed by

Smith and McLean [1978].

The method for calculating the sampling period to assess accurately the
grain size distribution of bedload is particularly important in studies of gravel
transport processes.
Normally, sampling periods are less than a minute or
rarely exceeding two to three minutes. Sampling periods have informally been
biased by' the amount of sediment collected, but that is not the only
determinant. The most challenging problem is the study of very low transport
rates, particularly those associated with questions of initial motion.
For
example, the study by Andrews and Erdman f1984J used a sample period of
about 4 minutes. According to our analysis ot theIr data, in order to collect
on average one 150 mm grain, typical of the coarser fraction found on the
surface, they should have sampled for about 80 minutes. Even if all 12 of the
4 minute samples they collected are combined, as they in fact did, their
sampling period is still too short by 30 minutes, at best. Regrettably, very
long sampling times at a point on the stream bed appear to be necessary in
many gravel-bedded rivers if accurate estimates of the bedload size distribution
are to be obtained.
The complex interactions that give rise to characteristic bed topography
and sediment sorting in rivers can be broken down into three components.
Channel curvature and spatial variation in bed topography arise from, and in
turn reinforce, spatial variation in the boundary shear stress field.
Cross-stream bed-slopes cause cross-stream movement of grains, and if
near-bed flow has a component up the slope, large and small grains may
become separated with the largest at the base of the slope and the smallest at

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Dietrich and \Vhiting

Vol. 12

45

the top.
Spatial adjustments in surface grain sizes control the local
relationship between boundary shear stress and sediment transport rate. Our
field studies shed some light on how each of these components can be
quantified and modeled. To our knowledge, however, no theory has included
all three of these effects to predict the bedload transport fields of individual
size classes through a bend. We have shown that accurate modelling of the
total boundary shear stress field in both a sand and gravel-bedded meander
with well-developed bar topography requires including all similarly-scaled
convective acceleration forces. In particular, forces arising from cross-stream
momentum transfer associated with a large vertically-averaged cross-stream
velocity are not negligible in our field sites. It may be that this term is only
negligible if flow entering a bend is skewed such that the high velocity core is
already near the outer bank, causing the net cross-stream discharge to be
small. This may have been the case in the experiments reported by Odgaard
and Bergs [1988]. Despite the potentially large errors in our analysis, both the
cross-stream structure and magnitude of the local downstream water slope
were reasonably well predicted. Comparison of the data for gravel-bedded and
sand-bedded cases, in which the total boundary shear stress was nearly the
same, showed that the magnitude of convective acceleration terms are similar.
Prediction from available equations which include all appropriate terms of the
mean velocity field for flow through sand or gravel bedded meanders should be
quite successful.
The cross-stream bed slope caused the largest particles in the bedload of
Muddy Creek to cross rapidly into the pool against the inward secondary
circulation which brought fine particles to the point bar top. These coarse
particles were first transported across a relatively flat point bar by the
shoaling-induced outward flow there. Coarsening of the pool bed also occurred
where sediment transport away from the outer bank caused scour to a gravel
lag before the cross-stream slope could sufficiently steepen.
In our
gravel-bedded site, the bedload also coarsened in the pool through the bend,
but the bed surface did not coarsen significantly and instead the bar top fined
considerably. This fining was associated with convergent transport and large
downstream reduction in the boundary shear stress which allowed coarse sand
that would otherwise travel in suspension to move as bedload. The lack of
coarsening of the bed surface through the pool may be due to the low excess
boundary shear stress generally found in gravel-bedded rivers which would not
allow a large cross-stream variation in boundary shear stress above critical
values.
In this case, the bed surface cannot coarsen significantly without
becoming immobile. This suggests the tipping model of Parker and Andrews
[1985] for predicting the zone of coarsest sediment through a meander may be
most applicable to channels with relatively larger excess boundary shear
stresses.
Comparison of the bedload transport and boundary shear stress fields in
Muddy Creek, Rio Grande del Ranchos, the laboratory flume of Hooke [1975],
in the south Esk River [Bridge and Jarvis, 1982] suggests a range of
relationships depending on grain size and heterogeneity. As Bridge and Jarvis
[1982] correctly point out, -the experiments by Hooke were for nearly uniform
size sediment and, in this case, boundary shear stress and bedload transport
fields must have very similar structure, i.e. the zone of maximum bedload
transport should and does closely track the zone of maximum boundary shear
stress as it shifts across the channel through the bend.
Shoaling-induced
outward flow and oversteepening of the cross-stream slope induces the required
net cross-stream transport. Hooke's results may be most applicable to large,
sand-bedded rivers with relatively fine beds, such as the Mississippi River,
where excess boundary shear stress is relatively high everywhere and grain size
changes may have small influence on bedload transport rates. The Muddy

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

46

Creek results show that, even in strongly heterogeneous bedload material, net
cross-stream transport occurs, although the zones of maximum bedload
transport and boundary shear stress do not always coincide. Here, the effects
of cross-stream bed tipping on bed grain size variation is large, but do not
cause such large critical shear stress variations that the bedload transport
maximum is prevented from shifting into the pool area. Although there are
some weaknesses with the Bridge and Jams [1982] data for the south Esk
river [see comments in Dietrich and Smith, 1983, 1984; and Dietrich, 1987], it
may be that tipping-induced cross-stream transport of large grains or scouring
into a lag gravel caused such large increases of critical boundary shear stress
in their coarse sand and fine gravel-bedded river that the bedload transport
maximum stayed close to the centerline as they propose.
The results at the Rio Grande del Ranchos bend may be typical of many
gravel-bedded rivers. Most of the bedload travels in a narrow bend near the
centerline and is considerably finer than the average bed surface in the bend.
The bedload transport field is greatly influenced by the surface grain size
across the bend. Although shoaling and cross-stream bed slope of the point
bar may cause significant cross-stream transport, the pool cannot greatly
coarsen without raising the critical boundary shear stress of the bed ~urface
above the local boundary shear stress.
Cross-stream sorting is perhaps
strongest where significant amounts of sand are present, such that changes in
mode of transport contributes to construction of the characteristic fine bar tail.
Unlike in sandy rivers, the long-term average bedload grain size distribution
that a gravel-bedded bend sorts is not obvious by inspection of the bed
surface or of the subsurface.
Acknowledgements

J. Dungan Smith, Tom Dunne, Wray Smith, Pat Irle, Leslie Reid, Rich
Spicer, Mary Power and Steve McLean assisted in the field work at Muddy
Creek. William Emmett, Robert Myrick, Peter Goodwin and Mary Power
assisted in the field work in New Mexico and George Ehlers assisted Whiting
at Solfatara Creek.
Useful discussions were held with Jon Nelson and J.
Dungan Smith. David Montgomery reviewed an earlier draft, Jacob Odgaard
provided many useful comments, and Lindy Foster processed the words.
Financial support for the field studies was provided by the Geological Society
of America, the Corporation Fund of the Department of Geological Sciences of
the University of Washington, National Science Foundation grants
ENG78-16977 and CEE-8307142 and American Chemical Society Grant
ACS-PRF-18427-AC2. The analysis of our field study was part of the Joint
United States-Japan Research on River Meandering sponsored by the Japan
Society of Promotion of Society and the National Science Foundation.
Notation

Cn
D,D n
Dx,Dso,Ds 4

an empirical constant relating the roughness parameter,


zo, to a representative grain size of the bed surface
drag coefficient
grain diameter; nominal grain diameter
grain diameter for which x, 50, or 84 percent of the
sample is fi nes

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

47

Dietrich and Whiting


E
K

g
h
M
n

N
p

Vol. 12

water surface elevation


empirical constant relating median grain size of bed
surface to critical boundary shear stress
gravitat i onal acceleration
depth of flow
mass of a grain
cross-stream coordinate, following the channel centerline
and positive toward the left bank

n/R
the percent of the total bedload that is in a given size
class
cross-stream and downstream discharge, respectively
radius of curvature of the channel centerline

Reynolds roughness number, u*~*


downstream coordinate, parallel to centerline
downstream water surface slope
components in force balance controlling downstream water
surface slope

Sn
Snl,Sn2,Sn3,Sn4
T

cross-st ream water surface slope


components in force balance controlling cross-stream
water surface slope
minimum bedload sampling period required to catch grain
size of interest
downstream and cross-stream components of fluid velocity,
<> implies vertically averaged

U',W '

w
ws
Z

ZQ
Zn

fluctuating component of downstream and vertical fluid


velocity
shear velocity defined as the square root of the boundary
shear stress divided by fluid density
width of channel, width of bedload sample
settling velocity of grains in still water
near vertical coordinate, perpendicular to the bed
roughness parameter including effect of saitating grains
roughness parameter for a static bed
an empirical constant relating the hopping height of
bedload particles to roughness parameter, zo, in the law
of the wall equation

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

48

DB
K,

v
P,Ps
( Tzs)b
( Tzn)b
Tb
Tc

</J

saltation height of bedload grains


von Karman constant
kinematic viscosity
density of fluid and sediment
downstream component of the total boundary shear stress
cross-stream component of the total boundary shear stress
boundary shear stress
critical boundary shear stress
friction angle of a resting grain; sediment size scale; Le.
grain diameter equals 2-</J
ReferenOO8

Andrews, E. D., and D. C. Erman, Persistence in the size distribution of


surficial bed material during an extreme snowmelt flood, Water Resour.
Res., 22, 191-197, 1984.
Bagnold, R. A., The nature of saltation and of 'bedload' transport in water,
Proc. Roy. Soc. Lond., A 332, 473-5-4, 1973.
Bluck, B. J., Sedimentation in the meandering river Endrick, Scottish J. Geol.,
7, 94-138, 1971.
/
Bridge, J. S., Flow and sedimentary process in river bends; comparison of field
observations and theory, in River Meandering:
Proceedings of the
Conference, Rivers '83, edited by C. M. Elliot, pp. 857-872, Am. Soc. Civ.
Eng., New York, 1984.
Bridge, J. S., and J. Jarvis, The dynamics of a river bend, a study of flow
and sedimentary processes, Sedimentology, 29, 499-541, 1982.
Carson, M. A., and G. A.. Griffith, Bedload transport in gravel channels, J.
Hydrology, 26(1), 1-151, 1987.
Church, M. A., D. G. McLean, and J. F. Wolcott, River bed gravels, sampling
and analysis, in Sediment transport in gravel-bed rivers: edited by C. R.
Thorne, J. C. Bathurst, and R. D. Hey, Wiley, Chichester, 43-88, 1987.
Dietrich, W. E., Flow, boundary shear stress, and sediment transport in a
river meander, Ph.D. dissertation, 261 pp., Univ. of Wash.-Seattle, 1982.
Dietrich, W. E., Settling velocities of natural particles, Water Resour. Res.,
18(6), 1615-1626, 1982.
Dietrich, W. E., Mechanics of flow and sediment transport in river bends, in
River Channels: Environment and Process, Richards, K. S., ed., Inst. Brit.
Geographers Spec. Pub. #18, Basil Blackwell Scientific Publications, pp.
179-227, 1987.
Dietrich, W. E., J. Kirchner, H. Iked~, and S. Iseya, The origin of the coarse
surface layer in gravel-bedded streams: the role of sediment supply, Geol.
Soc. Am. Abstracts with Programs. 19(7), 642, 1987.
Dietrich, W. E., and J. D. Smith, Influence on the point bar on flow through
curved channels, Water Resour. Res., 19(5), 1173-1192, 1983.
Dietrich, W. E., and J. D. Smith, Bedload transport in a river meander,
Water Resour. Res., 20(10), 1355-1380, 1984a.
Dietrich, W. E., and J. D. Smith, Processes controlling the equilibrium bed
morphology in river meanders, in River Meandering: Proceedings of the
Conference, Rivers '83, edited by C. M. Elliot, pp. 759-769, Am. Soc. Civ.
Eng., New York, 1984b.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

49

Dietrich and \Vhi ting

Dietrich W. E., J. D. Smith, and T. Dunne, Flow and sediment transport in a


sand-bedded meander, J. Geol., 87, 305-315, 1979.
Dietrich, W. E., J. D. Smith, and T. Dunne, Boundary shear stress, sediment
transport, and bed morphology in a sand-bedded river during high and low
flow, in River Meandering:
Proceedings of the Conference, Rivers '83,
edited by C. M. Elliot, pp. 632--639, Am. Soc. Civ. Eng., New York, 1984.
Diplas, P., and A. J. Sutherland, Sampling techniques for gravel sized
sediments, J. Hydraul. Div., Am. Soc. Civ. Eng., 114(5), 484-501, 1988.
Emmett, W. W., A field calibration of the sediment-trapping characteristics of
the Helley Smith bedload sampler, Prof. Paper 1139, U. S. Geological
Survey, 44 p., 1980.
Engelund, F., Flow and bed topography in channel bends, J. Hydraul. Div.,
Am. Soc. Civ. Eng., 100(11), 1631-1648, 1974.
Hills, R., Sediment sorting in meandering rivers, M. S. Thesis, Univ. of
Minnesota, 75 p., 1987.
Hooke, R. L., Distribution of sediment transport and shear stress in a meander
bend, Rpt. 30, Uppsala Univ. Naturgeografiska Inst., 58 p., 1974.
Hooke, R. L., Distribution of sediment transport and shear stress in a meander
bend, J. Geol., 83, 543-565, 1975.
Hunt, J. N., The turbulent transport of suspended sediment in open channels,
Proc., Roy. Soc. London (AJ, 224, 332-335, 1954.
Ikeda, S., Incipient motion of sand particles on side slopes, J. Hydraul. Div.,
Am. Soc. Civ. Eng., 108(1), 95-114, 1982.
Keller, E. A., Areal sorting of bed-load material, the hypothesis of velocity
reversal, Ceol. Soc. Am. Bull., 82, 753-756, 1971.
Leopold, L. B., and W. W. Emmett, Bedload movement and its relation to
scour, in River Meandering: Proceedings of the Conference, Rivers '83,
edited by C. M. Elliot, pp. 640-649, Am. Soc. Civ. Eng., New York, 1984.
Leopold, L. B., and M. G. Wolman, River meanders, Ceol. Soc. Am. Bull.,
71, 769-794, 1960.
Luque, R. F., and R. van Beek, Erosion and transport of bed-load sediment,
J. Hydraul. Res., 14(2), 127-144, 1976.
McLean, S. R., and J. D. Smith, Turbulence measurements in the boundary
layer over a sand wave field, J. Ceophys. Res., 84(12), 7791-7808, 1979.
Middleton, G. V., Hydraulic interpretation of sand size distribution, J.
Geology, 84, 405-426, 1976.
Nece, R. E., and J. D. Smith, Boundary shear stress in rivers and estuaries,
J. Water and Harbor Div., Am. Soc. Civ. Eng., WW2, 335-358, 1970.
Nelson, J. E., Mechanics of flow and sediment transport over nonuniform
erodible beds, Ph.D. dissertation, Univ. of Wash., Seattle, 227 p., 1988.
Nelson, J. E., and J. D. Smith, Numerical prediction of meander evolution,
EOS, 66, 910, 1985.
Odgaard, A. J., Meander-flow model, I.
Development, J. Hydraul. Engrg.,
112(12), 1117-1136, 1986.
Odgaard, A. J., and M. A. Bergs, Flow processes in a curved alluvial channel,
Water Resour. Res., 24(1), 45-56, 1988.
Parker, G., and E. D. Andrews, Sorting of bedload sediment by flow in
meander bends, Water Resour. Res., 21(9), 1361-1373, 1985.
Parker, G., K. Sawai, and S. Ikeda, Bend theory of river meanders, 2,
Nonlinear deformation of finite-amplitude bends, J. Fluid Meeh., 115,
303-314, 1982.
Petit, F., The relationship between shear stress and the shaping of the bed of
a pebble-loaded river La Rulles, Ardenne, Catena, 14, 453-468, 1987.
Smith, J. D., and S. R. McLean, Spatially averaged flow over a wavy
boundary, J. Ceophys. Res., 82(12), 1735-1747, 1977.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Boundary Shear Stress

50

Smith, J. D., and S. R. McLean, A model for meandering streams, Water


Resour. Res., 20(4), 1301-1315, 1984.
Struiksma, N., K. Oleson, C. Flokstra, and H. J. DeVriend, Bed deformation
in curved alluvial channels, J. Hydraul. Res., 23(1), 57-79, 1985.
Thorne, C. R., J. C. Bathurst and R. D. Hey, eds., Sediment transport in
gravel-bed rivers, Wiley, Chichester, 1987, 995 pp.
Wiberg, P. L., and J. D. Smith, Calculations of the critical shear stress for
motion of uniform and heterogeneous sediment, Water Resour. Res., 23(8),
1471-1480, 1987.
Yalin, M. S., An expression for bedload transportation, J. Hydraul. Div.,
Am. Soc. Giv. Eng., 89(3), 221-250, 1963.
Yen, C. L., and B. C. Yen, Water surface configuration in channel bends, J.
Hydraul. Div., Am. Soc. Giv. Eng., 97(2), 303-321, 1971.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Copyright 1989 by the American Geophysical Union.

Sedimentary Controls on Channel Migration and Origin of


Point Bars in Sand-Bedded Meandering Rivers
Hiroshi Ikeda
Environmental Research Center, University of Tsukuba, Japan

Abstract
Field observations in alluvial meandering rivers have revealed that plan form
and migration pattern of channels are strongly influenced by the distribution of
thick fine-grained cohesive deposits in flood plains. Meandering channels can be
classified into four types according to the influence of these cohesive deposits: (1)
fixed meanders, (2) restricted meanders, (3) confined free meanders, and (4) truly
free meanders. Laboratory experiments and field observations of point bars in low
gradient sand-bedded meandering rivers have revealed that, unlike in gravel-bedded
rivers, point bars develop in response to the pattern of flow through the bends, that
is point bars in these rivers are a result rather than a cause of meandering. Point
bar platforms, which are the flat tops of point bars, are gradually covered in these
rivers by inner bank accretions deposited out of suspension. Therefore, bends with
well developed point bar platforms lacking accretion deposits must be in a state of
active channel migration and, consequently widening.

Introduction
The River Teshio is the largest river in northern Hokkaido, Japan, and flows
into the Japan Sea. Its lower course, shown in Figure 1, is typical of Japanese rivers
in alluviating estuaries. The width of alluvial plain is not the product of lateral
erosion by the present river. Rather, it is a result of deposition of river-borne
sediment within a valley cut during the last Pleistocene low stand of the sea. The
channel characteristics of the lower Teshio River are shown in Figure 2. During the
1950s to the 1970s, many channel bends were artificially straightened. As a result,
channel gradient increased temporarily, which in turn caused channel widening
[Ikeda, 1983]. This enabled us to observe sediments in many cut banks.
We also observed depositional processes in and adjacent to the channel duri~
the past 10 years, especially just after the 1981 overbank flood fIkeda et al., 1984 .
We observed dunes along the lower Teshio River by echo-sounding surveys [Ike a
and Iseya, 1981; Iseya, 1982] and measured suspended sediment transport at
snow-melt floods to document the processes responsible for the sedimentary
structures of meander plain sediments. Moreover, we made some flume experiments
on the bed configurations in meandering channels [Ikeda, 1977; Kodama and Ikeda,
1.984; Ikeda et al' 1985] and also on the suspended sediment transport over dune
f
fields [Iseya, 1984J.
In the course of the U.S.-Japan cooperative research work on river meandering,
we observed some rivers in the U.S. and other countries. We compared flood plain
deposits in several different environments. As a result of these observations and our
51
Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sedimentary Controls

52

hills and terraces


wind dunes
alluvial plains
meander plains
~

abandoned
channels
artificially
cut-off chanels
10 distance from
the river mouth
(km)

5 km

I
---------~I

Fig. 1.

Meander plain in the lower Teshio River, northern Hokkaido, Japan. Drainage
area is about 5,000 km2. Mean discharge is about 200 m3 /s.
Mean annual
maximum discharge and the maximum discharge is about 1,500 m3 /s and 3,100
The hills are mainly composed of Tertiary unconsolidated
m 3Is, respectively.
fine-grained sedimentary rocks.

more detailed investigation of the Teshio River, we feel the importance of


sedimentary controls on channel shifts in meandering rivers should be emphasized,
as others such as Bluck [1971], Ferguson [1987], Friend [1983] and Mosley [1987], etc.
have noted. The purpose of this report is to describe the types of alluvial river
meanders, to propose an hypothesis on the formation of abrupt changes in channel
patterns, and to discuss the origin of point bars in order to suggest the essential
factors of river meandering~ Alluvial meandering channels are classified into four
types, as shown in Figure 3, based on the state of sedimentary controls on channel
shifting: (1) fixed meanders, (2) restricted meanders, (3) confined free meanders,
and (4) truly free meanders.
Fixed Meanders in Deltaic Plains
It is well known that most alluvial plains in the coastal locations of Japan were
formed during the late Holocene period after the post-glacial transgression. The
channel in a deltaic plain has rather straight reaches. But, at the river mouth of a
delta stream, a mid-channel shoal tends to develop. These often become bars and

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Ikeda

53

10

.:

\ oed

en 3 0 (\e
. :' :.....
...
.

......:

.t:.

0)

G)

J:

-5

-10

..

200

100

.s::

.... '.'
:-..

a
~: ...:.. ~

: :"

...~.

: - - .;1-

:_"': _ :: :

:: ::.:.::. '"

.:.:-:,-.\ ::..:

".

;-,':. ::~

:.:

o+-------r-----

..

10

.s::

a.
G)

...

...

.:':'.:'. :.:.:..:.....
.~
:.....:..... 1.......

2000+--------r----r-------:--------r--N

..

d
1000 '::.:

"t:.:..: ~~;.:..:...:- :..:: : :.~.\.;.:.::..:: ~:.r:..::::; :: .-.:~, ..-.:

::.._.:.

G)

o~

a.
E
o

Fig. 2.

Channel characteristics of the lower Teshio River.


(a) longitudinal profiles of
alluvial plain (former deltaic plain), present flood plain and mean bed heights; (b)
channel top width;
(c) mean channel depth;
(d) cross sectional area of the
channel; (e) grain size composition of bed materials (Based on data of Hokkaido
Development Bureau).

eventually islands which bifurcate the channels. Bifurcation may be developed also
within a few years in bays that suddenly become subject to alluviation due to
formation of a crevasse channel through the natural levee of active channels
[Russell, 1967, p. 521. The channel pattern of a delta stream, therefore, superficially
resembles subaerial braided or anastomosing flood plain channels. As a deltaic plain

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sedimentary Controls

54
+-0----

meander plain

")

b
meander-belt
deposits

c
fine-grained' older deposits
~E---- floodplain

d
fine-grained'deltaic plain deposits

Fig. 3.

Types of alluvial meanders classified by degree of sedimentary controls on channel


shifts. (a) truly free meanders, arisen from a deficiency of fine-grained sediment
supply, (b) confined free meanders formed by a high supply of fine-grained
sediment, (c) restricted meanders with confinement consisting of older cohesive
deposits, (d) fIXed meanders on deltaic plains.

extends seaward, one of these distributed channels is usually preserved as a main


sinuous course.
A more important change due to channel extension during delta growth is the
entrenchment of the channel bed. The entrenchment must occur because the lowest
part of the deltaic channel bed rises toward a seaward bar, so that as the channel
advances it must incise through the mouth bar and delta-front deposits, as shown in
Figure 4. These deposits are composed of colloids, clay, and finest silt, which,
although easily transported, form a cohesive resistant bank to the incised extending
channel. The resistant banks retard channel migration, leading to a "fixed"
channel. In the case of the middle course of the lower Teshio River (10-20 km from
the river mouth), the flood plain height of the present channel is nearly the same as
the former deltaic plain height, as shown in Figure 2a. Here overbank deposits are
carried away from the meander zone and across the deltaic plain, contributing to a
narrow flood plain. The meandering channel here does not appear to be in a fixed
configuration. However, it is actually strongly confined by cohesive deltaic deposits
lying beneath the thin flood plain deposits. At an earlier stage of flood plain
construction, the rapid migration of the channel is prevented by cohesive banks and
the low gradient of the stream. However, as the stream power of the flow in the
channel increases, due to slope increase as the channel extends and aggrades, the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Ikeda

55

_B
om
....

Bay

Subaerial levee

"\
\
\

\ :..:~! "~: t:;_ /' /


.

,,

~~~

-1 m

~~ .: : .:' .-~queous
-2 m

,
\

,
\

..

,,

""

" " ....

... ~,

Prodelta
Delta front

a
o
!

rnA
0-

natural levee deposits

=====.-;;;;;-.II!!-C"~_:":_--==:~~':"':::~"7~'.

/'

100 m
I

-1 -

-2-

b
Fig. 4.

~ '---'~d

channel bed deposits

flne-gralne
deltaic plain deposits

Diagrammatic representation of the sedimentation zones in the Koise delta, Ibaraki


Prefecture. (a) plan view, (b) channel cross section, showing channel incision into
mouth-bar or delta front cohesive deposits due to the extension of deltaic stream
[Based on data of Mikami et al., 1983].

fixed meandering channel changes its plan form gradually. This transition occurs in
the bends of the middle course of the lower Teshio River.
When the extending river channel encounters the coarser shoreline sediments
known as delta front sheet sand, that are transported laterally by waves [Reineck
and Singh, 1971, p. 270], the channel increases its width suddenly. This occurs
along the lower course of the lower Teshio River (below 10 km from the river
mouth).

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sedimentary Controls

56

A------1Il::-L--

_---+---+-""'t----

stough

/~:

point bar
~t-- ptatform

:::~/:I accretionary

A~T~:k
200 m

""'--.........~I

A
o

meander :{;hannel
plain

10 +-1--+----+1--t-----.,Ir---o
200
400 m

Fig. 5.

Plan form of channel and meander plain at 27 km station in the upper course of
the lower Teshio River.

Fig. 6.

Channel sinuosity in a sinuous confined trough increases when the meandering


channel is in phase with the sinuous confined trough (a), but the channel becomes
straight and wide when they are out of phase (c).

Restricted Meanders Formed by Incised Channels onto Deltaic Plains


In the upper course of the lower Teshio River (20-10 km from the river mouth),
the present river channel clearly cuts into the deltaic plain. The difference in height
between the former deltaic plain and the present meander plain, which is an area
cut and filled by the present channel, is several meters, as shown in Figures 5 and 6.
This channel incision onto a former deltaic plain may be due to uneven uplift of the
land. Lateral channel shift is limited within a narrow meander plain. Such
deformities [Mattes, 1941] or interruptions in the orderly development of meander
loops, resulting from local causes, were described by DeGeer [1906] and later
examined by Sundborg [1956] along the Klaralven. Lewin and Brindle [1977] also
described confined meanders.
Friend [1983] recognized the importance of
sedimentary supply for flood plain construction. The whole pattern of the upper
course of the lower Teshio River is that of restricted meandering, for the reason that
the width of the flood plain is not sufficient to permit full development of
appropriate loops.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

57

Ikeda

As the meandering channel migrates downstream in a sinuous confined trough,


the sinuosity of the restricted meandering channel changes in time, as shown in
Figure 6. The migration rate of the channel and, therefore, the rates of reworking
and accumulation of sediment in the restricted meander plain are also influenced by
the plan form of the confined trough. The rate of downstream passage of successive
bends within the meander plain in the upper course of the lower Teshio River
(Figure 6) is estimated to be on the order of 100 years.
meander plain
" (meander headlands)

confinements composed of older sediments

Fig. 7.

A meandering channel is compelled to bend abruptly when the channel. impinges


on confinements in a restricted or confined meandering channel. [Based on map of
Sundborg, 1956].

Impingement of the channel against confinements causes a unique channel plan


form of the restricted meandering, that is, the channel impinges the confined wall
and hugs the wall for some distance until a bend is initiated toward the opposite
trough wall, as shown in Figure 7. Although Ikeda [1985] and Parker et ale l1983]
claimed this behavior to be typical of all rivers, calling it the "Kinoshita curve," we
think this plan form is characteristic of restricted meanders. It seems that
concave-bank benches [Woodyer, 1975; Page and Nanson, 1982] are well developed
in restricted meandering rivers.
Confined Free Meanders Developed on Aggraded Flood Plains
Aggradation during the late Holocene has caused some channels to form thick
accumulations of flood deposits, that have set the channel free from the
confinements of former deltaic plain deposits. These channels have been called
freely-meandering channels.
It has been often considered that the
freely-meandering channels have been able to migrate over the entire part of their
wide flood plains during the late Holocene. Hence, the entire flood plain is an
As mentioned later, there are truly,
assemblage of meander plains.
freely-meandering channels which migrate freely over the entire flood plain, as
shown in Figure 3a; however, such streams are rare in Japan. In fact, most
apparently freely-meandering channels have two kinds of sedimentary confinements,
composed of thick, fine-grained flood plain accretions, which prevent free migration
of the meander belt.
One of these are clay plugs. Fisk [1952] showed that the local alignment and
configuration of the channel of the Mississippi River is governed to a considerable
extent by local variations in bank materials. He stressed the role of clay plugs, Le.,
deposits of silt, clay, and organic matter accumulated in old cut-off lakes [Fisk,
19471, and proposed that where thick clay plugs are present in considerable
numbers, they confine the course of a river to a limited meander belt. Clay plugs

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sedimentary Controls

58

are highly resistant to erosion by direct shear, but susceptible to slumping in large
blocks when undermined.
The other of these are flood basin accumulations. Although the meandering
channels are shifting laterally, during overbank floods fine sediment is carried
further across the flood plain and deposited. On a sufficiently large flood plain or a
relatively slowly migrating meander belt, thick accumulations of fine sediment
mixed with organic matter produced in a swampy environment will develop
adjacent to the meander belt; there serve to confine it to a region wider than
individual bends but much narrower than the entire flood plain. We propose here
the classification "confined freely-meandering channels" for such rivers, as shown in
Figure 3b.
Shifting of the entire meander belt in such channels occurs only by diversions or
avulsions, when the river escapes from its confined meander belt and cuts an
entirely new path. During the last thirty centuries, however, there have been only
two or three major diversions of meander belts in the lower Teshio River, as shown
in Figure 1; this is true of many other flood plains in Japan. The presence of
abandoned meander belts on flood plains is misleading, in that it suggests that the
channel migrated freely on the entire flood plain.
It has been argued that avulsion-dominated meandering rivers originate with
high coarse suspended load which causes a river to form high levees and aggrade its
bed above the flood plain [Allen, 1965, p. 126]. We would like to stress the presence
and role of cohesive sediments which restrict the channel to a meander belt.

Truly Frre Meanders Without Fine-Grained Flood Baa<lin


Accumulations

Wandering meanders (Figure 3a) can be observed on those plains without thick,
fine-grained flood basin accumulations which restrict the meander belt within a
narrow range. We investigated the Nuporomaporo River (Figure 8), a tributary of
the lower Teshio River, which is a gravel-bed meandering river in a wide flood plain
[Ikeda and Iseya, 1986]. There are very thin overbank accumulations and no thick
clay-plugs, which means the basin produces very little finer-grained material as
wash-load, or alternatively that very little of the fine load produced is deposited
because the channel is too steep.
N

Fig. 8.

Location of the Nuporomaporo River, a tributary of the lower Teshio River.


mountains, 2. hills, 3. alluvial plains.

Copyright American Geophysical Union

1.

Water Resources Monograph

River Meandering

Vol. 12

Ikeda

59

---

1969~ __ ~, 1971
-- -,,"- ",'-- "
" " " /'
..
.....

,,'

"

Fig. 9.

'---

/II

100

Rapid channel migration at the 2 km location in the Nuporomaporo River.

Cut-<>ffs of channels have frequently occurred in this river, as shown in Figures 9


and 10, and many abandoned channels are preserved in the flood plain, as shown in
Figure 11. l'he lack of tines prevents filling up of old channels; they remain
topographic lows in the flood plain. Such a meandering river, flood plain and which
migrates freely over the entire flood plain without confinements within it, should be
called truly free meanders. Why is the Nuporomaporo River not braided? We

1977

t:

,
0

Fig. 10.

'f
I

Channel cut-offs along the Nuporomaporo River.


successive years.

/tJ

5 km

Arrows indicate cut-offs between

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sedimentary Controls

60

200 m

Fig. 11.

Abandoned channels preserved on the flood plain of the Nuporomaporo River at


the 3 km location.

suppose that the Nuporomaporo River became meandering following a decrease of


sediment supply during the Holocene period.
In the Nuporomaporo River, impingements of the channel against bedrock bluffs
of a Tertiary formation produce channel deformities. Bank vegetation of thick
bamboo grass also often plays a significant role in the stabilization of banks for
maintaining the meandering in this river; tree stumps become bank obstruction in
moderate sized channels [Lisle" 1986].
Explanations for Abrupt Changm in Channel Patterns
Although there is a great range of channel patterns, from straight through
meandering to braided and anabranching, or anastomosing, [Brice, 1984], we can
often see abrupt and clear changes in channel patterns both in time and space. Why
does the channel pattern change abruptly? In some rivers, braided and meandering
courses alternate with time. It is possible that fluctuation of sediment supply causes
the alternation of braided and meandering pattern on the same reach. Bradley
[1984] reported how a meandering river became braided following the huge sediment
supply resulting from the eruption of Mount St. Helens. Evidence for a similar
channel evolution was observed among the terrace deposits in the Nishibetsu River,
eastern Hokkaido, indicating a large change in sediment supply due to pyroclastic
flow deposits [Kumagai 1980J.
Why does the channel pattern change abruptly from the meandering into the
braided by an increase in input of sediment? The idea of sedimentary controls on
channel shifting gives a plausible explanation. The rapid a~radation of the channel
floor, which resulted from the excess sediment supply, 'frees" the meandering
channel from its confinement. Although the braided channel pattern does not
always seem to develop in aggrading rivers only, a large number of braided channel
pat terns are found in aggrading rivers.
One more possible explanation for abrupt changes in channel patterns, indeed
one which does not rely on environmental disturbances, is based on the observation
that the lateral channel migration rate controls and limits the amount of the flood
basin accumulation, since lateral erosion would eliminate the deposits Wolman and
Leopold, 1957]. The rate of channel shifting across the flood plain is re ated to bank
stability, which in turn is a function of the texture of the flood plain deposits.
Braiding involves rapid channel movement, which prevents the accumulation of
thick fine material adjacent to the channel. Thus, braiding, once initiated, may be
amplified by a positive-feedback process. On the contrary, once a meandering

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

61

Ikeda

channel is formed, finer-grained sediments are well-preserved adjacent to the


meander belt. They prevent lateral migration of the channel, so that the vertical
accumulation becomes more dominant. This explanation is also applicable for
abrupt changes between meandering and fixed channels.
Although we have no way to test this hypothesis now, our hypothesis is suitable
to explain alternations of channel pattern from braided to meandering within a
short distance, in some rivers without tributaries or bedrock constraints. Our
hypothesis will also be applied to the appearance of point bar platforms later.
Origin of Point Bars and Appearance of Point Bar Platforms with Regions of
Channel Overwidening
Along the upper course of the lower Teshio River (Figure 12), emergent point
bars, sometiInes called point bar platforms, are visible here and there on the convex
sides of bends at low flow stage. The distribution of point bars corresponds to the
plan view of the channel. The bars remain fixed relative to the channel, and do not
migrate downstream independently of channel migration. A question thus arises as
to whether the bars developed due to the channel bends, or whether the channel
began to meander after the bars developed in a straight channel?

28 km

Fig. 12.

Plan form, low-flow channel (in black) and emergent point bars (in white) on the
meandering reaches of the lower Teshio River. Emergent point bars are visible
only at bends.

Many experiments indicate that a shallow stream with a steep slope, Le., one
with a large value of the channel-forming index S(W/D), where S = slope, W =
width, and D = depth, [Ikeda, 1973], will make alternate bars in a straight flume
[e.g., Kinoshita, 1961]. These bars usually move downstream steadily [cf. Ikeda and
Ohta, 1986]. We call such bars "migrating bars." If the bend curvature is less than
a critical value and the walls are rigid, these migrating bars travel downstream
through the sinuous channel. The bed relief reaches its maximum value when the
migrating alternate bars are in phase with the stationary bars forced by channel

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Sedimentary Controls

62

Fig. 13.

Vol. 12

Alternate bar height in a sinuous flume with small bend curvature reaches its
maximum value when the migrating alternate bars are in phase with the stationary
forced bars (a, e), and the bed surface becomes flat when they are out of phase
(c).

curvature, mentioned below. On the other hand, migrating and forced bars
counteract each other, and the bed surface becomes flat, when they are out of phase,
as shown in Figure 13. However, in more sinuous channels, I(inoshita and Miwa
[1974] found that these bars no longer migrate when the deflection angle of the bend
exceeds a critical value (ca. 20). Those migrating bars which become fixed in
position by the bend are termed fixed bars or stabilized bars, shown in Figure 14.
Straight flume

e=0

__:>:,.~r-~.",. ""f;Y
c=""

High

..,;';.:::;:::'i')-

Alternate
migrating bars

t
SW/D

1
Low

Sinuous flume

e < Be

Be

~.6~ fJ

~ ~ ~ 1:f!!:~P

~~~.~
Superposed on
forced bars

Fixed bars

---------

Fig. 14.

No-bars

Forced bars

Classification of alternate bars in straight and sinuous flumes.

In contrast, for flow with a low channel-form index, migrating bars are not
formed in straight channels. In sinuous channels of this type, stationary bars
develop along the convex banks. These bars are termed "forced" bars, because they
only develop in response to the pattern of flow through the bend. Thus we have two
kinds of alternate stationary bars. One type is the fixed bar, and the other type is
the forced bar. To elucidate the origin of river meandering and to control the
natural river channel appropriately, it is necessary to determine whether the bars in
sinuous channels are originated from fixed bars or forced bars.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Ikeda
A quick way to distinguish fixed bars from forced bars is to examine hydraulic
conditions of the flow over the bars. In many rivers, as the gravel content declines
and fine sand increases in content downstream in the bed and banks, the channel
narrows and becomes less steep. In this case bar formation gradually ceases
downstream, as the channel-form index becomes smaller [Ikeda, 1975]. We found
that bankfull flows in the lower Teshio River, as with many sand-bedded
meandering rivers in Japan, has a channel-form index that is too low to develop
migrating bars if the channel were straight, as shown in Figure 15. However, a
definite judgment cannot be made, because river discharge has changed in time.

1 0 __-...a.....----+-1-------'1'------11r------~--__41---------+_
:

o. --

"A

- - - . "_

~"

no bars
.......

,--

--

q/ I.

"

'~
.......

upper flow reg;m

-.. .:.;~' -... -.". f.

_0

--

......-

~.....- alternate

'.

'.

~4

barS
~

12
o 3

-~-~".

transition
0",

0.01

.....-

~ ~ ~

"0",

0.1

SW/D
Fig. 15.

@1

--

....- .""..
---multiple-row bars
I

10

Bankfull flows in some Japanese sand-bed meandering channels have too low a
channel-form index to develop bars if the channel were straight. Here shear
velocity U* is computed as .f9l[S, where 9 is the acceleration due to gravity, R is
hydraulic radius, and W, D, and S are mean width, depth and slope of
meandering channels, respectively; U* is critical shear velocity of the bed material,
determined from the Shields diagramc for median grain size of bed material. (1)
Tone River, (2) Arakawa River, (3) Edo River, (4) Teshio River.

A decisive way to determine whether bars will form is to straighten the channel.
Forced bars will disappear in the straight channel, but if the bars in the sinuous
channel were fixed, then the straightened channel will display alternate bars.
Examination of artificially straightened Japanese rivers has shown that the bars in
sand-bedded meandering rivers are forced bars, because all bars disappeared when
the channel was straightened, e.g. Figure 16 [Ikeda, 1977]. Many people believe that
alluvial meanders are caused by the formation of alternate migrating bars in
initially straight alluvial channels [/(eller, 1972; Leopold, 1982; Thompson, 1986].
Our conclusion rejects this hypothesis for narrow, deep channels. It is concluded
that the point bars in the lower Teshio River are forced bars. They are composed of
bedload materials, Le., coarse sand and gravel.
Over the point bar platform, gently sloping deposits called bank deposits
composed of silt and fine sand may deposit out of suspension. Point bar platforms
are gradually covered by those deposits, as shown in Figure 17 [Ikeda et al., 1984;
Iseya and Ikeda, 1988]. In the upper course of the lower Teshio River, we have
found that point bar platforms appeared only at rapidly migrating bends.
Elsewhere, the bank resistance along cut-banks is high enough to allow the
platforms to be completely covered with bank slope deposits, causing the channel to
have a minimum ~Tidth. This is illustrated in Figure 18 [Ikeda and Iseya, 1987).
Fisk [1952] has proposed that growth of point bars favors channel narrowing,
which in turn, causes accelerated erosion of the concave bank sufficient to maintain

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

64

Sedimentary Controls

1 km

Fig. 16.

Alternate bars form in a sinuous reach (a), but no bars develop in an artificial
straightened reach (b), in the lower Arakawa River, Saitama Prefecture, Japan.
Drawn from air photographs taken in 1962. Channel slope is 1/4000, width is 100
m, bankfull depth is 4 m, and the median grain size of the bed material is 0.5
mm for both reaches.

FACIES
PROFILE

ENVIRONMENT

OVERBANK

IN-CHANNEL
POINT BAR

PROCESS

Traction

THICKNESS

BANK

NATURAL LEVEE

Suspension

5-6 m

2-3 m

BEDDING

Epsilon

Tabular inclined

Wavy

STRUCTURE

Trough cross
stratification

Structureless or
climbing ripple
lamination

ParaDel and small


ripple lamination

TEXTURE

Pebble-bearing
coarse sand

Fine sand
and mud

Fig. 17.

BACK SWAMP

Suspension
2-3 m

Medium to fine
sand and mud

Horizontal
Thinly laminated
or structureless
Mud

Schematic model of flood plain sediments in the lower Teshio River.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

65

Ikeda

OVERBANK DEPOSITS

(%ff!(,{~ ]
i}:!(:/'/;'
.. .::f/': :fj.

"'IN-CHANNEl DEPOSITS

:0.

~"'~"""'BEDROCKS
~@L

J.. .

(shale)

BANK DEPOSITS

~;j;j;' }'CHANNEl-BED DEPOSITS


_____ -=-_-=.~.:-:;...~
===.: OLDER VALLEY-FILL
.......;. .~.-t

- - - -

- - -

=
-

- - -

DEPOSITS (silt & clay)

c
)

o
!

Fig. 18.

... OLDER FLOODPLAIN


DEPOSITS

100 m

A point bar platform appears at a rapidly migrating bend (c), but is covered with
bank slope deposits where the bank resistance along the cut-bank is sufficiently
high enough (a).
In case (a), the channel maintained a cross-section that is
narrow and deep. The location of each section is shown in Fig. 12.

its cross-sectional area. The higher the strength of bank-forming materials in the
cut banks, the narrower will be the channel width in bends. This hypothesis is
based on the idea that point bars are mainly developed by deposition due to high
bedload transport rate, and that bar growth is the cause of bank retreat along
opposite sides. If this is true, cut-bank retreat can be prevented by the excavation
of point bar deposits. Dr. W. Dietrich [personal communication, 19881 believes this
is a practice followed, successfully, by some engineers. It is not just the narrowing,
but the deflection of the high velocity core against the outer bank by the bar that
contributes to bank erosion.
Another explanation is that channel width is influenced by relative rate of
accretion and bank recession. Bar growth is viewed as due to overwidening of the
channel by recession of cut banks. Bends with well-developed point bar platforms
therefore are in a state of acti ve lateral migration. The bank retreat rate is affected
by channel curvature [Hickin and Nanson, 1975; Hanson, 1980]. If the channel is
straight, channel migration ceases and, therefore, point bars do not develop even if
the channel bank-forming material is cohesionless. In such a case, a straight
channel without bars would not necessarily imply a small amount of bedload. So
how do these channels start to meander? Most explanations of meandering usually
trace the transformation from initially straight channels. We think, however, that
meander bends are transmitted from upstream to downstream.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sedimentary ControIs

66

Conclusions
Alluvial meandering channels are classified into four types based on the state of
sedimentary controls on channel shifting: (1) fixed meanders on deltaic plains
resulting from channel extension towards the sea and channel incision into cohesive
delta-front deposits, (2) meanders restricted due to confinements composed of older
deposits, which restrict the meander plain width and prevent full development of a
meander belt, (3) confined free meanders, caused by confinements composed of thick
flood basin accumulations which prevent freely meander belt migration, and (4)
truly free meanders resulting from a lack of confinement within flood plains,
allowing the channel to migrate freely over the entire flood plain.
It has been assumed that alluvial meanders are caused by the formation of
alternate migrating bars in initially straight alluvial channels. We suggest,
however, that the point bars in sand-bedded meandering rivers are a result rather
than a cause of meandering. Explanations of meandering usually trace the
transformation from initially straight channels. However, transmission of meander
bends from upstream to downstream seems to be necessary to explain the
development of meandering.
Moreover, hydraulic models for development of meandering usually assume a
uniform distribution of bank material. In natural streams, however, this is rare.
Sufficient bank resistance is necessary to maintain a narrow, deep meandering
channel, because excessive bank retreat associated with bank materials may result
in a wide, shallow braided channel. Most actively meandering channels in sand
result from overwidening. It is also important to take into account this fact for the
analysis of flow and bed topography in meandering channels.
River channels change not only due to artificial disturbances, but also due to
more long-term natural disturbances. A careful analysis of meander anomalies
would yield considerable insight into present fluvial systems.

Acknowledgements
The research on which this paper is based was supported by Teshio Town,
Hokkaido. The writer wishes to gratefully acknowledge F. Iseya who performed
field work and flunle experiments. Thanks are also due to W. B. Dietrich for helpful
comments and the wording in English of an earlier version of the manuscript. This
research was supported by the Joint U.S.-Japan Research Workshop promoted by
NSF and JSPS.

References
Allen, J. R. L., A review of the origin and characteristics of recent alluvial
sediments, Sedimentology, 5, 89-191, 1965.
Bluck, _B. J., Sedimentation in the meandering River Endrick, Scot. J. Ceol., 7,
93-138, 1971.
Bradley, J. B., Transition of a meandering river to a braided system due to high
sediment concentration flows, in River Meandering, edited by C. M. Elliott, pp.
89-100, Proceedings of the Conference Rivers '83, New York, American Society
of Civil Engineers, 1984.
Brice, J. C., Planform properties of meandering rivers, in River Meandering, edited
by C. M. Elliott, pp. 1-15, ProceedingS of the Conference Rivers '83, New York,
American Society of Civil Engineers, 1-15, 1984.
DeGeer, S., Om KlaraJven och dess dalgang. Ymer, 26, 383-414, 1906.
Ferguson, R., Hydraulic and sedimentary controls of channel pattern, in River
Channels, edited by K. Richards, pp. 129-158, Blackwell, Oxford, 1987.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

67

Ikeda

Fisk, H. N., Fine-grained alluvial deposits and their effects on Mississippi River
activity, U.S. Army Corps of Engineers, Mississippi River Commission,
Vicksburg, Mississippi, 1947.
Fisk, H. N., Geological investigation of the Atchafalaya Basin and problems of
Mississippi River diversions, U.S. Army Corps of Engineers, Mississippi River
Commission, Vicksburg, Mississippi, 1952.
Friend, P. F., Towards the field classification of alluvial architecture or sequence,

Special Publications of the International Association of Sedimentologists, 6,

345-354, 1983.
Hickin, E. J. and G. C. Nanson, The character of channel migration on the
Beatton River, Northwest British Columbia, Canada, Bull. Geol. Soc. Am., 86,
487-494, 1975.
Ikeda, H., A study on the formation of sand bars in an experimental flume,
Geogr. Rev. Japan, 46, 435-451, 1973 (in Japanese with English abstract).
Ikeda, H., On the bed configuration in alluvial channels: their types and condition
of formation with reference to bars, Geogr. Rev. Japan, 48, 712-730, 1975 (in
Japanese with English abstract).
Ikeda, H., On the origin of bars in the meandering channels, Bull. Environ. Res.
Ctr., Univ. of Tsukuba, 1, 17-31, 1977 (in Japanese).
Ikeda, H., and F. Iseya, On the length of dunes in the lower Teshio River,
Transaction of Japanese Geomorphological Union, 2, 231-238, 1981.
Ikeda, H., River channel changes due to artificial cutoffs in the lower Teshio River,
Hokkaido, Human Culture and Environmental Studies in Northern Hokkaido,
Univ. of Tsukuba, 4, 13-22, 1983 (in Japanese).
Ikeda, H., F. Iseya, and K. Tezuka, Depositional facies of flood plain deposits in
the lower Teshio River, Hokkaido, Human Culture and Environmental Studies in
Northern Hokkaido, Univ. of Tsukuba, 5, 13-29, 1984 (in Japanese).
Ikeda, H., F. Iseya, and Y. Shinzawa, Christate bed configuration in a meander
bend, Bull. of the Environ. Res. Ctr., Univ. of Tsukuba, 9, 27-42, 1985 (in
Japanese).
Ikeda, H. and F. Iseya, Meandering in the River Nuporomaporo, Human Culture
and Environmental Studies in Northern Hokkaido, Univ. of Tsukuba, 7, 25-33,
1986 (in Japanese).
Ikeda, H. and A. Ohta, On the stationary bars in a straight flume, Annual Report of
the Institute of Geoscience, Univ. of Tsukuba, 12, 42-46, 1986.
Ikeda, H. and F. Iseya, Point bar formation in the lower Teshio River, northern
Hokkaido, Japan, Human Culture and Environmental Studies in Northern
Hokkaido, Univ. of Tsukuba, 8, 23-27, 1987 (in Japanese).
Ikeda, S., Bed topography in channel bends, in Hydraulics of Sediment Transport
(Ryusa-no-Suirigaku), edited by H. Kikkawa, pp. 221-247, Maruzen, Tokyo,
1985 (in Japanese).
Iseya, F., A depositional process of reverse graded bedding in flood deposits of the
Sakura River, Ibaraki Prefecture, Japan, Geogr. Rev. Japan, 55, 597--613, 1982
(in Japanese with English abstract).
Iseya, F., An experimental study on dune development and its effect on sediment
suspension, Environ. Res. Ctr. Papers, Univ. of Tsukuba, 56 pp., 1984.
Iseya, F. and H. Ikeda, Sedimentation in coarse-grained sand-bedded meanders:
distinctive deposition of suspended sediments, in Sedimentary Facies in the
Active Plate Margin, edited by T. Taira and F. Masuda, 1988, in press.
Keller, E. A., Development of alluvial stream channels: a five stage model, Bull.
Geol. Soc. Am., 83,1531-1540,1972.
Kinoshita, R., Investigation of channel deformation of the Ishikari River, Report of
Bureau of Resources, Dept. Science and Technology, 174 pp., 1961 (in Japanese).
Kinoshita, R. and H. Miwa, River channel formation which prevents downstream
movement of transverse bars, Shin-sabo, 94, 12-17, 1974 (in Japanese).
Kodama, Y. and H. Ikeda, Experimental study on the simulation of a free

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sedimentary Controls

68

meandering channel, Bull. Environ. Res. Ctr., Univ. of Tsukuba, 8, 29-36, 1984
(in Japanese).
Kumagai, M., Terrace topography along the Nishibetsu River, eastern Hokkaido,
M.S. thesis, Univ. of Tsukuba, 1980 (in Japanese with English abstract).
Leopold, L. B., Water surface topography in river channels and implications for
meander development, in Gravel-bed Rivers, edited by R. D. Hey, J. C.
Bathurst and C. R. Thorn, pp. 359-389, John Wiley & Sons, New York, 1982.
Lewin, J. and B. J. Brindle, Confined meanders, in River Channel Changes, edited
by K. J. Gregory, pp. 221-233, John Wiley & Sons, Chichester, 1977.
Lisle, T. E., Stabilization of a gravel channel by large stream side obstructions and
bedrock banks, Jaccoby Creek, Northwestern California, Bull. Geol. Soc. Am.,
97, 999-1011, 1986.
Mattes, G., Basic aspects of stream meandering, Trans. Am. Geophys. Union, 22,
632--636, 1941.
Mikami, Y., Y. Nagamine, M. Inokuchi, and S. Shindou, Environmental geoscience
of alluvial lowland around Lake Kasumigaura: Koise River delta sedimentation,
Tsukuba-no-Kankyo-Kenkyu, Univ. of Tsukuba, 7, 141-157, 1983 (in
Japanese).
Mosley, M. P., The classification and characterization of rivers, in River Channels,
edited by K. Richards, pp. 295-320, Blackwell, Oxford, 1987.
Nanson, G. C., Point bar and flood plain formation of the meandering Beatton
River, northeastern British Columbia, Canada, Sedimentology, 27, 3-29, 1980.
Page, K. and G. Nanson, Concave-bank benches and associated flood plain
formation, Earth Surface Processes and Landforms, 7, 529-543, 1982.
Parker, G., P. Diplas and J. Akiyama, Meander bends of high amplitude, J.
Hydraul. Div. Am. Soc. Civ. Eng., 109, No. HYI0, 1323-1337, 1983.
Reineck, H. E. and I. B. Singh, Depositional Sedimentary Environments,
Springer-Verlag, Heidelberg, 439 pp., 1973.
Russell, R. J., River Plains and Sea Coasts, Univ. of California Press, Berkeley, 173,
pp., 1967.
Sundborg, A., The river Klaralven, a study of fluvial processes, Geografiska
Annaler, 38, 127-316, 1956.
Thompson, A., Secondary flows and the pool-riffle unit: a case study of the
processes of meander development, Earth Surface Processes and Landforms, 11,
631--641, 1986.
Wolman, M. G. and L. B. Leopold, River flood plains: some observations on their
formation, U.S. Geol. Surv. Prof. Pap., 282C, 87-107, 1957.
Woodyer, K. D., Concave bank benches on the Barwon River, New South Wales,
Australian Geographer, 13, 36-40, 1975.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Copyright 1989 by the American Geophysical Union.

Flow In Meandering Channels With Natural Topography


Jonathan M. Nelson t and J. Dungan Smith

Geophysics Program, University of Washington, SeatUe

Abstract
A predictive model for velocity, surface elevation, and boundary shear stress
fields in natural meander bends is presented. The approach described herein retains
streamwise convective accelerations in the lowest order momentum equations and,
therefore, can be used to investigate flow in, channel bends with curvature and
topography that vary significantly in the streamwise direction, as is the case in most
natural streams and rivers. The total boundary shear stress field obtained from our
computations includes both skin friction and the form drag associated with pressure
forces on bars and bedforms. A simple method for removing the form drag
component is described, thereby permitting computation of the bottom stress
responsible for sediment transport. Results of our calculations are compared to
measurements obtained in Muddy Creek, a sand-bedded meandering stream in
Wyoming possessing the characteristics typical of natural channels.
Model
predictions of velocity, boundary shear stress, and surface elevation are shown to be
in excellent agreement with the measurements, primarily due to the careful
treatment of convective accelerations in the governing equations.

Introduction
Accurate specification of the spatial distribution of bottom stress in fluvial and
estuarine systems is of crucial importance in calculations of patterns of erosion and
deposition, primarily because erosion and deposition depend on the divergence of the
sediment flux and sediment transport rates are sensitive functions of boundary shear
stress. Unfortunately, direct measurement of boundary stress in such systems is
often very difficult. This is especially true in reaches where large convective
accelerations are produced by the presence of curvature and irregular bed
topography. The amount of data required to resolve the structure of the stress field
in systems of this nature usually makes a proper experimental study prohibitively
expensive and time-consuming. Most stream and river reaches are curved or have
irregular beds, so many important sediment transport problems are poorly
characterized as a result of insufficient or inaccurate knowledge of the stress field.
The difficulty encountered in measuring boundary shear stresses in nonuniform
flows indicates the desirability of a physically-based model which accurately
predicts these values. Many investigators [e.g., Engelund, 1974; De Vriend, 1977;

tNow at U.S. Geological Survey, Water Resources Division, Lakewood, CO


69
Copyright American Geophysical Union

80225.

Water Resources Monograph

70

River Meandering

Vol. 12

Flow in Meandering Channels

Ascanio and [(ennedy, 1983] have attempted to characterize the flow and bottonl
stress fields in curved channels. In order to simplify the governing equations, their
models have either ignored the convective accelerations due to downstream-varying
depth and curvature, treated them as perturbations about a lowest-order solution
that neglects these effects, or employed mathematically inconsistent expansions
based on concepts valid only under special conditions,. such as fully-developed bend
flow. Models of this nature will only be accurate when momentum fluxes associated
with bed topography and changes in radius of curvature are much smaller than the
pressure gradient and bottom stress terms in the vertically-averaged downstream
momentum equation. However, it has been clearly demonstrated, first by Yen and
Yen [1971] in a laboratory flume and later by Dietrich and Smith r1983] in a natural
meandering stream, that the vertically-integrated convective accelerations described
above are of the same order of magnitude as the accelerations associated with the
pressure gradient and shear stress divergence in meandering channels with typical
bar-pool topography. It is clear from these field and laboratory observations that a
valid model of flow in curved channels with naturally-occurring bed topography
must include these terms at lowest order.
Smith and McLean [1984] developed a model for flow in a meandering stream
based on a regular perturbation expansion about a zero-order solution that included
the convective acceleration terms described above. In order to demonstrate the
validity of their model, Smith and McLean compared predicted bottom shear stress
and water surface topography to values measured by Hooke [19751 in a curved
laboratory channel. The good agreement between predictions of the model and
Hooke's data indicate that the salient features of the flow are well described by their
approach. This is encouraging because Hooke's flume had bottom topography and
curvature characteristics similar to those of a natural meander bend. The centerline
of Hooke's flume was described by a sine-generated curve, which Langbein and
Leopold [1966] identified as a trace that closely defines the centerline of many
naturally-occurring meanders.
Furthermore, the bed of the flume was in
equilibrium with the flow and displayed the bar-pool structure typical of natural
sediment-transporting streams. However, while Hooke's flume was similar to
natural meanders in these two respects, it also had features that were definitely not
representative of natural situations. For example, the flume had vertical sidewalls
and was of constant width, unlike natural meandering channels. In addition, the
experimental run Smith and McLean attempted to reproduce with their model was
one in which the bed was stabilized, so there was no sediment transport and there
were no bedforms on the channel bottom. Thus, while the comparison of model
results and measurements for this laboratory case is a necessary first step, it is clear
that a field-scale test of the model should be made under sediment-transporting
conditions before it can be employed with confidence in natural streams. Moreover,
this test, like the previous one, must be made without adjusting any flow or
sediment transport parameters.
In this chapter, a consistent scaling of an appropriately closed set of momentum
equations is shown to produce a model similar to that of Smith and McLean [1984].
This model is then extended to the case in which width varies slowly In the
downstream direction and depth smoothly approaches zero at the banks of the
stream. The technique presented here also allows for the presence of bedforms and
sediment transport. In order to test this expanded version of the Smith-McLean
model, we apply it to a bend (the so-called IMR bend) in Muddy Creek, a
sand-bedded meandering stream in Wyoming. This site was studied extensively by
Dietrich [1982], and the data from this location permits calculated boundary shear
stresses and other flow variables to be compared to values measured in a system
possessing the dominant characteristics of natural meandering channels. The
calculations and subsequent comparison with data from the study site clearly
indicate both the importance of including appropriate convective accelerations at
lowest order and the overall veracity of the model.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

71

Nelson and Slnith


Mathematical Formulation

Fundamental Equations
Since bottom slopes in streams and rivers are usually gentle and the time scale of
discharge variations is large compared to the time required for a parcel of water to
traverse a typical meander, it is proper to assume that the pressure distribution is
hydrostatic and that the flow is quasi-steady. Using these approximations, Smith
and McLean [1984] obtained equations expressing conservation of mass and
momentum in a "channel-fitted" orthogonal curvilinear coordinate system.
Defining s, n, and z as the coordinates in the downstream, cross-stream, and
vertical directions, respectively, and defining R as the centerline radius of curvature,
these equations are as follows:

(1)
u

1 - N

au

Du

+ 1:.

_ ...::::K- DE

- 1 - N os
u

bv

1 - N

uv

(1 - N)R

[_1_

DTss

P 1 - N (JS

bv

bv

os + v on + w oz + (1

__ g DE

Du

os + v on + w 7JZ -

on

+ 1:.

+ DTns + DTzs _

on

OZ

2(1 -

N)R

(2)

u2
- N)R

[_1_ os
DTns + DTnn + DT zn + Tss on on (1 -

p 1-

Tns]

O=-.!.~-g
plFi

Tn n]

N)R

(3)

(4)

where u, v, and ware the velocities in the downstream, cross-stream, and vertical
directions, respectively, and where E and B are the elevations of the surface and
bottom. The quantity 1-N=(1-n/R) is simply the downstream metric associated
with the channel coordinate system, which was described formally by Smith and
McLean [1984] and used extensively by Dietrich [1982] and Dietrich and Smith
[1983].
.
In the above equations, Tss, Tns, Tzs, Tnn, Tzn, and Tzz are the independent
components of the deviatoric stress tensor. Rivers and streams are fully developed
turbulent boundary layers, so viscous stresses are negligible in comparison with
Reynold's stresses and can be ignored. If the existence of a scalar kinematic eddy
viscosity is assumed, the Reynold's stresses may be expressed in terms of the rate of
strain tensor. In the given coordinate system, this relationship results in the
following equations for the stress components:

Tss=2pK [1

~ N~-{1 ~ N)R]

Tnn = 2pK

[~]

Copyright American Geophysical Union

(5a)
(5b)

Water Resources Monograph

River Meandering

Vol. 12

Flow in Meandering Channels

72
T zz

= 2pK

[tJ

(5c)
(5d)

Tzn

= pK [~+ ~J

[1

Tns = pK 1 - N

IJv
os
+ (1

(5e)

- N)R +

On]
on

(5f)

In order to calculate the boundary shear stress, it is necessary to average the


continuity and momentum equations vertically by integrating them from the
channel bottom, z = B(s,n), to the water surface, z = E(s,n). Integral conservation
of mass is expressed by

)
<v>h
0 (
)
1 0 (
1 - N os <u>h - (1 - N)R + on <v>h

=0

(6)

where < > has been used to indicate vertically-averaged variables. Treating this
equation as an ordinary differential equation for <v>h and noting that this
quantity must equal zero at the banks of the stream, one finds the solution

<v>h=l-~N

-1.1
2"

~u>h)dn

(7)

where W denotes the stream width.


Vertical integration of the horizontal momentum equations yields the following
expressions:

) 2<uv>h
1 0 ( 2 ) 0 (
1 - NOS <u >h +on <uv>h -(1 - N)R

1[1
D
0
1 - NOS Tss>h) + on Tns>h)

= ~OE
1 - Nos + P

< T ns> h
1 ( ) OB ( ) OB ( )]
-2(1 -N)R+l - N Tss BOS + TnsBOfi- Tzs B

(8)

and

_1_0
1 - N OS

= - gh

uv

>h)+O 2>h)+u 2 > - <v 2 h


on v
( 1 - N) R

OE 1[ 1 D
0
on
+ Ii 1 - NOS Tns>h) + on Tnn>h)
<Tss - Tnn>h
(1 - N)R

+1

os

1 ( ) DB (
) OB (
)]
- N Tns B
+ Tnn B on - Tzn B

(9)

Equations (1) through (9) form the mathematical basis of the meander flow
model developed in this paper. The complete formulation of the model can be

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

73

Nelson and Smith

divided into two distinct parts and, for the sake of clarity, this approach will be
taken here. The two parts of the model are as follows:
(a) Using assumptions based on dimensional arguments, physical reasoning, and
field observations, Smith and McLean [1984] transformed (7), (8), and (9)
into a single integro-differential equation for (Tzs)B' the downstream
bottom shear stress, which can easily be solved numerically. This equation
may be arrived at directly from a consistent scaling of the verticallyaveraged momentum equations. In this paper, the scaling used to obtain
this equation is presented and then a simple and efficient numerical
technique for its solution is described. As in the original model of Smith
and McLean [1984], this solution to the vertically-averaged equations
includes effects due to downstream-varying radius of curvature and bottom
topography.
(b) A regular perturbation expansion is constructed based on the correct
ordering of terms in the full horizontal momentum equations, (2) and (3).
The lowest order of this expansion corresponds to the solution of the
vertically-averaged equations found in (a), Le., only the vertically-averaged
convective acceleration terms are retained at lowest order in the
downstream momentum equation. Using this expansion, we are able to
calculate approximate solutions for the complete (three-dimensional) shear
stress and velocity fields, and the full surface elevation field.

Scaling and Numerical Solution of the Vertically-Averaged Equations


In order to scale the equations it is convenient to introduce nondimensional
variables, denoted below with carets. We define mo, W 0, h o, and R o as length
scales typical of meander length, width, depth, and minimum radius of curvature,
respectively. In addition, Uo is used to represent the streamwise scale velocity, 10 is
used to scale changes in the surface elevation, and To is a typical bottom stress.
Employing these scales, the equations may be nondimensionalized using the
following definitions:

l!..2
Uov
Wo

w=

l!..2
Vow
mo

u = Uoll

v=

s = mos

n=WoD

z = hoz

R=RoR

E = 10E

h=hoh

B = hoB

Tij

ToTij

Fr =

Vo
(gh 0 )1/2

(10)

Inserting these variables into the vertically-averaged equations and dropping the
carets for convenience yields

[_1 0 U2>h)] + [l!..2]2 [0on uv>h)] _ [l!..2] [h o] [2

h o]
[mo
1 - N os

= Fr- 2

W0

W0

Ro

o][_1 0 Tss>h)]
[h][~OE]
+ [To][[h
mo 1 - N os
~
mo 1 - N os

Copyright American Geophysical Union

<uv>h ]
1 - N)R

Water Resources Monograph

River Meandering

Vol. 12

Flow in Meandering Channels

74

+ l.2
W0
+

[0on

TnS

ns>h ]
>h)] - R
h o [2<T
o ( 1 - N) R

~~ [1 ~ N(Tss)B ~] + ~~ [(Tns)B~] - (T~)B]

(11 )

[~~][~~][1 ~ N~uv>h)] + [~~r[kV2>h)]


+

[~~] [(iU~>~)R] - [~~] [~~r [(iV~>~)R]

= - Fr-

2[~~][h~] + ~J[~~ [~Tns>h)] + ~~ [k Tnn>h)]

+ Rh o [<Tss
(1
o

- Tnn>h]
- N)R

+~~ [(Tnn)B~]

+.h
[_1_ (Tns) OB]
mo 1 - N
B os

-(Tzn)B]

(12)

The nondimensional numbers appearing in these equations may be ordered with


respect to each other in order to determine the relative importance of the terms in
the equations. The first of these numbers is the depth-to-width aspect ratio,
ho/W o. In natural streams this ratio is typically a minimum of 0(10- 1 ), which will
be denoted as O( (). The ratio of average depth to minimum radius of curvature,
ho/R o, is also O( () because R o is of the same 'order as the width in well-developed
meanders. However, the ratio of ho to mo, the meander length, is 0((2), because the
lengths of meanders are typically an order of magnitude greater than their widths.
These three nondimensional numbers describe only the geometry of the meander.
The dynamical balance is described by the Froude number, the ratio of the
crossing-to-crossing drop in surface elevation to the meander length (lo/mo), and
the drag coefficient (T 0/ PU~). Since the drag coefficient in geophysical scale flows is
generally of 0(10- 2), this last nondimensional number is O( (2). The combination
Fr- 2 (lo/m o) must be of lowest order in the downstream equation because gravity is
the driving force for the flow. Similarly, Fr- 2 (lo/W o) must be of lowest order in
the cross-stream equation in order for the cross-stream pressure gradient to balance
the centrifugal force term. As a consequence of this scaling, the lowest-order
balances in the dimensional vertically-averaged equations are given by
1

1 -

0 (<u 2 > h) +on


0 (<uv>h ) -2 (1<uv>
h
Nos
- N)R
_~OE 1( )
- 1 - N os - P Tzs B

(13)

and
<u 2 >h _ h OE
( 1 - N)R - -g on

Copyright American Geophysical Union

(14)

Water Resources Monograph

River Meandering

Vol. 12

75

Nelson and Smith

This scaling supports the observation that topographically-induced convective


accelerations must be retained at lowest order in the downstream momentum
equation in order to describe flow in natural streams. Smith and McLean [19841
used a boundary layer expansion rather than the interior flow expansion employed
here and, as a consequence, they had to use observations of flow in meander bends
to justify the ordering of their terms. The more consistent scaling presented here
demonstrates that their ordering of terms was correct.
Using the observation that <uv> ~ <u><v> and the simple closure assumption
p<u 2> ~ p<U>2 = 0'( Tzs)B' where a is the reciprocal of the drag coefficient, Smith

and McLean [1984] transformed equations (7), (13), and (14) into a single_
integro-differential equation for (Tzs)B:

(0'( TzS )B )

- (1 -

1/2

N) Rh

fn IJ
-w

os (o'( T ZS) Bh 2 )

1/2

(15)

dn

2"

where E c is the surface elevation at the stream centerline.


The solution of (15) is subject to the constraint that downstream discharge must
be held constant. In order to apply this constraint, they defined a complete velocity
field as follows:
(16)
where ( = (z-B)/h and (0 = zo/h, and where Zo is the overall bottom roughness
parameter. The Zo used in this equation must represent all of the resistance to flow
in the bend, including grain roughness, sediment transport effects, and form drag
due to bedforms and bars. This parameterization is based on the fact that
ultimately all of these effects act at or near the bed, so that the various
momentum-removing processes extract momentum from the flow, in the same
manner as grain roughness alone. Our method for calculating the overall z is
presented in a subsequent section of this paper. In conjunction with the closure
assumption, (16) yields

[J f1((,(o)d(r

(17)

(0

Using the separable velocity profile of (16) yields a downstream flux of

2'"

-W

(0

f hf

2"

u d( dn

2'"

u*O'1 / 2 h dn

-W

2"

Copyright American Geophysical Union

(18)

Water Resources Monograph

River Meandering

Vol. 12

Flow in Meandering Channels

76

Equation (15) is solved by bringing the integral terms in iteratively as


inhomogeneous terms in a simple first-order linear differential equation. In other
words, we initially solve (15) assuming the integral terms are zero (thus reducing it
to a simple ordinary differential equation), then use the solution obtained thereby to
calculate the value of the integrals involving (Tzs)B on the right-hand side. These
values are then inserted into the equation as additional inhomogeneous terms and
the equation is again solved, yielding an improved estimate for the bottom stress
field. This procedure may be repeated until the solution for (Tzs)B converges,
usually about three or four times. Thus, the complicated integro-differential
equation is exchanged for a series of first-order inhomogeneous ordinary differential
equations of the form
(19)
where

'Yl

= 1ah

N + ~(ln(l - N))

DEc

12 -

-pg OS +

fN

1 D( 01 TzsI B )
1 - N
Os dN

Although the solution of (19) can be written down immediately in terms of integrals
which may be evaluated numerically [see Smith and McLean, 1984], it is simpler and
more efficient computationally to solve it directly using a finite difference
approximation.
Assuming that values for the downstream bottom stress are known at the
entrance to the bend, the finite difference form of (19) will yield the values of the
bottom stress at the gridpoints immediately downstream, provided that some value
for the centerline surface slope is prescribed. Of course, these values will not satisfy
the flux integral above, unless the initial guess for the centerline slope is an
extremely fortunate one. However, one may employ the calculated discharge in
combination with the desired discharge to improve the value of the centerline
surface slope by using a shooting technique. In other words, the finite difference
equation coupled with the flux inte~ral yields discharge as a function of centerline
surface slope at each cross-section (s = constant). Therefore, as long as the true
river discharge is known, the solution is completely specified. The simple shooting
method employed here generally converges on the correct slope in less than five
iterations. Once this procedure is followed at each section, the differential equation
is solved, so one may calculate the integral terms from the solution, reformulate the
differential equation with improved values for the inhomogeneous terms and repeat
the procedure, again stepping, downstream through the bend shooting for the
centerline slope values. This process is repeated until the entire field of (Tzs)B
values has converged.
The procedure described above requires that the values of the boundary shear
stress be known at the upstream section of the numerical grid. Unfortunately, due

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

77

Nelson and Smith

to the difficulty involved in measuring bottom stress, this upstream boundary


condition is usually unknown or only very roughly determined. However, the
equations are not very sensitive to the upstream boundary condition, and the
solution is essentially independent of the initial prescribed values beyond about one
meander length downstream, provided that the initial values are even remotely
reasonable and yield the correct value for the discharge. Therefore, in order to find
the solution in a given bend, it is only necessary to know the geometry and
topography of the given bend and the bend immediately upstream. Then, even if
the initial upstream values of (Tzs)B are inaccurate, the stress values calculated in
the bend of interest (and any subsequent bend) will be correct. If the upstream
bend is more or less similar in geometry to the bend of interest, it is sufficient to
perform the calculation through a few identical bends, each of which is the known
bend or its reflection about a line through the crossings.
This numerical procedure is well-posed for the case of constant width, in which
the banks of the stream correspond to lines of constant n. However, if the width
varies, as it does in most natural streams, the method by which the solution is
marched downstream along lines of constant n will no longer determine the solution
in the entire domain of interest. As long as the variations in the width are not
abrupt, however, this problem can be avoided by linearly straining the numerical
grid used so that lines of constant n correspond to the banks of the stream. The
calculation is then performed exactly as described above, except that the discharge
integral is calculated across widths which vary with downstream position. This
procedure is only valid if width variations are such that the cross-stream advection
of downstream momentum due to the width variations is small and can be
neglected, which is the same as saying that taking Dj Os outside the integral in the
numerical evaluation of the integral in (7) where it appears in the downstream
momentum equation will cause an error no larger than those associated with terms
already dropped from the equation as a consequence of the scaling arguments. Since
the limits of integration for this integral will be slowly varying functions of s in the
strained coordinate system, Liebnitz' rule may be used to show that the error
incurred by this procedure is of higher order than the rest of the terms in the
downstream equation providing that width variations over the length of the
meander are an order of magnitude smaller than the average width. If this
condition is not met, width-induced convective accelerations must be included at
lowest order, and a more complicated numerical procedure must be used. It is
important to note that the basic equations of the model are valid even for the case
of abrupt width changes; there is a large class of problems, however, for which width
is nearly constant and these are addressed here, chiefly for the sake of numerical
simplicity. For streams with complex geometries, the full set of convective
accelerations may be retained in ther vertically-averaged momentum equations. In
most situations, this may still be solved iteratively as described above, or the
equation may be solved with a standard two-dimensional numerical technique.

The Non-Averaged Equations


The solution of the vertically-averaged equations presented above yields the
downstream boundary shear stress field, the vertically-averaged velocity field, and
the centerline and cross-stream surface slopes, all based on the specification of a
vertical structure function for the downstream velocity. In order to calculate the
cross-stream bottom stress and velocity, as well as the error associated with the
assumption of a vertically similar downstream velocity profile at each point in the
stream (Le., that u = u* f( (,(0)), one must return to the non-averaged horizontal
momentum equations, (2) and (3). To scale these equations correctly, it is
necessary to examine the nature of vertical profiles of horizontal velocity in typical
streams and rivers.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Flow in Meandering Channels

78

Observations of velocities in channels with a wide variety of geometries indicate


that, except in a small region near the bottom, the deviation of the downstream
velocity from its vertical average is small compared to that average. For example,
in a logarithmic velocity profile it is easy to show that the deviations from the
vertical average are of order 10- 1 times the vertical average above a height d, where
d is of order 10- 1 times the flow depth. This is the same as assuming u*1 <u> ~
0(10- 1) or, equivalently, that the square root of the drag coefficient is 0(10- 1),
which is true in most boundary layer flows.
In scaling and expanding the full momentum equations, the assumption that u =
<u> + (u '(z) is employed, which is equivalent to assuming that deviations from
the vertical average are small. Using this reasoning and the definitions for the
geometric parameters as given above, it is appropriate to define the following set of
nondimensional variables:

= Uoy
s = mos

= Uou
E = 10E
z = hoz
= ku*z~( z) = (uohoK

= Uow
n = Woft

R=RoR

~
= ~
uZ
uZ

[<u>

+ (U/(z)]

~
au
110 az

(20)

Using these nondimensional quantities along with the estimated values for the
various aspect ratios yields the following equations:
_1_

1 - N

au2 + !. aUY + 1 auw _!. 2uy = --=L Fr- lJ! aE


as ( an {T az ( (1 _ N)R 1 - N { 2 mo as
2

+ L [K

au]
az

2
(1 -

,.

az

N)R

+(

[K av + Kli . + Kau]]
as (1 - N)R an

av

,. ay
u -- + v
(I - N
an

as

= _ Fr 2

[ILan Kavas + Laft [(1 -Ku N)R] + ILan Kauan + Laz KOwas

+ -1 w
,. ay
-+
(

az

+ O( (2)

u2
(1 - N) R

!..L aE I !L K av + (!L K fJW + O( (2 )


(Wo

aft

az az

(21)

az

an

(22)

As described above, the quantities Fr- 2 (101 (2m o) and Fr- 2 (101 (Wo) will be 0(1)
in natural streams because the pressure gradient terms provide the ultimate driving
force for the flow. Substitution of typical values also shows that these terms are
0(1), thus demonstrating the validity of the scaling used. The terms which are
multiplied by 11( and 1/ (2 will not produce terms of lower order than the pressure
gradient and stress because the cross-stream and vertical velocities are smaller
(higher order) than the downstream velocity, consistent with the aspect ratio scaling
used in the vertically-averaged equations. Before introducing the perturbation
expansion, the fact that deviations from the vertical average are small may be used
to separate the convective accelerations appearing at lowest order in the above

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

79

Nelson and Smith

equations into a vertically-averaged part and a remainder which will be of higher


order. For example, the first convective acceleration term in the downstream
momentum equation may be written as follows:

(23)
Using the scaling for the vertical profile of downstream velocity described above, it
is easy to show that the quantity in brackets on the right-hand side of the equation
above is small compared to the first term on the right-hand side or, in other words,
that convective accelerations not associated with topographic steering enter the
equations at one order of magnitude higher than those that are. This same scaling
justifies the afProximations <u 2> ~ <U>2 and <uv> ~ <u> <v> used by Smith
and McLean 1984] in the solution of the vertically averaged equations. Using
similar reasoning on the other lowest order convective acceleration terms in the
equation yields
__
1 _ o<u>2h +

h(l - N)

fh

as

o<u><v>h _! 2<u><v>
f (1 - N) R

ail

a ("
as

a (""

" ")

a ("")'
oz

"2) +1- - uv-<u><v> +1- - uw


+ f -1- - u2 -<u>
[1 - N
f an
f2

_ ~ [ <u>2 oh + !. <u><v> Oh] _ ~ (uv - <u>~v ]


h 1 - N
f
on f (1 - N) R

os

= _ _1_
1- N

oE + L K ou + f [L K ov + L [
os OZ oz
ail os an (1

+LKow_

oz as

2
A[K
(1 - N)R

OV +

as

KU
] +
- N)R

L K ou

oil on

KU .+Kau]]+O(f2)
(1 - N)R

ail

(24)

and
~ ov + v ov +!.w ov +

1- N

as

ail

oz

<U>2
+ f(U 2 - <u>2)
(1 - N)R
(1 - N) It

= _ DE + L K ov + (L K aw +

ail

OZ

OZ

oz an

O( (2 )

(25)

A regular perturbation expansion employing f as a small parameter is applied at


this point. The flow variables are expanded as follows:

V=VO+CVt+ f2V 2 ...

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Flow in Meandering Channels

80

(26)
where Vo = Wo = Wt = 0 from the simple aspect ratio scaling. The lowest--()rder
equations in dimensional form will be
........,..."._1~ b<uo>2h
h( 1 - N)
aS

+ 1 b<uo><vt>h

an

Ii

2<uo><vt>

- (1 -

N)R

=~~+~Ko~

(27)

and
<UO>2
_
bE o
( 1 - N)R - - g on

(28)

Consistency between (13) and (26) requires that


Tzs

= (Tzs)B (1 - z/h) = pu: (1 - () = PoK o ~

(29)

where (Tzs)B is the solution to the vertically-averaged equations found above. This
allows the assumed velocity profile to be related to the lowest--()rder eddy viscosity.
Using K o = u*h~( () in the above equation, where ~(() is the nondimensional eddy
viscosity, yields the following relationship between the lowest--()rder vertical velocity
profile and the eddy viscosity:

(30)
Thus, by assuming a vertical velocity profile (or alternatively, a nondimensional
eddy viscosity) solutions are found for the lowest--()rder downstream boundary shear
stress, velocity, and centerline and cross-stream surface slope, as detailed above in
the solution of the vertically-averaged equations.
The nondimensional eddy
viscosity used in the calculations performed to obtain the results presented here is
given by

~(()

= k((l _ ()1/2

(31 )

This eddy coefficient yields slightly greater vertical momentum transfer than an
eddy coefficient corresponding to a logarithmic velocity profile, which is given by
~(() =

k ((1 - ()

(32)

The eddy coefficient given by (31) is used because the enhanced near-surface
momentum transport it predicts is thought to represent the effect of boils. Smith
and McLean [1984] tried several eddy coefficients in their model of meandering
streams, and concluded that this choice was the most appropriate one.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

81

Nelson and Smith

It is important to note that althol!gh (27) and (28) were arrived at using scaling
that is only valid in the upper 90% of the flow depth or so, they yield the
appropriate value of the boundary shear stress if they are assumed to hold all the
way to the bottom. If the full horizontal momentum equations are scaled in the
lower part of the flow, where it is not reasonable to assume that deviations from the
vertical average are small, the lowest-order equation simply yields a constant stress
profile (Le., the variations in the stress are O( l) compared to the bottom stress,
because z/h<l). Thus, by extending the domain of the equations found to be valid
in the outer layer all the way to the boundary, some (but not necessarily all) higher
order terms are included in the solution near the boundary. The alternative to this
approach is to solve the appropriate equations in each domain and match them at
some intermediate level. Rather than use this more complicated approach, (27) and
(28) are applied throughout the flow depth. Moreover, the details of the flow in the
lowest layer will never be precisely modeled due to the presence of sediment
transport, ripples, and dunes, the effects of which are included here only
parametrically, through the specification of zoo
In order to calculate the lowest order cross-stream velocity and bottom stress, it
is necessary to resort to the O( l) equations, which are given by

b(l - N) 7JS uo><ut>h)


_ 2 <Ut><Vt>
(1-

0 Ut><Vt> + <Uo><V2> h)
+ li1 on

+ <Uo><V2> + _1_0 (u2 - <U >2)


1 - N

N)R

os

a (UOVI - <uo><Vt + (JZ


D (UoW2)
+ on
_ 1 [<uo>2 Dh
li 1 - N

>< > Dh]


os + < Uo.
VI
on

__ ~ DEI

1 - N os

+D

(JZ

oUo +

7JZ

_
2
2K ouo
(1 _ N)2R2 (1 -

_ (UoVI - <Uo><VI

(1 -

aK

OZ

N)R

OUI

lIZ

N) R

+ K ou o]
on [(1 KoU
- N)R
on

+a

Duo

(33)

on

and
2<UO><UI > (u~ - <uo>2) _
(1 - N ) R + (1 - N) R
-

aEI a
g on + (JZ

Ovl

K o lIZ

(34)

The cross-stream equation may immediately be integrated twice with respect to z,


since uo(z) is known. Using the boundary conditions that the surface is unstressed
and v = 0 at the bottom and Ko = u*hK( (), this procedure yields

-L
[ -u;h f (/") u* (1 - N)R 2 ~

VI -

[ h a(EI
g

an+ Eo) + 2h<uo><u


(1 - N) RI>] fI (/")]
~

where

Copyright American Geophysical Union

(35)

Water Resources Monograph

River Meandering

Vol. 12

Flo\v in Meandering Channels

82

M() =

(0

IW d(d('

('

Since the value of <VI> is known from (7), (35) may be integrated to set the value
of the terms multiplying f1 in (35). This yields

- g

D(E o

an+ El) - 2<uo><u


(1 - N) R1 >
(36)

where we have defined


F2

ff
1

d(

(0
Substituting the above solution back into (35) yields the cross-stream velocity and
stress, given by

VI

(1

u~h N)R [(X72 1 - 2] - [(1 _~)hall2

-t ~

(u*hal12 )dn] 1

(37)

2"
and

pu*(l -

f) fn osD (u*h a1/2)dn

- (1 - N)ha /2

(38)

-W

2"

Thus, without resorting to the O( t) downstream momentum equation, it is possible


to calculate lowest-<>rder solutions for all flow variables. The O( t) downstream
equation will yield corrections to these IC\"'\vest-<>rder solutions. Although this
equation may be solved in a relatively straightforward manner, it is worth
examining the veracity of the lowest-<>rder solution before adding this complication.
The higher-<>rder corrections brought in through the solution of (33) are principally
associated with the redistribution of downstream momentum by the helical part of
the cross-stream velocity, the changes in vertical structure associated with local
accelerations in the flow, and the effects of lateral friction.

Specification of the Roughness Parameter and Form Drag Partitioning


In order to specify f t ( (,(0), the nondimensional velocity profile, it is necessary to
determine a value for zo, the roughness parameter. In simple flows where the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

83

Nelson and Smith

effective roughness of the bottom is only due to the actual geometric roughness, Zo
may be determined easily using the experimental results of Nikuradse as presented,
for example, by Schlicting [1979]. However, this technique is rarely suitable for
natural streams, due to the presence of channel nonuniformity, bedforms, and
sediment transport. Each of these tends to increase the effective roughness of the
bottom acting on the flow outside the region where they have a direct dynamic
influence on the flow. Thus, a method is required whereby the total effective
roughness of the channel can be accounted for in terms of an overall Z00 The
boundary shear stress resulting from the calculations using these values for the
roughness will include the actual skin friction shear stress as well as the momentum
losses due to pressure distributions on bed and bank irregularities (e.g., ripples,
dune, bars, slump blocks, channel constrictions, and large-scale plant debris).
Sediment transport depends upon the skin friction shear stress, and since one of the
principal goals of this work is to make accurate predictions of sediment fluxes,
clearly a general method for reducing the overall boundary shear stress to the skin
friction value is necessary.
To treat both the determination of roughness lengths and the form drag
partitioning, a generalized version of the technique presented by Smith and McLean
l1977] is employed. Following their formulation, the total boundary shear stress is
written as a sum of skin friction and form drag components:

(39)
where T B is the magnitude of the total boundary shear stress vector, T SF is the skin
friction bottom stress (the stress acting to move sediment on the bed), T D is the
boundary shear stress equivalent to the dune or ripple form drag, and T CH
represents the form drag associated with all other topographic elements of the
channel. The general relationship between form drag and an equivalent bottom
stress may be written as follows:
1

T=

2 AX

2 PC n Ur A

(40)

In this expression, p is the fluid density, CD is a drag coefficient which must be set
empirically, Uris an appropriate reference velocity, Ax is the cross-sectional area. of
the obstacle perpendicular to the principal flow direction, and AB is the area of the
bed covered by the obstacle. The reference velocity is usually defined as the average
over Ax of the velocity that would exist if the obstacle were not present in the flow
(the undisturbed velocity). The reference velocity for a given obstacle is defined
here to be the vertical average of the unperturbed velocity over a two-dimensional
obstruction with cross-sectional area equivalent to the original obstacle. For the
case of a two-dimensional dune, this definition of the reference velocity reduces to
the average of the undisturbed velocity over the cross-section of the dune, as
expected. In contrast, for the case of a three-dimensional obstruction, the reference
velocity is taken to be the vertical average of the undisturbed velocity over an
equivalent two-dimensional obstruction. For example, in the case of a channel
constriction, the reference velocity is taken to be the vertical average of the
undisturbed velocity over a two-dimensional bump across the channel bottom. This
is done because the primary response to all obstructions is at the perimeter of the
channel, and the velocity varies logarithmically in a direction perpendicular to the
boundary. This height of this bump is chosen such that the cross-sectional area of
the bump is equivalent to that of the original constriction. This approach, which is

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Flow in Meandering Channels

84

consistent with the inclusion of the form drag effects through a total boundary shear
stress and roughness length, permits the use of empirically determined drag
coefficients for two-dimensional obstacles to be employed in all cases.
Using data measured over two-dimensional sand waves in the Columbia River,
Smith and McLean [1977] found CD = .21 for the case of separated flow over dunes,
and CD = .84 for unseparated flow. Although these values were originally
computed from data using a slightly different formulation for the reference velocity,
subsequent work has shown that the approach described herein is essentially
equivalent, and their values for the two-dimensional drag coefficients are
appropriate. Clearly, more experimental verification of these values is in order,
especially for the unseparated case, but these values are employed in the
calculations presented here.
For the case of two-dimensional bedforms in a turbulent flow, the unperturbed
velocity is simply a logari thmic profile characterized by the skin friction shear stress
and roughness. This follows from the fact that, if the bedform were not present in
the flow, the near-bed velocity profile would be given by

u=

1/2

~ In [(ZO)SF]
SF

(41)

where (zo)SF is the actual grain roughness if the sediment is immobile, and is a
roughness length proportional to the height of the saltation layer if bedload
transport is occurring. A semi-theoretical method for calculating the thickness of
the bedload layer and an experimental determination of the relation between that
thickness and the effective roughness length (see (50) below) has been provided by
Dietrich [1982]. More recently, the expression obtained by Dietrich has been
rederived theoretically by Wiberg [19871 and has been employed by Wiberg and
Rubin [in press] for upper and lower prane beds. Averaging (41) over the dune
height in order to obtain the reference velocity and noting that Ax = HDb and
AB = Ab, where HD, A, and b are the height, wavelength, and cross-stream width of
the bedform, from (40) the following expression is found for the form drag associated
with bedforms:

(42)
Above the region in which the dunes have a substantial effect on the spatial
structure of the flow, the velocity profile will be quasi-logarithmic. This outer
profile will be characterized by a shear stress equal to the sum of the skin friction
value and the dune form drag, along with a roughness length which describes the
extraction of momentum from the flow by both skin friction and form drag. This
roughness length is determined by matching the velocity profile given by equation
(41 J with the outer velocity profile at the height of the obstacle. Thus, one obtains

r~11/2 [H]
TO 112 In[ H ]
D
(ZO)SF =----...--- ~

~ln

Copyright American Geophysical Union

(43)

Water Resources Monograph

River Meandering

Vol. 12

85

Nelson and Smith


which leads to the following expression for (zo)n:

(44)
where

7 =[TS~S; Tof =[1 + ~~[ln(z7)SF) -Irr


0

12

12

(45)

Equations (42) and (44) provide the relationships necessary to reduce the total
boundary shear stress to the skin friction shear stress and to compute the total
effective roughness of a channel, respectively, provided that bedforms are the only
source of nonuniformity. This is rarely the case in natural channels, where form
drag associated with bars and bank variations typically is a significant proportion of
the total channel drag. To account for this effect, the form drag on each channelscale nonuniformity must be calculated using (40). The drag coefficient must either
be chosen as that of an equivalent two-dimensional obstruction, as described above,
or it must be found by analogy with laboratory results obtained for certain special
geometries. For example, in the case of a log obstructing the flow, well-known
classicai results for the drag coefficient of a circular cylinder at various angles of
attack may be employed. It is important to note that, by extracting the form drag
of channel-scale irregularities in this manner, this technique explicitly accounts for
the total drag of the channel. No attempt is made to distribute the drag force over
the channel bed, which would require a much more complicated approach.
In the case of a meandering stream with smooth banks and a nearly constant
cross-sectional area, the channel form drag is essentially that associated with the
point bar. The reference velocity for the channel-scale features producing form drag
is the outer velocity profile over the bedforms, as described above, averaged
vertically over the cross-section of a two-dimensional bump with the cross-sectional
area of the point bar. If Hb and .Ab are the height and wavelength of the point bar
and the cross-sectional shape of the point bar is approximated as a triangular
wedge, the bottom stress equivalent to the form drag of the point bar is given by

(46)
Because the flow typically does not separate over point bars, Cn = .84 is employed.
This expression allows the channel form drag to be related to the skin friction form
drag. To calculate the field of overall roughness lengths in the channel, the process
followed in obtaining (44) is used, with one minor difference. Rather than
matching at the height of the point bar or the height of the equivalent
two-dimensional obstruction, the matching is enforced at the water surface. This is
to avoid matching at points above the water surface, which would typically occur
over much of the meander bend if the height of the point bar was used. Physically,
this matching enforces the fact that the free surface inhibits the growth of wakes
associated with large scale obstructions. The matching yields

ZO

h ] 1CH
= h [ (zo}o

Copyright American Geophysical Union

(47)

Water Resources Monograph

River Meandering

86

Vol. 12

Flow in Meandering Channels

where

CB =

CB
[T T:

T
y

TSy
D

12

[1 + ~~2 ~~B [In[ 2(~~)D] _1]2

r
12

(48)

Equation (47) predicts values of the total roughness length at various locations in a
meander bend. This effective roughness includes the form drag of both the dunes or
ripples and the point bar. In cases where other types of nonuniformities are
important, expressions similar to (46) must be developed using geometrical
arguments. The resulting expressions for the various types of channel form drag are
summed, and (47) and (48) are employed to calculate the field of roughness lengths
in the channel.
An important by-product of the determination of the overall roughness is an
expression relating the overall boundary shear stress to the skin friction value.
From (45) and (48), we find
T

B_

T-'n'CH
SF

(49)

This equation allows the boundary shear stress values predicted by the numerical
model described above to be reduced to the value of boundary shear stress
responsible for the sediment transport on the bed. This value may be used in any of
various bedload equations to predict sediment fluxes. Although the procedure
described here may seem complicated, the mathematical structure of (44) through
(48) is such that many simple approximations may be made allowing computations
of the overall roughness and form drag partitioning even in cases where the dune
and point bar geometry are only very roughly known. The quantities in brackets in
each of these equations tend to vary only weakly in typical natural streams. Thus,
(47) is nearly equivalent to holding (0 constant. Holding (0 constant is equivalent
to using a single value of the drag coefficient for the entire reach, a technique that
has often been used without theoretical support.
Other simplifications are
addressed below in the context of a specific meander bend, but the full approach
described here is of general validity, and is applicable to a wide range of bed and
bank geometries.

The Study Site


Muddy Creek is a sand-bedded stream in western Wyoming characterized by the
strongly meandering structure typical of many natural streams, as shown in Figure
1. Between 1976 and 1982, a comprehensive field investigation was undertaken in
one of the meander bends of this stream, the so-called IMR bend. Some of the
results of this study have been presented by Dietrich [1982], and by Dietrich and
Smith r1983, 1984]. As a consequence of this work, a comprehensive set of data
exists tor this site, consisting of careful measurement of the stream geometry, as
well as detailed measurements of velocity, surface elevation and sediment transport
rates; this makes the Muddy Creek site a perfect test case for the model described
above.
The values of the mean depth, width, and meander length in the bend of interest
are about .40, 5.0, and 25 meters, respectively. These values yield a depth-to-width
aspect ratio of 0.08, which is 0(10- 1), and a depth-to-meander length ratio of 0.016,
which is 0(10- 2). A tlpical value {or the discharge is 1.0 m 3 Is, so the value of the
Froude number Q/Wlgh3 )1/2 will be 0.25. The crossing-to-crossing drop in the
surface elevation in the bend in Muddy Creek was about 4.0 cm, so the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Nelson and Smith

Fig. 1.

Vol. 12

87

Aerial photograph of the Muddy Creek study site taken in 1979. The bend with the
railillg along the outer bank is the IMR study bend.

nondimensional number Fr- 2 (lojmo) is 0(10- 2). These observations clearly


demonstrate that the scaling employed is valid in this typical natural meander.
The topography of the IMR bend is characteristic of stream meanders, with
fairly flat areas near the crossings (where R approaches infinity) and well-developed
bar-pool structure in the region of the minimum radius of curvature. A contour
map of the IMR bend is shown in Figure 2. The topography was measured by
making many cross-stream transects and averaging them together to remove the
effect of bedforms migrating through the section of interest, as described by Dietrich
[1982]. This procedure was repeated at enough downstream positions to ensure that
the bathymetry was well resolved.
The radius of curvature of the centerline was calculated from the channel
centerline digitized from maps constructed from surveys of the channel banks. The
resulting radius of curvature as a function of downstream position, shown in Figure
3, is quite close to a sine-generated curve.
Unfortunately, the topography and radius of curvature of the meanders
upstream of the bend were not measured in detail, so the upstream conditions were
not well specified. However, the bend immediately upstream of the IMR bend is
quite similar to the IMRbend in form and so it has been replaced with an image of
the study bend reflected about a line through the crossings. The justification for
this approach is described above in the discussion of the numerical solution. It is
also justified by the fact that the cross-sectional depth profile of the upstream
crossing is nearly identical to that of the downstream crossing reflected about the
centerline, as shown in Figure 4. Since the topography and the flow are in
equilibrium, and the topography and stream geometry are nearly identical at the
two crossings, it is reasonable to assume that the boundary stress fields at the two
crossings are similar and, therefore, that the solution at the downstream crossing

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Flo\v in Meandering Channels

88

DEPTH (eM)

Fig. 2.

Depth contours in the IMR meander bend in Muddy Creek after the removal of
mobile bedforms by time averaging. Depth contours are shown at 20 centimeter
intervals.

provides a good initial condition at the upstream one. Furthermore, the results of
the calculations support the validity of this approximation.
Results of the Model

Boundary Shear Stress and Sediment Transport


The magnitude and direction of the overall boundary shear stress calculated
using the model are shown in Figure 5. The highly nonuniform structure is
dominated by the presence of a jet-like region of bottom stress which traverses the
meander, following a path which is near the bar-side bank in the upstream part of
the meander, crosses the stream in the central region, and exits the meander along
the pool-side bank. This structure is typical of meandering streams, and is clearly
present in the data presented by Yen and Yen [19711, Hooke [1975], and Dietrich
[19831. Measurements of sediment transport published by Dietrich and Smith [1984]
are arso indicative of this type of structure in the boundary shear stress distribution.
Simple models of meandering streams that neglect streamwise variations in
topography and curvature predict cross-stream boundary shear stresses that are
everywhere directed toward the inner or bar-side bank. This conclusion has been
used extensively in theories of point bar stability, in which inward shear stresses are
often balanced against an outward (Le., down the face of the point bar)
gravitational force. However, as is clear from our calculations and from the
observations of Dietrich [1982], Yen f1967], and Onishi [1972], the boundary shear
stress over the point bar is actually directed downstream or even slightly outward.
This is easily explained by the presence of topographic steering, as discussed in
detail by Dietrich and Smith [19831. As the flow approaches the point bar, shoaling
of the bottom and deceleration of the fluid both act to produce a large negative
downstream convective acceleration. As can be seen from (13), this tends to reduce
the downstream surface slope, which in turn causes a relaxation of the cross-stream
surface slope. The result is an excess of centrifugal force over the bar that forces the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

89

Nelson and Smith


50.0

en
a::

~ 40.0
~

a::
=>

t-

30.0

=>

I.L.

en
=>

a:::

20.0

-J

a::

10.0

o 000000 00

00

OL---""'--~-.I._"""'_L.--""'-_....L.-..-..l._"""""--J

Fig. 3.

.20

.40

SIMa

.60

.80

1.0

Centerline radius of curvature for the study bend in Muddy Creek.

flow outward throughout the entire water depth. This clearly demonstrates the
importance of using a model that includes topographic steering terms in the
lowest-order equations. The magnitude of the downstream nonuniformity also
requires that the full equation for sediment mass conservation be employed in
calculations of point bar stability, rather than an equation expressing some
cross-stream balance.
It is important to note that the boundary shear stress values shown in Figure 5
include the form drag effects of bedforms and meander geometry. To determine
values of the boundary shear stress responsible for sediment transport, it is
necessary to remove these form drag effects using the model discussed above. In
applying this technique to the IMR bend, several simplifications were made. In
order to apply the form drag and overall roughness model described above, a value
for the skin friction roughness ((zo)SF) must be specified. In .Muddy Creek, the
value of this parameter is primarily related to the height of the saltation layer, since
most of the momentum extraction very near the bed is associated with grain
saltation. A theoretical model for the height of this layer has been presented by
Dietrich [1982], but the calculation of the thickness of the bedload transport layer
requires that the skin friction stress be known. However, the value of t5 , the
B

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Flow in Meandering Channels

90

240.0
r---

DISTANCE FROM CENTERLINE (CM)


120.0
0
-120.0

-240.0

-I~---r---~-____,r___-___r--~--r__-~____,n
---,-----,

20.0
A

O~OOO

-40.0
:::I:

t-

o..

6A>~~~~~

60.0

80.0
100.0
Fig. 4.

6 0
0

00006

~~~~~~~

o CROSS - STREAM DEPTH PROFILE


6

AT UPSTREAM CROSSING
CROSS-STREAM DEPTH PROFILE
DOWNSTREAM CROSSING

Cross-stream depth profiles at the upstream and downstream crossings of the study
bend in Muddy Creek. Note that one profile has been rotated about a vertical axis
through the channel centerline to facilitate comparison.

BOUNDARY SHEAR STRESS (DY/CM 2 )

"' ttrf1.

"'1"
1
" ' ',

,,""'~'"
"'tf-"'ft

"1,

'..~ '0,~I

""

'I,If "'/

: .~ )'~ "I. I,

~ ~ ~
~

Fig. 5.

'I'

Magnitude and direction of the overall boundary shear stress in the study bend in
Muddy Creek is shown. Note the presence of slightly poolward shear stresses over
the point bar. The maximum amplitude of about 90 dynes/cm 2 agrees quite well
with the maximum of 85 dynes/cm 2 measured by Dietrich [1982] near the same
location.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

91

Nelson and Smith

thickness of the saltation layer, is usually between one and three grain diameters,
depending on the ratio of the skin friction shear stress to the critical shear stress.
Postulating a linear relationship between b'B and (zo)SF based on dimensional
arguments, Dietrich empirically found a coefficient of proportionality of 0.08.
Subsequent theoretical work by Wiberg [1987] and further measurements by Dietrich
[personal communication] indicate that this value is about 0.10. Since (42) and (44)
depend only weakly on the value of (zo)SF' and the ratio between the saltation layer
height and the grain diameter D only varies slowly in the IMR study bend, the
following approximation is employed:

(50)
This approximation removes the necessity of iteration of the form drag model, and
is valid if active bedload transport (Le., TB ) Tc) occurs over most of the bed, as is
the case in the IMR bend. If transport only occurs in isolated parts of a channel,
then the form drag calculation requires iteration. In practice, one begins by
guessing reasonable values for the skin friction roughnesses (e.g., the Nikuradse
values), and then calculates the ratio given by (49) using (42) through (48). This
ratio allows the skin friction boundary shear stresses to be calculated and inserted in
the expression for the bedload layer height provided by Dietrich [1982]. Thus, an
improved value of (zo)SF is obtained and the process is repeated. This iteration is
well-posed and converges rapidly. However, as discussed above, this process is not
always necessary, and nearly identical results are obtained in the case of the IMR
study bend using (50), which does not require knowledge of the skin friction stress in
calculating the skin friction roughness.
To apply (42) through (49), the heights and wavelengths of the dunes or ripples
present on the bed must be known. However, these quantities are rarely measured
in detail, so a simple closure for these values is desirable. For the case of fullydeveloped quasi-two-dimensional dunes, the heights of the dunes may be
approximated as a constant proportion of the water depth. Both experimental
[Yalin, 1977; Jones, 1968] and theoretical work (Nelson and Smith, in press] show
that dunes tend to grow to a height of about one SIxth to one fifth of the flow depth.
Although the dunes in the IMR bend are often oriented somewhat obliquely to the
streamwise flow direction rDietrich, 1982], they are quasi-two-dimensional in the
sense that the near-bed f10w tends to go over them, rather than being steered
around them as in the case of true three-dimensional dunes. Thus, in applying the
form drag model to the IMR bend, it is reasonable to use

h
HON 5:5
N

(51)

The wavelength of the dunes may be approximated either of two ways: the ripple
index may be assumed to be roughly constant, or the average dune wavelength in
the meander may be employed. In the IMR bend, observations of the dune
geometry indicates that the second of these assumptions may be more valid. This is
a result of the fact that the dunes often have a cross-stream extent reaching from
the deepest part of the pool to well up the flank of the point bar. Thus, the flow
depth changes by a factor of two or three, but the wavelength is nearly constant.
The dune heights decrease as one moves up the flank of the point bar in response to
the shoaling of the flow; however, since the wavelength varies only slightly, the
ripple indices (Hoi A) change substantially. In fact, all the form drag calculations
were performed using both of the assumptions described above for the wavelength of
the dunes, and only very minor differences in the skin friction bottom stress values

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Flow in Meandering Channels

92

22

Fig. 6.

Numbering system and location of sections at which measurements were made.

were noted. All results shown herein use the bend-averaged value of the dune
wavelength, which was 147 cm.
Dietrich's measurements of the boundary shear stress responsible for sediment
transport were made at six sections in the bed, as shown in Figure 6. The
numbering system of the sections is the scheme used by Dietrich, and is employed
here to allow easy comparison of the model results to Muddy Creek observations
published in other papers. The methods used to measure the boundary shear stress,
flow velocities, and surface elevations are described in detail by Dietrich [1982]. The
measured values of sediment transport boundary shear stress are shown with the
model results in Figure 7. It is clear from this figure that the model predicts the
structure of the sediment transport boundary stress fairly well, especially
considering the rough assumptions used in the reduction of the overall stress to the
sediment transport value. This, of course, indicates the desirability of this type of
model for calculations of bedload transport. However, while the predictions are
reasonably accurate, there are some systematic discrepancies. For example, at
section 18, the predicted skin friction bottom stress is low near the inner bank and
high near the outer bank. Near the inner bank, there is substantial streamwise
shoaling of the flow. This shoaling produces a change in the streamwise vertical
structure of the flow and, more specifically, produces enhanced shear near the
boundary. The effect of this vertical structure change, which appears in the model
at the next order in the perturbation expansion, is to enhance the bottom stress.
Near the outer bank, there is relatively rapid streamwise deepening of the flow,
which results in less shear near the boundary, and a diminished boundary stress.
Thus, some of the differences between the measured and predicted bottom stresses
are associated with the assumption of a spatially invariant similarity structure for
the streamwise velocity. However, these errors may be accounted for by solving the
perturbation equations to one higher order and, by doing so, treating the effects of
spatial accelerations on the vertical structure of the streamwise velocity and shear
stress.
In order to calculate the downstream sediment flux fields from the sediment
transport boundary shear stresses discussed above, it is necessary to employ a
bedload equation. The results presented here are from the equation presented by
Valin [1963], which is given by

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Nelson and Smith

93

3000

22

15: ~~~
30'0~

.~O

15.0

~O~

OO~I

30.0

30.0

20

19A

~ 15:l0~-I-----,O~1
15.0

t
O~..I....---L-

~::L
30.0
15.0

Fig. 7.

o
~

O~
120.0

240.0

12

-120.0

O...L--...-J

-240.0

DISTANCE FROM CENTERLINE (eM)

Comparison between measured and calculated values of the boundary shear stress
responsible for sediment transport at six sections in Muddy Creek, Wyoming.

(52)
where D is the particle diameter, S is the local excess shear stress defined by (TSF Te)/ Te, and "'I = 2.45 (pI Ps)O.4( Te/(Ps-p)gD)o.5. Since by far the majority of the
sediment in Muddy Creek is made up of quartz sand grains, the sediment density Ps
is given by 2.65 gm/cm 3 . The grain diameters used in the calculations are the
median sizes measured and presented by Dietrich and Smith [1984], which vary from
about .03 to .20 centimeters in Muddy Creek, depending upon location in the
stream. The critical shear stresses Te were found from the grain size data using
Shields' diagram as presented by Smith [1977].
The calculated downstream sediment flux field is shown in Figure 8, along with
measurements of bedload transport taken in Muddy Creek using a bedload sampler
[Dietrich and Smith 1984]. Although there are some discrepancies, it is clear that
the general structure of the bedload transport is reproduced well by the model

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Flo,v in 11eandering Channels

94

1.0L
05
.

22

oC:, ~~~

1'O~

0..--0

20

0.5

o
1.0

G0.5

19A
0

~
,

18
0

:':L"
o

~o
14

...

oL--J

lOt
O:t-.. . . ----c~~ ,o~
0'------'--_""""""--_....1....-_""""--'_----'-_

12

240.0

120.0

DISTANCE

Fig. 8.

-120.0

- 240.0

FROM CENTERLINE (CM)

Values of downstream sediment flux at six sections in Muddy Creek. The data
points are those given by Dietrich and Smith [1984], while the solid line is the value
calculated using the skin friction bottom stress from the model in the Valin bedload
equation.

results and the Valin equation. The total downstream discharges of sediment at
each section calculated from the model results were fairly close to constant, with
about a 10% variation about the mean value. This is consistent with the
observation that no net erosion or deposition was occurring in Muddy Creek during
the period when measurements were taken.

Velocities and Surface Elevations


The Muddy Creek observations presented by Dietrich [1982] also include values
for the vertically-averaged downstream velocity. This data is shown with the model
results in Figure 9. This comparison clearly demonstrates that the lowest-order
vertically-averaged equations employed in the model represent the salient physics
well. The only noteworthy discrepancy between the measured and calculated
vertically-averag.ed velocities is found near the inner bank from sections 19B to 22.

Copyright American Geophysical Union

Water Resources Monograph

Nelson and Smith

SOO

River Meandering

L[0--00--"0

40:

- lr&
0

80.0

~400

~::> 80.0 ~
v 400

8o.0L
400

!~d-J

-OO-O~
0
0
~
!

,\.J
18

~
I

\,

,cro~OO
I

12

0~C(l"""000~

240.0

O---l

19B

20

cr~-o-----o

80.DL

22

O~

~-r:JJ

0/ '

40.0

o~oo

(:0--0-:)-...0---.0

400

95

80.0l~-o

Vol. 12

120.0

\
I
0

-120.0

0L--J
-2400

DISTANCE FROM CENTERLINE (CM)

Fig. 9.

Comparison of values of vertically-averaged velocity measured in the IMR bend with


those calculated from the model.

In this region, the model tends to underpredict the velocities. This minor error is
almost certainly due to the exclusion of vertical structure changes over the point
bar. By holding the vertical structure constant in the lowest-order equations, but
including the effect of topographic steering, the model tends to underpredict slightly
the volume flux over a three-dimensional obstruction. In reality, the routing of the
flow over an obstruction consists of a component of steering, which routes the flow
around the obstacle, and a component of vertical structure variation, which is
primarily associated with the flow going over the obstacle. By neglecting vertical
structure changes, the model underpredicts the flux of water up over the point bar.
However, this is only a minor error, as is consistent with the model scaling and, in
general, the flow field is predicted accurately. This conclusion is also supported by
comparison of the measured and calculated centerline surface elevations, shown in
Figure 10. The structure of the centerline elevation is typical of natural meandering
streams, with relatively small slopes near the crossings joined by a region of much
steeper slope near the minimum radius of curvature. This structure is produced by
the presence of the topographic forcing terms, since flat-bedded channels with small
width variations have essentially consant centerline surface slopes, as shown

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Flow in Meandering Channels

96
0.5

:z

-1.0

~
~

ex
>

--J

W-2.5

- 4.00L.....------L.-----L...--......L---.L...-.------L-----L..---l...---.1.....--~-----J
.1

Fig. 10.

.2

.3

.4

.5
S/Mo

.6

.7

.8

.9

1.0

Centerline surface elevation in the IMR bend at Muddy Creek. The solid line is the
elevation predicted by the model and the data points are from measurements made
by Dietrich [1982]. The measurements are accurate to within about .2 cm, as shown
by the error bars.

experimentally by Yen and Yen [19711 and verified using the model presented here.
The match between the measured and predicted overall head loss through the bend
indicates that the form drag is well-predicted by the model for the total channel
roughness presented above. Using the Nikuradse roughness values or the sediment
transport Zo values yields much lower total head loss through the bend, which
indicates the importance of including the form drag effects associated with bars and
bedforms usin~ the simple model described above.
Equation (14) yields the cross-stream surface slope which may then be used to
construct the overall surface elevation, which is shown in Figure 11. Both the
measurements and the model show a region upstream of the pool where the
downstream pressure gradient is zero or even slightly positive but the boundary
shear stress is still downstream. The principal balance in this area is clearly
between the pressure gradient and the convective accelerations included in (13),
which again demonstrates that it is inappropriate to introduce these terms as
perturbation quantities.
Figures 12 and 13 show sectional contours of the downstream and cross-stream
velocities obtained from the model and measurements, respectively. The production
of outward flow throughout the water depth over the bar due to the presence of
large fluid accelerations is clear in both cases. Furthermore. the lateral position and
value of downstream and cross-stream velocity maxima are in quite good
agreement. The most noticeable discrepancy between the model results and the
data is found in the vertical position of the downstream velocity maxima. The
observations indicate that the peak velocity is often below the water surface, while
the model can only predict velocity maxima at the surface. The submersion of the
velocity jet is almost certainly due to the momentum redistribution effect of the
cross-stream circulation terms, which only appear in the model at the next order.
Nevertheless, the overall agreement and the fact that the surface velocities above
submerged velocity maxima are only about 10-20% lower in magnitude than the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

97

Nelson and Smith

1.0
-2.0

_5.0

22
0

_--'--_-'--_""--o-:'_--'--_.?__
..

1.0~

0,

- --0-"--..&...10_-0--

_--'--_.....-_""----I~____L,_ _""-I_......L.-_....&.I_---L.:IO~.....I

~_2~:~
~-5.0~

~ ~: .: ~
1.0~

-2.0

:: : :

o
Q

14

I '

240.0

.....L...._-'--_........_.l.....-----I

-2.0

-5.0

18

- 5.0 - - '_--'--_.....'_---'-

1.0~

19A
o

=:

>

Fig. 11.

OA....--.I

20

-2.0~
0
-5.0

o _.....P_.. .

O""-,_O_.....

O--o~

12

120.0
0
-120.0
-240.0
DISTANCE FROM CENTERLINE (CM)

Comparison of calculated cross--stream elevations with the measurements made by


Dietrich at six sections in the IMR study bend. Elevations are relative to the
centerline elevation at the upstream crossing.

maximum indicate that the circulation-induced convective acceleration terms are


correctly placed at higher order. At any rate, this error is a minor one, especially in
light of the fact that it has only a very small effect on the sediment transport, which
is the most sensitive and important field one hopes to predict.
Summary of the Meander Flow Model
In summary, it has been demonstrated that the simple model for flow in curved
channels with typical topography presented in this chapter does a good job of
reproducing velocities, s~rface elevations, boundary shear stresses, and sediment

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Flow in Meandering Channels

98
MEASURED
DOWNSTREAM VELOCITY

PREDICTED
DOWNSTREAM VELOCITY

,: 22
300

( )~O.
'\(0

&0.

200

100

-100

-200

-300

,::l2?~,J
300

E
u

I00

_ _---.Io

I...-.I---Io~

---'-

""""'

--'--'

.100

-200

-300

-100

-200

-300

100 L.......l----L----L----'-----L----L..----L..----L..
300
200
100
-100

-200

-300

:I:
~

Q..

200

100

-100

-200

-300

,:l~B~"J
300

200

100

200

100

-100

-200

3 00

-100

-200 -300

UJ
C

300

200

100

,:J

300

50

300

200

100

5:~2 .~6Q~ ~

loot, 1...!.-!..-.--Ia
300

200

100

,]

.100

- 200

-300

DISTANCE FROM CENTER (em)

Fig. 12.

Measured and predicted contours of downstream velocity at six sections in the


Muddy Creek bend. The contour interval is 20 em/sec for the predicted velocities
(on the right), and 5 em/sec for the measured velocities. The left-hand column is
reproduced with permission from Dietrich [1982].

transport fields measured in a natural meandering stream. The primary reason for
this accuracy is the inclusion of convective accelerations due to downstream-varying
topography and radius of curvature in the lowest-order equations. The scaling
presented here indicates that these terms must be included at the lowest order in
order to construct a model which is applicable to naturally-occurring curved
channels. This result has important implications in the areas of sediment transport

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

99

Nelson and Sluith


MEASURED
CROSS-STREAM VELOCITY

PREDICTED
CROSS - STREAM VELOCITY

50

22
300

50

20
100 '-3--"010-.............'-0....1.-.....
' 0-0....1.-............1.-......
'0-0...1.-_2...1.- -.3.....
00......
00
20
0

-)1
'I

"\-=15........

~'-IO

1,-

0'"

198
100

300

I
300

200

100

100

-200

-300

,:l~B~~!=,= J

o;~

u
J:

5:l20~J

100

200

100

300

-100

200

100

-100

-200

-300

a.
LLJ
a

50

50

100 300

200

100

-100

200

5:r4~

100

I
300

200

100

-100

-200

-300

100 ........--.....-2.....0-0...0.-....1.-............1.-...1.-..1.-"'---"'---.0.-..........

-300

50

100 L......L- ....L..-....L- .....I...-....L,0-0.....I...-.....I...-.....I...-.,.....


00-..L...-....--'-.-3o.l-o~
2 OO
300
20 O

,::r~ ,~~:, ]
300

200

100

-100

-200

-300

DISTANCE FROM CENTER (em)

Fig. 13.

Measured and predicted contours of cross-stream velocity in Muddy Creek in 1978.


Contours are every 5 em/sec. Incomplete contours on the sections in the left-hand
column are the result of missing data. The left-hand column is reproduced with
permission from Dietrich [1982].

and meander stability, and it is certain that the coupling of this model with an
appropriate sediment transport algorithm will provide interesting information on
the formation and evolution of meandering streams and rivers [see Nelson and
Smith, this volume].

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

100

Vol. 12

Flo\v in ~1eandering Channels


Notation

B
Cd
D
E
ft
f2

elevation of channel bed above arbitrary datum


drag coefficient
grain diameter
elevation of water surface above arbitrary datum
vertical structure function for streamwise velocity
vertical structure function for nonsimilarity part of
cross-stream velocity

F2
Fr
g
h
HB
Hn
k
K
10
Mo
n
N

integral of f2 over the flow depth


Froude number
gravitational acceleration
local flow depth
bar height
dune height
von Karman's coefficient (= 0.40)
eddy viscosity
surface elevation drop scale
meander length along centerline
cross-stream coordinate
cross-stream coordinate divided by the centerline radius of
curvature

p
Q
Qs
R
s
S
u.
u
v
w
W

pressure
water discharge
sediment discharge per unit width
centerline radius of curvature
streamwise coordinate
excess shear stress
shear velocity
streamwise velocity
cross-stream velocity
vertical velocity
channel width
vertical coordinate
roughness length

Zo
ll'

closure parameter between vertically-averaged velocity and


bottom stress

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

101

Nelson and Smith


'reB
'rn
~
l

(
(0
K,

A
AB

Ps
T
Tc

Vol. 12

form drag reduction ratio for channel-scale nonuniformities


form drag reduction ratio for bedforms
height of the bedload layer
perturbation parameter, 0(.1)
vertical coordinate nondimensionalized by local depth
nondimensional eddy viscosity
nondimensional eddy viscosity
wavelength
bar wavelength
fluid density
sediment density
stress
critical shear stress for the initiation of sediment motion

Referenoos
Ascanio, M. F. and J. F. Kennedy, Flow in alluvial river curves, J. Fluid Meeh.,
133(1), 1-16, 1983.
De Vriend, H. J., A mathematical model of steady flow in curved shallow channels,
J. Hydraul. Res., 15(1),37-54,1977.
Dietrich, W. E., Flow, boundary shear stress, and sediment transport in a river
meander, Ph.D. dissertation, 261 pp., Univ. of Wash., Seattle, 1982.
Dietrich, W. E. and J. D. Smith, Influence of the point bar on flow through curved
channels, Water Resour. Res. 19(5), 1173-1192, 1983.
Dietrich, W.E. and J. D. Smith, Bedload transport in a river meander, Water
Resour. Res., 20(10), 1355-1380, 1984.
Engelund, F., Flow and bed topography in channel bends, J. Hydraul. Div., Am.
Soe. Giv. Eng., 100 (HYl1), 1631-1648,1974.
Hooke, R. L., Distribution ot sediment transport and shear stresses in a meander
bend, J. Geol., 83, 543-565, 1975.
Jones, D. F., An experimental study of the distribution of boundary shear stress and
its influence on dune formation and growth, M.S. thesis, Univ. of Wash., Seattle,
1968.
Langbein, W. B. and L. B. Leopold, River meanders - Theory of minimum variance,
U.S. Geol. Surv. Prof. Pap., 422-H, Hi-H15, 1966.
Nelson, J. M. and J. D. Smith, Mechanics of flow over ripples and dunes, J.
Geophys. Res., in press.
Onishi, Y., Effects of meandering on sediment discharges and friction factors of
alluvial streams, Ph.D. dissertation, 158 pp., Univ. of Iowa, Iowa City, 1972.
Owen, P. R., Saltation of uniform grains in air, J. Fluid Meeh., 20(2), 225-242,
1964.
Rozovskii, I. L., Flow of water in bends of open channels, Israel Program for
Scientific Translation, originally published by Academy of Sciences of the
Ukranian SSR, 233 pp., 1957.
Schlicting, H., Boundary Layer Theory, 7th 00., McGraw-Hill, New York, 1979.
Smith J. D., Modeling of sediment transport on continental shelves, in Goldberg,
E.D., ed., The Sea: Ideas and Observations on Progress in the Study of the Sea,
Wiley and Sons, New York, 1977.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Flow in Meandering Channels

102

Smith, J. D. and S. R. McLean, Spatially averaged flow over a wavy surface, J.


Geophys. Res., 83,1735-1746, 1977.
Smith J. D. and S. R. McLean, A model for flow in meandering streams, Water
Resour. Res., 20(9), 1301-1315, 1984.
Wiberg, P. L., Mechanics of Bedload Sediment Transport, Ph.D. Dissertation, 132
pp., Univ. of Wash., Seattle, 1987.
Wiberg, P. L., and D. M. Rubin, Bedload roughness produced by saltating grains,
J. Geophys. Res., in press.
Wiberg, P. L. and J. D. Smith, A theoretical model for saltating grains in water, J.

Geophys. Res., 90(4),7341-7354,1985.

Yalin, M. S., Mechanics ofSediment Transport, Pergammon, 1977.


Yalin, M. S., An expression for bedload transportation, J. Hydraul. Div., Am. Soc.
Giv. Eng., 89(HY3), 221-250, 1963.
Yen, C., Bed configuration and characteristics of subcritical flow in a meandering
channel, Ph.D. dissertation, 123 pp., Univ. of Iowa, Iowa City, 1967.
Yen, C. and B. C. Yen, Water surface configuration in channel bends, J. Hydraul.
Div., Am. Soc. Giv. Eng., 97(HY2), 303-321, 1971.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Copyright 1989 by the American Geophysical Union.

Sediment Transport and Sorting at Bends


Syunsuke Ikeda

Department of Foundation Engineering


Saitama University

Abstract
Equilibrium sorting of bed material transported as bedload and bed
topography are considered for uniformly--curved channels. A formula for lateral
bedload transport is developed, including the effects of secondary flow and
gravity due to lateral bed slope. The coarsening of sediment size toward the
outside bank is predicted. The major agent for the sorting is associated with
the tendency for coarser grains to feel a larger ratio of lateral gravitational
force to fluid force than finer grains. The bed topography is calculated so as
to allow for coupling with the lateral sorting.
The results indicate that
sorting suppresses local scour near the outer bank.
An application to an
actual river supports the present theory.
Introduction
Meandering rivers with heterogeneous bed material carry different sediment
sizes in different manners (suspended load or bed load), different proportions,
and different directions. This yields a consistent pattern of longitudinal and
lateral sorting at bends; sediment size is finer at the point bar and tends to
be coarser toward the outside bank, as observed by Jackson [1975], Dietrich et
ale [1979] and Bridge and Jarvis [1982]. In addition, the sediment size is
coarser at the upstream part of the point bar than the downstream part.
The first mechanistic model of the lateral sorting in curved channels was
presented by Allen [1970], in which the grain size distribution is predicted from
the transverse force balance between the upward secondary flow and the
down-pull of gravity exerted on the grains placed on an idealized empirical
Bridge [1976] and Bridge and Jarvis [1982] calculated lateral
bed profile.
sorting by combining the lateral force balance and Engelund's r1974] empirical
bed topography, the lateral slope of which is uniquely correlated with the
dynamic coefficient of Coulomb friction between the moving particles and the
bed. However, !(ikkawa et ale r1976] and Odgaard [1981] have shown that the
lateral bed slope increases with the square root of the Shields Stress.
In
addition, Hooke and Chase [19781 have noted that the model of Bridge [1976]
does not satisfy the condition of bed load continuity for each grain size range.
Deigaard [1980] proposed a model in which the sorting is accomplished by
a transverse downslope movement of bed load (coarser materials) and a
transverse upslope movement of suspended load (finer materials). He noted
that his model cannot predict lateral sorting in the case of pure bed load
103
Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sediment Transport and Sorting

104

transport because Engelund's [1974] lateral bed formula, which does not
account for the effect of the Shields stress, is used in his analysis.
Odgaard [1982] developed a model of static sorting, for which it is assumed
that the bed materials are in critical shear stress everywhere.
Recently, Parker and Andrews [1985] have treated the sorting in sinuous
rivers with nonsuspendable bed material, in which the bed topography and the
probability density of sediment size are solved for a fully mobile condition in
terms of linear analysis. Ikeda et ale [1987] have treated a uniformly--curved
channel case, in which the bed topography has been calculated coupled with
lateral sorting to satisfy the continuity condition of sediment size.
A complete model necessarily incorporates all agencies responsible for the
sorting, Le. the different lateral bed load transport rate for different sediment
sizes, the lateral transport of suspended load due to the secondary flow and
the turbulent diffusion [Ikeda and Nishimura, 1986], and the vertical sorting
due to armoring or pavement [e.g., Parker and Klingeman, 1982).
Mechanism of Sorting
Consider first a wide, uniformly--curved channel with uniform and
nonsuspendable bed material. A constant water discharge, which exceeds the
critical Shields stress, will cause scour at the region near the outside bank and
deposition near the inside bank, as observed consistently at river bends. An
equilibrium bed topography will be realized when the lateral fluid force due to
the secondary flow and the lateral gravitational force exerted on the grains are
just in balance. When a finer grain is placed on the lateral slope at the
channel center, it will move up the slope as it progresses around the bend,
since the up-pull fluid force is larger than the down-pull gravitational force.
A coarser grain will move down the slope, since the former is smaller than the
latter for the grain because the fluid force increases with the square of the
diameter, while the gravitational force increases with the cube of it.
Now, a heterogeneous bed material with a continuous size distribution is
considered in the same channel. Each grain size will shift its locus to find a
stable position where these two forces are in balance, as the grain moves
around the bend. Thus, an equilibrium bed topography is formed coupling
with the lateral sorting. It is, therefore, deduced that the bed material will
be sorted completely in a mathematical sense, and the bed material size and
the local depth of flow are determined uniquely at a given lateral locus for the
uniformly--curved channel.
This is the basic mechanism for sorting in
uniformly--curved bends.
The sorting in sinuous channels is different from that in uniformly--curved
channels in the sense that the former allows a sediment size distribution at an
arbitrary position of the channel [Parker and Andrews, 1985].

Flow in Bends
The flow in bends is characterized by the secondary flow induced by the
difference of centrifugal forces between the upper and the lower layers of flow.
If it is assumed that a curved channel has a sufficiently large radius of
curvature such that the secondary velocity components, Ur and U z, are much
smaller than the longitudinal velocity component, uEh then the non-linear
convective terms such as UrOur/ Or and uzOur/oz in the Reynolds equations
become second order in magnitude and can be neglected. Then, assuming a
vertically constant eddy viscosit;:, f, the vorticity equation, which defines the
secondary flow, is described by lKikkawa et al., 1976]

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

105

Ikeda

(1)
in which \11 = stream function of the secondary flow, rand z = lateral and
vertical coordinates, respectively, with the z-axis being taken upward from the
free surface of the water.
For wide channels the partial derivatives with
respect to r become small everywhere except near the banks.
Then, (1)
reduces to

(2)
in the central region of a bend.
viscosity

A vertically constant value for the eddy

(3)
is used, which is the approximation used by Lane and Kalinske [19411 in
deriving a concentration distribution of suspended sediment. Here u* = focal
shear velocity and h = local depth of flow.
The distribution of Ue in a
channel cross section is given by

~~ = F(r) [~* + 1[In ~ + I]]

(4)

in which '" = Karman constant, U* = shear velocity at channel center, U =


vertically averaged value of ue at channel center, and F(r) = lateral
distribution of uO. An accurate distribution of F(r) must be calculated by
taking account of the lateral transfer of longitudinal fluid momentum due to
secondary flow, the difference of local bottom shear stress in the lateral and
the metric effect (the longitudinal slope of the free surface being inversely
proportional to r for uniformly curved channels).
The theory requires a
nonlinear expansion of the fluid momentum equations [Johannesson and Parker,
1988; Ikeda et al., 1988], and therefore is not treated herein. Laboratory
[I(ikkawa et al., 1976] and field [Rozovskii, 1961] observations have revealed
that F(r) is crudely described by
F(r)

f(r)

(5)

in which f(r) = function of r, which is unity everywhere except near the


banks, where f = 0 (see Figure 1) and R = radius at channel center.
Substituting (3), (4), and (5) into (2), Eq. (2) was solved under the
boundary conditions
\I!

[p.\I!

0, ""lfi:T

0 at z

0, h

(6)

the result is

(7)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sediment Transport and Sorting

106
f

r
B

R -2

R+~

Fig. 1.

Structure of modification function, fer).

in which Ur = lateral velocity component of the secondary flow = lW /r8z,


Cf = (U*/U)2, '" = 1 + z/h, and Fl (",) and F2 (",) are functions described
by
F1
F2 (",)

(",)

15
= 2"

=-

1 ",2
15 [ ",21n '" - 2

[ ",2 1n2 ",

",2 In '"

+ 21

15]
54

(8)

19]
",2 - 54

(9)

The field data obtained by Rozovskii [1961] in the Desna river, a branch of
the Dnieper, are used to test the theory. The profiles of the longitudinal flow
and the secondary flow were measured at different radii of a section (section II
in his monograph). These distributions are described reasonably well by (4)
and (7), as seen in Figure (2).. Equations (4) and (7) have also been tested
against laboratory experiments and provide good agreement [I(ikkawa et al.,
1976].
Fluid Force Exerted on Sediment
Consider a sediment particle
which a cylindrical coordinate is
coordinate, r denotes the radial
coordinate form the bed. Let
which is assumed to be small so

Particl~

moving on the bed of a curved channel, for


defined such that rO denotes the longitudinal
coordinate and n denotes the upward normal
denote the lateral slope of the bed inclination,
that
~

tan

sin

(10)

A typical value of is 0(0.1), and is much larger than the longitudinal slope
of the bed, .
The existence of the lateral gravitational force and secondary flow yields a
situation such that the sediment particle moves along a course that deviates
by an angle fJ from the 0 direction. The secondary flow creates a near bed
fluid velocity which is skewed an angle 0 to the 0 direction.
The fluid force exerted on the particle can be split into two parts, Le. the
drag force, D, and the lift force, L. Assuming sphericity for the particle, the
drag force is described by

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

107

Ikeda

0
2
E

"

'r-

1....'

..c

"t::"

('

(a)

"'A

J..~

/'>.~~

0.5 m/s

0
I

0
2
E

c
'r-

4
/...'

..c

"-

('

"-

(t
(b)

,,~

"'A<~ ".,..~'

8
359

406

385

426

r in m
Fig. 2.

Test of Eqs. 4 and 7 by field data obtained at a bend of the Desna River
[Rozovskii, 1961].

D
in which p

= !.2

pC

!4 d2 (ub -

mass density of fluid, CD

(11)

V )2
p

drag coefficient, d

particle size,

ub = near bed fluid velocity, and Vp = particle velocity. The longitudinal


and the lateral components of D are described, respectively, by

f}

D
r

D ubO - vpf)
u - v
b

br ub -

(12)

pr

vp

Copyright American Geophysical Union

(13)

Water Resources Monograph

River Meandering

Vol. 12

Sediment Transport and Sorting

108

in which ubO, Ubr = 0 and r components of Ub, respectively, and Vpo, Vpr = 0
and r components of Vp, respectively.
From the definitions, the following
relations hold:
U

tan 8 = ubr

(14)

bO

tan {J

= --E.!.

(15)

vpO

The lift force oriented upward normal to the bed is described by

L =
in which C

lift coefficient.

k pC L ~ d2
J

if

(ub -

(16)

V )2

Comparison of (16) with (11) yields at once


(17)

L =aD

in which a = ratio of lift coefficient to drag coefficient.


The drag coefficient for spheres placed on top of close-packed spheres,
defined against ub (vp = 0), was found to group around the drag coefficient
curve for a sphere in free fall [Coleman, 1967]. Therefore, for sufficiently large
Reynolds numbers C is given by

(18)
Several studies [e.g., Chepil, 1958; Coleman, 1967; Davies and Samad, 1978]
reported negative lift for low Reynolds number. Christensen's [1972] analysis
indicated that the lift force should be positive for the rough flow range, Le.
for large Reynolds numbers. Chepil [1958] reported a positive constant ratio
of

0.85

(19)

for large Reynolds numbers.


Lateral Bed Load
The equations of motion for a sediment particle moving on a plane inclined
at angles t/J and J in () and r directions, respectively, are

DO + (M - m)g sin t/J - p[(M - m)g cos J - L]


Dr

v ()
!=0
p

vpr

(M - m)g sin J - p [M-m)g cos <p - L] v

(20)

(21)

in which M = mass of the particle, m = mass of fluid which occupies the


same volume as the particle, J.L = dynamic coefficient of Coulomb friction, and

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

109

Ikeda

= gravitational acceleration.

Since tanb' ( 1 and tan,B ( 1 in

(15), respectively, the following approximations result:


v pO

1 D
,

~ D

(14) and

(22)

Since by assumption 't/J ( J ( 1 , it follows that


sin 't/J

0, cos J

(23)

Substituting (11), (17), (22), and (23) into (20) and assuming sphericity for
the particle, the following relation is derived:

(24)
in which Rs = submerged specific gravity of sand particles = 1.65 for most
natural sands. Substitution of (11), (13), and (24) into (21), with the aid of
(10) and (23), yields an expression for vpr/vp
4 Jle n

-hr- +

tan J

~ = _3_1_+_0'_I-"_JRSid_R-.;s~d

vp

~3 /lC n
vp
4 + all .JRsgd
1

(25)

J.t

+-0'-1-"

"::"'""1.......

Eliminating Vp from the right hand side of (25), and using (24), (25) reduces
to

-.E

vp

u +
=~
uh

t _a_n_<P
C

3 Jl n
4" 1 + O'Jl

(26)

uh
rrr-;;;r
1.l'-sgu

Evaluating ub at the critical condition of motion for which Vp


follows that

o in (24),

it

(27)
Here ubc
near bed fluid velocity at incipient motion.
Iwagaki [1956]
introduced a sheltering coefficient in calculating the critical tractive force, the
meaning of which is that a stationary sand particle placed on the bed is
sheltered by other stationary sand particles. Equation (24) is derived under
the assumption that the particles are in motion, and (27) is obtained as the
limiting case Vp -1 0 in (24). Therefore, a sheltering coefficient A is introduced
in (27) such that

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sediment Transport and Sorting

110

pRsgd

_ ,\ 14

(28)

~~ C (l + aIJ)

u bc -

Here ,\ takes the value [Iwagaki, 1956]

,\ =

0.59

(29)

Substitution of (28) into (26) yields

--E!.

vp

= --E! +
ub

1\

J.L

ap ~ tan J

(30)

ub

The lateral bed load per unit longitudinal length, qr' is equated to
qr _

qo

v pr

___

v pr

(31)

N_

vpO - vp

in which qe = longitudinal bed load per unit width.


Assuming that the
resistance coefficient remains unchanged regardless of fluid velocity, the near
bed fluid velocity can be correlated to the Shields stress as follows:
(32)
in which T*c

critical Shields stress, and T*

Shields stress, given by


(33)

Substitution of (14), (31), and (32) into (30) yields

qr

qo

tan 0

+
1\

(XI!:.

J.L

JT*c

T*

tan

if!

(34)

Engelund [1974] proposed a similar lateral bed load equation


qr

-qo =

tan 8

+ u tan 0

(35)

fA'

Equations (33) and (34) indicate that the deflection angle of the sediment path
from the down-ehannel direction increases as the sediment size becomes large,
if the critical Shields stress remains unchanged with respect to the sediment
size. This indicates that the coarser particles move preferentially toward the
outside bank, which is the basic mechanism of sorting at bends as described
previously.
Therefore, Engelund's relation, (35). is invalid in treating the
sorting, as Parker and Andrews [1985] noted.
Equation (34) is tested in a straight, laterally-inclinable wind tunnel
(Figure 3), for which any desired lateral slope can be obtained between 0 to

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

111

Ikeda

Fig. 3.

Laterally inclinable wind tunnel used to measure the. lateral bed load.

40 .
Details of the measurement procedure are described in Ikeda [1982a,
1982b], and are not repeated herein.
For a straight channel it can be
reasonably assumed that tan 0 = 0 in (34). Then, it follows that

(36)
In the wind tunnel experiments both qe and qr are measured to test the
above relation.
The dimensionless, longitudinal, volumetric bed load for a
level bed, q*e = qe / ,JRsgd 3 , is plotted against T*/ T *c in Fi~ure 4, from which
an appropriate expression for q*e is found to take the form lParker, 1984]

(37)
Substitution of (37) into (36) provides the relation
q*r
tan(j)

0.00595 1

+ o:p
Ap,

[I:L [I:L _1]]1/2


T*c

T*c

(38)

in which q*r = qr/~.


As an example, relations between q*r and T*
measured in the wind tunnel for several values of > are depicted in Figure
5, in which it is found that q*r is approximately proportional to tan>e Using
these data, q*r/tan> is plotted against T*/T*c, from which the most probable
formula for the lateral bed load is found to take a form similar to (38);

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sediment Transport and Sorting

112

0.5

10

30

T*/T*C
Fig. 4.

Longitudinal bed load relation for sand with d

= 0.42

mm.

(39)
The functional relationship between qr/qa and T*/T*c proposed in (36) is thus
shown to be correct.
It should be noted that the longitudinal bed load function defined in (37)
is valid for air flows only. Therefore, another bed load formula such as that
of Einstein [1950], should be used for water flows.
The deflection angle of the near-bed fluid velocity from the down--ehannel
direction, 6, is now estimated. For hydraulically rough boundaries, which are
usually found in rivers, the distribution of longitudinal fluid velocity is
expressed by

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

113

Ikeda

o
o

39

-&

tC

"

10- 2

SiC
r:::::r

()

5
10
15
20

()

25

0
@)
@)

30
5

Fig. 5.

10

Lateral bed load relation for various side slopes for sand with d

-Uo = f -r [1U*

'"

Z + h
In --,:-KS

+ Ar]

= 0.42.

(40)

in which Ar = 8.5, and ks = roughness height which is of the order of the


bed material size. It should be noted that (40) is a more primitive form of
(4). Then, ub9 is given by

(41)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sediment Transport and Sorting

114

The near bed velocity component of the secondary flow, ubr, is given from (7),
(8), and (9) as
U

Therefore, tano

tr = f

= Ubr/Ub9
tant5

[~r ~R

(- 4.167

.J!Jj-)

2.640)

(42)

is given by

f .!..
R

!!. _1_
r

K,A r

[_ 4.167 _1_

/Ci

2.640]

(43)

K,

Bed Material Size Distribution


For heterogeneous bed materials, the grain size, d, becomes a function of r,

as mentioned earlier, and a continuity equation for sediment size is required to


close the problem.
Consider a
differential

element

of

bed

surface

with

an

area

rdOdr ./1 + (dh/dr)2 (hatched area in Figure 6a), in which a body of uniform
grains with a diameter, d, gathers as a result of lateral sorting, as described
earlier. The volume of sand grains therein, dV, is

dV

(1 - p)

+ [~r

(44)

rdOdrdz

in which p = porosity of bed material, d (J = central angle of the area, and


dz = thickness of the thin surface layer. This volume must be equal to the
total volume multiplied by the probability density of the size, d, in the
original mixture before sorting (hatched area in Figure 6b), such that

(l-p) 1 +

[~r

R+!!

fB

rdOdrdz = ted) dd

(l-p) 1 +

R--2

[~r

rdOdrdz (45)

in which B = channel width, and t(d) = probability density of bed material


size, d, which satisfies the following relation:
00

t(d) dd = 1

(46)

o
Since dh/dr is the order of 0.1 for natural rivers, it is reasonably assumed
that (dh/dr)2 ( 1 in (45). Then, (45) reduces to
dd

or =

I(d)

r
BR

(47)

For uniform grains, t(d) reduces to the Dirac delta function and is therefore

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

115

Ikeda

(a)

4>(d)

d.

mln

d d+dd

max

(b)
Fig. 6.

Probability density of heterogeneous bed material: (a) sorting in uniformly


curved channel; (b) probability density of original mixture.

infinite. Thus (47) indicates that the bed material size is constant laterally.
The size distribution can be obtained by integrating (47).
Bed Topography

A stable bed topography in uniformly curved channel is realized when the


lateral bed load vanishes everywhere, from which (34) reduces to

dh - f !.

or -

!!r

- 1.941 + 1.226 _1_

{Ci

Copyright American Geophysical Union

(48)

Water Resources Monograph

River Meandering

Vol. 12

Sediment Transport and Sorting

116

in which dh/dr = tan, is used for the derivation, and the numerical constants
in (34) are evaluated as a = 0.85, P, = 0.43, and ,\ = 0.59 [Kikkawa et al.,
1976]. Assuming a constant resistance coefficient, as used in deriving (32), the
local Shields stress, T*, is described, with the aid of (5), by
(49)

in which fI
yields

Substitution of (49) into

vertically averaged value of u.

- 1.941 + 1.226
5 . 382 ..[T;;,

(48)

_1_

rei

(50)

Equation (50) reveals that the lateral slope of line bed in bends increases with
the sediment particle Froude number, u*/,JRsgd, as evidenced by much
laboratory and field data [Odgaard, 1981]. For uniform bed material, (50) can
For
be integrated directly to determine the lateral bed topography.
heterogeneous bed materials, however, the grain size, d, becomes a function of
the lateral coordinate, r, as described by (47). Since (dh/dr)2 is neglected in
(47), the lateral distribution of sediment size can be determined independently
without coupling with the bed topography.
Substituting the sediment size
distribution thus obtained into (50), the bed topography can be defined by
integrating (50) with the boundary condition
h

H at r

(51)

in which H = flow depth at channel center.


Equations (47) and (50) are the basic equations which serve to define the
distributions of bed material size and bed topography.
In solving the
equations the following independent variables must be specified in advance: S,
R, B, Cf, I(d) and T*c, in which S = free surface slope along channel
centerline.
Application
A complete set of data at bends which can specify all the independent
variables could not be found. For example, the data obtained by Rozovskii
[19611 at the Desna River do not include the median size and the distribution
of tne bed material and therefore cannot be used to test the theory.
Jackson's [1975, 1976] data obtained for Helm bend in the lower Wabash
River, Illinois and Indiana, provide information for all hydraulic variables
except U* which can be crudely equated to 0.1 U. The lower Wabash River
maintains a fairly stable discharge at Mt. Carmel, Illinois, varying form 200
m3 /s to 3,000 m 3 /s during November, 1972 and September, 1973.
The
bankfull discharge, which is dominant in maintaining channel width [Andrews,
1984; Ikeda et al., 1988], is about 1,200 m 3 fs. Therefore, the cross-sectional
survey at Helm Traverse II for 1,270 m /s on May, 1973 seems to be
appropriate to test the theory, for which U = 1.25 mise The corresponding
geometric properties are H = 3.4 m, R= 610 m, and B = 250 m (see Figure

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Ikeda

500 m

117

II

Fig. 7.

Plan view of Helm bend of the lower Wabash River [Jackson; 1975, 1976].

7). The probability density of bed material size is depicted in Figure 8, for
which the median size, dso , is seen to take a value of 0.9 mm. Since U*
dso/ v is estimated to be 110, the bed is in assumed to be fully rough, for
which the critical Shields stress, T*c, is found to be 0.066 from the Shields
curve [see e.g., Vanoni, 1975]. All independent variables that are required to
integrate (47) and (50) have thus been specified.
The bed material size distribution measured at Helm Traverse II during
August, 1972 and September, 1973 [Jackson, 1976] is depicted in Figure (9), in
which the sediment size originally described by a phi-scale (= -log2d) is
converted to d-scale. The calculated size distribution using (47) is shown by
a solid line in the same figure. It is seen that sediment size increases very
rapidly near the outer bank, a feature explained reasonably well by the theory.
A laboratory test also supports the present theory {Ikeda et al., 1987].
The lateral bed topography predicted by (50), Including the lateral sorting,
is depicted in Figure 10 by a solid line. A broken line in the figure indicates
the calculated bed topography, assuming that the bed consists of a uniform
bed material of d = 0.9 mm. It is found that the coarsening of bed material
size toward the outside bank reduces the lateral bed slope considerably in the
outer area. The inclusion of sorting in the theory reduces the maximum depth
at r/R = 1.16 by 36% for this case. A theory that includes lateral sorting
thus predicts the measured bed topography much better than a theory that
assumes a uniform bed material. Therefore, for natural rivers which usually
have wide sediment size distribution [see e.g. Kellerhals, 1967; Bray, 1979], the
bed topography must be calculated by including the effect of lateral sorting.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sediment Transport and Sorting

118

E
E
s:::

0.5

'r-

-0

0&

o
0.1

10

d in mm
Fig. 8.

Probability density of bed material size at Helm Traverse II, the lower
Wabash River.

The effect of suspended load on the bed topography must also be estimated
because T* is estimated to be 1.07 along the centerline at this bend, and
because it is reported that sediment is transported actively as suspended
load [Jackson, 1976].
The concentration of suspended sediment increases

-0

o
0.8

0.9

1.1

r/R
Fig. 9.

Lateral distribution of bed material size [Jackson; 1975, 1976].

Copyright American Geophysical Union

1.2

Water Resources Monograph

River Meandering

Vol. 12

119

Ikeda

''''', -

2
.2

......

Measured bed topography

......

",

" ,,
" ,,

,,
,,
,,

4
5

0.8

0.9

1.1

" ' ....


1 .2

r/R
Fig. 10.

Bed topography at Helm Traverse II [Jackson; 1975, 1976].

toward the bottom, and the familiar distribution is expressed by


C

= cb exp [-

~:

(z

h)]

in which c = local concentration of suspended sediment, cb =


concentration of suspended sediment, Ws =. settling velocity of
sediment, and fm = vertically--eonstant eddy diffusivity for suspended
The eddy diffusivity, fm, is known to be slightly larger than
viscosity, f [e.g. Vanoni, 1975]. It is assumed herein that

(52)
near-bed
suspended
sediment.
the eddy

(53)
which implies that fm =
suspended sediment can be
conditions in unidirectional
[Ikeda and Nishimura, 1985]

1.16f (see (3)).

The near-bed concentration of


estimated with the data obtained for equilibrium
flow, for which an empirical relation is proposed
such that

c = 2.31xlO- [u*]
4

1 60

Ws

, (54)

The vertically-integrated lateral volumetric flux of suspended sediment per unit


longitudinal length is described by

qcr

=f

u c dz
r

-h

Substitution of (7) and (52) with the aid of (8) and (9) yields

Copyright American Geophysical Union

(55)

Water Resources Monograph

River Meandering

120

Vol. 12

Sediment Transport and Sorting


10- 2

- - - - - - - - , . - - - - - - - . , . -....

1...-""""'--------'-------.. . .

10- 6
10-1

10

Fig. 11.

qcr
Uh

= _ f2

r2 h f
W r cr

(56)

in which fer is a function of u*/ws and Cf, and the functional relationship is
depicted in Figure 11.
The condition for equilibrium in this case is described by
qr

qcr

Substitution of (34) and (56) into (57) yields, with the aid of (43)

Copyright American Geophysical Union

(57)

Water Resources Monograph

River Meandering

Vol. 12

121

Ikeda

2
..........

..c

.......

Eq. 43
4

Eq. 58

D.H

0.9

1.1

1.2

r/R
Fig. 12.

Effects of suspended sediment on the bed topography.

qo [- 1.941

1. 226 _1_]

dh _ f r h

or -

It

5.382qe

{Cj

+f~

Uhler

(58)

~r;-

Comparison of (58) with (48) indicates that the lateral bed slope increases due
to the lateral convective transport of suspended sediment. Equation (58) is
applied to the Helm Traverse II, and the result is depicted in Figure 12 in
which it is assumed that the sediment size is a uniform 0.9 mm. It is found
that the inclusion of the suspended sediment increases the maximum depth of
scour by about 10%.
Lateral sorting, however, is expected to reduce the
amount of suspended load as described below. At a typical position rlR =
1.1, the sediment size is about 2 mm from Figure 9, for which cb is calculated
to be about 30% of that for 0.9 mm. Since the effect of suspended load on
the lateral bed slope at bends increases in proportion to cb, the suspended
sediment is inferred to increase the maximum depth of scour by about 3% if
lateral sorting is considered, which is much smaller than the value 36% caused
by the variation of lateral bed load due to sorting.

Conclusions
A formula for lateral bed load on a side slope is derived, which indicates
that the lateral bed load increases in proportion to the lateral slope and is
inversely proportional to the square root of the Shields stress. A wind tunnel
test supports the formula. A model for defining bed topography and sorting
in bends is presented using the proposed formula for lateral bed load. The
model is applied successfully to a natural river. The present study indicates
the following: (1) The major mechanism for lateral sorting is that coarser

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sediment Transport and Sorting

122

bed material feels a larger ratio of transverse gravitational force to transverse


fluid force than finer material; (2) sediment size shows a coarsening toward
the outside bank; and (3) the sorting reduces the lateral bed slope considerably
in the outer area of bends (near the outside bank) because the coarser
materials prevent scouring. For example, a practical application to the lower
Wabash River indicates that the maximum depth of scour is reduced by about
36 percent due to sorting.

Acknowledgements
Financial support was provided by a Grant-in-Aid for Scientific Research
from the Ministry of Education and Culture of Japan, Grant No. 59020003.
The present study is also supported by Joint United States-Japan Research on
River Meandering during May 1985 and October, 1987.

Notation
Ar
B
CD,C L

Cf

DO,D r
d
Fl, F2

a numerical constant

8.5

channel width
drag and lift coefficient, respectively
resistance coefficient

(U./U)2

local concentration of suspended sediment


bottom concentration of suspended sediment
drag force acting on bed material
() and r component of d, respectively
sediment size
funct ions

of

TJ

for

secondary

flow

in

(8) and (9),

respectively
F(r)
f(r)

lateral distribution of longitudinal fluid velocity


normalized distribution of F(r) as seen in Fig. 1

gravitational acceleration

depth at channel center

local depth

bottom roughness height

lift force acting on bed material

mass of sediment particle

mass of fluid which occupies the same volume as the


sediment particle

porosity
vertically-integrated lateral volumetric flux of suspended
sediment per unit longitudinal length

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

123

Ikeda
qO,qr

longitudinal

and

lateral

volumetric

bedload

per

unit

length, respectively
q*O,q*r
R
Rs

dimensionless values of qe and qr, respectively


radius of curvature at channel center
submerged specific gravity of bed material

local radius of curvature

vertically-averaged value of longitudinal fluid velocity at


channel center

U*
ub

shear velocity at channel center

u bc

near bed fluid velocity at incipient motion of bed material

ubo,u br
uO,u r
u*
V

vpO ,vpr
W

s
Z

near bed fluid velocity


longitudinal and lateral component of
longitudinal

and

lateral

respectively

Ub,

component

of

fluid

velocity,

respective I y
local shear velocity
velocity of sediment particle
longitudinal and lateral component of

Vp,

respectively

settling velocity of suspended sediment


vertical coordinate taken positive upward from the free
surface

G'

ratio of lift coefficient to drag coefficient

(3

deviat i on of sediment particle path from the longitudinal


direction

deviation of direction of near bed fluid velocity from the

eddy viscosity

longitudinal direction
{

eddy diffusivity of suspended sediment

1/

dimensionless vertical coordinate

longi tudinal coordinate

'A"

Karman's constant

a sheltering coefficient

Jt

dynamic coefficient of Coulomb friction

mass density of fluid

T*c

z/h

0.4

0.59

0.43

critical Shields stress at incipient motion of bed material

lateral slope of bed

stream function of secondary flow

probability density of sediment size distribution

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Sediment Transport and Sorting

124
fer

dimensionless lateral transport of suspended sediment


longitudinal slope of bed

Referenoos
Allen, J. R. L., Studies in fluvial sedimentation:
A comparison of fining
upwards cyclothems, with special reference to coarse-member composition
and interpretation, Journal of Sedimentary Petrology, 40, pp. 298-323, 1970.
Andrews, E. D., Bed-material entrainment and hydraulic geometry of gravelbed rivers in Colorado, Geological Society of America Bulletin 95. pp.
371-378, 1984.
Bray, D. I., Estimating average velocity in gravel-bed rivers, Journal of the
Hydraulics Division, Am. Soc. Giv. Eng., 105, No. HY9, pp. 1103-1122,
1979.
Bridge, J. S., Bed topography and grain size in open channel bends,
Sedimentology, 23, pp. 407-414, 1976.
Bridge, J. S. and J. Jarvis, The dynamics of a river bend: A study in flow
and sedimentary processes, Sedimentology, 29, pp. 499-541, 1982.
Chepil, W. S., The use of evenly spaced hemisphere to evaluate aerodynamic
forces on a soil surface, Transactions, Am. Geophys. Union, 39, No.3, pp.
397-404, 1958.
Christensen, B. A., Incipient motion of cohesionless channel banks,
Sedimentation, H. W. Shen, Ed. and Pub., Fort Collins, Colorado, Chap. 4,
1972.
Coleman, N. L., A theoretical and experimental study of drag and lift forces
acting on a sphere resting on a hypothetical stream bed, Proceedings of the
Twelfth Congress, International Association for Hydraulic Research, Fort
Collins, Colorado, pp. 185-192, 1967.
Davies, T. R. H. and M. F. A. Samad, Fluid dynamic lift on a bed particle,
Journal of the Hydraulics Division, Am. Soc. Giv. Eng., 104, No. HY8, pp.
1171-1182, 1978.
Deigaard, R., Longitudinal and transverse sorting of grain sizes in alluvial
rivers, Series Paper 26, Technical University of Denmark, 1980.
Dietrich, W. E., J. D. Smith, and T. Dunne, Flow and sediment transport in
a sand bedded meander, Journal of Geology, 87, pp. 305-315, 1979.
Einstein, H. A., The bedload function for sediment transport in open channel
flow, Technical Bulletin 1026, U. S. Department of Agriculture, Soil
Conservation Service, Washington, D. C., 1950.
Engelund, F., Flow and bed topography in channel bends, Journal of the
Hydraulics Division, Am. Soc. Giv. Eng., 100, No. HY11, pp. 1631-1648,
1974.
Hooke, R. L. and C. E. Chase, Flow, bed topography, grain size, and
sedimentary structure in open channel bends: A three-dimensional model:
A discussion, Earth Surface Processes Landforms, Vol. 3, 421-422, 1978.
Ikeda, S., Incipient motion of sand particles on side slopes, Journal of the
Hydraulics Division, Am. Soc. Giv. Eng., 108, No. HY1, pp. 95-114, 1982.
Ikeda., S., Lateral bed load transport on side slopes, Journal of the Hydraulics
Division, Am. Soc. Giv. Eng., 108, No. HY11, pp. 1369-1373, 1982.
Ikeda, S. and T. Nishimura, Bed topography in bends of sand-silt rivers,
Journal of Hydraulic Engineering, Am. Soc. Giv. Eng., 111, No. 11, pp.
1397-1411, 1985.
Ikeda, S. and T. Nishimura, Flow and bed profile in meandering sand-silt
rivers, Journal of Hydraulic Engineering, Am. Soc. Giv. Eng., 112, NO.7,
pp. 562-579, 1986.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

125

Ikeda

Ikeda, S., M. Yamasaka, and M. Chiyoda, Bed topography and sorting in


bends, Journal of Hydraulic Engineering, Am. Soc. Giv. Eng., 113, No.2,
pp. 190-206, 1987.
Ikeda, S., G. Parker, and Y. Kimura, Stable width and depth of straight
gravel rivers with heterogeneous bed materials, Water Resources Research,
Am. Geophys. Union, 24, No.5, pp. 713-722, 1988.
Ikeda, S., M. Yamasaka, and K. Sakarori, Structures of flow in curved
channels, Proceedings, Japan Society 0 Civil Engineers, (to be submitted),
1988.
Iwagaki, Y., Fundamental study on critical tractive force, Proceedings, Japan
Soc. Giv. Eng., 41, pp. 1-21, (in Japanese), 1956.
Jackson, R. G., Velocity-bed form-texture pattern of meander bends in the
lower Wabash Ri ver of Illinois and Indiana, Geological Society of America
Bulletin, Vol. 86, pp. 1511-1522, 1975.
Jackson, R. G., Unsteady-flow distributions of hydraulic and sedimentologic
parameters across meander bends of the lower Wabash River,
Illinois-Indiana, U.S.A, Proceedings of the International Symposium on
Steady Flow in Open Ghannel Flows, Cranfield, England, pp. G4.35-48,
1976.
Johannesson, H. and G. Parker, Velocity redistribution in meandering rivers,
Journal of Hydraulic Engineering, Am. Soc. Giv. Eng. (submitted for
publication), 1988
Kellerhals, R., Stable channels with gravel paved beds, Jour. Waterways
Division, Am. Soc. Giv. Eng., 93, No. WWI, pp. 63-84, 1967.
Kikkawa, H., S. Ikeda, and A. Kitagawa, Flow and bed topography in curved
open channels, Journal of the Hydraulics Division, Am. Soc. Giv. Eng., 102,
No. HY9, pp. 1327-1342, 1976.
Lane, E. W. and A. A. Kalinske, Engineering calculation of suspended
sediment, Transactions, Am. Geophys. Union, 22, 1941.
Odgaard, A. J., Transverse bed slope in alluvial channel bends, Journal of the
Hydraulics Division, Am. Soc. Giv. Eng., 107, No. HYI2, pp. 1677-1694,
1981.
Parker, G. and P. C. Klingeman, On why gravel bed streams are paved,
Water Resources Research, Am. Geophys. Union, 1409-1423, 1982.
Parker, G., Discussion of lateral bed load transport on side slopes, by S.
Ikeda, Journal of Hydraulic Engineering, Am. Soc. Giv. Eng., 110, No.2,
pp. 197-203, 1984.
Parker, G. and E. D. Andrews, Sorting of bed load sediment by flow in
meander bends, Water Resources Research, Am. Geophys. Union, 21, No.9,
pp. 1361-1373, 1985.
Rowvskii, I. L., Flow of water in bends of open channels, Israel Program for
for Scientific Translations, Jerusalem, Israel, 1961.
Vanoni, V. A., ed., Sedimentation Engineering, Manuals and Reports on
Engineering Practice, No. 54, Am. Soc. Civ. Eng., 1975.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Copyright 1989 by the American Geophysical Union.

Sediment Control by Submerged Valles. Design Basis


A. Jacob Odgaard and Anita Spoljaric
Department of Civil and Environmental Engineering
and Institute of Hydraulic Research
The University of Iowa, Iowa City

Submerged vanes are small-aspect ratio flow training structures installed on


the streambed, usually oriented at 10 to 20 to the local primary flow
direction. Vane height is typically 0.2 to 0.5 times the local water depth
during design-flow conditions. In curves of river channels, they eliminate the
centrifugally induced secondary motion typical of flows in curved channels and
the root cause of bank undermining. In shoaling channels, they generate a
secondary motion which redirects the sediment and provides depth control.
The key to vane performance is the horizontal force that they exert against
water flow and its effect on near-bed flow, and on circulation induced in flow
downstream from vanes. These features determine the number and layout of
vanes of a given design required to eliminate the problem of either bank
erosion or shoaling.
Herein, a summary is given of the theoretical background and design basis.
The summary includes a description of recent attempts to verify and improve
critical design relations.
Laboratory experiments are described in which
measurements were taken of: (1) transverse (lift) and streamwise (drag) force
components on vanes, and (2) induced velocity field downstream from the
vanes. The variables tested were: (1) vane orientation (angle of attack), (2)
ratio of vane height to water depth, (3) vane shape, and (4) Froude number.
The data were analyzed within the framework of aerodynamic theories for
finite wings. Relatively simple design relations developed on the basis of wing
theories were tested and verified.

Introduction
River meandering, and in particular the bank erosion attendant to the
growth and migration of the meander loops, is a national problem in many
countries. In the United States, its significance has been documented by the
U.S. Army Corps of Engineers rI98I] in a report to Congress. The Corps'
report, which is based on work conducted under theStreambank Erosion
Control Evaluation and Demonstration Act of 1974 (Section 32 of Public Law
93-251), shows that one of the principal obstacles to amelioration of the bank
erosion problem is the cost and inadequacy of the currently available riverbank
protection methods. These involve either: (1) armoring of concave banks by
one means or another, ranging from protecting them with stone or concrete to
enhanced vegetative cover; or (2) reducing the near bank velocity by dikes or
other structures that increase the local channel roughness or shelter the banks.

127
Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Submerged Vanes

128

A review of various bank protection measures has been presented also by

Jansen et al., [1979].

The problem of bank erosion is related to the problem of in-stream


shoaling.
The sediment entrained by a river from bank erosion is often
transported downstream and deposited in areas where it interferes with
navigation and diminishes channel capacity. In-stream shoaling is responsible
for much of the dredging done annually by the U.S. Army Corps of Engineers
and others to maintain navigation depths and channel conveyance capacities.
This problem is especially acute along the crossings between successive river
bends. In recent years, the Corps of Engineers alone has dredged an annual
average of approximately 65 million cubic yards to maintain navigation in the
Nation's waterways. The problem has other facets, because environmentally
acceptable areas for dredge-spoil disposal are often remote from the dredging
site, or not available at all.
The Corps of Engineers' report [1981], as well as the extensive body of
published literature on the subject, clearly demonstrate that there is a need for
development of new concepts and corresponding approaches for safeguarding one
of the Nation's most valuable resources, its agricultural land, from river bank
erosion; and for alleviating the related problem of deposition of the eroded
material downstream where it poses other major, expensive sediment
management problems.
Recent analytical and experimental studies conducted at the Iowa Institute
of Hydraulic Research (IIHR) have led to development of a new concept for
bank protection [Odgaard and Kennedy, 1983], which is more economical and
more acceptable to environmental interests than traditional concepts; and
appears also to hold promise for alleviation of shoaling along river crossings
The concept consists of flow training by small,
between successive bends.
submerged vanes. The vanes are vertical, small-aspect ratio plates or panels
installed on the streambed, usually oriented at 10 to 20 to the local primary
flow direction. Vane height is typically 0.2 to 0.5 times the local water depth
during design-flow conditions. Theoretical and experimental studies at IIHR,
as well as field studies, have shown that such vanes, if installed correctly, can
significantly reduce erosion of streambanks.
Bank erosion generally occurs in curved river channels, where the
interaction between the vertical gradient of velocity and the curvature of flow
generates a secondary or spiraling motion of flow. The secondary flow moves
high velocity, surface current outward and low velocity, near-bed current
inward. The increase in velocity near the outer bank increases flow depth a~d
undercuts the bank causing bank erosion. By directing the near-bed current
toward the outer banks, submerged vanes counter the spiral flow and, thereby,
inhibit bank erosion. The vanes make the water and sediment move through
the channel curve as if it were straight. The effectiveness of the vanes has
been demonstrated in a bend of East Nishnabotna River, Iowa [Odgaard and
0

Mosconi, 1987].

A preliminary investigation conducted at IIHR has demonstrated that the


vanes are also effective in deepening the center portion of a straight channel.
In this application, the vanes generate a secondary current which imparts a
transverse shear stress to the bed and causes a transverse transport of
sediment. As demonstrated by Odgaard and Spoljaric [1986], the vanes can be
laid out to accomplish this without changing overall sediment transport. This
is judged to be an important finding in regard to application of vanes for
alleviation of shoaling problems.
The key to the vane performance is the horizontal force that they generate
and its effect on near-bed flow, and on circulation induced in the flow
downstream from vanes. These features determine the number and layout of

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

129

Odgaard and Spoljaric

vanes of a given design required to alleviate the problem of either bank


erosion or shoaling.
In this paper, a summary is given of the theoretical
background and design basis. The summary includes a description of recent
attempts to verify and improve critical design relations.
Theory
Although the literature indicates that vanes, or panels, of a form similar to
that developed at IIHR have been used in the past for river--ehannel
stabilization [Potapov, 1947, 1950, 1951; Chabert et al., 1961; Jansen et al.,
1979], little is documented on vane design and performance. The first known
attempts to develop a theoretical design basis are those of Odgaard and
Kennedy [1983], and Odgaard and Spoljaric [1986]. Odgaard and Kennedy's
efforts are aimed at the problem of streambank erosion, whereas Odgaard and
Spoljaric's work is focused on shoaling problems, in particular maintenance of
navigation channels and flood-flow capacities.
In the case of streambank
erosion, the vanes function by eliminating the centrifugally induced secondary
motion in the flow; in the case of shoaling, they generate a secondary motion.

River-Bend Bank Protection


In a bend flow, the secondary current is produced by the nonuniform
vertical distribution of centrifugal force (per unit mass of fluid), which in turn
is a result of the primary flow that is nonuniformly distributed over the
depth. The secondary current exists because the faster moving fluid near the
free surface experiences a larger centrifugal force than the slower moving
near-bed fluid.
Theoretically, elimination of the secondary current is
accomplished by incorporating into the governing flow equations a body force
of appropriate form.
The governing equations are either those of
moment-of-momentum or those of linear momentum.

Moment-oj-momentum. In fully developed bend flow, the moment (about,


say, the channel centroid) resulting from the nonuniform distribution of
centrifugal force is balanced by the moment of the radial component (produced
by the secondary flow) of the boundary shear stress. This balance is described
and analyzed by Zimmermann and Kennedy [1978] and in greater detail by
Falcon and Kennedy [1983]. Any force (structure) exerting a torque on the
flow that balances the centrifugally induced torque will prevent the secondary
flow and its effects (bed warping, bank undermining).
For a channel cross
section that is rectangular (the desired shape) and a primary flow velocity
profile given by the power law,
(1)

in which fi = depth-averaged mean velocity; d = flow depth; n = exponent;


and u = point velocity at height z, the centrifugal force torque about the
section centroid acting on a volume element with included angle tP is

Tc

=f

ro

ri

~2

[z -~]

r; dzdr

= ~ pf/ n(~~2)

Copyright American Geophysical Union

b;d2

(2)

Water Resources Monograph

River Meandering

Vol. 12

Submerged Vanes

130

\~~

T
v~
7~7n::A:: ~
T

-=-

SECTION A-A

Fig. 1.

Centrifugal-force torque, and torque produced by riverbed vane.

in which r = radius of curvature; ri, ro = innermost and outermost section


radii; p = density; and b = width of section = ro - ri (Figure 1). The
velocity profile exponent, n, is related to Darcy-Weisbach's friction factor, f,
as n = ~ [Zimmermann and Kennedy, 1978]; K, = von Karman's constant
(~0.40).
Hence, n = K,u//i,Sd, in which S = longitudinal slope of water
surface. As the radial-pressure-gradient force is nearly constant over depth, it
exerts no moment about the centroid.
Therefore, the secondary flow is
essentially produced by T c. It should be noted that, because T c is computed
by integration over the flow section, the details of the cross section shape
utilized are not important.
Use of a shape other than rectangular and
reasonable corresponding lateral distribution of the primary velocity yields
expressions for T c that do not differ significantly from (2).
The submerged vanes are installed to counter T c.
The vane-induced
torque, Tv, is generated by the lateral ("lift ") force, F v, that each vane exerts
on the flow. The increment of force exerted on the flow by a vane element of
height dz is
dF v

= 21

cL pu 2L dz

(3)

in which c = lift coefficient; and L = vane length. The torque generated by


L
N independent vanes about the section centroid is then

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Odgaard and Spoljaric

131

Tv = N

I
o

[~-

z] dF v =

pNL

2
L u

[~-

z] dz

(4)

in which H = vane height. To eliminate the secondary flow, Tv must balance


Tc. Equating Tc and Tv yields
NHL

f1D

~ [11

n+ 1
n(n+2) r

CL

[!!]- 2
u

[1

_~]d d [~]H ]

-1

(5)

This equation yields total vane area (NHL) needed per unit surface area of
channel bed.
If cL is constant over the vane and equal to

cL '

(5) is reduced to

(6)
in which
F =

[~2/n [(n+l) - (n+2) ~r

(7)

The function F has its minimum with respect to H at HId = 2(n+1)/(N+2)2:


. _
n+2 [2( n+1~] -2/n
F mIn - n(n+1) (n+ 2)

(8)

It is of importance to note that F is relatively insensitive to variations of HId


over a fairly large range of HId values. For example, for n=4, F is within
20% of its minimum value (the design value) when HId is within the range
0.12 < HId < 0.48 (Le., when the water depth is between 2 and 8 times vane
height).
In other words, the vanes function nearly optimally over a wide
range of river stages.
Note also that discharge is not a primary design
parameter. Discharge is used only for the determination of n.
Linear momentum. In fully developed bend flow, the radial component of
the linear-momentum equation is essentially a balance between centrifugal force
and radial pressure-gradient force,

(9)
in which p = pressure. The radial shear-stress component resulting from the
secondary current has little effect on this balance and is neglected herein. The
secondary current exists because the pressure-gradient force is nearly uniform
over depth, while the centrifugal force is nonuniformly distributed. It follows
that if the flow is affected by a force distribution fc(z) that eliminates the
nonuniformity of the centrifugal force, the flow will be free of secondary
current. The momentum balance then reads

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Submerged Vanes

132

1.0
""C

"

0.8

0"
w
(D

0.6

>
0

(D

0.4
~

:r:

z /d

w 0.2
:r:

O..-:::;;...;"---:lo..........I..lIo.--:ll.....-.:L..-.------L..-.......

1.0

CENTRIFUGAL FORCE) u2/~2

Fig. 2.

Centrifugal-force profile, and force distribuiton


centrifugal-force deficit and secondary current.

fc(z)

= ~r -

fc

required

to

U2

eliminate

(10)

p -r

The most obvious force distribution to add is one which results in a pressure
gradient of {}p/{}r = pUJ./r. Equation (10) is then reduced to
fc(z)

fi2-- ru=p-

(11 )

This distribution changes sign over depth, and it calls for a vane that
produces an outward force on the flow near the bed and an inward force near
the water surface as indicated in Figure 2. With the velocity distribution
given by (1), the force is zero at z = Zo, where

~o

[n~lt

(12)

The total near-bed (outward) force required for a given bend section is then
ro

Feb

Zo

f f

fi2 -

u
2

p - - r - r; drdz

= pu_ 2

2
n+2

[n]
n
d
n+l
r;b r

(13)

in which
Zo

Feb

fe(z)dz

Copyright American Geophysical Union

(14)

Water Resources Monograph

River Meandering

Vol. 12

133

Odgaard and Spoljaric

:r:

.........
N

1.0

~ 0.8
<t:

>

~ 0.6

-.J

<t:
~

I
(!)

W
I

0.2

OL..----------L---------L-----I~

0 1 2
LIFT COEFFICIENT, C
L

Fig. 3.

Ie L

Distribution of lift coefficient for elimination of centrifugal-force deficit; n = 4, and


HId O.3-Q.4.

The total near-surface, inward force required is of the same magnitude.


The vane shape required to eliminate the centrifugal-force deficit is
obtained by equating the horizontal lift force on vane panels of height dz,

(15)
and the centrifugal-force deficit for a volume element of size (b)(r,)(dz) at
distance z from the bed
dF c

(b)(r~)(dz)fc

= pbr~

fi2 -

u2

-r-- dz

(16)

Equating dF v and dFc yields


CL

~~t

[* - 1]

(17)

or, with u given by (1) and NL by (6)

:L = ~

[*f2-D)/n [[n~d2 [~r/n _ []21n]

(18)

Figure 3 shows the variation of cL over vane height as computed by (18)


using n = 4 and Hid = zo/d. This distribution calls for a vane with angle of

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Submerged Vanes

134

attack G' increasing from 0 at z = Zo to 1r/2 at z=o. If H > zo, the vane
must have a negative angle of attack above elevation zo,
In this approach, total elimination of secondary current requires a vane
extending to the water surface.
The vane must be twisted; its angle of
incidence must vary from top to bottom.
Such a vane affects vertical
momentum by inducing a vertical velocity component, which further increases
induced circulation. In other words, a vane with such a twist may generate a
stronger than necessary secondary current.
The same follows from the
moment-of-momentum analysis, because the twist gives rise to a vertical lift
force (in addition to the horizontal), whose moment about the section centroid
may be significant.
The dynamic features of a twisted vane are currently
being examined.

Shoaling Amelioration
Vane layouts for control of sediment moveluent in straight channels are
designed on the basis of either the vorticity equation or the transverse
component of the momentum equation.
The vorticity equation describes
downstream decay of the vortex induced by the vane. The vortex is centered
near the vane's top elevation and it gives rise to a near-bed tangential
(transverse) velocity of
(19)
in which r = circulation at ~ = 0; ( = eddy viscosity; Vb = transverse
= downstream distance
velocity at distance h from the vortex line; and
along the vortex line. At ~ = 0, vb ~ u b tan G', and (19) is reduced to

(20)
in which u b = near-bed value of u. The transverse velocity vb determines
the transverse components of bed shear stress and sediment transport.
An alternative equation for vane-induced transverse near-bed velocity is
obtained by the transverse component of the momentum equation.
A
They
simplified solution was obtained by Odgaard and Spoljaric [19861.
assumed that: (1) the vane induced transverse velocity profile is tinear and
can be expressed as

~s = 2 [~-~]

(21)

in which Vs = transverse velocity component at the water surface; (2) the


eddy viscosity profile is parabolic,
(22)
in which u* = friction velocity:: /TTiJ, and
traflsverse variation of v s is given by

= bed shear stress; and (3) the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

135

Odgaard and Spoljaric


V

vsc cos

[1r dY] ; -

d 5 y 5 2"
d
2"

(23)

(in consequence of the helical motion induced by the vane) in which Vsc =
centerline value of vs.
With these assumptions, the downstream decay of
maximum vane induced near-bed transverse velocity component vb (= - vs ) is
-vb = (tan 0') exp
ub

in which x

[2K
~

+K

~]

(24)

downstream distance measured along the channel centerline.

Critical Design Parameters


The key to vane performance is the horizontal (lift) force that they
generate and its effect on near-bed flow and on circulation induced in the flow
downstream from vanes. These features determine the number and layout of
vanes of a given design required to alleviate the problem of either bank
erosion or shoaling.
Note that the required lifting surface is inversely
proportional to the lift coefficient c . In their original approach, Odgaard and
L
Kennedy [1983] used a lift coefficient given by

(25)
This relation is obtained by applying the Kutta condition [Sabersky and
Acosta, 1964; Bertin and Smith, 1979] to ideal flow around a large aspect ratio
(2-D) flat plate at small angle of incidence with flow. Equation (25) was
subsequently modified by Odgaard and Mosconi [1987] to read
C

21rO
= ---

+ !.
H

(26)

This equation was proposed based on Prandtl's lifting line theory for finite
The decrease of cL with decreasing aspect ratio, predicted by this
wings.
equation, is .due to the tip vortex (the vortex trailing the upper edge of the
vane), which induces a downward motion (downwash) in the fluid passing over
the vane. The downwash has the effect of turning the free stream velocity, so
that the effective angle of attack is reduced.
For a finite span foil in
undisturbed free stream flow, Prandtl, by assuming an elliptic spanwise
circulation distribution, determined the reduction in 0' to be given by (eLI If)
(L/H). As the riverbed vane is a "wall-attached" foil with only one edge with
tip vortex, Odgaard and Mosconi assumed that its effective angle of attack is
reduced by only half the amount determined by the lifting line theory, and,
The corresponding induced
hence that its lift coefficient is given by (26).
drag, cD' is

(27)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

136

Submerged Vanes

I
I

SIDE VIEW

I-

3.41 H

IIItOO73H I
I
I

_IT

I.

TOP VIEW

SIDE VIEW

2.15 H

IT

TOP VIEW

'. JH

I?l~

SIDE VIEW

~PH

I_

1~2IH

0.7 H

SIDE VIEW

II

2.35 H

I_

TOP VIEW

2.33 H

0.24 H

.PH

TOP VIEW

Fig. 4.

Vanes Tested.

Equation (26) is derived under the assumption that:


(1) the fluid is
inviscid, (2) the oncoming flow is uniform, and (3) the aspect ratio of the
vane is not small. In reality, vanes are placed in a fully developed, turbulent
boundary layer, and their aspect ratio is much less than one. By applying
improved numerical analysis, Spoljaric [1988] developed a solution for such a
condition. Her analysis was based on the vortex lattice method, a power-law
velocity profile, and the low Froude number approximation (which allows the
free surface to be taken as a rigid boundary). Calculations were made for
flow depths between two and four times vane height, which is the range for
which prototype vane systems are designed. Three different types of vanes
were analyzed: a vertical flat plate, a cambered foil with vertical surface, and
a cambered foil with twisted surface. The types are shown in Figure 4. The
cambered foils were selected because cambering is known to increase lift of
large aspect ratio wings in uniform free stream. The twisted foil was designed
to roughly conform to the theoretical design objective of increasing vane lift
force toward the bed (to counter the increase of centrifugal-force deficit).
Samples of Spoljaric's calculations are seen in Figure 5. They show that
c is a function of not only a and H/L but also of d/H and vane shape. The
L
variation of c with n is a result of the normalization with u. Angle 0'0 is
L
the angle between the flow. direction and the direction of the trailing vortex
line.
The calculations were made with 0'0 values selected based on
measurements. (For cambered foils, 0'0 was measured to be somewhat less
than a).
The lift coefficient increases as d/H decreases.
For example, for the
cambered, vertical vane set at angle of attack of a = 15 in a boundary layer
with n = 4, c increases from 0.49 to 0.72 when d/H decreases from 4 to 2.
L

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

137

Odgaard and Spoljaric


~

-l

+
-0
U

4 , . . . . - - - r - - r - - - r - - r - - r - - - - - - , ,....-.....,.---,....---r--r---r----, ,....--r---~--r--~_-----.

THIN, FLAT VANE

CAMBERED VANE W/O TWIST

CAMBERED VANE WITH TWIST

a =0

a o= a

ao=a

<t

d/H=

-l

d/H

2
---

1"

CD

---

d/H =

2
---

~ 2
z

LIFT

UJ

U
L;:

L.

~
u

1
DRAG

C)-

<t

a:
o
o
Z

<t

.~

-l

Fig. 5.

10

20

10

20

10

20

30

ANGLE OF INCIDENCE WITH FLOW, a (DEGREES)

Lift and drag coefficients computed by vortex lattice method. The


lift coefficient computed by lifting line method.

21rO'

line is the

This effect is due to increasing suppression of the tip vortex, and consequently
decreasing effect of downwash, as the top of the vane approaches the free
surface.
As with large aspect ratio wings, shape is an important parameter.
Cambering is seen to significantly increase lift, in particular at small angles of
attack. In addition, cambering decreases lift slope, making cL less sensitive to
angle of attack. This is of practical significance as vanes have to perform
under conditions in which flow direction can vary considerably.
Drag is also a critical parameter. As was shown by Odgaard and Mosconi
r1987], the induced drag has a direct influence on energy slope of the channel.
It is well known that even small changes in energy slope can have dramatic
effects on channel stability; therefore, induced drag must be kept to a
minimum. It follows that vane efficiency can be measured in the same way as
for wings, by ratio cL/c D . For large-aspect-ratio wings, cambering increases
this ratio.
According to Prandtl and Tietjens [1934, page 151], cambered
large-aspect-ratio wings at angles of incidence of 12 -15 have, for the same
drag, twice as large lift coefficient as uncambered wings. However, Spoljaric's
calculations show that cambering of the low-aspect-ratio vanes does not
increase their cLI CD ratio.
In (20), the critical parameters are f and h. The vane induced circulation
decays because of turbulence generated by both the channel resistance and the
wake behind the vane.
Channel resistance is predictable.
Assuming a
parabolic f profile (22), the depth averaged channel eddy viscosity is
0

(28)
However, there are no guidelines for determining the effect on decay of wake
induced turbulence. The same applies to the location of the trailing vortex
line (h and 0'0)'

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Submerged Vanes

138
Experiments

The experiments were conducted to test the validity of the aforementioned


design basis.
Although sophisticated numerical techniques are available for
calculation of the flow field around finite wings, their application to the vane
problem is uncertain. A low aspect ratio vane in free-surface flow induces a
very complex three-dimensional flow pattern.
Four different flow features
interact with one another: (1) a tip vortex induced at the top edge of the
vane, (2) a shedding of vortices at the rear end of the vane, (3) a horseshoe
vortex generated at the junction of the vane and bed, and (4) boundary-layer
separation on the vane. Because of the low aspect ratio, all of these features
dominate the flow pattern around the vane. Each of the features have been
studied in the past.
For example, the tip vortex, vortex shedding and
separation are classical topics in aerodynamics, and the horseshoe vortex is an
issue in studies of thick-body junctions (cylinders and struts attached to a
larger solid surface).
In aeronautics and naval engineering) advanced
techniques are available for determining details of each of these flow features
[Devenport and Simpson, 1986; Nakayama, 1985; El-Ramdy, 1975; Selic, 1975;
Levy and Shamroth, 1986; and Mansour, 1985]. However, there are no known
studies or reports on the flow around a bed mounted vane that is induced by
a combination of all four features.
The flow around a vane is further
complicated by effects of the free surface.

Test Set-Up
The schematic of the set-up is shown in Figure 6. The experiments were
conducted in an 8 m long, 61.5 cm wide, rectangular, glass-walled tilting
flume.
Discharge control was obtained by gate valves and orifice meter.
Uniformity of flow was achieved by baffles and screens at the inlet and by
slope adjustments using the tilting mechanism of the flume. Levels of water
surface and channel flow were measured with point gages.
The flume was provided with a series of 0.61 m by 1.22 m sand coated
sheet-metal panels that were inserted in the flume to form a false bottom
8 cm above the bottom of the flume. The sand on the panels was 0.3 mm
quartz sand. One of the panels had a 0.45 m by 0.35 m cutout in which a
frame supporting a sand coated plate was suspended such that it could move
freely in any direction in the horizontal plane. The plate surface was coplanar
with the surrounding panels, with a 1.5 mm clearance on all four sides. The
vane was mounted on this plate. With flow through the flume, the frame of
the vane mount exerted a force on a miniature 50 g compression load cell,
positioned for either lift or drag measurement.
The velocity data were taken with a small two-component electromagnetic
flow meter (the Montedoro-Whitney Model MVM-1) with range from 0 to
0.4 m/s. Positioning was computer controlled.

Flow Conditions
The design parameters (lift, drag, and induced circulation) are dependent
on HIL, dlH, blH, ll', Fr, Re, f, and vane shape. However, due to limitations
in the set-up, notably channel width and resistance, and to practical
constraints on HIL [Odgaard and Mosconi, 1987], the test variables were
reduced to dlH, ll', Fr, and vane shape.
Depth and velocity were varied within the ranges 10-25 cm and
10-30 cmls, respectively.
The friction factor ranged from 0.016 to 0.020,
which corresponds to a power-law exponent n from 8 to 9.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

139

Odgaard and Spoljaric


LIFT

DRAG
CHANNEL WALL
(GLASS)

II:Iml

'.

Il!ml

SUSPENDED
VANE MOUNT

FLOW

VANE

~----~--~~

LOAD CELLS

TOP VIEW

ELEVATED CHANNEL BOTTOM


WITH ARTIFICIAL ROUGHNESS

WATER SURFACE

FLOW

LOAD CELL

CHANNEL BOTTOM
SECTION A-A

Fig. 6.

Schematic of set-up for lift and drag mesurements.

Lift and Drag


Figure 7 shows measured lift and drag coefficients for the aforementioned
three vane types.
Each point in these diagrams is the average value for
velocities within the range 10 to 30 cm/s (individual values were within 10%
of the average values). The most notable finding is that, with the exception
of the thick, flat vane, the data are in relatively good agreement with the
lifting line formula (26). This is quite surprising considering the complexity of
the flow around the vane. However, the agreement confirms that ll' and H/L
are the primary variables for the lift coefficient.
Cambering and twist
obviously do not improve the lift characteristics of the vane.
This is
explained by boundary-layer separation around the vane, which was observed
(by dye injections) to be more pronounced around cambered vanes than around
thin, flat vanes.
The theories do not account for this feature.
Flow
separation reduces the pressure difference between pressure and suction side
and thus reduces lift.
The twisted vane also generates a vertical lift
component; however, this could not be measured in this set-up.
Figure 7d shows lift coefficients measured on a flat vane with thickness of
about 21% of vane height, about the same as that of the prototype vanes in
the East Nishnabotna installation [Odgaard and Mosconi. 19871. This vane is
clearly less efficient than any of the other shapes tested. Tne cL values for

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Submerged Vanes

140

(0)

THIN, FLAT VANE

CAMBERED VANE WITH TWIST

d/H =

d/H=

3
4
\J 2-4

0/' 0

LIFT

/ff

<t

I......

Ji

0/'0

(27)

/'

...J

/'

1 -

Ji'

....J

"-7

I>

.......
o

/- 3

;:
}/

:J

~6/ 3 ~ ~ ~

tt-,

/'

- 2

/~

(c)

en

0 ....
.-.:::;;;;.......L....-_.....&.....-_..I.....-----'_---'-_----' 0 '

4 .----____..-----r-----".....--------.-----.-----, 4 .----~I-~--I.--------..I---r--'---. 4

UJ

d/H =

UJ

o
u

THICK, FLAT VANE

CAMBERED VANE W/O TWIST

I.L
I.L

o
6

d/H =

2
3

. /

(!)

<t

0::

o
o
Z

<t
~

I.L
....J

2 ,....

/'

/'

/'

If-

1 -

/'

(b)
I

10

20

/'

0 ......
' 30 0

"-,,~/'

o
- 1

(d)

1-----'-'--'''-------'-,--"'''''---,----' 0

.......

ANGLE OF INCIDENCE WITH FLOW,

Fig. 7.

10

20

30

(DEGREES)

Mesured lift and drag coefficients. The 21rll' line is the lift coefficient computed by
the lifting line method (26).

this vane are 25 to 40% lower than those predicted by (26). Increase of vane
thickness causes an increase of the size of the separation zone on the rear side
of the vane and thus a decrease of lift. By comparing Figures 7b, 7c and 7d,
it would appear that if, for practical reasons, the vane must be made with
thickness, it should be airfoil shaped.
The measurements confirmed that cL increases with decreasing dill ratio.
For example, for the thin, flat vane at an angle of attack of 20, cL increases
from a value of 0.45 at dill = 4 to a value of 0.64 at dill = 2.
As
mentioned earlier, this trend is due to suppression of the tip vortex by the
free surface, which results in an increase of the pressure difference between the
vane surfaces.
The measured drag coefficients are seen to agree with those predicted by
(27). There was no measurable dependence of CD on dill.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

141

Odgaard and Spoljaric

The data do not suggest that a Froude number dependence is present


under the flow conditions tested (dill > 2, and Froude number less than
0.25). Note that the vortex lattice method applied above is based on the low
Froude number approximation, and, therefore, does not predict any dependence
on Froude number.

Velocity Measurements
As mentioned earlier, one of the main purposes of determining the velocity
pattern downstream from the vanes is to obtain guidelines for the layout of
vane systems. llaving determined, based on lift coefficient, the number of
vanes required to eliminate or generate a certain secondary circulation within a
given control volume, the next step in the design is the determination of the
pattern in which vanes are to be placed (lateral versus longitudinal spacing
and alignment along vortex lines). For that, it is necessary to have means to
estimate the area of influence of each vane, rate at which induced circulation
decays downstream from a vane, and direction of trailing vortex.
Figure 8 shows samples of measured profiles of streamwise and transverse
velocity downstream from vanes set at an angle of 15 with the channel axis.
The complete set of data is in Spoljaric's thesis. The profiles clearly show
that the flow around the vane is dominated by vorticity formed at the top
edge.
The change in sign of transverse velocity and the depression of
streamwise velocity are seen to occur slightly below the top level of the vane
in consequence of the acceleration of flow over the vane. This observation
suggests that (24), which is based on the assumption that the induced vortex
is centered at mid-depth, is inadequate for a description of vane induced
circulation.
This is also indicated in Figure 9, which shows transverse
near-surface and near-bed profiles of v.
For dill = 3, the transverse
velocities near the surface are significantly weaker than those near the bed.
The profiles in Figure 9 show that the transverse extent of the vortex is less
than 2 flow depths.
Figure 10 shows the lateral position of the trailing vortex in downstream
sections, as determined by the points of maximum depression in streamwise
velocity. The path delineated by these points determines the location of the
downstream vanes which have to pick up and reinforce the circulation
generated by upstream vanes. The direction of the path, described by angle
ll'o with the channel axis, was found to depend on ll', dill and vane shape.
For vanes with vertical surfaces and dill = 2, the path was nearly in the
direction of flow (ll'o = 0). For dill = 3, the vertically faced vanes deflected
the vortex line toward the direction of the vane; ll'o approached ll'/3 for ll' ~
15 . For the twisted vane, ll'o was significant for both dill = 2 and 3. The
dependence of ll'o on dill is related to the acceleration of the flow over the
vane. When dill = 2, the velocitx of the flow increases more as the flow
passes over the vane than when dIll = 3. As the vane induced transverse
velocity component (for given ll') is nearly the same in both cases, the
resultant velocity vector above the vane is less deflected when dill = 2 than
when dill = 3. This information on ll'o was used also for the aforementioned
numerical lift and drag calculations, which are sensitive to ll'o.
The vertical position of the trailing-vortex core varied slightly with dill
and vane shape. The vortex of the twisted vane was generally closer to the
bed than that of the other vanes (Figure 11), probably because of the larger
zone of separation on the suction side. An overall average of hIll of 0.8 was
indicated.
The velocity data were used also to determine eddy viscosity land
circulation around the vane r o. The eddy viscosity determines induced drag
By using the
and decay of circulation.
Circulation r 0 determines lift.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Submerged Vanes

142
0.6

0.8

1.0

1.2

-0.5

0.5

...---r---....-------j....----.----.,...._....--.......-_....--_._-......-_._---.-........-----..----.
y/d

aid

24

THIN. FLAT VANE

---0--- 48
_-&-7.1

0
0

d/H=2

.--0--- 119

---..._. 190

BOTTOM OF FLUME

'.

,
---0---

20
40

.......

15.8

._.-{:r._. 5.9
9.9

--::>---

...... -

0.3
0.3

a
y/d

0
0.2

>

'--6--

1.6
3.2
48

0.5

CD

7.9
--~-..._._..... - 12.7

05
0.5

W
CD

--0----

---0---

<x:

SZ

~/

J:

f/

BOTTOM OF FLUME

d/H=3

SZ

(>

0
--0--

---0--._.~._.

aid

y/d

1.3

2.6

0.1
0.4

CAMBERED VANE
WITH TWIST

40

0.5

d/H' 3

6.6

0.8

SZ

THIN. FLAT VANE

SZ

BOTTOM OF FLUME-

CAMBERED VANE
WIO TWIST
d/H=2

02
03

aid

....J:

y/d

SZ

a
aid

--0--

J:
"-

sz

SZ

;' BOTTOM OF FLUME

TOP.~~.~ .. _

0
0.6

0.8

1.0

BOTTOM OF
FLUME

1.2

DOWNSTREAM VELOCITY COMPONENT. u/u

Fig. 8.

LBOTTOM OF FLUME

0
-0.5

0.5

TRANSVERSE VELOCITY COMPONENT, v/u

Streamwise and transverse velocity components downstream from vanes at 15 degree


angle of incidence with the flow. The profiles were taken near the trailing-vortex
core.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

143

Odgaard and Spoljaric


-05

iii

03[
02 -

iii

iii

05

iii

iii

iii

THIN. FL.\T VANE

dlH.~.~
'. ... ~.~:~

01

iii

rlH1

-0.5

~781 [

05

~1~i~f~i~i~i~i~~~~~~~i~l~i~
CAMBERED VANE
WIO TWIST

02

d/H'2

1
01

..

~
w

-01

-0.1

-02

-0.2

-03

- - 0 - 2.4

---0--_.~._.

-0.4

--0--

---0---

20

48
71

._.~._.

40
5.9

---0---

---0---- 11.9

ct

I-

t1M-0281

.19.0
I

it

-04

99

-05

-0.5

-03

........__.. 158

!:J THIN':/~~~"H"57.~[
o

03

r1H'1623

iii

..J..'....L'--L,...'.....
1 ...JI........I......I.-I..-.I..-'--I..-.I..-'--~....L!--L,...I.....
' ...JI---JI

L..'

iii

iii

iii

CAMBERED VANE
WITH TWIST

0.3

z/H2.392

0.2

d/H'3

1
0.1

-0.1

-0.1

-0.2

-0.2

-0.3

1.6
---<)--- 3.2
48

-0.4

---0---

--~-1.9

-0.5

---0---

ZIHO.281

---._12.1

-0.5

-0.3

- - 0 - - 1.3
26
_--6-_40

--0-

._~.-

0.5

,/H.0.309

6.6

to.
J. .

O.5

-0.5

'"

0.5

TRANSVERSE VELOCITY COMPONENT, v/u

Fig. 9.

Transverse near-surface and near-bed profile of v downstream from vanes at 15


degree angle of incidence with the flow.

Kutta-Joukowski theorem, and thus assuming that r 0 is constant along bound


vortex lines, lifts are calculated which are in good agreement with those
measured [Spoljaric, 1988]. With the exception of a small region immediately
downstream from the vanes, the f values calculated from the velocity data,
and plotted in Figure 12, are essentially equal to those obtained by (28); that
is, vane induced turbulence has little effect on the decay of circulation. In
Figure 13, measured maximum near-bed transverse velocities downstream from
the vanes are compared with those computed by (20) using h/H = 0.8 and
l = lo = (",/6) u*d.
The agreement is seen to be reasonably good.
Conclusions

The laboratory experiments showed that (26) and (27) provide reasonable,
gross estimates of horizontal lift and drag on thin, flat, small-aspect ratio

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Submerged Vanes

144
a

d/H

(DEG)

15

20

0.6
TH IN, FLAT VANE

.......
:>.

0.4

z
:J

Q::

L&J

0.2

IZ

L&J
U

:E
0
Q::

lL&J

0.8

I-

en 0.6

0
L&J

en 0.4
Q::

WITH
TWIST

L&J

>
en
z
0.2
Q::

I-

14
01 STANCE DOWNSTREAM FROM VANE, x/d

Fig. 10.

Transverse shift of vortex core as determined from depression in streamwise velocity


profile.

riverbed vanes at angles of incidence up to about 20. Although low Froude


number wing theory predicts that cambering of the vane should increase lift,
the experimental data show little improvement.
This is explained by
boundary-layer separation around the vanes, which is more pronounced around
cambered vanes than around thin, flat vanes. Vane thickness increases the
size of the separation zone on the rear side of the vane which causes a
decrease of lift. Flat vanes with a thickness comparable to that of the vanes
used in the East Nishnabotna installation have lift coefficients 25 to 40% lower
than those predicted by (26).
The theory predicts a significant increase of lift with decreasing d/H ratio
within the range 2 5 d/H 5 4 (the range for which vane systems are
designed). This increase is due to suppression of the tip vortex by the free
surface, which results in an increase of the pressure difference between the
vane surfaces. The experimental data confirmed this trend. The difference
between theory and experiments again is attributed to boundary-layer
separation, which is not accounted for in the theory.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

145

Odgaard and Spoljaric


d/H

VANE TYPE

THIN, FLAT

CAMBERED WIO TWIST

CAMBERED W TWIST

1.4

a = 15 0

:I:
.......
oS:;

W 1.2

0::
0
U

1.0

0::

> 0.8
lL.

0
~

:I:
~

0.6

w
:I:

0.4

10

12

14

DISTANCE DOWNSTREAM FROM VANE, x/d

Fig. 11.

Elevation of vortex core as determined from transverse velocity profiles.

d/H

(DEG)

15

20

\J
0

THIN, FLAT VANE

\J~

>t:

--~--tr--1n----t~----r--'

~ 1.0
u

en

>
>o
o
UJ

O--........------......."""'-.......... - -......- .......

-.jlooooo-.......- - - - . j - - - - . . .

10

12

14

16

DISTANCE DOWNSTREAM FROM VANE, x/d

Fig. 12.

Eddy viscosity computed from measured velocity profiles downstream from vane,
normalized with no-vane channel viscosity, fo.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

146

Subnlerged Vanes

o
1.0

d/H =2

t - - -.......~

0.8

a
c

C
-.0
::J

0.6

">

.0

r-

0.4

G
o

0.2

r-

-l
W

MEASUREMENTS;
THIN, FLAT VANE
CAMBERED VANE WIO TWIST

!:J.
o

>

CD
I

a::

<{

en
a::

1.0

>
en
z

0.8

d/H

!:J.

=3

<{

0-

a::
t-

:E
:::>
:E

0.6

<{

0.4

MEASUREMENTS;
0.2

!:J.
o

THIN, FLAT VANE


CAMBERED VANE W TWIST

O--........-.................- - - - - - I - - - - - I _ I . . . . . - . . I . . . -.......................L.........L..---L.----L----I
o
2
4
6
8
10
12
14
16
DISTANCE DOWNSTREAM FROM VANE, x/d

Fig. 13.

Measured and predicted downstream variation of maximum transverse near-bed


velocity component.

The theory applied is based on the low Froude number approximation, and,
therefore, does not predict any dependence on Froude number. The data do
not indicate that a Froude number dependence is present under the flow
conditions tested (d/H > 2 and Froude number less than 0.25).
The velocity data show that the transverse extent of vane induced
circulation is generally less than about 2 flow depths. This suggests that the
lateral spacing of vanes in a layout should not exceed 1.5 to 2 flow depths if
the layout is to produce a common circulation. The longitudinal spacing of
vanes depends on the rate of decay of induced circulation. This decay was
found to be described reasonably well by (20), with the eddy viscosity based
on channel flow data and a value of h/H of about 0.8. Vane alignment is
another critical design parameter. As downstream vanes must pick up and
reinforce the circulation induced by upstream vanes, they must be aligned in
reference to the trailing-vortex line produced by upstream vanes.
The

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

147

Odgaard and Spoljaric

experiments showed that vanes with vertical surfaces generally produce a


vortex line in the direction of flow when d/H = 2, whereas the line tends to
drift in a direction which approaches 0:/3 when d/H ~ 3.
The different
behavior of the vortex line at different d/H ratios is due to differences in
acceleration of the flow over the vane.
In the East Nishnabotna installation rOdgaard and Mosconi, 1987]' the
vanes were installed with 0:0 = o. As these vanes are vertical, flat plates
designed for d/H = 3 and 0: = 15, the aforementioned findings suggest that
their line-up is incorrect. Moreover, as its design is based on (26) and vane
thickness is considerable, the installation is probably inadequate.
As a final comment, it is emphasized that the design basis discussed is for
vanes on a horizontal, rigid bed; and it was tested herein under the same
conditions.
In reality, vanes are placed on movable beds, and their
performance is undoubtedly strongly dependent on bed forms such as dunes
and ripples. Studies are currently underway in which their performance on
movable beds is determined.

Applications

River-bend bank protection. Design a vane system for a 200 m long, 50 m


wide channel segment.
The centerline radius of curvature is 200 m, and
design (bankful) depth and discharge are 3 m and 200 m 3 /s, respectively.
Channel slope is 0.0005.
The design relations are (6) and (26). Substituting (26) into (6) yields
NHL = L.
f1b
0:1r

[1 + ~]
H

~
r

(29)

The bend angle, ;, is approximately 1r/3, and the area of the segment is
r;b = 10,472 m 2 As the design velocity is fi = 200/(bd) = 1.33 mIs, the

velocity-profile exponent is n = "'fi/ /i,Sd = 4.4.


Therefore, H
Required vane area, NHL, decreases when H/L increases.
should be as large as possible, but less than 0.5 d, where d = depth of
erosion causing flow. Assuming that an appropriate value is H = 1 m, (29) is
reduced to
NHL

= a25

[1

L]

+ Ii (m2)

(30)

The lateral vane spacing, fly, must be less than 2d. By selecting fly = 5 m,
there are 9 vanes in each lateral array. The streamwise spacing, flx (measured
along the centerline) is then determined by
flx

200

(31)

Nt+ 1

in which

Nt -- aHL
2. 78 [1

L
-H]

Copyright American Geophysical Union

(32)

Water Resources Monograph

River Meandering

Vol. 12

Subluerged Vanes

148
Table 1 shows values of

/:U{

for different selections of a and L.

Table 1. Streamwise vane spacing, /).x (meters).

(deg)
10
15
20

L/H
2

8.0
11.8
15.4

9.0
13.2
17.2

9.6
14.0
18.3

Hence, if 4 m long, 1 m high vanes are selected and set at an angle of 20


degrees, a total of 90 vanes are required to eliminate the secondary current.
(Note that total elimination of secondary current across the entire section may
not be necessary. Bank erodibility should be considered as well.) The vanes
should be aligned with reference to a vortex line at an angle of a/3 with the
channel centerline.

Shoaling reduction. Design a vane system for depth control in a 100 m


wide, straight channel. The system is to deepen the central portion of the
cross section by producing two symmetric, counter-rotating secondary currents
with near-bed velocity directed outward toward the banks. Design depth and
Channel slope is 0.0005.
discharge are 3 m and 300 m 3/s, respectively.
Assume that the sediment size is such that the objective is achieved when the
system generates an overall average vb of 0.15 u from the centerline toward
b
the banks.
The design relation is (20). With a design velocity of fi = 300/(bd) =
1.0 mIs, the velocity exponent is n = 3.3 and eddy viscosity Co = lrn*d/6 =
,,2fid/6n = 0.024 m 2/s.
The vortex line is at h/H = (h/d)(d/H) = 0.8.
, Assuming aIm high vane (20) is reduced t.o

Vb/(U b tan a)

1 - exp [-2.22 (d/x)]

(33)

The streamwise distance between vanes, ~, is defined by


6x

0.15 ub =

h f d[a]
o

Vb

yielding the distances listed in Table 2


Table 2

ax

10
15
20

6.3
20.4
36.6

(deg)

(m)

Copyright American Geophysical Union

(34)

Water Resources Monograph

River Meandering

Vol. 12

149

Odgaard and Spoljaric

The lateral distance between vanes l:1y should be less than 1.5 d. A distance
of l:1y = 4 m is appropriate.
It follows that each lateral array should have about 11 vanes on each side
of the centerline. If the vanes are angled at 15 with the centerline, there
should be about 20 m distance between each lateral array. The vanes should
be aligned with reference to a vortex line at an angle of 0:/3 with the channel
centerline.

Acknowledgement
This material is based upon work supported by the National Science
Foundation under Grant MSM-8611147.

Notation
b
cn
cL
d
f
fc
F

channel width
drag coefficient
lift coefficient
flow depth
Darcy-Weisbach friction factor
function (10)
function (7)

Fc
Fv
g
h
H
L
n
N
p
r
S
Tc

integral of fc
vane lift force
acceleration due to gravity
height of vortex core above bed
vane height
vane length
velocity-profile exponent, equal to K,{8/f) 1 12
number of vanes
pressure
radius of curvature
downstream slope of water surface
centrifugal force torque

Tv
u
fi

v
x

vane-induced torque
velocity component in downstream direction
average over depth of u
velocity component in cross-stream direction
downstream coordinate

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Submerged Vanes

150

y
z

cross-stream coordinate
vertical coordinate
downstream distance along vortex line

vane angle of attack


von Karman's constant (~ 0.40)
fluid density
eddy viscosity

e
K,

P
i

bend angle
circulation

initial, outer
inner
surface
bed

r
Subscripts

s
b

References
Bertin, J. J. and M. L. Smith, Aerodynamics for Engineers, Prentice-Hall,
Inc., Englewood Cliffs, N.J., 1979.
Chabert, J., M. Remillieux, and I. Spitz, Application de la circulation
transversale a la correction des rivieres et a la protection des prises d'eau,
Proceedings of the 9th Convention IAHR, Dubrovnik, Yogoslavia, pp. 1216-1223,
1961.
Devenport, W. J. and R. L. Simpson, Some time-dependent features of turbulent
appendage-body juncture flows, XVI Symposium on Naval Hydrodynamics, July,
Berkeley, 1986.
EI-Ramdy, Z., Investigation of the development of the trailing vortex system
behind a swept-back wing, Report No. MElA 75-3, Department of Mechanical
and Aeronautical Engineering, Carleton University, Ottawa, Canada, 1975.
Falcon, M. A. and J. F. Kennedy, Flow in alluvial-river curves, J. Fluid Mech. 133,
1-16, 1983.
Jansen, P. Ph., L. van Bendegom, J. van den Berg, M. de Vries, and A. Zanen,
Eds., Principles of River Engineering, Pitman Publishing Ltd., London, England,
1979.
Levy, R. and S. J. Shamroth, Numerical analysis of the viscous flow field resulting
from a hull-sail interaction, XVI Symposium on Naval Hydrodynamics,
Berkeley, July 1986.
Mansour, N. N., Computation of tip vortex off a low-aspect-ratio wing, AIAA
Journal, 23(8), 1985.
Nakayama, A., Characteristics of the flow around conventional and supercritical
airfoils, J. Fluid Mech., 160, 155-179, 1985.
Odgaard, A. J. and J. F. Kennedy, River-bend bank protection by submerged
vanes, J. Hydraul. Eng., 109(8), 1161-1173, 1983.
Odgaard, A. J. and C. E. Mosconi, Streambank protection by submerged vanes, J.
Hydraul. Eng., 113(4),520-536,1987.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

151

Odgaard and Spoljaric

Odgaard, A. J. and A. Spoljaric, Sediment control by submerged vanes, J. Hydraul.


Eng., 112(12), 1164-1181, 1986.
Potapov, M. V. and B. A. Pyshkin, Method Poprechnoi Tsirkulyatsii i ego
Primenenie v Gidrotekhnike, (Method of Transverse Circulation and its
Application in Hydraulic Engineering), Akademyia Nauk SSSR, Moskow,
Leningrad, USSR, 1947.
Potapov, M. V., Sochineniya v trekh tomakh - I, (Collected Writings in three
volumes - I), Gasudarstvennoe Izdatel'stvo Sel'skokhozyaistvennoi Literaturi,
Moskow, USSR, 1950.
Potapov, M. V., Sochineniya v trekh tomakh - II, (Collected Writings in three
volumes - II), Gasudarstvennoe Izdatel'stvo Sel'skokhozyaistvennoi Literaturi,
Moskow, USSR, 1951.
Prandtl, L. and O. G. Tietjens, Applied Hydro- and Aeromechanics, McGraw-Hill,
New York, 1934.
Sabersky, R. H. and A. J. Acosta, Fluid Flow, MacMillan Publishing Co., Inc., New
York, 1964.
Selic, Z. R., Measurement of vortex properties during wing-vortex interaction,
ASRL TR 178-3, Aeroelastic and Structures Research Laboratory,
Massachusetts Institute of Technology, Massachusetts, 1975.
Spoljaric, A., Mechanics of submerged vanes on flat boundaries, Ph.D. dissertation,
Univ. of Iowa, Iowa City, 1988.
U.S. Army Corps of Engineers, The Streambank Erosion Control Evaluation and
Demonstration Act of 1974 Section 32, Public Law 93-251, final report to
Congress, Waterways Exp. Stat., U.S. Army Corps of Eng., Vicksburg, MS,
1981.
Zimmermann, C. and J. F. Kennedy, Transverse bed slopes in curves alluvial
streams, J. Hydraul. Div. Am. Soc. Giv. Eng., 104 (HY1), 33-48, 1978.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Copyright 1989 by the American Geophysical Union.

Analysis of a 2-D Bed Topography Model For Rivers


Nico Struiksma
DELFT HYDRAULICS,

and

The Netherlands

Al~sandra

Crosato

Delft University of Technology, The Netherlands


Summary
From an analysis of a time-dependent 2-D model for river bed topography,
results are obtained which deepen the understanding of the processes of river
bed deformation.
With a non-steady state analysis the occurrence and behavior of
propagating alternate bars are described.
Due to the relatively large
propagation velocity of these bars, this type of bed perturbation cannot give
an explanation for the much more steady meandering process, which is
characterized by the point bar-pool configuration.
A steady state analysis
turns out to be more appropriate to describe the meandering process.
In terms of wave length and longitudinal damping rate, this analysis
provides a good description of the phenomena involved.
For conditions
prevailing in meandering rivers. it appears that the result of the interaction
between water and sediment motion depends on the ratio of two characteristic
adaptation lengths which govern the two independent equations for the flow
and for the bed deformation, respectively. In addition, it is shown that the
degree of non-linearity of the sediment transport with flow velocity is also an
important parameter.
Finally, the results of the analysis are compared with data from a straight
flume experiment with movable bed, in which at the inflow a steady
perturbation was imposed, and with data from a curved flume experiment with
movable bed and fixed banks.
Introduction
An important class of rivers exhibit a relatively stable meandering
planform. The main characteristic of the bed deformation in these rivers is
the point bar and pool development in the inner and outer bend, respectively.
As early as the 1940s, the Dutch engineer Van Bendegom [1947] studied the
phenomena of this typical bed deformation and was most probably the first
who produced, starting from horizontal bed, a complete 2-D bed topography
computation in a river meander.
However, due to lack of appropriate
computer facilities at that time he never repeated this.
Nowadays, with a variety of number crunching machines, substantial
progress is made in this field.
At DELFT HYDRAULICS, for instance, a
computer program has been developed which makes it possible to simulate the
153
Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

154

2-D Bed Topography NIodel

1 km

prototype 1966

computa tion

- - -

adopted bank alignment


discharge : 2000 m 3 / s

CONTOUR LINES OF WATER DEPTH IN m


Fig. 1.

Example 2-D bed topography computation for the Rhine Branches in the
Netherlands.

time dependent 2-D bed topography evolution in rivers. The reliability of this
model has been proved several times [Struiksma et al., 1985; Struiksma, 1985].
An example of a result from this model is given in Figure 1.
Apart from the possibilities to compute the bed topography in the case of
fixed banks, which is important for, e.g., river engineering, an interesting
spin-off is a better understanding of the phenomena which govern the typical
bed deformation in meanders. This paper will focus on this aspect.

Mathematical Model
Generally for the computation of time dependent river bed deformation, a
quasi steady approach is followed, Le., an interaction between a steady water
motion and an unsteady bed evolution is considered. This approach is widely
accepted for morphological computations in flows with a small to moderate
Froude Number [Jansen, 1979]. It still allows for a varying discharge: the
regime can be modeled by a stepwise approximation of the discharge
hydrograph so that during one computational time step the discharge is kept
constant. In addition, it is assumed that:
- the banks are non-erodible,
- the vertical profile of pressure is hydrostatic,

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

15.5

Struiksma and Crosato


-

shallow flow approximations are appropriate,


the spatial variation of the hydraulic roughness (Chezy coefficient) can
be neglected,
the rate of sediment transport is determined by local conditions
(dominant bed load), and
the influence of grain sorting is insignificant (uniform bed material).

Fig. 2.

Coordinate system for depth averaged flow.

Flow Model
The two-dimensional depth averaged momentum and continuity equations for
steady shallow flows, using the Cartesian coordinate system depicted in Figure
2, read:

Ou
Ou
0
1
u OX + v oy + g OX (h + zb) + pn T bx

10
=P
OX

(hT xx) +

10
p
oy

(hT xy) (1)

Ov
u OX +

10
=P
OX

(hT xy) +

10
p
Of

(hT yy ) (2)

Ov
Voy
+

0
oy

(h + zb) +
Ohu

7JX

1
pn

Tby

Ohv oy-

(3)

in which u and v = depth averaged velocity components in the x and y


direction, respectively; h = water depth; Zb = bed level; Tbx and Tby = bed
shear stress components in the x and y directions, respectively; g =
acceleration due to gravity; p = mass density of the fluid; and x and y =
spatial coordinates. The factors T xx, T yy and T xy account for the horizontal
momentum exchange due to viscosity, turbulence and non-uniformity of the
vertical velocity distribution. The third among these is dominant in a curved
flow because of the spiral motion. It is called secondary flow convection and
in river bends it results in a shift of the flow lines toward the outer bend [De
Vriend, 1981]. For the prediction of flow pattern and bed topography, these
factors have to be incorporated. However, they will be neglected hereafter for
the sake of simplicity.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

156

2-D Bed Topography Model

The components of the bed shear stress in the momentum equations are
related to the depth-averaged velocity using Chezy's relation:

Tbx

TbY

= pgu

(u2

= pgv

(u2

v2 )1/2

(4)

C2

v2 ) 1/2

(5)

C2

in which C = Chezy's coefficient. According to (4) and (5), the direction of


the bed shear stress coincides with the direction of the depth averaged
velocity, which is only true for noncurved flow. However, it is assumed that,
for curved flow, the influence of this deviation on the main flow is negligible.
Important for the motion of the sediment along the bed is the expression
for the direction D of the bed shear stress, which includes the effect of the
spiral motion [/(och and Flokstra, 1980]:

8=

arctan

[~]

- arctan

[A ~J

(6)

in which A = a coefficient which weighs the influence of spiral motion, having


the followinll expression: A = 2(~- 2 (1 - Ji/ ~C) when the velocity profile is
logarithmic lJansen, 1979]; ~ = Von Karman constant; and ( = calibration
coefficient.
Equation (6) was derived for locally fully developed spiral flow with R*
denoting the effective local radius of curvature of the streamline. To compute
R* the inertia of the spiral motion is taken into account, introduced by the
equation [compare Rozovskii, 1961]:

A 01

r~

= !!.
R

(u 2

v2 )1/2 with Ar

= f3 ~
~

(7)

in which Ar = adaptation length of spiral motion; {3 = a given coefficient; s =


the streamwise coordinate; R = local radius of curvature of the streamline
which is determined from the computed flow field; and I = a measure of the
intensity of the spiral motion. The effective radius of curvature R*, in (6) is
now defined by:
R*

h (u 2

+
I

v2 )

1/2

(8)

De Vriend [19811 proposes a value of about 1.3 for the coefficient {3, noting,
however, that the transverse bed shear stress will adapt faster to changing
curvature than the intensity of the spiral motion.
Therefore, it is
recommended to use {3 = 0.6 lalso see Kalkwijk and Booy, 1986].
Sediment Motion Model
In the sediment motion model it is assumed that the effects of suspended
load and grain sorting are insignificant. Then the development of the bed
level Zb can be described by the continuity equation for the sediment:

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

157

Struiksma and Crosato

~
+ 7JX
lJS x + ~
iJt
iJy

= 0

(9)

in which Sx and Sy = components of the volumetric sediment transport,


including pores, per unit length in the x and y direction, respectively. These
components can be expressed as:
Sx

Sy

= Se

(10)

Se cos a

and
sin a

(11)

in which Se = total effective volumetric sediment transport, including pores,


per unit of length, and a = direction of transport.
For the computation of the effective transport, the influence of the bed
slope is taken into account according to the following relation:
Se

St [ 1 -

e gC2

~]
iJs

(12)

in which e = calibration coefficient; and s = streamwise coordinate. The


transport St can be computed with various transport fomulas,e.g., Engelund
and Hansen [1967] and Meyer-Peter and Muller [1948].
The order of
is about 0.03 [Olesen, 1987].
magnitude of coefficient
The direction a of the sediment transport does not coincide with the
direction b of the bed shear stress. This is due to the gravity force acting on
the grains moving along a sloping bed. Along the lines of Van Bendegom
[1947J the following formula is used [see also Koch and Flokstra, 1980]:

. ~

tana =

S 1 nu

1 ~

- f(1!} iJy

1 ~

(13)

cos 6 - f[1J) iJx

in which f( fJ) = weighing function for the influence of the sloping bed; and
= the Shields parameter, defined by:

o=

u2

+ v2

2 ~

Dso

(14)

in which ~ = relative submerged density of the sediment; and Dso = the


median grain size.
For a review on the weighing function f( 0) reference is made to Odgaard
[1981].
In this paper the. following formula is adopted [e.g., see also

Zimmerman and !(ennedy, 1978]:

f( 0)

= O.~5 .[0

(15)

in which E = calibration coefficient.


The physical phenomena incorporated In the presented model can be
summarized as follows:

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

2-D Bed Topography Model

158

Vol. 12

depth averaged main velocity:


steady, inertia, bottom shear stress,
without horizontal momentum exchange;
vertical distribution of the main flow: logarithmic;
secondary flow intensity: including inertial growth and decay (Le., Ar f
0);
vertical distribution of the secondary flow: as in fully developed curved
flow with a logarithmic main velocity profile;
bed shear stress: magnitude and direction according to fully developed
curved flow without secondary flow convection;
magnitude of the sediment transport:
straight channel formula,
expressed in terms of the total bottom shear stress and corrected for
slope effects;
direction of the sediment transport: due to direction of bed shear stress
and transverse bed slope; and,
bed level:
time-dependent computation based on the local sediment
balance.

Linear Analysis

Introduction
The bed deformation can be explained on the basis of a first order linear
analysis of the mathematical model.
Furthermore, for the steady state
(azbl at = 0) such a first order analysis can provide a fairly good estimate of
the equilibrium bed topography in rivers with a simple geometrx.
Many scientists, among others Hansen [1967], Callander l1969], Engelund
and Skovgaard [1973], Parker [1976] and Freds0e [1978], have assumed that
alternate bars are the initial stage of meander formation in alluvial rivers and
carried out a stability analysis on unsteady uniform perturbations of the
channel bed.
However, considering the relatively high propagation velocity of alternate
bars, a non-propagating perturbation offers a better explanation for initiation
of meanders [Olesen, 1983].
For this reason the unsteady solution of the
present model will be adopted to investigate the alternate bars occurrence in a
straight channel, while the steady solution will be adopted to estimate the
equilibrium bed topography.
This can finally be related to initiation of
meandering.
Meander development was studied with another approach by Ikeda, Parker
and Sawai [1981].
They assumed meandering to be caused by a "bend
instability" and developed a model that includes bank erosion. Some years
later Blondeaux and Seminara [1985] developed a "unified bar-bend" stability
analysis and discovered a resonance phenomenon that is assumed to control the
bend growths. This phenomenon was found to occur when curvature "forces a
natural solution" represented by a uniform bed perturbation of the alternate
bar type having very low propagation velocity (it can be considered
approximately steady). The discovery of Blondeaux and Seminara is more or
less in accordance with the assumption of considering a steady perturbation to
investigate the meandering channel's bed topography.
Flow and bed topography are almost always significantly influenced by
transitional phenomena, see Struiksma et ale [1985] and Dietrich and Smith
[1983], which are due to the redistribution of flow and sediment, downstream
of any change of conditions.
Mathematically, this implies downstream
variations of the dependent variables, to be described with the steady linear
solution (provided that the non linear effects are negligible).

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

159

Struiksma and Crosato

The linear solution is found by considering first-{)rder perturbations of a


zero order solution. The critical perturbation is harmonic (wave length two
times the channel width) in the transverse direction and can be exponential,
harmonic or both in the downstream direction. The critical perturbation in
the unsteady linear solution is purely harmonic.
The steady linear analysis can also provide a strongly simplified version of
This was
the mathematical model rDe Vriend and Struiksma, 1983].
demonstrated to be a valid and simple tool for prediction of the bed
topography in alluvial channels.
In addition, it can be used as a basic
element in a meander migration model.
In order to facilitate the linear analysis, only straight channels are
considered. This simplification provides a simple zero order solution, but it
restricts the applicability of the results of the analysis.
However, the
phenomena involved are basically the same in a bend as in a straight reach,
so the conclusions of the analysis also apply to curved channels. In strongly
curved bends the spatial variation of the zero order solution may give rise to
some quantitative influence on the results of the first order analysis [Olesen,
1987].
The adopted coordinate system is cartesian, with the positive x-axis
The
downstream along the centerline of the channel, (see also Figure 2).
channel width is assumed to be constant.

Steady Flow Model


The basic flow equations are simplified by linearization with the further
assumption of a low to moderate Froude number which allows for a rigid lid
A second assumption of large radius of curvature-to-width
approximation.
ratio (mildly curved channel approximation) is necessary when a curved
channel, and not a straight one, is considered [Olesen, 1987, and Struiksma et

al., 1985].

Every quantity is assumed to be given by the sum of two terms, a zero


order .term, that corresponds to the reach averaged value of the quantity
(exept for water level), plus a first order perturbation term, e.g.:

= h o + h'; U = Uo + u ' , V = v', Zw = Zwo + zw / , etc.


(where Zw = h + Zb) in which the perturbations must comply with h' h o,
and Zwo = i xo + a constant, with ixo = water surface slope for normal flow
and z'w = hoe
h

The linear analysis of (1) through (3) leads to the following simplified
equations:

Zero order:
uo

= -...9...
Bh o

(16)

OZWO _

~ -

Uo

(17)

21\w

that corresponds to the Chezy relation, in which:


water discharge; and B = channel width.

AW = (C 2 ho )/2g; Q

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

2-D Bed Topography Model

160

Fig. 3.

Schematization of the channel's bed deformation.

First order:
I

1 IJuI + ~ 8z w + 1 u I
1 hi - 0
uo OX
u o OX
'X; uo - 2X; ~ I

IJvI

Dh '

uo OX

tJz w

7JY

1 Ou'

2X; uo
1

IJvI

~~+-"1J::-+-X;-=0
110 UX
UO uX
Uo vl

(18)
(19)
(20)

The perturbations are assumed double harmonic (see Figure 3), expressed by:

hi

= Ii

exp i(kx

kaY - ,t)

(21)

(the other quantities have a similar form) where the amplitude Ii is a complex
number of which the modulus is the amplitude and the argument is the phase;
k and ; are complex numbers; k is the complex wave number in the
longitudinal direction, the imaginary part of which describes the development
of the amplitude of the perturbation in the flow direction, the real part
represents the wave number. Hence:

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

161

Struiksma and Crosato

_ 21r
Re (k) - L'
with L = longitudinal wave length of the perturbation; kB is the wave number
in the transverse direction. Impermeable side walls put constraints onto it,
namely:

m1r
kB =B'
with m = 1,2,... (mode that determines the transversal pattern of the bed);
m = 1 in the case of meandering river.
Inserting the double harmonic perturbations into (18), (19), and (20) leads
to:

hk+Uk+Vk =0
Uo
Uo B

(22)

U . k
Uo 1

ur-

,gzw .

U 1
h 1
Uo 'X; - ~ 2X;

(23)
(24)

i =

The amplitude of the streamline curvature [1/R


1981] becomes:

=-

(l/u)( lJv/ ax); De Vriend,

(25)
Time Dependent Bed Deformation Model
The bed deformation equation is linearized in a similar way.
transport per unit width in flow direction is given by:
Sx

So

Sx with:

Sx ( So

The sediment
(26)

where

, = So [u'
c
b Uo + eg

Sx

Dh '] '
7Ji

which holds good for small streamline curvatures.


The sediment transport in transverse direction is given by:

Sy

Sy

Soa'

Zero order:
The sediment transport equation is approximated by:

Copyright American Geophysical Union

(27)

Water Resources Monograph

River Meandering

Vol. 12

2-D Bed Topography Model

162
b

(28)

So :: Uo

The exponent b originates from the linearization of the sediment transport


formulas, viz.:

aS

o
So ~

b - Uo
-

(29)

and is assumed to be a constant. It implies that the sediment formula is


approximated by a power law of the flow velocity (where b expresses the
nonlinearity of the sediment transport rate).
Adopting the Engelund &
Hansen transport formula b is constant and equal to 5, whereas adopting the
Meyer-Peter & M liller transport formula b is variable and depends on the
Shields parameter.
If the value of the Shields parameter lies close to its
critical value (incipient motion), the parameter b is large and strongly
variable. This is the case for most gravel bed rivers.

First order:
For small deviation angles of the bed shear stress and for relatively small
longitudinal bed slopes, the expression for tanO', (13), becomes:
tanO'

= u-

Ah
R*

Dh

+ f(1J) OY

where:

~:

AI

(u2 + v 2 )

(30)

112

which in linearized form becomes:


I

Q'

=Uo
VI

ARh*] ,

f(80 )

Dh'
OY

(31)

This is then coupled with (7) of which the linearized form, with the
assumption that the streamline coordinate s is coincident with the
x~oordinate, is:
\

1\ r

IJI' + I' IJv


"lJX""
- - h 0"lJX""'

wIth:

I'

~R =~;

UOllO

and Ar

= /3

Ch

_0

.[g

(32)

The conservation law of sediment in its linearized form is:

(33)
Considering the double harmonic character of the deformation (31), (32), and
(33) yield:

(34)
iAr

+ i = - i v hok

- I. h'" L
So

b
= I. -Uo

.. k

WI'th :-R = ~

~ C2 k2 h'"

~ -

UOllO

..
I. kBll'

Copyright American Geophysical Union

(35)
(36)

Water Resources Monograph

River Meandering

Vol. 12

163

Struiksma and Crosato

Combination with the linearized flow (22) through (24) only yields non-trivial
solutions if:

~ 1
+~
~_~_w

(bK2

+ 3 -~~
b) +
kB

(K

)1

bK3 _

DEN

i3A ~ K2 - A k

DEN

(1 +

1\
B

K3

(37)

K)

where:

DEN

= ~ (2 +
B

K - k
-

As

K2)

i (K

K3)

_ 2B

~ -

mL

= (m~)2

ho

[~]

f( 80 )

(38)

In the above relations ; is the complex celerity that is equal to zero for the
steady state solution.
The damping is governed by the first two terms of the right hand lnember
of (37), representing the influence of the bed slope on the sediment transport
direction and rate, respectively.

Unsteady State Solution


The unsteady linear solution (of Equation 37) is supposed to describe the
alternate bar characteristics (celerity, growth rate, wave length). The unknown
is the complex celerity;, that is determined for given values of the
longitudinal alternate bar wave length. Its real and imaginary parts represent
the propagation celerity and the growth rate, respectively. It is assumed that
the alternate bars that will develop along the straight channel are uniform and
characterized by the maximum growth rate.
In Figure 4 growth rate and
celerity are given as a function of the longitudinal wave length for the
conditions of the straight flume experiment described in the section on
Comparison With Measurements, Straight Flume Experiment.
In Figure 4, point P corresponds to waves that are not growing in time
and have a relatively small celerity (close to zero). Blondeaux and Seminara
[1985] discovered in this area a resonance phenomenon that is assumed to
control the bend growth.
The influence of curvature and the spiral flow adaptation, the latter
represented by the terms containing Ar, were shown to play an important role
For this reason it is not
in the alternate bar formation [Olesen, 1987].
possible to simplify the mathematical model for the unsteady state case any
further.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

2-D Bed Topography Model

164
5
...['J)

"-

lO

of max. growth rate

....-l

>.

+oJ

cu

C)

~-

....-l

cu
ro

+oJ

s...

..c
~
0

s...

bll

10

4
L (m)

-1

- 2

Fig. 4.

L--------l'--------l..

----:....-

......L....-_---L-_...L.....-_ _~

Growth rate and celerity of alternate bars as a function of their wavelength.

Steady State Solution


For ; = 0 (37) becomes a sixth degree complex polynomial equation, where
the unknown is the complex wave number k:

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

165

Struiksma and Crosato

+ (kA w)2

[1 + !2 (3A -

G) h o + At x; "X;

(b-3) At
2A;

- 2

e]

+ i(kA w) [(b;3) - ~ _~] - ~ = 0

(39)

where:
G

1
E
= f(iJ o) = 0.85 ~

For conditions prevailing in meandering rivers, among the six solutions of


the polynomial, generally four of them are purely imaginary and the other two
have the same absolute values in their real and imaginary parts, but their real
parts have opposite signs.
The general solution has the form:

h'

=h

exp i(kx

(40)

kBy)

For conditions valid for meandering rivers and along the side wall (y
B/2) the solution can be written in the form
h'

hte- ktx

h2e-k2X

hae-kax

h 4e-k4X

hse-k5X sink 6 (x

xp)

(41)

where ht, h2, etc. are amplitudes and Xp is the phase lag, (the other
parameters have a similar solution). The exponential solutions are generally
very strongly damped and are noticeable near the boundaries only, see Figure
5.
By assuming fJ = 0 (At = 0) and
= 0, the influence of secondary flow
inertia and the longitudinal bed slope are neglected in (39).
These
assumptions are justified when the wave length of the considered bed
deformation, L, is much longer than the relaxation length of secondary flow,
Ar, and when longitudinal sediment transport rate is independent of the bed
slope, which is also legitimate for relatively large wave lengths. With these
assumptions the polynomial equation becomes of fourth degree.
The solution of this fourth degree equation is very close to the one given
by the complete model. For this reason we can conclude that spiral motion
adaptation and longitudinal bed slope do not play an important role in the
phenomenon. For a deeper insight reference is made to Olesen [1987].

Steady State Solution: Strongly Simplified Model


A strongly simplified model is obtained by taking the streamline curvature
(see Figure 2) identical to the curvature of the channel center line [De Vriend
and Struiksma, 1983; Struiksma et al., 1985]. Hence:
complete model; Equation (39)
strongly simplified model

Copyright American Geophysical Union

(42)
(43)

Water Resources Monograph

River Meandering

Vol. 12

2-D Bed Topography Model

166

-05

-10J---------I~-M~----+-----~

-1.5L-

-----L..

....L.-

o
------+

---l

12

distance /width

total solu'tion
sum of two ind<2ntical harmonics
<2xponClntial parts
u~ and h~ orcz the steady disturbances
at thcz inflow

Fig. 5.

Representation of the longitudinal bed profile response to a disturbance at the


upstream boundary.

where R is the radius of curvature of


curvature of the channel centerline.
substitution of the harmonic disturbances
evaluated at the channel centerline (n =

au + -ruAw = ~
11
2Aw

""!J':'"

uS

uo
~
IlO

Rc
- -B'Jr ~1
S

the streamline, R c is the radius of


The linearized flow equation, after
(m = 1 in this case does apply) and
0), reads:
B
uo - ~
uo ( 2 2'Jr Aw

(J

[1 ]
-

Rc

(44)

in which (J = weighing coefficient of the secondary flow convection [De Vriend,


1981; Struiksma et al., 1985]. This secondary flow convection neglected in the
foregoing, is introduced here in a simple and approximative way, via the
source term. From the equation above it is easy to observe that Aw is the
governing parameter of the main flow adaptation. Henceforth it is called "the
main flow characteristic adaptation length."
The corresponding independent bed level equation can be derived from (20),
(31) and (33) for ah/at = 0 and without using the momentum equation of the
flow. It reads:

aii + X;
ii

7JS

= (b _ 1) ho
uo

au +

7JS

1r

h~B

Copyright American Geophysical Union

[L]
Rc

(45)

Water Resources Monograph

River Meandering

Vol. 12

167

Struiksma and Crosato

Similar to the main flow, the water depth behavior is governed by an


adaptation length, As, which is called "the bed characteristic adaptation
length. "
The combination of (44) and (45), leads to a system of second order
differential equations. The equation for the water depth is given below (the
one of the main flow is identical in its homogeneous part):

a2 li + ali

[1X; _(b:x3J]
2

+ li

A f( Oo~ hoB
1rAs w

[L]
R

7JST

os

(46)

The source term is a, function of local parameters only and is equal to zero in
the case of a straight channel. The roots characterize two identical growingdamping harmonic perturbations, yielding a solution for the water depth of the
form:

(47)
in which L = dampling length; L p = wavelength; and Sp = phase lag. The
n
solution for the main flow velocity is given by a similar expression.
In some cases the roots are purely imaginary, which implies that the
harmonic character of the solution vanishes. The complex roots mainly depend
on the ratio of the adaptation length of bed topography development and the
adaptation length of main flow, As/ Aw [Struiksma et al., 1985; De Vriend and
Struiksma, 1983]. The ratio is called interaction parameter.
Figure 6 compares the solutions of the complete model, (39), (far enough
The relative damping
from the boundaries), and of the simple model.
coefficient and the relative wave number are given as function of the
interaction parameter, As/ Aw.
In view of the good agreement between the two linear models for As/ Aw
larger than about 0.25, which is the case for most natural rivers, it can be
concluded that the effects included in the simple model; Le., the lateral
redistributions of the flow and of the sediment transport, are highly significant
for the bed development (wave length and damping) downstream of a local
The model of Engelund [1974], which also
bed and/or flow disturbance.
includes the redistribution effects, is, in principle, based on this approach and
it has been rather successfully applied to the experiments of Hooke [1975].
The linear analysis shows that the wave lengths and the damping rate of
the steady bed oscillations are quite sensitive to the changes of the exponent b
As already
derived from the selected transport formula, see Figure 7.
mentioned according to the Engelund & Hansen transport formula, b is
constant and equal to 5. However, according to the Meyer-Peter & Muller
formula, b is a function of the Shields parameter and can be relatively large
and strongly variable, when the parameter is close to its critical value
(incipient motion). So the value of b can be very uncertain.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

2-D Bed Topography Model

168
periodic range
2.5

a..

b=5

-oJ

:) 2.0
t=
C\I

"'0

1.5

c
0

-oJ

1.0

0.5

-0.5
0.1

Fig. 6.

0.2

0.5

1.0

5.0

2.0

10.0

Relative wave number and relative damping coefficient as a function of ~s/~w.


3,..--~-r---.....-----.

3,....------,..---,------,

~I a

t'(...J

-1....---+--1----+-----1

-2L.....-_.....L.-_.&..-..........._ - - - - I

0.2

0.5

1.0

2.0

------+ As lAw
damping c()(lfficiilnt

Fig. 7.

5.0

-2L.....-_....I...-_L.....----L._-01

0.2

0.5

1.0 2.0

5.0

------+ As I A w
WCNfl

numbczr

Variations of the relative damping coefficient and the relative wave number due to
changes of the exponent b.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

169

Struiksma and Crosato

Comparison With Measurements

General
The results of the foregoing analysis of the flow, the sediment transport
and the bed configuration (steady and unsteady case), (39) and (37) have been
incorporated in a computer program.
Some coefficients can be adjusted to
calibrate the model: l, {3, E and {. They weigh the influence of curvature,
spiral motion inertia, sloping bed (on the direction of sediment transport) and
longitudinal bed slope (on the sediment transport rate), respectively.
The simple model simulates the equilibrium bed topography (the steady
state case) only. It was incorporated in a computer program simulating the
change in planform of meandering rivers ICrosato, 1987]. For that purpose the
flow and bed topography model is coup ed with a bank erosion model. The
simple model is calibrated with the variation of three coefficients: l, E and (1.
They weigh the influence of curvature, sloping bed and secondary flow
convection, respectively.

Straight Flume Experiment


The computational results given by the complete model and by the simple
one are compared with the bed topography measured along a straight flume
during an experiment carried out at DELFT HYDRAULICS. The experimental
conditions were a well-defined steady flow, while at the upstream boundary a
disturbance was induced by restricting the inflow section with a plate (see
Figure 8). The obstacle yielded a steady deformation of the flow and of the
bed level at the inflow section.
Table 1 gives the relevant conditions of the experiment, while flow and bed
behavior characteristics are summarized in Table 2.
Table 1.

Conditions of the Experiment

Flow
Water Chezy Froude Shields Sediment
Flume Discharge Water
depth velocity surface coeff. number para- transport
width
meter
slope
F
00
S
C
is
h
u
B
Q
0

(m3 /s)

(m)

.1

(m/s) (.I .) (m 2 /s)

(m)

0.60 6.85x10

-3

0.26

0.044

Table 2.

3.00

22.6

(m 3/h)

(-)

(-)

0.39

0.37 2.0x10

-3

Flow and Bed Behavior Characteristics

Secondary Interaction
Exp. b
Exp. b
Flow
Bed
flow
parameter Meyer-Peter Engelund
adaptation adaptation
length
adaptation
length
k Muller &Hansen
,\

(m)

1.14

* {j

= 0.6

,\ s **

,\ r *

,\ s 1,\ w
(-)

(-)

(-)

0.86

0.19

0.75

4.84

5.00

(m)

(m)

**

= 0.5

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

2-D Bed Topography l\10del

170

6.--------,--------r------r--------.--------..:....-.,
4 t - - - - - - - - - + - , - - - vezlocity (5 maasurezmeznts ) - - - - - - + - - - - - - 1

.~

,
t-------t~--""""'\--+___.:_:_"""'i'i"':"'~~~~__=___~__r__""I:--_+-,; ,~J'\,
\

------I

E~

u
.~
tL:

-2t-t---;------t-t------+------+-------I----

l-4~---+------+------+--l-~:------~-----~--....,...._--~-----J

-6L.....-

10

15

20

25

- - + distancez in m
Fig. 8.

Longitudinal time averaged bed level and flow disturbance profile.

Fine and almost uniform sediment was used, having the following
characteristics: D10 = 0.162, Dso = 0.216, Dgo = 0.271 mm.
It can be assumed that for the measurements of bed topography the
duration of the experiment was sufficiently long to ensure the establishment of
equilibrium conditions.

Steady state. For the steady state case the "noise" of bed forms (ripples
and alternate bars), that were strong, especially in the downstream part of the
flume, had to be smoothed by averaging a large number (> 20) of independent
soundings. The time-averaged longitudinal bed level profiles measured at O.lm
from the side walls showed the development of a steady, mildly damped wave
originated near the disturbance, all along the flume. The wavelength of this
steady wave was much larger than the one of the observed alternate bars
which were migrating downstream. The time-averaged velocity profile (only
five independent sets of measurements available) had the same wavy behavior
as the bed, but with a phase lag (see Figure 8).
The observed steady wavelength was about 6.6m, while the dalnping
coefficient 1/LD was approximately 0.09m- 1 .
With the Meyer-Peter and Muller transport formula (b = 4.84), and
without any calibration of coefficient E, for which a value of 0.5 was chosen

(corresponding to f(D o ) = 1.7 ~; see Van Mierlo, [1986]), the simple model
yields a wavelength of 6.5m and a damping coefficient of O.18m- 1.
Most
probably a calibration of E (that weighs the influence of the sloping bed on
sediment transport direction) would give even better results. In Figure 9 the
comparison between the measured and the computed equilibrium bed
topography is given.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

171

Struiksma and Crosato

4,.----------,-------'T--------,..-------,--------,

21----#-~____t-----__+-------+------+-------i

.S

.c

-401:--------~5:----------:-'-10-----~15~-----2-:L-O------..J25

------. distanccz in m

Fig. 9.

Comparison between measured and computed equilibrium bed profile with the simple
model.

In the area close to the upstream boundary the other solutions of the
complete model cannot be neglected (see Figure 5), so the simple model is not
representative in this area.
The results given by the complete model are quite similar to the ones
given by its simplified version. In general the computed wave lengths tend to
be longer than the ones given by the simple model, that is from 6.8 to 7.4m,
The damping
depending on the values of the calibration parameters.
coefficient varies from 0.11 to 0.23, see Table 3 and Table 1. Coefficient E
had the same value, 0.5, as in the computations with the simple model, in
order to have comparable results.
Table 3 and 4. Computed Wave Length and Damping Coefficient with
different Values of the Calibration Parameters f, f3 and
(Meyer-Peter and Miiller Transport Formula).

Table 3

Table 4

0.000

0.025

0.050

7.4
7.3
7.2
6.8

7.3
7.2
7.1
6.8

7.3
7.2
7.1
6.8

0.000

0.025

0.050

0.13
0.12
0.11
0.18

0.15
0.14
0.13
0.20

0.17
0.13
0.15
0.23

e
f

1.3
0.6
0.0 {

05
.
0.0

1.3
0.6
0.0 {

05
.
0.0

The computed results are not strongly influenced by the variations of the
calibration coefficients. This means that the influences of curvature, the spiral
motion inertia and the longitudinal bed slope are not playing the major roles
in the phenomenon.
The complete model gives additionally four exponential solutions which are
induced at the up and downstream boundaries. Their initial amplitudes are
relatively small to negligible. They are strongly damped and are, therefore,
only noticeable close to the boundaries.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

2-D Bed Topography 110del

172

- - - . distanctZ in m

10

15

20

25

0~~"W""""'~--+'I"'-"""""""~~""""""'++......,..-----4

4~-----Ir-t:---:l,.......-.~~
........~~~~'IttPI--_____t

1
.c

71 ~~----Ilo,o-.rr-~:I"'7""'-~~----,,~~ .......~----t'-_____t
76 ............,.....-..----::;~-=--:I~~::lIr-~++Jl~~~....;;..;;....-_____t

+ 2 cm amplitudcz of transvczrscz bczd Iczwl


+ dczformation, longitudinal instantanczous
profilczs

~g~~I

I
altczrnattZ

Fig. 10.

bars propagation

Longitudinal profiles of alternate bars.

Unsteady state. The observed alternate bars and their propagation are
given in Figure 10.
Their average wavelength is 3.9m, and their average
celerity is about 2.9x10- 5 mis, corresponding to 10 cm/h.
The complete model (unsteady state case) gives quite a good prediction of
the alternate bar wavelength. Calibration of coefficients optimizes the results,
as shown in Table 5.
Some parameters which are insignificant in the
prediction of the equilibrium bed topography (steady state), turn out quite
important here: curvature and secondary flow Inertia, weighed by coefficients
land {J respectively (see Table 5).
The computed alternate bar celerity is proportional to the sediment
transport rate, which is strongly dependent on the adopted transport formula.
An additional weighing coefficient for the sediment transport rate is needed, as
the Meyer~Peter and Muller and especially the Engelund-Hansen formula
predicted much higher transport rates than measured.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

173

Struiksma and Crosato


Table 5.

Alternate Bar Wavelength as Function of the Calibration


Coefficients (Meyer-Peter and Muller Transport Formula)

0.000

0.025

0.050

3.1
2.6
2.0

3.5
3.0
2.5

3.7
3.4
3.1

Y:

JJ

1.3
0.6
0.0

}Oo5

(For f = 0 and fJ = 0 only


negative growth rate, no
physical meaning for f = 0
and fJ 1= 0)

Curved Flume Experiment


The redistribution effects are also highly significant for the bed development
in curved channels, and in this case their importance can easily be assessed, as
they are responsible for the point bar-pool configura.tion. .Figure 11 shows the
bed level along the outer bank of a bend of constant curvature downstream of
a straight inflow section, as predicted by the complete model. In this case
mathematical models that disregard the redistribution effects, predict a
constant transverse bed slope all along the bend, and correspondingly a
horizontal rather than a harmonic bed level profile along the banks.
The equilibrium bed topography was measured in an experiment carried out
at DELFT HYDRAULICS in a curved laboratory flume with fixed banks,
Struiksma et al., [1985]. The planform of the bend (see Figure 12), with a
radius of curvature of 12m and a bend length of 29.32m, does not fit into the
characteristics of freely meandering rivers. For example, it is in contrast with
the finding of Leopold et al., [1964] that the meander length is about ten
times the \vidth.
-----+ distanca I width
-4

0.8

12

16

1.0

.r=.

....a.
<::i
"0

1.2

<::i

....0

.~

axp.
growing

axp.

1.4

straight

b<lnd

1.6

Fig. 11.

Bed level longitudinal profile along the outer bank of a channel bend having
constant curvature.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

2-D Bed Topography Model

174

band

Ie

0.50

0.75

.r.

+-'
Q.

1.00

"0
~

>

1.25

+-'

.Q
~

1.50
-4

--....

12

16

20

28

24

distanca I width

computed with complate modal


maasurad
computad by local parametars (zaro order)

Fig. 1.2.

Prediction of longitudinal bed level profiles along the banks.

Table 6 gives the relevant experimental conditions.


The sediment was
characterized by a median grain size D50 = 0.45mm and a gradation parameter
l1g = 1.19.
For measurement of bed topography, duration of this experiment
is assumed, to have been sufficiently long to ensure equilibrium conditions.
The noise of bedforms was smoothed by averaging a large number (> 20) of
independent soundings.
Table 6.

Experimental Conditions

Flow
Flume Discharge Water
Water Chezy Froude Shields Sediment
depth velocity surface coeff. number para- transport
width
slope
meter
q
B
ho
C
F
S
Uo
is
Do
.1
2
2
3
(m)
(m)
(m/s) ( / .. ) (m /s) (-)
(m /s)
(m /s)
(-)

1.5

0.047

0.08

0.39

2.36

28.4

0.44

Copyright American Geophysical Union

0.26

7.2xl0- 6

Water Resources Monograph

River Meandering

Vol. 12

175

Struiksma and Crosato

In Figure 12 the computed longitudinal bed level profile and the measured
one are shown.

Summary of Conclusions
The predicted equilibrium bed topographies in the two experiments agree
well with the measured ones. In the straight flume experiment the complete
and the simple versions of the linear model give rather similar results (the
simple model in this case gives even a better prediction). For this reason we
can conclude that the simple model takes into account the principal
parameters influencing the phenomenon. It will be a useful and simple tool in
predicting the channel's bed topography for large scale problems.
The unsteady linear solution of the complete model gives a rather good
prediction of alternate bar wave lengths. The influence of curvature and of
secondary flow inertia are shown to play an important role in the phenomenon.
In this case the complete model cannot be further simplified.
Discussion
From the steady state analysis it appears that the interaction parameter,

As/ Aw, that is a characteristic of the channel, is important for the equilibrium

bed topography configuration. This is readily explained by the simple model


described in the section on Linear Analysis, Steady State Solution: Simplified
Model. In this section two independent equations, one for the flow and one
for the bed deformation, are given. They are governed, in their homogeneous
parts, by the adaptation lengths Aw and As, respectively. This is enough to
deduce that the ratio of the two adaptation lengths is a significant factor in
the interaction between flow and sediment motion.
Also in the complete
model, (39), it is dominantly present. That the difference between the results
of the simple and complete model is relatively small is demonstrated in Figure
6 for conditions prevailing in meandering rivers.
According to (38), the adaptation length of the bed deformation, and
consequently the interaction parameter, is mainly governed by the width over
depth ratio. This has important consequences in river engineering. Reducing
the width of a river, for example, will decrease the interaction parameter and
will lead to a larger mean water depth.
Consequently, the river bed will
become more stable, for the damping will increase (see Figure 6). If alternate
bars are present they will also be less pronounced. This can be explained
with (37) in which the first term of the right hand member will become
larger, so that the damping will increase.
The influence of the sloping bed on the direction of the sediment transport,
(13), is represented in the interaction parameter by the weighing function, f( (J).
From calibrations of the complete and of the simple model on laboratory
[Struiksma et al., 1985] and on prototype data [Struiksma, 1985] a striking
discrepancy between the two cases was found. For a good reproduction of the
laboratory conditions it was necessary to give the calibration coefficient E in
(15) the value 0.5, whereas for the prototype the value of E giving the best
results was 1.0. This discrepancy is not yet fully understood. Most probably
it is due to the bed forms, which are usually exaggerated in laboratory
experiments. According to Van Rijn [1984] this is caused by the influence of
the ratio depth over grainsize.
The influence of the spiral motion on the direction of the sediment
transport, weighted by the coefficient A in (6), shows a similar discrepancy.
The relative calibration coefficient, f, requires the value 0.5 for laboratory
conditions and the value 1.0 for prototype conditions. Also in this case this

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

2-D Bed Topography

176

~10del

difference may be explained from the exaggerated bed forms obtained under
laboratory conditions.
In the simple model, described in the section on Linear Analysis, Steady
State Solution: Strongly Simplified Model, the convection of momentum by
the secondary flow is represented roughly by a coefficient multiplying the
transverse friction term. The secondary flow momentum convection effect is a
shift of the maximum velocity position toward the outer bank.
From
calibrations of the model its influence could be assessed, and it did not appear
to be negligible.
In the river bed deformation models, the applied sediment transport
formula is also important. In the non-steady state, (37), this is expressed by
the rate of transport, So, and the degree of non-linearity with flow velocity,
the exponent b.
The rate of transport only determines the time scale of
propagation of bed perturbations (alternate bars) and is therefore not present
in the steady state model, ~39). An accurate estimate of this time scale is
difficult, due to the "classic' imperfection of the available transport formulas.
For example, in the straight flume experiment described in the section on
Comparison With Measurements, Straight Flume Experiment, a sediment
transport of 2 l/hour was measured, whereas the Engelund and Hansen and
Meyer-Peter and Muller formula estimated a rate of 10 and 6 l/hour,
respectively. This difference was also observed from the measured propagation
velocity of the alternate bars which was much more in accordance with the
measured transport than with the computed transport.
The degree of non-linearity of the sediment transport as a function of the
flow velocity, expressed by the exponent b in (28), is of the same importance
for the bed deformation as the width over depth ratio in the interaction
parameter, As/Aw (see Figure 7). While the interaction parameter cannot be
estimated accurately because of the uncertain magnitude of the weighing
function, f{ (J), also the degree of non-linearity is difficult to be estimated.
Only some general remarks can be made.
The value of b can be
approximated by comparing the Shields parameter with its critical value for
incipient motion. Consequently, use has to be made of sediment transport
formulas which contain the critical Shields parameters, e.g., Meyer-Peter and
Muller [1948], Ackers and U'hite [1973], etc. When the Shields parameter is
close to its critical value, the exponent b is relatively large and highly
variable. A large value of b will, according to Figure 7, result into a less
stable bed deformation. The damping coefficient can become negative. Such a
situation can easily occur in gravel bed rivers.
The presented steady state analysis can also be used for planform
classification of rivers. If the combination of interaction parameter, As/ Aw, and
the degree of non-linearity of sediment transport, b, leads to negative
damping, more channels will be formed, complying with m > 1 in (38). A
major disadvantage of the proposed classification method is that it requires a
previous knowledge of the channel characteristics. To cope with this problem
in a rough way it is possible to combine the method with regime equations for
the river width and depth. This will lead to a discrimination criterion similar
to that of Leopold and Wolman [1957], however, including sediment properties
(density and grainsize) as an extra factor. For a comprehensive description of
this method reference is made to Struiksma and Klaassen [1988].

Acknowledgement
The content of this chapter is based on the results of a joint effort of
HYDRAULICS and the Delft University of Technology, within the

DELFT

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

177

Struiksma and Crosato

framework of the Applied Research Program (TOW), sponsored by


Rijkswaterstaat (Netherlands Governmental Water Control and Public Works
Department). Part of the work was undertaken while the second author was
in receipt of grants of Fondazione Ing. A. Gini (Padua, Italy) and Istituto
Veneto di Scienze Lettere ed Arti (Venice, Italy).
The authors gratefully
acknowledge Dr. H. J. De Vriend for his comments and support.
Notation
A

weighing coefficient of spiral flow intensity

B
b

width
exponent:

sediment transport rate


Chezy coefficient

D
E
F
G

expressing the degree of nonlinearity of the

grain size
calibration coefficient for the influence of the transverse
bed slope on the sediment transport direction
Froude number
weighing function of the influence of the transverse bed
slope on the sediment transport direction

acceleration due to gravity

h
I

water depth
spiral flow intensity

is
K
k
k

water surface gradient


ratio between longitudinal and transverse wave numbers
longitudinal wave number
transverse wave number
wave length

damping length

Lp

wave length

mode of number of channels

discharge

R
~

radius of curvature of streamline


effective radius of curvature of streamline
sediment transport including pores per unit length

Se

effective sediment
length

St

sediment traansport computed with a transport formula


including pores per unit length

transport

including

Copyright American Geophysical Union

pores

per

unit

Water Resources Monograph

River Meandering

Vol. 12

2-D Bed Topography Model

178
S

streamline coordinate

Sp

phase lag

Zb

bed level

Zw

water level

fJ

angle of direction of sediment transport


coefficient in expression for the relaxation length of the
spiral motion

submerged relative density of sediment

angle of direction of bed shear stress


calibration coefficient for spiral flow intensity

e
()

coefficient for influence of bed slope on sediment transport


rate

Ar

Shields parameter
Von Karman constant
relaxation length of spiral motion

As

relaxation length of bed deformation

K,

Aw

p
U

Ug

relaxation length of main flow


density of fluid
coefficient weighing the secondary flow convection
geometrical

standard deviation of the

sieve curve of

sediment
Tb

bed shear stress


complex celerity of bed perturbations
superscript for perturbation
superscript for amplitude of perturbation

subscript indicating zero-order

subscript indicating channel's centerline.

t
u
v

time
flow velocity in x-direction
flow velocity in y-direction

spatial coordinate

Xp

phase lag
spatial coordinate
Referenres

Ackers, P. and R. White, Sediment transport: New approach and analysis, J.


Hydraul. Div., Am. Soc. Giv. Eng., (HY11), 1973.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

179

Struiksma and Crosato

Blondeaux, P. and G. Seminara, A unified bar-bend theory of river meanders, J.


Fluid Mech., 157, p. 449, 1985.
Callander, R. A., Instability and river channels, J. Fluid Mech., 36, part 3,
p. 465, 1969.
Crosato, A., Simulation model of meandering processes of rivers, extended abstracts,
Euromech 215 Conference, Sept. 15-19, Genova, Italy, 158-161, 1987.
De Vriend, H. J., Steady flow in shallow channel bends, Communication on
Hydraul., Dept. Civil Eng., Delft University of Technology, The Netherlands,
1981.
De Vriend, H. J. and N. Struiksma, Flow and bed deformation in river bends,
Proc. of the Conf. River '83, p. 810, New Orleans, Louisiana, 1983.
Dietrich, W. E. and J. D. Smith, Processes controlling the equilibrium bed
morphology in river meanders, Proc. of the Con! Rivers '83, p. 759, New
Orleans, Louisiana, 1983.
Engelund, F., Flow and bed topography in channel bends, J. Hydr. Div., Am. Soc.
Civ. Eng., 100,(HYll), p. 1631, 1974.
Engelund, F. and E. Hansen, A monograph on sediment transport in alluvial
streams, Copenhagen, Danish Technical Press, 1967.
Engelund, F. and O. Skovgaard, On the origin of meandering and braiding in
alluvial streams, J. Fluid Mech., 57, part 2, p. 289, 1973.
Freds0e, J., Meandering and braiding of rivers, J. Fluid Mech., 84, part 4, p. 609,
1978.
Hansen, E., On the formation of meanders as a stability problem, Coast. Engn.
Lab., Techn. Univ. Denmark, Basis Res., Progress Report 13, p. 9,1967.
Hooke, R. L., Distribution of sediment transport and shear stress in a meander
bend, J. of Geol., 83(5), p. 543, 1975.
Ikeda, S., G. Parker and K. Sawai, Bend theory of river meanders: Part 1, Linear
development, J. Fluid Mech., 112, p. 363,1981.
Jansen, P. Ph (00), Principles of River Engineering, Pitman Publishing Ltd.,
London, 1979.
Kalkwijk, J. P. Th. and R. Booy, Adaptation of secondary flow in
nearly-horizontal flow, J. Hyd. Res., 24(1), p. 19, 1986.
Koch, F. G. and C. Flokstra, Bed level computations for curved alluvial channels,
Proc. of the XIX Congress of the Int. Ass. for Hydr. Res., 2, p. 357, New Delhi,
India, 1980.
Leopold, L. B. and M. G. Wolman, River channel pattern: Braided, meandering
and straight, U.S. Geol. Surv. Prof. Pap. 282-B, 1957.
Leopold, L. B., M. G. Wolman, and J. P. Miller, Fluvial Processes in
Geomorphology, Freeman and Co., San Francisco, 1964.
Meyer-Peter, E. and R. Miiller, Formulas for bed load transport, Proc. 2nd
Congr. Int. Assoc. Hydraul. Res., 2, pap. 2, p. 39, Stockholm, 1948.
Odgaard, A. J., Transverse bed slope in alluvial channels bends, J. Hydr.
Div., Am. Soc. Civ. Eng., 107(12) p. 1677, 1981.
Olesen, K. W., Alternate bars in and meandering of alluvial rivers, Proc. of the
Conf. Rivers '83, p. 873, New Orleans, Louisiana, 1983.
Olesen, K. W., Bed topography in shallow river bends, Ph.D. Thesis,
Communications on Hydraulic and Geotechnical Engineering no. 87-1, Delft
University of Technology, The Netherlands, 1987.
Parker, G., On the cause and characteristic scales of meandering and braiding in
rivers, J. Fluid Mech., 76(3) p. 457, 1976.
Rozovskii, I. L., Flow of water in bends in open channels, Israel Program for
Scientific Translation, Jerusalem, 1961.
Struiksma, N., Prediction of 2-D bed topography in rivers, J. Hydr. Eng., Am.
Soc. Civ. Eng., vol. III, 1985.
Struiksma, N. and G. J. Klaassen, On the threshold between meandering and
braiding, Int. Conf. on River Regime, Wallingford, England, May 18-20, 1988.

Copyright American Geophysical Union

Water Resources Monograph

180

River Meandering

Vol. 12

2-D Bed Topography Model

Struiksma, N., K. W. Olesen, C. Flokstra, and H. J. De Vriend, Bed deformation in


curved alluvial channels, J. Hydr. Res., 23(1) p. 57, 1985.
Van Bendegom, L., Some considerations on river morphology and river
improvement, De Ingenieur, 59(4), (in Dutch, English transl.: Nat. Res. Counc.
Canada, Technical Translation, 1054, 1963), 1947.
Van Mierlo, M. C. L. M., Influence of a sloping bed on the sediment transport
direction, Rep. on Experim. Invest., R657-XXIX, Q186, August 1986, Delft
Hydraulics, The Netherlands, 1986.
Van Rijn, L. C., Sediment transport Part III: Bed forms and alluvial roughness,
J. Hydr. Eng., Am. Soc. Giv. Eng., 110{12), 1984.
Zimmerman, C. and J. F. Kennedy, Transverse bed slopes in curved alluvial
streams, J. Hydr. Div., Am. Soc. Giv. Eng., 104(NY1), p. 33, 1978.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Copyright 1989 by the American Geophysical Union.

Linear Theory of River Meanders


Helgi Johannesson and Gary Parker

St. Anthony Falls Hydraulic Laboratory


Department of Civil and Mineral Engineering
University of Minnesota, Minneapolis, Minnesota 55414 USA
A mathematical model is developed for the calculation of flow field and
bed topography in curved channels with an erodible bed. A small perturbation
approach is used to linearize the governing equations, which retain the full
coupling between the flow field, the bedload transport, and the bed
topography. This coupling is shown to increase significantly the lateral bed
slope in the upstream part of a channel bend, even beyond the value for fully
developed bend flow which is approached in the downstream part of a channel
bend.
This coupling is also shown to give rise to resonant behavior for
certain combinations of input variables; the common origin of the two
phenomena is explained. The predicted flow field and bed topography compare
very well with laboratory data. Further, assuming the banks to be erodible,
The results
the model is used to predict wavelengths of river meanders.
compare favorably with both laboratory and field data.
Introduction
The fluvial processes encountered in natural meandering rivers have long
intrigued engineers. Not only does the geometric shape of river bends give
rise to a complicated three-dimensional flow field, but equally importantly, the
flow field reshapes the bends through the process of bank erosion.
Ikeda, Parker and Sawai [1981] proposed a model for simulation of lateral
migration of rivers in which the bank erosion rates are related to near-bank
values of the depth-averaged primary flow velocity. An essential part of their
analysis is the calculation of the lateral distribution of the depth-averaged
primary flow velocity.
Since the original work of Ikeda et al. [1981], much progress has been made
in furthering the understanding of the complicated flow field and the resulting
bed topography that arise in curved channels. Several researchers [Leschziner
and Rodi, 1979; Kalkwijk and De Vriend, 1980; De Vriend, 1981; and De
Vriend and GeldoJ, 1983J have emphasized that an important cause of primary
flow velocity redistribution in meandering rivers is the convective transport of
primary flow momentum b~ the secondary flow. This mechanism, however,
was neglected by En!1elund l1974] and Ikeda et al. [1981] and more recently by
Smith and McLean l19841, Blondeaux and Seminara [1985], Struiksma et at.
[1985], and Odgaard 119861. The flow field model of Ikeda et al. [1981] was
rederived by Johannesson and Parker [1988a,b], taking into account the
above-mentioned convective transport of primary flow momentum by the
secondary flow.
This was seen to be precisely the mechanism needed to
resolve the contradiction between the theoretical results of Ikeda et al. [1981],
181
Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of River Meanders

182

Tamai and Ikeuchi [1984], Johannesson rI985], and Blondeaux and Seminara
[1985], and the experimental results of [(ikkawa et al. [1976]. These theoretical
analyses predicted that for the classical case of subcritical developed bend flow
in a channel of constant curvature, flow over a flat non-erodible bed always
realizes a higher downstream velocity near the inside bank than near the
outside bank, whereas the experimental results clearly indicate that in this case
the highest velocity is often near the outside bank. The resulting flow model
of Johannesson and Parker [1988a,b] further accounts for the phase lag
between the secondarx flow strength and the channel curvature. The flow
models of Gottlieb l1976], Kitanidis and Kennedy [1984], and Ikeda and
Nishimura r1986] also include this effect, which was neglected by Engelund
[1974) and Smith and McLean [1984).
Recently, Blondeaux and Seminara [1985] and Struiksma et al~ r1985] have
emphasized the importance of the coupling between the flow fiefd and the
Struiksma et al.
sediment transport when calculating the bed topography.
[1985) observed that a significant part of the lateral bed slope in bends is due
to an "overshoot" effect induced by the redistribution of water and sediment
in the first part of the bend. Both their experimental and theoretical findings
indicate that the lateral bed slope at or just downstream of the entrance to a
bend can be substantially higher than the slope obtained if the flow is
assumed to be fully adapted to the bend curvature, as may be the case farther
called
"overdeepening"
herein.
downstream.
This
phenomenon is
Overdeepening cannot be simulated using the model of Ikeda et al. [1981],
since therein the transverse bed slope is assumed to be a function of only local
channel curvature, rather than being calculated under the restriction imposed
by the continuity equation of sediment transport.
Herein, the model of Johannesson and Parker [1988a,b) is generalized to
include an erodible bed, which is the last step in the recent rederivation of the
model of Ikeda et al. [1981]. The bed topography is calculated through a
transverse force balance relation on a sediment particle moving along an
inclined bed. The continuity equation of bedload transport is also satisfied, so
that the coupling between the flow field, sediment transport and bed
topography is retained. It is shown that the overdeepening of Struiksma et al.
[1985J and the resonance detected by Blondeaux and Seminara [1985] are
closely related phenomena, both arising from the same above-mentioned
coupling.
The predicted flow field and bed topography is compared with
laboratory data. The agreement is very good. A more detailed comparison of
the several theories can be found in the companion paper, Parker and
Johannesson [1989].

Finally, using the bank erosion model of Ikeda et al. [1981], the new model
is used to calculate the most unstable wavelength of river meanders. The
results compare favorably with both laboratory and field data, and are shown
to be significantly better than those of Ikeda et al. [1981].
Governing Equations
Let ii and v denote fluid velocity in the sand ii directions, respectively
(Figure 1). The following velocity structure is assumed in the central region
of a wide channel;
ii

u(s,ii)T(()

v=

v(s,ii)T( ()

(la,b)

v(s,ii,()

where u and v denote the vertically averaged values of ii and

Copyright American Geophysical Union

v,

T( () is a

Water Resources Monograph

River Meandering

Vol. 12

183

Johannesson and Parker


(0)

y
CENTERLINE

A'

(b )

Fig. 1.

___....._"'"'--_

DATUM

Definition of variables and coordinate system.

dimensionless velocity shape function averaging to unity, ( = z/h where z is


distance upward normal from the bed, h is upward normal depth, and v is the
transverse secondary flow [e.g. Kalkwijk and De Vriend, 1980; Smith and
McLean, 1984]. Since v must average in the vertical to v, it follows that
1

fT(()d(

foo( = 0

1 ;

(2a,b)

Engelund [1974] evaluated the zeroth order dimensionless velocity shape


function in the central portion of the channel, T( (), as

T(()

(3)

where

x1{rri
= ~

X = Xl

-1

0:

= 0.077

(4a.,b)

and Cf is a bed friction factor given by


Cf

= Ts

pii 2

Copyright American Geophysical Union

(5)

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of River Meanders

184

for straight channels.


Herein, 1-8 denotes the bed shear stress in the
downstream direction. Equation (3) with the coefficients X and X determined
1
from (4a,b) is very similar to the logarithmic law.
The multiple length scales that characterize the flow in a meander bend
are:
the mean flow depth H, the channel half-width b, the meander
wavelength A, measured along the channel centerline, and the minimum radius
of curvature rm. It is assumed in the following analysis that since b/A 1,
the downstream momentum balance can be stated in depth-averaged form,
although the same approximation cannot be made when stating the transverse
momentum balance.
The governing equations for the flow field and the bed topography in a
sinuous channel, with slowly varying erodible bed, can then be written in
dimensionless form as:
Streamwise momentum balance (depth-averaged)

au + vA on
au +
os

1
T 2 [ 1+nC
u

-k [~ (uh of

- - 1+nC

F- 2

ae - 'Y ns
os
T

+ 1~~C

uh

(vT + v) + (vT + v)

Tvd()

A] _

1+nC uv

(6a)

Tvd(]

Transverse momentum balance

l~nc

uT

= - F-

(vT + v) -

l~nc

~ + l\ ji(vT + ii)

u2 T

(6b)

Fluid mass conservation (depth-averaged)

~+~

[{1

nC)vh] = 0

(6c)

Sediment mass conservation

1ft + 1~~C {~ + ~ [(1

+ nc)qnl} = 0

(6d)

Streamwise sediment transport relationship

(6e)
Transverse sediment transport relationship

(6f)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

185

Johannesson and Parker


~

where T is a Coriolis coefficient whose variation from unity is small and will
be neglected [e.g. Johannesson and Parker, 1988a,bl, C is centerline channel
curvature, ~ is the water-surface elevation, 1/ is the oed-surface elevation, Ts is
the bed shear stress component in the
s direction, (qs,qo) denotes the
volumetric sediment transport per unit width in the (s,n) direction, and M and
(J are 0(1) coefficients. The variables have been made dimensionless as follows:
(s,n) =

= (u, v,v)

(u,v,v)

(~,1J,h)

(qs,qo)

(81'/)

b
1=iI

(7a,b)

Ts
pU2

(7c,d)

- U
to

(7e,f)

_ y'RgDS Ds
o - (l-p)UH

(7g,h)

Ts

(g s , go )
,.JRgD s Ds

F=~
@

be

{'Cr

(7i,j,k)

where U, H, and qso denote reach averaged values of u, ii, and qs, p is the
density of water, g is acceleration due to gravity, D s is the mean particle
grain size of the bed material, R is the. submerged specific gravity of the
sediment, and p denotes sediment porosity.
It is important to realize that (6a) through (6f) are valid only in the
central portion of the channel at a distance greater than about one channel
depth from each bank. Close to the banks, the assumption that the vertical
velocity component is negligible, and the corresponding reduction of the
equation of vertical momentum to the hydrostatic condition is no longer valid.
This does not seriously diminish the range of problems to which this model
can be applied, since for natural rivers the ratio of depth to half-width is
Rib NO.1, making the model valid over about 90% of the channel width.
In order to fully. specify the problem, additional relationships are needed for
determining Ts, M, and {J.
The bed stress in the s-direction is evaluated with the use of a friction
factor, Cr;

(8)
If the Engelund-Hansen [1967] sediment transport formula is used, M (see
derivation in Appendix) is given as

2 + 3

1-

=5

(lower regime
dune-eovered bed)

(9a)

(upper regime flat bed)

(9b)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of River 11eanders

186
where

r* -

1"s

(10)

- pRgDs

is the downstream Shields stress, T~ denotes critical Shields stress, and T~


(grain Shields stress) is the part of 1* active in the bedload process. The
appropriate friction law to use in connection with the Engelund-Hansen [1967]
sediment transport formula is

[Cr

rT~]-1/2

a- 1*

T* -

[1

T~ ii ]

2.5 In 2:-5":rr D s

(lower regime
dune-eovered bed)

(upper regime
flat bed)

(lla)
(lIb)

In order to determine the flow regime, the approximate procedure of Brownlie


[1983} may be used. He analyzed a large number of both experimental and
field data and found that upper regime flow exists if one or both of the
following conditions are satisfied:

_u_ >
~
Finally, the expression for
1984]

1.74(1)-1/3

I > 0.006

(12a,b)

(J takes the form [Kikkawa et al., 1976; Parker,

(13)

where G'* is the ratio of lift coefficient to drag coefficient for a spherical sand
particle placed on a rough bed, J.t is the dynamic angle of Coulomb friction,
and f* is an order-one coefficient to be determined from data.
The only
difference between (13) and the corresponding equation derived by [(ikkawa et
ale [1976] and Parker [1984] is that therein it is assumed that all the shear
stress is active in the bed load process rather than just TO. It will be shown
by comparison with data that the use of T~ gives somewhat better results.
The values recommended in the literature to be used for a* and Jt vary
considerably (e.g. [(ikkawa et ale [19761 use p = 0.43; Wiberg and Smith f1985]
use J.t = 1.73). Rather than engaging in a fruitless discussion over which are
the appropriate values to use, the suggested values of [(ikkawa et al. [1976] are
chosen (lr* = 0.85, Jt = 0.43), and the coefficient f* is added in order to
account for inaccuracies introduced by this choice. The value of f* will be
determined from the data of Zimmennann and [(ennedy [1978].
The boundary conditions on (6a) through (6f) are stated subsequently, after
the equations have been linearized.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

187

Johannesson and Parker


Analysis for a Sinuous Channel
The channel centerline is assumed to follow a sinuous shape
- = -1 = - -dO = -1 O'(ks)
-C
f
fm

(14)

ds

is the centerline radius of curvature, with a minimum magnitude


f m, 0 is the angle between the down--ehannel direction and the x-axis (Figure
1), 0' denotes an order-one dimensionless curvature, k = 21r/ X and X are
characteristic meander wavenumber and wavelength, respectively, both based on
centerline arc length.
A dimensionless measure of the maximum centerline curvature is defined as
where

b
\110 = -

(15)

fm

Using the above relation, (14) can be rewritten, in dimensionless form, as


C = \I1oO'(

(16)

(17)

where
>
denotes phase, and k

= hi<.

ks

For example, for a sine-generated curve

o=

(18)

Oocos(ks)

where 00 is the angle amplitude; it follows from (14), (15), and (16) that
\110

= kOo ;

0'

sin>

(19a,b)

From (17), it follows that

{)

7J8 = k (j(fJ

(20)

Parker and Johannesson [1989] (see Figure 2 therein) show that for typical

meander bends, k and


reduced wavenumber

are of the same order of magnitude, so that the


r

= -kl

(21)

is order-one.

Parker and Johannesson [1989] also show from (6c) that if

such that

lV

it follows that v is order-one.

Copyright American Geophysical Union

is rescaled

(22)

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of River Meanders

188

Equations (6a) through (6f) reduce with the aid of (3), (8), (20), (21), and
(22) to

(23a)

1~nc uT cr( lvT


C

- - u2 T
l+nC

+ II)' +

=-

(lvT

-2 ~

+ II) ~

(iVT

+ II)
.

(23b)

Ii (-ev + X1v)

r(uh) , + ~ [(1+nC)vh] = 0

~
[lV +
qso u

er (u M), + 0 [(I+nC)
7Jii

x1iJ(O) _
u

(23c)

!J. !l!J..]] = 0
'Y on

(23d)

where ' = oj 8J.


At this point, an expansion for small curvature, C, is introduced.
Recalling that C = woO', where
0'
is an order-one function of J, and
expanding in wo, it is found that

(u,v,v)

(h,e,TJ)

(1,0,0)

+ wo(Ul,Vl,V1) + ...

(l,er - I*s, TJr - I*s)


1*

(qs,qn) = qso[(I,O)

(24a)

+ wo(ht,el,TJ1) + ...

if 1

(24b)
(24c)

+ wO(qSl,qn1) + ...]

(24d)

where {r and TJr are reference elevations for which H = er - ijr. Substituting
into (23a) through (23d), the following result is obtained at zeroth order
e

-2

1*

rei - nO'

; u

1 ; qs

(25a,b,c)

qso

At O(wo), (23a) through (23d) yield


rui

+ 2Ul = -

-2

e1 -

'111 -

on

fo Tv d (
1

(26a)
(26b)
(26c)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

189

Johanllesson and Parkef

Table 1. Experimental Results


Run
Number

H
(m)

(l/s)

(mm)

Ds

F
(7i)

16.fCi
25a)

(44)

As

meas. (34c) (9a)

Kikkawa et ale [1976]


0.055

2.00.10- 3

0.90

0.63

13.84

3.54

2.98

0.063
30
M2
Struiksma et ale [1985]

2.00.10- 3

n on
v.t7V

nv.u~
~1

11)

.4

0.080

2.36.10-3

0.45

0.44

0.45
0.45

Ml

Tl

25

47

T2

61

0.100

2.03.10- 3

T3

74

0.091

4.19.10- 3

(f

U~ + ~

3.93

4.81

1::'.4

~':>.\J"t

I!-f
"t.~l

If.~~

3.90 4.85

9.10

1.85

3.54

4.61

0.41

9.11

2.90

3.85

4.58

4.22
4.23

0.57

8.86

2.30

5.14

6.21

3.82

[(Vi + X/t(O) -

~~] =

no

(26d)

In the above relations, only the lowest-order terms in (; are considered (Le.
the underlined terms are neglected) since (; is typically fairly small. This is
illustrated in Table 1 of Parker and Johannesson [1989]; e.g. for 'Y = 10 and
Cr = 0.005, (; = 0.05.
The appropriate boundary conditions for (26b) have been derived by
Johannesson and Parker [1988b] and will not be repeated herein, since their
solution of (26b) will be adopted directly. Further conditions needed to fully
specify the problem are the requirements of channel walls that are
impermeable to both water and sediment;
Vt

qnt

at

:l:

(27a,b)

and the requirements that the total discharge of water and sediment and the
average river slope are unaffected by the perturbed quantities, to wit
1

I(Ut + h.)dn = 0

21t t

IUidn = 0

i(e~dnd<P = 0

(28a,b,c)

It can be deduced from the governing


solution for et, TJt, and Ut is a function of J
n
is odd, plus a function of J which
cross~ectionally averaged values of
TJt,

e,
t

equations that the form of the


and n, for which the variation in
indicates the variation in the
and Ut from zero.
This fact,

together with (26a) and (28a,b,c) allows the integral conditions given by
(28a,b,c) to be simplified to
Ut

et

TIt

at

(29a,b,c)

The above linearized version of the governing equations (26a,b,c,d) is very


The
similar to the model of Blondeaux and Seminara lI985], (32) therein.
most important difference is that Blondeaux and Seminara [1985] did not
neglect the underlined terms in (26b).
These terms a.re of importance for

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of ill ver Meanders

190

alternate bars, which, however, is not the subject of this analysis. Secondly,
Blondeaux and Seminara [1985] used the depth-averaged form of the transverse
momentum equation, which does not allow for the calculation of the secondary
flow. They then determined the secondary flow using an expression derived
for fully developed flow in circular channels. Equation (26b), on the other
hand, can allow for a more accurate calculation of the secondary flow. Most
importantly, it allows for calculation of the phase lag between the channel
curvature and the secondary flow strength [Johannesson and Parker, 1988b].
Thirdly,
This phase lag tends to be rather small for natural channels.
Blondeaux and Seminara [1985] did not account for the convective transport of
primary flow momentum by the secondary flow (the last term on the right
hand side of (26a)). This term can be of considerable importance, as has been
shown by Johannesson and Parker [1988a). Fourthly, Blondeaux and Seminara
[19851 expanded the friction factor, whIch is taken to be constant herein.
Final y, Blondeaux and Seminara [19851 included the dependence of the
streamwise bedload function on .local depth, which influence is neglected
inherently herein by the power law assumption given by (6e). As regards the
final two points, the assumptions of the present analysis correspond to (45) of
the companion paper, Parker and Johannesson [1989].
Reduction
Further progress requires solution of (26b) for the water-surface elevation
and the secondary flow, together with the boundary condition given by (29b).
Johannesson and Parker [1988b] obtained the following results:
~1 =

(30a)

F X nO'
20

where
(30b)
is a coefficient very close to unity (Le. between 1.01 and 1.11 for 10 < 1/{rSi
< 30), and
(31a)
where

GoW =

h [[X

+ ~o

(5 -

i X + ~5](X + () - ~ x (2 -! X(3 - h (1-X)(4


2

riu (6] - X [X + ( - ~ (2]


20

(31b)

denotes the vertical structure of the secondary flow in the case of developed
bend flow. The function O's, which satisfies
r

W+

O(1s

0(1

Copyright American Geophysical Union

(31c)

Water Resources Monograph

River Meandering

Vol. 12

191

Johannesson and Parker

quantifies the strength of, and the phase shift in, the secondary flow due to
changing curvature in a sinuous channel. In the above relation

(32)

For natural channels, the value of r/o, and thus the phase shift between
the secondary current strength and the channel curvature predicted by (31c)
and (32) is typically rather modest, Le. on the order of 10 degrees, so that (Js
does not deviate substantially from the local normalized dimensionless
curvature (J.
The solution for the bed profile TIt is now decomposed as
(33)
where Tl
satisfies the portion of (26d) that represents the direct effect of the
IC
curvature-induced secondary flow on the bed, and the boundary condition
given by (29c). This results in
Tl IC

=-

A n

(34a)

Us

where

A = _ 1 Xl VI ( 0)
f:J

(34b)

Us

is the transverse bed slope parameter introduced by Ikeda et ale [1981].


Reducing with (31a),(31b),(7k), and (4a), it is found that

.1.-'
A = - /lc Go(O) =

12

7iT13 iI5

X
X

+7

+~

For "typical" values X = 1 and (3 = 1.5, for example, a value of A


realized.
We likewise decompose as follows
Ut=uIC+u IF

; ht=hIC+hIF;

Vt=vIC+vIF

(34c)
4.82 is

(35)

where u lC , hlC' and vIC are terms generated directly by the


curvature-induced bed topography, TJ IC ' and u 1F ' h 1F , and vIF are extra
terms, needed to satisfy sediment continuity. The subscripts "e u and "F" are
selected so as to indicate the response of the system that is .characterized by
local curvature forcing (e.g. point bars), and the component characterized by
the response of the free system (e.g. alternate bars) to this forcing.
Substituting into (26a,c,d) and making use of (30a) and (31a), the problem is
decomposed to:

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of River Meanders

192
C-Problem

(36a)

(36c)
Here the overbar denotes integration for 0 to 1 in (, Le. vertical averaging.
The boundary conditions on vIC (27a) and u lC (29a) are
vlci

0=:1:1

= 0

Ulci

0=0

(37a,b)

= 0

F - Problem

which can be reduced to

r(M-l) uiF

rT/iF -

{flT/

r ~=-

r(M-l) uic

r hic

(3gb)

r = f3/(er) = f3/(,2Cf) is an 0(1) constant [see Parker and


Johannesson, 1989]. The boundary conditions on vIF (27a), T/ IF (27b), (26d),

where

and (29c), and u lF (29a) are:


VIFI

0=:1:1

T/ IF I

0=0

(40a,b)

0
0

UIFI

0=0

(40c,d)

Once the solution for the C-Problem is obtained, the F-Problem (39a,b)
together with the boundary conditions given by (40a,b,c,d) becomes fully
defined.
The C-Problem, as stated herein, corresponds to the model of Ikeda et ale
[1981], modified three ways. The most important difference is that Ikeda et
ale [1981] neglected the convective transport of primary flow momentum by the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

193

Johannesson and Parker

secondary flow.
Secondly, Ikeda et ale [1981] assumed the strength of the
secondary flow to be in phase with curvature. This corresponds to setting (1s
= (1 in (31c). Thirdly, Ikeda et ale [1981] did not take into account the
metric coefficient which introduces the free vortex term -(1 in (36b ).
The
F-Problem, on the other hand, was completely neglected by Ikeda et at. [1981].

Solution for Constant Curvature


For the case of fully developed flow in a channel of constant centerline
curvature, C, (1 = 1 and \lIo = C, (15) and (16), and the streamwise
Further, vIC = 0, (36c, 37a), and there is no
derivatives (8/ fJ vanish.
contribution from the F-Problem (39a,b). Equations (30a), (31c), and (34a)
then reduce to
(41a,b,c)

Johannesson and Parker [1988al noticed that, in the absence of the


redistribution term due to the secondary flow in (36b), the general solution for
u
would take the form
lC
(42)

= u lC at n = 1. They then obtained an approximate solution of


ICb
this form to (36b) using a "moment method." Their solution for the case of
constant curvature reduces . to

where u

u ICb

1 2

= 2(F

20

(43)

+ A + As)

where

As

(?\

TGo

181

[~] 2 ~1 [2X 2 + h +

ts]

(44)

Equations (42) and (43) were used by Johannesson and Parker [1988a] to
simulate the measured values of u
taken by Kikkawa et ale [1976], (Exp. Fl,
lC
F2, F3 for which A = 0), and Struiksma et ale [1985], (Exp. Tl and T2, for
which the measured values A = 3.54 for Exp. Tl and A = 3.85 for Exp. T2
were used as an input in (43). The agreement was very good, and indeed
much better than if the corresponding model of Ikeda et ale [1981] is used.
The expression derived for the transverse bed slope parameter A (34c) has
yet to be compared with data. This is done in Figure 2. The -input data for
all the experiments used is summarized in Table 1, except for the data of
Zimmermann and Kennedy [1978], which is summarized in Table 1 therein.
Some calculated results are shown in Table 1. The flow conditions for all the
erodible bed experiments were estimated to be in the lower regime (dunecovered bed), (12a,b).
Note that values for r~, r*, and r~ are needed in
order to compute A and M.

The values of T~ are obtained from a Shields

diagram [Vanoni, 1977, Figure 2.43].


and r~ is estimated from (1Ia,b).

Measured values of

Copyright American Geophysical Union

T*

are used (10),

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of Ri vel' l\1eandcrs

194

o ZIMMERMANN AND
KENNEDY (1978)

KIKKAWA ET At. (1976)


A RUN M1
6 RUN M2

o
Q

w
a:

a:

~
CI)

<

STRUIKSMA ET AL (1985)
RUN T1
() RUN T2
RUN T3

::>

w
~

<

f" ;:: 1.19

COEFFICENT OF DETERMINATION

= 0.76

o IL...-.....a...-----L....-----'_......_---.L----'-_.l.-------'-----'-------'

(a)
Fig. 2.

5
A

10

(CALCULATED)

OIL...-.....a...-----L....-----J'---.....&........----I...---'-_.L.-~_a_----I

(b)

5
A

10

(CALCULATED)

Comparison of transverse bed slope parameter A, as calculated by (34c), and as


measured. a) Calibration of f*. b) Validation.

The dimensionless curvature \110 = be, assumed to be small in this


analysis, varied from 0.182 to 0.333 for the experiments of Zimmermann and
I(ennedy [1978]. It took the respective values of 0.111 (b = 0.5 m and r =
4.5 m) and 0.0625 (b = 0.75 m and r = 12.0 m) for the experiments of
Kikkawa et ale [1976] and Struiksma et ale [1985] (Exp. T1, T2, T3). These
values are perhaps sufficiently small for the analysis to be valid.
The calculated transverse bed slope parameter A (33c) is plotted in Figure
2 versus the measured one. The data of Zimmermann and I(ennedy [1978] was
used to determine the value of f* to be 1.19 (Figure 2a). The scatter is small
and much less than if the total shear stress, r*, is used instead of T~ in the
calculation of fJ (Figure 3a), as recommended by Kikkawa et ale r1976] and
Parker [1984].
The calibrated equation is verified b~ simufating the
measurements of Kikkawa et ale [1976] and Struiksma et ale l1985] (Figure 2b).
The results are very good, and again much better than if the total shear stress
r* is used for the calculation of fJ (Figure 3b).

Solution for a Sine-Generated Curve


For a sine-generated channel, the centerline curvature is specified by
(19a,b). Equations (30a) , (31c), and (34a) , then reduce to
(45a,b)
(45c,d)
The solution of the C-Problem (36a,b,c) can be written as

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

195

Johannesson and Parker

o ZIMMERMANN AND

KIKKAWA ET Al. (1976)


A RUN M1
RUN M2

KENNEDY (1978) 0

o
a:

o&

:::J

o
w
a:

8(()

00

&

:::>

~
w

STRUIKSMA ET Al. (1985)


RUN Tl
() RUN T2
RUN T3

00

f..

= 0.535

COEFFICEf'.ff OF OETEAMlNAllON = 0.42


O~.....L..----L.._I---"""",,-------L..----l""-----....r...----,---.L------'"

(a)
Fig. 3.

O---L....-----L..--l""-----"""--_._----L_"""---'-----L..-...I

10

(b)

(CALCULATED)

10

(CALCULATED)

Comparison of transverse bed slope parameter A, as calculated by (34c), (r* is used


instead of
in the calculation of fJ by (13, and as measured; a) Calibration of

1"0

f*.

b) Validatiop.

(46a)
(46b)
where

ac = -

rrh {r[x

20

(F

+ 2) -

+ (A + As)cos2 0"SL]

+ (A + As )sin20'SL}
bc ~

1
rrn
{2[X20 F

- r[rX20 +
Cc

:::

r(b c

- 1 + (A + As)COS20"SLJ

A + As)sin20"SL]}

(46c)

F X

20

A cos 2 0"SL)

de = r(lic - ~ A sin20"SL)

(46d)

(46e)
(46f)

Having solved the C-Problem, the F-Problem can be stated as follows

(47a)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of River Meanders

196
r(M-l)uiF

r71iF -

(/211

1F
r lJii'l
=

n(D 1cos>

D2sin

(47b)

where
(47c,d)
The boundary conditions are given, as before, by (40a,b,c,d).
Having so successfully used the "moment method" on the C-Problem, and
knowing that the transverse bed profile in an actual river is often well
described by a linear profile, the "moment method" is also used to simplify
the F-Problem.
The first moments of the lateral distribution of ulE and 11 lE are defined as:
1

n ulF

= ~f

ulFndn

-1

nT/IF

=~f

T/IFndn

(48a,b)

= -

1, the

-1

Multiplying (47a,b) by n and integrating from n


following result is obtained if (48a,b) are applied:

1 to n

(49a)

(49b)
The lowest order term in the Fourier series for 11

1F

is
(50a)

where 111Fb

711F at n = 1.

This, together with (48b), gives


nT/IF

= [~] 2 T/IFb

(50b)

Equation (49b), together with (40b) and (50a,b), reduces to

Although the solution to the F-Problem (if Fourier expansion is used)


indicates that the variation in the transverse direction goes as sin( 1r/2 n)
rather than of as a linear function (note that the difference is small), a linear
solution, that preserves the first moments n ulF and nn 1F ' is obtained by
, respectively, by u lFb /3 and 111Fb/3 in (49a) and (51).
replacing n ulF and n
711F
This gives

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Johannesson and Parker

197
ruiFb

r(M-l)uiFb + r71iFb +

2u 1Fb

=-

711Fb

[;fr 7l1Fb = Dl

(52a)
COS cP

+ D2sin cP

(52b)

The solution of (52a,b) can be written as


(53a)
(53b)
where
D l + Jt D 2
b- F -- -rJ2 J2
+ r 2 J2

aF = -

2a F - r

bF

(53c,d)
(53e,f)
(53g,h)

An interesting feature of the solution to (52a,b) deserves some discussion.


For certain combinations of input variables, the solution to the F-Problem
shows resonant behavior.
The solution, of course, fails to be valid in the
neighborhood of this resonance. The model of Blondeaux and Seminara r1985]
also shows the same behavior.
The condition for the occurrence or the
resonance can be determined from (52a,b). Using (52a) to eliminate 1J
from
1Fb

(52b) one gets

ui;b + }

[3 - M + [;]2 r Ju iFb + 2[;f ~ u1Fb


(54)

Resonance occurs if both of the following conditions are satisfied: the damping
coefficient must vanish, Le.
(55a)
and the wavenumber of the undamped homogeneous equation, k res , must equal
the wavenumber of the forcing function, k, to wit

k2res = 2 [1r]
2" 2 r

f2

= k2

(55b)

Equation (55a,b) is satisfied if Jt and J 2, as defined by (53g,h) both approach


zero, which, as seen from (53c,d), will obviously lead to resonance.
In Figure 4, the results for the amplitude of the component of Ut in phase
with curvature, Le. be + b as predicted by the present model, are compared
F

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of River Meanders

198
30

r----------r---------r------_----__.

/Y=
V
f:

20

= 16.95
= 0.77
/3 = 1.08
M = 3.69

1/fC';
40

I!

:In/

,/411,.,/

10
u..

~.o

'a

Y =30

y= 20

e,:'

()

l.o

-10

- - PRESENT THEORY
(As = 0, crSL = 0)
-20

..... BLONDEAUX AND


SEMINARA (1985)

-30 """----_ _- - 1 .

0.2

----1...

0.4

.....1.-_---..;._ _- - - - '

0.6

0.8

k
Fig. 4.

Comparison between the present theory and the theory of Blondeaux and Seminara
[1985] for the amplitude of the portion of "1 in phase with curvature.

with the results of Blondeaux and Seminara [1985], as reported in Figure 2


therein. The input variables have been selected so as to agree with those
used by Blondeaux and Seminara [1985]. That is, r* = 0.25, Ds/H = 0.005,
and F = 0.77.
Assuming an upper regime flat bed, (lla) gives l/JCi =
16.95. Using the sediment transport formula of Meyer-Peter-Miiller ((80) in
Appendix) gives M = 3.69 if r~ = 0.047. Also note that (3 is taken to be
1.08 as predicted by Blondeaux and Seminara [198. 51, whereas if (13) is used, (3
= 1.16. Further, As and Us have been taken to be zero and (1, respectively,
to make the comparison as precise as possible. The agreement is seen to be
excellent. The present theory predicts resonant conditions to be at , = 33.3
and k res = 0.136, from (55a,b), whereas the theory of Blondeaux and Seminara
rt985] predicts the conditions to be approximately , = 40 and k res = 0.15.
The overall structure of the two solutions is also seen to be very similar.
Resonance and Alternate Bars

Blondeaux and Seminara [1985] explained the origin of the resonance


phenomenon. Their conclusion was that, for a certain combination of input
variables, curvature forces a perturbation of alternate bar type to the flow. It
may appear contradictory that the present model, which does not retain all of
the terms of a full model of alternate bars, shows the same resonant behavior.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

199

Johannesson and Parker

In order to clarify this, a simple alternate bar stability analysis for a straight
channel is performed using the present model in the companion paper [Parker

and Johannesson, 1989].


Parker and Johannesson [1989] consider only alternate bars, Le. first-mode

braiding. Higher modes of braiding may be studied by replacing sin (rn/2) in


(77a) therein with sin(m rn/2) for m = 1,3,... and cos (m ; n) for m
2, 4, ... , where
m
denotes the number of braids.
This results in
perturbations of the form
sin (m
Ut

ib

~ n)

(m odd)

[ cos (m ~ n) (m even)

j U

tb

= ulOelrtcos(kt)

(56a,b)

In the above relations, a is the exponential growth rate, c is the downstream


wavespeed, and UtO is an 0(1) coefficient. Note that instability requires that a
> o. If the analysis of Parker and Johannesson [1989] is repeated for this
case (but with P = P t = 1, Mt = 0, therein), it is found that

= ( Qo

qso

rn [M - 1- [; mrr[l + ~]]
rn +

c = ( Qo qso

[6 - 2M

r2 ]

(57a)

(57b)

The notation differs slightly from Parker and Johannesson [1989].


It is
apparent from (57a) that m = 1 is always the most unstable mode, indicating
that the simplified model can never predict braiding. In the following, m will
then be taken to be one which is the appropriate choice for alternate bars.
It can be easily shown that if the resonance conditions, given by (55a,b),
are satisfied, it follows from (57a,b) that a = c = o.
This confirms the
conclusion of Blondeaux and Seminara r1985] that the resonance disturbances
are steady and non-amplifying alternate bar perturbations forced by curvature.
It is of interest to see how far this simple bar instability analysis can be
stretched.
It is seen from differentiation of (57a) that a/( lQoqso) is a
monotonic function in k for all I. The model therefore cannot be used to
predict the wavelengths of alternate bars. The model can, however, be used
to predict if alternate bars exist or not. A typical neutral curve (a = 0) is
shown in Figure 4 of Parker and Johannesson [1989], together with the results
of the full theory of Colombini et ale [1987]. As seen from case b) of that
figure, the present model gives surprisingly accurate results for low values of
k and predicts accurately the critical value IC below which the channel is
stable for all k.
An expression for IC is obtained from (57a) and the
condition that a < 0 for all r. This yields
(58)
Further, note from Figure 4 of Parker and Johannesson [1989] that in the
neighborhood of resonance (, = 23.7 and k res = 0.108 from (55a,b)) the
~resent theory gives results very similar to the full model of Colombini et ale
lI987].

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of Rivel' l\1eandcrs

200

From the above analysis, it is apparent that the present model has the
essential features of alternate bars needed to produce resonance in a sinuous
channel, although some terms of importance in bar instability analysis have
been neglected.
Resonance and Overdeepening

Struiksma et ale [1985] explained the overdeepening in terms of the


wavelength and dampIng obtained with a linear analysis of the governing
equations for the steady state case, assuming zero curvature. If the present
model is used to carry out this analysis, the governing equations are the same
as for the alternate bar analysis, except that unsteady terms are dropped.
Dropping the unsteady terms in (79) of Parker and Johannesson [1989],
applying (45) therein, and replacing Ub therein with Utb of (56) herein, one
gets
,',
u ib

+r1

[3 - M

r
+ [1r]
2" 2 r ]u,ib + 2 [1r]
2" 2 f'2'"

tb

(59)

Equation (59) is seen to be identical to the homogeneous part of (54), which


was used to explain the resonance phenomenon. The damping coefficient Cd
and the wavenumber k res are, as before, given by (55a,b).
Defining the discriminant d of (59) to be
d

[c

2 -

k 2res /k2 ]

(60)

one gets if d < 0 (negative discriminant) that all solutions of (59) have the
form
Utb

= Ae

- c 4>
d

sin( <p+ d )

(61)

where A and d are arbitrary coefficients of no interest in this analysis, which


can be determined from boundary conditions.
The wavenumber
k
is
obtained from the condition that (1/2)R = 1. This, together with (60) and
(55b) gives
(62a)
and, for the sake of completeness, the expression for the damping coefficient
(55a) is written again as
(62b)
The relations for the wavenumbers and the damping are shown in Figure 5a,b
as a function of r for the case M = 5. These are identical to the result
obtained by Struiksma et ale [1985] for their simplified model (Figure 6
therein).
The similarity between the resonance phenomenon predicted by Blondeaux
and Seminara [1985] and the overdeepening predicted by Struiksma et ale [1985]

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

201

Johannesson and Parker

RESONANCE
CONDITIONS
Ea. 55b

k/2

ck/2

Ea. 62a
-1

Fig. 5.

-1L....-...............L......L.............L..L.L.L._.....L...-..............................

I---L-""""---L..........L..L.I.L._""""'--.............~........

a) Damping coefficient, and b)


straight channel for the case M

10

0.1

10

0.1

wavenumber of a steady perturbation of Ut in a

= 5.

is now apparent.
If the resonance conditions, as given by (55a,b), are
satisfied, the damping coefficients, as given by (62b), will be zero and the
wavenumber, as given by (62a) will be equal to the resonant wavenumber.

Solution for an Arbitrarily Shaped Channel


In order to be able to compare the model to a variety of experimental and
field data, a solution is sought for the case of arbitrary curvature.
A general solution to the C-Problem (36a,b,c), obtained by Johannesson
and Parker [1988a,b], can be summarized as:

Us

= us(O) e-{O/r)f/J +

$e-{o/r)f/J Je(o/r)f/J~(f/J')df/J'

(63)

lllCb(f/J) = lu1Cb(0)

+ !.
r

Ix

2
X u(0)]e- f/J/r - X u( f/J)
20

20

(F + 2) - l]e-2f/J/r J u( f/JI )e2f/J' /rdf/J'


2

20

+A + As)e-2f/J/r J
o

et,

us( f/JI) e2f/J'/ rdJ'

(64)

Note that
VI, 11 ' and h
are given by (30a), (31a), (34a), and (36a),
1C
lC
respectively.
Using again the moment method on the F-Problem, (39a,b) reduce to

ru Fb

2u 1Fb

=-

11lFb

Copyright American Geophysical Union

(65a)

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of River Meanders

202

=Using (65a) to eliminate 71

1Fb

(M - l) ru icb

(65b)

rhicb

from (65b) gives

[3 - M+ [;fr]uiFb + 2[;f ~ u1Fb

ui~b + }

(66)
Defining new variables YI and Y2 to be
YI

u 1Fb

Y2

(67a,b)

uiFb

a set of equations, equivalent to (66), can be written as


Y{

Y2

=-

(68a)

Y2

}[3 - M+ [;]2 r ]Y2 - 2[;f


+ }[(M -

l)uicb - hicb]

Y.
(68b)

Equations (68a,b) are solved with the IMSL subroutine DVERK, which uses a
Runge Kutta-Verner fifth and sixth order method. The boundary conditions
used are
(69a,b)
Note that from (65a) and (67a,b), it is seen that setting YI(O) = Y2(O) = 0
corresponds to selecting u1Fb(O) = 1J1Fb (O) = 0, which is often the appropriate
choice, Le. when the upstream end of the study reach is at the downstream
end of a straight portion of the river channel. Having solved (68a,b) for YI
and Y2, u 1Fb and 711Fb are determined as follows
U

1Fb

YI

711Fb

=-

rY2 - 2YI

(70a,b)

The general model was used to simulate the experiments of Struiksma et ale
[1985] (Exp. Tl, T2, T3).
The results are shown in Figure 6.
What is
compared is the longitudinal variation of depth,

ii,

and the depth-averaged

primary flow velocity, ii, at 0.375 m from left and right bank (Le. n = 0.5).
The agreement is seen to be excellent.
It is particularly notable that the
model predicts quite well the overdeepening at the beginning of the bend for
all three runs. The results from the present linear theory are seen to be as
good as or better than those obtained from the full nonlinear model of
Struiksma et ale [1985].
The discontinuities in the predicted depth and
velocity at the bend entrance and exit are of no concern. They are due to
the discontinuity in the channel curvature and the fact that the transverse
slope of the water-surface is assumed to be a function of local curvature only
(30a).

Copyright American Geophysical Union

Water Resources Monograph

0.05 ,

IO'10~

..

/1

.'

,-,\ ..

l"",

RIGHT

.. "
.~
'l

'

'~I

;
.... \,,"

\..
\

I~

0.15

River Meandering

en

U)

a:

0.4

en
a:
w

U)

ex:

....w
w
!

II

25
1 (METERS)

Fig. 6.

en
en
0
~

0.-

0.10

~
~

::r.-'
~

~I

E~P. !2

j oj

BEND
I(

~I

EXP. r 3

25

50

S (METERS)

0
0

At1\ -

0.375 METERS FROM RIGH A"'V LEFT BANK


AM) AlONG It-E CHAN\EL AXIS

~'~~~~j MEASURED

DEPTH PROFILES

- - COMPUTED DEPTH PROFILES

0.375 M:TERS FROM RIGHT At-V LEFT BAt-.I<

0.4

\:J

lEFT

EXP. T 1
50

0
t:::r"

(t)

RIGHT

I~

50

BEND

'i

....W
!

I:J

'-l

I.e:

25

0.5

w
en
a:
w
Q.

RIGHT

(.)

0.3

en

BEND

I-

Z
0
U

I(

a:
w

/\ ...... '.\ ,/"""\,~

0-

S (METERS)

Q.

0.05

(METERS)

....

O.,J

I
50

'''v''/\'

/','
, I
'/

25

~
~

I.e:

EXP. T 1

a:
w
....
w 0.10

(t)

BEND

0.5

...'

0.05

Vol. 12

0.3

BEND
I(

25

tI
I EXP.

RIGHTI

LEFT

MEASURED VELOCIT IE S

- - COMPUTED VELOCITY PROFILES

T2
50

(METERS)

Longitudinal variation in depth, h, and depth-averaged primary flow velocity, u, as


measured and predicted.

Copyright American Geophysical Union

t."'-'

0
v.J

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of River Meanders

204
Table 2.
Run
Number

Cdk / f
(62b)

Wavelength of Overdeepening

8[~rr

X(m)
meas.

X(m)
(62a)

28.2
44.5
25.6

19.5
19.5
19.5

17.4
20.6
20.1

lres(m)
(55b)

Struiksma et ale [1985]


T1
T2
T3

1.16
2.17
1.19

15.7
15.7
17.7

It is of importance to note that the overdeepening results solely from the


F-Problem; if it is neglected, the transverse bed slope simply increases
monotonically from zero at the bend entrance to its fully developed flow value
at some distance further downstream. The relation between overdeepening and
resonance, as discussed earlier, can thus be easily tested with data.
The
measured wavelength of the overdeepening should approximately correspond to
the wavelength of a steady damped disturbance, as given by (62a). If the
damping coefficient Cd is small, the wavelength so computed should be about
equal to the resonant wavelength, as given by (55b). This is illustrated in
Table 2. The predicted wavelength (62a) is shown to be very close to the
estimated wavelength of the overdeepening (respective difference of 11 %, 6%,
and 3%) for Exp. Tl, T2, T3). Further, since the scaled damping coefficient
squared (2Cdk/ f)2 is substantially smaller than the scaled resonant wavenumber
squared (k res /{)2 = 8(1r/2)2r (e.g. Table 2), the estimated wavelength is also
close to the resonant wavelength, as predicted by (55b) (respective difference of
19%, 19%, and 9% for Exp. Tl, T2, T3), Table 2. It should be emphasized
that only for the case of zero damping does the resonant wavelength become
equal to the wavelength of a steady disturbance (55b), (62a). Although that
condition will rarely be exactly met, the calculations clearly show the close
analogy between resonance and overdeepening.
Bend Instability
Let us consider the stability of an initially straight river to an arbitrary
fluctuation of the channel centerline around the x-axis
y

= yoeaotsin( 4>

- wot)

(71)

where y has been nondimensionalized with the channel half-width b, ao is


the dimensionless exponential bend growth rate, and wo is a dimensionless
circular frequency, such that wo/k is a dimensionless bend migration speed.
The channel curvature is then found to be
(72)
where \110 = k 2yoe aot and (J = sin( 4> - wot).
Following the analysis of Ikeda et ale [1981], the dimensionless rate of bank
erosion is assumed to be given by

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

205

Johannesson and Parker

(73)
where Eo is a positive coefficient of bank erosion, and ac' bc' aF' and
given by (46c,d) and (53c,d), respectively. Equation (71) yields

= [-

cos( tP - Wot)

bF are

+ ~ sin( tP - Wot)] lifo

Direct comparison of (73) and (74) yields the following expressions for
Wo

(74)
0'0

and

(75a,b)
The wavenumber k which maximizes instability (Le. maximizes 0'0) is
found using the IMSL subroutine ZXGSN, which uses the golden section search
method.
The results are compared with data (75 field cases and 89
experiments) in Figure 7. This is the same data as used by Ikeda et ale
[1981] for the generation of their Figure 5, except that experiments in channels
with non--erodible banks are excluded.
In order to make the calculation
procedure uniform A and M are taken to be 2.89 (same as the value used by
Ikeda et ale r1981i) and 5 (typical value for M, e.g., Struiksma et ale [1985]),
respectively, for at! of the data, since the available information usually did not
allow for calculation of these parameters. As seen from Figure 7, the present
theory predicts the right orders of magnitude of meander wavelengths both for
the laboratory and the field data.
The scatter is substantial, but indeed
considerably less than that obtained by Ikeda et ale [1981].
Even more
important than the reduced scatter is the fact that the model can explain the
scatter as being partly due to the sensitivity of the predicted wavelength
values to the value of M. On the other hand, a change in A does not modify
the results much.

Conclusions
A mathematical model for calculating the flow field and bed topography in
curved channels with an erodible bed is presented. The results can be viewed
as a complete rederivation of the flow and bed topography model of Ikeda et
ale [1981]. A small perturbation approach is used to linearize the governing
equations, which retain the full coupling between the flow field, the bedload
transport, and the bed topography. This coupling, missing from the model of
Ikeda et ale [1981] is shown to be responsible for the overdeepening observed
and predicted theoretically by Struiksma et ale [1985] and the resonance
detected by Blondeaux and Seminara [1985].
The overdeepening and the
resonance are further shown to be essentially the same phenomenon.
The present theory is in general agreement with the work carried out at
the Delft Hydraulics Laboratory [Struiksma et al., 19851 and at the University
of Geneva lBlondeaux and Seminara, 1985]. All of the above are, however,
quite different from the theory of Odgaard [1986], from which the continuity
equation of sediment is absent and in which the overdeepening is explained as

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of Ri ver ~1canclcrs

206

10 5 ............---...,.----.,....----~---_r___---_r___--____,.
BORDAS (1963)
QURAISHY (1973)
() WOLMAN & BRUSH (1961)
SCHUMM & KAHN (1972)
A ACKERS & CHARLTON (1970)
o SHARMA (1973)
ANDERSON ET AL. (1975)

00

Q
QQ

.
.- ...

Q
Q

~0

0 0
~

00
6.

o @0
4~0<> 0

"

.
~

TILTING FLUME DATA


CHANNEL DATA
o CHITALE (1970)
SCHUMMM (1969)
Q WOLMAN (1955)
LEOPOLD & WOLMAN (1957)
LANGBEIN & LEOPOLD (1966)
o KINOSHITA (1961)

o TOWING

,~
()

()
()

10-1
10- 1

ot......-

~---"""""'------""""-----""'-------""'-----------I

101

10 2

10 3

10 4

10 5

PREDICTED WAVELENGTH (METERS)

Fig. 7.

Wavelengths of river meanders as measured and as predicted.

being due to an inertial term u( av/ os) in the transverse momentum balance.
The model, using the bank erosion relation of Ikeda et ale [1981], predicts
wavelengths of river meanders that are in general agreement with both
laboratory and field data. The agreement is indeed better than if the full
theory of Ikeda et ale [1981] is used. More importantly, the present theory
can explain the large scatter in Figure 7 as being partly due to the sensitive
nature of the results to the value of the exponential, M, in the streamwise
bedload transport relation.
Although the predicted flow field and bed topography compare well with
laboratory data, care should be taken when using it to simulate field cases.
All the laboratory data used for comparison herein was obtained using fairly
uniform sediment (O'g $ 1.6). This is the appropriate data to use for testing
the theory, in which only the mean particle grain size Ds is used to
characterize the sediment. Preliminary results, not yet reported, indicate that
the model may overpredict scour in the field. The theory should be expanded
in the future to encompass sediment sorting and armoring which hopefully will
improve the predicted bed topography for actual rivers.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

207

Johannesson and Parker


Appendix

Herein, the derivation of the expression for M as ~iven by (9a,b) is


discussed.
Parker fI976], Blondeaux and Seminara [1985J and Parker and
Johannesson [19891 did not assume a simplified streamwise bedload transport
relation corresponding to (6e). Therein, formulas are given for linearization of
arbitrary bedload transport relations which include the expansion of the friction
factor. However, in the present analysis, the friction factor Cf is taken to be
constant as done by Ikeda et ale [1981] and Struiksma et ale [1985].
For
example, if the Engelund- Hansen [1967] bedload transport relation is used, the
full theory yields the result that the coefficient multiplying UI in the linearized
streamwise momentum equation (26a), taken to be two herein, should be 2P,
where

P = - - - - - -1- - - - - - -

(1 + 5JCfG )2(1 - T~/T~) - 5JCfG

(76a)

(lower regime dune-covered bed)

(upper regime flat bed)

(76b)

Similarly, the coefficient multiplying hI in (26a), taken to be one herein, can


be shown from the full theory to be PI, where

(1 + 5JC fG )2(1 - T~/T~)


P1 = - - - - - - - - - - - - (1 + 5~)2(1 - T~/T~) - 5JCfG

(77a)

(lower regime dune-covered bed)


PI

+ 5JCi

(upper regime flat bed)

(77b)

Returning to the expression for M, the full theory gives

M = 2 + 2"

2
(1 + 5~)2(1 - T~/T~) - 5JCfG

(78a)

(lower regime dune-covered bed)

M = 5

(upper regime flat bed)

(78b)

which is simplified herein to the result given by (9a,b). Finally, for the sake
of completeness, it should be noted that the full theory gives an extra term
-trM1hi on the left hand side of (26d). Ml, taken to be zero herein, is given
as

5~

M1 = 2" - - - - - - - - - - - - - - - -

(1 + 5~)2(1 - T~/T~) - 5JCfG


(lower regime dune-covered bed)

Copyright American Geophysical Union

(79a)

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of River Meanders

208

15
MI = 2

JCi

(upper regime flat bed)

(7gb)

If the Meyer-Peter-Miiller bedload transport relation is used the full theory


gives the following expression for M;
,

M = 3 1-

TelT*

(upper regime flat bed)

(80)

to which no simplification is made herein.


The present theory can easily be generalized to include the full expressions
for P, PI, M, and MI. This has been done in the companion paper by Parker
and Johannesson [1989]. There is, however, considerable advantage of using
the simplified expressions in the present model, especially considering that
there is no general agreement on which is the most appropriate bedload
transport relation to use. Further, it is important to realize that the error
introduced by the above simplifications may often be much less than the
variation obtained by using several different bedload transport relations.
Notation

C
Cr

c
C

Cd

Ds
DI,D2

de
F
f*
Go
g
H

coefficient of transverse bed slope, as defined by (34c)


coefficient defined by (44)
coefficients defined by (46c), (53c), and (53e) ,
respectively
half-width of the channel
coefficients defined by (46d), (53d), and (53f),
respectively
dimensionless centerline curvature = bC
centerline curvature
dimensionless Chezy friction factor = gHI/U 2
wavespeed of alternate bars defined by (57b)
coefficient defined by (46e)
damping coefficient defined by (55a)
median size of bed material
coefficients defined by (47c,d)
coefficient defined by (46f)
Froude number = U/ .;gH
calibration coefficient in (13) equal to 1.19
function of ( defined by (31 b)
acceleration of gravity
reach averaged value of ii

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Johannesson and Parker


h

ii
hi

h1C,h 1F
I
1*
JI,J2
k

k
k res
M, MI

n
ii

P, PI
p

Q
Qo

qso
qsl,qnl
R

209

dimensionless channel depth = ii/H


channel depth
perturbation of h
defined by (35)
average channel water-surface slope
(b/H)I
coefficients defined by (53g,h)
dimensionless meander wavenumber
meander wavenumber

bk

resonant wavenumber defined by (55b)


dimensionless coefficients in the equation of sediment
continuity, defined in Appendix.
dimensionless cross-stream coordinate = nIb
cross-st ream coordinate
first moments of the later distributions of u 1F and 1J1F ,
respectively, defined by (48a,b)
dimensionless coefficients in the equation of downstream
momentum balance, defined in Appendix.
sediment porosity
water discharge
ratio between the scale of sediment discharge and the
flow rate, defi ned by (7h)
dimensionless volumetric bedload transport per unit
width in the (s,n) direction
volumetric bedload transport per unit width in the (s,n)
direction
reach averaged value of qs
perturbations of qs and qn
submerged specific gravity of the sediment

kIf

s
S

T
U
u

radius of curvature of the channel centerline


minimum magnitude of r
dimensionless downstream coordinate = sIb
downstream coordinate
dimensionless velocity shape function
reach averaged value of ii
'dimensionless vertically averaged downstream velocity

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Linear Theory of River ~1eandcrs

210
ii
ii
Ut
Utb
T/IC,T/IF
0

00

u1C,u 1F

downstream velocity
vertically averaged downstream velocity
perturbation of u
value of Ut at n = 1
defined by (33)
angle between the centerline down--ehannel direction and
the x-axis
angle amplitude of channel centerline
defined by (35)

v/

transverse velocity
vertically averaged transverse velocity
dimension I ess vertically averaged transverse velocity
perturbation of v
defined by (35)

V
Vt
v IC'V IF

x,y,z
Yt,Y2
0'

0'

Cartesian coordinates,
bed

z being

directed upward from the

defined by (67a,b)
coefficient = 0.077 (4), not to be confused with 0' given
by (57a)
exponential growth rate of alternate bars given by (57a)

(3

exponential growth rate of meander bends given by (75a)


ratio of lift coefficient to drag coefficient = 0.85
coefficient defined by (13)

(3/ ( 12 )

b/H

0'0
0'*

cri tical value of " below which alternate bars are not
present
coeffi cient defined by (32)

(b/H)Cf

Ie

z/h

1Ir

dimensionless bed elevation


bed elevation
reference elevation for the bed

T/t

perturbation of T/

T/

11

meander wavelength

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

211

J ohannesson and Parker

v
ii

Vol. 12

dynamic coefficient of Coulomb friction = 0.43


dimensionless secondary flow velocity = ii/U
secondary flow veloci ty
perturbation of v
eddy viscosi ty
dimensionless water-surface elevation
water-surface elevation
reference elevation for the water-surface
perturbation of
density of water
dimensionless centerline curvature
geometric standard deviation of the bed material
dimensionless secondary current cell strength
phase lag of the secondary flow
dimensionless bed shear stress in the down--ehannel
direction
bed shear stress in the down--ehannel direction
Shields stress defined by (10)
critical Shields stress
grain Shields stress given by (11 a,b)
ks
X - 1/3

(J

1"8
T*
T*
C
T*
G
<1>

a/1Ci
coefficient defined by (30b)
b/f m
circular frequency of meander bends given by (75b)

Acknowledgements
This research was funded by the National Science Foundation, through
Grants No. MSM-8311721-o2 and INT-8412678, and the Legislative
Commission on Minnesota Resources.
References
Ackers, P., and F. G. Charlton, The slope and resistance of small meandering
channels, The Institution of Civil Engineers, Proceedings Supplement XV,
Paper 7362S, 349-370, 1970.
Anderson, A. G., G. Parker, and A. Wood, The flow and stability
characteristics of alluvial river channels, Project Report No. 161, 8t.

Copyright American Geophysical Union

Water Resources Monograph

212

River Meandering

Vol. 12

Linear Theory of Rivel' IVIeanders

Anthony Falls Hydraulic Laboratory, University of Minnesota, 116 pp.,


1975.
Blondeaux, P., and G. Seminara, A unified bar-bend theory of river meanders,
Journal of Fluid Mechanics, 112, 363-377, 1985.
Bordas, M., Contribution a I'etude des relations entre Ie debit generateur et
les meandres d'une riviere, Grafica da Universidade do Rio Grande do SuI,
Porto Alegre, Brasil, 1-78, 1963.
Brownlie, W. R., Flow depth in sand-bed channels, Journal of Hyd1"aulic
Engineering, 109(7), 959-990, 1983.
Chitale, S. V., River channel patterns, J. Hydraulics Div., Am. Soc. Giv. Eng.~
96(1), 201-221, 1970.
Colombini, M., G. Seminara, and M. Tubino, Finite-amplitude alternate bars,
J. Fluid Mech., 181, 213-232, 1987.
De Vriend, H. J., Velocity redistribution in curved rectangular channels, J.
Fluid Mech., 107, 423-439, 1981.
De Vriend, H. J., and H. J. Geldof, Main flow velocity in short river bends,
J. of Hydraulic Eng., Am. Soc. Giv. Eng., 109(7), 991-1011, 1983.
Engelund, F., Flow and bed topography in channel bends, J. Hydraulics Div.,
Am. Soc. Giv. Eng., 100(11), 1631-1648, 1974.
Engelund, F., and E. Hansen, A monograph on sediment transport in alluvial
streams, Danish Technical Press, Copenhagen, 1967.
Gottlieb, L., Three-dimensional flow pattern and bed topography in
meandering channels, Series Paper 11, Institute of Hydrodynamics and
Hydraulic Engineering, Technical University of Denmark, 1-79, 1976.
Ikeda, S., and T. Nishimura, Flow and bed profile in meandering sand-silt
rivers, J. of Hydraulic Eng., Am. Soc. Giv. Eng., 112(7), 562-579, 1986.
Ikeda, S., G. Parker, and K. Sawai, Bend theory of river meanders. Part 1.
Linear development, J. of Fluid Mech., 112, 363-377, 1981.
Johannesson, H., Computer simulated migration of meandering rivers, M. S.
Thesis, University of Minnesota, U.S.A., 1-115, 1985.
Johannesson, H., and G. Parker, Inertial effects on secondary and primary flow
in curved channels, Ext. Memo. No. 208, St. Anthony Falls Hydraulic
Laboratory, University of Minnesota, 1988a.
Johannesson, H., and G. Parker, Secondary now in a mildly sinuous channel,
J. of Hydraulic Eng., Am. Soc. Giv. Eng., in press, 1988b.
Kalkwijk, J. P. Th., and H. J. De Vriend, Computation of the flow in shallow
river bends, J. of Hydraulic Res., 18(4), 327-342, 1980.
Kikkawa, H., S. Ikeda, and A. Kitagawa, Flow and bed topography in curved
open channels, J. Hydraulics Div., Am. Soc. Giv. Eng., 102(9), 1327-1342,
1976.
Kinoshita, R., Investigation of channel deformation in Ishikari River, Report of
Bureau of Resources, Department of Science and Technology, Japan, 1-174,
(in Japanese), 1961.
Kitanidis, P. K., and J. F. Kennedy, Secondary current and river-meander
formation, J. of Fluid Mech., 144, 217-229, 1984.
Langbein, W. B., and L. B. Leopold, River meanders-theory of minimum
variance, U.S.G.S. Professional Paper 422H, 1-15, 1966.
Leopold, L. B., and M. G. Wolman, River channel patterns:
Braided,
meandering and straight, U.S.G.S. Professional Paper 282B, 1-85, 1957.
Leschziner, M. A., ~nd yv. Rodi, Calcu!ation of stron~ly curved open channel
flow, J. Hydraulzcs Dzv., Am. Soc. Gzv. Eng., 105(10), 1297-1314, 1979.
Odgaard, A. J., Meander flow model. I: Development, J. Hydraulic Eng.,
Am. Soc. Giv. Eng., 112(12), 1117-1136, 1986.
Parker, G., On the cause and characteristic scales of meandering and braiding
in rivers, J. Fluid Mech., 76, 457--480, 1976.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

213

Johannesson and Parker

Parker, G., Discussion of "Lateral bed load transport on side slopes," by S.


Ikeda, J. Hydraulic Eng., Am. Soc. Giv. Eng., 110(2), 197-199, 1984.
Parker, G., and H. Johannesson, Observations on several recent theories of
resonance and overdeepening in meandering channels, this volume, 1989.
Quraishy, M. S., The meandering of alluvial rivers, Sind University Research
Journal (Science Series), VII(l), 95-152, 1973.
Schumm, S. A., River metamorphosis, J. Hydraulics Div., Am. Soc. Giv. Eng.,
95(1), 255-273, 1969.
Schumm, S. A., and H. R. Khan, Experimental study of channel patterns,
Bull. of Geological Soc. Am., 88, 1755-1770, 1972.
Sharma, H. D., Extract of the work done at D.P. Irrigation Research
Institute, Roorkee, on meandering, braiding, and avulsion of rivers and their
prevention, V.P.I.R.1. Report, Roorkee, Vttar Pradesh, India, 1973.
Smith, J. D., and S. R. McLean, A model for flow in meandering streams,
Water Resour. Res., 20(9), 1301-1315, 1984.
Struiksma, N., K. W. Olesen, C. Flokstra, and H. J. De Vriend, Bed
deformation in curved alluvial channels, J. Hydraulic Res., 23(1), 57-79,
1985.
Tarnai, N., and K. Ikeuchi, Longitudinal and transverse variations of the
depth-averaged flow fields in a meandering channel, J. of Hydroscience and
Hydraulic Eng., 2(2), 11-33, 1984.
Vanoni, V. A., Sedimentation Engineering, ASGE Manual and Reports on
Engineering Practice--No. 54, 1-745, 1977.
Wiberg, P. L., and J. D. Smith, A theoretical model for saltating grains in
water, J. Geophysical Res., 90(C4), 7341-7354, 1985.
Wolman, M. G., The natural channel of Brandywine Creek, Pennsylvania,
U.S.G.S. Professional Paper 271, 1-56, 1955.
Wolman, M. G., and L. M. Brush, Factors controlling the size and shape of
stream channels in coarse noncohesive sands, U.S.G.S. Professional Paper
282G, 183-210, 1961.
Zimmermann, C., and J. F. Kennedy, Transverse bed slopes in curved alluvial
channels, J. of the Hydraulics Div., Am. Soc. Giv. Eng., 104(1), 33-48,
1978.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Copyright 1989 by the American Geophysical Union.

Studies on Qualitative and Quantitative


Prediction of Meander Channel Shift
Kazuyoshi Hasegawa

Department of Civil Engineering, Faculty of Engineering,


Hokkaido University

Introduction
River engineers have considered the problem of meander channel changes
for many years, motivated by the practical problem of land erosion. It was
not until Friedkin's [1945] work, however, that a serious study was begun.
Herein a review of studies on qualitative and quantitative predictions of
channel shift is presented. In addition, the writer's recent work toward the
development of a universal bank erosion coefficient is discussed.
As excellent reviews on the problem of meandering have already been
presented by Callander [1978], Ashida [1982], etc., and as reference books by
Allen [1984], Gregory [1980] etc. are available, only the minimum number of
papers are reviewed herein.

Qualitative and Quantitative Studies on Channel Change

Feature of Meander Channel Change


I(inoshita [1961] investigated alluvial meandering rivers in Japan and found
the following results: 1) Alternating bars appear to be a universal feature of
alluvial rivers. 2) Two meander patterns are found in the rivers-one consists
of a sinuous channel meander with two alternate bars within one channel
meander wavelength, and the other consists of a tortuous meander containing
three or more alternate bars within one channel meander wavelength. 3) In
the case of sinuous meanders, the whole bend migrates downstream at a
Individual parts of a tortuous
uniform rate, as indicated in Figure 1.
meander, however, show differential shift, so that loop is deformed so as to
create apparent "immovable points," as shown in Figure 2.
Kinoshita
suggested that the difference between the two kinds of meander development is
related to the way in which the primary flow impinges upon the banks, as
indicted in Figures 3 and 4. Erosion occurs along the bank upon which the
primary flow impinges. This impingement may be caused by alternate bars
and bend shape. These findings have proved useful for enabling qualitative
predictions of channel shifts.
Other studies of meander growth patterns have been done by Brice [1974,
1983], Hickin [1974], Hickin and Nanson [1975], Jackson [1975] and others.
Though these investigators used different terms, they commonly classify
meander patterns as (a) simple symmetrical, (b) simple asymmetrical, (c)
215
Copyright American Geophysical Union

Vol. 12

Water Resources Monograph

River Meandering

Vol. 12

Meander Channel Shift

216

Equi-directional and equi-distant change

Fig. 1.

Channel change patterns in a sinuous meander [after Kinoshita, 1961].

Itaya Farm

Fig. 2.

Channel change
Kinoshita, 1961].

pattern

Deepest point

in

tortuous

meander

(Uryu

River)

Front of bar

Primary
flow
Fig. 3.

Flow and bed form in a sinuous meander [after Kinoshita, 1961].

Copyright American Geophysical Union

[after

Water Resources Monograph

River Meandering

217

Hasegawa

Fig. 4.

Vol. 12

Flow pattern in a tortuous meander [after Kinoshita, 1961].

compound symmetrical, (d) compound asymmetrical, and (e) "lobing" or


"double heading". Furthermore, they detected the following growth patterns:
(a) extension, (b) translation, (c) rotation., (d) enlargement, (e) lateral
movement, and (f) complex change.
These classifications, however, are
essentially taxonomical in nature, since the conditions necessary for generating
the patterns are not related to the flow mechanics.
Hickin's f1974] investigation of channel shift used relic arcuate scroll bar
patterns visible on aerial photographs. A dendrochronological technique was
used to date the scroll bars. Lines drawn orthogonal to relic scroll bars, and
thus old channel banks, allow for tracing of the pattern of shift, from which
the true migration rate can be measured.

Observations of Channel Migration Rates


Migration rates of natural river channels have been measured by many
investigators [e.g. Allen, 19841. Hooke [1980] obtained a regression relation
between the migration rate ot a meander loop Y (mjyear) and the upstream
drainage area A (km2 ) by using her own data as well as those of others:
or

=
Y

0.0867

0.00114A

0.0245A 045

The approximate relation of Y to the square root of A implies that Y should


increase in proportion to the channel width B, because the square root of A
can be expected to be in proportion to B. This suggests that bank erosion

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Meander Channel Shift

218

rates are similar for basins of all sizes if normalized in terms of channel
widths per year, as Hooke pointed out.
This relation is, however, rather
broad, since many factors other than A which influence migration rates are not
accounted for therein.
After investigating the relation between Y/B and
various parameters expected to influence channel migration rate (section length
of the site, width/depth ratio, percentage of silt and clay in the bank
material, channel slope, bend radius of curvature, bank height, existence of a
gravel layer, etc.), Hooke [1980] found that the percentage of silt and clay in
the bank material was the dominant factor for 11 streams in Devon. It was
also found, however, that erosion may be influenced by a complex combination
of other factors.
On the other hand, Hickin and Nanson [1975] and Nanson and Hickin
[1983] obtained channel migration rates Y for the Beatton River by applying
the methods of dendrochronology to trees growing on the scroll bar. They
found the following relationship:
Y
Y

= 2.0
= 0.2

B/r

(B/r ~ 0.32)

riB

(B/r > 0.32)

Hickin and Nanson [1984] further suggested that the coefficients relating Y
and B/r are closely related to the texture of the bank materials. Their work
is significant in that they evaluate bank erosion coefficients based on actual
data for migration rates. It is nevertheless difficult to consider the radius of
curvature of the channel as the sole factor governing bank erosion.
Quantitative Studies on Channel Changes
The study by Ikeda et ale [1981] and the previous work by Ikeda et ale
[1976] can be called a pioneering work toward a mechanistic analysis for
meander channel changes. The analytical basis for their theory consists of the
assumption that the bank erosion rate at any point is in proportion to the
near-bank perturbation of depth-averaged flow velocity (that is, the difference
between the near-bank and the centerline value). This idea is in accord with
Kinoshita's [1961] phenomenological interpretation of bank erosion, since the
velocity perturbation becomes positive near banks toward which the primary
flow converges, and becomes negative near banks away from which the flow
diverges. Parker et ale [1982, 1983] advanced this theory by considering a
non-linear formulation, and succeeded in deriving the development process of a
meander loop from a simple symmetrical form to a compound asymmetrical
form.
Furthermore, Parker and Andrews [1986] obtained a higher order
solution for the time growth of meanders by using a finite-amplitude method.
They demonstrated there is no stable high-amplitude state to which meander
bends generally tend.
Channel shift predictions in field meandering rivers were attempted by
Hasegawa et ale [1978], Parker [1983], Beck [1983], and Johannesson and Parker
[1985J, by using the same fundamental equation system as in the papers
quoted in the paragraph above. These studies indicated that if a bank erosion
coefficient and some other appropriate parameters could be evaluated
adequately, rather good predictions can be obtained. However, it was known
by Johannesson and Parker that the calibrated coefficients for bank erosion
change depending on whether the bank being eroded is in a forested area or
for
not.
For forested areas, the coefficient is about half of the value
non-forested areas. Very recently, Odgaard [1987] carried out a similar study
of stream bank erosion, and obtained a constant, connecting the erosion rate
with the near-bank perturbation of the primary flow velocity. His data also

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

219

Hasegawa

demonstrate that mature trees in a bend lower the constant. The influences
of the geotechnical properties of the bank materials on the erosion coefficient
is also important, as is discussed in a later section.

Summary of Channel Shift Mechanisms; Future Topics


In Figure 5, a chart due to Hasegawa [1983] illustrates the
interrelationships among factors governing channel changes in alluvial rivers.
The arrow marks indicate the influences among various factors; the degree of
association is expressed by the thickness of the arrow.
The individual
relationships in the chart more or less follow the findings summarized in the
previous sections.
According to the chart, a sinuous channel planform (i.e. channel meander)
causes a secondary flow within a cross section (a / ).
The secondary flow
changes the bed topograRhy (b,C), but does not strongly influence bank erosion
or accretion directly l b ').
Channel planform change occurs rather by
convergence of the thread of high velocity to one side bank, i.e. by the
Thus, the influence of
deviation velocity uB illustrated in Figure 7 (B ').
channel planform on bank shift is seen to be indirect: the planform gives rise
to bank erosion or accretion by means of the associated variation in bed
topography.
However, channel planform does play an important local role in the
formation of planimetric flow separation (a) along the inside bank. When this
occurs, it can yield strong deposition of suspended sediment. This phenomenon
can be viewed as a direct effect of channel planform on bank shift.
Recently, the above interpretation of the mechanisms of channel shift has
been confirmed by detailed field measurements at natural meander river sites.
Several important questions, however, remain:

(1) Is it correct that the direct effect of secondary flow on bank erosion
(b / ) is weak? Associated with this problem is the issue as to whether or not

bank erosion is governed by the differential primary flow velocity (Figure 7) as


Fukuoka and Yamasaka [1982] suggest. The major controls on bank erosion
remain unclear at present. The investigation of Lapointe and Carson [1986] on
the Rouge River suggests that at the bend entrance, the nascent secondary
flow helix causes outer bank erosion, whereas in the zone of developed bend
flow, the intensity of near-bank primary velocity, as shown in Figure 7, is the
main control on bank erosion.
They furthermore emphasize that "greater
bank--erosive stresses may generally be accompanied by stronger helix flows".
Therefore, the relations illustrated in Figure 5 are considered to be basically
correct, but probably need some modifications for more detailed analysis.
(2) It is known in general that deposition of suspended sediments in the
separation zone helps construct point bars along or on the floodplain behind
Existing knowledge concerning the accretion of bank
the inner bank (B ').
sediment is, however, extremely limited.
Kinoshita [1987] investigated sedimentary structures along the inner bank of
an old oxbow loop (Horseshoe Swamp) in the Teshio River, Hokkaido, Japan,
where scroll bars are well developed. By excavating a trench 3 m deep and
50 m long, orthogonal to the bankline, it was possible to visualize the
depositional sequence. Figure 6 shows a view of the entire trench. Kinoshita
analyzed the sediment structures under the surface of the inner bank shown in
Figure 7, and deduced that a scroll bar may be formed from an embryonic
sand bar (at the core of each scroll bar). Each such sand bar is generated by
the deposition of suspended sediment swept inward due to the action of
Iarge--scale, near-bank separation vortices stretching downstream from the apex

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

t-.:l

o
D

Flow Convergence

Sand \~aves
Generation

Spatial

Banks

I Differential
r----

c:::>

Equilibrium
Bed

.L:.
+.J
"C

3:

Qj
c:::
c:::
to

.L:.
U

Confi ned
Pl an forms

i'

r::::>

Sinuous
Pl anforms

c'

Accretion

c::::::>1

-~

s=

(t)
~

0..
(t)

Fig. 5.

Chart of mechanisms associated with channel shift of alluvial rivers.

oP"

r.n

e:

:;::::a

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Hasegawa

Fig. 6.

Vol. 12

221

Banded accretion structures seen on a trench wall at Ubushi, along the


Teshio River.

of an inner bank. According to Nanson and Hickin [1983], a very long time,
even over 100 years depending on circumstances, is required for a scroll bar to
grow sufficiently to adjust (reduce) channel width. During this time, retreat
of the outer bank may stop or weaken. This suggests that accretion along an
inner bank, often regarded to play a negative role as regards channel
migration, may in fact have an important positive role. By reducing channel
width, it may increase velocity, so generating a scouring force along the outer
bank. Inner bank deposition may thus partially control the migration rate.
Therefore, the mechanism (a) in Figure 5, and the process of accretion of
suspended sediments should be investigated in more detail.
Furthermore, the following important and interesting problems remain:

(3) Improvements in the solution for flow in bends, especially one in


which the convection terms are retained, are necessary for obtaining a more
precise estimate of the bank erosion rate (B ').

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

t-.:)
t-.:)
t-.:)

Hatched area: Fine sand layers thickly concentrated


Dotted area: Layers with similar tendency

Ground surface of the right bank


at the time of channel cut off

5
E

a;
>

Q)

..."0

c:

s..

(.!)

3
-=T

40

30

20

10

Distance from the water margin in transverse direction (m)


~

Fig. 7.

Accretion structures observed on the trench investigated at East Ubushi on


the Teshio River (presented by Rumoi Construction Head Office of Hokkaido
Development Bureau, after Kinoshita, [1987]).

ro

0.-

ro
"""S

aP"'
~

~
~

en
P"'

S;

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Hasegawa

Vol. 12

223

(4) A more exact prediction of bed topography, including alternate bars,


would improve the accuracy of the flow solution.
After carrying out sixty replicate experiments of meander initiation and
development under the same conditions, Wallis [1978] concluded that random
behavior associated with meanders is due to shoals and ripples of sediment in
the bed, which cause small lateral deflections in the flow. Therefore, it is
very important to predict the precise bed topography in order to obtain the
correct flow solution and meander shift.
(5) The relation between scouring force and erosion rate must also be
determined.
Cunge [1983) remarks that there is no satisfactory operational
model simulating meandenn phenomena, and that the physical link between
some "erodibility coefficientsft and reality is lacking. It is therefore important
to elaborate upon the physical mechanism of bank erosion. This very problem,
discussed by Hasegawa [1988], will be taken up in the following section.

Equations Governing Bank Erosion; Universal Erosion Coefficient

Equations Governing Bank Erosion


In order to estimate bankline change correctly, the continuity equation of
sediment should be applied to a control volume including the bank water
margin. Where the symbols are defined in Figure 8, this equation can be
writ ten as follows,

(1)
Here, s = downstream distance along the channel centerline, n = transverse
horizontal distance, r = the planform radius of curvature along the channel
centerline, and TJ = vertical bed displacement about an average bed, such that
positive TJ corresponds to a downward displacement.
Furthermore, qs and
qn = the s- and n-eomponents of the volumetric bed load transport per unit
width, respectively, A = the porosity of the bed and bank sediment, and t =
time.
Equation (1) is now integrated from nT to n B of Figure 8, after neglecting
the third term in the parenthesis on the right-hand side. The following
relations are applied at the water margin:

(2)

(3)
where ( denotes the rate of channel shift normal to the centerline.
The
following equation governing bank migration rate can thus be derived from (1),
(2), and (3):

(4)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Meander Channel Shift

224

B
2

Fig. 8.

B
2

.--+--~----t""------u

Definition sketches of the coordinate system and symbols used.

Here, Ho = - 1/(nB ) is the average water depth, and hf = the height of the
emergent part of the outer bank above the mean water surface.
Assuming the bank material to consist of sand and silt, then s- and ncomponents of the bedload can be expressed by the followin,g formulas. Herein
qs is given by the formula of Meyer-Peter and MUller [1948J:

(5)
The transverse bedload transport rate qn can be obtained by multiplying qs by
the tangent of the angle of deviation of the direction of motion of a sand
particle to the s-axis. For example, the formula of Hasegawa [1981] yields:

(6)
Here, p and Ps = the density of water and sand material, respectively, d
typical diameter of a sand grain, g = acceleration of gravity, K = a
dimensionless coefficient, T* = nondimensional tractive force, T *c

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

22.5

IIasegawa

nondimensional critical tractive force on a plane bed, v = the transverse


component of the flow velocity averaged over the water depth, Uo = the
cross-sectionally averaged downstream flow velocity, and 'l/J = deviation angle
of direction of the near-bottom flow from the s-axis due to secondary
currents.
Furthermore, T = ~T*c/(JLs Ilk T*0), JLs and Ilk = the static and
dynamic Coulomb friction coefficients of sediment particles, respectively, and
T *0 = the cross-sectional mean value of T*.
The dimensionless tractive force T* can be denoted in terms of the depth
averaged flow velocity in conjunction with the shallow water flow model:

U2 --

1
C
T* - (Ps/p - l)gd r

C
ro

(Ps/p - l)gd

[H]
IiO

1/3

U2

(7)

Here, U = the s-eomponent of the mean flow velocity averaged over the water
depth at a point; Cr and Cro = the local and mean resistance coefficient of
the flow, respectively; and Hand Ho = the local and mean water depths,
respectively. The minus one-third power dependence in (7) is derived from
the assumption that the local resistance coefficient follows the ManningStrickler formula.
Next, the terms U and H are perturbed about a base state Do, Ho; U =
Uo + u, H = Ho + 11 + e ~ Ho + 11, as illustrated in Figure 8. Substituting
into (5), (6), and (7), and furthermore neglecting higher order terms, the
following can be deduced for the bank erosion rate.

~ _
~ -

gs 0
(1-I\)H o

a [!L]]
_
os
Ho
~ n-n

{~[a [.!!...-] _ 1
~ 7J3 Uo
()
~

~o

T tan Ok [3J*

'--v---/

[~ rP*

it - ~f]

+ 1]

- [uv

n=n

'-...r/

(8)

+ tan'l/J] n-n
_ .}

~,....--J

In the above relation, the integral of Oqs/ as, with respect to n occurring in
(4) has been approximated by the product ot the integrand and the term (n B n ); also hr/Ho has been neglected insofar as hr is relatively small in many
T
alluvial rivers. Furthermore, Ok = i~ the average transverse slope angle of the
concave
bank
in
an
incipient
state
of
erosion,
and
qso =
K~(ps/p - l)gd 3 (T*0-T*c)I.5, T*o

Cr Uo/{{Ps/p - l)gd), J* = T*o/{T*o-T*c).


According to the above equation, the bank erosion rate is essentially
controlled by the ratio qso/llo. This means that the bigger the value of qso
(Le., the larger the river), the larger is the absolute migration rate that can
be expected. Likewise, a shallower river is subject to faster channel migration.
Let us discuss (8) in more detail, in terms of the effectiveness of terms (1)
- (6) in determining bank erosion rates. The terms may be explained as
follows: <1> longitudinal rate of change of flow velocity, <2> longitudinal
rate of change of bed elevation, <3> the relative magnitude of the perturbed

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Meander Channel Shift

226

component of flow velocity near the bank (corresponding precisely to the term
used by Ikeda et ale [1981]), <4> relative depth of bed scour, <5> relative
bank height, and <6> the relative magnitude of the transverse component of
near-bottom flow velocity.
In the case described in Figure 8, the terms <1> and <3> often take
positive values; <6> also becomes a positive, since the transverse flow near
the bed typically has a negative (Le. inward) direction. These factors can be
said to cause the erosion.
The other factors often take negative values,
effectively delaying the erosion process.
In order to determine which of the above six factors dominate in terms of
order of magnitude, the terms are compared usin& the following constants: tan
Ok = J-lk = 0.7, J.ls = 1.0. Assuming that T*o/T*c ) 1 at bankfull stage, it
follows that the values tP* ~ 1 and T ~ 0.5 should be used. Furthermore,
considering a regular meander channel with erosion taking place at the concave
bank, the condition u(n T ) > 0 can be taken to hold at the top of the concave
bank.
Let Ho/B = 8 1; then assuming u(nT )/U o - 0(8), and B/L - 0(8), the
order of magnitude of each term can be estimated as follows:

0(82), term <3>

term <1>

0(82), term <2>

0(8),

term <4>

0(8), term <5> = 0(82), term <6> = 0(82).

The observed values of tan 8k and 7J(nT )/H o in the Ishikari River were
used for the purposes of the above evaluations. It is seen that terms <3>
and <4> are dominant in (8). Indeed the term <3> directly causes erosion,
confirming the assumption ( (X uB ' where uB ~ u(n T ) (see Fig. 8). On the
other hand, term <4> acts to delay erosion. As is evident from (4), the
height of the banks does not directly arrest erosion, but rather works only to
decrease the erosion rate. Therefore, the term <4> should not exceed <3>,
and can be left out of consideration. This notwithstanding, one cannot ignore
the influence of the total height of the banks for all reaches of all rivers.

Universal Bank Erosion Coefficient


If ( is approximated only by term <3> of (8), the following equation can
be obtained, setting u(n T ) ~ u B .
i- ':. -

39so T tanOk * u
(I-X) H o U o

(9)

Substituting the relations based on averaged values with respect to (5) and (7)
in (9), and taking 10 to be an average bed slope along the channel, (10) can
be obtained;

(=

{Ci 10

3KT tanOk

] u

(I-A)(Ps/p - 1)1?);

(10)
B

The terms in the brackets on the right-hand side of (10) are, with
exception of the parameter T*0 implicit in T and tP*, dependent
geotechnical characteristics of the banks. The parameter within the
can thus be taken to be a constant reflecting the essential nature of

Copyright American Geophysical Union

the sole
on the
brackets
the bed

Water Resources Monograph

River Meandering

Vol. 12

227

Hasegawa

and banks of each river, since T and J* may not change significantly as
erosion takes place during a flood. Representing the bracketed parameter as
E*, (10) can be rewritten as follows:
(11)
In a general expression for bank erosion, the slope 100 defined along a valley
axis is more useful than channel slope 10 = X3 I , because 10 changes as
00
meanders deform. Here X is the ratio of the mean flow velocity Uo in a
meandering channel of given configuration to the mean flow velocity U which
00
would occur if the channel were straight and oriented along a valley axis, but
otherwise unmodified. Thus, (11) is rewritten as:

(12)
The bank erosion coefficient introduced by Ikeda et ale [1981] is

(13)
from which it follows that

(14)
It is clear that Eo is not solely dependent on the properties of the bank soil.
The real migration rate (p during floods in excess of some standard stage
near bankfull discharge should be expressed as,
(15)
Here Pr is the fraction of time the flow is in excess of this standard stage,
herein taken to be bankfull discharge. This definition facilitates comparison
among different rivers.
A bank erosion coefficient that corresponds to the migration rate defined by
(15) may be expressed as follows:

(16)
Comparison of ( with Estimated Values of uB
The near-bank perturbed downstream flow velocity u B is defined in Fig. 8.

Hasegawa et ale [1978], and, independently, Parker [19831, derived an equation

to estimate the downstream variation of uB in a channel with arbitrary


meander form by improving Engelund's [1974] solution. Their result may be
expressed in the following form, with the use of the coordinate system shown
in Figure 8.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Meander Channel Shift

228
o

80

Un

m/year

ffi/S

.4

I
I

0 t---------

- 8

,
,
I
,

r------I-----I-.-----1

1932 -

-.2

\ I
\ I

1939

\
~

-.4

-12

-.6

80

81

82
Fig. 9.

Ishikari
Ri ver

.2

Ii

Comparison between measured ,


Ishikari River.

f if

and calculated u

79
B

KM

along a reach of the

UB

Uoo

= -XC(S) 2" +
B

exp{-

2i

{(A + 2)X2 + F2 X5 }

00

Cf

(s - S/)} C(S/) ~ ds l

(17)

00

Here H denotes the mean water depth in an otherwise identical straight


00
channel oriented along the valley axis. Also, F2 = U2 j(gH ), and A = a
00
00
scouring factor expressing the degree of the transverse bed inclination caused
by secondary flow, and defined in Ikeda et ale [1981] and Parker [1982].
In order to test the relations developed herein, actual migration rates (
were determined from maps of different dates. The main sources of data were
maps of channel shift of the Ishikari River collected and arranged by
I(inoshita [1961]. For example, Figure 9 shows a comparison of u calculated
B

from (17) at bankfull flow with observed values of ( along a reach extending
from 79 to 81 km upstream of a standard point on the Ishikari River during
the period 1899-1911. The correlation between ( and u B is fairly good.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

229

Ifasegawa

This adds credence to the assumed validity of (17), (11), and (12). However,
the clear relation seen in Figure 9 is less evident over much longer reaches.
The correlation coefficients between measured (and calculated u B obtained in
40 reaches of the Ishikari, Uryu, and Mukawa Rivers range from 0.15 to 0.99.
The reaches with correlation coefficients in excess of 50% occupy less than 50%
of the total length of reaches considered, when several reaches for which data
were difficult to quantify due to braiding, cutoffs, tributary junctions, artificial
influences, etc. are included. This behavior has been studied only for the
three abov~mentioned rivers in Hokkaido; the results may, however, be
applicable to other alluvial rivers.

Characteristics of Bank Erosion Coefficients


Estimated from the Universal Formula
As mentioned in the previous sections, the characteristics of bank resistance
to erosion are expected to differ in each reach even along the same river.
Herein, the relation between the coefficient E*p and nature of the bank soil is
investigated. For the purpose of this study, a total of 32 reaches (including
several identical places at different dates) for which the correlation coefficient
between ( and uB exceeds 0.4 were chosen from among data for the Ishikari,
Uryu and Mukawa Rivers. The mean velocities and depths at several gauging
stations in the realigned channels at bankfull stage were used to estimate U
00
and H in the computation for uB ' because the realigned channel passes near
oo
the valley axis. The bed slope 10 was determined by dividing the difference of
bed elevations from beginning to end of each meander reach by the channel
length. Likewise, X was computed as the cubic root of the ratio of the length
along the valley axis to the length along the channel. The values of Pr were
estimated by equating the bankfull discharge with the mean yearly snowmelt
flood discharge. These values were estimated from the daily discharge records
spanning 15-29 years in the Ishikari River, and spanning 29 years in the Uryu
an Mukawa Rivers.
Eo was estimated from (13) by applying the method of least squares to
everyone or two meander reaches. This value was then transformed into E*p
by means of (14) and (16).
At this point, it is necessary to decide upon a parameter with which to
characterize the nature of the bank soil. Figures 10 and 11 illustrate the
distributions of the average N value of the standard penetration test at various
points, and for various bank material, along the Ishikari and Uryu Rivers.
These data were taken from boring tests at the time of levee construction.
The Ishikari River may be divided into two alluvial reaches, delta and
transition, near the 62 km point shown in Figure 10. The former reach is
rich in clay and silt, rendering the rate of channel shift rather modest. In the
latter reach, however, fine sand is abundant, and vigorous channel shift is
observed.
In Figure 10, the average N values are seen to decay in the
downstream direction (in spite of local variations), in accordance to the
constitution of the banks.
The data for channel shift were collected at the
places numbered in the transitional reach shown in Figure 10.
Since the
surveyed points are several hundred meters distant from the present low water
channel, it cannot be guaranteed that the points were in accord with the
channel bank when the data were taken.
Figure 12 indicates the relation between Nn and E*p.
Although the
plotted points show scatter, the data may be grouped into two sets. One

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

l~

c,...J

+J
II)

No

Q.J

Delta Reach

-1~4----_-

.~
+J

ro

9
1'1

40

s...

+J

Q.J

OJ

0...

'U

s...

ro

30

so

.~

0:
.... 0:

>
~

10

V)

000

>,

20

'U

N
so

clayey
silt
si 1 ty (':.,

ro

::>

~
Z

Region No.
for
r--1 data sampling
1

r-;--r

.~.

gravel

clay

ro

rI

.s::.

san~
,~

+J

n-I
N

<c::: OI;FEFR
fi ne

4-

r--l

si 1 t

so

'U

V)

Transitional Reach

10

Q.J
O'l

ro

s...

106.6KM

QJ

>

c::(

40

60

80

100

Distance along the improved channel from river mouth

120

KM

s=

(t)

0.-

Fig. 10.

Distribution of averaged N-values along the Ishikari River.

ot:r

~
~

en

e:

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

231

Hasegawa

No

+oJ
III

OJ

r--1

l-

3 2 1
I I I

654

Reg i on No.
for
data sampling

40

+oJ
10
~

+oJ

OJ

OJ

0-

30

"0
~

sandy
gravel

10

"0

10
+oJ
V)

gravel

20

40

22.1 KM

OJ
:::J

~
>
I

10

Z
"0
OJ

0)

10
~

OJ

>

KM

20

10

Distance along the improved channel


Fig. 11.

Distribution of averaged N-values along the Uryu River.

group is from the region of confluence of three major tributaries along the
Ishikari River, and the lower reaches of the Uryu River. Within this group,
clayey soils prevail in banks. The other group is from the middle to lower
reaches along the Ishikari River, and along the lower reaches of the Mukawa
E*p
10

gravel silt River


sand clayey

!:l

Ishikari
Uryu
Mukawa

0
0

Fig. 12.

12

16

20

24

28

No

Relation between univeral erosion coefficient and averaged N-value.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

l\1eander Channel Shift

232

Both groups exhibit a


River, where gravelly sand prevails in the banks.
tendency for E*p to decrease as ND increases. This means that the greater
the compressive strength of the bank materials is, the stronger is the resistant
strength of the bank. It should be noted, however, that a gravelly sand bank
is more easily eroded than a clayey bank. Furthermore, it is seen that the
lower curve in Figure 12 drops steeply near the smallest values of ND' because
E*p takes smaller values in the delta reach.
Since the values of Pr treated in this study do not differ largely among the
rivers and reaches under consideration, the relation between E* and ND can be
expected to be scarcely different from the relation between E*p and N in
D
Figure 12. Furthermore, E* indicates values from 10- 3 to 10- 2 in the Ishikari
River, and near 10- 3 in the Uryu and Mukawa Rivers. These values are
nearly equal to or greater than the value of 1.03x 10- 3 determined by Pa7'ker
et ale [19861 for the Beatton River, Canada.
It can be judged from these facts that E* and E*p maintain values of a
similar order of magnitude among a variety of rivers. These parameters thus
provide a universal means for characterizing the coefficient necessary to predict
bank erosion rates.

Notation

A
B
C
Cf, Cfo
d

Eo
E*
E*p

E*/Pr

g
H, Ho
H

00

hf
10
1

00

K
N

scouri ng factor
channel width
curvature
local and mean resistance coefficient of flow.
typical diameter of a sand grain
bank erosion coefficient
general bank erosion coefficient
general coefficient for the real erosion rate during a flood
gravitional acceleration
local and mean water depth of a channel
mean water depth in a straight channel along a valley
axis.
height of a bank over the water surface
average bed slope along a channel
average valley slope
coefficient in the relation for bed load
spatial mean of N-values averaged over depth from bank

surfaces to the deepest point of the channel bottom


value defined according to the standard penetration test

transverse horizontal axis

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

I-Iasegawa

Vol. 12

233

n--eoordinate corresponding to the water margin of the


eroded bank and the deepest point of the cross section.
Pr

qso

s
t
V

fraction of time that the flow exceeds bankfull discharge.


components of sediment discharge in the s- and
n-directions, respectively
mean of qs over the cross section.
radius of curvature along the channel centerline.
axis along the channel centerline; the origin is taken to
be
at the apex of a bend.
time.
s-direction component of mean flow velocity averaged

over the water depth.

(= y
(p = (/Pr
TJ

mean flow velocity over the cross section.


mean flow velocity in a straight channel along the valley
axis.
perturbation of V about its transverse average.
near-bank value of u.
n-direction component of the depth-averaged flow
velocity.
bank erosion rate (channel migration rate)
real erosion rate during a flood.
displacement of the bed surface about the sectionally
averaged bed surface elevation.
average transverse slope angle of the concave bank.
porosity of the bank and bed material.

p, Ps

static and dynamic Coulomb friction coefficient of


sediment particles.
displacement of the water surface from the sectionally
averaged level.
density of water and sand grains.
nondimens ional tractive force.
nondimensional critical tractive force.
average of T* over the cross section.
T*o/ (T*o - T*c).

Vo/V 00
deviation angle of a near-bottom stream line from s-axis.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

234

Vol. 12
~1eander

Channel Shift

Acknowledgements
This study was accomplished under the U.S. - Japan Cooperative Science
Program, as part of the research project: Development and Application of the
Theory of River Meandering.
The Japan Society for the Promotion of Science provided funds for research
performed in Japan. This aid is gratefully acknowledged.

References
Allen, J. R. L., Sedimentary structures, Their character and physical basis,
Developments in Sedimentology 30, Elsevier.
Ashida, K. (head of the committee), Report of the Committee on ThreeDimensional Patterns of Flood Flow and Channel Configurations, Special
Committee of Japan Soc. Civ. Eng. (in Japanese), 1982.
Brice, J. C., Evolution of meander loops, Geol. Soc. of Am. Bull., 85, 1974.
Brice, J. C., Planform properties of meandering rivers, Proceedings of River
'83 Specialty Conference on River Meandering, New Orleans, 1983.
Beck, S. M., Lateral Channel Stability of the Pembina River Near Rossington,
Canada, Research Council of Alberta, Edmonton, Canada, 1983.
Callander, R. A., River meandering, Ann. Review Fluid Mech., 10, 1978.
Cunge, J. A., Feasibility of mathematical modeling of meanders, Proceedings of
River '83 Specialty Conference on River Meandering, New Orleans, 1983.
Engelund, F., Flow and bed topography in channel bends, Jour. Hydr Div.,
Am. Soc. Civ. Eng., 100, No. HYll, 1974.
Friedkin, J. F., A laboratory study on the meandering of alluvial rivers, U. S.
Army Corps of Engineers, Waterways Experiment Station, 1945.
Fukuoka, S. and Yamasaka, M., Theoretical study on meander development
caused by bank erosion and deposition, Proceedings of Japan Soc. Civ.
Eng., 327, 1982.
Gregory, K. J. (Editor), River Channel Changes, Wiley.
Hasegawa, K. and Itoh, H., Computer simulations on meander channel
changes, Proc. of the Hokkaido Branch of Japan Soc. Civ. Eng., 34 (in
Japanese) 1978.
Hasegawa, K., Bank-erosion discharge based on a non-equilibrium theory,
Trans. of Japan Soc. Civ. Eng., 13, 1981.
Hasegawa, K., A Study on Flows, Bed Topographies and Plane Forms of
Alluvial Meanders, Thesis, Dr. Engrg., Hokkaido Univ. (in Japanese), 1983.
Hasegawa, K. Universal erosion coefficient of meander banks, Jour. Hydr.
Engrg., Am. Soc. Civ. Eng. (submitting 1988)
Hickin, E. J., The development of meanders in natural river channels,
Am. Jour. Science, 274, 1974.
Hickin, E. J. and Nanson, G. C., The character of channel migration on the
Beatton River, Northeast British Columbia, Canada, Geol. Soc. Am. Bull.,
86, 1976.
Hickin E. J. and Nanson G. C., Lateral migration rate of river bend, Jour.
Hydr. Eng., Am. Soc. Civ. Eng., 110, No. 11, 1984.
Hooke, J. M., Magnitude and distribution of rates of river bank erosion, Earth
Surface Processes, 5., 1980.
Ikeda, S., Hino, M. and Kikkawa, H., Theoretical study of the free meandering
of rivers, Proc. Japan Soc. Civ. Eng., 255 (in Japanese), 1976.
Ikeda, S., Parker, G. and Sawai, K., Bend theory of river meanders. Part 1.
Linear development, Jour. Fluid Mech., 112, 1981.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

.235

llasegawa

Jackson, R. G., Velocity - bed form - texture patterns of meander bends in


the lower Wabash River of Illinois and Indiana, Ceol. Soc. Am. Bull., 86.,
1975.
Johannesson, H. and Parker, G., Computer simulated migration of meander
river in Minnesota, Project Report No. 242, St. Anthony Falls Hydraulic
Laboratory, University of Minnesota, prepared for Legislative Commission
on Minnesota Resources, State of Minnesota, 1985.
Kinoshita, R., Investigation of channel changes on the Ishikari River, Report
for the Bureau of Resources, No. 36, Science and Technology Agency.
Japan (in Japanese), 1961.
Kinoshita, R., Investigation on alluvial actions in a flood and experimental
study on an optimum designed channel with complex cross section, Report
for "Studies on control of flood flows in alluvial rivers and improvement of
the safety for river training", Grant-in-Aid for Developmental Scientific
Research of MESG (Principal Investigator Tsutomu Kishi), in Japanese,
1987.
Lapointe, M. F. and Carson, M. A., Migration patterns of an asymmetric
meandering river: The Rouge River, Quebec, Water Resour. Res., 22, No.
5, 1986.
Meyer-Peter, E. and Muller, R., Formulas for bedload transport, Proc. 2nd
IAHR Meeting, Stockholm, 1948.
Nanson, G. C. and Hickin, E. J. Channel migration and incision on the
Beatton River, Jour. Hydr. Eng., Am. Soc. Giv. Eng., 109, No.3., 1983.
Odgaard, A. J., Streambank erosion along two rivers in Iowa, Water Resour.
Res., 23, No.7, 1987.
Parker, G., Sawai, K, and Ikeda, S., Bend theory of river meanders. Part 2.
Nonlinear deformation of finite-amplitude bends, Jour. Fluid Mech., 115,
1982.
Parker, G., Diplas, P. and Akiyama, J., Meander bends of high amplitude,
Jour. Hydr. Eng., Am. Soc. Giv. Eng., 109, No. 10, 1983.
Parker, G., Theory of meander bend deformation, Proc. of River '83 Specialty
Gon! on River Meandering, New Orleans,1983.
Parker, G., and Andrews, E., On the time development of meander bends,
Jour. of Fluid Mech., 162, 1986.
Wallis, I. G., The random component in stream meandering, Water Resour.
Bull., Am. Water Res. Assoc., 14, No.3, 1978.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Copyright 1989 by the American Geophysical Union.

Finite Amplitude Development of Alternate Bars


Shoji Fukuoka

Head, River Hydraulics Division, River Department,


Public Works Research Institute, Ministry of Construction, Japan
Abstract
This paper presents a review of research concerning finite amplitude
alternate bars. An understanding of finite amplitude bars leads to a resolution
of the mechanics of river meanders.
Emphasis is placed on nonlinear
interaction between the flow, sediment discharge, and bed profile, and on
clarifying how alternate bars stabilize and equilibrate in shape.
Introduction
In a natural river, the bed profile and planform vary according to the
mutual relationship of the flow, sediment discharge and planform, Le.,
longitudinal and transverse profiles of the channel. It sometimes happens that
these relationships adjust themselves well and a state of equilibrium appears
The
such that there is no change in time in the shape of the channel.
processes leading to the equilibrium state can generally be divided into the
following two. The first is the process, such as seen in the laboratory, where
the channel gradually approaches a certain shape under given flow conditions.
The second is the process where the flow discharge changes with time, as in
rivers, and is divided according to the time scale of flow variation into two
subcases. In the first of these subcases, the time scale of discharge variation
is large, and the equilibrium shape corresponding to each flow stage is attained
at each time. The shape of the channel varies around the mean value because
of discharge variation.
In the second subcase, the time scale of discharge
variation is small and the shape of the channel cannot attain the equilibrium
shape corresponding to the flow at each time. The shape varies about the
average shape with a time lag relative to the flow variation. In the first case,
the channel shape at each time can be presumed to be its equilibrium shape.
In the second case, processes leading to the equilibriurn shape, namely, the
processes of development (or decay) dominate; the actual equilibrium is
computed from an average.
The flow in a river course is largely governed by the planform and bedform
of the channel.
In some instances, there appears a concentration of flow
toward a certain region along the river bank. The planform of the channel
associated with the concentration of flow along the river bank is the meander;
the corresponding bedform is the alternate bar. To carry out safe and proper
flood control measures, it is therefore necessary to determine the equilibrium
shapes of alternate bars and meanders, as well as the flow in channels with
such morphological features.
To this end, one must first understand the
occurrence and development (or decay) of alternate bars or meanders, and the
equilibrium shape of bars and meanders under conditions of steady flow. A

237
Copyright American Geophysical Union

Vol. 12

Water Resources Monograph

River Meandering

Vol. 12

Finite Amplitude Developlnent

238

considerable number of studies have been made on the occurrence and


development of alternate bars and meanders in rivers. Over the two years
from 1981 to 1982, the Committee on Hydraulics, Japan Society of Civil
Engineers compiled and reported upon the results of these studies in
the volume "Studies Concerning the Three-Dimensional Structures of Flood and
With this report as a starting
Channel Process" [Task Committee, 1982].
point, this article includes a review of research papers concerning finite
amplitude development of alternate bars, with particular emphasis on clarifying
how alternate bars equilibrate.

Alternate Bars in a Channel with Fixed Banks (Forced Meander)


The bedform and planform of a channel affect each other via the flow and
However,
sediment discharge in the channel, as indicated in Figure 1.
materials composing a natural channel normally differ between the river bed
and the bank regions; the time scales required for bed and bank changes also
differ. Specifically, when the banks are composed of cohesive materials, the
planform takes a relatively long time to reach equilibrium because of the great
resistance to erosion. In the bed area, meanwhile, the bed is liable to conform
to the flow because bed materials which can be deemed noncohesive
predominate; it is considered that an equilibrium bedform corresponding to the
channel's planform and discharge (discharge is assumed constant) is attained
It is, therefore, possible for laboratory
before the planform changes much.
tests and theoretical analyses concerning the processes of development of
alternate bars and their equilibrium shape to be conducted in a channel with
fixed banks. This is particularly true in large rivers in Japan, where rivers
are revetted so as to preclude channel shift.
(8)

(A)

CD

Bed form

Fig. 1.

Flow and
sediment
discharge

(C)

Plan form of
channel

@)

Interaction between flow and river morphology.

Flow and Sediment Discharge over Alternate Bars


Even if the alignment of a bank is rectilinear, the flow in a channel tends
to meander, as shown in Figure 2. Concentration and divergence of the flow
are repeated cyclically if alternate bars form on the bed.
In a flood,
revetment may be damaged at places where the flow is concentrated;
sometimes channel shift may result l Committee Report, 1984; Task Committee,
1982].
The points of flow attack thus become weak points of a levee.
Understanding the flow over alternate bars makes it possible to evaluate the
external forces working on the river banks, and is thus important in planning
for river improvement.
Hasegawa et al. [1983] have shown that the flow in a channel with
arbitrary periodic planform and alternate bar pattern can be determined with

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

239

Fukuoka

Fig. 2.

Meandering flow over alternate bars.

considerable accuracy by using plane two-dimensional analysis.


They have
obtained the following expressions for flow velocities u and v, and water level
by using Fourier series to express the downstream change of
variation
channel centerline curvature ro(s) as

e,

r~(s) = j~O {R~j

sin jkS

R~j

cos jkS}

(1)

and the bed profile as

E sin

i=O j=O

[i~N + ;

hie] (Asij sin jkS + Acij cos jkS)

(2)

The symbols are defined in Figure 3. Substituting these expressions into the
linearized shallow water equations, it is found that
U

Uo

[1 +.E .E sin
1=0 J=O

[i~N

+ ; hie] {aij sin jkS + bij cos jkS}] (3a)

Fig. 3.

Definition of symbols.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

240

Vol. 12

Finite Amplitude Development

[! - ; Oie]

v = Uo .!: .!: cos


1=0 J=O

{ = ho!:

!: sin

i=O j=o

[! + ;

{Cij sin jkS + dij cos jkS}

(3b)

0ie] {eij sin jkS + fij cos jkS}

(3c)

where

jk

2Cfo

4F
r

-2Cfo

jk

fo '

jk

Cfo

-Cfo

jk

(-l)ii1rl

jk

(-l)ii1rl
2F

jk

(-1 )i+1 i 1rl


2

?
F

bij

2F r

Cij

dij

jk

eij

jk

fij

2F r

Dio(-l) (i+1)/2 2Cfo


}2 1r2 lRcj

-C fo ,ASIJ

aij

(-1)i+1 i1rl

C ,Acij
fo

-efo'

Dio(-l) (i+1)/2 2Cfo


}2 1r2 lRsj

b'io(-1) ( i +1)/2 4
} 1r R c j

(4)

Dio(-l)( i+1)/2 4
} 1r R s j

- jk Asij
- jk Asij
Here, ro, Rsj and Rcj are the radii of curvature nondimensionalized by the
mean water depth, hOe Also, K = 21rh o/L, q = 2h o/B, S = s/ho, N = 2'fJ/B,
Die = {I - (-l)i+l }/2, Dio = {I - (-1)il/2, and, = 4/3.
Figure 4 shows the equilibrium bed profile in a strongly sinuous channel
with a geometric shape characterized by L = 432 cm, B = 22.0 cm, Rctho =
46.9 cm, and RC3ho = - 115 cm.
The figure also shows measured and
predicted values of the flow velocity vector. The theoretical values generally
explain the measured flow field well. In some regions, a tendency opposite to
the measured distribution of flow velocities is observed. Especially at cross
section 10, where the curvature of the channel is maximum, the two do not

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Fukuoka

241

ME-5
Unit:mm

Ca)

SO (cm/s)

(b)

y=4/3
SO (cm/s)

(e)
Fig. 4.

Bed geometry and flow velocity vector.

correspond well. This is probably because the secondary flow caused here by
the centrifugal force difference between the upper and lower layers of flow is
strong; secondary flow is ignored in the present shallow water analysis.
Furthermore, there can arise the problem that flow velocity responds
sensitively to small variations of river bed. Also, fluctuating flow velocity at
high wave numbers may be overestimated because of the use of linearized
equations. For these reasons, it is considered that, if the curvature of the
channel is not extremely large and the bed profile changes gently, flow in the
channel can be determined rather accurately by means of a linearized
two-dimensional plane flow analysis.
Fukuoka and Yamasaka [1985] analyzed flow over alternate bars, taking
even the nonlinearity of the flow into consideration. Their study was made
taking note of the fact that the nonlinearity of flow plays an important part
in stabilizing the wave height of alternate bars. If the nonlinearity of flow is

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Finite Amplitude Development

242

taken into consideration, the analysis becomes complicated, and it is difficult


to find a solution for an arbitrary geometrical shape such as that of Hasegawa
et ale [1983]. Thus, a straight channel is assumed for the planform, and the
following expression, which is a good representation of the shape of alternate
bars, is used.

Ho =
Since an equilibrium
by Hasegawa [1983],
water equations by
infinitesimal quantity
u
Uo

alsinN cos(kX - ) - a2cos2N

(5)

bed profile satisfies the relation 0(a2) ~ o(a;), as shown


the following solution is obtained by solving the shallow
the method of perturbations, using al as a primary
and a2 as a secondary infinitesimal quantity.

+ sinN{A 1llcOs(kX - ) + B ll1 sin(kX - )} + a2cos2N U2202

+ a~ [{U200 + A 202cos2(kX - + B202sin2(kX - >)}

~O

cos2N {U2201 + A 222cos2(kX - + B222sin2(kX - >)}]

al cosN{Cll1 cos(kX - )

+ Dll1sin(kX - )}

+ al sin2 N{V220 + C222 cos2(kX - ) + D222sin2(kX - )}


~

= alsinN{E111cos(kX -

(6)

(7)

+ F111sin(kX - >)}

+ a~sin2N[{600 + E202cos2(kX - + F 202sin2(kX - >)}

cos2N {6201 + E222cos2(kX - + F222sin2(kX - >)}]

(8)

Here, l = 1rho/B, Y = y/h o, and X = x/hoe According to the linear analysis,


the perturbation of flow velocity and water level are of the same mode as that
of the bed profile. In (6) to (8), however, higher modes appear in u, v, and
due to the effect of nonlinearity.
For example, Figure 5(a) shows a measured bed profile for alternate bars,
and Figure 5(b) compares the measured and computed distributions of flow
velocity near the tip of the bars, for the case where the bed profile is
reasonably approximated by (5). Figure 5(c) shows the computed flow velocity
vector field. The level of agreement between the distribution of flow velocities
in the transverse direction is somewhat low, because the bed profile is
approximated by the simple expression (5). The analysis, however, generally
agrees with observations in magnitude and direction, and adequately explains
the overall situation of concentration and divergence of flow over alternate
bars.
If the flow meanders and alternating flow concentration and divergence
occurs, the fluid force working on the bank is not spatially uniform, and
points of flow attack are formed where forces are locally large. If the bank is

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Fukuoka

243

t t 1, \ \, \ t
Experimental Condition

Fr=1.23
io =1/50

t t i ,\\, \ t
1tA \\\ \ \ \

1/\\\\\\ \

y\ \\\\\\,

21h o =2.92x10- 2

7l'~o =0.117

9.9m- ' '."

\ \ \ \ \

\\\\\\\tt
a1=0.99
a2=0.41

II \ ,

\1 \\1

~ \ \ \ \, tt t
II t It t\1 /I
tf f f t tt

t'

1111 II ~\II
t I ! II t tV t

(a)

8.0m -

1111111\1
tlllll! I~
tt !III II i
II til/II /
trtt~lll~

\11/1/ /II
II ~i t till
t.

tty t \ \ \ \ \

0.5

-Ef ~
Fig. 5.

-0.5

(b)

(c)

Computed values and measured values of flow velocity distribution.

composed of an erodible material, erosion occurs and the channel planform


changes.
However, the mechanism of bank erosion is still not thoroughly
understood, even though the flow field in the channel can be calculated.
Therefore, it is still difficult to predict where points of flow attack may occur
Fukuoka and Yamasaka [1984] investigated the
under these circumstances.
hydraulic indicators for bank erosion to be used in a straight channel with
alternate bars and erodible banks, assuming that bank erosion starts at a given
point of flow attack. The transverse slope of the bed, streamline deviation,
water level and flow velocity are appropriate hydraulic indicators of a point of
flow attack. Figure 6(a) shows where the maximum value of each indicator
appears according to (5) to (8). Figures 6(b) and 6(c) illustrate the nature of

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Finite Amplitude Development

244

F
~
r----

Maximum
Maximum
Maximum
Maximum

transverse slope
streamline deviation
water level
flow velocity

\1Y\1V

(a)

_5

Movable

\_Movable

(b) Fujita's experiment

V'

Movable

Flow

Fixed

(c) Ikeda's experiment


Fig. 6.

Correspondence between theoretical indicators and measured point of flow attack.

bank erosion in a straight channel with alternate bars. It cali be seen from
these figures that the point of flow attack can be theoretically determined as
the point of greatest streamline deviation. However, to determine the strength
of flow attack, it is necessary to consider the mechanism of bank erosion.
This problem must be left to the future.
According to the above two analytic methods, flow in a channel can be
determined if the equilibrium bed profile is given. However, if the bed profile
is unknown, it is impossible to use these methods to determine the flow. To
predict the variation of channel planform, therefore, it is necessary to compute
the equilibrium profile of a river bed under the constraint of a given channel
planform. The change of bed topography is directly related to the continuity
equation of sediment

Q!1 =
Of:

1
[OqBX
+ ~BY]
r=-x
OX
y

(9)

Generally, alternate bars in a straight channel do not cease downstream


migration; their equilibrium shape can be given by the following expression:
1/

= f1(x -

ct, y)

= f1(x,

y)

Copyright American Geophysical Union

(10)

Water Resources Monograph

River Meandering

Vol. 12

Fukuoka
The relationship between the equilibrium shape and the sediment discharge
distribution is thus
(11)
The sediment discharge distribution over alternate bars is difficult to measure,
and its temporal and spatial change is not yet thoroughly understood. Thus
the process leading to an equilibrium bed profile through temporal and spatial
chan~e of sediment discharge has received little attention.
Fukuoka et ale
[1983J and Uchijima et ale [1984, 1985] investigated temporally varying bars by
photographing and analyzing the movement of bed material with a video
camera. They explained the process of development leading to the equilibriulTI
of alternate bars from the relationship between the flow velocity vector and
the sediment discharge distribution. Figure 7(a) shows the bed profile, 7(b)
the near-bed flow velocity and sediment discharge vectors, and 7(c,d) the
distribution of near-bed flow velocity and sediment discharge in the
longitudinal and transverse directions 20 minutes after the start of flow, when
the bed profile had nearly reached equilibrium. The directions of the flow
velocity and sediment discharge vectors in the vicinity of the bed generally
agree except near the alternate bar front. In the vicinity of the front, the
sediment discharge vectors deviate from the flow velocity vector toward regions
of lower bed elevation in accordance with the gravitational force due to the
transverse gradient of the bed. Where the transverse gradient is almost zero,
the relationship between the sediment discharge in the transverse direction and
the flow velocity can be taken to be

(12)
Thus, if u changes spatially, a direct comparison of qBy and v is almost
meaningless. However, if u varies little, the transverse sediment discharge qBy
changes nearly in proportion to the transverse flow velocity v. According to
Figures 7(c) and 7(d), the spatial change of u is small on alternate bars. So,
the distribution of qBy generally agrees with the distribution of v except where
the transverse gradient of the bed is large. Figure 8 shows the flow velocity
distribution over developing bars 14 minutes after the start of flow. From the
comparison of this with the equilibrium flow velocity distribution (Figures
7(c,d)) one can see that there is no great difference in the distribution of
longitudinal flow velocity, but that there is a distinct difference in the
distribution of transverse flow velocity.
(Note that different representative
scales are used for the transverse and longitudinal flow velocities.) Specifically,
during the developing stage, transverse flow velocities are unidirectional within
a cross section, but at the equilibrium stage, their directions change along the
front line. On the whole, the flow is seen to be in the direction of the front
edge. This is probably because the flow direction was taken in the vicinity of
This change in the
the bed where it is influenced by secondary flow.
distribution of transverse flow velocity is closely related to the process leading
to equilibriurn.
The spatial distribution of bed load transport has seldom been measured in
rivers. Dietrich and Smith [1984] measured the distribution of longitudinal and
transverse bed load transport rates, the process of sediment transport and the
path difference according to grain size in a bend of a small meandering river

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Finite Amplitude Development

246
St.Ne.
1

___ Bed at time of flow


_
velocity measurement
(a)
(fixed bed)
St.Ne.
1
2
3
.:J -::=I __ :J

4
I"

.-- Near bed flow


velocity

234

4--=8 ---=9

(Vy)-

(Qbx) .....

6
(Qby)

10

10

t--l 1.0

11

12
,,:1,7

_-4

11

12

(Vx
, QbX)
Vo QbO

10

1--1

0.2 (Vy

(d)
Fig. 7.

12

t--t (1cm)

.-::1

(c)

11

Sediment discharge
(10 particles/em 's)

(b)

(Vx)--

10

Bed at time of
sediment discharge
measurement

_--:I

---::I

11

Vo

12
,

13

13

Qb Y )
Qbo

Distributions of near-bed velocities and sediment discharge over alternate bars in


equilibrium stage.

with an equilibrium bed profile (Muddy Creek, Wyoming). They studied how
these factors affect the bed profile and the planform of the channel. They
deduced from their observations that an equilibrium bed profile appears when
a net outward sediment discharge occurs in a region with maximal bed shear
stress.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

247

Fukuoka

St.No.

St.No.

(Vy)--

Fig. 8.

678

10

11

12

13

11

12

13

~ 1.0(~'~)
Vo Qbo

(Qbx)

(Vx) - -

(Qby)

10

(VV

0.2

y , Qby \

QbO)

Distributions of near-bed flow velocities and sediment discharge on alternate bars


at developing stage.

Analysis of Dominant Wave Length and Equilibrium Wave Height


of Alternate Bars
The analysis described above (see Flow and Sediment Discharge over
Alternate Bars) does not yet constitute a full calculation of equilibrium shape
of alternate bars because the bed profile, flow, and the sediment discharge still
remain to be quantitatively related. A quite complex analysis is necessary to
The method of dimensional analysis is
determine the equilibrium shape.
therefore often used to predict equilibrium wave height and length.
Ikeda [1984] obtained the following expression for wave height ZB;

ZB

no = 9.34

[B]Q

--{)A5

exp 2.53 erf

IOg10

0 . 594

1.2]

(13)

as indicated in Figure 9. Here B = channel width, ho = water depth, and d


= grain size; these are taken as the main parameters governing equilibrium
wave height ZB based on experimental data. In the region 6 < B/h o < 40,
the relation

ZB
no
=

1.51 Cf

[Bno]

1.45

(14)

holds approximately. Using the same parameters for wavelength L of alternate


bars, the following relation is found within the limits of Fr ~ 0.8

!: _
B -

5.3

[~] --{).45 ~
d

ho

as indicated in Figure 10.

Copyright American Geophysical Union

(16)

Water Resources Monograph

River Meandering

248

Vol. 12

Finite Amplitude Development

"V

o
A

Kuroki et al.
Fujita
Ikeda
Yoshino
Kinoshita

Eq.13

10 0

I o - _........_

........."""""-.........................._

.............

........_....Io.......-.~

10'

10

B/ho
Fig. 9.

o
r:,.
"V
D

o
A

Wave height for alternate bars.

Iguchi
Chang et al.
Kuroki et al.
Fujita
Ikeda
Yoshino
Kinoshita

Eq.15

10' .~_........_ .......................


10

"""-I................_ _.............- -........................

Fig. 10.

Wave length for alternate bars.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

249

Fukuoka

Jaeggi [1984], noting that it is through the action of deep scour rather than
wave height itself that the alternate bar affects channel processes, has
experimentally found the following relation

(16)
between the wave height ZB and the maximum depth of scour below the mean
river bed elevation SD. As to the wave height, an expression was obtained by
simply substituting the relation for relative wave height ZB/Zo and relative
roughness d/h o of a dune with corresponding values for alternate bars, ZB/B
and. d/B, but little verification is available for this expression.
Equilibrium heights and lengths derived from dimensional analysis are
important for practical use, but questions arise as to whether they incorporate
the dynamics. It is therefore necessary to study the equilibrium shape from
the viewpoint of dynamics. Fujita et ale [1982] first studied the equilibrium
wave height of alternate bars in the context of change in sediment discharge.
They assumed that the cross section of alternate bars is approximated by the
trapezoidal shape shown in Figure 11, and that sediment scoured in the shaded
part A is transported to the shaded part B and accumulated there.
Furthermore, on the basis of the following experimental observations: (1) the
flow velocity in the transverse direction is proportional to wave height ZB' (2)
the wave height Zk of the semi~ylindrical transverse shape obtained by
averaging over a wavelength of alternate bar is proportional to the wave
height ZB of the alternate bars, and (3) the transverse gradient of the average
river bed profile is proportional to the square of Zk, they deduced the
following expression for equilibrium wave height ZB at which the lateral
sediment transport rate became zero, on the average:

ZB
B

[0.0051] [B]
1-!!K

liO

2/3

[hQo] -

1/3

Ud

Fig. 11.

Trapezoid-approximated semicylindrical transverse shape.

Copyright American Geophysical Union

(17)

Water Resources Monograph

River Meandering

Vol. 12

Finite Amplitude Development

250

Here Ug and Ud are the downstream velocities of sediment and flowing water,
respectively.
Equation (17) was deduced on the basis of simple physical
considerations, and the assumption that nonlinear change in the transverse
gradient versus the change of wave height plays an important part in
determining the wave height of alternate bars. Fujita et ale 's [1982] expression
incorporates experimental results and some hypotheses, but is an interesting
achievement pointing to the essence of the phenomenon.
The mechanism by which the wavelength attains equilibrium was studied
by Fujita, !(oike, and Muramoto [1985], using a simple dynamic model. When
alternate bars first become manifest, the migration speed of upstream and
downstream bars differs, and the wavelength of the upstream bar increases
until the travel speed of both becomes equal. The following relation
1

qBx

c=r=-xZ-

(18)

generally relates the wave height, the travel speed of alternate bars and the
sediment discharge. So, using the subscript "0" for the upstream bar and the
subscript "1" for the downstream bar, the change of half-wavelength travel
speed can be described by the relation

(19)
The temporal change of ZBl and ZBO can be expressed as

(20a)
(20b)
using the analysis of Fujita et ale [1982].
They determined the temporal
change of wavelength by solving equations (20) and (19) simultaneously under
suitable initial conditions and attempted to determine the equilibrium wave
length as t approaches infinity. Since, however, (20) is a linear expression,
the wave heights ZBO and ZBl diverge with time. Thus, the state dfB1/dt =
o corresponds to vanishing migration rate, and the foregoing results for
equilibrium wave height taking nonlinear relationships into consideration are
incapable of analyzing equilibrium wavelength.
However, their attempt to
dynamically explain the mechanisms is useful in setting the direction of future
studies.
To stabilize the wave height of alternate bars, it is necessary to equilibrate
the tendency to increase it against the tendency to suppress it. Fukuoka and
Yamasaka [1985] assumed that the factors to develop and stabilize bars consist
of the nonlinearity of flow for the bed profile, and the nonlinearity of sediment
discharge for the flow. They performed a theoretical analysis of equilibrium
wave height that took these two nonlinear relations into account. As a result,
they shed new light on the problems of bar development and equilibrium shape
which had not previously been adequately handled by analytic methods. The
shape of alternate bars is taken to be given by (5) and the nonlinear solutions

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

2.51

Fukuoka

for flow velocity are given by (6) and (7). The' nonlinearity between the flow
velocity and sediment discharge is considered in the relations for sediment
discharge, incorporating the effects of longitudinal and transverse gradients of
the river bed:

{
-

T*co
T*r

. { [-TT**xr

1 - (sin 2 Ox

sin 2 Oy)

(21)

T *co.
T *co.
+- slnOx ] 2 + [T- *co + - slnOy] 2
Its
T *r
J.Ls

- ~::o J

1 - (sin 2 Ox

+ sin2 0y) }

(22)

Here, T* is the dimensionless tractive force for a level bed (i.e. in the absence
r

of bars) given by
T

*r

P Cro Uo _
hOI O
(Ps - p) gd - ~

and the longitudinal and transverse dimensionless


respectively expressed by
T*X

*y -

(23)
bed shear stresses are

p Cr
(Ps - p)gd

u2

p Cr
(Ps - p)gd

u2

+ v2

(24a)
(24b)

From the condition that the distribution of sediment discharge calculated from
(6), (7), (24), (21), and (22) must satisfy the continuity equation (9), the
following relations for growth rate of the amplitudes at and a2 in (5) are
obtained:

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Finite Amplitude Development

252

(25)
da 2

<IT

RIal

R 2a 2

R 3a l

R 4a 1a 2

(26)

Also, the phase <p in (5) satisfies

~=

Sl

+ s2a~ +

(27)

S3a2

In the above relations, time t has been made dimensionless as follows:

[f!. -

T2

*r

1]

gd3 .

(1 - A)h~

(28)

The coefficients Qt, Q2, ... , 83 are determined bX the coefficients in the
expression for nonlinear flow velocity of (6) and (7):
Aut, Btu, ... , C222.
Also, the static friction coefficient of sediment, J.Ls, the tractive force ratio,
T* IT* , depth width ratio l(=1I'"ho/B), and the dimensionless wave number
r
co

k( =211'" ho/L) play a role in determining the coefficients.

In (25) - (27),solutions for at and a2 satisfying both datjdT = 0 and da2/dT = 0 may exist,
depending on the hydraulic conditions.
After attaining this bed profile,
alternate bars do not change in shape, and migrate downstream with constant
speed. The wave height is then in equilibrium, and may be expressed by
(29)
\vhere at is the amplitude of the first mode of the bed profile, as defined in
(5). At equilibrium, the two amplitudes are given by, respectively,

(30)

Q1

Q 2d

= ----2
Q3

(31)

as found from (25) and (26). Here, W t = (Q2R4 - Q3R3), W2 = (QtR4 +


Q2R2 - Q3Rt) and W3 = QtR2. Figure 12 shows comparison between the
theoretical equilibrium amplitude a
(made dimensionless with water depth)
lT

determined from (29), with the use of measured wave number k, and measured
The theoretical wave height
dimensionless equilibrium amplitude ZB/2ho.
agrees well with observations. The mechanism that stabilizes wave height can
be explained from a detailed study of the meaning of the coefficients in (25)
and (26), Qt
R4, as follows. With the development of wave height, the
nonlinear relation between the bed profile and the flow becomes strong. The
tractive force in the center of the channel increases, and the tractive force
tv

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

253

Fukuoka

1.5

Za
2h o

../

~(<tJ~.

0.5

0/
/

/<:/

~:- 0

/.

Fig. 12.

.eI'
/

..
/ .....vo. .
o

1.0

Fukuoka and Yamasaka

Muramoto and Fujita

o
0.5

Fujita, Muramoto et al.

1.0

1.5

Theoretical values and measured values of equilibrium wave height.

near the banks decreases on the average. Thus, the nonlinear relation between
the flow and the sediment discharge in the bank regions necessary to develop
alternate bars gradually diminishes.
Consequently, the tendency for
development of wave height weakens (Q2 < 0) and equilibrium is approached.
The action for suppressing the development of wave height is mainly a
gravitational effect due to the transverse gradient of the river bed. (If the
first term, Qlal, of (25) is divided into the amplitude-nhancing and
suppressing terms, and expressed as Qlal = (Qle - Q1S)al, the suppressing
action is Q1Sal)' It increases almost linearly with wave height. Meanwhile,
the second term, Q2a~, in (25) acts to weaken amplitude growth Qleal + Q2a~;
it gradually weakens with the increase of wave height. Therefore, the wave
height attains equilibrium when the two tendencies are generally balanced
3
(Figure 13).
The nonlinear term Q2al indicates the suppressive effect of deposition and
scour along the bank?- and results froln a decrease in velocity near the bank.
This velocity decrease arises from the nonlinear relationship between the flo\v
and bed profile, and leads to a decrease in the lateral variation of sediment
discharge through nonlinear interaction between flow and sediment discharge.
In the foregoing analysis of equilibrium wave height, the wavenumber k is
considered known, and no information on wavelength is provided therein.
Fukuoka, Yamasaka, and Shimizu [1985] determined the equilibrium dominant
wavelength, and the conditions for bar formation, from a further extended
analysis of the foregoing nonlinear model. The wavenumber of an equilibrium
bed profile that dominates appears to be one such that the steepness is the
largest among the various wavenumbers composing the bed profile. Figure 14
shows how equilibrium steepnessalk (alternate long. and short dashed line) and

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Finite Amplitude Development

254

Main suppressing action

Q,sa,

"'C

c:
co
c: c:

.-0.2
...

O,ea, +02a,3

~~

/
/

g'.~

.0..

Main developing action

,/

./

Q)a.

/1

I
I
I

,/

(1)

,/

~g-

Cl cn

,/

,/

./

,/

'/

Fig. 13.

Mechanism stabilizing wave height.

initial wave steepness growth rate, f(l/al)d(alk)/dT]al = 0, namely, Qlk


The values of kmin and k max
(broken line) change with wavenumber.
correspond to the lower and upper limits on wavenumber at which Ql in (25)
is positive. Therefore, the range of possible wavenumbers is kmin < k < k max .
The wavenumber with the largest equilibrium wave steepness, k 01 (dominant
wave number according to the nonlinear analysis), and the wavenumber with
the largest initial growth rate of wave steepness, k0 2 (the dominant wave
number according to the linear analysis) nearly agree. It is therefore possible
to conclude that the equilibrium wave steepness becomes larger as the initial
wave steepness growth rate increases. Figure 15 shows the region in which
In that figure,
alternate bars form, along with the dominant wavenurnber.
1(=7r ho/B) is the ordinate and wavenumber k(=27r ho/L) is the abscissa. The
alternate long and short dashed line represents the dominant wavenumber
according to the nonlinear analysis, while the broken line represents the
dominant wavenumber as determined from the maximum growth rate of initial
wave steepness (i.e. the linear analysis). As indicated in Figure 14 as well,
the two dominant wavenumbers are nearly seen to agree, and both increase in
general proportion to I (i.e., the wave length is proportional to the river
width).

X10- 4

==0.1
io=1/50
2 P==0.25

10-2

k
Fig. 14.

Change of initial growth rates of wave-steepness and equilibrium wave-steepness.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

255

Fukuoka

Rmax
......................................................................
...

__

-.

10- 1

10- 2

10- 3

Fig. 15.

10- 2

10- 1

Occurrence region and dominant wave number.

In Figure 16, the theoretical dominant wave number, kT obtained by the


nonlinear analysis is compared with the measured wave number, kObo Figure
17 shows a comparison of the equilibrium half-wavelength, a 1T , determined

from (29) using a given theoretical dominant wavenumber, and the measured
half-wavelength, ZB/2ho. The theoretical wavenumber is seen to be somewhat
larger than the measured wavenumber. This is probably due not only to a
lack of analytic accuracy, but also reflects the fact that variation of
wavelength and flow velocity in the streamwise direction is not included in the
present theory [Fujita et al., 1985]. To this point, the analysis has been
limited to single-row alternate bars, but the analysis can be expanded to
double-row bars by merely replacing f used in the previous analysis with 2/.
Thus, the region of occurrence of double-row bars, and their dominant
wavenumber, can be obtained by scaling down by a factor of two in the f
direction the results for single-row bars shown in Figure 15. Consequently, if
1 < f max /2, it is possible for single-row and double-row bars to coexist.
Which of these prevails can be decided in terms of the magnitude of wave
steepness in analogy to the determination of dominant wavenumber. Thus it
is seen that, of the two different types of bars, those with multiple rows with
larger wave steepness prevail. Figure 18 shows how the division line between
single- and double-row bars thus determined, and a line delineating the limit
of formation of bars corresponding to I max , shown in Figure 15, relate to
measured data. In the drawing, the thin line is the division line of the case
where the growth rate of the initial wave steepness is used as the standard for

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Finite Arnplitude Development

256

kob

o
D

<D
D

o
10- 2
Fig. 16.

Fukuoka et al.
Hasegawa
Kinoshita
Ashida et al.
Muramoto et al.
Fujita et al.

kT

Theoretical dominant wave number and measured wave number

deciding the prevalence of either the single or multiple row configuration. It


can be seen that both nearly satisfy the regional division of measured data.
Theoretical studies in the past on the dominant wave number and the
formation of regional division are based on linear analyses. These, therefore,
used the condition of initial growth rate maximum for an infinitesimal
disturbance as a standard for deciding, the dominant wave number, and the
prevailing type of bar [Committee Report, 1984]. In (25), the limits a2 -+ 0
and at -+ 0 agree with those obtained from conventional linear analysis. The
regime diagram for alternate bar formation was prepared in accordance with
criteria similar to those used in the past, using the linear formulation

(32)
It is found to be in general agreement with the diagram of I(uroki and Kishi
[1982, 1984]. Some differences arise at large T*/ T*c, due to the fact that the
effect of transverse bed gradient on sediment discharge is treated differently
rFukuoka et al., 1985]. Colombini et ale r1987] obtained an analytical solution
tor equilibrium wave height of alternate bars by a perturbation method that
differs from that of Fukuoka and Yamasaka [1985].
At the critical
width-depth ratio f3c = Bc/h for the formation of alternate bars, Qt of (32)
becomes zero. For small values of t (when a is very small), they obtain

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

257

Fukuoka

Zs
2h o

/
/

I
~

Fig. 17.

0/

.~/

60

C:J. /

o0

& /0
~
0
0
o~
.~
oj LD 0

10-

6. .

o
(1)

Fukuoka et al.
Hasegawa
Kinoshita
Ashida et al.
Muramoto et al.
Fujita et al.

Comparison between equilibrium wave height calculated by giving theoretical


dominant wave number and measured wave height.

(33)
which is identical in form to (32). The method of the multiple scales is
introduced for values of ~(=B/hc) differing from {3cj
{3 = (3c(l

+ a)

k = kc

+ a

~=~+a~

k1

(34a,b)
(34c)

Here, a = ({3 - (3c) / {3c, k is the wav~number, and kc is the critical


wavenumber for the formation of alternate bars. Non--quilibrium terms which
bring about modification in the fundamental mode of the bed profile are
produced via third-order nonlinear interactions of the equations governing the
flow and sediment over alternate bars. Then, (33) modified to take account of
nonlinear terms, becomes

(35)
When a is sufficiently small ({3 = (3c), the first term of the left-hand side of

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

258

Finite Amplitude Developluent

_ T*r
T*co

Single row} ...


dou bl e row divIsion line
Non linear Linear

Formation limit line

::J

ro

10'

o
::J

'\\\

o <3

\',

:E

:=l

0-

0 0

..
~

_.

.
""0

r0(3
~

~cp

\ oCDQ~~
~
~ 'I~
<99\..~O~ ~\~
to DOg a\~ i
0

:E

....,-0

<p

l~"

,\

_.

ce..

::J
to

0- ro~
<3

~.

ro

'\\

to

::J

en

0-

~
~

(linear, non linear)


z 1/1000 ' \\ 1/100

0
())

-<D-

-0-

Fukuoka et al.
Hasegawa
Kinoshita
Ashida et al.
Muramoto et al.
Fujita et al.

t6. 0

10

?I '
//j
1;/
J

---' /

D~

-<Ir

10'

10

10 2

0.2
-B
10

ho

Fig. 18.

Bar formation regional division chart.

(35) is in balance with the first term of the right-hand side. Consequently, a
is of the order of (}'1/2.
Now, U, v, ~, and 11 are written in the form of a series expansion in (}1/2 :
u

U (}'1/2

+ u2 O' + u3 a;3/2 + ...

(36a)

o+

V 10'1/2

+ v2 O' +

V a;3/2

+ ...

(36b)

~= o +

~ 10'1/2

+ ~2O' +

~ a;3/2

+ ...

(36c)

~= o +

11

+ 112 0' +

11 aJ/2
3

+ ...

(36d)

Uo

v
Uo

0'1/2

By introducing (36) into the equations of motion for shallow water and the
continuity equations for flow and sediment, and solving the equations of each
order from zeroth to third order, a differential equation governing the complex
amplitude A(T 1) of the fundamental model is derived:
dA +
<IT1

f A
1

+ "( A A
2

= 0

The equilibrium amplitude of the fundamental mode is given as

Copyright American Geophysical Union

(37)

Water Resources Monograph

River Meandering

Vol. 12

259

Fukuoka

a=

(38)

(}'1/2

by setting dA/dT t = 0 in (37) and taking account of (36).


The analysis of Columbini et ale [19871 considers nonlinearity of the flow
and bed profile up to third order. The bed profile is expressed in terms of
Therefore, their analysis is more
higher harmonics than the form of (5).
comprehensive than that of Fukuoka and Yamasaka [1985]. However, it is not
clear from their analysis how essential the third-order solution is for the
stabilization_ of wave height, and by what mechanism the development of wave
height is suppressed toward equilibrium.
The above studies concern the equilibrium shape of alternate bars in
straight channels.
Regarding the equilibrium bed profile in meandering
channels, Hasegawa r1983] has carried out an analysis, taking secondary flow
into consideration. A plane two-dimensional flow analysis of (3) is performed,
while the transverse sediment discharge is determined by
qBn

qBso

[~o +

tan1jJ

(T*

~]

(39)

This relation includes the effects of secondary flow in the cross section and the
transverse bed slope.
In the above expression, represents the angle of
deviation of the near-bed flow caused by secondary flow. This is expressed as:

(40)

T * is .Jr. /(JjsJ-tkr*) and f Q!1 indicates the gravitational effect due to


o
0
an
transverse bed slope.
From the condition under which (39) satisfies the
continuity relation for sediment discharge in meandering channels, equilibrium
bed profiles where Cij = 0 are obtained as functions of N ij and Mij. Here,
Nij quantifies the intensity of secondary flow due to the planform of the
channel.
To this, the intensity of secondary flow in a curved channel is
applied mutatis mutandis. Mij, meanwhile, quantifies the secondary flow due
Since the undulation of a river bed
to the undulation of the river bed.
changes temporally and spatially, Mij is a quantity that normally cannot be
obtained without solving for the three-dimensional flow. The wave height Xij
cannot be deterlnined without a knowledge of the magnitude of Mij. Thus,
Hasegawa has contrived to back-ealculate Mij from experimentally obtained
equilibrium bed profiles. Ml1, the principal term of Mij, is found to be a
function of (B/h o)(B/d)-1/3. In the analysis of Hasegawa et ale [1983], the
suppressive action of secondary flow in the cross section is identified as the
mechanism stabilizing wave height.
This is an interesting idea, but their
assumption that Mij, the magnitude of secondary flow due to the unevenness
of the river bed, is constant regardless of the magnitude of wave height Xij,
and their notion of the equilibrium point of wave height leaves room for
theoretical improvement [Fukuoka and Yamasaka, 19841.
Numerical simulation is a powerful tool for predicting the flow and bed
topography in rivers where planform and longitudinal bed slope cha!1ge

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Finite Alnplitude Developn1ent

260

unit;mm
Time
200
400
600
800
1,000
1,200
1,400
1,600
1,800
2,000sec

I I I I I I

o 20406080100cm
Fig. 19.

Time variations of calculated bed configuration of meander angle

= 20

degrees.

variously.
Shimizu, Itakura, and Yamaguchi [1987] performed a numerical
simulation of bed topography of river channels using a two-dimensional model.
They solved numerically the equations of motion, the continuity equation, and
the sediment continuity equation for two-dimensional flow by the use of the
fini te-difference method.
Figures 19 and 20 show the computation results for the temporal change of
bed configurations in channels with different angles of meander.
The lines
------. f Iow

. unit; mm
Time
200
400
600
800
1,000
1,200
1,400
1,600
1!800
2,000sec
I I I I I
o 20406080100cm
I

Fig. 20.

Time variations of calculated bed configuration of meander angle

Copyright American Geophysical Union

= 30

degrees.

Water Resources Monograph

River Meandering

Vol. 12

261

Fukuoka

liB
7
6

Calculated

Kinoshita
(1974)

3
2
0
Fig. 21.

10

20

Comparison of critical meander angle


calculated values.

30

40

Be

between experimental,

theoretical,

and

show the position of scour holes traveling downstream with time. In the case
of angles of 30 (Figure 21), the scour holes are almost stationary, and
alternate bars stop migrating. These numerical results provide a reasonable
explanation for the results of [(inoshita's [1974] experiment and Hasegawa's
[1983] study, as shown in Figure 21.
Their results on the behavior of
alternate bars are very interesting from the viewpoint of river engineering and
river planning. They have applied this numerical simulation technique to the
prediction of flood flow and bed configuration therein, and have shown it to be
promising.

Conclusion
This article is restricted to the explanation of the stabilization mechanism
and equilibrium shape of alternate bars. An understanding of finite amplitude
alternate bars will facilitate the investigation of river meanders, because river
meander development is caused by the mutual interaction between the bedform
and planform of rivers. This is an important problem to be resolved in the
future.
NOTATION

at
a2

-TJt/ho
T/2/h o
channel width

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Finite i\lnplitude Devclopn1ent

262
C

Cf
Cfo

veloci ty of bar
friction factor
friction factor for a level bed

Fr

sand diameter
Froude number

ho
io
K

flow depth for a level bed


longitudinal bed slope for a level bed
to
proportionality
constant
relating

Ko
k

k max
kmin
ko
kT
L
I
IB
Mij
N

Nij
n

Q1,2,3
qBx
qBy
qBn
qBs

R 1,2,3,4

nonequilibrium

sediment transport
coefficient defined by (20)
wave number (= 21rh o/L)
maximum wave number
minimum wave number
dominant wave number
wave number calculated theoretically
hal f meander length
1rho/B
half bar length
secondary flow due to undulation of river bed
2T]/B
intensity of secondary flow due to the channel planform

normal direction to the main stream


coefficients of nonlinear flow velocity
longitudinal sediment transport rate
lateral sediment transport rate
sediment transport rate in - the direction normal to the
streamwise direction
streamwise sediment transport rate
coefficients of nonlinear flow velocity

Rsj

radii of curvature dimensionalized by the flow depth

Rcj
ro
S

radii of curvature dimensionalized by the flow depth

Sn
s
S1,2,3
T

rad i i of curvature dimensionalized by the flow depth


s/h o
maximum scour depth below the mean river bed
streamwise direction
coefficients of nonlinear flow velocity
dimensionless time defined by (28)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

263

Fukuoka
T*

u
Uo
Ug

v
X

Xij

relative shear stress


time
flow velocity in the streamwise direction
flow velocity for a level bed
streamwise velocity of sand
flow velocity in the lateral direction

x/h o
bar height

longitudinal Cartesian coordinate

y/h o
lateral Cartesian coordinate
equilibrium bar height

Y
ZB
ZK

semi-eylindrical transverse shape obtained by averaging


over a wavelength

(fJ-pc) / f3

f3

f3e(l+a)

f3e

cri tical width-depth ratio

'Y
Die

numerical value

{I

(-l)i+l }/2; when i: odd number, Die


0; when i:
=1
tan By / (d1]/dn)
bed level variation from the mean bed
first mode of bed variation
lateral bed profile averaged over one bar length
angle of meander at which alternate bars cease migration
longitudinal inclination angle of bed
lateral inclination angle of bed
phase angle
porosity of sand

jj

viscosity of water

even number, Die


t

1]
1]1
1]2

Be
Ox

By

J.ts

static fr i ction coefficient of sand

jjk

dynamic friction coefficient of sand


water level variation from mean flow depth
density of water

Ps
T

T *e

density of sand
dimensionles s shear stress
dimensionless critical shear stress

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Finite Amplitude Developnlent

264
T*co

dimensionless critical shear stress for a level bed

T*x

dimensionless shear stress for a level bed


dimensionless longitudinal shear stress

T *y

dimensionless lateral shear stress

T *r

'l/J

angle declination of the near-bed flow vector from the


streamwise direction due to secondary flow

References
Colombini, M., G. Seminara, and M. Tubino, Finite amplitude alternate bars,
Jour. Fluid Mech., 181, 1987.
Committee Report, Three-dimensional structures of flood and channel process,
Proceedings of Japan Society Civil Engrs, No. 345//1-1 (in Japanese), 1984.
Committee Report, Task Committee on threEHlimensional structures of flood
and channel process, Committee on Hydraulics and Hydraulic Engineering,
Japan Society Civil Engrs. (in Japanese), 1982.
Dietrich, W. E. and J. D. Smith, Bedload transport in a river meander, Water

Resour. Res., 20(10), 1984.

Fujita, Y., Y. Muramoto, S. Horiike, and T. Koike, On the mechanism of


alternating bar development, Proceedings of the 26th Japanese Conference
on Hydraulics (in Japanese), 1982.
Fujita, Y., T. Koike, and Y. Muramoto, The mechanism of wave length
development of alternate bars, Proceedings of the 29th Japanese Conference
on Hydraulics (in Japanese), 1985.
Fukuoka, S. and M. Yamasaka, Equilibrium height of alternate bars based on
non-linear relationships among bed profile, flow and sediment discharge,
Proceedings of the Japan Society Civil Engrs., No. 357/11-3 (in Japanese),
1985.
Fukuoka, S. and M. Yamasaka, Equilibrium wave height and flow on alternate
bar based on nonlinear relationships of bed form, flow and sediment
discharge, 21st IAHR Congress, Melbourne, Australia, 1985.
Fukuoka, S. and M. Yamasaka, Prediction of a point of flow attack in the
straight channel with alternate bars, Proceedings of the 39th Annual
Conference of Civil Engineers II (in Japanese), 1984.
Fukuoka, S., M. Yamasaka, and Y. Shimizu, Dominant wavelength and
formation conditions of alternate bars based on nonlinear analysis,
Proceedings of Japan Society Civil Engrs., No. 363/11-4 (in Japanese),
1985.
Fukuoka, S., K. Uchijima, M. Yamasaka, and H. Hayakawa, Distribution of
sediment transport rate over alternating bars, Proceedings of the 27th
Japanese Conference on Hydraulics (in Japanese), 1983.
Fukuoka, S. and M. Yamasaka, Discussion to a study on flows and bed
topographies in meandering channels, Proceedings of the Japan Society of
Civil Engrs., No. 351/11-2 (in Japanese), 1984.
Hasegawa, K., A study on flows and bed topographies in meandering channels,
Proceedings of the Japan Society Civil Engrs., No. 338 (in Japanese), 1983.
Hasegawa, K., I. Yamaoka, Y. Watanabe, and S. Sasajima, Theory and
experiment of the flow over bars in a winding channel, Proceedings of
Hokkaido Branch, Japan Society Civil Engrs., No. 39 (in Japanese), 1983.
Ikeda, S., Prediction of alternate bar wavelength and height, Jour. Hydraulic

Engrg., Am. Soc. Giv. Engrg., 110(4), 1984.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

26.5

Fukuoka

Jaeggi, M. N. R., Formation and effects on alternate bars, Jour. Hydraulic

Engrg., Am. Soc. Civ. Engrg., 110(2), 1984.

Kinoshita, R., Investigation of channel deformation in Ishikari River, Rpt.


Bureau of Resour., Agency of Science and Tech., Japan (in Japanese), 1974.
Kinoshita, R. and H. Miwa, Channel topography for stabilizing the position of
alternate bars, Shin Sabo, 94 (in Japanese), 1974.
Kuroki, M. and T. Kishi, Regime criteria on bars and braids in alluvial
straight channels, Proceedings, Japan Soc. Civil Engrs., No. 342 (in
Japanese), 1984.
.
Kuroki, M. and T. Kishi, Study on regime criteria of river morphology,
Proceedings, 26th Japanese conference on Hydraulics (in Japanese), 1982.
Shimizu, Y., T. Itakura, and H. Yamaguchi, Numerical simulation of bed
topography of river channels using two-dimensional model, Proceedings of
31st Japanese Conference on Hydraulics (in Japanese), 1987.
Uchijima, K. and H. Hayakawa, Geometrical shape and sediment discharge of
developing alternate bars, Proceedings of the 39th Annual Conference of
Civil Egineers, II (in Japanese), 1984.
Uchijima, K. and H. Hayakawa, Near-bed velocity distribution and sediment
discharge distribution of developing alternate bars, Proceedings of the 40th
Annual Conference of Civil Engineers, II (in Japanese), 1985.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Copyright 1989 by the American Geophysical Union.

Alternate Bars and Meandering:


Free, Forced and Mixed Interactions
G. Seminara and M. Tubino
Istituto di Idraulica, UniversitA di Genova

Foreword
This paper attempts to give a systematic treatment of the linear and
weakly nonlinear aspects of the development of flow and bottom perturbations
of lengthscale of the order .. of channel width in. developing meanders.
The
content of the paper summarizes some of the works published in recent years
by the Genoa group on sediment transport. Some as yet unpublished results
are also presented, part of which have been the subject of preliminary
contributions to Euromech 215 (S. Margherita Ligure (Genoa), Sept. 1987).
Most of the above material has been presented by the senior author (G. S.)
as an invited contribution at the Joint U.S.-Japan Final Meeting on River
Support for participation at the
Meandering (Kauai Island, October 1987).
meeting was provided by the direction of the U.S.-Japan joint project which is
gratefully acknowledged.
We are also deeply grateful to Mrs. Tina D'Agostino for carefully typing
our manuscript.

Introduction
The origin of fluvial meandering is still an intellectual challenge in spite of
a longstanding interest of the scientific community in this subject [Guglielmini,
1697; [<elvin, 1876]. However, the interacting efforts of geomorphologists and
hydraulic engineers in the last two decades have strongly increased the level of
understanding, moving from a state of descriptive empirical knowledge to the
gradual buildup of rational models.
The starting point of this relatively rapid. development may be traced back
to the cornerstone contributions by Kinoshita [1961] and Leopold and Wolman
[1957]. The cause for the development of the alternating sequence of pools
and riffles experimentally observed by the above authors in both straight and
meandering reaches of rivers was then the subject of several theoretical
The basic idea, apparently first
investigations developed in the seventies.
conceived by Callander [1968], was the recognition that, under appropriate
conditions, the flat cohesionless bottom of a turbulent stream flowing in a
straight channel loses stability to a perturbed configuration characterized by
3-D infinitesimal growing and migrating perturbations of lengthscale of the
order of the channel width. In particular the bottom perturbations tend to
form an alternating sequence of deep and shallow reaches, a pattern which was
interpreted as the precursor of meanders. The above idea was then set as the
267
Copyright American Geophysical Union

Vol. 12

Water Resources Monograph

268

River Meandering

Vol. 12

Alternate Bars and Meandering

basis of several increasingly refined 2-D and 3-D linear stability theories
[Hansen, 1967; Hayashi, 1970; Sukegawa, 1971; Engelund and Skovgaard, 1973;
Parker, 1976; and Freds0e, 1978]. These contributions aimed at improving the
understanding of the physical mechanism underlying the instability process and
predicting the marginal stability conditions, Le., the conditions for bar
formation in the space of flow and sediment parameters, the wavelength and
the wavespeed of perturbations selected, Le., those corresponding to the
maximum growth rate. The linear 'bar theory' will be outlined below (see
Linear Development: Bar Theory).
In all the above contributions, the sinuous migrating thalweg produced
within the straight banks of the channel as a result of the 'alternate bar'
mode of instability was taken as implying incipient meandering.
In other
words, although no attempt was made to analyze the occurrence of bank
erosion as a result of the establishment of an alternating sequence of flow
accelerations and decelerations close to the banks, it was tacitly assumed that
a sinuous migrating thalweg within a straight channel should somehow evolve
into a sinuous channel thus leading to meandering. The latter implication is
by no means obvious and will be questioned in the following.
Parallel to the above deterministic approach an alternative research line
.was developed based on the idea that meandering could be explained as a
stochastic process [Langbein and Leopold, 1964]. The stochastic approach will
not be discussed in the following. The interested reader is referred to the
original paper by Langbein and Leopold [1964] and to Callander's [1978] review
for a critical discussion.
The state of the art at the end of the seventies is well described in
Callander [1978, p. 155]: "what causes meanders is still a question waiting a
complete answer, although the case for dynamic instability is strong".
In the early eighties, a different point of view was taken in an important
paper by Ikeda, Parker and Sawai [1981] developing ideas originally put
forward by Ikeda, Hino and Kikkawa [19761. Meander formation was associated
with a planimetric instability, the destabilizing mechanism being bank er9sion
associated with secondary flows -induced by channel sinuosity. In other words,
the conditions were sought for the straight planimetric configuration of a
channel with flat cohesionless bottom to lose stability to a perturbed
configuration of the channel axis. The conclusion of this work (bend theory)
was that for small channel sinuosities the wavelengths selected by the 'bend'
mechanism are close to those predicted by 'bar' theories. This would vaguely
support the idea that the initial instability leading to the formation of
alternate bars might proceed into a planimetric instability and thus to the
development of a meandering channel with an initial wavelength close to. that
of alternate bars.
The new perspective introduced by the latter contribution has stimulated in
recent years further investigations aimed at clarifying the distinct role played
by the 'bar' and 'bend' mechanisms in the process of meander formation.
/(itanidis and /(ennedy [19841 emphasize the role played by secondary flows in
the initiation and early development of meanders both in alluvial and
rock-incised channels but conclude that it is not presently possible to ascertain
"through validation of existing theories with available data which mechanism
(or mechanis~s) is responsible for the initiation and development of meanders".
The first attempt to understand the relationship between the 'bar' and
'bend' mechanisms within the context of a unified approach is due to
B?ondeaux and Seminara [1983, 1985]. .(The latter paper will be referred to in
By examining the structure of the dispersion
the following as BS.)
relationship of the bar theory relating the complex growth rate of 'bar'
perturbations to their wavelengths, it turned out that, for a wide range of
values of the width ratio of the channel, there exists a class of 'bar'

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

269

Seminara and Tubino

perturbations characterized by nearly vanishing growth rate.


These
quasi-nonamplifying and quasi-nonmigrating 'bar' perturbations were found to
be forced by the development of channel sinuosity, thus leading to a
quasi-resonance phenomenon. The wavelengths selected by the bend-resonance
mechanism were found to be about three times as large as those predicted by
traditional 'bar' stability theories.
The above analysis, which will be presented below with some developments
and corrections (see Forced Bars in Curved Channels), reconciled the theory
with the physical idea that the type of perturbations responsible for
meandering initiation must be steady in order to determine such a slow
process as that of bank erosion. On the contrary, migrating alternate bars
exhibit a relatively large migration speed as compared with the speed
associated with bank erodibility.
The latter observation was also
independently made by Olesen [1983], who suggested that steady spatially
growing 'bar' perturbations are a more reasonable candidate to explain
meander formation. However, Olesen [1983] did not explain what mechanism
would excite the development of such perturbations which are not 'naturally'
The results of BS have
growing, their temporal growth rate being zero.
indicated that the forcing effect of curvature can provide the excitation
mechanism for steady spatially periodic perturbations.
If the above picture is correct, one then expects to see two different
phenomena occurring in the process of meander formation in alluvial channels:
the first is the development of migrating alternate bars (which will be called
'free' bars in the following), a relatively fast process compared with the second
slower process, the development of channel sinuosity (Le. of 'forced' bars) with
a wavelength 2-3 times as large as that of the initial migrating 'free' bars.
This sequence was indeed observed by Lewin [1976] on a gravel bed river, the
development of which, from an artificially straightened configuration to a
meandering pattern, was followed for a period of about one year. The channel
was active only occasionally during high stage flows; the observations showed
that a meandering pattern developed with a wavelength which "...came to
exceed twice the spacing of initial bars...These features suggest a loosening of
the initial dimensional control of bar spacing... " [Lewin, 1976, p. 284]. Lewin's
field observations confirm that the bend mechanism follows the initial
development of migrating alternate bars and acts in the sense predicted by BS.
The resonance phenomenon discovered by BS has been recently confirmed
by Johannesson and Parker [see paper in this volume]. However its actual
relevance to the process of meander formation requires the investigation of
further effects which have so far been either neglected or oversimplified. Let
us restrict ourselves to the understanding of a still fairly idealized model of
meander formation:
a laboratory experiment performed under steady flow
conditions, with no transport in suspension and uniform grain size distribution.
The model proposed by BS would then be appropriate to describe the above
conditions except for the fact that it does not allow for the coexistence of
migrating alternate 'free' bars 'naturally' developing with steady point bars
'forced' by channel curvature. Such a coexistence is not only indicated by
Lewin's field observations but has also been detected in the interesting
laboratory investigations by Gottlieb [1976] and j(inoshita and Miwa [1974J
(hereafter referred to as KM) (see also Dietrich and Whiting's paper in the
present volume).
KM investigated the behavior~ of alternate bars in a meandering channel
formed "from straight segments at .an angle a to each other" such that the
resulting wavelength was either equal to or a fraction of that of alternate
'free' bars which had been previously found to form in a straight channel with
identical flow and sediment. characteristics. Two distinct flow regimes were
experimentally detected depending on a falling above or below a threshold

Copyright American Geophysical Union

Water Resources Monograph

270

River Meandering

Vol. 12

Alternate Bars and Meandering

value ll'c which was experimentally determined with an accuracy of about 1.5
For ll' < ll'c, the train of 'free' bars continues to migrate even after reaching
an apparently 'naturally stable' state where 'free' bars are perfectly in phase
with steady 'forced' bars. For ll' > ll'c, alternate 'free' bars cease migration,
this state being clearly detected from the disappearance of any bed oscillation
The* value of ll'c was found to vary
in time at any given cross section.
strongly with channel meander wavelength L m
KM's results suggest the idea that it is the interaction between migrating
alternate ('free') bars and steady ('forced') bars which is responsible for the
suppression of the former perturbations. Furthermore, for suppression to occur,
the amplitude of 'forced' point bars, which increases with ll', must exceed a
threshold value dependent on the meander wavelength. This points out the
need for a theoretical interpretation of the process able to provide also a
predictive tool for engineering purposes. In order to set up such a theory, a
finite amplitude representation of both migrating 'free' bars and steady 'forced'
bars is preliminarily required. Weakly nonlinear theories suitable to this aim
are presented below (see Free Bars in Straight Channels, and Interactions
Between Free and Forced Bars). A theoretical approach able to explain how
the interaction between steady and migrating perturbations prevents the
persistence of the latter is also outlined (see Suppression of Free Bars in
Meandering Channels).
On the other hand, one may take the dual viewpoint and wonder whether
the above interaction can give rise to a steady effect able to affect bend
instability, Le. to modify the amplitude and phase shift that steady 'forced'
bars exhibit in the absence of coexisting 'free' bars.
A theory suitable to
successfully attack this problem is outlined below (see Interactions Between
Free and Forced Bars).
There are still various aspects of the natural phenomenon which are ignored
in the above models.
Some of them, namely flow and bar unsteadiness,
transport in suspension, and grain sorting in bends, have been given some
attention in the recent literature [Tubino and Seminara, 1987; Seminara and
Tubino, 1985; Ikeda and Nishimura, 1985; Parker and Andrews, 1985; Ikeda et
al., 1987]. Some of them are briefly discussed in the following, along with
those features of the physical process which have so far been either completely
ignored or roughly modeled like the mechanism of bank erosion, side wall
effects, and entrance effects.
Finally, "Perspectives" (see below) is devoted to some concluding remarks.
0

Free Bars in Straight Channels

Introductory Remarks
In the following, we will call 'free' or 'free alternate' bars those bars which
develop spontaneously in channels as a result of an instability process. The
word 'free' emphasizes the spontaneous character of the latter phenomenon as
opposed to the development of point bars in curved channels which is
controlled by the forcing effect of curvature.
A sketch of a sequence of developed free bars is given in Figure 1. Their
main features are fairly steep consecutive diagonal bar fronts, deep pools at
the downstream face of each front along the channel banks followed by
relatively gentler riffles along the upstream faces of the fronts, bar heights of
the order of the average flow depth, bar lengths of the order of a few channel
widths, and migration speed much smaller than the average flow speed.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Seminara and Tubino

271

1 - - - - - - - - - - - 21(;/

Fig. 1.

AN

-------------l

Sketch of a sequence of developed free bars.

In gravel bed rivers free bars form at flood stage and undergo minor
variations at lower stage. In sand bed streams bars coexist with dunes and
significant variations of the bar structure may occur as the flow stage varies.
Various questions arise: i) Can we predict under what conditions free bars
form? ii) How are wavelength and migration speed of free bars related to
flow and sediment characteristics?
iii)
Can one explain the longitudinal
asymmetry embodied in their fronts? iv) Does the amplitude of the bar grow
under steady flow conditions until it reaches a finite 'equilibrium amplitude'?
v)V\That is the significance of such a concept under unsteady conditions?
Questions i) and ii) are also relevant to the problem of meander formation:
they will be addressed below (see Linear Development: Bar Theory) where
linear bar theory will be briefly outlined.
Questions iii) and iv) will be
tackled by means of a nonlinear theory which will be presented (see Free
Interactions and Amplitude Equation):
this will allow us to predict the
maximum scour associated with the development of free bars in straight
channels. The practical implications of the latter results are fairly obvious as
regards highly developed densely populated countries where channelization and
artificial straightening of rivers, motivated by land reclamation needs, often
create the conditions for the appearance of free alternate bars with consequent
risks of bank and bridge failures. Finally, question v) will be discussed in the
section on Unsteady Alternate Bars: it will appear that the growth rate of
bars varies during the flood and may even become negative close to the peak.
this implies that the concept of 'equilibrium amplitude' must be revised under
unsteady conditions.

Formulation of the Problem


Preliminaries
We consider flow in a straight alluvial channel with nonerodible banks and
constant width 2B*. Since we wish to investigate finite amplitude effects on
the development of free alternate bars, we need a careful description of the
geometrical characteristics of the flow.
Let us then refer to an orthogonal reference frame (s*,n*,z*) such that s* is
a longitudinal direction, n* is a transverse horizontal direction, z* is orthogonal
to the (s*,n*) plane and the corresponding metric coefficients are (h *s ,l,l)
(Figure 2). Let the bed surface be defined by the following relationship
vt *

z* - 1]*(s*,n*)

(1)

The plane locally tangent to the erodible boundary intersects the reference
plane (s*,z*) along a straight line. We denote by
the unit vector of the

!1

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

272

s*

tangent plane

Fig. 2.

Sketch of reference frame (vp*

= particle

velocity, u *

-b

= average

bottom stress).

latter while
denotes the unit vector of the upward direction normal to the
tangent plane~ Finally, let 7-2 be the unit vector of the Cartesian axis lying
on the tangent plane and orthogonal to 7-1 and v. It is readily shown that
(~,
reads:
-

!l,!2)

"

1/

[- ----i 7J*s*

- 7I7n* ; 1 ]

11 +
~

s -' - - - - - - - - - -h=

(1]* *)2
,n

--i 1]* *]

[1 ; 0 ; h
s

[--i
1]* *ph
J
s

,8

(2a,b)

,8

7J*s* 71\*
[- --i
hs "

; 1+

[--i
7J*s*]
hs '

2 ;

TJ*,n*]

(2c)

In the present chapter, the adopted coordinate system will be Cartesian so


that h *s = 1. The above relationships will be employed in order to derive
the Cartesian components of bedload after the direction of bedload transport
along the tangent plane will have been established.

Governing equations
Let us assume the channel width (2B*) to be large enough for side wall effects
to be negligible in the central region of the flow. Furthermore, let us recall

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Seminara and Tubino

273

that alternate bars contribute little to flow resistance at least for active gravel
beds [Shen, 1962; Jaeggi, 1984; Parker and Peterson; 1980; Bray, 1979]. This
implies that flow separation is not likely to be a crucial feature of free bars
development.
This premise then encourages one to assume that an
approximate representation of the flow field by means of a depth averaged
model which is obviously unable to predict separation might still be suitable
to model the gross features of flow structure and bed topography. The St.
Venant equations of quasi-steady shallow water flow in a straight channel with
slowly varying erodible bottom may be written in the form
VU
VV

,0

,0

UU

,8

= - H

+UV ,8 =-H ,0

(VD) ,0 + (UD) ,8

~
D

(3)

_ !lIJl

(4)

,8

(5)

(6)
where (U,V) are depth-averaged axial and radial velocity components, Ts and
Tn are bottom stresses, H is water surface elevation, D is local depth, Qs and
Qn are sediment flow rate components, Fo is the unperturbed Froude number
and t is dimensionless time.
Furthermore, Qo is the ratio between the scale of sediment discharge and
the flow rate and {J is width ratio defined as

Qo

il~

ds

- l]gd:t:
* *

{J

(1 - P )DoUo

B
= ..

(7a,b)

Do

where Ps and d*s are density and diameter of the sediment modelled as
uniform, p is water density, g is gravitational acceleration and p denotes
sediment porosity; finally Uo* and Do* are averaged speed and depth for the
uniform unperturbed flow.
.
The variables have been made dimensionless in the form:
* * *
* 2
(H ,D ,TJ) = Do(FoH,D,TJ)

*
(U *,V *) = Uo(U,V)

* *

(s*,n*) = B *(s,n)

(Ts, Tn)

*2
pUO (Ts, Tn)

(& ]

* * = d *{
(Qs,Qn)
(Qs,Qn)
s p - 1 gd *}1/2
s

t*=

*
.!!....- t
Uo*

(8a,b)
(8c,d)
(8e,f)

Boundary conditions
We ignore the side wall boundary layers and require the channel walls to
be impermeable both to the flow and to the sediment. Thus
V

Qn

(n

1)

Copyright American Geophysical Union

(9a,b)

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

274

Bottom stresses
Having ignored separation, we model the flow structure as slowly varying
both in space and in time. This suggests that the bottom stress, modelled as
aligned with the depth averaged velocity vector, can be expressed in terms of
a local friction coefficient C defined by the relationship
2 1/2

! = (rs,rn) = (U,V)(U + V)
2

(10)

In the following, the local bed configuration will be assumed to be plane so


that the following logarithmic formula will be employed for the friction
coefficient:
- 1/2

6.

/AJ

(11)

2.5v'12.5ds

where the roughness parameter has been put equal to 2.5d *s after Engelund and
* o*, has been
Hansen [1967] and a nondimensional sediment diameter, ds = ds/D
introduced.
We point out explicitly that this procedure is equivalent to assuming that
the turbulent structure is in equilibrium with the local conditions, its spatial
and velocity scales being the local values of depth and friction velocity,
respectively. This model is likely to be approximately adequate anywhere but
in the separation zone.

Sediment transport
For the sake of simplicity, we assume sediment to be transported mainly as
bedload. An extension to the case where a significant fraction of sediment is
transported in suspension is conceptually straightforward.
Sediment transport is assumed to be determined by local flow conditions,
its direction deviating from the direction of average bottom stress under the
action of gravity. In nondimensional form we write (see Figure 2)
Q
-

(Q T 1,QT2,Q II )

(coso,sino,O)if>

(12)

where if> is the equilibrium sediment load function.


For relatively small values of 0 the experiments of Ikeda [1982a] (see also
Parker's [1984] discussion) suggest the following formula:
sino

sinX - _r_ (F~H - D)

f3/D

(13)
,D

!1

to be appropriate, where X is the angle between the bottom stress and the
direction, 0 is Shields parameter and r is a coefficient (presumably dependent
on the particle Reynolds number) which various authors suggest to take as a
constant ranging between .3 [Olesen, 1983] and .6 [Engelund, 1981].
We wish to stress the empirical and approximate character of (13) although
successful attempts to derive an expression of the type (13) have been
proposed in the literature [I(ikkawa et al., 1976; Engelund, 1981; Parker and
Andrews, 1985] all of them based on somewhat 'averaged' models of sediment
grain dynamics along curved paths. Furthermore, (13) can only be appropriate

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

275

Seminara and Tubino

within a linear context. The extension of (13) to the weakly nonlinear case
would involve complications which do not seem justified at this stage given
the uncertainty which is still present in the estimate of the coefficient r. The
latter has often been based on measurements of transverse bed slope in fully
developed flow in constant curvature rectangular channels with erodible bed
and on estimates of sinX based on theoretical models which are linear and
apply to fairly wide channels. The latter conditions are only approximately
satisfied in the experiments.
The Meyer Peter Muller formula in the form given by Chien [1954] will be
employed to evaluate the equilibrium sediment load function . Thus:

= 8( 0 -

Ocr)3/2

Ocr

(14a,b)

.047

Finally, the longitudinal and transverse components ofQ are derived from QTl
and Q using (2). We find:
T2

1- ~

Qs = QTl [

[iff] - 11,sJ
~ [iff]
Q T2

Qn = QT2[l -

(15)

(16)

having neglected fourth order terms in the products between the longitudinal
and transverse slopes.

Linear Development: Bar Theory


We now attempt to answer questions i) and ii) posed above.
This is
possible within the context of a linear theory which investigates the conditions
required for the unperturbed uniform flow to lose stability to flow and bottom
perturbations small enough for linearization to be a valid approximation
(strictly infinitesimal). Details of the theory are given in BS and Colombini,
Seminara and Tubino [1986, 1987J (hereafter, the latter paper is referred to as
CST). In the following, we out ine the main analytical and physical features
of bar theory.
Let us examine disturbed flows of the form:
(U,D,H,V)

(1,I,Ho,O)

A(U1,D1,H1,V.)

(17)
(18)

with A infinitesimal.
In (18), Co and o denote friction coefficient and
bedload function of the undisturbed uniform flow.
On substituting from (17) and (18) into the differential system (3-6),
performing the linearization and using the relationships (10-16) to express Ts1,
T n1, QS1 and Q n1 in terms of (U 1, V1,H1,D 1) a linear partial differential problem
is obtained for the latter unknown functions.
We now restrict our attention to a class of linear perturbations which
admits of a Fourier representation both in the longitudinal and in the
transverse direction (normal mode approach) and write:
(U1,D1,H1,V.)

exp(nt)(Sm(n)f1,Sm(n)d1,Sm(n)h1,Cm(n)g.)E1(s,t)

c.c.

(modd)

Copyright American Geophysical Union

(19a)

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

276

(m even)

+ c.c.

(19b)

Here c.c. (or an overbar) denotes the complex conjugate of a complex number
and we define
Sm(n) = sin( mnn/2)

Cm(n) = cos( 1rIIln/2)

(20a,b)

Em = exp[mi(As - wt)]

(20c)

with A, w, and n real quantities which denote wavenumber, angular frequency,


and growth rate of the perturbations, respectively, while m defines the number
of braids which perturbations tend to form (m=1 for the alternate bar mode).
The linear homogeneous algebraic system obtained by substituting from (19)
into the governing differential problem requires a solvability condition which
defines a dispersion relationship between flow and sediment parameters, which
takes the form of equation (50) of BS, and can be written in the general form
f(r!,w,A,,B ; O,d s )

=0

(21)

For given values of f) and ds (21) allows one to define 'neutral' conditions by
requiring that the amplification factor r! of the bar perturbation should vanish.
In the plane (A,,B) this condition determines a neutral curve which may exhibit
a minimum at A = Ae and ,B = ,Be. A typical neutral curve is plotted in
Figure 3 for the case of alternate bars (m = 1).
In order to give a simple physical interpretation of bar instability, it is
convenient to assume the dimensionless amplitude of bottom elevation 1/1 =

25 _ _ . _ - - - -

-------r----,,...------,

STABLE

-------I..-------I

5 L . . - - - - -.........
0.0

Fig. 3.

Ac

A typical neutral curve for the case of alternate bars (m

Copyright American Geophysical Union

1.5

= 1, (J = .3,

ds

= .01).

Water Resources Monograph

River Meandering

Vol. 12

277

Seminara and Tubino

(F~ht- d t) to be real (which is equivalent to choose appropriately the origin


of the longitudinal axis). We then write:
(b) (b) (b) (b) _
(ft,gt,ht,d.,qs ,qn ,t s ,tn ) -

(22)

exp[-i( 8t,82,83,84,85,86,~,88)]
having defined
(Qs.,Qnt, Tst, Tnt) = exp(Ot )(Smq~b),Cmq~b),Smt~b),Cmt~b))Et(s,t)

(m odd)

C.C.

(23)

with a similar expression for m even. Thus each 8 gives the phase lag of
bottom profile with respect to each of the above quantities. The values of the
above lags can be obtained, for given f), d s, and {j, by solving the linear
algebraic system obtained by substituting from (22) into the governing
differential system with 111 assumed to be real.
Figure 4, which shows the
results of our calculations for the alternate bar mode (m=l), is readily
interpreted:
the peak of longitudinal shear stress, longitudinal velocity and
longitudinal component of sediment transport is located along the rising part of
the bottom profile; the peaks of transverse velocity and transverse shear stress
are located around the crossings, Le. about a quarter of a wavelength ahead of
the peaks of bottom profile; flow depth is in opposition with respect to bottom
profile.

2.0

b
1C

05=0 7
01
03

1.5

04
06

1.0

02=0 8

.5

0.0 + - - - - - - + - - - - - - - t - - - - - - + - - - - - - +
0.0

Fig. 4.

.5

1.0

The phase lags of various flow and sediment properties with respect to bed profile
in the linear bar theory are plotted versus the wavenumber (/3
10, 0
.1, d s
.01, m
1).

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

278

Finally, the phase lag 06 of transverse sediment transport depends on the


relative importance of two contributions to Qnl. The former is associated with
gravity, is proportional to transverse bed slope and negative; whence, in the
absence of further contributions, transverse sediment transport would be in
opposition with respect to bottom elevation (06 = 1r). The second contribution
to Qn1 is proportional to the transverse shear stress associated with secondary
flow; a glance at Figure 4 shows that for this component. of Qnb b6 just
exceeds 1r/2. The vectorial composition of the above two contributions to Qn1
leads to values of 06 falling in the range (( 1r/2)';'1r), as shown in Figure 4.
Using (22-23), sediment continuity is readily shown to lead to the following
relationship for the growth rate

!L = _ Alq~b)1 sines + mn~q~b)1 cos


Qo

1] 1

06

(24)

1] 1

Thus bar instability (i.e. the sign of n) depends not only on the phase lag 05
of longitudinal sediment transport with respect to bed profile (as in the case
of two-dimensional mesoforms) but also on the phase lag 06 of transverse
sediment transport with respect to bed profile. The above discussion shows
that, when sediment transport mainly occurs as bedload, its longitudinal
component is dominated by fluid friction and is always destabilizing. On the
contrary, both contributions to transverse sediment transport are stabilizing, in
particular that associated with gravity which is proportional to m 2 .
The
latter effect inhibits the development of higher order modes and its balance
with the destabilizing effect previously discussed determines the number of
braids (m) selected by the instability process.
It may be worthwhile to point out that, if the effect of gravity is ignored,
destabilizing effects prevail on stabilizing effects for any value of m and higher
order modes are increasingly unstable [Engelund and Skovgaard, 1973J.
The predicted values of /3e as a function of () for typical va ues of the
roughness parameter ds and assuming the perturbed bed to be plane are
plotted in Figure 5. How accurate the theoretical predictions for /3e and Ae
are, as compared with experimental observations, can be inferred from Figures
3 and 4 of CST.

Free Interactions and Amplitude Equation


Within the context of a linear theory, the amplitude A of the perturbation
cannot be determined being an arbitrary infinitesimal factor.
We now
formulate a nonlinear theory which allows us to predict the development of
perturbations within the unstable region of Figure 3, and thus to answer
questions iii) and iv) posed above. It will appear that finite amplitude effects
inhibit the exponential growth predicted by linear theory leading to an
asymptotic
'equilibrium anlplitude'
in
agreement
with
experimental
observations.
The first idea underlying the present nonlinear analysis (whose details are
given in CST) is the recognition that the nonlinear development of the
perturbations is 'slowly varying' in time. In fact, the growth rate n predicted
by linear theory is O(Qo) (see (6)) and thus is quite small. More precisely,
within the unstable region of the plane (A,{J) of Figure 3, close enough to the
neutral curve, n is proportional to Qo[(,8-/3e)//3e], which is a small parameter
even for values of /3 much larger than Pe.
In the following we assume:

(25)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

279

Seminara and Tubino


14

~c

ds

12

.001

10

.005
.010
8

.020
.030

:8~8
6

.05

Fig. 5.

.20

.35

The critical value of the width-ratio Pc as a function of


values of d s . The unperturbed bed is assumed to be plane.

.50
(J

is plotted for given

and take advantage of the slowly varying character of the process by


employing a multiple scale technique [see, for instance, Nayfeh, 1973]. Let T,
then, denote a 'slow' time variable associated with the growth of perturbations
such that
T

(t

(26a,b)

and let us assume that the amplitude A of the fastest growing linear
perturbation is indeed a function of T to be determined.
The second main feature of the theory is that, having restricted the
analysis to a neighborhood of the neutral configuration, we can assume
nonlinearity to be 'weak'.
This implies that the generation of higher
harmonics from self-interactions of the fastest growing linear perturbation with
itself is such that the order of magnitude of harmonics of increasing order
decreases with some power of the small parameter (. The order of magnitude
eX of the amplitude A of the fundamental is then determined by requiring that
nonlinearity affects the physical balance controlling the growth of the
fundamental perturbation. In the neighborhood of the critical conditions, the
above balance is neutral at leading order O( (x) where destabilizing effects
associated with friction equal the stabilizing effect of gravity.
The linear neutral balance is modified at the order at which the
fundamental is reproduced by self interactions.
A simple thought on the
cascade process which controls the generation of higher harmonics suggests that
this occurs at 3rd order where the growth rate of the fundamental,
proportional to ([d(A(x)/dT], must be balanced by terms linear in the
amplitude function A, which are of the form O(A(x), and by nonlinear cubic

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and 11eandering

280

terms associated with nonlinear interactions and proportional to (A2 Ac 3x ). In


the absence of the latter contribution, balance would be accomplished for
arbitrary x and A ()( exp(OT/ c) as expected, in the unstable region, for linear
perturbations. Provided x = 1/2, the effect of the cubic terms alters the
above balance and leads to a nonlinear amplitude equation for A(T).
We then take x = 1/2 and expand the solution in the form:
(U,D,H,V)

(l,l,Ho,O) +

E (Up,Dp,Hp,Vp)(cl/2) + O(c 4 / 2 )

(27)

p=l

(Co,O,<Po,O)

+ E

p=l

(Tsp,Tn p,QsP,Qnp)(t 1 / 2 )

+ O(c 4 / 2 ) (28)

O(C 1 / 2 )
On substituting from (27) and (28) into the differential system (3-6) and
equating likewise powers of c, at O( c1 / 2) we obtain the linear differential
problem with f3 replaced by f3c. This system admits of a solution of the form
(Ut,Dt,Ht,Vl)
(29)
As already mentioned, substitution of (29) into (3-6) leads to a homogeneous
algebraic linear system for (ft,dt,ht,gl) whose solvability condition reduces to
the dispersion relationship (21). The complex amplitude function A(T) is left
undetermined at this order.

O( c)
The next order problem is obtained by substituting from (27-28) into (3--6)
and equating terms O( c).
Analyzing the form of the nonhomogeneous terms arising from self
interactions of the leading order perturbation, it follows that the O( c)
perturbation consists of four contributions: a harmonic of order 2 both in the
longitudinal and transverse direction, a harmonic of order 2 in the longitudinal
direction and of order 0 (Le. constant) in the transverse direction, a harmonic
of order 2 in the transverse direction and of order 0 (Le. constant) in the
longitudinal direction, a distortion of the basic uniform flow.
Thus, the following structure of the perturbation at O( c) can be assumed:
(U2,D2,H2)

{A E2[C2(n)(f22 ,d22,h22) + (fo2,do2,h o2)] + c.c.}


+ AA{ c2(n)(f2o ,d2o,h 2o ) + (foo,doo,hoo)} + (O,O,hoo)

V2

{A E2(S2(n)g22 + g02) + c.c.} + AA(S2(n)g2o +goo)

(30a)
(30b)

Notice that a correction hoo,s (due to the nondimensionalization employed) is

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

281

Seminara and Tubino

required for the basic water surface slope having expressed the width ratio in
the form of (25).
On substituting from (30a,b) into the O( l) differential problem, after some
manipulations, one ends up with linear nonhomogeneous algebraic systems for
the coefficients of the expansion (30).
The above systems are to be supplemented by appropriate boundary and
integral conditions.
i J The requirement of V2 vanishing at the walls implies:

(31)

gOO=g02=0
ii)
iii)

iv)

The condition of vanishing sediment flow rate at the walls is


automatically satisfied by the present solution.
The condition that flow discharge per unit width should not be altered
by the development of perturbations leads at O( t) to the following
relationships:
2foo + 2doo + ftdt + ftdt = 0

(32)

2fo2 + 2do2 +ftdt = 0

(33)

Finally, the condition of constant average reach slope requires at O( t)


that
2

Fohoo - doD = 0

(34)

o( t 3 / 2 )
The 3rd order problem is obtained by substituting from (27-28) into (3--6)
and equating terms O( t 3 / 2 ).
Again, by analyzing the form of the forcing terms, it turns out that self
interactions reproduce the fundamental perturbation and give rise to third
harmonics both in the transverse and longitudinal direction so that we can
write:

+ higher harmonics
V3 = {E t Ct(n)g3t(T) + c.c.}

higher harmonics

(35a)
. (35b)

Amplitude equation
The nonhomogeneous algebraic system obtained for (f3t, g3b h 3t, d3t) is
such that its homogeneous part is identical to the linear algebraic system
governing the linear perturbation (ft, gt, hI, d t ). Since the latter admits of a
nontrivial solution, a solvability condition is required at third order which can
be shown (see CST) to lead to the following nonlinear ordinary differential
equation for the amplitude function A(T):

(36)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

282

where 0'1 and 0'2 are complex coefficients expressed in terms of the O( (1/2) and
O( f) solutions. Physically, the condition (36) tells us how the linear neutral
balance between destabilizing and stabilizing effects occurring at the critical
conditions (Ac,{3c) is modified by the growth of perturbations within the
unstable region.
In particular, (36), being of the Landau-Stuart type, exhibits the following
important features:
If the cubic term is neglected, one recovers the usual exponential
behavior of A(T) predicted by linear theory.
Nonlinear effects inhibit growth and lead to an equilibrium amplitude
Ae reached as T -1 00, provided the real parts of 0'1 and 0'2 have
different signs. In fact, (36) is readily solved in the following closed
form:

IAI

-Re ( 0'1 )
Re(a2)-Re(a1) CA exp[-2Re(a1)T]

with CA arbitrary constant.


Re( at} > 0, I A I

ex

(37)

Thus, in the unstable region where

exp[Re( (1)T] as T

-1

00

(as expected from linear

theory) while IA I -1 v-Re(at)/Re(a2) as T -1 00, which is meaningful


provided Re( 0'2) < o.
Finally, the correction Lv of the angular frequency of alternate bars at
equilibrium is defined by the relationship
Ae =

IA I

exp(iwT)

(38)

and is readily found to be expressed in the form:

(39)
It may be appropriate at this stage to mention that an independent attempt
to tackle the present nonlinear problem was proposed by Fukuoka and
Yamasaka [1985]. The perturbation procedure employed by the latter authors
was based on decoupling the flow equations from the bottom equation,
assuming a structure for the bed profile where the contributions due to the
2nd order components 22 and 02 were ignored and the amplitude a2 of second
order perturbations was assumed to be independent of the amplitude a1 of the
fundamental, solving for the flow perturbations where all second-order terms
were now retained and deriving two independent amplitude equations for a1
and a2. The weakly non-linear character of their expansion is evident since a2
2
was assumed to be O(al), but it was not linked with the condition that {3
must fall within a neighborhood of the neutral conditions nor was the slowly
varying character of bottom development recognized by the latter authors. In
view of CST's analysis, it appears that a1 and a2 are indeed dependent;
furthermore, ignoring the contributions associated with the 2nd order
components 22 and 02 prevents sediment continuity to be satisfied.

Nonlinear Characteristics of Steady Alternate Bars


The solution of the differential problems at the various orders of
approximation and some tedious algebra allow us to calculate the coefficients
0'1 anda2 of the amplitude equation (36) for any given set of parameters {3, 0,

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

283

Seminara and Tubino


5,....-----~----r__---___,__--r__-r__---___.

Jaegg i

(1983)

PVC

Jaeggi (1983) sand


o SUkegawa (1971)

Kinoshita (1961)
0 Muramoto. Fu j ita

<>
<>

(1978)

Ashida. Shiomi (1966)


6 Ikeda (1982)
A Fuj ita. Muramoto (1985

6
6

<>
E

.0

6\
0

::r:

o '.
o

1.---

--l

I.

-----l-_

2 3 4

(Hbm)exp
Fig. 6.

Comparison between the maximum bar height predicted by (40) and experimental
results of various authors.

and ds. The real parts of at and' 0'2 were always found to have different
signs. The maximum height of alternate bars at equilibrium H
defined like
BM

in Ikeda [1982b] as the difference between the maximum and minimum bed

* can then be computed from the


elevations within a bar unit (scaled by Do)
following relationships
HBM = max( TJ) - min( TJ)

TJ

02

cos( 2AcS

b02 )]

2-

+ Fohoo} + O(A e {3/2)

(40a)

(40b)

where TJ = 1 and TJ , TJ , TJ , ~2, and b0 2 are functions of 0 and ds, which


122
20
02
are reported by Tubino and Seminara [1987].
The agreement of formula (40) with experimental results by I(inoshita
[1961], Ashida and Shiomi [1966], Chang et al. [1971], Sukegawa [1971],
Muramoto and Fujita [1978], Ikeda [1982b], Jaeggi [1983], and Fujita and
Muramoto [1985] is explored in Figure 6. We refer the reader to CST (where

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

284

a simplified form of (40) was derived and the latter data were not included)
for details of the above comparison; here we simply state that the agreement
is found to be quite good except for data corresponding to values of (J close to
the critical value for sediment motion (Jcr. The latter uncertainty is justified.
Indeed, the theoretical value of {Jc falls quite rapidly close to () = (Jcr so that
a relatively small error on (Jcr leads to a relatively large error on {Jc and thus
on H .
BM
We point out that (40) establishes a correlation between the height H
of
BM
bars and the width ratio {J with (J and d s parameters. The role of {J and ds
as relevant variables for the bar height had already been pointed out by Ikeda
[1982b] on dimensional and empirical grounds.
A strong linear correlation between the maximum relative scour 7J and the
M
maximum bar height HBM is found when our theoretical predictions are
applied to the experiments mentioned above. We find a value of .57 of the
average ratio 17M/HBM, very close to the experimental value (.5) reported by
Ikeda [1982b].
1. a

cr

Q)

,0

:c

.5

""

,'*

...........

Linearized

,0

:c

Actual

Fujita & Muramoto

(1985)

0.0
0.0

.5

1.0

T/(T)eq
Fig. 7.

Comparison between the time development of the maximum bar height as


predicted by (37) and that experimentally observed by Fujita and Muramoto
[1985].

The work by Fujita and Muramoto [1985] allows an interesting comparison


between the time development of bar amplitude predicted by the present
theory (37) and that experimentally observed. The agreement is found to be
satisfactory, as demonstrated in Figure 7.
Notice that the value of the
constant CAin (37) was chosen such that the value of the amplitude at T =
o would fit the first experimental point [run H.3 of Fujita and Muramoto,
1985]. Also notice that the present theory, linearizing the growth rate in a
neighborhood of the neutral conditions, tends to overestimate the rate of bar
amplification. The agreement between theoretical predictions and experiments
improves if the actual value of n rather than its linearization is employed in
the calculations (dotted line in Figure 7). It is of interest to point out that
the presence of an inflection point in the theoretical curve for A(T) is
experimentally confirmed: this feature is typical of the nonlinear development

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

285

Seminara and Tubino

.03

ds

------_..J .040
.050

----------J

.02

.030
.020

-----------J .010

-------_.J
-----_-.J

.01

o . 00

'---_~_~_.........L.___--'_

.20

.05

Fig. 8.

____'__ _. L . . . . . . . . _ _ . . . . I . . . . __

.35

__L..__

.005
.001

___I

.50

The O( f) correction of the normalized Chezy coefficient CIC o averaged within a


bar unit is plotted versus (J for given values of ds.

of perturbations in hydrodynamic instability and is presumably also present in


the case of dunes, although it has not been recognized in theoretical attempts
to describe the unsteady development of dunes [Freds0e, 1979; Tsujimoto and

Nakagawa, 1984].

Figure 8 shows the O( i) correction of the normalized Chezy coefficient C/ Co


averaged within a bar unit. It appears that flow perturbations give rise to
quite small frictional effects. However, notice that expansion losses are not
accounted for in the theoretical predictions.
Finally, a prominent feature brought up by nonlinear effects is the
formation of diagonal fronts and the increased steepness of the bot tom
downstream of the fronts as demonstrated in Figure 9 of CST. Furthermore,
a bell-shaped cross section of the type experimentally observed by Fujita and
Muramoto [1985, Figure 1.6] is produced by the effect of second harmonics.

Unsteady Alternate Bars


We finally tackle question v) posed earlier, summarizing some preliminary
results of a recent work [Tubino and Seminara, 1987]. We investigate how the
instability process which causes bar formation and the nonlinear phenomenon
which controls the development of finite amplitude perturbations are affected
by unsteadiness of the basic flow, which is always characteristic of rivers. On
physical ground, one may expect that a strong influence of unsteadiness is
likely to be displayed if the time scale for bar growth is comparable with the
characteristic period of the basic unsteadiness.
Moreover, under the latter
conditions, the concept of 'finite equilibrium amplitude' established in the
previous section can only keep its significance provided it be referred to some
'reference flow stage' which will have to be defined. Moreover, the question

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

1\1 ternate Bars and Meandering

286
-1

ds
-2

:8~8

.030
.020

.010

~ -3

.005

t3
0
M

bO -4

.001

-5

-6

" - - - - " " " " - - - _ . . . . . . - _ - - - - ' - - _ . - - - - I . . . _ - - - - - L_ _. . I . . . . - - _ . . . . . I - _ - - - ' - - _ - - - - I

.05

Fig. 9.

.20

.35

.50

The growth rate of bar amplitude in the slow time scale is plotted versus the
Shields stress 0 for given values of the grain ratio d s .

arises of whether bars develop as a result of a single flood event or as a


cumulative process.
In order to answer the above questions, let us model the way a flood wave
is 'felt' by a relatively short reach of the channel with length of a few channel
widths.
One can readily show that, within the above spatial scale at the

* where u*
lowest order of approximation in the small parameter u(=u*B*/Uo)
denotes a characteristic angular frequency of the basic unsteadiness, a flood
wave is 'felt' as a 2-D locally uniform channel flow varying in time such that
Do*

-*
= UoUO(T)

Do*

-*
= DoDo(T)

(41a,b)

-* and Do,
-* respectively, characteristic speed and depth of the basic flow
with Uo
and T 'slow time variable' defined as ute
During the rising stage Do* increases so that the instantaneous width ratio {3
and the nondimensional grain size ds decrease while Shields parameter ()
increases. In Figure 5 the evolution in time of the basic flow is represented
by a family of hyperbolae, namely {3 = (/38)/ (), /3 and 8 being the reference
Depending on the particular
values of width ratio and Shields parameter.
struct ure of the basic flow, the path in Figure 5 may fall entirely, partly or
may not fall within unstable regions. Gravel bed rivers which become active
only at relatively high flows will have only a relatively short portion of their
trajectory in the plane ({3, ()) such that {3 > {3c( (),d s) and () > ()cr.
We now analyze the growth of bar perturbations with respect to a basic
Using the same
locally uniform unsteady flow in the the form of (41).
notations and scalings as before and assuming that a 2-D approach is still
suitable to investigate alternate bar development (an assumption which is

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

287

Seminara and Tubino

likely to fail at low stage where the effects of flow separation are presumably
more important), the problem can be formulated as before (see Formulation of
the Problem) provided {3 is now taken as a function of time.
We now restrict our attention to the weakly nonlinear regime such that the
average width ratio /3 falls into a neighborhood of Pc.
Furthermore, we
assume unsteadiness to be weak and write

/3 =

flc(l

+ ()

fl

= /J[1 -

(42a,b)

tID 01 ( T)]

with 6 and ( small parameters and DOl ( T) the unsteady part of the basic flow
depth.
The above restriction makes the problem amenable to analytical
treatment. Notice that the steady results of the previous section suggest that
the analysis is likely to hold for values of ( sufficiently large to be of practical
significance.
The growth of perturbations is described by the following
expansion:

+ ([A

(T)(d22C2
(3/2

d02)E2

[d 31 (T)SlEl

c.c.]

c.c.

6(1/2

IA(T) 12 (d20C2

doo)]

[d31D01( T)A(T)SIEl

c.c.]

(43)

higher harmonics

where A(T) is the amplitude function to be determined.


Similar expansions may be set up for the further unknowns U, V and H.
The structure of (43) reveals that the interaction of the fundamental with the
unsteady part of the basic flow reproduces the fundamental itself at O( 6(1/2).
If 6 = ~(, with ~ an 0(1) constant, the solvability condition required for the
nonhomogeneous differential system at O( (3/2) is directly affected by the latter
interaction and the following 'amplitude equation' is derived

dA
<IT

A[O'I

~O'OD01(T)]

.2

+ AlAI

(44)

0'2

where 0'1 and 0'2 are identical with the complex constants of the 'steady'
amplitude equation (36) (recovered from (44) for vanishing ~) and 0'0 is a
further complex constant dependent on the O( (1/2 ,f) perturbation solution.
Equation (44) shows that, within the present 'weakly unsteady' context, the
effect of unsteadiness is felt at the linear level, Le., it affects the instantaneous
growth rate and phase of the perturbations.
In order to let the actual parameter controlling the influence of
unsteadiness clearly emerge, it is convenient to rewrite (44) in the form:
d IA I = ~
---a-r
U

{[I + ~

where r denotes 'real part'.


parameter is

O'Or
0'1 r

D01( T)] I A I

O'2r
0'1 r

A1

(45)

Equation (45) shows clearly that the relevant

(46)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

288

2.0 r - - - - - - - - - - - , , . - - - - - - - - r - - - - - - r - - - - - - - ,

STABLE

1.5

UNSTABLE

1.0

.5

NO SEDIMENT MOTION
0.0

--J'---

L....-

----L

----I

----L..

.5

0.0

Fig. lOa. Flow hydrograph reconstructed from


discharge per unit width,

Lewin

1.0

[1976]

'6 = .1, ds = .025, P = 12.8,

(qo
(f

dimensionless flow

= 2xlO- 4).

D may

be easily estimated from Figure 5 and from Figure 9, which gives air
as a function of 0 for some values of ds. Indeed: i) if iT ( 1, bars develop
on a time scale much faster than that associated with the basic unsteadiness
so that bars reach their steady 'equilibrium amplitude' corresponding to
instantaneous flow characteristics; ii) if iT ) 1, bar development is much slower
than the unsteady component of the basic flow, which is thus unable to affect
0(1), as often occurs in nature, flow unsteadiness does
the former; iii) if iT
affect bar development.
Figures (10a,b) show an application of the present theory to a series of
floods which occurred on an artificially straightened reach of Ystwyth River
and are reconstructed from Lewin [1976]. It appears that bar development is
essentially a cumulative process arising from various events, each exhibiting a
phase lag between basic flow and bar amplitude.
N

Future Developments
Various features of bar development in straight channels still require further
elucidation.
The role played by suspended load has been ignored in our analysis. The
linear contribution of FredS0e [1978] suggests that the presence of suspended
load affects the morphological regime, Le. the bar stability criterion, when the
Shields stress attains values larger than (.3-.4). The above influence appears
to be important as, in the absence of suspension, the bed was found to be
stable for 0 ~ 1, whereas instability persisted in the upper regime when
suspension was accounted for. A further problem which may also be related
to the effect of suspended load is the coexistence of dunes and bars. There is
no doubt that the latter is observed in nature. A simple way to account for

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Selninara and Tubino


1.0

289

r-------r------r------~--_..,__,

IA I

//

/~/1~

\
\
\

~~~~~~

~~~~~

.5

\
\

"

\
\
\
\
\

"I"

I
- - - - I-~. ~,---L-.

----='

0.0

11//

I
L..-....-_1

j"

j"

...--:::::'

/
.L-...-';""'/

-.L-

-.L-

.5

0.0

........

1 .0

Fig. lOb. Time development of bar amplitude for the hydrograph of Fig. lOa:
actual amplitude, AT
instantaneous response, A
linear response.
L

IA I

the presence of dunes is to ignore its interaction with bars and simply employ
Engelund and Hansen '8 [1967J approach to estimate the total load. However,
dunes are likely to interact with secondary flow induced by bars so as to
affect the dynamics of transverse sediment transport in a manner which has
never been thoroughly analyzed.
This points out the need for further
experimental and theoretical work.
Finally, strongly nonlinear effects may be involved both in the case of
steady and unsteady bar development when values of f3 are attained which are
much larger than f3c:
The weakly nonlinear expansions can be readily generalized so as to
account for the effect of higher harmonics and a numerical procedure can be
employed to determine the amplitude of each harmonic as a function of time.
The feasibility of this approach is currently being investigated.
Finally, the role of grain sorting has been neglected and will require some
attention in the future.

FORCED BARS IN CURVED CHANNELS


Introductory Remarks
The subject we investigate in this chapter concerns the prediction of the
pattern of bottom topography induced by secondary currents forced by
curvature in weakly meandering channels. The above investigation will be set
For the moment, we will
as the basis of a theory of meander formation.
ignore the possibility for free bars to coexist with forced bars.
It has long been recognized that in uniformly curved channels streamline
curvature leads to an unbalance between transverse pressure gradient,
practically uniform along the vertical, and centrifugal forces which decrease
from the free surface to the bottom. This leads to a secondary circulation,
directed outwards at the surface and inwards close to the bottom,

Copyright American Geophysical Union

Water Resources Monograph

290

River Meandering

Vol. 12

Alternate Bars and l\tleandering

characterized by transverse shear stresses able to compensate for the above


dynamic unbalance [Rozovskij, 1957].
If the channel bed is cohesionless, the action of bottom stresses associated
with the secondary flow makes sediment particles deviate towards the inner
bank, thus establishing a transverse slope of the bottom which reaches an
equilibrium condition when the transverse bottom stress acting on the particle
is balanced by the downslope component of gravity. As a result of the latter
mechanism, the bottom deepens along the concave bank and becomes shallower
alon~ the inner bank [Van Bendegom, 1963; Engelund, 1974; I(ikkawa et al.,
1976 .
I channel curvature varies in the longitudinal direction as in meandering
streams, the effect of longitudinal convection of transverse momentum gives
rise to some delay of secondary flow and bed topography with respect to
curvature [Gottlieb, 1976; Seminara and Tubino, 1985 (hereafter referred to as
ST)].
A similar effect arises in the entrance region of uniformly curved
channels.
The interaction of secondary flow with the main longitudinal flow is more
subtle than it may appear. Within the framework of a linear theory, which
accounts for curvature effects linearly and ignores the influence of side walls,
the effect of secondary flow is felt by the main flow only indirectly. More
precisely in uniformly curved channels:
curvature of the channel axis implies higher bed slope in the inner
bend than in the outer bend: this effect is independent of secondary
flow and would lead to higher downstream velocity along the inner
bank, as found by Rozovskij [1957] and Engelund [1974];
deepening of the outer flow arising from superelevation of the free
surface and from scouring induced by secondary flow enhances the
vertical gradient of longitudinal stress thus accelerating the outer flow.
A longitudinal variation of curvature of the channel axis again leads to
some delay associated with longitudinal convection of downstream momentum
and with longitudinal variation of the free surface slope. A small effect of
vertical convection of longitudinal momentum is also present. All the above
effects are accounted for in Gottlieb r1976] and in ST.
The direct effect of secondary ilow on the transverse distribution of the
main flow occurs through transverse convection of longitudinal momentum as
pointed out by I(alkwijk and De Vriend [1980]. However, in order for the
latter mechanism to operate, a transverse variation of longitudinal velocity is
required to interact with the secondary flow. In a wide straight rectangular
channel, far from the side walls, the transverse distribution of longitudinal
velocity is flat. Thus no transverse transport of longitudinal momentum may
be induced by secondary flow at a linear level; the first contribution arises at
O( v 2 ) if v is a dimensionless measure of curvature effects to be precisely
defined. On the other hand, if the width ratio of the channel is not large
enough, the effect of the side walls is felt in the central region of the channel
leading to some transverse variation of the downstream velocity and to some
momentum redistribution.
An attempt to quantify the latter effect has
recently been proposed by Johannesson and Parker [see paper in this volume].
The 3-D and and 2-D models of flow and bed topography in mildly curved
wide channels developed herein are employed in order to study meander
formation as a planimetric instability of the type first investigated by Ikeda et
ale [1981]. This will lead us to detect the resonance phenomenon mentioned
earlier (see the Introduction to this paper) which will arise both from the 2-D
and the 3-D models. The original results of BS corrected from an algebraic
error will also be extended so as to account for some effects which had been
previously neglected.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Seminara and Tubino

291

Formulation of the Problem


Geometrical preliminaries
Let us adopt a curvilinear coordinate system such that the longitudinal
coordinate follows the direction of the main flow. We define the channel axis
in parametric form:
*
*
*
P(s*):(Xa(s*),Ya(s*),Za(s*))
with s* coordinate denoting the arclength (see Figure 11).

Fig. 11.

Geometrical sketch of the coordioate system.

*
Let us assume the slope of the channel axis S :: sin Os = - (dZa/ds*)
to
be constant and denote by (s*,n*,z*) an orthogonal coordinate system such that
n* is horizontal and z* is directed upward.
The metric coefficients of the
latter system are readily obtained in the form:
*
hs

[*2
X ,s *
*

hn
*
hz

*2
Y ,s *

[*2

X ,0 *

[*2
X ,z *

*2 ] 1/2
Z ,s *

*2

*2 ] 1/2

*2
Y ,z *

*2 ] 1/2
Z ,z *

* -X * n*
,*
a
cos Os Y a

(47a)

(47b)

cosOs

(47c)

+ Y ,0 * + Z ,0 *

having employed the following relationships for


(X*,Y*,Z *) in terms of (s*,n*,z*)
X

n* *- 1
cos (J.s ro (s*)

the Cartesian coordinates

n*
,*
*
= Y a - cos Os X a

Copyright American Geophysical Union

(48a,b)

Water Resources Monograph

River Meandering

Vol. 12

292

Alternate Bars and Meandering

Z*

= Za* +

z*cosOs

(48c)

*- 1

and having assumed that ro (s*), the curvature of the channel axis, be
positive when the center of curvature lies along the negative n* axis. The
latter is also characterized by torsion ~*(s*), the value of which is given by
the following relationship:

~ (s*)

*-1

tg( Os)ro

(49)

(s*)

In particular, if ro* is constant, the axis becomes a helic so that ~* is also


constant. It is apparent that, if the slope of the channel is very small, as is
typical of meandering rivers, torsion is negligible and the metric coefficients
have the form (47) with cosOs set equal to 1.

Governing equations
We again assume the channel to have rectangular cross section and to be
wide enough for side wall effects to be negligible in the central region of the
flow. However, we model the flow field as 3-D: three-dimensional effects are
indeed crucial to the development of secondary flows.
If Reynolds stresses are modelled through an eddy viscosity ~, Reynolds
equations can be written in the following dimensionless form:
Nuu

,8

vu

,n

+ fJwu ,z + vN
r0

= - NP +

2N

,8

JCO
7

[v,[NU

+ fJJCO [v, [u ,Z +

+
Nuv

,8

+ vv

= -

,n

,Z

vw ,n

w ,8]]

liN

r0

v]]

vN

r0

,8

+ N~fJ
F0

,Z

Nv,s -

~:

vN
ro

u]}

(50a)

u2

+ fJJCO [v, [v ,z + ~fJ


7JCO (v, v)
,n ,n

- 27JCO v'
Nuw ,8

7JN

,8

If- [~ + ~:N]{V1[U.n +

+ {3wv -

P,n +

uv

[N u

,8

+ vN
ro

v - v ]
,n

+ fJww ,z
Copyright American Geophysical Union

w,n ]] ,z

(50b)

Water Resources Monograph

River Meandering

Vol. 12

293

Senlinara and Tubino


= -

pP ,z + 2{3 /CO (v,w)


,z ,z

ICj

-!l.r + NICj [v,(fJu ,z + Nw,s )] ,s


F0

[~ + ~:] [v,(w,n +
Nu ,s +

NQ s,s

(50c)

/3v,z)]

[~
uu

+ VN]
v +
r0

[8on + VN] Q

/3w ,z =

f 0

(50d)

(50e)

where

ro

= f2

Ro

fO

ro

V*

* *
/CO Vo Do

(51a,b)

vn

z*

=*

(51c,d)

Do

with Ro* typical curvature radius of the axis. Furthermore, in (50), (u,v,w)
* P- is the mean
are the 3-D components of the local velocity scaled by Do,
pressure plus 2/3 of turbulent kinetic energy, and v is a curvature ratio
defined in the form
(52)
Further notations and scalings are as before (see Free Bars in Straight
Channels).
Finally, sediment transport is modelled as in the section
Formulation of the Problem (see Free Bars in Straight Channels).

Boundary conditions
At the bed we impose no slip:
u=v=w=O

(z

TJ

(53a,b,c)

+ zoD)

where Zo is the nondimensional conventional reference level for no slip under


uniform conditions.
At the free surface, the kinematic and dynamic conditions are to be
2
satisfied. In nondimensional form at z = FoH we find:

/3w -

vFoH ,n - NuFoH ,s
W

,n +/3v ,z =0

- P+
Nw ,s

2/CO v-r
(Ju ,z

Copyright American Geophysical Union

,z

(54a,b)
(54c,d)

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

294

where the nearly horizontal character of the free surface (at least for relatively
low Froude numbers) has been accounted for.
In the following we will ignore the side wall layers. Thus we will not
force the no slip conditions at the side walls but rather assume the following
kinematic conditions
2

FoH

vdz

(n

'fJ

1)

(55a,b)

Forced Secondary Flow and Bed Topography


in Constant Curvature Wide Channels
We now introduce appropriate assumptions which allow the problem
formulated in the previous section to be reduced to a simplified form amenable
to analytical treatment.
We consider a constant curvature channel and denote by Ro* the curvature
radius of the channel axis. Furthermore, we assume:

fJ ) 1

v ( 1

(56a,b)

Both positions are well-representative of natural streams, although (56a) is too


strong a restriction to describe the meandering process in the advanced stage.
From (56b) the possibility is inferred of studying the central region of the flow
independently of the boundary layers adjacent to the side walls.
By this
scheme, which goes back to Engelund [1974], essentially the existence of two
distinct length scales for transverse variation of flow quantities is assumed:
the channel width which is relevant to the core region and the flow depth
which is relevant to the wall region. Furthermore, the latter flow regions are
assumed to interact weakly. This approach appears to be reasonable inasmuch
as flow and sediment fields can be assumed to be small perturbations of the
uniform field in a wide straight channel.
However, the limits of this
procedure still need be ascertained by comparison with results of fully
nonlinear models of the complete flow field. The restrictions (56) also suggest
the use of a simple turbulence closure obtained by perturbing the eddy
viscosity distribution appropriate to the uniform case.
Before proceeding to derive the linearized model, we notice that it is
convenient to stretch the z-eoordinate defining the variable ( in the form
( = 1

z - FoH(n,s)
D(n,s)

(57)

such that ( C [0,1].


Taking advantage of (56a), we expand the solution in powers of v in the
form
(u,v,w,P ,v-r,H,D)

(uo,O,o,Po,V-ro ,Ho,1)

(58)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

295

Seminara and Tubino

where the observation that the z-component of velocity is 0(,8- 1) smaller than
the remaining components in the core region arises from an order of magnitude
analysis of the equation of continuity (50d).
Substituting from (58) into the flow differential problem (50) at O(vo ) , we
recover the uniform flow solution.
Proceeding at O(v) and neglecting terms O(VCO v/,8) and O(v,8-2), we find
the differential problem governing the secondary flow in the central region:
[V

+ :::

TO

,~,~

u'

_0

,8VCO

1
- -uo- r = - -

,8VCO,8

[H

,8VCO

l,n

2
- uo]

1,~

-!4
Fo

(59a,b)

u'

_1 H
+ ---.!L..Q. [(1 - ()D - F~H ]
,8VCO 1,8,8JCO
1,8
1,8

(59c)

1)D 1, 8
+ F~Hl
]UO'
,8

(59d)

r ,n +:=: +
,~

\11 ,8

= [(( -

where the following expression has been employed


(60)

which can be readily derived by assuming that v * be proportional to the local


T

value of the friction velocity u* and of the flow depth D* keeping the vertical
distribution of the uniform case.
The linearized form of the boundary conditions to be associated with (59)
is obtained from (53-55) and reads
(61a,b,c)

w=r-~-o
....

.:. - FouoH

1,8

= w,~

= 0

(( =

(61d,e,f)

1)

,~

(n

rd( = 0

(62)

1)

'0
We point out that the integral condition (62) on the transverse flow rate
arises because the side wall layers where the flow gradually adjusts to the
no-slip condition have been ignored in the present scheme.
Finally, the
linearized form of (50e) reads:

(v \l1)
[~
VCO ro

,~ ~o

+ f 2D1] +
,8

[[~]
Uo

~o

R(F~Hl -

Copyright American Geophysical Union

Dt}]

,n ,n

(63)

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and IVlea.nclcring

296
where

_r_

/3/UO

(64a,b,c)

where 00 and <Po are Shields parameter and bed load function associated with
the uniform configuration.

Fully developed flow


We now examine the simplest case, that of fully developed flow, occurring
in the downstream portion of the bend, far enough from the initial section for
entrance effects to be no longer felt by the flow field.
Thus we set 8/ Os = 0 in (59) and (63) and reduce the governing
differential problem to the following simplified form:

[v 1Or O /,]

I'

,') ,')

= -1

/3/CQ

(65)

[h o - uo]
1

J fod( = 0

(66a,b,c)

'0
(67)

[VrQlIIo},

[VrQlIIo),o

(\lI o),

= - /CQ(do - 1)

(\lI 0,,) '=1

(68)
(69a,b)

having expressed the solution in the following form:


f

f o(()

(\lI,Hl,Dt)

(\I1o((),ho,d o)n

=: =

(70a,b,c)

The differential equation (65) with the conditions (66a-e), assuming a structure
for v TO' can be solved for fo( () and hOe We point out that (65) expresses
mathematically the need for a secondary circulation producing transverse shear
stresses able to compensate for the dynamic unbalance between centrifugal
forces and transverse pressure gradient. The condition (66c) determines the
latter such to give rise to vanishing transverse flow rate. Using Dean '8 [1974]
form of v TO' the distribution of ro is found as given by ST (Figure 4, pg. 28).
The condition (67) expresses the requirement that the transverse sediment
flow rate should vanish: this determines the transverse slope of the channel,
Le. the value of do. Figure 12 shows the predicted dependence of do on () and
ds assuming the undisturbed bed to be plane.
Finally, (68) shows (as anticipated in the Introductory Remarks to Forced
Bars in Curved Channels) that the main flow is only affected by secondary
flow through the interaction of the latter with the bottom.
Indeed a
perturbation of flow depth gives rise to a perturbation of the eddy viscosity

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

297

Seluinara and Tubino


25.

do

3D

ds
.050

20

20.

.010
.001

15.

10.

5.

.05

Fig. 12.

.50

.20

The predicted value of dO is plotted as a function of Shields stress


values of the grain ratio d s.

(J

for given

(which is felt through the term - v'COd o in (68)). Further contribution to the
latter comes from the perturbation of the longitudinal velocity as shown in
(60). Finally, the transverse variation of longitudinal slope is felt through the
term v'CO in (68). The latter equation with the boundary conditions (69) is
readily solved in the form:
\110

2( do - 1)uo( ()

(71)

having neglected (0 with respect to (do - 1).


From l70b), (71) and Figure 12 it is clear that:
in the fixed bed case (do = 0) the sign of \110 is negative whence the
predicted thread of high velocity is located along the inside of the
bend where n is negative;
in the mobile bed case deepening of the outer flow pushes the thread
of high velocity towards the outside of the bend.
The former prediction is not confirmed by the experimental observations of
Rozovskij [1957] and I(ikkawa et al. [1976] which exhibit an increasing value of
the longitudinal velocity in the transverse direction.
A recent successful
attempt to explain this contradiction by accounting for the effect of transverse
redistribution of longitudinal momentum is due to Johannesson and Parker [see
paper in this volume]. Although we feel that the singular perturbation nature
of the latter effect still requires elucidation, various arguments suggest that the
main point of Johannesson and Parker's paper, namely the important role of
Indeed
side walls in controlling momentum redistribution, is quite correct.
I(ikkawa et ale 's [1976] results of run Fl (characterized by {) = 10) show a
slight outward decrease of longitudinal velocity in the central region of the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

.Alternate Bars and l\1eandering

298

channel, a trend which changes quite sharply as the side walls are approached.
This feature is decreasingly evident in the results of runs 2 and 3
characterized by the values 9.1 and 7.9 of (3 respectively. Thus it appears
that side wall effects are felt in the core region even for relatively large values
of the width ratio. On the other hand, the O{v2 ) contribution to momentum
redistribution in the core region, which is associated with terms of the type
v2 ro\l10, may be partially responsible for the above effect. Indeed, the leading
order contribution vro{duo/dn) which vanishes in the core region for (3 -1 00 is
at most O{ vr0) in the same region for moderate values of fJ. Thus the two
contributions become comparable for values of v such that v
O(l/\l1o).
Equation (71) and Figure 12 suggest that the latter condition is satisfied for
relatively low values of v.
N

Entrance effects
A question of both practical and conceptual interest is
how fast flow and bed topography develop starting
conditions.
A simple estimate of the characteristic length involved
process is readily obtained on purely dimensional ground.

that of ascertaining
from given initial
in the development
In fact, both (59a)

and (59c) show that the lengthscale Le for flow adjustment is B*/ ({3J(SOfJ TO)
having denoted by fJ TO a typical value of the dimensionless eddy viscosity. If
the above argument is applied to [(ikkawa et ale 's [1976] experiment Ft, using
the constant value .077 suggested by Engelund [1974] for fJTO' one finds Le ~
16 B* which appears to be close enough to the experimental value 13 B*.
Of course, the detailed structure of the adjustment process can be
ascertained by solving the differential system (59-{)4).
We are currently
investigating this problem with the aim of explaining the 'overshooting'
phenomenon experimentally detected by Struiksma et al. [1986], which still
needs an overly convincing explanation.
Finally, we point out that the approximations on which the formulation of
the linearized problem (59-{)4) was based rule out the possibility for the
upstream influence of bend flow to be detected. Indeed this follows from the
parabolic character of the above differential problem which has thus lost its
original elliptic nature.

Forced Secondary Flow and Bed Topography in Weakly Meandering


Wide Channels: 3-D Model and Resonance
Let us now consider wide channels characterized by variable curvature and
focus our attention on the case, vastly explored in the literature, when the
channel axis is described by a sine-generated curve. Thus let us set:
- 1

1'0 (s)

el

C.C.

exp(iAms)

C.C.

(v ( 1, (3

1)

(72a,b,c)

where Am is the dimensionless meander wavenumber and v is now the ratio


between the half-width of the channel B* and twice the radius of curvature

*
(Ro/2)
at the bend apex. Under the above assumptions, the study of flow
and bed topography can be formulated in a similar way as discussed in the
previous section.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

299

Sen1inara and Tubino

Fully developed flow


We examine the case of fully developed flow, Le. we assume that entrance
effects are no longer felt by the flow; the meander under investigation should
then be only one of a sequence of identical meanders.
We can then expand the solution in powers of v in the form:
(u,v,w,P,vT,H,D) = (uo,O,O,Po,v TO ,Ho,l)

(73)
so that we recover at O(Vo ) the uniform flow solution while at O(v) we find a
differential problem identical with (59-64) provided oj Os be replaced by (iA m).
The solution of the latter has been determined by ST in a more general
context where transport in suspension was also accounted for.
Previously,
Gottlieb [1976] had obtained a 3-D solution for the case of bedload transport
only.
We refer to ST for details of the mathematical procedure employed to
solve the O(v) differential problem. It suffices here to say that we expanded
the solution in the form
00

('l1,3,D t ,Ht)

ro

('l1 on,O,d on,h on)

+ b

('l1m,3m,dtm,htm)sin(Mn)

00

b
m=l

r m cos(Mn)

(74a)

m=l

=~

(2m

1)

(74b,c)

In words, we split ,the solution into a 'slowly varying' component


(ro, 'l1on,don,h on) corresponding to the fully developed solution for the constant
curvature case evaluated locally and a further component needed to satisfy
flow and sediment continuity which are now unbalanced due to the effect of
longitudinal variations of flow and sediment load.
The above idea will be
further employed in the next section where, following Engelund's [1974]
pioneering work, the latter component of the solution will be obtained from a
2-D model of the flow field. We point out explicitly that even multiples of
1r/2 are discarded from the solution (74) in order to let the component of r
with non-zero depth average vanish at the side walls. However, the structure
of the above solution can obviously be valid only \vithin the core region since
The differential problems
it does not satisfy no-slip at the side walls.
obtained on substituting from (74) into the governing equations were solved in
ST analytically. in terms of m-independent solutions which were found
numerically.
In Figure 13 we show a comparison between the bed topography predicted
by the present model for the experimental conditions of Ikeda and Nishimura
[1985]. In the latter, a very small transport in suspension was also present (of
the order of few percent of the total load) and the bottom was ripple-covered.
We assumed a dune covered bottom and employed Engelund and Hansen's
[1967] approach to evaluate frictional effects and sediment load.
Bed
topography appears to be predicted fairly well by the theory of ST.
The main observation arising from Figure 13 is the downstream shifting of
the point bar with respect to the bend apex. A physical interpretation of the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

300

0.4

O.t.

0.4

0.2
0.4
0.2

o -0.2

Fig. 13.

Comparison between the bed topography (dimensionless scour) predicted by ST's


[1985] model (B) and Ikeda and Nishimura's [1985] experimental findings (A)
.2).
(contour intervals

above was given in ST examining how the effect of a variable curvature affects
the mass balance imposed by the sediment continuity equation.

Resonance
Further results of the 3-D model are found in ST. Here we point out an
important effect, which was first detected by BS within the context of a 2-D
model, and does emerge also in ST's 3-D model. For given values of (J and
ds , there exists a fairly wide range of values of the width ratio (j such that
the response of flow and bed topography to the forcing effect of curvature
exhibits a sharp peak. This is shown in Figure 14 where the depth-averaged
value of \11 evaluated at the outer wall as obtained from ST's model is plotted
versus Am and compared with the results of the quasi-2-D model discussed in
the next section.
The cause for the appearance of the above peak is readily understandable if
the system 'liquid flowing in a channel with erodible bottom' is interpreted as
a forced system, the forcing effect being associated with a periodic spatial
variation of curvature. Any mechanical system subject to perturbations, in the
absence of external forcing, exhibits a 'natural' response which leads to
damping of the perturbations provided the system is stable or allows for their
amplification if the system is unstable. The natural responses of the system

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Senlinara and Tubino

301

.3
2Da
1
-.1

Ii
1"

~=-

[ 20
30 -----_.
40 - _ . -

~=

[ 20
30 -----_.
40 - _ . -

~ =-

[ 20
30 -----_.
40 - - . -

-.3

\.

-.5
.3
2Db
r-l

II

9;

-.1

\'/.
"/\

a.>

0::

-.3
-.5

.3
3D
1

f\

)~\\
I

-.1

t.

~~';"
-.3

.--=-=--:. -=-=-::-.:..-=-::- .

, I

-.5
0.0

Fig. 14.

.1

.2

.3

.4

Am

.5

The component of the perturbation of longitudinal velocity in phase with curvature


evaluated at the outer wall is plotted versus the wavenumber ,\m for given values
of fJ (2Da
BS's [1985] model, 2Db
present 2-D model, 3D
ST's [1985]
model).

considered herein were analyzed in the previous chapter and consist of


sediment waves that we called 'free' bars. A crucial point to understand is
that the class of potential natural responses of the system encompasses a wide
spectrum of sediment waves, each characterized by a different wavelength,
migration speed and temporal growth rate, satisfying the dispersion relationship
(67). In particular, Figure 15 shows that a range of wavenumbers exists such
that the corresponding free bars are quasi-steady, Le., their amplification rates
and migration speeds are close to zero. In the absence of any forcing effect,
Le. in a straight channel, these bars would not naturally develop; on the
contrary those which exhibit the maximum amplification rate are found to
grow spontaneously. However, if the system is forced by a periodic variation
of curvature, as occurs in meandering channels, provided the wavenumber of

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and 1\1eandering

302
.3

Q
Qo

-00
Qo

----~=40

<=
Resonant
wavenumber
range

~=30
~=20

o.

- .3

l . . . - - _ - ' - - _ . . . . L . . . . - _ . . . . & . . . . - _........._ . . . . . . L - _ - - - ' - - _ - - - - L . _ - - - - ' - _ - - - - ' _ - - - - I

0.0

Fig. 15.

.2

.4

.6

.8

-2. 5

1.0

The normalized growth rate (O/QO) and angular frequency (- w/Qo) of free bars
are plotted as functions of the bar wavenumber A for given values of the width
.25, d s
.005).
ratio 13( (}

the latter falls in the range of quasi-steady free bars, a quasi-resonant


response occurs which leads to the peaks of Figure 14.
This explanation
becomes mathematically obvious observing that under the latter conditions
0 = W = 0) the homogeneous part of the linear differential system for
r, \II ,Dl,H t ) becomes identical with that describing the linear development of
ree bars.
We point out an interesting feature arising from Figure 14: increasing the
meander wavenumber through the resonant range leads to a continuous, though
quite rapid, increase of the phase lag of longitudinal speed with respect to
curvature from a value falling in the range (0 .;- 1r/2) (when both Re(\II)n=1

and Im(ljI)n=l are positive) to a value in the range (11" .;.

11") (when both

Re(\II)n=1 and Im(\II)n=1 are negative).


This feature was also present in
Engelund's [1974] results, although it was not associated with resonance and
occurred for higher values of Am.
The relevance of the resonance phenomenon discussed above to the process
of meander formation is discussed below (see Bend Theory of Meander
Formation).

Forced Secondary Flow and Bed Topography in Weakly Meandering


Wide Channels: 2-D Model and Resonance
Although the formulation and solution of the 3-D model proposed in the
previous section does not exhibit a prohibitive mathematical complexity, at
least at the linear level, still it would be highly desirable to achieve results of
comparable accuracy by means of a simpler depth-averaged model.
This

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

303

Seminara and Tubino

would provide a tool suitable to investigate more complex phenomena involving


nonlinear interactions like those which will be discussed later (see Interactions
Between Free and Forced Bars).

Formulation
A proper formulation of a 2-D model of flow and bed topography in
curved channels requires accounting appropriately for the 'dispersive' effect
associated with transport of longitudinal momentum by the component of
transverse velocity with zero depth average.
This was first pointed out by
Kalkwijk and De Vriend [1980] who suggested the following decomposition of
the velocity field:

v[fO(z)el

(75a)

uo(z)U(s,n)

(75b)

+ c.c.]U(s,n) + uo(z)V(s,n)

where (U,V) are the depth-averaged longitudinal and transverse components of


velocity and uo(z) is the uniform unperturbed velocity distribution.
Notice
that the vertical distribution associated with the depth-averaged components of
u and v is exactly uo(z) only at a linear level and in the limit Am -i 0, as can
be shown from the fully 3-D differential problem discussed in the previous
section.
Thus the validity of assumption (75) for finite Am and at higher
order in v is only approximate; in this section we explore its limits at a linear
level.
By substituting from (75) into the differential system (50), performing
depth integration and neglecting terms o[(/CO/ P)v2 ] we find the following
differential equations for the depth-averaged component of the flow field
UU

,8

VU

,0

Ts = - v 9t( s)
+ /l.I...n

H ,8

[n

[n

nTn = - v2 9t(s)

[n[vv

1 2
2
- v9tl(S)n(U D),o - v2 ~(s) n (U D),o

uv

,8

VV ,0

//sel(S){~

H ,0

{3

[(DU ),s

(F oH _ D) ,t

,8

VU ]
,0

2U

,0

uv]

2]
+

(76)
H ,0

+ PnT n] -

+ 2(UDV),nl} - //(iA mkl el + c.c)U

- 2//2 ~(s) [~ (UVD),n


(DU)

[/l.I...ns

+ UV] -

//2

sea(s)

8,8

0,0

) - - Q
-

(77)

(U D),n

S~
+ //se(
v n:ge s)

+ (DV) ,0 = - v9t(s)[VD + n(DV) ,0]

Q (Q

(Q

(78)
- nQ 8,8 ) (79)

where

9t(s)

9t2(S)

k 1e2

=
+

el
kl

+ c.c.
+

c.c.

= k1el + c.c.
9t3(S) = k2e 2 + k3 +

(80a,b)

9tl(S)

Copyright American Geophysical Union

c.c.

(80c,d)

Water Resources Monograph

River Meandering

Vol. 12

30:1

A.lternate Bars and


ek = exp[k(iAmS)]

~feandering

(80e)

and the parameters k1, k2 and k 3, arising from the velocity decomposition (75),
have the following form:

kt

=f

1
2

k3 = fifo 1 d( (81a,b,c)

uorod(

'0

'0

The above equations


conditions:

must

be supplemented

v = Qn =

(n

by

:i:

the

following

boundary

(82a,b)

1)

The constitutive relationships (10) and (12) are now affected by the component
of secondary flow with zero depth-average and read:
T

(Ts,T n )

sina

(U,V + vU(k4et + c.c.))Cv'u 2 +

=h -T

I :1

p~

(FoR - D)

v2

(83)
(84)

,n

where
(85)
222
The dependence of the parameters ,81<1, p k2, P k3 and p k 4 on Am and d s as
obtained using the results of the 3-D model discussed in the previous section
is shown in Figure 16 (a-d) and compared with I(alkwijk and de Vriend's
[1980] results and Engelund's [1974] suggestion for ,81<4. Notice that the above
quantities slightly depend also on the width ratio (3.

Linearization
We again consider the case of fully developed flow in a sequence of
identical meanders characterized by typical curvature ratios v small enough for
linearization to be a valid approximation. We then expand the solution for
(U,V,H,D) in powers of v in the form
(U,V,H,D,)

(I,O,Ho,l)

v{[Ut(n),Vt(n),Ht(n),Dt(n)]exp(iAms)

c.c.}

O(v2 )

(86)

The linear differential problem for (Ul,Vl,Hl,D t ) is then found by substituting


from (86) into (76-85) and reads:

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Selninara and Tubino


a

305

.8

.6

-----------------------

.4

.2

__ __ .-_. __ .__ __ ._- -_.-_.-C ' - ' - ' , - ' - ' - ' --'-'-'-'-'

0.0

._ _ . _ _ ._ _ ._ _ . _ _ ._ _ ._ _ . _ - . - - . _ _ .

-.2 -3

-2

-1

10

.--

._._._._._.~.~~:::::--

.---'--'--'

-10

--.--.
---'

-20
-3

Fig. 16.

-2

Log 10 (d s )

-1

(a~) The parameters ,8kt, ,B2 k2, p2 k3, ,8k4 are plotted versus the grain ratio d s
for given values of the wavenumber Am (= 0.,.2,.4)(.8 = 30).
- - - - real part
_ e _ e _ imaginary part
- - - - Fig. 16a - Kalkwijk and De Vriend [1980]
- Fig. 16b - Engelund [1974].

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

306
c

30

25

20

15

10

-3

-2

:=:=:=:=:=::=:::=:==:===:==:

-5,
l----------------------

-10

Fig. 16. cont.

Copyright American Geophysical Union

-1

Water Resources Monograph

River Meandering

Vol. 12

Seminara and Tubino

307

(87)

Vi

R(FoH 1,n - D1,n ) - k4

(n = 1)

(88a,b)

where L is identical with the linear 'bar' differential operator with zero
complex growth rate and A replaced by the meander wavenumber Am. We
point out that the terms proportional to k1 and k 4 in (87) had been neglected
in BS. Furthermore, an algebraic error was contained in equation (3.9d) of
BS (see Corrigendum JFM in print). Figure 14 shows a comparison between
the amplitude of the depth-averaged longitudinal component of the
perturbation velocity evaluated at the outer wall as obtained from ST's 3-D
model, BS's 2-D model corrected from the algebraic error and the present
model. It appears that:
the resonant effect exhibits similar features in all the models;
the effect of momentum redistribution accounted for in the present 2-D
model affects the amplitude of the flow response but leaves its
dependence on Am unaltered;
3-D effects shift the peak in the response to slightly larger values of

Am.

Bend Theory of Meander Formation


The models for flow and bed topography in weakly meandering channels
developed in the previous sections can be set as the basis for a theory of
meander formation provided a model is established whereby meander growth is
somehow associated with flow perturbations with respect to the straight
channel configuration assumed to be in equilibrium.
The first attempt in this direction was proposed by Ikeda et ale [1981] who
assumed that the rate of bank retreat (e) be proportional to the difference
between the near-bank depth averaged longitudinal velocity and the reach
averaged velocity. In the initial stage of the process of meander formation,
the rate of bank retreat would then be proportional to the curvature ratio v.
This is consistent with the observations of Brice fI982], Hooke r1980], Hickin
and Nanson [1975] and Nanson and Hickin [1983 (see also Odgaard [1987])
who demonstrated that the width-radius ratio is a dominant factor in
controlling the rate of bank retreat. We point out that other assumptions,
ITs - TsO I as proposed in BS, would still be consistent with the
like e
above observations.
Indeed, any component of flow perturbations is
proportional to v for weakly meandering channels.
It will appear in the
following that, if meander formation is associated with resonance, different
assumptions do not lead to large differences in the values of selected
wavenumbers.
However, the detailed knowledge of the mechanism of bank
erosion is crucial to meander development in advanced stage and will have to
be further investigated in order to model the latter phenomenon (in this
respect see Hasegawa's paper in this volume).
The model for flow and bed topography employed by Ikeda et ale [1981]
was based on Engelund's [1974] approximate procedure and could not produce
the resonant peaks showed by ST's 3-D and BS's 2-D models for reasons
which have been discussed in BS [see also Johannesson and Parker, paper in
this volume].
(X

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

308
a

.2

/'

---- --- --- - -- - /

/'
,/

.1

~,;:::P'

/--

/~'

'

0.0
0

10

.3

{}=
d s=

Am

20

30

50

.50
.005

.2

/:::2=:'

------- -----------....,

.1

40

'

0.0

o
Fig, 17

10

20

30

40

50

(a-b) Selected values of meander wavenumber as a function of the width ratio f3


for given values of Shields stress (J and grain ratio d s (the unperturbed bed is
assumed to be plane).
- - - - BS's 2-D model (corrected)
- - - - present 2-D model
- - - - - - ST's 3-D model
- - Ikeda et al. '8 model [1981].

Employing a 2-D model with the initial bottom configuration assumed to


be plane and the bank erosion criterion mentioned above BS showed that Ikeda
et ale '8 (1981) bend instability is associated with 'resonance' in that flow and
bottom perturbations (in particular near bank longitudinal shear stress) peak at
values of meander wavenumbers corresponding to a particular quasi--{3teady
'bar-type' natural response of the system, as discussed above (see Forced
Secondary Flow and Bed Topography in Weakly Meandering Wide Channels).
Similar conclusions are derived from ST's 3-D model.
In Figure 17 we

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Seminara and Tubino

309

compare the wavenumbers corresponding to maximum bend growth as given by


BS's 2-D model corrected from the algebraic error with those found by the
present 2-D model and by ST's 3-D model. Also shown are Ikeda et ale '8
l1981] results. It appears that the effects neglected in BS do not affect the
selection of the preferred meander wavelength whereas some variations are
brought up by three-dimensional effects which tend to increase the values of
preferred wavenumbers in the plane bed case and to decrease them in the
dune case.
We point out that the actual role played by resonance in the process of
meander formation still requires to be further elucidated. We are presently
investigating the possibility that steady, spatially growing perturbations may
develop in the straight channel as a secondary bifurcation and be the
triggering mechanism of bend instability.

Interactions betwoon Froo and Forced Bars

Introductory Remarks
As discussed in the Introduction to this paper, the problem of whether free
bars can coexist with point bars which are 'forced' by channel curvature bears
both a conceptual and practical importance.
Conceptually the above question is related to the problem of meander
formation. If the view is taken to look at the latter as a 'bend instability'
the as yet untackled question arises of the effect that the presence of alternate
'free' bars in the initially straight channel may have on bend growth and on
the selection of the most unstable meander wavelength. The assumption that
the propagating character of alternate bars implies that no steady effect can be
produced so as to affect bend instability will be shown to be 'a priori'
incorrect.
The practical relevance of the problem is twofold:
in order to predict bed topography in meandering channels the possible
interaction between alternate bars and point bars needs to be
investigated;
in order to suppress alternate bar formation in artificially straightened
channels it has been suggested (KM) that the channel should be given
a 'sufficient' sinuosity. The important results reported by KM have
shown that, at least under the conditions considered in their
experiments, a critical sinuosity exists for any given set of flow and
sediment characteristics such that alternate bars are indeed suppressed.
The theory outlined in the next sections will be seen to be suitable to
attack both problems.

Formulation of the Interaction Problem


The critical values of v for alternate bars suppression observed by KM
never exceeded 0.1.
This suggests the suitability of a 'weak meandering'
approach to investigate the interaction problem.
Let us also restrict our
attention to the weakly nonlinear regime defined in a neighborhood of the
critical conditions for alternate bar formation by (25-26). From CST's results
we can expect the analysis to work successfully even for ( 0(1).
Following the lead of BS and CST we then seek a mixed solution of the
problem in the form
N

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

310

l\lternate Bars and Meandering

{1/2[Af1S1Et

{[A (f22C2

f02)E2

v2 [U2(n)e2

c.c.

v 2 {1/2 [AFto(n)Et

+
+

c.c]

c.c.

Uo(n)]
c.c.]

v[U1(n)et

+
+

AA(f2oC2

c.c.]

foo)]

V{1/2 [AFll(n)etEt

v{[et(AAFot(n)

AFll(n)eti~~t

Fot(n))

c.c.]

c.c.]
(89)

Similar expansions are assumed for V, H, and D.


In (89), A and ware the alternate bar critical wavenumber and angular
frequency .predicted by linear theory and A(T) is a slowly varying amplitude
function of time (T = it), describing bar growth, to be determined.
Some important qualitative features of the problems posed in the previous
section are implied by the structure of the expansion (89).
In fact, the
nonlinear interactions involving the fundamental O( {1/2) alternate bar mode and
the fundamental O(v) point bar mode lead to O( {,V2 ,V{1/2) second order free,
forced and mixed modes. Further interactions lead to reproducing
the fundamental alternate bar mode at 0(V2 {1/2);
the fundamental point bar mode at O(V{).
The former effect is 'forced' by curvature and is comparable with the
'natural' effect of self-interactions of the free bar mode provided v2 {1/2
O( {1/2), Le. v
O( {1/2). This appears to be physically sensible implying that
the interaction is significant when forced and free bars have comparable
magnitudes.
The latter effect simply implies that from nonlinear interactions between
propagating and steady modes, a steady contribution arises. This is O(V{) and
whether it may be comparable with the fundamental O( v) effect of the forced
bar mode when the amplitude of alternate bars is large enough needs to be
investigated.
N

Suppression of Free Bars in Meandering Channels


The former effect mentioned above can be investigated quantitatively by
assuming that v = ki{1/2 with ki
0(1), substituting from the above
expansions into the governing differential problem, solving the ordinary
differential systeills obtained at various orders and imposing the solvability
condition required at O( {3/2 ,v2 {1/2).
This procedure can be performed
analytically [Tubino and Seminara, 1988] and leads to deriving the following
amplitude equation for A(T)
N

(90)
where all is a function of 0 and d s arising from the effect of mixed
interactions while at and a2 are coefficients identical with those of the
amplitude equation (36).
As ki 4 0, (90) tends to the amplitude equation derived by CST for the
straight channel case. A supercritical equilibrium amplitude of alternate bars
2

exists provided R e ( at + k i all) /Re ( a2) is negative. This condition was found
to occur for vanishing ki by CST.
Thus if curvature tends to sup~ress
alternate bars we expect that sgn[Re ( a11)] =1= sgn[Re( at)] so that R e( O't + k i 0'11)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

311

Selninara and Tubino


70

NON MIGRATING

35

8
0

MIGRATING

0.0

.2

.1

.3

.4

.5

.6

Am

.7

Fig. 18a. The critical value ere predicted by the present theory is compared with KM's
[1974] experimental findings.

10.

7],
5.

--- ---

o.
.1

-5.

.3

.5

.7

_ _ _ _ Real part
_______ Imag inary part

-10

Fig. 18b. The fundamental forced component of bottom elevation is plotted versus AId
showing that the resonant peak corresponds to maximum free bar suppression.
Data as in KM's (1974) experiments.

changes sign for some critical value k e of ki (Le. of sinuosity). Preliminary


calculations appear to support this expectation.
Quantitative comparison is pursued with KM's results in Figure 18a.
Calculations were performed using the values of 0, d s and (j of KM
experiments, assuming the relative density of coal to be 1.5 and the
unperturbed bed to be plane, predicting the bar wavelength by our linear
theory and finally choosing KM's values of the meander wavelengths.
It
appears that the predicted trend of the critical value Q'e of the angle between
KM's straight segments as a function of Am agrees with the experimental
observations.
This may be readily understood in terms of point bars
resonance. Indeed Figures 18 show that the minimum value of Q'e for free
bars suppression occurs for values of Am corresponding to the resonant peak for
the point forced bar. Thus, the present results can be also considered as an
indirect demonstration that resonance is experimentally observed.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

312

Let us finally come to some conclusions. For values of Ami A such that a
critical value k c defined as "-[Re(G't)]/[Re(G'u)] exists, the critical value Vc of the
curvature ratio able to suppress alternate bars is given by the following
relationship
(91)
The above formula predicts a relatively wide variation of Vc with (3. This
result is physically sensible: indeed the larger the amplitude of free bars, the
more sinuous the channel should be in order to suppress them.
We notice that the variation of Vc with (3 predicted by (91) could not be
detected by KM, who designed their experiments such that (3 was held
constant.

Injluence of Free Bars on Forced Bars


The modification of the steady point bar arising from mixed interactions
can be investigated by substituting from the expansion (89) into the governing
differential problem and solving the differential system obtained at various
orders up to O(Vi).
Results obtained following this procedure can be summarized as follows:
i)
the O( Vi) correction of steady forced bar associated with
interactions is comparable with the amplitude of the fundamental
O(v) component only for values of (3 sufficiently large to have high
amplitude free bars but relatively low amplitude forced bars. The
latter condition cannot occur if (3 falls into the resonant range;
ii)
a resonant peak appears in the O( Vi) correction but is shifted
towards values of Am higher than those characterizing the resonant
peak of the fundamental.
Result i) points out a difficulty which does not seem to have been
appreciated before. Is it possible to validate theoretical models of flow and
bed topography in weakly meandering channels where the interaction of
free-forced bars is not accounted for by performing comparison with
experimental results (like those by Gottlieb [1976]) referring to experiments
where free and forced bars are found to coexist? In fact, even filtering out
the unsteady part of the signal as Gottlieb [19761 did, the steady component
thus obtained is not purely associated with the forced bar but also with the
O( Vi) effect of mixed interactions. In light of the above results, this difficulty
appears to become decreasingly important as i ~ 0 and disappears when the
experiments are performed in channels narrow enough to prevent free bar
formation or sinuous enough for alternate bar suppression.
Result ii) suggests that mixed interactions do affect meander growth.
Detailed calculations and comparison with experimental results are being
performed.
Figure 19 makes statements i) and ii) quantitative for a particular case.
The presence of free migrating bars in the originally straight channel may
also influence meander formation and the selection of meander wavelength, in a
different and more subtle way, by affecting the threshold conditions' for the
development of steady, spatially growing disturbances which may arise as a
secondary bifurcation in the straight configuration and then trigger bend
instability. We are currently investigating along these lines.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

313

Seminara and Tubino


M

II

-.c

~=

.4

+J

.,-1

g .0

20.

~= .1

>,

__ -_

d s = .01

I----=:;;;~.......~-----.;:::......t.~_-_-_-_-_-_-_~_I-_---_-_-_-+_-_-_-_-_-_-_-_~

r-t

aJ

:>

r-t

co

-.4

',-1

_ _ _ O(V) camp.
______ .0 (V8) camp.
_ _ . _ 0 (11) + 0 (V8 )

B-.8

"d
.,-1

bO

= 30.

.4

~= .1

CH

.01

.0

',-1

+J

co

-e -.4
:J

+J

~-.8

______ .0 (V8)

_ _ _ 0(11)

camp.
camp.

_ _ . _ 0 (~')

0 (VS)

Q)

0:::

Fig. 19.

0.0

.1

.2

.3

.4

Am

.5

The order (v) and (Vf) components of the perturbation of longitudinal velocity
evaluated at the outer wall are plotted versus the wavenumber ~m.

Perspectives
The theoretical framework developed in the previous sections appears to
draw a consistent picture of the various phenomena involved in the initial
process of meander formation in alluvial channels.
Let us consider an
originally straight channel with cohesionless boundary and uniform grain size.
Unless the channel is too narrow or too wide, it undergoes an instability
process which, on a relatively fast time scale, leads to the formation of
As channel widening proceeds,
migrating perturbations, alternate bars.
provided the width ratio exceeds a second threshold value above which spatial
disturbances may grow, the channel undergoes a second instability process, a
planimetric one. The latter process, the details of which still require to be
elucidated, is not primarily determined but is definitely affected by the
presence of free bars.
As the amplitude of meander increases, so does
sinuousity, so that free bars tend to decrease their amplitude till they
disappear, leaving only forced bars to induce bank erosion. The above picture
definitely needs further refinements to be able to describe features of the
process which have so far been neglected, namely transport in suspension, grain
sorting, entrance and wall effects, and unsteadiness. Further experimental and
field verification of the whole picture is also needed through carefully designed
experiments.
However, the big challenge for the near fu t ure appears to be the
development of models able to account for strong _ nonlinearities of bank

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

314

Alternate Bars and Meandering

erosion, which still needs further mechanically sound elucidation.


Seminal
ideas in this respect have been put forward in recent years by Parker et ale
[19821, Dietrich and Smith [1983], and Smith and McLean [1984].
The cooperative effort of engineers and geomorphologists will still be an
essential ingredient of our fascinating story.

Notation

amplitude function of free bars


equilibrium amplitude of free bars
bed surface
half width of the channel
friction coefficient
arbitrary

constant

in

the

solution

of

the

amplitude

equation (37)

Cm
C

= cos(mnn/2) (m = 1,2,... )
friction coefficient of uniform unperturbed flow
torsion of channel axis
flow depth

Do*

averaged depth of uniform unperturbed flow

-*
Do

reference flow depth in unsteady flow

Dot
do

unsteady part of the basic flow depth


flow depth in fully developed flow in constant curvature
channel

d*s
*
*
d s (=ds/Do)
D

(p=I,2,3)

sediment diameter
roughness (grain) parameter
dt,d22,do2,d2o,doo,dat,dtm

(m=I,2, ... ) components of flow

depth at various orders


Ek

e
ek

Fo
ft ,f2

= exp[ki (As - wt)]

(k=I,2, ... )

rate of bank retreat


= exp[k(iAms)
(k=1,2, ... )
unperturbed Froude number
coefficients of (63) (see (64a,b))

gravitational acceleration

water surface elevation

HBM

difference between maximum and minimum dimensionless


bed elevations within a bar unit
water surface elevation of uniform unperturbed flow

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

315

Seminara and Tubino


ho

water surface elevation in fully developed flow in constant


curvature channel

* * *z
hs,hn,h
Hp (p=I,2,3)

metric coefficients
hl,h22,ho2,h2o,hoo,hoo,hal,htm (m=I,2, ... )
water surface elevation at various orders

kc

critical value of ki for free bar suppression

ki

(=.V/t: 1/2 )

k1,k2,ka,k 4
Le

*
Lm
m

parameter which defines the number of braids


(see (51b))

transverse coordinate

p
Po

of

momentum redistribution coefficients


lengthscale for flow adjustment
meander wavelength

P(s)

components

parametric form of the channel axis


mean pressure plus 2/3 of turbulent kinetic energy
unperturbed value of P

first order component of


sediment porosity

= (Qs,Qn) = (QTl,QT2,Qv) sediment flow rate

Qo

ratio between the scales of sediment discharge and flow

P1

rate (7a)
Qnp (p=I,2,3)

q~b) components of the transverse sediment flow rate at

Qsp (p=I,2,3)

various orders
q~b) components of the longitudinal sediment flow rate at
various orders

R
Ro*

9l, 9l1, 9l2, 9la


r

coefficient (see (64c))


typical curvature radius of the channel axis
(see (80a,b,c,d))
empirical coefficient (13)
radius of curvature of the channel axis

ro
S

channel slope

Sm
s

= sin(mnn/2)
(m=I,2,... )
longi t udinal coordinate

T (=t:t)

slow time variable describing the time development of


free bars

t
U

time
depth averaged longitudinal component of velocity

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and l\feandering

316

parameter controlling the effect of unsteadiness of free bar


development

Uo*

averaged speed of uniform unperturbed flow

-*
Uo

reference value of flow velocity in unsteady flow

uo(z)

longitudinal component of local velocity (3D)


uniform unperturbed velocity distribution
friction velocity

Up (p=1,2,3)

v
v
V p p = 1,2,3)
w

X*,Y*,Z*
* * *
Xa,Ya,Za
z

zo

UO,Fl1, Fl1,FI0,Fot,Fol,fl,f22,fo2,f20,foo,f31
longitudinal velocity at various orders

components

of

depth averaged transverse component of velocity


transverse component of local velocity (3D)
gl,g22,g02,goo,g31

components

of

transverse

velocity

at

various orders
z-component of local velocity (3D)
Cartesian coordinates
Cartesian coordinates of the channel axis
coordinate orthogonal to the plane (s,n)
dimensionless reference level for no slip condition
angle between the straight segments of I(inoshita and

Miwa's [1974] experiments


critical value of a for free bar suppression
complex coefficients of the amplitude equation for A (see
(36)
comp I ex coefficient of the amplitude equation for A in
unsteady flow (see (44)
complex coefficient of the amplitude equation for A in
meandering channel (see (90)

*
/3 (=:B*/Do)
1J

width ratio
reference value of /3 in unsteady flow

/3e

critical value of /3 for free bar formation

O(v) component of local transverse velocity

(=:0/ t) 0(1) constant in unsteady flow

rm

o
o

(m=O,l,... ) coefficients of the expansion for


angle between particle velocity and

!' direction

small parameter in unsteady flow (see (42b))


phase lags of bottom profile with respect to various flow
characteristics

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

317

Scnlillara and Tubino

phase lags of the O( f) solution for the bottom elevation


(='(~Pc) / Pc)

(0

11
11M
111, 'f/22, 'f/20, 1102

7J
Ocr

00
Os
A

Ac

Am
V

*
(=.B*/Ro)

(see 40b)
perturbation parameter of free bars
references value of ( for no slip condition
bot tom elevation
maximum relative dimensionless scour of free bars
components of bottom elevation at various orders
Shields parameter
reference value of 0 in unsteady flow
critical value of 0 for sediment motion
unperturbed Shields parameter
inclinat i on of channel axis
free bar wavenumber
critical value of A
meander wavenumber
curvature ratio
critical value of v for free bar suppression
eddy viscosity
unperturbed eddy viscosity

O(v) component of vT
O(v) component of z-component of local velocity
':'m

(m=I,2, ... )

coefficients of the expansion for B


water density

Ps
(J

T (=.ut)

! =
Top

(Ts,T o )

(p=I,2,3)

sediment density
characteristic angular frequency of a flood wave
slow time variable in unsteady flow
bottom stress
b
ti ) components of transverse bottom stress at various
orders

TSp

(p=I,2,3)
TSO

!l,!2,'::
t

to
X
I
1 m (m=O,I,... )

t~b) components of longitudinal bottom stress at various


orders
unperturbed longitudinal bottom stress
unit vectors
bed load function
unperturbed bed load function
angle between bottom stress and

T1

direction

O(v) component of longitudinal local velocity


coefficients of the expansion for I

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Alternate Bars and Meandering

318

w
W

(prime)

c.c. or
r (subscript)

growth r ate of free bars


free bar angular frequency
nonlinear correction for the angular frequency of free bars
dimensional quantities
or s)
first derivative (with respect to
complex conjugate
real part

Abbreviations
BS
CST
ST
KM

Blondeaux and Seminara [1985]


Golombini, Seminara and Tubino [1987]
Seminara and Tubino [1985]
[(inoshita and Miwa [1974]

References
Ashida, K. and Y. Shiomi, Study on the hydraulic behavior of meanders in
channels, Disaster Prevention Research Institute Annuals, [(yoto Univ., 9,
457-477, 1966.
Blondeaux, P. and G. Seminara, Bed topography and instabilities in sinuous
channels, Proc. Am. Soc. Giv. Eng. Rivers '83, New Orleans, 747-758,
1983.
Blondeaux, P. and G. Seminara, A unified bar-bend theory of river meanders,
J. Fluid Mech. 157, 449-470, 1985.
Bray, D. I., Estimating average velocity in gravel bed rivers, J. Hydr. Div.,
Am. Soc. Giv. Eng., 105, (HY9), 1103-1122, 1979.
Brice, J. C., Stream channel stability assessment, Rpt. FHWA/RD-82/021,
Fed. Highway Admin., U.S. Dept. of Trans., Washington, D.C., 1982.
Callander, R. A., Instability and river meanders, Ph.D. Thesis, Univ.
Auckland, 1968.
Callander, R. A., River meandering, Ann. Rev. Fluid Mech., 10, 129-158,
1978.
Chang, H., D. B. Simons, and D. A. Woolhiser, Flume experiments on
alternate bar formation, J. Waterways, Harbors, Goastal Engrg. Div., Am.
Soc. Giv. Eng., 97, 155-165, 1971.
Chien, N., The present status of research on sediment transport, J. Hydr. Div.,

Am. Soc. Giv. Eng., 80, 1954.

Colombini, M., G. Seminara, and M. Tubino, Equilibrium amplitude of


alternate bars, Proc. 3rd Int. Symp. on River Sedimentation, Jackson, 1986.
Colombini, M., G. Seminara, and M. Tubino, Finite amplitude alternate bars,
J. Fluid Mech., 181, 213-232, 1987.
Dean, R. B., Aero Rpt. 74-11, Imperial College, London, 1974.
Dietrich, W. E. and Smith, J. D., Influence of the point bar on flow through
curved channels, Water Resour. Res., 19(5), 1173-1192, 1983.
Engelund, F., Flow and bed topography in channel bends, J. Hyd. Div.,
Am. Soc. Giv. Eng., 100(HY11), 1631-1648, 1974.
Engelund, F., The motion of sediment particles on an inclined bed, Tech.
Univ. Denmark ISVA Prog. Rpt. 53, 15-20, 1981.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

319

Sernillara and Tubino

Engelund, F., and E. Hansen, A Monograph on Sediment Transport in Alluvial


Streams, Copenhagen: Danish Technical Press, 1967.
Engelund, F. and O. Skovgaard, On the origin of meandering and braiding in
alluvial streams, J. Fluid A/ech., 57, 289-302, 1973.
Freds0e, J., Meandering and braiding of rivers, J. Fluid Mech., 84, 609--624,
1978.
modification of
Freds0e, J., Unsteady flow in straight alluvial streams:
individual dunes, J. Fluid Mech., 91, 497-512, 1979.
Fujita, Y. and Y. Muramoto, 1985, Studies on the process of development of
alternate bars, Bulletin of the Disaster Prevention Research Institute, 35,
55-86, 1985.
Fukuoka, S. and M. Yamasaka, Equilibrium wave height and flow on alternate
bar based on non-linear relationships of bedform, flow and seditnent
discharge, 21st IAHR Gongress, Melbourne, 1985.
Gottlieb, L., Three-dimensional flow pattern and bed topography in
meandering channels, Tech. Univ. Denmark ISVA Series Paper 11, 1976.
Guglielmini, D., Della Natura Dei Fiumi, 1697.
Hansen, E., On the formation of meanders as a stability problem, Hydr. Lab.
Tech. Univ. Denmark Basic Res., Prog. Rpt. 13, 9-13, 1967.
Hayashi, T., The formation of meanders in rivers, Proc. Japan Soc. Giv. Eng.,
180 (in Japanese), 1970.
Hickin, E. J. and G. C. Nanson, The character of channel migration on the
Beatton River Northeast British Columbia, Canada, Geol. Soc. Am. Bull.,
86, 487-494, 1975.
Hooke, J. M., Magnitude and distribution of rates of river bank erosion, Earth
Surf Processes Landforms, 5, 143-157, 1980.
Ikeda, S., Lateral bedload transport on side slopes, J. Hydr. Engrg., Am. Soc.
Giv. Eng., 108, 1369-1373, 1982a.
Ikeda, S., Prediction of alternate bar wavelength and height, Rpt. Dept.
Found. Engrg. and Gonstr. Engrg., Saitama Univ., 12, 23-45, 1982b.
Ikeda, S., M. Hino, and H. Kikkawa, Theoretical study of the free meandering
of rivers, Proc. Japan Soc. Giv. Eng., 255, 63-73 (in Japanese), 1976.
Ikeda, S. and T. Nishimura,. Bed toPOgraPhr in bends of sand-silt rivers, J.
Hydr. Engrg., Am. Soc. Gzv. Eng., 111(11 , 1397-1411, 1985.
Ikeda, S., G. Parker, and K. Sawai, Bend t eory of river meanders, Part 1 Linear development, J. Fluid AJech., 112, 363-377, 1981.
Ikeda, S., M. Yamasaka, and M. Chiyoda, Bed topography and sorting in
bends, J. Hyd. Engrg., Am. Soc. Giv. Eng., 113(2), 190-206, 1987.
Jaeggi, M., Alternierende [<iesbanke, Mitteilungen der Versuchsanstalt fur
Wasserbau, Hydrologie und Glaziologie, Zurich: E.T.H., 1983.
Jaeggi, M., Formation and effects of alternate bars, J. Hydr. Engrg., Am. Soc.
Giv. Eng., 110, 142-156, 1984.
Kalkwijk, J. P. Th. and H. J. De Vriend, Computation of the flow in shallow
river bends, J. Hydr. Res., 18(4), 327-342, 1980.
Kelvin, Lord, On the winding of rivers in alluvial plains, Proc. Roy. Soc.
London, A 25, 5-8, 1876.
Kikkawa, H., S. Ikeda, and A. Kitagawa, Flow and bed topography in curved
open channels, J. Hydr. Div., Am. Soc. Giv. Eng., 102(HY9), 1326-1342,
1976.
Kinoshita, R., Investigation of channel deformation in Ishikari River, Rpt.
Bureau of Resources, Dept. Science and Technology, Japan, 1-174, 1961.
Kinoshita, R., and H. Miwa, River channel formation which prevents
downstream translation of transverse bars, Shinsabo 94, 12-17 (in Japanese),
1974.
Kitanidis, P. K. and J. F. Kennedy, Secondary current and river-meander
formation, J. Fluid Mech., 144, 217-229, 1984.

Copyright American Geophysical Union

Water Resources Monograph

320

River Meandering

Vol. 12

Alternate Bars and 1/leandering

Langbein, W. B. and L. B. Leopold, Quasi-equilibrium states in channel


morphology, Amer. J. of Science, 262, 782-794, 1964.
Leopold, L. B. and M. G. Wolman, River channel patterns:
braided,
meandering and straight, U.S. Geol. Survey Prof. Paper 282-B, 1957.
Lewin, J., Initiation of bed forms and meanders in coarse-grained sediment,
Geol. Soc. Am. Bull., 87, 281-285, 1976.
Muramoto, Y. and Y. Fujita, The classification of meso-scale river bed
configuration and the criterion of its formation, Proc. 22nd Japanese Gonf.
on Hydraulics, Japan Soc. Giv. Eng., 275-282, 1978.
Nanson, G. C., and E. J. Hickin, Channel migration and incision on the
Beatton River, J. Hydr. Engrg., Am. Soc. Giv. Eng., 109(3), 327-337, 1983.
Nayfeh, A. H. (ed.), Perturbation Methods, New York:
J. Wiley & Sons,
1973.
Odgaard, A. J., Streambank erosion along two rivers in Iowa, Water Resour.
Res., 23(7), 1225-1236, 1987.
Olesen, K. W., Alternate bars and meandering of alluvial rivers, Commun.
Hydraul., Delft Univ. of Technology, Rpt. 7-83, 1983.
Parker, G., On the cause and characteristic scales of meandering and braiding
in rivers, J. Fluid Mech., 76, 457-480, 1976.
Parker, G., Discussion of: Lateral bed load transport on side slopes, by S.
Ikeda, J. Hydr. Engrg., Am. Soc. Giv. Eng., 110, 197-199, 1984.
Parker, G., and E. D. Andrews, Sorting of bedload sediment by flow in
meander bends, Water Resour. Res., 21(9), 1361-1373, 1985.
Parker, G., and A. W. Peterson, Bar resistance of gravel-bed streams, J.
Hydr. Div., Am. Soc. Giv. Eng., 106(HYI0), 1559-1575, 1980.
Parker, G., K. Sawai, and S. Ikeda, Bend theory of river meanders, Part 2:
Nonlinear deformation of finite-amplitude bends, J. Fluid Mech., 115, 1982.
Rozovskij, I. L., Flow of water in bends of open channels, Kiev: Acad. Sci.
Ukranian SSR, 1957.
Seminara, G. and M. Tubino, Further results on the effect of transport in
suspension on flow in weakly meandering channels, Golloquium on The
Dynamics of Alluvial Rivers, Genova, 1985.
Shen, H. W., Development of bed roughness in alluvial channels, J. Hydr.
Div., Am. Soc. Giv. Eng., 88(HY3), 45-58, 1962.
Smith, J. D., and S. R. McLean, A model for flow in meandering streams,
Water Resour. Res., 20(9), 1301-1315, 1984.
Struiksma, N., K. W. Olesen, C. Flokstra, and H. J. De Vriend, Bed
deformation in curved alluvial channels, J. Hydr. Res., 23(1), 57-79, 1986.
Sukegawa, N., Conditions for the occurrence of river meanders, Trans.
Japan Soc. Giv. Eng., 2, 257-261, 1971.
Tsujimoto, T. and H. Nakagawa, Unsteady behavior of dunes, Int. Gonf. on
Hydr. Design in Water Resour. Engrg., Southampton, 1984.
Tubino, M., and G. Seminara, Unsteady alternate bars, Atti del Seminario
Leggi morfologiche e loro verifiche di campo, Cosenza, 1987.
Tubino, M., and G. Seminara, Free-forced interactions in developing meanders.
Part 1: Suppression of free bars, submitted to J. Fluid Mech., 1988.
Van Bendegom, L., Some considerations on river morphology and river
improvement, Nat. Res. Counc. Canada, Technical Translation 1054, 1963.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Copyright 1989 by the American Geophysical Union.

Evolution and Stability of


Erodible Channel Beds
Jonathan M. Nelsont and J. Dungan Smith

Geophysics Program, University of Washington, Seattle, WA

98195

Abstract
A computational model for flow and boundary shear stress fields in natural
channels is combined with a bedload transport algorithm in order to
investigate the genesis and finite amplitude characteristics of riverine bars.
Since the time scales associated with bar growth and migration are typically
much larger than the time scales of the flow, the flow is treated in a
quasi--steady manner, and the modification of the bed topography due to
convergences and divergences in the sediment flux field is coupled to the flow
model using a simple iterative technique. The topographic evolution algorithm
yields accurate predictions of the development of point bars in curved channels
and indicates that, while the origin of point bars is primarily due to
curvature-induced secondary flow, the ultimate stability of these features is
related to topographically-induced streamwise convective accelerations, as well
as to gravitational modification of sediment fluxes by bar slopes.
The
technique presented herein is also used to investigate the mechanics of
alternate bars in straight channels. Comparison of the nonlinear theory with a
simple linear stability analysis for these features is used to demonstrate the
importance of the nonlinear effects, and to provide a clearer physical
understanding of the alternate bar instability. In contrast to point bars, the
initial instability of alternate bars is shown to depend on a simple topographic
steering response, while the finite amplitude characteristics of these features
depend on gravitational effects and the production of secondary flow associated
with the curvature of flow streamlines.
The evolution model yields good
predictions of the finite amplitude morphology of both point and alternate
bars, as demonstrated by comparison of model predictions and measured
bathymetry for several experimental studies.
Introduction
By definition, equilibrium channel morphology is attained when no net
deposition or erosion is occurring on the channel bed for a prescribed flow and
sediment discharge. This condition is reached when the amount of sediment

t Now

at U.S. Geological Survey, Water Resources Division, Lakewood, CO

80225.

321
Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

r~rodi ble

322

Channel Beds

arriving in a given length of time at each location on the bed is precisely


equal to that departing from that point or, in other words, when the
divergence of the sediment flux is zero everywhere on the bed. If the channel
geometry or flow is modified in a manner such that the flow is not in
equilibrium with the channel, erosion and deposition must occur on the bed
until stability is reinstated.
Both the mechanism by which the channel
adjusts to a perturbation from equilibrium and the time scale over which
adjustment occurs is of fundamental importance in many problems of concern
to hydrologists and geomorphologists.
In order to investigate the evolution and stability of natural channels, the
interactions between the flow field, the channel bathymetry, and the sediment
transport must be characterized in a realistic manner. These interactions may
be loosely divided into three categories.
First, there is a strong coupling
between the flow field and channel geometry due to topographic forcing of the
flow field. The spatial distributions of velocity and boundary shear stress are
strongly affected by bathymetry through the production of convective
accelerations associated with channel nonuniformity. These effects are perhaps
the most often neglected in previous work on stable channel configurations,
and yet, as is made clear below, they are of primary importance in many, if
not most, natural channels. The second type of interaction considered herein
is that between the sediment transport over a bed and the slope of that bed.
Clearly, both cross- and downstream slopes modify sediment fluxes over the
bed through the gravitational forces that act on the sediment particles as they
move. These forces are important in the determination of stable bar and bank
slopes. Lastly, and perhaps obviously, there is a strong coupling between the
flow over the bed and the transport of sediment. For the case of bedload
transport, only the boundary shear stress pattern is important, while in the
case of mixed or suspended load transport, the bottom stress, velocities, and
turbulent diffusivities all are important for predicting patterns of erosion and
deposition. Any attempt to understand the basic processes by which a channel
adjusts to its equilibrium form must include a physically-based sediment
transport model that includes the effect of bed slope, as well as a fluid
dynamical model that accurately treats the response of the flow to topographic
nonuniformities.
In this paper, a method whereby predictions of the evolution and stability
of natural channels may be made is presented. The basic idea behind the
approach presented here is shown in the form of a flow chart in Figure 1. To
predict the temporal development of channel morphology, one begins by
assuming that the flow discharge and the initial geometry of the stream are
known. Using these values, a numerical flow model is employed to obtain the
distributions of velocity and bottom shear stress in the flow through the initial
channel. Since the time scales associated with the flow are much shorter than
the time scales associated with the slow evolution of the bed topography, the
fluid dynamical model can be constructed using the assumption that the flow
is steady. However, in order to treat topographic steering of the flow by
channel nonuniformities, appropriate convective accelerations must be retained
in the model equations. This is discussed in more detail below in the context
of specific flow calculations.
The boundary shear stresses and velocities obtained from the flow model
can be used along with the particle size distribution in order to predict the
flux of suspended load and bedload. As mentioned above, the effect of the
bed slopes on the bedload sediment fluxes must be included. As shown in
Figure 1, these sediment fluxes can be used along with an expression enforcing
conservation of sediment mass in order to predict erosion and deposition on
the bed. This pattern of scour and fill can then be employed to predict how
the bed will change in form in a specified time interval. Thus, the calculated

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Nelson and Sluith

323
T=O

CHANNEL PLANFORM
DISCHARGE, GRAINSIZE AND
INITIAL BATHYMETRY

NUMERICAL MODEL
FOR FLOW IN
NATURAL CHANNELS
'tTOTAL
FORM DRAG CORRECTION

'to SKINFRICTION
SEDIMENT TRANSPORT
CALCULATION INCLUDING
GRAVITATIONAL CORRECTION

EROSION AND DEPOSITIONAL


PATTERNS ON THE BED

NO

T= T+~T
NEW BATHYMETRY
AFTER SOME SMALL
INCREMENT OF TIME PASSES

Fig. 1.

IS CHANNEL AT
QUASI-EQUILIBRIUM

Flow chart depicting the iterative technique used in predicting equilibrium bed
morphology in channels. Quasi-stability refers to the condition in which any bars
present are not changing in height, wavelength, or shape; in some cases, these
well~eveloped features migrate in a streamwise direction.

rates of erosion and deposition are employed to predict what the bed will look
The flow over this new bed
like at some small increment of time later.
geometry is calculated using the fluid dynamical model, and the entire
procedure is repeated. The procedure is continued until the flow variables
calculated over the evolved topography yield sediment fluxes for which no
erosion or deposition occurs, or until all the erosion and deposition goes into
migration of the bar forms. When this condition is met, the topography has
evolved to the equilibrium condition.
As of this writing, there have been two attempts to construct fully
nonlinear flow and bed deformation models similar to the one presented here.
The first of these is presented by Shimizu and Itakura [1985], while the second
is very briefly discussed by Struiksma et al. [1985].
Each of these two
formulations considers only the solution of the St. Venant equations, while the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

324

model presented herein uses a solution of consistently derived set of vertically


averaged equations combined with a similarity vertical structure as a
zero-order solution to the three-dimensional momentum equations. As a result
of this, these other models must employ theoretical expressions for
fully-developed bend flow to calculate the helical part of the cross-stream
velocity distribution. Thus, the near-bed veering of the shear stress vector is
treated in the models presented by these other workers using results obtained
for well-developed bend flow.
This is inappropriate and, although it may
perform adequately in meander bend calculations, clearly it does not account
for the production of helical circulation in straight channels, as discussed
below. Furthermore, these models do not provide any information about the
vertical structure of the flow, so they cannot be employed without modification
to examine any phenomena related to the variations of velocity and stress in
the vertical l e.g., suspended load transport).
Nevertheless, they clearly
represent substantial improvements over the use of linear theories for
predicting bed topography and, as such, are a step in the right direction.
The work described herein begins with a discussion of the fluid dynamical
and sediment transport models that are to be utilized.
After building this
framework, the general scheme for predicting bed evolution and equilibrium
topography described above is applied in two fundamental and relatively
well-understood cases in order to strengthen confidence in the technique.
First, the method is used to predict the evolution and stability of a point bar
and pool in a curved channel with an initially flat cross-sectional shape.
Second, this technique is shown to predict the formation and eventual stability
of alternate bars downstream of a perturbation in an initially straight channel
with a flat bed.
In conjunction with the finite amplitude, alternate bar
evolution theory, a linear stability analysis of the governing equations is
presented, both in order to clarify the physical processes responsible for
alternate bar formation, and to demonstrate the failings of the linear theory.

The Fluid Dynamical Model


The numerical flow model used in the bed evolution calculations is a
generalization of the meander flow model originally presented by Smith and
McLean [1984], and subsequently expanded and tested in Nelson and Smith
(this volume, referred to hereinafter as NSl).
In the meander model, the
technique of solution is centered on a perturbation expansion about a threedimensional zero-order velocity field comprised of a solution to an approximate
set of vertically averaged equations in conjunction with a properly scaled
similarity vertical structure function. In essence, the solution to the vertically
averaged equations in combination with a drag coefficient closure produces a
complete boundary shear stress field; the shear velocities obtained from this
stress field are then multiplied by a similarity vertical structure function to
obtain the zero-order streamwise velocity field.
Simplification of the
vertically-averaged equations arises from careful scaling arguments specifically
valid for typical natural meander bends. In order to construct a model of
general validity, the scaling constraints employed in the meander model are
altered. However, most of the physical assumptions employed in the meander
model are also used in the more general flow model developed in this paper.
These are restated briefly here:
1. Streamwise velocities are assumed to be much larger than cross-stream
velocities, which are, in turn, much greater than vertical velocities.
This assumption is valid in natural streams, except near abrupt
obstructions and vertical banks.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

325

Nelson and Smith


2.

The pressure field is assumed to be hydrostatic or, in other words,


vertical accelerations are assumed to be negligible in the vertical
momentum equation. As in the case of assumption (1), this is valid
except near abrupt obstructions in the flow or in the vicinity of steep
banks.
3. The flow is considered to be fully turbulent and the exchange of
momentum associated with the turbulent fluctuations is described using
a physically-appropriate scalar kinematic eddy viscosity.
This eddy
diffusivity for momentum depends only on the local shear velocity and
the distance from the boundary; thus, production and dissipation are
assumed to be in balance locally and any advective influences on the
turbulence field are neglected.
This is almost always a good
approximation near the bed and in the interior of the flow, where this
assumption may break down, the vertical structure of the velocity is
relatively insensitive to changes in the turbulence structure.
The
assumption, however, specifically excludes treatment of the local effects
of wakes.
4. The total roughness of the channel is assumed to be representable in
terms of a field of values for Zo, the roughness length. These values
represent momentum extraction associated with grain roughness, as well
as that due to sediment transport and form drag on channel
nonuniformities (e.g., bedforms and bars).
5. The lowest-order structure of the velocity field is assumed to be
characterized by strong shear near the boundary and relatively weak
shear (nearly constant velocity) in the region away from the boundary.
Thus, the downstream velocity is assumed to be nearly equal to the
vertically-averaged velocity over most of the flow depth. This leads to
the assumption that the vertically-averaged convective accelerations in
the downstream momentum equation are reasonable approximations to
the full convective acceleration terms.
Fortunately, near the bed,
where this assumption breaks down, the accelerations are small
compared to the stress divergence. In conjunction with (2) and (3)
above, this assumption yields the result that, to lowest order, the
velocity profile measured along the direction of a vertically-averaged
streamline is well-described by a similarity velocity profile.
6. The effects of lateral friction are neglected in the model formulation.
These terms clearly have an influence on the flow very near the banks,
but since most rivers and streams are wide relative to their depth, this
effect is not of zero-order importance in determining the overall flow
field.
Furthermore, these terms are not related to the genesis and
growth of bars, although they may have some weak effect on finite
amplitude bar morphology.
These assumptions form the foundation of the flow model employed herein
and will be adhered to henceforth.

Solution of the Routing Problem


As in a meander model, the first step in this theoretical model is the
solution of the vertically-averaged equations.
Emyloying the channel-fitted
coordinate system presented by Smith and McLean 1984] in conjunction with
assumptions (1) and (5), the full vertically-averaged equations expressing
conservation of mass and momentum (Equations 6, 8, and 9 in NS1) are
reduced to the following balances:
1

I-N

{)
os

<v>h
<u>h) - (l-N)R

{)

+ on

v>h)

Copyright American Geophysical Union

(1)

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

326
_1_

I-N

1
I-N

a (<u2 >h ) + on
a (<uv> h)
os

_ 2<uv>h _ ~ BE _ !. ( )
(l-N)R - I-N os P Tzs B

(2)

a (<uv>h ) + an
a (<v2 >h ) + (< u2(l-N)R
>-<v 2 > )h
8E
os
= -gh on
- p1 ( Tzu) B ()
3

Following the conventions enlployed in NS1, sand n are the streamwise and
cross-stream coordinates, respectively, R is the radius of curvature of the
channel centerline, and I-N is the downstream metric of the coordinate
system. Furthermore, u and v are the streamwise and cross-stream velocity
components, respectively, E is the surface elevation above some arbitrary
datum, and (Tzs)B and (Tzn)B are the streamwise and cross-stream components
of the bottom stress. Surface stresses are assumed to be zero and < > is
used to represent vertically-averaged quantities.
In order to solve these equations, it is necessary to specify a closure
between the velocity and the components of the bottom stress. Extending the
approach taken in the st:eamwise equation of the meander model, the following
closures are utilized:

p<U>2

a( Tzs)B

p<u><v>

a( TZn)B

(4)

where the assumption that the cross-stream velocities are small compared to
the streamwise velocities has been employed to simplify the closure.
The specification of a proceeds in a manner analogous to that used in the
In that model, the fact that vertically-averaged
meander flow model.
convective accelerations are good approximations to the full convective
accelerations over most of the flow depth was employed in obtaining a linear
stress profile at lowest order.
The linear stress profile along with the
definition of the eddy viscosity was used to arrive at a similarity structure for
the lowest-order streamwise velocity field. This similarity profile was shown
to depend only on the functional form chosen for the eddy viscosity and the
roughness of the bed. In the more general model developed here, the same
physical ideas are used to argue that a linear stress profile is a good
lowest-order approximation to the stress profile measured along the direction of
the vertically-averaged streamlines of the flow. In other words, the vertical
structure of the velocity field along the direction of the vector defined by
these two components of averaged velocity is well-described by a similarity
profile.
This reasoning along with the definition of the eddy viscosity
immediately leads to the following equations:

Uo

Vs

TZS) ]
(T

1/2

[(;::)B]

ft ( (,(0)
ft ( (,(0)

u*ft ( (,(0)

v*ft ( (,(0)

(5)

(6)

In these equations, the subscript on the streamwise velocity shows that,


following the conventions used in Chapter 2, this is the lowest-order
streamwise velocity. The subscript s on the cross-stream velocity component
indicates that this quantity is the similarity part of vi, the first-order cross
stream
velocity.
This
point
will
be
made
clearer
below.
Nondimensionalization of the lowest-order eddy viscosity using Ko = u*h~( ()
yields

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

327

Nelson and Smith

1((,(0)

(
1 : { d(

(7)

(0
Thus, as in the meander model, specification of a non-dimensional eddy
viscosity and the bed roughness provides the lowest-order vertical structure for
the velocity field.
Unless otherwise noted, the eddy viscosity used in the
calculations presented here is given by Equation 30 in NSl, and the drag
partitioning algorithm described in NSI is employed to calculate local
roughness lengths. Using (4), (5), and (6), it is easy to show that

(8)
and that
U* =

<u>
arrr

v* =

<v>
arrr

(9)

Equation (4) above allows the components of the boundary shear stress in
and (3) to be expressed in terms of the components of the
vertically-averaged velocity and 0'. The value of 0' proceeds directly from the
specification of a lowest-order eddy viscosity, or, equivalently, the choice of a
similarity vertical profile for the velocity.
However, in order to solve the
vertically-averaged equations, the quantities <u 2 >, <uv>, and <v 2 > must be
expressed in terms of the vertically-averaged velocities, <u> and <v>. This
is accomplished by expanding the horizontal velocities in terms of their vertical
average and the deviation from that average, and then inserting these
expansions in the various vertical averages appearing in (2) and (3). Using
primes to denoted deviations from the vertical average, the following results
are obtained:

(2)

<u><u> + <U'2>
<v 2 >

<uv>

<v><v>

<u><v> + <u'v'>

<v'2>

(10)

In the meander model, the last terms on the right-hand sides of the above
equations (the vertical correlation terms) were assumed to be small, and were
neglected in the vertically-averaged solution. As a result of neglecting these
effects, correction terms appeared in the perturbation equations for the vertical
structure (the last four terms on the left-hand side of Equation (33) in NS 1J.
This approach was justified by the measurements taken by Dietrich [1982,
which demonstrated that this approximation was reasonable in typical meander
bends.
Here, this same approximation is used in the solution of the
vertically-averaged equations, but a technique whereby higher order corrections
to this closure may be made is also described. It is worth noting here that, if
the similarity structure were exactly correct, setting the vertical correlation
terms equal to zero would be a very good approximation. For the case of a
logarithmic vertical profile and typical values for (0, the value of these terms
is typically a maximum of 5% of the value of the uncorrelated part of the
vertical averages.
However, the deviations from similarity weaken this

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

328

assumption somewhat, making a technique allowing for the inclusion of these


effects of some value.
Dropping the vertical correlation terms allows the vertically-averaged
equations to be solved in a straightforward manner. These solutions along
with the similarity vertical structure are then inserted as forcing terms in the
full (non-averaged) momentum equations in order to calculate the deviations
about the linear stress profile and the similarity velocity profile. As developed
more fully below, the variations about the similarity assumption can be used
to reformulate the vertically-averaged equations incorporating the vertical
correlation terms as forcing terms. In contrast to the original meander flow
model, this procedure effectively moves the entire flow routing problem to the
vertically-averaged equations, and does not require corrections to the local
streamwise or cross-stream discharge in the non-averaged equations.
Inserting the drag closure and separating the correlated and uncorrelated
parts of the vertical averages in (2) and (3) produces the following results:

a
I-Nos

_1_

a
on

u>2h) +

I-Nos

u><v>h)

-_

(l-N)R

BE

_1_

u><v>h) _ 2<u><v>h + F'

os -

I-N

on

<u>2

(11)

all

v>2h) _ u>2-<v>2)h
(l-N)R

-gh 8E

G'

<u><v>

on -

(12)

ali

where

G'

F'

_1_

_1_

a
os

I-N

I-N

7f8

u'2>h)

u'v'>h)

a
on

a
an

u'v'>h) _ 2<u'v'>h

( I-N) R

v'2>h)

u'2>-<v'2h
(l-N)R

In obtaining solutions of these equations, the vertical correlation forcing terms


F' and G' are initially set to zero but may be brought in iteratively. In
order to solve these equations, one also nc "'ds the integral form of the equation
expressing conservation of mass given by
n

<v>h

= l~N

-w
2"

u>h)dn

(13)

The three equations immediately above are easily solved numerically using
any of a variety of techniques. Primarily for historical reasons, the solutions
shown here are found using the same method described in detail in NSI for
the meander flow model.
In short, (12) is used to rewrite the pressure
gradient in (11) in terms of the centerline surface slope and various integral
terms. Equation (13) is then employed to eliminate the vertically-averaged
cross-stream velocity.
Using the drag closure, the final result of these
manipulations is an integro-differential equation for (Tzs)B.
This equation,

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

329

Nelson and Smith

along with an integral constraint on the streamwise discharge, is easily solved


using a simple finite difference approximation in conjunction with an iterative
scheme for determining the integral terms (see NSI for a complete description
of this approach). This technique offers some advantage over more standard
numerical techniques in that, because the iterative scheme exploits the fact
that the vertically-averaged streamwise velocity is typically substantially
greater than cross-stream velocity, it tends to be much more cost effective
than a full finite difference or finite element solution.
However, this
advantage is gained at the expense of some generality in a geometric sense.
For instance, the technique is inappropriate for the case of very rapid width
variations in the streamwise direction. It is important to point out, however,
that (11) - (13) are valid in general, and that specific solution techniques can
The only
easily be constructed for even the most complicated geometries.
inputs required to obtain a solution to this set of equations are the stream
geometry, the discharge, and the distribution of roughness lengths on the bed.
Solution of the vertically-averaged equations yields values for <u>, <v>,
(Tzs)B' (Tzn)B' and the surface elevation, E. Using the similarity velocity
structure and the linear stress profile predicts values for both velocity and
stress throughout the flow. In some sense, this lowest-order model corresponds
to pure topographic steering; the solution predicts the routing of flow over a
complex nonuniform bed, but it does not yield predictions of variations in the
vertical structure associated with spatial accelerations. In order to calculate
the deviations about the similarity structure, it is necessary to resort to the
full momentum equations.
Deviations from Similarity

Although deviations about the similarity profile may be relatively small,


these effects are often of primary importance in determining depositional
patterns. For instance, the helical flow typically observed in parts of meander
bends clearly plays a critical role in the initial growth of point bars.
By
definition, helical circulation is flow with no net cross-stream discharge, so this
important physical effect is not predicted by the solution for <u> and <v>
described in the previous section. In the meander flow model, this problem
was rectified by using a perturbation expansion about the vertically-averaged
equations. The helical component of velocity and the associated cross-stream
stress term were found from a balance of the cross-stream stress divergence,
first-order pressure gradient, the first-order averaged centrifugal force term,
and the deviation from the vertical average of the zero--order centrifugal force
term, as shown in Equation (34) in NSI. Using the more complete solution to
the vertically-averaged equations described above essentially moves the entire
flow routing problem to zero order, so the vertical average of the str.eamwise
perturbation velocity is zero. In this case, Equation (34) in NSI reduces to
the following balance:
(u 2 - <u >2)

O
0
_...,......-----,--__
_
(l-N)R

a
on + 1fi

g GEt

{)vt
oz

(14)

Adding the zero--order balance given by Equation (28) in NSI to the above
equation yields the result that, to first order, the cross-stream momentum
equation is given by

u~

OE

(l-N)R -

- g

on +

(fi K o

{)vt

oz

Copyright American Geophysical Union

(15)

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

330

Using the similarity solution for Uo and <Vt> in conjunction with the
boundary conditions that the free surface is unstressed and the velocity goes to
zero at the bed, this equation is easily solved for the cross-stream velocity and
stress (see NSl). The expression for the cross-stream velocity is of the form
(16)
In (16), the function f t is the similarity velocity profile defined in (7), and
gt is a vertical structure function describing the well-known helical part of the
cross-stream velocity, as given by the first term on the right-hand side of
Equation (37) in NSI. The integral of gt from Zo to the water surface is
zero; by definition, there is no net cross-stream discharge associated with the
helical circulation. Note that the second term on the right hand side of (16)
is simply the similarity part of the cross-stream velocity field, as defined in

(6).

There are two interesting asymptotes to (16). If no streamwise variations


occur in the topography or curvature, then from (13) it is clear that <Vt> =
0, which results in a cross-stream velocity field which is entirely helical (Le.,
there is no net cross-stream discharge).
This solution corresponds to the
classical case of well-developed bend flow, wherein streamwise variations in
topography and curvature are considered negligible [cf. Rozovskii, 1957;
Engelund, 1974]. The orientation of the velocity vector is inward (Le., toward
the center of curvature) near the bed, and is directed outward near the free
surface. The production of inward velocity and, therefore, inward boundary
shear stress, is intimately connected to the production of point bars on the
inner banks of curved bends, as discussed in detail below.
The second case of interest is that of a straight channel (R-+oo). In this
case, the vertical structure of both the downstream and cross-stream velocities
are given by the similarity profile ft.
For a typical choice of the eddy
viscosity, f1 = ~ In( (/ (0), so the direction of the velocity vector is constant in
the vertical, and its magnitude is distributed in a logarithmic manner. This
solution corresponds to the case of "pure" topographic steering, wherein the
flow is steered by downstream variations in the topography, but, in contrast to
the case of well-developed bend flow, the orientation of the velocity vector
does not vary in the vertical.
The mathematical model which led to (15) is specifically designed for
application to meander bends, and the argument which leads to this
formulation is inappropriate for the case of straight or weakly sinuous channels
with streamwise topographic nonuniformiti~s.
In this situation ("pure"
topographic steering), the inclusion of the effects of streamline curvature solely
through the specification of the channel radius of curvature omits the
production of helical flow and the associated perturbation boundary shear
stress. This is a consequence of assuming that the curvature of the channel
centerline is a good approximation to the curvature of the streamlines, a
hypothesis that is relatively accurate in the case of fully-developed meander
bends, but which is a poor approximation in straight channels with irregular
bathymetry.
As the flow is steered around the bars or other topographic
obstructions in a straight channel, the induced streamline curvature will
produce convective accelerations in the cross-stream equations that are not
included in (15). These convective accelerations produce cross-stream surface
slope and a helical components-stream velocity, just as the centrifugal force
term in (15) produces cross-stream superelevation and helical circulation in a
meander bend.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Nelson and Smith


In order to develop a more general model, which includes the salient
physical effects in both curved and straight channels with irregular bathymetry,
it is useful to examine the ordered equations which result from the
perturbation expansion employed in NSI for the situation where the radius of
curvature is infinite (a straight channel). For this case, the three lowest--{)rder
cross-stream momentum equations in the perturbation expansion are given by

0(1) =
O({):

O( {2):

=-

=-

Uo

I-N

(17)

BEt
a
an
+ Iii

+ Vt an
Ovt +

{}vt
os

BE 2
a
an
+ tfi

{}v 2

OZ

{}vt
7JZ

W2

(18)

{}v 1
IJZ-

{}vt

(19)

tfi Kt 7JZ

When R ~ 00, the O( 1) solution tells us nothing about the flow field, it only
provides the lowest-<>rder value of the cross-stream surface slope, which is
zero.
The O( () equation may be solved immediately using the boundary
conditions described above; the result is simply the similarity part of the
cross-stream velocity.
This solution is identical to the non-helical component of the cross-stream
velocity field given by (16). However, in the case of a straight channel, the
production of .a helical component of the cross-streanl velocity occurs at one
order higher in the perturbation expansion, in (19). The first two terms in
(19) make up the lowest-<>rder convective accelerations associated with the
curvature of the streamlines in the horizontal plane.
As such, they are
directly analogous to the centrifugal force term in (15), as can be verified
using either a coordinate transformation or simple formulae for the geometry of
curves in space. The third term on the left-hand side of (19) is associated
with streamline curvature in the vertical, while the right-hand side of equation
encompasses the momentum fluxes associated with the higher-<>rder pressure
gradient and stress divergence.
To include the effects of helical flow in the case of a straight channel, the
first two terms on the left-hand side of (19) are moved down one order. By
doing so, the effect of streamline curvature in a straight channel is brought
into the model in the same manner that is included in the case of the
meander bend calculations.
Note that, in doing this, the perturbation
expansion is not violated; if the meander scaling presented in NSI had been
carried out in a Cartesian coordinate system, these terms would have appeared
in both the vertically-averaged equations and (15), essentially replacing the
centrifugal acceleration terms. Employing (5) and (6), it is easy to show that
moving these cross-stream spatial accelerations to O( () yields the following
replacement for (15):
Uo

I-N

{}VS

os

Vs

Ovs

an

Uo

(l-N)R -

- g

BE + a
on
Ui

K {}vt
0

7JZ

(20)

In practice, (11), (12) and (13) are solved for the values of <uo> and <Vt>,
and then these values are employed to find Uo and Vs using (5) and (6).
Inserting these expressions in (20) and integrating employing the boundary
conditions described above yields analytical expressions for the cross-stream
velocity and stress. The equations for the velocity and stress are given by

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

332

(21)
and

(Tzn).

pu*v*(l-()

+ [~~~ ~ +

hv*

~ + (1~~)R]g2('(O)

(22)

where v* = <vl>/a1/2 and g2 is given by analogy with'Equation (38) in NSI.


In typical meander bends, the first two terms in the bracketed expression
in (21) above are much smaller than the last term. Thus, (21) is essentially
equivalent to (16) for meander bends with typical sinuosity. However, in the
case of very low sinuosity or straight channels, the first two terms in the
bracketed expression will be larger than the last term.
In this case, the
dominant part of the cross-stream velocity will be the component associated
with topographic steering, but the effect of streamline curvature in producing
helical flow is not neglected, as it would be if the meander model described in
Chapter 2 were applied without modification. As will be made clear below,
the helical circul:1tion generated by the curvature of the primary flow
streamlines plays a significant role in determining the morphology of alternate
bars in straight channels. Thus, by including the convective accelerations in
the cross-stream equation in the manner described above, a model is
formulated which includes the. salient. physical processes active in both low and
high sinuosity channels.
The above solution for the cross-stream velocity may be used along with
the similarity streamwise velocity to calculate the forcing terms originally set
to zero in the vertically-averaged equations. If these terms are calculated and
inserted in (2) and (3), the resulting solutions for the boundary shear stress
and vertically-averaged velocity will include the effects of helical redistribution
of streamwise momentum. Although this effect may only be important in few,
if any, natural streams, it has been shown to be operative in certain
specialized laboratory experiments. For example, in extremely long bends with
constant curvature and no streamwise bathymetric variations, this effect is
responsible for the tendency of the high velocity core to move toward the
outer bank as the bend is traversed. Usually, however, this weak effect is
totally overwhelmed by convective accelerations stemming from the streamwise
nonuniformity present in most natural streams.
There is also a vertical structure equation describing the deviations of the
streamwise velocity and stress from, the similarity profiles.
This equation
predicts changes in the streamwise vertical structure associated with
accelerations and decelerations, as well as with the redistribution of
downstream momentum by the helical component of cross-stream velocity.
Although this equation may be solved via simple vertical integrations, as in
the case of (20), there are important physical implications which must be
considered. By bringing in this equation, a perturbation streamwise stress is
produced.
Unlike the bottom stress that is predicted by the
vertically-averaged solution, this perturbation stress is not phase locked to the
vertically-averaged velocity field.
This allows the prediction of bedform
instabilities, as explained by Smith [1970]. Since the primary aim of this work
is to investigate the formation and evolution of bars, there is a substantial
disadvantage in adding this complicating effect to the model. Furthermore, it
is possible that the assumption of local equilibrium used in the turbulence
closure is not accurate enough to be employed in predicting the streamwise
-perturbation velocity and stress.
Thus, the equation for the perturbation

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

333

Nelson and Smith

velocity and stress in the streamwise direction is not solved here for two
reasons.
First, it - is advantageous to consider the bar problem completely
separate from the bedform problem and second, the accurate treatment of some
of the effects appearing in this equation may require a more sophisticated
turbulence closure than that employed herein.
In summary, all flow calculations presented in this paper are based on the
numerical solution of (11), (12), and (13). These solutions, which are obtained
using a prescribed channel geometry, discharge, and roughness distribution,
yield the lowest-order values for the streamwise surface slope, velocity, and
boundary shear stress, as well as the value of the vertically-averaged
cross-stream velocity. These values are inserted in (21) and (22) in order to
calculate the cross-stream velocity and stress.
Convective accelerations
associated with streamwise nonuniformities are included in the lowest-order
downstream momentum equations, as required by a consistent scaling of the
full momentum equations. The flow solutions obtained from the algorithm
described may be expected to be good approximations of the flow fields
present in natural channels.

The Sediment Transport Algorithm


The results of the flow model described above may be used to calculate
sediment fluxes associated with bedload transport and suspended load
transport. In the case of bedload transport, the particles making up the bed
travel by rolling or saltation, the sediment fluxes are calculated using a
bedload equation. This equation predicts the flux given the particle size, the
critical shear stress T c, and the skin friction boundary shear stress acting on
the bed. For the case of transport by suspension, the sediment fluxes must be
found by the numerical solution of an advective-diffusion equation employing
the diffusivities and velocities from the flow model. This requires knowledge
of the settling velocities of the particles in suspension, which may be found
using the results of Dietrich [1982]. Specification of a reference flux lower
boundary condition, wherein the upward flux of particles off the bed into
suspension is calculated as a function of the boundary shear stress, is also a
requisite part of this technique.
The solution of the advection--1iiffusion
equation will yield both the horizontal and vertical distribution of suspended
sediment, as well as the exchange of sediment with the bed (Le., the erosion
and deposition). In cases where the vertical stratification of the water column
by sediment is large enough to damp the turbulence in the flow, a
stratification correction such as that presented by Smith and McLean [19771
and subsequently tested by Gelfenbaum and Smith [1986] must be employed
and the flow and suspended sediment models must be coupled iteratively. In
the present work, the evolution of bed morphology in the case where sediment
transport occurs exclusively as bedload is examined; however, it is important
to note that the approach described is of general validity, and may be
extended to the case of mixed or suspended load transport.
The results presented in this raper are based on the use of the bedload
equation presented by Valin [1963 in conjunction with a simple algorithm for
including the effect of bed slope on the sediment fluxes.
If TSF is the
magnitude of the vector skin friction boundary shear stress on the bed, then
the sediment flux in the vector direction of the stress is given by the Valin
equation as
Qtot

0.635

[~F]

1/2

DS [l -

In(l

Copyright American Geophysical Union

/,S)]

(23)

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds


where D is the particle diameter, S is the local excess shear stress defined by
(TSF - Tc)/ Tc, and 'Y = 2.45 (p/Ps)O.4( Tc/(Ps-p)gD)o.5. The variables p and Ps
are the density of the fluid and the sediment, respectively, and are taken to
be 1.0 gm/cm 3 and 2.65 gm/cm3 .
For the case in which the downstream bottom stress is much greater than
the cross-stream value, as is consistent with the model scaling presented above
and with measurements made in natural rivers and streams, the components of
the vector sediment flux in the cross- and streamwise directions are given by
Qs

(TS) ]1/2
-----!DS

0.635 [

(Tn) ]1/2
-----!-

Qn = 0.635 [

r1 - ~

(TS):: ]1/2

[(Tn)

DS

+ "s)

In(1
1 -

In(l

]
+ "s)

(24)

(25)

where the subscripts nand s refer to the streamwise and cross-stream


direction, respectively, and (Ts)SF and (Tn)SF are the downstream and
cross-stream skin friction boundary shear stresses.
In the sense used here,
"skin friction" denotes the value of the boundary shear stress after the form
drag associated with various types of channel irregularities has been removed
(see NS1).
If the actual bottom stress values predicted by the flow model are inserted
in (24) and (25), the calculated sediment fluxes will not include the
modification of those fluxes by gravitational forces, since the bedload equation
does not take the effect of bed slope into account.
The slope effect is
expected to be relatively large in cases where the boundary shear stress is only
slightly larger than the critical shear stress for the initiation of particle motion
(Le., low transport stages).
This is due to the fact that, at these low
transport stages, sediment particles tend to roll, rather than saltate, and
therefore remain in contact with the bed most of the time [Wiberg and Smith,
19851. Conversely, we expect the gravitational correction to be relatively small
in the case of saltation (higher transport stages), wherein the direction of
travel of the moving particle is almost entirely determined by the direction of
flow. While the actual numerical error incurred by neglecting the effect of
bed slope may be negligible compared to the total sediment flux, the small
variation in the direction of the sediment flux vector on the bed can be
important in the calculation of stable channel morphologies, especially in the
case of low transport stages.
The effect of gravitational effects on bedload sediment fluxes is often
invoked in the case of well-developed bend flow. In this situation, which has
been studied in detail both experimentally and theoretically [Rozovskii, 1957;
Engelund, 1974; Odgaard, 1981], channels bends are approximated as circular
arcs and bathymetry is assumed to vary only in the cross-stream direction.
The constraints that streamwise changes in the radius of curvature and
topography are negligible are typically satisfied only near the apices of very
long meander bends, and are inappropriate in many reaches of natural
channels.
However, this relatively simple case does demonstrate the
importance of gravitational effects.
In the meander coordinate system
described above, the equation relating bedload sediment fluxes on the bed to
the rate of erosion and deposition is given by

OB _
1 [V -+] _
Of - - c
oQ B

[1

1-N

~s

OS

OQn
+ OIl
-

Copyright American Geophysical Union

Qn]

(l-N)R

(26)

Water Resources Monograph

River Meandering

Vol. 12

335

Nelson and Smith

where B denotes the elevation of the bed with respect to an arbitrary datum
and where cB is the concentration of sediment in the bed (c ~ .65). If the
B
bed is in equilibrium, the left-hand side of (26) will be zero. Furthermore,
for the case of well-developed bend flow, all derivatives with respect to swill
be zero, since there is no streamwise variation in the flow. For this case, (26)
reduces to a simple first-<>rder ordinary differential equation for Qn. Applying
the boundary condition that there be no sediment flux at the stream bank
yields the result that, if the bed of a well-developed bend is stable, Qn must
everywhere be equal to zero. However, noting that v* is zero for the case of
well-developed bend flow, it is easy to see from (22) that the cross-stream
bottom stress will always be non-zero and directed toward the inner bank.
This result appears paradoxical.
The boundary shear stress and the
resulting sediment flux clearly have a cross-stream component directed toward
the inner bank; however, the stability criterion requires that the cross-stream
sediment flux be zero. This calculation indicates that there will continuously
be a flux of sediment away from the pool side bank and toward the inner or
point bar side bank, resulting in an ever-deepening pool and a point bar
In
which grows until the critical shear stress is not exceeded at its top.
reality, the pool will deepen and the point bar grow only until the
gravitational effect associated with the sloping face of the point bar is great
The balance that has been used extensively in
enough to make Qn = o.
simple models of point bar stability in the case of well-developed bend flow
[cf. Engelund, 1974; Zimmerman and Kennedy, 1978]. It is crucial to note
that this balance is obtained only in the very special and restricted case of
well-developed bend flow; in most natural channels, the presence of streamwise
nonuniformity, troughwise flow in the lee of bedforms, and multiple grain sizes
may invalidate this result, as discussed by Dietrich [1982]. Nevertheless, the
gravitational modification of the sediment flux on the bed can be of
considerable importance in the determination of the stable channel morphology.
Simply inserting the boundary shear stresses obtained from the flow model into
the bedload equation neglects this effect and is undesirable for the work
presented here.
To include the effects of bar and bank slopes in the calculations of channel
evolution, the gravitational forces acting on sediment particles are
parameterized in terms of an equivalent gravitational shear stress vector,
This pseudo-stress is defined as a vector in the local plane of the bed directed
along the line of steepest descent. On a slope equal to the bulk angle of
repose of the sediment (<Po ~ 30), the gravitational pseudo--stress is required
Thus, even if there is no fluid
to be equal to the critical shear stress.
mechanical shear stress acting on the sedin1ent particles, motion will occur if
the slope of the bed is greater than the bulk angle of repose. Using the fact
that ;~ = 0 on a flat bed and postulating that Tg, like the gravitational force,
dependS linearly upon the sine of the angle of bed slope, one obtains

;g.

..

Tg -

0'0

sin 0'0
Tc

Arctan

VB

TVBT

(27)

(I VB I)

(28)

sln(f)o

If water surface slopes are much less than the local slopes of the bottom
topography, as is often the case in natural allu',rial streams and rivers, then B,
the bed elevation, may be replaced by h, the local flow depth, in (27) and
(28).

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

336

= T c sin a

T
G

a=ARCTAN (IVHI)

sin ~o

VH
IV HI

epo =MASS ANGLE OF REPOSE

T c = FLAT BED CRITICAL SHEAR STRESS

Fig. 2.

Schematic diagram of the gravitational pseudo-stress


gravitational correction to the sediment flux field.

vector

used

in

the

To include the gravitational pseudo-stress vector in calculations of bedload


sediment fluxes, this quantity is added in a vector sense to fluid dynamical
This is shown in
boundary shear stress predicted by the flow model.
schematic form in Figure 2. The resulting bottom stress vector is used to
By using the
calculate sediment fluxes on the bed from (24) and (25).
...
generalized vector form for T g, this model explicitly treats the effect of bed
slopes in both the streamwise and cross-stream directions. For example, if the
fluid dynamical boundary shear stress is oriented in the direction of steepest
descent on a sloped bed, the addition of ~g produces an enhanced sediment
flux in that direction, as is physically appropriate. This effect is not included
in models for making gravitational corrections to sediment fluxes which
consider only the slope of the bed perpendicular to the flow direction [e.g.,
Odgaard, 1981; Kikkawa et al., 1976; Parker, 1978]. In addition, this simple
approach only requires the specification of two relatively well-known empirical
constants, the bulk angle of repose and the critical shear stress. This model
has also been employed for investigating the migration of channel banks. If
the bank slope exceeds 30, this simple gravitational correction will predict
some small sediment flux down the bank, even if the fluid stress is zero. This
flux continues until the bank slope is decreased to the mass angle of repose.
Thus, this model has the effect of maintaining the bank slopes at about the
mass angle of repose. For cases in which stabilized vertical banks are present,
the gravitational pseudo-stress is set equal to zero, but in the case where the
banks are constructed of cohesive (but not fully stabilized) material, the value
of <Po is set equal to the observed bank angle. Thus, the cohesivity of the
banks is characterized entirely by the bulk angle of repose.
In the results 'presented here, the critical shear stress is set using the
results of Shields las shown in Wiberg and Smith, 1985] for the case of a
single ~rain size, and is specified using the theoretical model of Wiberg and
Smith l1987] for the case of mixed grain sizes, while the bulk angle of repose
is taken to be 30 for noncohesive sediment.
Although other models for
making slope corrections to sediment fluxes are available in the literature, they
are typically more complicated than the one offered here, often rely on the
determination of empirical constants which are specific to each application, and

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

337

Nelson and Smith

do not address the basic physical problem in a mechanistic manner. Clearly,


future work in this area must be aimed at developing a theoretical model for
calculating sediment fluxes due to rolling and saltation on an
arbitrarily-sloping (Le., sloping both in the direction of and perpendicular to
the fluid stress vector) bed made up of natural grains. Because results of this
nature are currently unavailable, the simple, well-defined technique discussed
above is used, rather than employing techniques with more degrees of freedom
which, although they may yield more accurate results in specific cases,
parameterize the physical processes incorrectly.
In summary, the sediment transport algorithm employed in the evolution
model is based on the calculation of streamwise and cross-stream sediment
fluxes using (24) and (25).
The components of the stress used in these
equations are calculated from the vector sum of the fluid dynamical boundary
shear stress predicted by the numerical model and a gravitational pseudostress, given by (27) and (28). The calculated sediment fluxes are inserted in
(26) in order to determine the rate of change of bed elevation. These erosion
and deposition rates predict the temporal evolution of the bed topography. By
using suitably small time increments, this calculation can be used to predict
the adjustment of the channel to an equilibrium condition, as described in the
introduction to this paper.

Evolution and Stability of Point Bars


The salient features of point bar development on a flat bed are related to
the fact that channel curvature forces both streamwise and cross-stream
convergences of sediment transport. The cross-stream sediment convergences
and divergences are initially forced by the production of helical circulation, as
has been well-described by others (e.g., Rozovskii, 1957).
The streamwise
effects are produced as a result of variations of the radius of curvature in the
downstream direction. This variation produces a change in the cross-stream
surface slope which, due to coupling through the elevation field, inevitably
forces a change in the streamwise surface slope (and boundary shear stress).
The relative importance of cross-stream divergences to streamwise
divergences of sediment flux determines the position of the point bar in a
curved reach. As the point bar grows, large spatial accelerations are produced
as a result of the topographic nonuniformity. These play a crucial role in the
finite amplitude stability. Essentially, the Bernoulli response of the flow to
the point bar produces a surface elevation rise over the upper end of the point
bar. This elevation change weakens the helical circulation, as well as the
associated inward boundary shear stress.
Thus, the point bar grows until
topographic steering effects become large enough that the flow (and sediment)
is steered around the upstream end of the point bar. This idea was discussed
in detail by Dietrich and Smith [1983].
In Figures 3 and 4, bottom topography and shear stress fields are shown
for a typical evolution run using the finite amplitude model .described above.
For the case shown, the planform geometry of the channel is given by a
sine-generated curve, which was identified by Langbein and Leopold [1966] as a
simple shape which adequately characterizes many natural channel bends. The
radius of curvature for a sine-generated curve is given by
R

[1r{l
sin
Mo

1rs] - 1

Mo

(29)

where Mo is the meander length along the channel centerline, s is the


is the angle in radians between the downvalley
streamwise distance, and
direction and the channel centerline at the crossings (where R is infinite). For

{l

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

338

o.o~
o.o~

Fig. 3.

Evolution of bottom topography for a typical point bar case.


The channel
planform is given by a sine-generated curve with n
45
the width-to-depth
ratio is 12.5, and the down-valley meander wavelength is about 10 widths. The
Froude number is 0.6 and the transport stage is about 3. The contour interval is
one fourth of the mean depth.

the evolution case shown in these figures, Mo was chosen to be twelve widths
This yields a downvalley wavelength of about ten
and n was taken as 45
for these meander bends. The width-to-depth ratio for this test case is 12.5,
the Froude number is 0.6, and the transport stage (ratio of reach-averaged
skin friction shear stress to critical shear stress) is about three.
The evolution of the point bar and pool bathymetry shown in Figure 3 is
typical of point bar formation in general.
The growth of the bar on the
initially flat bed is primarily produced by the convergence of sediment toward
the inner bank. This convergence of sediment flux arises as a result of the
production of helical circulation, as described above. The pattern of deposition
is only weakly modified by the occurrence of streamwise convergences and
divergences of sediment, which result from modification of the streamwise
pressure gradient by the varying radius of curvature.
If the depositional
feat ures were produced only as a result of the helical circulation (or more
0

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Nelson and Smith

Fig. 4.

Vol. 12

339

Evolution of the vector boundary shear stress field for the conditions given in
Fig. 3.

correctly, the cross-stream stress associated with the helical flow), the point
bar would grow most rapidly at the minimum radius of curvature, which is
located at the bend apex. However, the fact that the radius of curvature
decreases as the apex of the bend is approached from upstream and decreases
downstream of the apex results in increasing streamwise stress (erosion) near
the inner bank in the upstream half of the bend, and decreasing streamwise
boundary shear stress (deposition) near the inner bank in the downstream half
of the bend. When this pattern of erosion and deposition is superimposed on
the more dominant depositional pattern induced by the cross-stream stresses
all along the inner bank, the final result is that the point bar tends to grow
most rapidly slightly downstream of the bend apex.
In other words, the
streamwise effects tend to augment point bar deposition downstream of the
bend apex, and tend to lessen the deposition along the inner bank upstream of
the bend apex.
The skewing of the bottom stress vectors toward the inner bank is clear in
the lowest boundary shear stress map in Figure 4, which is for the initial flat
bed. These predictions also show the streamwise variations in bottom stress
discussed above, with increasing stress near the inner bank upstream of the
apex, and decreasing stress along the inner bank downstream of the bend apex.
The shear stress distribution on the flat bed also shows the weak tendency for
the high velocity core to cross the channel. In this train of bends, the high
velocity core enters each bend near the inner (point bar side) bank and slowly
crosses the stream as the bend is traversed. The high velocity core does not

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

340

actually cross the channel centerline until the downstream crossing is reached.
This is in marked contrast to the situation found in the case of equilibrium
topography.
As the point bar grows (both upward and outward), convective
accelerations and modification of the surface elevation field are inevitably
produced in response to the topographic nonuniformity this feature presents to
As described above, and in more detail by Dietrich and Smith
the flow.
[19831, the result of this modification is, in simplest terms, the "steering" of
the flow around the point bar. This steering causes the high velocity core to
cross the channel centerline further upstream and more abruptly than in the
flat-bedded case, as is clear from Figure 4. In fact, for the equilibrium case,
the high velocity core crosses the stream just downstream of the bend apex,
rather than at the downstream crossing.

Fig. 5.

Equilibrium bathymetry and boundary shear stress calculated from the model for
the conditions given in the caption in Fig. 3.

In Figure 5, expanded plots of the equilibrium bathymetry and boundary


shear stress are given along with the equilibrium sediment flux field. The
steering effect has important implications for the production of equilibrium
morphology in channel bends. In the past, models of point bar stability have
been almost exclusively based on balancing the component of inward boundary
shear stress associated with the helical circulation against a gravitational force
that is directed down the face of the point bar [e.g., Allen, 1970; Engelund,
1974; I(ikkawa et al., 1976]. These models are predicated on the existence of
"well-developed" bend flow, wherein all streamwise variations in topography
and curvature are neglected.
Unfortunately, both simple scalings and
measurements taken in natural streams indicate that this is commonly a very

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Nelson and Smith

341

poor approximation, except perhaps in very limited segments of meander


bends.
The steering effect produces outward flow over the upstream end of
the point bar; thus, both gravitational and fluid d!ag forces act to move
sediment outward in this locale. Clearly, the balance producing stability in
this area is one between the streamwise convergence of sediment flux and the
cross-stream divergence of sediment flux, rather than a purely cross-stream
balance. In other words, the steering of the high velo ~it "'t ~ore forces steering
of the path of high sediment flux, as one would expect. rlalancing circulationinduced inward boundary shear stress against a downslope gravitational force
for a single sediment size results in particle trajectories with no cross-stream
component, which is inappropriate.
In Figures 6 and 7, contours of streamwise and cross-stream velocity are
shown at five equally spaced sections through the predicted equilibrium bend.

. 50

Fig. 6.

.40

. 30

.20

. 10

.00

-. 10

-. 20

-. 30

-.40

-.50

Contours of streamwise velocity at five sections in the test bend described in the
text.
The sections are equally spaced in the bend, with the lowest plot
representing the upstream section. Velocity values are normalized by the mean
velocity, and contours are drawn at intervals of 0.2. The vertical coordinate is
nondimensionalized with the reach-averaged depth, and the cross-stream coordinate
is normalized with the width of the channel.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

342

!~i

t:

CR05S-STRERM VEL0CITY

:8:: !':-: ::'::,,:: :":",':-:":'-:':::::: :-:-:-:.-:: :.:::::

.50

.40

.30

.20

. 10

. 00

-. 10

-.20

-.30

-.40

2.5
.50

.40

.30

.20

. 10

.00

-. 10

-.20

-.30

-.40

.40

.30

.20

. 10

.00

-. 10

. '10

.30

.20

i:

iJ

2.5
.50

iJ
2.5

.50

.00

Fig. 7.

. 40

'!I"

. 30

.20

. 10

-.20

-. 10

-.30

-.20

!I'

.00

-. 50

:~~:::: :~: ~]
-.30

(II::':::: :'~' :',,


2.5
. 50

-.50

:-:-:~: :::::.~: ., : ::J

:J

'==:','=' ': I~')'


. 10

:J

-. 10

"I

-. 20

-.30

-. '10

-.50

-. '10

-.50

,I' ,

",I",
-. '10

-. 50

Contours of cross-stream velocity at five sections in the test bend. Velocities are
normalized by the mean streamwise velocity and contours are drawn at intervals of
0.05.

The sections at the bottom of these plots corresponds to the upstream


crossing, while the top section represents the downstream crossing. The inner
bank is located on the right-hand side of these sections, as is clear from the
topographic profiles. In the contours of the streamwise component of velocity,
the steering of the high velocity core is readily apparent. The highest velocity
region is near the inner bank at the upstream end of the bend, crosses the
channel relatively abruptly near the bend apex, and exits the bend near the
outer bank.
In accord with this prediction, the cross-stream flow contours
exhibit helical flow only in the deeper part of the pool region, rather than
uniformly throughout the bend.
Since the principal sediment flux balance
producing equilibrium over the point bar is between the decreasing streamwise
flux and the increasing cross-stream flux, gravitational effects are essentially
negligible. Thus, stability is obtained with a relatively flat point bar top, as
is commonly observed.
In the pool downstream of the bend apex, helical
circulation and an inward component of boundary shear stress is present, as is

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

343

Nelson and Smith

a relatively steep transverse slope. In this region, the bed stability depends
primarily on the gravitational effects, as is the case in well-developed bend
flow. However, this balance is restricted to a very small region of the bend,
and the importance of including streamwise nonuniformity should be clear.
The balance in this region is also strongly controlled by physical effects which
have been included only parametrically through the gravitational correction.
For example, the complex near-bed flow field induced by bedforms exerts a
significant influence on the channel-scale bed morphology, as described by
Dietrich [1982].
More precise predictions of channel morphology depend
crucially upon the development of physical models which specifically treat the
effects of bed slopes on sediment transport, both for the case when the bed is
locally uniform and in the case when bedforms are present.
To build confidence in the evolution model, it is necessary to reproduce
bathymetry and flow fields measured either in laboratory flumes or natural
channels for which equilibrium conditions have been obtained. The next few
paragraphs describe model predictions and their comparison to measured data
for several cases.
First, model predictions are made for the mobile bed
experiments performed by Hooke [1974, 1975].
Hooke's experiments were
performed at four different discharges in a laboratory bend. The planform of
After
the flume was given by a sine-generated curve with n = 55.
presenting and discussing these cases briefly, model predictions are shown alon~
with experimental results obtained by Whiting and Dietrich [pers. comm.J.
Like Hooke's flume, the planform of their flume is given by a sine-generated
curve.
However, in this latter case, the amplitude of meandering is very
small, with n = 10. These five different situations offer a comprehensive
test of the model and allow both its strengths and weaknesses to be examined.
In the experiments performed by Hooke r1974, 1975], a mobile-bedded
laboratory flume was used to investigate the adjustment to equilibrium
topography in a meandering channel. Experimental runs in this laboratory
bend, which is shown schematically in Figure 8, were completed at four
different discharges: 10, 20 35, and 50 liters/sec. In Figures 9 through 12,
results of the evolution model are shown for each of these runs. The flow

HOOKE FLUME

INLET
FLOW METER

PIPE

oI

2M
I

UPSTREAM
CONTROL SECTION

Fig. 8.

DOWNSTREAM
CONTROL SECTION

Schematic diagram of the flume used in the Hooke [1974, 1975] experiments. Note
the presence of the upstream control section, which may have precluded the
establishment of a "natural" upstream boundary condition.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

344

Helix Strength

....

........

"t,

""

HOOKE RUN 10 2/5


Fig. 9.

Predicted equilibrium conditions for the Hooke 10 liter/sec run. In this and the
following figures, contours of topography, helix strength, normalized bottom stress,
and normalized sediment flux are at intervals of 2 em, 5, as follows: discharge,
10 liters/sec; mean depth, 5.2 em; mean velocity, 19.2 em/sec; slope, .0021.
Topographic contours are relative to the mean depth. Flume width and sediment
size are given in all cases as 100 em and .03 em, respectively. Calculated values
of bottom stress and sediment flux are normalized using the values measured by
Hooke.

conditions for each run are given in the captions of these figures. The plots
show calculated values for the bed topography, the helix strength, the
normalized boundary shear stress, and the normalized sediment fluxes. Also
included are vector plots of the boundary shear stress and sediment flux fields.
The normalization factors employed in the contour plots of stress and sediment
flux are the average values measured by Hooke, rather than the average of the
values calculated here. Thus, these plots indicate both the spatial structure of
these fields and the overall comparison with the magnitude of the measured
values. To facilitate the comparison with Hooke's measurements, his graphs
are shown in Figures 13 through 16.
In these model runs, the overall
roughness was constrained such that the experimental water surface slope was
reproduced in the numerical solution, thus ensuring that all of the form drag
effects were taken into account.
This roughness was distributed spatially
according to the approximation (0 = constant, which is generally a good
approximation (see NSl). The form drag was removed from the predicted
total boundary shear stress using the form drag model described in NSI.
In general, the comparisons between the measured and predicted topography
are reasonable. Although there are clearly some local discrepancies in each of
the four cases, the only consistent difference between the predicted and

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

345

Nelson and Smith

Bed Topography

Helix Strength

HOOKE RUN 202/5


Fig. 10.

Predicted equilibrium conditions for the Hooke 20 liter/sec run. See Fig. 9 for
contour intervals.
The conditions for this run were as follows:
discharge 20
liters/sec; mean depth, 7.3 em; mean velocity, 27.5 em/sec; slope, .0021..

measured topography is in the location of deepest scour. All of the Hooke


results indicate that this point occurs slightly upstream of the apex of the
bend, while the model results place this point somewhat downstream of the
bend apex. This is true even when the agreement for the shape and location
of the point bar is quite good, as in the 35 lIs case.
There are two potential reasons for this discrepancy, although neither of
them is entirely conclusive.
First, the results of the model were found
assuming periodic boundary conditions and an initially flat bed in all four
cases. However, as is clear from the schematic of Hooke's flume, the upstream
condition in these bends probably were not representative of the situation
obtained if a train of bends were located upstream of the test bend. The flow
passes over a control section halfway through the bend immediately upstream
and cannot be expected to adjust to the appropriate upstream condition in
only half of a meander bend. However, a computational run performed for the
35 lIs case using a uniform upstream condition half a bend upstream of the
study bend still did not produce a pool as far upstream as that found in the
experimental case.
Further investigation was not possible without a more
detailed description of the upstream conditions present in the experimental
runs. Another possible reason for the difference in the position of maximum
scour may be that the initial bed conditions for the experimental runs were
not the same as those employed in the calculations.
Computed evolution
sequences began with an initial bed that was completely flat.
This was
probably not the case in the experiments, since Hooke reports that a rounded
radius was constructed of cement at the intersection between the bed and the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

346
HOOKE RUN 35 i/s

Bed Topography

Helix Strength

HOOKE RUN 35 ~/s


Fig. 11.

Predicted equilibrium conditions for the Hooke 35 liter/sec run. See Fig. 9 for
The conditions of this run were as follows:
discharge, 35
contour intervals.
liters;/sec; mean depth, 9.5 em; mean velocity, 36.8 em/sec; slope, .0022.

outer bank. Furthermore, he reports that this region was often exposed when
the bed reached equilibrium and, in fact, that the cement floor of the flume
was occasionally exposed in dune troughs. Thus, there may have bren some
external control on the scour depth. In the computations, the scour depth
was not allowed to exceed the depth at which the flume bottom would be
exposed, but this end result may not have been achieved in the model in the
same manner as in the flume.
Helix strength is defined as the angular difference between the near-bed
velocity vectors and the surface velocity vectors. In general, the predicted
values of helix strength are about the same in maximum amplitude as those
measured. However, there is a consistent difference in the spatial distribution
of helicity. In the measurements, the helical flow tends to be confined to the
region near the outer bank, while the calculations tend to predict helical flow
over most of the point bar. It is important to point out that a positive value
of helix strength does not imply inward flow near the bottom, since this
represents only the difference between the angular orientations of near-bed and
surface velocities. The noted discrepancy may be at least partially associated
with the near-bed flow modification by bedforms, as well as changes in
streamwise vertical structure due to spatial accelerations and decelerations,

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Nelson and Smith

Vol. 12

347

Bed Topography

Helix Strength

HOOKE RUN 50 2/5


Fig. 12.

Predicted equilibrium conditions for the Hooke 50 liter/sec run. See Fig. 9 for
The conditions for this run were as follows: discharge, 50
contour intervals.
liters/sec; mean depth, 12.8 em; mean velocity, 39.4 em/sec; slope, .0022.

Bed Topography

Fig. 13.

Experimental measurements taken by Hooke for the 10 liter/sec case. See Fig. 9
for run conditions. Boundary shear stress and sediment fluxes were not given for
this discharge. Reproduced from Hooke [1975].

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

348

Helix Strength

Bed Topography

Fig. 14.

Experimental measurements taken by Hooke for the 20 liter/sec case.


for run conditions. Reproduced from Hooke [1975].

See Fig. 10

effects that were not included in the fluid dynamical model. Credence is led
to the idea of bedforms affecting the flow by rneasurements Hooke made of
flow over a stabilized bed. The bedforms were removed from the bed before
stabilization, so their effect on the helical flow was removed. Measurements
over the smoothed, immobile bed showed higher helix strength present over a
larger portion of the bend than found in an identical case where bedforms
were present. However, some of these differences may also be associated with
the difficulty in measuring small angular deviations in the flow over the water

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Nelson and Smith

Fig. 15.

Vol. 12

349

Experimental measurements taken by Hooke for the 35 liter/sec case.


for run conditions. Reproduced from Hooke [1975].

See Fig. 11

depth.
Hooke estimated that the error in these measurements varied from
1 to 5, depending on the turbulence level. The face of the point bar
was covered with bedforms, so it may have been quite difficult to make
accurate measurements in this region due to the boils. If the measurements
were only accurate to 5, the lower values of helix strength may have been
essentially unobservable. However, this helicity may explain the location of
the scour pool in the measurements relative to the predictions, and certainly
should not be discounted. At present, this issue can probably be resolved only
experimentally, or perhaps by inclusion of the streamwise vertical structure
changes in the model.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

350

h~
,<-V

Helix Strength

C\J

Fig. 16.

Bed Topography

Experimental measurements taken by Hooke for the 50 liter/sec case.


for run conditions. Reproduced from Hooke [1975].

See Fig. 12

Overall, the agreement between the predicted bottom stress and sediment
fluxes and those measured is quite good. The most notable differences are
associated with lateral boundary layer effects. Typically, the numerical results
predict high stresses (and sediment fluxes) right up to the bank whereas the
measurements show the maxima to be slightly away from the bank. This is
not surprising, since the development presented herein neglects the momentum
exchanges associated with lateral friction. A careful treatment of the lateral
boundary layers requires both consideration of the lateral diffusion of
momentum by turbulence and the characterization of bank roughness. This
task is not part of the work presented here, although it is clear that these

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

351

Nelson and Smith


RUN 5-25

------

STUDY SECTION

~
------~
......-::=:::= w

------...~

------~--~-~
STU D Y SEC:-=------.
T 10 N---------1----I

w = 10 0
W = 25 em
2.0em
= 8W

Hinit=

Fig. 17.

1m

= 1670

=33

em 3 /see

em/sec

Fr= 0.74
D=O.06cm

Schematic view of the flume employed by Whiting and Dietrich [pers. comm.].

effects must be considered in the future if this model is to he extended to


investigate planform evolution of channels.
Although there are clearly some discrepancies which point out important
areas of research in the future, the comparison to Hooke's measurements
indicates the power of the evolution technique.
In order to explore the
predictions of the model in a somewhat different case, the model was run for
the low amplitude meander bends studied by Whiting and Dietrich [pers.
comm.l. As shown in Figure 17, the flume planform for this experiment was
given by a sine-generated curve with n = 10. Various other geometric and
dynamic parameters pertinent to Whiting and Dietrich's run S-25 are also
given in Figure 17.
In contrast to Hooke's flume, which had a bend
wavelength of about 10.5 widths, this experiment was performed with a bend
wavelength of only 8 widths. Furthermore, while the ratio of minimum radius
of curvature to width in the Hooke flume was 2.2, this parameter was 7.3 in
the Whiting and Dietrich experiment. Thus, one expects streamwise variations
in boundary shear stress and sediment flux to be larger relative to crossstream effects in the Whiting and Dietrich case than in the Hooke case.
In Figure 18, the predicted topographic evolution of run S-25 is presented.
The initial bed was taken to be flat, and, since significant bedforms were
absent, form drag effects were neglected.
Equilibrium topography was
predicted to occur after roughly one hour in this small scale experiment. As
expected from the simple scalings in the previous paragraph, the point bars
tend to form well downstream of the bend apices. In the discussion of point
bar stability above, it was noted that on a flat bed the initial effect of
cross-stream sediment convergences due to helical flow was to produce
deposition near the inner bank around the apex of the bend. In contrast, the
streamwise convergences and divergences of sediment flux associated with the
streamwise variations in the radius of curvature tended to produce erosion near
the inner bank upstream of the bend apex, and deposition downstream. The
short, low-amplitude bend used in run 8-25 is a case where the streamwise
effects dominate the cross-stream convergences.
This produces a point bar
downstream of the bend apex, forced primarily by the streamwise convergences
of sediment transport, rather than by inward boundary shear stress associated
with helical flow.
In Figure 19, the predicted equilibrium topography is shown along with the
bathymetric contours measured by Whiting and Dietrich. As in the Hooke

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Erodible Channel Beds

352

~-

Vol. 12

0.0

o.o~

--

Fig. 18.

Topographic evolution predicted theoretically for the Whiting and Dietrich case
8-25. Run conditions are given in Fig. 17. Contours are drawn at intervals of
0.2 cm. Time increases upwards in this plot and the time increment between plots
is about 8 minutes.

cases, the agreement between the predicted and measured topography is


reasonable, but there are some noteworthy discrepancies. First, the numerical
model underpredicts the steep slopes of the bars at the upstream end of the
pool regions. The measured slopes approach angles high enough to expect that
grain flow mechanisms may play a part in the sediment transport. These
slopes also are high enough that changes in vertical structure of the flow are
almost certainly important. Neither the flow or sediment transport models
employed are appropriate for these steep slopes.
The model also fails to
predict some of the small scale structure observed near the upstream end of
the pool region. This is probably a result of the grid scales employed, rather
than a failing in the model formulation. The computational grid for this case
consisted of thirteen points in the cross-stream direction and eleven sections
per bend in the streamwise direction. Thus, while the cross-stream structure
should be well-resolved, some detail in the streamwise variations may have
been neglected.
In Figure 20, various flow variables are plotted for the equilibrium case.
The boundary shear stress is normalized by the reach-averaged value of 10
dynes/cm2 Sediment fluxes are normalized by the value of 0.04 cm2 /sec
measured by Whiting and Dietrich. Local measurements of flow variables are
not presently available, so a comparison was not possible. The helix strength
results for this case are of particular interest. The largest values for the
angular deviation between the bed and the surface are found somewhat

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

353

Nelson and Slnith

RUN 5-25

. . ::::::~~;;;~~1~:::~ . :.-:::_::~~~~~~:::: _ :
CALCULATED

OBSERVED
Q = 1670 em 3 /s

W=10
h = 2 em

S =.005
T

=10

d50 = .06em

W =25em

Fig. 19.

Comparison between measured and predicted equilibrium topography for run 5-25.
Contours are drawn at 0.2 em intervals in both cases.

downstream of the bend apex. This demonstrates the importance of treating


streamline curvature in low amplitude bends, rather than just channel
curvature. As is clear from the vector bottom stress plot, the flow curvature
is actually greatest downstream of a bend apex. Topographic steering near the
bend apex essentially compensates for the channel curvature, resulting in
relatively weak curvature of the flow. As will become clearer in a subsequent
discussion, the stability in this low amplitude case is actually somewhat similar
to that found in the case of alternate bars, with the exception that the
channel curvature tends to trap the bars and prevent migration.

Evolution and Stability of Alternate Bars


In this section, the initial development and the finite amplitude growth and
stability of alternate bars are treated. In some sense, these features are the
fundamental ones related to bar instability in river channels, since they arise
spontaneously from any small perturbation to a straight, uniform channel. In
order to investigate the genesis of these forms, the first part of this section
contains a linear stability analysis of the governing equations.

Linear Stability Analysis


In the case of point bars, the initial instability of the flat bed is produced
by the curvature of the channel planform. Thus, for- that case, curvature

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

354

RUN 5-25

f1

Fig. 20.

=.

--.

~~~=:=:

I 25cm I

--

-.

Predicted flow variables for run 8-25 at equilibrium conditions. Contours intervals
for topography, normalized bottom stress, normalized sediment flux, and helix
strength are 0.2 em, 0.1, 0.2, and 5 , respectively.

essentially acts to force the observed instability. In the case of alternate bars,
this is no longer true. Alternate bars are a manifestation of a fundamental
instability in the coupled flow-sediment transport equations - no forcing is
required. By analogy with harmonic oscillations, the alternate bar instability
is the "free" response of the system, while point bars are a forced response.
Of course, these two problems are not decoupled - there is a genetic link
between the alternate bars and point bars. Examination of that link, however,

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

355

Nelson and Smith

requires the incorporation of a bank erosion calculation into the bar evolution
model, which is beyond the scope of this paper.
To investigate the physical effects responsible for the initiation of the
alternate bar instability, and to determine the initial wavelength of these
features, it is useful to employ the techniques of linear stability analysis. This
analysis is performed using the vertically-averaged equations expressing mass
and momentum conservation for the fluid in conjunction with the equation
expressing conservation of sediment mass and a bedload equation.
The
analysis presented here uses (1), (11), (12), and (26) in conjunction with the
Meyer-Peter-Mueller bedload equation modified to include the critical shear
stress, which is given by

(29)
where
'Y

8
pl/2 (Ps-p )g

In this equation, Q is the volume flux of sediment per unit width along the
direction of the vector boundary shear stress, the magnitude of which is Th.
The choice of this bedload equation rather than the Valin equation, which is
used in the evolution calculations, is due to the mathematical simplicity of the
modified Meyer-Peter-Mueller equation. This choice simplifies the formulation
of the linear stability analysis without removing any salient physical effects.
The five equations listed above are linearized about a steady uniform flow in a
straight channel using

<u>
<v>

Uo

fUt(s,n)

+ ...

= fVt(s,n) + f2v2(s,n) +
E = Eo + fEt(s,n) + .
h = ho + fht(s,n) + .

(30)

This leads to the following set of coupled linear equations:


aut
all t
h Ovt
hOOS+UOOS+
oan=O
aut
Uo OS

=-

BEt - Cd
g OS

Ovt
Uo OS

=-

(31)

[2 Uo Ut - ,Uo. h]t

BEt
an
-

ho

C Uo
d

liO Vt

1( TO-Tc)3/2] (}vt
Uo

an

Copyright American Geophysical Union

(32)
(33)

(34)

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

356

where TO = edUO and the fact that B = E - h has been used to eliminate B
from the analysis.
Note that either h or E could have been eliminated
instead, but the choice of B results in slightly simpler algebra, and the final
result is clearly unaffected by the choice made here.
2

Fig. 21.

Three-dimensional plot of the doubly harmonic bed perturbation employed in the


linear stability analysis.

To examine the growth and migration characteristics of alternate bar


perturbations of various wavelengths, the perturbation quantities are expressed
in terms of complex exponential functions, as shown in (35).
This is
equivalent to perturbing a flat bed with a small amplitude doubly harmonic
wave, as shown in the three dimensional plot in Figure 21.

(35)
where the circumflex designates a complex amplitude. Inserting these in the
system of four equations above yields an algebraic system of four equations in
four unknowns. To obtain a nontrivial solution to these equations, a condition
relating u to the drag coefficient, the streamwise wavelength of the bars, the
transport stage, and the width-to-depth ratio must be satisfied.
Before
finding this relation, it is useful to define the following nondimensional
quantities

u =-

Uo

v- =vuo

Ii
no

E
no

uo
Fr = -::r::-rrr
gUOH'"

/2

W --

cbuhp
3

,ToT

1 -

Tc

TO

(36)

Using these definitions and (35) in the system (31) through (34) gives the
algebraic system below.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

357

Nelson and Smith


1
-Cd

iO:~Fr2]

i/3/Fr2

-w

[I]

= 0

(37)

This homogeneous system only yields nontrivial solutions if the determinant of


the matrix above is zero. This condition yields the desired expression for the
growth and migration rates of the infinitesimal alternate bar perturbations as a
function of dynamic and geometric parameters.
The equation for the
nondimensional frequency is given by
( 3 0:3

w=

2 2 2
T)
i(3(/3 - 0: 2 )Cd - 3/3 CdT )

+ 0:/3

(38)

The growth rate and migration velocity of the alternate bar perturbation
are given by O"i and AO"r/27r, respectively, where the subscripts rand i refer to
the real and imaginary parts of 0". Equivalent nondimensional growth and
migration rates are given by Wi and wr/o:, respectively. Since 0: is positive
definite, the sign of the migration rate is always determined by Wr, and the
wavelength for which no migration occurs corresponds to Wr = o. Thus, it is
possible to ascertain the fastest-growing wavelength and the migration
characteristics of alternate bar perturbations simply by examining Wi and Wr.
The values of these nondimensional parameters are plotted versus Alb for a
typical case in Figure 22. As expected, a fastest-growing wavelength is found
for the alternate bar perturbations, corresponding to the peak in Wi. Since Wr

.03

Wj---

\
\

Wr---

q~ 0 ~~--,........-.. . . . . . ---,-_--,--.......;:...o..~"""-----o...._..o.....---,

10

AlB

15--20-25

Sf H = 30
Co

-.03

= .006

=(1-T"c I'b

) =.5

Fr= 0.8
Fig. 22.

Real and imaginary parts of w plotted versus the ratio of bar wavelength to width
for typical values of the width-to-depth ratio, drag coefficient, transport stage, and
Froude number. These results are predicted by the full linear analysis, including
the pertinent convective accelerations and the free surface effects.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

358

.03

Wj--W r ---

,
,
1
\

3~ 0 1--~:::::=I5-""""".....L.1O-.l.....-...L...!5-----1-

.L.-----l..-----J

2O

AlB

.--

25

__ - - - -

//"-

\.-'/ BI H =30
-03

C o =006
T =.5

Fr = 0.5

UO ~J Uo~--.O

as,

Fig. 23.

as

Results of the stability analysis for the case in which the linearized convective
accelerations in the cross and downstream momentum equations are neglected.

is positive at the selected wavelength, the fastest-growing features will travel


downstream.
The primary purpose in presenting the linear stability analysis is to
ascertain the physical processes responsible for the initial instability of
alternate bars.
To pursue this goal, it is useful to introduce "switches",
denoted SI and S2, in the equation for w. The variable SI is set equal to zero
when results are desired for the case in which the streamwise convection of
streamwise momentum (the first term in (32)) is neglected in the analysis, and
set equal to one otherwise. Similarly, S2 is set equal to zero when results of
the stability analysis are desired that do not include the streamwise convection
of cross-stream momentum (the first term in (33)), and is set equal to unity
otherwise. Using this technique, it is possible to precisely identify the roles of
the various terms in creating the observed instability. Using the definitions
for SI and S2, the equation for w becomes
(39)
If SI = S2 = 1, (39) reduces to (38), as expected. If both of these two
variables and the Froude number are set equal to zero, the expression for w is
pure real. Thus, if the rigid lid approximation is made and both streamwise
and cross-stream convective accelerations are neglected, the alternate bar
instability is removed from the analysis. In order to explore the roles of these
three effects in producing the alternate bar instability, results of the stability
analysis are presented which include each of these effects separately.
In Figure 23, results of the stability analysis are shown for the same
conditions as the results shown in Figure 22, but with the convective

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Nelson and Smith

Vol. 12

359

accelerations removed from the problem (Sl = S2 = 0).


Thus, this case
includes only the destabilizing effect of the free surface response to the bed
perturbation. An instability is found with the fastest-growing wavelength at
about four channel widths. The instability is quite weak compared to that
found in the full analysis and; because Wr < 0, the fastest-growing features
propagate upstream, rather than downstream.

.03

\
\
\

Wj--

Wr---

Ot--+-J---L.-....L....----L....----l~.....a::::_----L.-...L....---L...----J

'-

10

.3

AlB

15 --20 --25

8/H=30
Co = .006

-.03

T =( 1-rc /rb ) =.5


Uo ~, Fr---'O

as

Fig. 24.

Results of the stability analysis for the case in which free surface effects are
neglected (Fr
0) and the streamwise convection of cross-stream momentum is
neglected.

In Figure 24, the real and imaginary parts of ware shown for the case in
which the rigid lid approximation is employed and the convective acceleration
in the cross-stream equation is neglected (Sl = 1, S2 = Fr = 0). Again, an
instability is found, the fastest-growing wavelength of which occurs at about
Although this is slightly longer than the wavelength
six or seven .widths.
defined from the full analysis (Figure 22), the growth and migration curves for
this case are very similar to the full analysis and are, in fact, asymptotic to
the curves in Figure. 22 as the ratio of wavelength to width becomes large. If
the linear analysis is applied neglecting the free surface deformation,
eliminating the streamwise convective acceleration term, and including the
cross-stream term (Sl = Fr = 0, 82 = 1), the results shown in Figure 25 are
As in the other two cases, an instability is found with a
obtained.
fastest-growing wavelength near the one selected in the full analysis. The
magnitude of the growth rate in this case is much larger than that found in
the full analysis (note the change in scale on the vertical axis), indicating that
the neglected terms have a damping effect on the instability produced by the
streamwise convection of cross-stream momentum. As in the case of the full
analysis, the fastest-growing features propagate downstream.
The important conclusion from these three cases is that any of the three
effects described above is sufficient to produce an alternate bar instability, and
each selects a fastest-growing wavelength near that found in the full analysis.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

360
.10

BI H = 30
Co = .006
T =.5
Uo

hJ

as

Fr---..O

.04

3... 0

~......L...-~r---.1....-......L...-----:I:====:~

..L..-~

3
.02

AlB
Wj---

Wr ---

Fig. 25.

Results of the stability analysis for the case in which free surface effects are
neglected (Fr = 0) and the streamwise convection of downstream momentum is
neglected.

These features are not particular to the values of the parameters chosen for
these runs, rather, they are typical of the entire range of conditions found in
natural streams.
The analysis performed including only the streamwise
convection of streamwise momentum yields results most similar to the full
analysis, and clearly dominates the large response found including only the
cross-stream advection. However, the selected wavelength in the full analysis
is dependent on all three of these effects, and all must be retained in order to
make correct predictions of the wavelength of infinitesimal alternate bars.
The figures presented and discussed above provide some insight in the fluid
dynamical effects that play a role in the alternate bar instability, and this
insight can be used to provide a simple, physical understanding of the
instability. This may be obtained by considering the response of the flow to a
single perturbation located on one side of a straight channel. In other words,
one considers the flow around a symmetric bump with some specified
streamwise length and a cross-stream width of half the channel. As the flow
approaches the bump, the convective accelerations induced by the spatial
nonuniformity will produce an alteration in the pressure gradient (surface
elevation) field. This is precisely the same effect one observes upstream of
any obstacle, and is analogous to a stagnation pressure. Thus, the surface
elevation increases on the upstream side of the bump in a process which is
intuitively understood in terms of the Bernoulli response of the flow to the
obstruction presented by the bump.
This effect produces a streamwise
deceleration of the flow on the upstream side of the bump, and an
accompanying production of cross-stream flow.
The production of the
cross-stream flow is an inevitable result of the surface elevation increase on
the upstream side of the bump, which produces a cross-stream pressure

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

361

Nelson and Smith

gradient force. This reasoning reduces to the fact that the flow is "forced"
around the bump, just as flow is steered around a stick inserted vertically into
flowing water. Since the streamwise flow decelerates on the upstream side of
the bump and the cross-stream flow increases away from the bank, one
expects a convergence of sediment flux due to the streamwise flow field, and a
divergence of sediment flux due to the cross-stream flow. For relatively long
bumps, the streamwise convergences are greater than the cross-stream
divergences of sediment, and one expects deposition on the upstream side of
the bump. The same argument leads to the fact that erosion will occur on
the downstream side of the bump. This means that the longer features will
migrate upstream, in accord with Figure 22. In the case of shorter features,
the cross-stream steering is such that cross-stream divergences of sediment
outweigh the steamwise convergences on the upstream side of the bump. This
results in downstream migration, with erosion on the upstream part of the
bump and deposition on the downstream side. In either case (relatively long
or short features), the pattern of sediment fluxes is such that one expects a
transition from erosion to deposition (or vice versa) near the apex of the
bump. The location of this point relative to the top of the bump determines
whether the bump will grow or not. For very short features, the streamwise
advection of cross-stream momentum will be such that there is still significant
cross-stream velocity and sediment flux at the apex of the bump. Thus, in
this case, erosion will occur at the crest, and the bump will not grow. As the
wavelength is increased, the point of transition from erosion to deposi tion
shifts upstream. This is essentially due to the lessening of the inertial lag
between the topography and the flow. For sufficiently long wavelengths, the
result is deposition on the crest of the bump and, therefore, growth of the
bump. However, as the feature becomes even longer, the rate of growth will
begin to decrease asyrnptotically to zero, since the topographic steering effect
becomes weaker and weaker. Thus, one expects a maximum rate of growth at
some intermediate wavelength. This wavelength
is short enough that the topographic steering produces significant adjustments
in the pressure and velocity fields, but long enough that the inertial effects do
not act to shift the locus defining the transition from erosion to deposition
beyond the obstacle crest. These simple arguments are in agreement with the
results of the linear stability analysis and are obtained primarily through
insight gained from this analysis. The linear theory also provides a method
whereby the dependence of the alternate bar growth and migration on flow
and sediment transport .parameters may be investigated. In Figure 26, the
real and imaginary parts of ware plotted versus the ratio of wavelength to
width for several values of the drag coefficient, Cd. These values for the drag
coefficient are chosen to cover the range of values typically found in natural
streams and rivers. In all cases, a fastest-growing wavelength is identified,
but this wavelength tends to increase as the stream roughness decreases.
Thus, for Cd = 0.01, the wavelength of fastest growth is about four widths,
and in the case of Cd = 0.002, this wavelength is about seven widths.
Furthermore, the wavelength of fastest growth is more poorly selected as the
bed becomes smoother, as evidenced by the flattening of the curve of Wi as Cd
decreases. This analysis suggests that, in streams that are relatively smooth,
the wavelength of alternate bars may be more susceptible to alteration as a
result of finite amplitude effects or some external forcing. This point can be
further developed by examining the dependence of the fastest-growing
wavelength on the ratio of width to depth.
In Figure 27, the results of the linear analysis are shown for several values
of the width-to-depth ratio. The value of the fastest-growing wavelength is
only weakly dependent on the ratio of width to depth, with the longer features
being associated with the relatively narrow, deep streams. This wavelength

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

362

Vol. 12

Erodible Channel Beds


3.0

8/H = 30
CD = .002, .006, .01

-3.0

Fig. 26.

T= .8

Fr = 0.5

Results of the full stability analysis for three values of the drag coefficient, holding
other variables constant.

tends to be more poorly selected in streams with low width-t<Hlepth ratios.


In conjunction with the result described above, wavelength selection is expected
to be quite weak in relatively smooth streams with low ratios of width to
depth, and is expected to be much more robust in rough streams that are
wide and shallow. As further discussed below, this has important implications
for meander development.
3.0

3'-

01--+......l.----l-......L...-----l..-""'------'--..I....-.........-""---

25

B/H

=10,20,30

C o =006
-3.0

Fig. 27.

T= .8

Fr = 0.5

Results of the full stability analysis for various values of the width-to....depth ratio,
holding other parameters constant.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Nelson and Smith

363

Bf H = 30
C o =006

.03

\\

T=.2 \ \

T =.2, .5, I.

Fr= .8
\\

\ T=.5

3
5

10

AlB

15

20

25

Wj-Wr ---

-.03
Fig. 28.

Results of the stability analysis for various values of T, where T is defined as


unit.l minus the inverse of the transport stage.

The dependence of the linear results on the modified transport stage, T, is


shown in Figure 28. Clearly, the value of T has only a weak effect on the
specification of the fastest-growing wavelength, although it does have a
substantial effect on the migration rate. This has two important implications.
First, this means that the transport stage of the sediment is not very
important in calculating the infinitesimal alternate bar wavelength. Second,
this means form drag is not an important factor in the determination of the
wavelength.
In the linear analysis, the boundary shear stress available for
sediment transport was explicitly assumed to be equivalent to the total
boundary shear stress given by the drag coefficient closure. In fact, if ripples
or dunes are present on the bed, some of this friction is associated with
momentum extraction by pressure forces on the bedforms. If a form drag
correction factor relating the overall boundary shear stress to the skin friction
shear stress is included in this analysis, it appears only in the expression for
T.
Because T has only a weak effect in determining the wavelength of
instability, this complication is unnecessary. For the cases shown on Figure
28, the transport stage (defined as the ratio of boundary to critical shear
stress) varies from 1.25 to 00, with only a very small variation in the
fastest-growing wavelength. For typical streams, the variation in T with and
without the form drag correction is quite small (about 10% for the Muddy
Creek site described in NSl), so the effect on varying the selected wavelength
is negligible.
In Figure 29, the effect of Froude number variations on the determination
of alternate bar wavelengths is depicted. For all three cases, which span the
range of Froude numbers typically found in natural flows wherein bars occur,
the migration of the fastest-growing features is downstream.
The selected
wavelength increases with increasing Froude number, corresponding to the
increasing dominance of inertial effects. Interestingly, the Froude number only

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

364
.03

\\ \

~
~

Fr=O
Fr=.5

Wi - -

w,---

Fr=2.0

3: 0 I"--+-&-+---+--"---.....I...-----I-----=-~~-.L-...l....-.....J
~

10

AlB

15

20

25

81 H = 30
Fr=2.0

-.03

Co = .006
T =.5
Fr= 0.,.8,2.

Fig. 29.

Results of the linear stability analysis for various values of the Froude number.

effects the flow and pressure gradient patterns in a smoothly varying way;
there are no sudden transitions or critical values of the Froude number. This
is due to the predominance of the three dimensional nature of the flow around
alternate bars. Intuition based on the extension of two-dimensional theories
(wherein the steering of the flow is disallowed) clearly would lead to erroneous
results.
Figures 26 through 29 yield information that can be used to approximate
the fastest-growing alternate bar wavelength in most, if not all, natural
situations. However, it is of paramount importance to note that the selected
wavelength is the one associated with the infinitesimal perturbations, and may
be altered by finite amplitude effects. This has been neglected in almost all
previous treatments used to identify the wavelength of alternate bars [e.g.,
Callander, 1968; Parker, 1976]. Previous analyses assume that the wavelength
identified by the infinitesimal amplitude theory is length of the well-developed
bar forms.
This has been shown, however, to be incorrect experimentally;
observed alternate bars often change in wavelength as they evolve.
In Figure 30, predicted bar wavelengths from the linear theory are shown
along with values measured in flume eXl?eriments by Whiting and Dietrich
[pers. comm.], Fujita and Muramoto [1985J, and Fukuoka et ale [1983]. This
same relationship is shown in nondimensional form in Figure 31. As is clear
from these figures, the linear theory systematically underpredicts the finite
amplitude wavelength for these cases, tYJ?ically by 30-40%. The linear theory
presented by Blondeaux and Seminara l1984] also underpredicts the observed
wavelengths for these cases, although their analysis used a different sediment
transport relation and included a gravitational correction. In fact, using data
from a wide variety of conditions, Blondeaux and Seminara found that the
linear theory underpredicted observed bar lengths in about 75% of the
experimental cases.
Nevertheless, their computations did show that some
wavelengths are slightly overpredicted by their theory. These cases may be

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Nelson and Slnith

365

500
E

400

J:
t-

LINE OF PERFECT
AGREEMENT

w 300

reO

-.J
W

>

<r

3 200
o
W

t-

D~.fOO~

o Whiting
o Fujita and

0::

() 100
W

a..

Muramoto

Fukuoka

I'--_----'-_ _---L..._ _.....L...._ _" " ' - _ - - - - - '

100

200

300

400

MEASURED WAVELENGTH
Fig. 30.

500
(em)

Predicted versus measured values of alternate bars for the experimental results of
Fujita and Muramoto [1985], and Whiting and Dietrich [pers. comm.]. Predicted
values are found from the linear stability analysis.

for situations where the infinitesimal wavelength is strongly selected (Le., rough
streams with large width-to-depth ratios), so that the infinitesimal theory
performs adequately. Nevertheless, the cases shown in Figures 30 are poorly
treated by the linear theory, and the experimental observations described by
Fujita and Muramoto [1985] support this conclusion.
12

LINE OF PERFECT
AGREEMENT

:)

w
~

:)

-t~~~

0::

0-

CD

"

o Whiting

,<

o Fujita
~

and Muramoto
Fukuoka

12

AlB MEASURED
Fig. 31.

Predicted versus measured values of wavelength over width for various flume
experiments.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

366

Run H-2-1
t= 7'45"

Y>0XS // /Z

45 1 53 11

0.45

:>';
1

0.15

0.70

y:

1.65

120
_~:x_)<~

300

,/

3.15

0.85

1.00

10
X(m)

3.20

3.05

I!!

0.80

2.80

2.90

3.20
2.60

0.30 0.75

0.80
1.80

1.40

1.75

0.35 0.50

:sz'SZ:"[

2.35

1.70

><::~
I

0.70
0.80

2.80

0.95

200

.Z<~

0.50 0.40

1.05

<:~'\

0.35

wave height of bar (em)

0.35

76 56"

7\~

3.35
2.30
I

15

0.85
,

20

After Y. Fujita and Y. Muramoto

Fig. 32.

Sketch of. alternate bar evolution in an initially flat-bedded channel. Reproduced


from the work of Fujita and Muramoto [1985]. See Figure 35 for the experimental
run conditions for run H-2. Numbers next to the banks represent bar heights
in em.

In Figure 32, the temporal evolution of a series of alternate bars from an


ini tially flat bed is shown for one of the cases studied by Fujita and
Muramoto [1985].
This sketch very clearly shows the alteration of the
wavelength of the bars as they grow to their well-developed state. The initial
wavelength of the evolving bars is predicted very well by the infinitesimal
theory, as one would expect. However, finite amplitude effects produce an
alteration in this wavelength. This effect is not an isolated observation. In
Figure 33, the wavelength evolution of alternate bars is shown for all the
flume studies performed by Fujita and Muramoto.
In all cases, the
wavelength increases substantially as the bars evolve. Thus, it is not at all
surprising that the linear theory predicts poorly the well-developed wavelengths
of the bars.
This observation also has important implications for the the
formation of meander bends.
In their excellent treatment of the bar-bend instability problem, Blondeaux
and Seminara rlg85] demonstrate that the alternate bar linear analysis
substantially underpredicts the wavelengths of meanders.
If true as stated,
this would be surprising, since meander bends are typically associated
genetically with the formation of alternate bars.
However, in light of the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

367

Nelson and Snlith

1.2

o
e

()

1.0

0.8 o
CD

c::><

"m 0.6

,I'

CD

<I

<D

e e

0.4
0.2

<I.
e

CD

0-------1""'-------1.------'-------'-----1
0.2 0.4 0.6
0.8
1.0
TIT e
After Y. Fujita and Y Muramoto

Fig. 33

Bar wavelength evolution measured by


experimental investigation. Each symbol
run. T and lb denote elapsed time and
subscript E refers to these variables in the

Fujita and Muramoto [1985] in their


corresponds to a different experimental
bar wavelength, respectively, while the
equilibrium case.

ideas discussed above, this may be explained as follows.


Meanders are
expected to form .at the wavelength of well~eveloped alternate bars. If the
wavelength of the bars increases as they grow to their equilibrium morphology,
there is no reason to expect that the linear theory should identify the meander
wavelength. Furthermore, the results of the linear theory show that the small
amplitude wavelength is very poorly selected in smooth streams (low Cd) with
These characteristics are typical of
relatively small width-to-depth ratios.
meandering streams. Thus, the linear theory is expected only to identify the
wavelength for small amplitude features, while the meander wavelength is set
One possible
by the finite amplitude wavelength of the alternate bars.
exception to this case is in a situation where the bed is initially flat, but
there is some bank irregularity. The bank irregularity is fixed in a spatial
sense, and therefore must force stationary bars. As is clear from the stability
analysis, stationary alternate bars are still growing features (see Figure 22),
although they amplify at a rate less than that of the fastest-growing features.
Colombini, Seminara and Tubino [1986] constructed a weakly nonlinear
theory of alternate bar development. While this theory seems to predict bar
heights well, one of the conclusions of their analysis is that the adjustment to
the wavelength found in the linear theory due to the weakly nonlinear effects
is negligible. However, this is definitely not in accord with the experimental
evidence, and indicates the need for a fully nonlinear model, such as the one
discussed below.
Although the presentation of the linear theory here has been a lengthy
digression, it leads directly back to the central subject of this paper: the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

368

finite amplitude model. Clearly, the infinitesimal amplitude treatment yields


important insight into the development of alternate bars.
However, this
theory yields only uncertain values of the bar wavelength, and little or no
information about the finite amplitude shapes, migration rates, and heights of
the bars. Since these quantities are precisely what one typically wants to
know for natural channels, the need for a finite amplitude bar evolution theory
is clear.

Finite Amplitude Evolution


In this section, results of the finite amplitude evolution model are presented
for the case of alternate bars for two cases studied by Fujita and Muramoto
[19851. In the first case, denoted by Fujita and Muramoto as run C-2} the
calculations concentrate on examining the evolution of the wavelength of the
bar forms. The second calculation, which is for Fujita and Muramoto's run
H-2, examines the full evolution problem, concentrating on the processes
producing the stable morphology of the migrating features.
In Figure 34, the predicted evolution of run C-2 is shown. For this case,
the initial bed of the straight channel was taken to be flat, with the exception
RUNC-2
BED TOPOGRAPHY

,40cm,

J
T=O

T= 23min.

T=39min.

T= 1hr.18min.

~,;;;:;:gj::::~,_:,:~H:~~~
T= 2 hrs. 36 min.

~~~:'::':"-"'-;;'~&~~
Fig. 34.

T= 5hrs. 51 min.
Evolution sequence for flume experiment C-2, performed by Fujita and Muramoto
[1985]. The conditions for this run were as follows: channel width, 40 cm; mean
depth, 1.26 em; discharge, 1.95 liters/sec; slope, 0.0093; sediment size, 0.1 em.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

369

Nelson and Smith

of a small perturbation introduced at the cross-section furthest upstream.


Subsequently, the computational grid was moved such that the initial point of
perturbation remained at the upstream end of the computational grid. This
effectively removes the migration of the bars from the problem, thereby
allowing more precise observation of the alteration of bar wavelength as time
passes. The experimental conditions for run C-2 are given in the caption of
Figure 34.
For these conditions, the linear analysis predicts a fastest-growing
wavelength of about five widths. Fujita and Muramoto [1985] report values of
bar wavelength averaged over the entire inspection reach, as well as the value
of wavelength found over what they call the "averaged" reach. The value
found in the averaged reach apparently is the mean wavelength of a smaller
number of bars well downstream of the flume entrance. Even in the case of
the mean wavelength averaged in the inspection reach, the region near the
upstream end of the flume was not considered.
For run C-2, the initial
half-wavelengths reported in the evolution process were 107 and 126 cm in the
inspection reach and the averaged reach, respectively. The flume was 40 cm
wide, so these measurements yield initial wavelength-to-width ratios of 5.3 and
6.3, which are in reasonable agreement with the results of the infinitesimal
analysis. These values also are in good agreement with the results of the
~played in Figure 34, from which a value of 5.5 widths is found for
the earliest case in which a wavelength can be defined.
The tendency for the bar wavelength to increase with time is clearly shown
in Figure 34. At equilibrium, the wavelength is about 7.5 widths, an increase
of almost 40% over its initial value. After one hour and six minutes elapsed,
Fujita and Muramoto report half-wavelengths in the inspection and averaged
reaches of 177 and 206 cm, respectively, corresponding to wavelength-to-width
ratios of 8.8 and 10.3. After two hours and forty-two minutes, they report
half-wavelengths in both the inspection and averaged reach of 145 cm,
corresponding to a wavelength over width ratio of 7.25. This latter value is
very close to the equilibrium wavelength observed in the calculations shown in
The calculations do not reproduce the overshooting of the
Figure 34.
wavelength observed by Fujita and Muramoto [1985], and the physical reason
for this behavior is not known at present. However, at least in this case, it
is clear that there is a marked increase in wavelength from the initial
instability to the well-developed bar form, and this increase is predicted by
the finite amplitude model presented herein.
The time required to reach equilibrium in the numerical computations was
about three times larger than that required in the experiments, although the
predicted sediment fluxes were quite similar. The primary difference seems to
In the experiments, bar trains were
lie in the initiation of bar growth.
distinguishable on the flume bed only a few minutes after the beginning of the
run. In the numerical experiments, it took well over an hour before a train of
bars was observed. This difference is probably due to the nature of the initial
In the numerical model, a very small depth
perturbation to the system.
perturbation is introduced in order to start the evolution process. Typically, a
bump with a height equal to about 10% of the mean depth is introduced at
one cross-stream section over approximately 20% of the width. This is a very
small perturbation on the uniform flow situation, so the induced erosion and
deposition rates are extremely small.
In the experimental case, there are
many sources of initial perturbation: small-scale bedforms, "patchiness" in the
sediment transport field, and even waves on the free surface. The lack of all
these random perturbations in the numerical calculations result in a much
longer "spin-up" time for the system.
Some of the overprediction of time to reach equilibrium may also be due to
the fact that vertical structure changes are neglected in the streamwise velocity

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

370

field.
The omISSIon of these vertical structure changes results in an
overprediction of the throughput of sediment at each bar front compared to
sediment fluxes going into bar front migration. This occurs because the model
underpredicts the rate of decrease of the boundary shear stress on a
downstream-sloping face.
This means that the rate of deposition on these
faces will be underpredicted, and that, as a result, the rates of migration and
growth of the bars will be underpredicted.
In Figure 35 predicted bed evolution is shown for run H-2 from Fujita and
Muramoto [1985]. In this case, the computational grid is held fixed relative to
the bed, so migration of the bars can be observed. As discussed briefly above,
the calculation is started with a very small perturbation. If this is not done,
the numerical solution will consist simply of the maintenance of the uniform
flow and bed topography. The perturbation in the case shown was located in
the center of the streamwise reach near the lower bank. The depth change
was about 10% of the mean flow depth, and the cross-stream and streamwise
scale of the perturbation were 0.2 widths and 1.0 widths, respectively. Other
weak departures from uniform flow were also tested as initial conditions,
including small perturbations to the surface elevation field and the upstream
stress boundary condition, and the equilibrium topography was found to be
independent of the form of the initial perturbation, provided that the
introduced departure from uniformity was small.
The instability that leads to the formation of alternate bars has been
discussed in depth in the section on the linear theory, and is related primarily
to topographic steering of the flow around a bump. As the bars grow in
amplitude, finite amplitude effects not included in the linear analysis become
important and are, in fact, crucial to the establishment of equilibrium bar
morphology. The features present in the finite amplitude evolution model that
distinguish it from the linear analysis presented above are (1) the gravitational
correction, (2) the production of helical flow due to streamline curvature, and
(3) the full nonlinearity of the equations governing the flow and sediment
transport.
Each of these three is important in calculating equilibrium bar
shapes.
The lengthening of the bar wavelength discussed above, which is also easily
observed in Figure 35, is related to the nonlinearity of the convective
accelerations. In the discussion of the infinitesimal analysis, the tendency for
relatively short bars to grow slower or decay was related to the development
of an inertial lag between the bar and the flow due to the streamwise
advection of momentum.
For finite amplitude bars, the linear theory
underpredicts this inertial lag, because it underpredicts the magnitude of the
convective accelerations relative to the zero-order pressure gradient and stress.
Thus, the wavelength of the fastest-growing wave must increase as the degree
of nonlinearity of the momentum flux terms increases.
In Figure 35, the tendency for the bars to form diagonal fronts across the
This behavior is also shown in Figure 36.
In this
channel is clear.
three-dimensional depiction of the equilibrium topography, both the tendency
for diagonal front formation and a steepening of the bar fronts is clear.
Initially, the scour pools tend to be roughly symmetrical, in reasonable
agreement with the topographic form assumed in the linear theory. However,
as the bars grow, the downstream faces of the diagonal fronts become much
steeper than their upstream surfaces. The steepening of the lee sides of the
alternate bars is brought about in precisely the same manner as steepening of
bedform lee faces.
The nonlinearity in the stress and sediment transport
relationships produce more rapid migration of the bar tops as they grow to
finite amplitude.
This produces deformation of the originally symmetrical
shape, as found by Exner [reported by Leliavsky, 1955, p. 24]. In the case of
two-dimensional bedforms, this process continues until flow separation and

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

371

Nelson and Smith

"'==0.0

Fig. 35.

Vol. 12

--J

aa
~.~

Model predictions of the evolution of bed topography for flume experiment H-2 of
Fujita and Muramoto [1985]. The conditions for this run were as follows: channel
width, 50 cm; mean depth, 2.11 cm; discharge, 4.02 liters/sec; slope, .0056;
sediment size, 0.1 cm. The time interval between each plot is about one hour.

grain flow mechanisms become important. However, in the case of alternate


bars, the flow is not directed perpendicularly to the diagonal bar front, and
other mechanism are brought into play, like the production of helical flow.
The pattern of flow through a train of alternate bars is fairly well
described by the meandering of the high velocity core from scour pool to scour
pool. As the bars grow and the pools deepen, this meandering becomes more
pronounced.
This produces secondary circl:lation, as discussed in the
development of the flow model. Streamline curvature in the horizontal plane
produces a helical component of velocity perpendicular to that streamline, as
well as an associated perturbation bottom stress. In Figure 37, an expanded
view of the equilibrium bed contours are shown along with equilibrium bottom
stress and helix strength results.
In the bottom stress vector plot, the
tendency of the flow to meander through the bars is evident. The production
of secondary flow due to this meandering is demonstrated in the contours of
helix strength.
The secondary flow tends to produce a cross-stream stress
component directed up the face of the bars. This opposes the tendency of
both the gravitational and topographic steering effects to force sediment fluxes
into the pool region. If the secondary circulation terms are suppressed in the
model, much greater flow curvatures occur in conjunction with much more

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

372

Fig. 36.

Erodible Channel Beds

plot of the equilibrium topography for run B-2. Note that the
wid th has been exaggerated relative to the wavelength in order to make the
thre~imensional structure of the bar easier to discern.

Three~imensional

BED T0P0GRRPHY

80TT0M STRESS

HELIX STRENGTH
~~-~;~~-~;~

---------~

:::::: __ -----

C--7.\)-

o
Fig. 37.

~
0

Equilibrium values of bathymetry, bottom stress, and helix strength predicted by


the evolution model for run H-2. Contours for bottom topography are at intervals
of 0.5 em, while helix strength contours are shown at 10 intervals. The mean
value of the boundary shear stress is about 11 dynes/cm2 .

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

373

Nelson and Slnith

concentrated jets of high velocity and sediment flux.


In other words, the
secondary circulation acts to counteract the tendency for the flow to meander
more and more strongly as the bars grow larger.
This point can be made clearer by analogy with point bar evolution and
stability. In the point bar case, helical flow due to the curvature forced by
the channel geometry causes the initial instability.
As the bar grows,
topographic steering effects act to oppose the tendency of the helical flow to
cause erosion near the outer bank and deposition near the inner one. The
final stability requires the inclusion of gravitational effects, topographic steering
effects, and the secondary flow.
In the case of alternate bars, the same
sensitive balance eventually is obtained, but in a slightly different sequence.
The initial instability is produced primarily by topographic steering but, as the
bars grow, secondary circulation effects become important in determining the
sediment transport fluxes on the bed.
In this case, curvature of the
streamlines due to the topographic steering produces the helical flow, rather
than the channel curvature.
The prediction of helical flow adjacent to
alternate bars also agrees with the experimental measurements made by
Leopold [1982], who observed both the presence of helical circulation and the
occurrence of regions wherein the helical flow was in the opposite direction on
either side of the channel centerline. This behavior is predicted to occur just
downstream of the bar top, as shown in the central portion of the helix
strength plot in Figure 37.
As in the case of run C-2, the fully developed wavelength of the alternate
bars is well-predicted by the model.
Fujita and Muramoto report an
equilibrium bar half-wavelength of 213 em. Since the flume width for run
H-2 was 50 cm, this yields a ratio of wavelength to width ratio of about 8.5.
The value predicted by the evolution model is 8.2. The linear theory predicts
that the fastest-growing wave has a ratio of wavelength to width of five,
demonstrating the importance of the nonlinear effects.
The sketch of run H-2 presented by Fujita and Muramoto [1985] and
reproduced here as Figure 35 indicates that the maximum amplitude of the
observed bars (defined as the difference between the topographic minimum and
maximum on any single bar) was about three cm, in good agreement with the
heights predicted by the model. Their sketch also indicates that the rate of
bar migration was about 0.1 cm/sec. In contrast, the numerical calculation
predicts a. migration rate of about 0.03 em/sec. This discrepancy is almost
certainly due to the inability of the sediment transport and flow models to
deal with the processes important on the steep downstream faces of the
bedforms. As already discussed above, better treatment of this region requires
both the inclusion of streamwise vertical structure changes in the model, and a
more rigorous treatment of the effects of streamwise bed slopes in the sediment
transport model.
Summary of the Bar Evolution Model
In this paper, coupling of a fully nonlinear flow model with sediment
transport calculations has been described.
The resulting evolution model
reproduces the growth and stability of point bars in curved bends, as well as
the growth of alternate bars in straight channels. The insight gained from
examining results for these two fundamental bar types is easily generalized to
more complex bar forms. In fact, although the results presented herein have
concentrated on these two simple cases, the mathematical formulation is quite
general. Thus, this model can be employed to investigate bar behavior in
channels with complicated planform geometries.
Importantly, the method
presented allows calculation of both equilibrium topography and the temporal

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

374

changes that occur in cases where the flow and the bed are not in equilibrium.
Thus, this technique can also be used to investigate the behavior of complex
bars for which discharge variations play a role [Andrews and Nelson, this
volume].

Notation
b

channel width in linear analysis

elevation of channel bed above arbitrary datum

Cb

concentration of sediment in the bed

Cd

drag coefficient

grain diameter

elevation of water surface above arbitrary datum

f1

vertical structure function for streamwise velocity

Fr

Froude number

g1

vertical structure function describing curvature-induced

g2

vertical structure function describing curvature-induced

part of cross-stream velocity


part of cross-stream bottom stress
g

gravitational acceleration

local flow depth

von Karman's coefficient (= 0.40)

eddy viscosity

Mo

meander length along centerline

cross-stream coordinate

cross-stream coordinate divided by the centerline radius


of curvature

Q
R

sediment discharge per unit width


centerline radius of curvature
streamwise coordinate

S1

switch (= 0 or 1) used in linear analysis

S2

switch (= 0 or 1) used in linear analysis

excess shear stress

time coordinate

unity minus the transport stage

u*
u

shear velocity
streamwise velocity

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Nelson and Smith

375

cross-stream velocity

W
z

vertical velocity
channel width
vertical coordinate

Zo

roughness length

G'

closure parameter between vertically-averaged velocity

on
G',{3, I

and bottom stress


magnitude of bed slope along path of steepest descent
nondimensional constants used in linear stability analysis

(
(0

(see (36))
empirical constant in Meyer-Peter-Mueller bedload
equation
perturbation parameter, 0(.1)
vertical coordinate nondimensionalized by local depth
nondimensional roughness length

'"
A
p

nondimensional eddy viscosity


wavelength
fluid density

Ps

sediment density
frequency in linear analysis

I
t

(J

stress
critical shear stress for the initiation of sediment motion
bulk angle of repose
angle between sine-generated meander and average

streamwise direction at crossing


nondimensional frequency in linear analysis

T
Tc

<Po

Vol. 12

References
Ascanio, M. F., and J. F. Kennedy, Flow in alluvial river curves, J. Fluid
Mech., 133(1), 1-16, 1983.
Blondeaux, P., and G. Seminara, A unified bar-bend theory of river meanders,
J. Fluid Mech., 157, 449-470, 1985.
Callander, R. A., Instability and river channels, J. Fluid Mech., 36, 465-480,
1969.
Colombini, M., G. Seminara and M. Tubino, Finite-amplitude alternate bars,
J. Fluid Mech., 181, 213-232, 1987.
De Vriend, H. J., A mathematical model of steady flow in curved shallo\v
channels, J. Hyd. Res., 15(1), 37-54, 1977.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Erodible Channel Beds

376

Dietrich, W. E., Flow, boundary shear stress, and sediment transport in a


river meander, Ph.D. dissertation, 261 pp., Univ. of Wash., Seattle, 1982.
Dietrich, W. E., and J. D. Smith, Influence of the point bar on flow through
curved channels, Water Resour. Res., 19(5), 1173-1192, 1983.
Dietrich, W. E., and J. D. Smith, Bedload transport in a river meander,
Water Resour. Res., 20(10), 1355-1380, 1984.
Engelund, F., Flow and bed topography in channel bends, J. Hydraul. Div.
Am. Soc. Civ. Eng., 100(HYll), 1631-1648, 1974.
Fujita, Y., and Y. Muramoto, Studies on the process of development of
alternate bars, Bull. Disast. Prevo Res. Inst., Kyoto Univ., 35(3), 55-86,
1985.
Fukuoka, S., K. Uchijima, M. Yamasaka, and H. Hayakawa, Distribution of
sediment transport rate over alternate bars, Proc. of the 30th Japanese

Conference on Hydraulics, 1983.

Gelfenbaum, G. R., and J. D. Smith, Experimental evaluation of a generalized


suspended-sediment theory, in: Knight, R. J., and J. R. McLean (Eds.),
Shelf Sands and Sandstones, Canadian Society of Petroleum Geologists,
Memoir 2, p. 133-144, 1986.
Hooke, R. L., Shear-stress and sediment distribution in a meander bend,
Uppsala Univ. Naturgeografiska Inst. Rapport 30, 1974.
Hooke, R. L. Distribution of sediment transport and shear stresses in a
meander bend, J. Geol., 83, 543-565, 1975.
Ikeda, S., Prediction of alternate bar wavelength and height, J. Hyd. Eng.,
11 O( 4), 371-386, 1984.
Kikkawa, H., S. Ikeda, and A. Kitagawa, Flow and bed topography in curved
open channels, J. Hyd. Div. Am. Soc. Giv. Eng., 102(HY9), 1327-1342,
1976.
Langbein, W. B. and L. B. Leopold, River meanders - Theory of minimum
variance, U.S. Geol. Surv. Prof. Pap., 422-H, H1-H15, 1966.
Leliavsky, S., An Introduction to Fluvial Hydraulics, Constable, London, 1955.
Leopold, L. B., Water surface topography in river channels and implications
for meander development, in Gravel-Bed Rivers, ed. R. D. Hey, J. C.
Bathurst, and C. R. Thorne, John Wiley, London, 1982.
McLean, S. R. and J. D. Smith, A model for flow over two-dimensional bed
forms, J. Hyd. Eng. Am. Soc. Civ. Eng., 112, 300-317, 1986.
Nelson, J. M., Mechanics of flow and sediment transport over nonuniform
erodible beds, Ph.D. dissertation, Univ. of Washington, Seattle, WA, 1988.
Odgaard, A. J., Transverse bed slope in alluvial channel bends, J. Hyd. Div.
Am. Soc. Giv. Eng., 107(HYI2), 1677-1694, 1981.
Onishi, Y., Effects of meandering on sediment discharges and friction factors
of alluvial streams, Ph.D. dissertation, 158 pp., Univ. of Iowa, Iowa City,
1972.
Parker, G., On the cause and characteristic scales of meandering and braiding
in rivers, J. Fluid Mech., 76, 457-480, 1976.
Rowvskii, I. L., Flow of Water in Bends of Open Channels, Israel Program for
Scientific Translation, originally published by Academy of Sciences of the
Ukraine, SSR, 233 pp., 1957.
Schlicting, H., Boundary Layer Theory, McGraw-Hill, New York, 1979.
Shimizu, Y., and T. Itakura, Practical computation of two-dimensional flow
and bed deformation in alluvial streams, Civil Engineering Research
Institute Report, Hokkaido Development Bureau, Sapporo, 1985.
Smith, J. D., Modeling of sediment transport on continental shelves, in
Goldberg, E. D., ed., The Sea: Ideas and Observations on Progress in the
Study 0/ the Sea, Wiley and Sons, New York, 1977.
Smith, J. D., and S. R. McLean, Spatially averaged flow over a wavy surface,
J. Geophysical Res., 82, 1735-1746, 1977.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

377

Nelson and Smith

Smith, J. D., and S. R. McLean, A model for flow in meandering streams,


Water Resour. Res., 20(9), 1301-1315, 1984.
Struiksma, N., K. W. Olesen, C. Flokstra, and H. J. de Vriend, Bed
deformation in curved alluvial channels, J. Hyd. Res., 23(1), 57-77, 1985.
Wiberg, P. L. and J. D. Smith, A theoretical model for saltating grains in
water, J. Geophysical Res., 90(4), 7341-7354, 1985.
Wiberg, P. L., and J. D. Smith, Calculations of the critical shear stress for
motion of uniform and heterogeneous sediments, Water Resour. Res., 23(8),
1471-1480, 1987.
Valin, M. S., An expression for bedload transportation, J. Hyd. Div. ASGE,
89{HY3), 221-250, 1963.
Valin, M. S., Mechanics of Sediment Transport, Pergamon, 1977.
Yen, C., Bed configuration and characteristics of subcritical flow in a
meandering channel, Ph.D. Dissertation, 123 pp., Univ. of Iowa, Iowa City,
1967.
Yen, C., and B. C. Yen, Water surface configuration in channel bends, J.
Hyd. Div., Am. Soc. Giv. Eng., 97(HY2), 303-321, 1971.
Zimmerman, C., and J. F. Kennedy, Transverse bed slopes in curved alluvial
streams, J. Hyd. Div., Am. Soc. Giv. Eng., 104(HYl), 33-48, 1978.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Copyright 1989 by the American Geophysical Union.

Observations on Several Recent Theoriffi of Resonance and


Overdeepening In Meandering Channels
Gary Parker and Helgi Johannesson

St. Anthony Falls Hydraulic Laboratory


Department of Civil and Mineral Engineering
University 01 Minnesota, Minneapolis
Three recent linear treatments of alternate bars and flow in meandering
channels are compared in terms of a unified framework. This framework is
used to clarify the relation between such self--excited topographic effects as
alternate bars and curvature-driven phenomena such as point bars.
The
equation of depth-averaged transverse momentum balance, which may be
driven by topography or curvature, is decoupled from the equation governing
All three
secondary flow, which is driven essentially by curvature alone.
theories are found to yield similar results for the phenomena of resonance and
overdeepening.
Introduction
Recent progress in the study of meandering has led to new insight into the
interaction between oscillations in flow and bed topography driven by channel
curvature, and those arising freely as alternate bars in straight channels. The
coexistence and interaction of point bars and alternate bars in channel bends,
alluded to by Engelund [1975], has been confirmed. In particular, Blondeaux
and Seminara [1985] have identified a resonant condition, according to which
curvature can force an erodible-bed channel at its "natural" wavenumber
(associated with alternate bars), resulting in pronounced scour and fill.
Struiksma et ale [1985] have likewise identified an oscillatory "overshoot," or
overdeepening effect, according to which pronounced outside scour may be
observed at the entrance of a bend contiguous with a straight reach upstream.
Resonance and overdeepening proved to be major topics of discussion of the
final workshop of the Joint U.S.-Japan Seminar on River Meandering, October
20-22, 1987, Kauai, Hawaii. It is the thesis of the present paper that both
phenomena flow from the same source, Le. a bar-bend interaction that can be
understood in a wholly linear context. Five schools of thought, all represented
at the workshop, are considered herein with the intention of elucidating this
interaction.
1. The Delft school [Struiksma et al., 1985; Crosato, 1987].
2. The Genova school [Blondeaux and Seminara, 1985; Columbini et al.,
1987].
3. The Iowa school rOdgaard, 1986a,b].
4. The Minnesota school [Johannesson and Parker, 1988 a,b, 19891.
5. The Washington school [Smith and McLean, 1984; Nelson, 19881.
This paper serves as a companion to Johannesson and Parker [1989] in the
same volume. The essential comparison is between the Delft, Genova, and
379
Copyright American Geophysical Union

Vol. 12

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

380

Minnesota schools. The notation is that of Johannesson and Parker [1989],


but with occasional variations and simplifications.
In order to facilitate
comparison, reference is made to the notation of other theories in the following
way: DS B refers to the notation B of the Delft school.
Governing Equations
The basic equations used by Johannesson and Parker [19891 are presented
in this section, and compared with those used in the analyses 0 Seminara and
Blondeaux [1985], Struiksma et ale [1985], and Smith and McLean [1984].
Reference is also made to Odgaard [1986a].
A linear theory is formulated about a steady, uniform base flow with
constant depth H and depth-averaged downstream velocity U (DS Uo; GS U*),
in a channel with constant width 2b (DS B;
2B*).
The actual fl8w
considered is that over an erodible bed; it may differ modestly from the base
flow due to bed topography such as alternate bars, or channel curvature
(Figure 1).
The sidewalls are taken to be inerodible.
Downstream and
transverse coordinates, made dimensionless with b, are sand n, respectively;
the vertical coordinate , has been made dimensionless with local depth ii (( =
z/h, where z is a coordinate upward normal from the bed). The channel is
assumed to be wide in the sense

as

b
'""I=ii> 1

(1)

Here '""I = GS {3, DS B/(2h o).


Flow velocities are made dimensionless with U; the downstream and
transverse components are, respectively, uT and v*T+v, where u and v*
denote depth-averaged values, and T is a normalized vertical profile of
downstream velocity satisfying the condition

Fig. 1.

Definition diagram.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

381

Parker and Johannesson


1

J Td( = 1

(2)

Also, v is the helical secondary flow rKalkwijk and De Vriend, 1980; Smith and
McLean, 1984]. It follows from the above definitions and (2) that

J vd( = 0

(3)

o
The profile T is approximated as a function of ( alone, obtained from the
center region (away from the sidewalls) of the base flow.
Where rc denotes (dimensional) local channel centerline radius of curvature,
local dimensionless curvature C( = GS vlro) is given by

(4)
A bed friction coefficient Cf is defined such that
1"8
pU 2

Cf u2

(5)

where 1'8 is dimensioned downstream bed shear stress. Here Cf = .Q.S. C, DS


g/C2. The depth-averaged equation for steady downstream momentum balance
may then be stated as
1

l+nC u

au
os

+ v

au
on

- n1 [aon

+ l+nC uv

* _ _F
-

l+nC

u2

11

ae _

os

uh Tv + 2C
l+nC uh Tv]

Here F = UI JiH is base flow Froude number,


elevation, made dimensionless with H, and

Also, Tv denotes
depth-averaging, Le.

-2

dispersion

(6)

e denotes

water surface

= ,cf
parameter,

(7)
and

the

overbar

denotes

Tv

= J Tvd(

(8)

In deriving (6), the classical slender-flow approximations (al ()z ) al an, al ()z )
81 where nand s are dimensioned) have been used to simplify the Reynolds
stress terms before depth-averaging. After depth-averaging, the approximation

as,

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

382

1 has been made. Except for the dispersion terms [I(alkwijk and De
Vriend, 1980], (6) is essentially identical to the corresJ?onding relations in
Blondeaux and Seminara [1985] and Smith and McLean lI984]. Struiksma et
ale [1985] present a Cartesian version of the relation, also neglecting dispersion,
but Crosato's [19871 simplified formulation retains it.
The particular form for T used by Johannesson and Parker [1989] is that
of Engelund's [1974] slip-velocity method;

~ ~

(9)
where X and X are parameters not far from unity, given by
1

(10)

(Note that 0.77 5 X 5 1.92, 0.44 5 X 5 1.59, and 1.04 ~ T2" ~ 1.01 for 10 5
25).
This 1 approach leads to a constant eddy viscosity, and the
following equation for steady transverse momentum balance tnot layeraveraged);

C~12 ~

1~nc Tu

(Tv* + v) + (Tv* + v)

=Here

= ~ ;

F -2 {)~

on +

!! (-v*
n

(Tv* + v) - 1~nc T 2u2

+ Xii)
1

(11 )

different assumptions for T (e.g. logarithmic) lead to slightly

different formulations for shear stress due to the secondary flow (fX v in the
1
present case).
The slip velocity method leads to the following boundary
conditions on v [Johannesson and Parker, 1988b];

v(l)

v(O)

xv(O)

(12)

A form essentially identical to (11) can be found in Smith and McLean [1984];
they use a modified logarithmic form for T. Blondeaux and Seminara [1985]
and Struiksma [1985] do not explicitly solve for the secondary flow v; their
version of (11) is obtained by setting v = 0, T = 1 therein. Odgaard [1986a]
uses a power law for T.
The steady, layer-integrated equation of water continuity is

(l+nC)v*h

(13)

The equation of sediment continuity is


(14)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

383

Parker and Johannesson

Here TJ denotes bed elevation, made dimensionless with H. Also qn and qs


denote dimensionless downstream and transverse sediment transport rates per
unit width, respectively, made dimensionless with qo, where qo denotes the
corresponding dimensioned downstream sediment transport rate per unit width
of the base flow.
Furthermore, t = tqo/[(I-p)bH], where t denotes
dimensioned time and p denotes (constant) bed porosity.
Equation (13) is common to Smith and McLean [1984], Blondeaux and
Seminara [1985], and Struiksma [1985] as well. Smith and McLean [1984] treat
the flow only, and do not compute bed topography. Thus (14) is omitted
therein, but retained in the other two analyses. Odgaard [1986a] also retains
water continuity but does not treat sediment continuity.
In order to complete the above set of equations, it is necessary to relate qs
and qn to flow parameters.
Parker [19841 and Parker and Andrews [1985]
developed the following linearized relationship between qn and qs from the
work of Kikkawa et ale [1976] and Ikeda [1982] [but see also Hasegawa, 1981];

(15)
where
(16)
Here a* = 0.85, Jl is the Coulomb coefficient of friction (0.40
denotes the Shields stress associated with the base flow,
r*

so

0.60), and r*so

(17)

Tso

P Rg D s

In the above relation, Tso denotes the value of 1"s of the base flow, Ds denotes
the sediment grain size and R = (Psi ~1) denotes sediment submerged specific
gravity, and r* denotes critical Shields stress. If the bed is covered with

r:o

s),

dunes,
must be replaced with TO = Tal(p Rg D
where Ta is the
"grains stress," Le. the residual bed stress in the absence of dunes, of the base
flow. Johannesson and Parker [1989] evaluated the constant f* as equal to
1.19 for the case Jl = 0.43.
Struiksma et ale r1985] and Crosato J1987] use a form of (15) in which f3 =
DS I/(fs O), where , is between 1 an 2 and 0 can be identified with
Blondeaux and Seminara [1985] set fJ = GS r/O'-1/2, where r' is near 0.54 and
0' can be identified with TO; this assumption is essentially equivalent to (16).
Let Cfo denote the value of Cf associated with the base flow, and let

r:o.

(0

'Y Cfo

&.

(18)

Furthermore, let X denote (dimensioned)


Here (0 = GS X and DS ~
o
alternate bar or meander wavelength. The following dimensionless bar or bend
wavenumber can be defined;

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

384

10 2 , . - - - - - - , - 1 - - - -............1 -----.......-1--------.

;~~0.
o

10

CHITALE (1970)
SCHUMM (1969)
Q WOLMAN (1955)
LEOPOLD & 'MJLMAN (1957)
LANGBEN & LEOPOLD (1966)
CiiI KINOSHITA (1961)
1

.....
1
10

rn

00

--"'1
2

10 3

l
Fig. 2.

10.2 '--

o
0

@
~.~ ~

10- 1 -

.1.-

10

----'

c (METERS)

The rescaled dimensionless wavenumber r = k/ (0 versus meander wavelength ~ for


75 field cases. In most cases r is seen to be of order unity.

(19)

Here k = GS Am, DS Bk/2. In Figure 2, adapted from Ikeda et ale [1981]


(27rb)/k is plotted against k/ (0 for a wide range of self-formed alternate bars
and natural meander bends; it is apparent therein that the scale relationship
k
-

(0

0(1)

(20)

is of general validity. Furthermore, for natural streams, (0 is typically on the


order of 0.1 or less, as illustrated for several meandering streams in the table.
These two scaling relationships are crucial as regards illustrating the
relation between the treatment of the Genova school and the others. It may
be noted that the dimensionless transverse coordinate n has been defined so
that it varies from -1 to 1, i.e. over an 0(1) range. Likewise, if downstream
phase tP is introduced such that

tP

= ks

Copyright American Geophysical Union

(21)

Water Resources Monograph

River Meandering

Vol. 12

385

Parker and Johannesson


then J varies by the 0(1) amount 21r over one wavelength.
renormalizing k into the 0(1) wavenumber r, where

Further

k
r =-

(22)

o = lor "lJiP
0
os

(23)

lo

it follows from (21) and (22) that

According to (23), the equation of water continuity (13) is rendered as

on

ouh + 0 ( l+nC ) v*h


lor O(f)
Insofar as

(24)

is small and all parameters except v* have been scaled to be

lo

0(1), it is apparent that v* must be rescaled such that


v*

0
on

(l+nC)vh

(25)

loY

resulting in

Buh

r O(f) +

=0

(26)

Likewise, (6) and (11) rescale to


1
I
l+nC ruu + v
-2

-rF

= 1+nC

~I
~ -

f
fo

au

on

+ l+nC uv

2
1 [a
nu - ~
on uh Tv

2c
-]
+ 1+nC uhTv

(27)

(28)
Here I = oj oJ. Finally, with the transformation lot -+ t, the equation of
sediment continuity, (14), transforms with the aid of (9), (12), and (15) to

!Jlt

l~nc {rq~ + ~[(1+nC)qs[~ + Xl~:~) - r ~]]}

=0

(29)

Here

~ = 12gfo

is termed the coefficient of gravitational diffusion.

Copyright American Geophysical Union

(30)

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

386

In fact, the parameter r is typically 0(1) in natural streams. It can be


seen directly from Table 1 that ,2Cfo is 0(1), and f3 is seen to vary from
about 2.67 to 0.67 as 1"* /1"* varies from 1 (critical) to 16. The parameter r
so

is equivalent to GS

W and DS ~[~J.
o

The above equations are subject to the conditions of impermeable sidewalls;


that is,
(31)
They can be closed by relating qs and Cf to flow parameters. Parker and
Anderson [1977] have shown that these may be eXfressed in terms of up to
two independent hydraulic variables.
Parker [1976 expressed qs and Cf as
functions of u and energy slope; herein the completely equivalent formulation
of Blondeaux and Seminara [1985], according to which
qs

qs( Ts,h)

Cf

(32a,b)

Cf( Ts,h)

is used to facilitate easy comparison.


It may be noted that terms involving v contain the small parameter (0 in
the denominator in (27) and (29).
The scaling yielding such a result is
justified subsequently.

Linearization
The quasi-steady approximation of erodible-bed theory is implicit in (26)
(29), in that time dependence is retained only in the equation of sediment
(29) and (32) are now perturbed about the
continuity.
Equations (26)
base-flow solution;
N

=
=

up ; h

Vp ; v

e = eo - rF2

=1+
=0+

J + ep ; TJ

hp ;

Ts

vp ; qn

TJo -

=
=

Tsp ;

qnp

F2

qs

qsp

(33a,b,c,d)
(33e,f,g)

J + TJp

(33h,i)

The perturbed quantities, denoted with the subscript p, are assumed to be


TJ,
small, Le. Up (1. Note that insofar as h =

e-

(34)
Substituting (33) into (12), (26) - (29), (31), and (32) and linearizing
under the additional assumption of small centerline curvature
(35)

C ( 1
yields
ru' = - rFP

{' p

nC - 2Pup

Plh p -

L(0 iffari p

Copyright American Geophysical Union

(36)

Water Resources Monograph

River Meandering

Vol. 12

387

Parker and Johannesson

(37)
ru'p + rh'p + ~
= 0
iJn

(38)

(39)

subject to the boundary conditions

de = 0

Vp

(40a,b,c)

o
v p In=::I:l

[\::(0) _r ~]

(41a,b)

In =::I:l

The coefficients P, Pi, M, and M1 in the above relations result from


linearizing (32a,b);

I ac f
P1=1---m:P
Cfo Ull l 1
M

2P

8Qs l
7JT;1

M1

(P 1

1) 8Qs l - &lsI

7JT; 1

OIl 1

(42a,b)

(42c,d)

The subscript "1" denotes the base flow (h = 1, Ts = 1). The values of the
coefficients differ depending on the choice of load and resistance relations, but
typically P
1, P1
1, M > 3, M1 < M [e.g. Parker, 1976]. A sample
evaluation can be obtained in terms of the load relations of Meyer-Peter and
Muller and the resistance relation of Keulegan, in the forms used by Blondeaux
and Seminara [1985]. In the present notation, these take the respective forms:
N

qs

1 _

-Ti- T-;l ]3/2


Tso

--~---:*---

[1 - T~r/2

3/2
Ts

-1/2

Cf

[H ]

2.5ln 1<; h

(43a,b)

where k s denotes a roughness height, yielding [by either the above method or
that due to Parker, 1976],
P

1 ;

P1

M1

5/Cfu

2.5/Cfu M

Copyright American Geophysical Union

(44a-d)

Water Resources Monograph

River Meandering

Vol. 12

00
00

TABLE-Various Parameters for Several Meandering Sand-Bed Streams at Bankfull Flow


-1/2

lliver

Reference

Cf

Root +

Johannesson [1985]
Johannesson [1985]

Zumbro+
Minnesota*
Red Lake+
Minnesota*
Pembina

Genesee~
Muddy Creek
Fall

Johannesson [1985]

fo

f3

21.8

9.5

0.020

0.13

6.5

0.39

2.03

15~4

10.1
11.7

0.043

0.22

5.1

0.34

0.79

0.067

0.34

5.1

0.90

1.16

15.5
11.4
16.2
9.7

0.082
0.016
0.103
0.075
0.10
0.31

0.23
0.25
0.13
0.17
0.30
0.28

2.8
15.6
1.3
2.3
3.0
0.90

0.80
0.67
0.93

0.63
3.65
0.56

1.58
1.93

2.62
1.01

13.2
13.7
Parker [1982]
26.7
Beck [1983]
12.5
Beck et ale [1983]
11.4
Dietrich 3 Smith [1983] 7.73
Thorne et ale [1985]
4.53
Johannesson [1985]

*Different

reaches

6.0
6.3

+D s estimated as 0.5 mm

~Ds unavailable

S'
CIJ

::s
~
::s
n

(0

~
::s

0.-

<

(0
~

0.-

(0
(0

"'C

(0

::s
S

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

389

Parker and Johannesson

If, however, Cr were assumed to be constant, P and PI would equal unity,


From (44a d), this is
clearly the case in the limit of small Cr. The values for Cr of field streams
in Table 1 suggest that in a simple theory, PI might be set equal to unity
This
and MI neglected compared to M, without a gross loss of accuracy.
simplification, Le.

M would be given by (44c), and Ml would vanish.

PI

Ml

(45a-c)

is adopted by Struiksma et ale [1985], Crosato [1987], and Johannesson and


Parker [1989]; Blondeaux and Seminara [1985] retain the general formulation of
(42) and (431. Odgaard [1986a] does not consider sediment continuity, and so
. has no parameters comparable to M and Ml; (15) therein implies that P = PI
= 1. Note that M = DS b, QS f1; P = GS SI/2; PI = GS 1 - S2.
Secondary Flow

a Curvature-Driven Phenomenon

88

The perturbed flow field described by (36)


(39) could be driven by
curvature, by topographical effects independent of curvature (such as alternate
bars), or by a (linear) summation of the two effects. It is shown herein,
however, that the following simplifications hold:
a) The equation of transverse momentum balance (37) can be reduced to
two decoupled equations, one for layer-averaged transverse velocity Vp
and one for secondary flow V p
b) The secondary flow is essentially driven by curvature alone, and is
unaffected by bed topography at the linear level. It can be neglected
in a straight channel.
To establish the above, (37) is first layer-averaged, yielding
N

=-

r fo Tv ' + r T2" f2 v' - VC


pop

2
F- 8f..p - foX vp(O) - f2 Vp

Oil

(46)

Subtracting (46) from (37) yields


r fo(TVp - TVp)' - (T2 - T2")(C - rf 2 v') = foX [ii p
o

(47)

+ vp(O)]

The source term driving the secondary flow in (47) is C - r f2 v'.


o

The

leading terms in the equation, r Tfo v', and its depth average represent the
p
redistribution of secondary flow momentum by streamwise convection; as such,
they cannot change the order of magnitude of Vp. Thus, for the purposes of
an order of magnitude estinlate, (47) and (40a,b,c) are solved dropping these
leading terms. This yields the s'olution

Vp
Go(()(C - rf 2 v')
l O P

loX

where

G o (()

=f

(T2 - T2")d(Md('

(' 1

f f f
o

o ('

- T

(48a)

(T2 - T2")d('''d(Md('

(M

Copyright American Geophysical Union

(48b)

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

390

The evaluation for Go in the case that T is given by (9) is presented in


Johannesson and Parker [1989]; this yields the auxiliary result

fOX/"p(O) = -

~ [x + ~]

k (C - f(~ v~)

(49)

First, the case of a straight channel (C = 0) is considered. It is apparent


from (49) that in this case the secondary flow must be driven by bed
topography, through vp; (49) then yields the scale estimate
(50)
(recall that X and X are not far from unity).
1

Using the previously-justified

assumption 1'2" ~ 1, then, layer-averaged transverse momentum balance (46) in


a straight channel (C = 0) reduces to

r(~ v~

=-

o!- - (~

F-2 8cp

Vp

[2]

+ a 45 r(~ v~

(51)

Here (50) has been employed in the order estimate.


It is apparent from (48)
(51) that the residual topographically-driven
secondary velocity V p in straight channels is very weak, scaling as only on the
The
order of four percent of the mean transverse velocity v p in (51).
secondary flow term in (46) can thus be dropped in a linear treatment of
alternate bars in a straight channel. This assumption has been implicitly used
in almost every linear analysis of straight-ehannel alternate bars, including
Callander [1969J, Parker [1976], Fredsoe l1978], and Blondeaux and Seminara
[1985]. Heretofore, however, a proper justification has been lacking.
Returning to the case of curved channels, it can be seen from Table 1 that
(2 typically scales as 0.01 or smaller for natural channels; values are even
tV

smaller for many laboratory channels [Johannesson and Parker, 1988a,b].


for channels of modest curvature such that

Thus

(52)

tV

0.1

it follows that r(2 v' scales as an order of magnitude smaller than C in (47);
yielding

rTlOv ' P

(T2 -

V)C

loX

[ii p + vp(O)]

(53)

The secondary flow is seen to be driven solely by curvature, and is essentially


uninfluenced by topographical effects. The treatments of Ikeda and Nishimura
[1986], Smith and McLean (1984] and Johannesson and Parker [1988a,b] are all
In accord with this conclusIon at the lowest order at which secondary flow is
computed. Odgaard [1986a] stands alone in linking Vp strongly to Vp.
In regard to the applicability of (52), it is worth noting in passing that
nine of the sixteen bends on the Beatton River studied by Hickin and Nanson
[1975] possess values of C between 0.07 and 0.16; (2o is equal to 0.0018 at
bankfull flow.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

391

Parker and Johannesson

Johannesson and Parker [1988a,b] have formally shown that the dropping of
the leading term in (47) for most natural meandering channels (but not for
several experimental meandering channels) is justified. The steps leading to
(49) are thus validated in a more rigorous sense. Again, approximating ~ =
1 in (46) and dropping the o( (2) term in (49), the equation of depth-averaged
o
transverse momentum balance becomes
I

r f~ vp

=-

-2

ae

(Ji1 - f~

VP

1 [X + "72] 45
2C
+ X3"

(54a)

The final term, due to the secondary flow, is clearly negligible compared to C
itself, yielding

=-

vI - C

(2

F -2 ~ -

(2
0

vp

(54b)

This is the relationship employed by Blondeaux and Seminara [19851.


It
admits of both curvature-driven and topographically-driven solutions (or Vp,
and reduces to (51) in the limit as C -+ O. Comparing (47), (51), and (54), it
can be seen that for the range of natural channels considered, Vp is essentially
driven by curvature alone, and the equations for Vp and Vp can be decoupled.
The linear analysis of Blondeaux and Seminara [1985] and Johannesson and
Parker [1989] share this feature, and it is implicit at the linear level in the
analysis of Struiksma et ale [1985].
Expansion
It is clear that a complete theory of meandering should encompass alternate
bars in a straight channel as well as curvature-driven flow and bed topography
in a meandering channel.
In a linear theory, this can be done in a
straightforward way. Let

b
o=-(l

rm

(55)

where rm is the minimum (dimensional) centerline radius of curvature on a


given reach.
Thus an order4>ne dimensionless local centerline curvature
function u( s) can be defined from (4) and the above relation such that

ou

(56)

denote the curvature-driven or "bend" part of Up, Le. the part


Let e.g. u
pB
that vanishes as o -+ o. The following expansion may be envisaged;
U

pB

= o

UBI

-to

~ UB2 ...

(57)

The residual curvature-independent part of Up associated with bed topography


induced by e.g. straight-ehannel alternate bars, u pT ' is then given by
(58)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

392
Now, where
and

'6

is some measure of straight--ehannel bed topography amplitude,

(59)
then upT can be expanded as

upT = t5uTl

(60)

+ b'2uT2 + ...

The perturbed quantities Vp, hp {p, and TIp in (36) - (39) can be expanded
likewise, yielding
up

'l/Jo

UBI

SU

{p

'l/Jo

e Bl

t5eTl

+ ...

Tl

hp

'l/Jo h Bl

TIp

'l/JoTlBI

+ bhTl +

(61a,b)

(61c,d)

t5T1Tl

where from (34)

hTI

e Tl -

TlTl

(62a,b)

In accordance with the previous section, however, the secondary flow v is


expanded in curvature alone,
Vt

'l/Jo vBI

(63)

+ ...

Equations (56), (61), and (63) may be substituted into equations (36), (38)
(41), (51), (53), and (55). Herein, only the linear terms in 'l/Jo and t5 are
retained, so the subscript "1" is dropped to reduce clutter. The terms in "po
and t5 can be decomposed by considering the equations obtained in the limit as
"po . . . 0, and subtracting them from the full equations. The following "T" and
"B" problems are obtained:
"T" problem (associated with bed topography):
N

(64)

r(2

v'
T
ru I

=+

rh I

-2

BeT

~
~1

lJv

(2 V

+ OilT =

(65)

(66)

(67)
(68a,b)
"B" problem (bend):

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Parker and Johannesson

rUB

=-

393

rF

-2

eB I

rf 2 v I
o

nO' -

0'

=-

PUB

for TV~ - (T2 - 1'2)0'

-2

ru~ + rh~ +

Pt h B -

aeB -

ml

fOX

urv

1
B
fo ---oil

(69)

(70)

f2 V

[VB

+ vB(O)]

(71)

Ov

o!- =

(72)

(73)

(74a,b)

(74c,d,e)

It is seen by inspection that the T-problem is homogeneous in nature.


Any perturbations must be self--excited, Le. alternate bars. The B-problem is
clearly driven by inhomogeneous curvature terms in (69) rv (71).
The
equations presented above are similar to those of Blondeaux and Seminara
[1985] but are generalized so as to specifically compute the secondary flow and
associated dispersion.

Alternate Bars in a Straight Channel


The linear theory of alternate bars in a straight channel is embodied in the
"T" problem of the previous section. It has been treated by e.g. Callander
[1969j, Engelund and Skovgaard [1973], Parker r1976], Ozaki and Hayashi [1983],
etc., and an essentially complete solution of tbe linear problem can be found
in Fredsoe [1978], Kuroki and [(ishi [1985], and Blondeaux and Seminara
[1985]. The purpose of the present section is to develop a very simplified
theory that allows for a subsequent comparison between various theories of
resonance and overdeepening.
A complete linear stability analysis of the "T" problem, Le. (64) rv (68a,b),
can be found in Blondeaux and Seminara [1985].
The simplified treatment
outlined herein is an extension of the approach of J ohannesson and Parker
[1989] for curved channels, using the previously justified simplifying assumption
of small f2. It is of interest to explore the consequences of the assumption.
o

Equation (65) is seen to approximate to


(75)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

394

Integration of (66) in n and if> yields, with the aid of (62b) and (68), the
condition

(uT +

eT -

11T )dn

(76)

-1

according to which channel water discharge is unchanged by alternate bars.


order to study alternate bars, the following perturbations are introduced;

(UT

'

l1T

'

eT )

= [U b(</>} , l1b(</>} ,
VT

= Vm( </>} cos [~

eb(</>}]

Sin[~

n]

In

(77a)

n]

(77b)

where "b" and "m" denote "near-bank" and "middle", respectively.


of (64)
(68), (76), and (77) yields

Reduction

eT
r

~+

2P

~ + {r2(l+Mdu~ +

Ub

(78)

r[2P(l+Mt} - (M-l}Pt

+ 2rp [~r
The appropriate form for

Ub}

+ [~fr]Ub

= 0

(79)

for a stability analysis is

(80)
Here u* is an 0(1) coefficient, and a and c are normalized amplitude growth
rate and downstream migration speed, respectively. That this corresponds to
alternate bars is readily seen by the implied form for the bed perturbation;

l1T = 11* eat cos(</> - ct - </>11}

Sin[~

n]

(81)

where 11* is an 0(1) coefficient and if>11 is a phase shift.


Substitution of (80) into (79) yields

0'

c -

_ 2PJl - r 2P1(M-1)
J2 _
r 2 + 4p2
r

r r2 ~
+ p2

r[Jl + 2PJ2] _
r 2(1+Ml)
r 2 + 4p2 - r

2P~2P (1+M!l - (M-1)Pl]


r + 4P 2

4P

r -2
2~

(82a)
(82b)

where
(82c)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

395

Parker and Johannesson

[;t r

J2 = 2P(l+Mt) - (M-l)Pt +

(82d)

The qualitative characteristics of a and c are readily analyzed with the aid
of (45), Le. P = 1, PI = 1, MI = 0, without any loss of generality.
Accordingly,

Jt

r2 -

[;r

2r

(83a,b)

a=----------&.------'---

S0
c

as

S0

r2

as

of downstream

TIr:r+f2

~ M-l

r r

r2

Stable and unstable regions, and regions


migrating bars, are delineated as follows;

r2

S 2(M-3)

2 (3-M)

(83c,d)

+4
and

upstream

(84a)
(84b)

A stability diagram in wavenumber r and gravitational diffusion coefficient


can be determined from (82a,b). In order to allow comparison with the
complete theory of Blondeaux and Seminara r1985] and Colombini et ale [1987],
parameters associated with Figure 2 of the latter paper are employed. That
is, r* = 0.3; Ds/H = 0.01; and (j = 0.548. Colombini et ale [1987] employ
so
-1/2
(43a,b) for load and resistance relations; they yield efo = 15.22; F = 1.071;
M = 3.557; P = 1; PI = 1.328; and MI = 0.584 for this case. Regions for
which a S 0 and c S 0 obtained for these values are shown as case a) in
In case b) of the same figure, results obtained during the
Figure 3.
simplifications (45), Le., P = 1; PI = 1; and MI = 0, are shown.
In order to allow for. a direct comparison with Figure 2 of Colombini et ale
[1987], (82a,b) are transformed into the (-k plane, using (18), (27), and (30).
In Figure 4, the results of case a) and case b) are plotted along with the line
of neutral stability of Colombini et ale rI987]. The simplified theory presented
herein is seen to grasp the essence of the instability mechanisms, including the
range

(85)
or from (30)
1 -

- if

<

1r

I1JC- I / 2
fo

"2 v'PI(M-i)

(86)

within which the channel is too narrow for the formation of alternate bars at
any wavenumber.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

R,esonance and Overdeepening

396
1.50

DOWNSTREAM MIGRATIN'1
(!J

1.25

STABLE

-~
a:

(!J

.. ..

......

. ...

1.00 1-<
W

..... .- ...-

-- - --- -

a:

....

,. ,. ...-

(J)

a.
:::>

0.75

'

... .... .....

,,-

"

"

UNSTABLE

0.50

---

""

I:: I

0.25 -

Fig. 3.

... CASE A

-:.:i

..f
u:.' :
o

--

.:i.... I

I
II

CASE B
INCIPIENT RESONANCE
I
4

Alternate bar stability diagram in the k-f plane. Case a) denotes the results fronl
(82arv b). In case b), the simplifications of (45) have been used.

It is seen from a comparison of cases a) and b) in Figure 4 that the error


in employing (45) is rather modest. The difference between case a) and the
line of Colombini et ale [1987] therein, however, illustrates the effect of
dropping the O( (2) terms in (65). The form of the leading (inertial) term
therein includes the combination r{2, suggesting that the approximation is valid
o
for r
0(1), but fails when the r{2 is no longer small. This is verified in
Figure 4. As a result, the simplifi~d theory fails to predict stability at all
values of "y for large wavenumber k, and the amplification spectrum a(k) fails
to possess a maximum value.
The simplified theory presented herein is seen to provide a cogent but
incomplete explanation of alternate bars. Its formulation was motivated by a
preview of some numerical calculations presented at the final U.S.-Japan
Workshop on River Meandering, Hawaii, 1987, and subsequently incorporated
in a thesis [Nelson, 1988].
It is of interest to note that if Jl and J2 vanish in (82a,b), both the
growth rate a and migration speed c vanish, and the problem is rendered
time-independent. Under these conditions,
N

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

397

Parker and Johannesson


45

r--~"'--:"--r-------r----r--------,.----r------,

FULL THEORY
(COLOMBINI ET AL.)
....... CASE A
- - CASE B

INCIPIENT RESONANCE

25

y
20

UNSTABLE
.....

10

...................................................................
STABLE

OL.-----L.....----.l--_--L....._ _----'-

0.25

0.50

0.75

--'-_ _

----I

1.00

1.25

1.50

k
Fig. 4.

Alternate bar stability diagram in the k-, plane.


above.

1r

illL

Cases a) and b) are described

r = [~r[(M-l)Pl - 2P(l+M.)]

rres = 2 ~ (I+M1J

(87a,b)

where rres is a natural resonant wavenumber, and (79) reduces to


u

b+

ub

=0

(88)

It is clear that if the homogeneous equation (79) were to be forced under


constraints (87a,b), Le., a the natural wavenumber and with vanishing
damping, the spatial oscillatory system would go into resonance.
The
condition of incipient resonance is noted on Figures 3 and 4 for cases a) and
b). That this is indeed the resonance of Blondeaux and Seminara [1985] will
be demonstrated subsequently. Suffice it to remark that since, according to
(30), r > 0, it is seen from (87b) that a necessary condition for incipient
resonance is

M > 2P(I+Mt) + P t

Copyright American Geophysical Union

(89a)

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

398
or for the special case of (45),

M > 3

(89b)

Blondeaux and Seminara [1985], Struiksma et ale [19851, and Johannesson and
Parker [1989] all quote values of M sufficiently high for resonance; Struiksma

et al. use a value of 5. It can also be seen in Figures 3 and 4 that incipient
resonance occurs at 0(1) values of r, so that a theory that drops O( (2) terms
o

in (65) can still delineate it accurately.


The theory presented herein for alternate bars is thus seen to be a
simplified version of that of Blondeaux and Seminara [1985]. The relation to
Struiksma et ale [1985] can be seen by considering perturbations that may grow
To this end, the temporal terms in (79) are
spatially but not in time.
neglected, r is allowed to be complex, such that
r

rr

iri

(90)

and a perturbation of the form


(91)
is introduced, using (21) and (22).
The spatial perturbation thus has
normalized wavenumber rr and normalized spatial growth rate -rio
Substitution into (79) and reduction yields

rj

=~

(l~~d

r~ = (l~Md [2[~fpr - i (l~~d]

(92a,b)

Equation (92b) defines a relationship between the parameter rand


normalized wavenumber that is plotted in Figure 5 for the cases a) and b)
previously defined in the context of Figure 3. The parameters rr and ri are
plotted against r therein according to (92a,b).
A range of spatial growth
(ri < 0) exists for the conditions J2 < 0, which reduces for case b) to

r < [~] 2(M_3)

(93)

Case b) of Figure 5 is precisely equivalent to Figure 6 of Struiksma et ale


[1985] with the transformations
rr -- -DS k rAw
\
., 2"
ri -- -DS klAW,
\
. r -- -DS ~
8 X;
Aw
2"

(94a,b,c)

allowing for the trivial difference that M = 3.577 in the former figure and
= 5 in the latter one. Stntiksma et ale [1985] also employ (45).
It is seen by inspection of (92a,b) that incipient resonance is again given
by J1 = J2 = 0, or (87a,b),. in which r is replaced by rr. The points of
incipient resonance are noted on Figure 5.

M(= DS b)

Flow in Sinuous Channels


The curvature-driven part of erodible-bed open--ehannel flow is embodied in
the bend, or "B" problem of (69)
(74). A solution for the secondary flow
N

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

399

Parker and Johannesson


5 ,...----"""T1---""T'"1---"T"""1---1r--/------,.1-----.

INCIPIENT RESONANCE
/ /
(rj = 0)
/

. CASE A
_ _ CASE B

4 -

/ / /
/

r~
I

3~

.....
.

/
/ /

;' ;'"..:. ~.;,.~. ~:.::.:.~<.:.... . .


/ ..' / / . . \.
t
I.

2 -

I .....
1 ~,.;.

J::

.....

\~
\
r

.....

/ / ...
/

/.......

...

I'

.
....

".

..

,~
~

, .

./

-1

......I

234

L....-_ _.......

..I--I

II.......-_ _.......I_ _- - - - - I

r
Fig. 5.

Plot of rr and ri versus

for steady alternate bars.

VB can be obtained directly from (71), (74c N e), and the assumed form (9) for

Building on the base of Ikeda and Nishimura [1986],


Johannesson and Parker [1988b] have illustrated that the leading term (orTv~

the primary flow.

in (71) is negligible in most natural channels, as has been implicitly assumed


by others [e.g. Engelund, 1974; Smith and McLean, 1984]. Accordingly, (71)
and (74c eJ yield
N

(95)
where Go is given by (48b). Substitution of (9) into (48b) results in the
Engelund [1974J solution. In analogy to (49), further reduction with (10), (18),
and (30) provides the result
2
A - 7.51 X +"7
- -f3-~

[Johannesson and Parker, 1989].

(96a,b)

The minus sign in (96a) indicates that the transverse bed shear stress due
to secondary flow is directed toward the inside of a bend. It can thus act to

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

400

scour the outside and deposit on the inside to a degree proportional to A. It


is clear from (96a) and previously cited estimates for /3 and r that the term
X v(O)jlo is 0(1), and has been correctly retained in (39) and (73).
The
1
parameter A is equivalent to the GS ajr, and the DS AfsfJ.
In fact, A has a simple physical interpretation as a bed scour factor [Ikeda
et al., 1981]. Substituting (96a) into (73), it is found that

a~

~B

at + r MU~ - rMlh~ + Oil - r

a [Au
on

a~] =

+ Oil

For steady, developed flow in a bend of a uniform curvature (u


(97) yields the solution
7J

(97)

constant),

= - Anu

(98)

which identically satisfies the boundary condition (74b). Thus A describes the
strength of the outer pool and inner bar produced by secondary flow. From
Similar
(96b) and (16), A is seen to increase the base-flow Shields stress 1"*.
so
forms have been ~resented by other researchers [see Odgaard, 1981].
A
comparison of (96a) in conjunction with (16) against data is presented in
Johannesson and Parker f1989].
Johannesson and Parker [1988a] have obtained an approximate formulation
of the dispersion term TVB in (69), using (9), (95), and (48b);

(99a,b)
It is useful to develop a formal solution of the "B" problem in several
For the case of steady, developed flow in a bend of uniform curvature,
(70) admits the solution

ste~s.

(100)
From (62a), (98), and (100), then,
hB

(A + F2)nu

(101)

in this case.
It is assumed for the moment that even in a channel of spatially varying
curvature u, (98), (100), and (101) hold locally as an approximation; in this
case the parts of the solutions for 7JB , eB' and hB driven by local curvature
("e") are, respectively
7Je

=-

Anu

he

(A+F2)nu

(102a,b,c)

From (69), (72), (99a), and (102a-e), the equations governing the portions of
uB and VB driven by local curvature (u e and v e ' respectively) are

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Parker and Johannesson

401

=-

rub + 2Pu c

rnO"

+ [Pt(A + F2) + As - l]nO'

Ov
-of
= - r(F2 +A)nO"

rub +

(103)

(104)

where from (74a)


(105)
It is quickly seen, however, that the solution to (102)
(105) cannot satisfy
sediment continuity, (73). The "B" problem must then be completed in the
form
N

(106a e)
N

where
hF
and from (69), (70), (72)

(74a,b), and (102)

(106f)

~F - T}F

(106),

(107)
v

1'(2

rU F,

=-

= - F

r h'F

-2

8~F

un

Ov

(2 V

+ onF =

(108)

(109)

(110)

r(M-1)ub + r(l+M t )hb


8T}F

onln=l

(llla,b)

= 0

Thus the solution for the secondary flow (95), and the summation of the "C"
problem of (102)
(105), and the "F" problem of (107)
(111) is equivalent
to the original I B" problem.
In fact, the "C" problem is equivalent to Engelund's [1974] second
approximation. It is this part of the solution that was used by Ikeda et ale
[1981], Parker et ale f1983], and Parker and Andrews [1986] in their treatment
of bend growth.
The "F" problem is seen to be identical to the "T" or
alternate bar problem of (64)
(68), except for the curvature forcing that
N

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

402

It is this forcing by
enters through the sediment continuity equation.
curvature of the homogeneous alternate-bar problem that gives rise to the
possibility of both the resonance of Blondeaux and Seminara r19851 and the
overdeepening of Struiksma et ale [1985]; thus the appellation "F' problem
Resonance and Overdeepening in Steady Sinuous Flow
Herein the analysis is restricted to steady flow. Blondeaux and Seminara
[1985] have solved the complete "B" problem for steady flow in a sinusoidal
channel. Herein the stead~ problem is treated in a way that closely follows
Johannesson and Parker 1989], using the same approximation in the "F"
problem as was used for a ternate bars, Le., neglect of the O( (2) terms in the
o
depth-averaged transverse momentum equation. As opposed to the case of
alternate bars, it will be demonstrated that no important feature of the
phenomenon is lost in this way as regards bend flow.
The "e" problem is already partially solved; Le. (102a,b,c).
For the
special case of a sine-generated channel, Le.
.
(J

(103)

(112)

sin>

(105) may be further solved to yield


ue = n(a e cos>

(113a)

+ be sin

(113b)
where
-

ae

=-

2P

+ A

r r2

A=

P 1(F2

(114a,b)

4p2

and

+ A) + As - 1

(115)

terms in (108)
In analogy to (78), the consequence of neglecting the O( (2)
o

of the "F" problem is that r may be taken to vanish.


demonstrated as follows: from (108)

This may be

(116a,b)

Le.
is a function of > only. For steady flow, the conditions of constant
r
water and sediment discharge applied to (104), (109), and (110) yields with
the aid of (102), (105), and (111), the relations
1

f
-1

(u r

+ hr)dn = 0

(Mu r - Mthr)dn

-1

Copyright American Geophysical Union

(117a,b)

Water Resources Monograph

River Meandering

Vol. 12

403

Parker and Johannesson


from which it follows that

(118a,b)
Integrating (107) across channel width and reducing with (116) and (118), it is
found that {F is constant. Any such free constant can be absorbed into the
constant
of (33h), resulting in

eo

(119)
The steady "F" problem (106)

(111) thus reduces to

ru~ + 2Pu F + Pt 7J
F

rU FI - r7JFI +

Ov

roF - r

rMu~ + rMt7J~ +

Fl n =l

827JF

0fi2
0

Ov

roF =
=-

(120)

r(M-1)u

c+

r(l + Mt)h

87JF

roln=l =

(122)

(123a,b)

with the auxiliary relation


(124)
For the special case of a sine-generated channel, the inhomogeneous terms in
(122) are given by
-r(M-1)u

C+

r(l+Mt)h

c= n(Dtcosq, +

D2sinq,)

(125)

where
Dt = r[(l+M t )(A + F2) - (M-1)bc]

D2 = r(M-1)a c

(126a,b)

(126) can be solved exactly, in analogy to Blondeaux and


The solution is presented in the Appendix. A very accurate
approximate solution can, however, be obtained by means of a first-order
Galerkin approximation. In analogy to (77), it is seen that the assumed forms
In fact (120)

Seminara [1985].

(U F, l1F, {F)

= [UFb(lP) , l1Fb(l6), {Fb(lP)]sin [; n]

(127a)

vF = VFm(lP) oos[; n]

(127b)

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

404
satisfy the boundary conditions (123a,b).
bank and middle. Likewise, placing

Here "b" and "m" again denote


(128a,b)

it follows from (102c) and (103) that uCb and h Cb are independent of n.
Approximating n by the first term in its Fourier sine series yields

n
Equations (120)

Sin[; n]

(129)

(122) then reduce with the aid of (127) and (128) to

r2(l+Ml)uFb

=-

~~

r[2P(l+M 1)

P'[- r(M-l)u

Cb

(M-l)P 1

+ r(I+M 1)h

[;rr]u b + 2rp[;rU b

cb]

(130)

Here, the inhomogeneous term in (130) is found with the aid of (129) to be

- Pl[- r(M-l)u

Cb

r(l+M 1)h

cb] =

~ PdD1cos> + D2sin>}

(131)

for the case of a sine-generated channel.


It should be noted that in the companion paper by Johannesson and
Parker [1989], a moment method is used in which 8/1r2 is replaced by unity in
(129) and (131). The results obtained therein for the "F" solution differ from
the solution obtained herein by this factor.
The analogy between (79) of the alternate bar theory and (130) is striking.
In particular, (130) appears as the steady-state, spatially oscillating alternate
bar equation, forced by curvature. In the case of sinusoidal forcing, Le. (131),
resonance can be expected if bend wavelength and channel width are chosen
such that rand r are equal to the resonant values associated with the
alternate bar theory, and given in (87a,b).
The solution to (130) and (131) is
(132)
where

a F -

8 P J I D1 + r J 2 D2
1
J2
2 J22
1 + r

b F -

PI

8 -r J 2D1 + J I D2
J2
2 J2
1 + r
2

(133a,b)

and Jl and J2 are given by (82c,d). Recalling that the condition of vanishing
Jl and J2 is equivalent to (87a,b), it is verified tbat the "Fit problem does
indeed display resonant behavior.
The complete solution to u of the "B" problem in the case of a
sine-generated channel is then seen from (106a), (113a), (114) and (132) to be

uB

[ae n + ar Sin[; n]]

cos>

+ [be

+ br sin[~

The near-bank streamwise velocity perturbation is thus given by

Copyright American Geophysical Union

n]]Sin>

(134)

Water Resources Monograph

River Meandering

Vol. 12

405

Parker and Johannesson


10

r------~A_--__r_----r___---_,.__---___.

0.--------+------------------1

-5

'5 8
-10

...... CASE A
CASE 8

-15

-20
0

Fig. 6a.

Plot of b

versus r illustrating resonance. Case a) was obtained from the exact


B
solution of (120) N (126). In case b), the approximation embodied in (129) has
been used to obtain a solution. In both cases, the simplification (45) has been
used. The agreement is excellent.

(135)
where
(136a,b)
It is seen by comparison with (112) that the coefficient fiB in (136b)
corresponds to the part of the velocity perturbation that is in phase with
curvature.
Blondeaux and Seminara [1985] provide a plot of fiB versus
wavenumber, as obtained from their complete theory, in their Figure 2. The
load and resistance relations used are (43a,b); T*so = 0.25, Ds/H = 0.005, and

f3 = 1.08.

- 1/2

These result in the values of efo = 16.96, F = 0.770, M =


3.70, P = 1, P1 = 1.295, M1 = 0.545 and (using their parameters), A = 6.48.
Furthermore, As is set equal to O.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

406
10

....-----..,.-----~---~---____r_---__,

CASE C
..... BLONDEAUX &
SEMINARA (1985)

b8

0 1------+------------------1

-5

1....-...---.. . . . .- --....1....----....1....--------.. . .

-10

0.1

0.2

0.4

0.3

0.5

k
Fig. 6b.

Plot of b
Figure 6a
(M1 = 0,
compared
agreement

versus k illustrating resonance. Case c) is the same as case b) of


B
except that the computed values of M1 and P1 are used instead of (45)
P1 = 1). It embodies the approximate solution obtained herein. It is
with the exact solution of Blondeaux and Seminara (1985).
The
is excellent.

First, in Figure 6a, fiB is plotted versus r for the case of I = 30 (r =


0.345), using the above input parameters, except that the values of (45), Le.,
Pi = 1 and Mi = 0 are used. For these values, the result from (114b),
(133b), and (136b) is plotted along with the result obtained from the exact
(126), in which the approximation (129) is not used. The
solution to (120)
agreement between the approximate and exact theories is seen to be excellent,
validating the approximation (129). Similar agreement is obtained with other
input parameters. Resonant behavior is apparent near the value rres = 1.30,
corresponding to the "natural" wavenumber predicted from (87a).
In Figure 6b, fiB is plotted versus k for all of the above input parameters,
Le. without using (45), for the approximate theory herein. The results are
compared with those of Figure 2 of Blondeaux and Seminara f1985J, for I =
30. The agreement is quite good; it is seen that no essentia information is
lost, even at large r, due to the approximation embodied in neglecting O( (2)
o
terms in (108), or that embodied in (129).
The treatment of Struiksma et ale [1985] can also encompass resonance.
Their original analysis does not indicate this, nor is the reader supplied with
enough information to verify this for himself. The simplified version of their
work presented in Crosato [1987], however, is readily reduced to the relations
N

hCb

rUCb + 2u Cb

r2u~b +

[3-M +

= ~1r

A(J

= - ~ ru' + ~

(137)

(A

[~rr]ru~b + 2[~rruFb

As - l)u

= (M-l)ru Cb - rhCb

Copyright American Geophysical Union

(138)
(139)

Water Resources Monograph

River Meandering

Vol. 12

407

Parker and Johannesson

ASYMPTOTE

DEPTH ALONG
OUTER BANK

1----S
STRAIGHT
REACH

Fig. 7.

CURVED
REACH

Illustration of overdeepening.

where A = DS AfsO, M = DS b, As = DS E-l and r = DS (8/ 1r2)(A w/ As).


A comparison with (102c), (103), (128), and (130) reveals a correspondence
that is nearly exact for the choices P = p. = 1, M. = o.
The only
differences are the neglect of F2 compared to A, and the extraneous 2/ 1r, in
(137) and (138). Thus (139) exhibits the same behavior as that shown by
Blondeaux and Seminara [1985] and Johannesson and Parker [1989], including
resonance.
A strong manifestation of resonance in a periodic, sinuous channel requires
that the meander wavelength and width-depth ratio, Le. rand r, fall in a
As yet, there is
rather narrow band near the values specified by (87a,b).
little direct experimental or field evidence for resonance. Kinoshita and Miwa
[1974] have, however, observed the tendency for alternate bars to migrate
downstream progressively more slowly as the sinuosity of a meandering channel
is increased. Beyond a critical sinuosity, the alternate bars cease migrating
downstream, and meld with bend point bars to produce alternate zones of deep
scour and extensive fill. The phenomenon cannot be described completely at
the linear level, but the experiments strongly suggest that resonance plays a
role in the synchronization of the steady bar and bend patterns.
There is, however, considerable indirect evidence for resonance in the
phenomenon of "overshooting," or overdeepening, described by Struiksma et ale
lI985]. The process is schematized in Figure 7. Water flows from a straight
reach to a curved reach (taken to be of constant curvature therein for
simplicity) of an erodible-bed channel. One might expect that the transverse

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

408

slope would increase monotonically in the downstream from zero in the


straight reach, approaching the value given by (98) asymptotically. In fact,
the transverse bed slope is often greater than the asymptotic value in the
upstream part of the bend, giving rise to overdeepening. The overdeepening is
part of a spatially oscillatory pattern that may persist in the downstream
direction. The oscillation represents the free (alternate bar) response to an
impulsive forcing caused by a step increase in curvature. It is governed, at
least at the linear level, by (130), Le. the same relation that governs
resonance; 1JFb is predicted from uFb via (120). The wavelength of oscillation
can be predicted from (87a). Struiksma et ale [1985] present several successful
predictions of overdeepening in experimental channels, using a nonlinear model.
Johannesson and Parker [1989] demonstrate that overdeepening can be
predicted with equal or better success from the linear model herein; see Figure
6 in that paper.
The analysis of Blondeaux and Seminara [19851 clearly encompasses
overdeepening, although they did not comment upon the phenomenon in any
way.
Conclusions
The conclusions of this paper can be summarized as follows:
1.

The following dimensionless scalings are typically valid in natural


channels
r

0(1)

O( 1); {~

<1

Here r = 21rb/( {oX) is a dimensionless measure of alternate bar or


meander wavelength, r = /3/( 'rio) is a gravitational diffusion coefficient
in the relation for transverse sediment transport, {o = '}Cfo, where 'r is
the ratio of half-width to depth, and Cfo is a base flow bed friction
coefficient.
2.

For typical natural channels, the equation of transverse momentum


balance can be decomposed into a relation governing depth-averaged
transverse velocity, and a relation governing helical secondary flow.
The two are not directly coupled to each other. Helical secondary
flow may be taken to be purely curvature-driven, and unaffected by
bed topography at the linear level, whereas the depth-averaged
component may be strongly affected by bed topography. Secondary
flow may be neglected in a linear treatment of alternate bars.

3.

Alternate bar instability can be predicted even in a theory that


neglects as O( (2)
the linear inertial and frictional terms associated with
o
depth-averaged transverse velocity. The approximate theory provides
critical values of "y for the inception of alternate bars that are in
essential agreement with the more complete theory of Blondeaux and
Seminara [1985] and Colombini et ale [1987] (and thus with data). It
fails, however, at high wavenumber r, and thus can predict neither a
finite "critical" value of r corresponding to the lowest value of "y at
which instability occurs, nor a value of r associated with maximum
instability.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

409

Parker and Johannesson


4.

The same approximation, Le. neglect of o( (2)


terms associated with
o
depth-averaged transverse velocity, yields a linear treatment of bend
flow that is much simpler than, but differs in no essential way from,
that of Blondeaux and Seminara [1985]. Neither the neglect of o( (2)
o
terms, nor the approximation of the transverse structure of the "F"
problem as sinusoidal, Le. (12), affect the accuracy of the solution in
any substantial way. The method is elucidated in more detail in the
companion paper, Johannesson and Parker [1989].

5.

The theories of the Delft .school [Struiksma et al., 1985; Crosato, 1987],
the Genova school [Blondeaux and Seminara, 19851, and the Minnesota
school ["Johannesson and Parker, 1989] all describe in a similar way
both resonance and overdeepening. Prior credit for elucidation of the
former phenomenon accrues to the Geneva school; in the latter case it
accrues to the Delft school.

This research was funded by the National Science Foundation (Grant No.
MSM-8311721--Q02 and INT-8412678) and the Legislative Commission on
Part of the research was conducted at the Disaster
Minnesota Resources.
Prevention Research Institute, Kyoto University, Japan, and Genova
University, Italy, during the senior author's sabbatical leave. Sincere thanks is
extended to Kazuo Ashida and Giovanni Seminara in this regard.

Appendix
The exact solution to (120)
(45), Le., P = P t = 1, Mt = o.
and Ml is straightforward.
uF

Ut(n)cos4>

+ U2(n)sin4>

vF

Vt(n)cos4>

+ V2(n)sin4>

1]F

Et(n)cos4>

+ E2(n)sin4>

(126) is given as follows, for the case of


The generalization to other values of P, Pt

Here

an2+
r (r

dVt +
dV2
-3 an
r an

3 dV2

r dVt

an

U2

+ 9)

r (r 2 + 9)

Vip

Bit n 2

+ Bi2

=
i

1,2

1,2

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

R.esonance and Overdeepening

410

a1

D1

D2

r2

3(M-3)

= 2" (r2 + 9)
1

D1

= - 2" (r2 + 9)

2r

+ r [r

D2

2r

+ r [r

B l1 -

(r 2 + 6)

B21

+ (r 2 + 6)

In the above relations, D1 and D2 are given by (126).

B21]

B l1]

Furthermore,
i

aa = 21'(r2 + 4)

a4

jr 4

+ (M2 - 6M + 13)r2 + 4(M-3)2


as

r(M-1)

1,2

1,2

A*11

cosh

~1 COS~2

A*12

sinh

~1

Copyright American Geophysical Union

sin

~2

1,2

Water Resources Monograph

River Meandering

Vol. 12

411

Parker and Johannesson

Notation
dimensionless bed scour factor in (98)
dimensionless coefficient of dispersion in (99)
dimensionless coefficients of the "B", "C", and
solutions in (135), (113), and (132), respectively

"F"

channel half-width
dimensionless

coefficients

of the

"B",

"C",

and

"F"

solutions in (135), (113), and (132, respectively

C
Cf, Cfo
C

Ds
Dl, D2
F

f*
Go

b/ re;

dimensionless channel centerline curvat ure

bed friction coefficient, base flow bed friction coefficient


dimensionless streamwise migration rate
grain size
coefficients in (125)
= U/

JfJf,

base flow Froude number

order-one coefficient in (16)


function in (48) defining vertical structure of secondary
flow

gravitational acceleration

depth of the base flow

ii

local depth

ii/H, dimensionless depth

hp

h B , he' h F , h T

perturbed dimensionless depth


scaled dimensionless depth perturbations of the "B", "C",
"F", and "T" problems, respectively
near-bank value of he
coefficients
conditions

k
ks
M, M1

defined

in

(82)

that

vanish

at

21rb/ Xm; dimensionless wavenumber


roughness height in (43b)
dimensionless coefficients in the equation of sediment
continuity, defined in (42)

resonant

transverse coordinate made dimensionless with b

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Resonance and Overdeepening

412

P, PI
P
qo
qs, qn
qsp, qnp
R

rc
rm
r

dimensionless coefficients in the equation of downstream


momentum balance, defined in (42)
bed porosity
downstream volumetric sediment transport rate per unit
width of the base flow
downstream and transverse volumetric sediment transport
rates, made dimensionless with qo
perturbations of qs, qn, respectively
sediment submerged specific gravity
centerline radius of curvature
minimum centerline radius of curvature

kl

(0,

(90)
rr, ri
rres
s
t, t
T

U
u

rv

scaled dimensionless wavenumber; real except in


(94)

real, imaginary part of r in (90)


(94)
resonant value of r, as specified by (87a)
streamwise coordinate made dimensionless with b
time, dimensionless time
N

function specifying vertical structure of primary flow


depth-averaged downstream velocity of the base flow
depth-averaged downstream velocity made dimensionless
with U
perturbed dimensionless downstream velocity
scaled dimensionless streamwise velocity perturbations in

ucb

v*

the "B", "C", "F", and "T" problems, respectively


near-bank value of Ue
transverse depth-averaged velocity made dimensionless
with U

perturbed dimensionless transverse velocity


scaled dimensionless transverse velocity perturbation of the
"B", "C", "F", and "T" problems, respectively

coordinate upward normal from the bed


perturbation amplitude growth rate in (80)
coefficient in (16)
coefficient in transverse sediment transport relation (15),
defined in (16)

PI (1'(0);

coefficient of gravitational diffusion

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

413

Parker and Johannesson


b/H; ratio of half-width to base flow depth
amplitude of bed topography
b/H; dimensionless amplitude of bed topography
f, fo

(
TJ, TJo

f Cf, f Cfo

z/h; dimensionless upward normal coordinate


bed elevation made dimensionless with H, reference value
of TJ at s = 0
perturbation of TJ
scaled perturbations of TJ of the "B", "C", "F", and "T"
problems, respectively
centerline arc wavelength of alternate bar or point bar
coefficient in (16)
helical
component
of
transverse
velocity
made
dimensionless with U
perturbation of v
scaled perturbation of v of the "B" problem
water surface elevation made dimensionless with H,
reference value of at s = 0
of the "B", "C", "F", and "T"
scaled perturbations of
problems, respectively

(]'

TSO
TS

Tsp
T*

so
T*

G
T*
c

dimensionless
order-one
channel
C/ o;
curvature
dimensioned downstream bed shear stress
value of Ts of the base flow

centerline

downstream bed shear stress made dimensionless with Tso


perturbation of Ts
Tso/(p Rg Ds ); Shields stress of the base mode
grain stress component of T*
so

critical value of T*so


ks, streamwise phase coordinate
order-one coefficients defined in (10)
b/f m; measure of curvature amplitude

References

Beck, S. M., Lateral channel stability of the Pembina River near Rossington,
Canada, Report, Research Council of Alberta, Edmonton, Alberta, Canada,
1983.

Copyright American Geophysical Union

Water Resources Monograph

414

River Meandering

Vol. 12

Resonance and Overdeepening

Beck, S. M., D. A. Melfi, and K. Yalamanchili, Lateral migration of the


Genesee River, New York, Proceedings, River Meandering, Rivers '83
Conference} ASCE, New Orleans, 510-517, 1983.
Blondeaux, P., and G. Seminara, A unified bar-bend theory of river
mechanics, J. Fluid Mech., 157, 449-470, 1985.
Callander, R. A., Instability and river channels, J. Fluid Mech.} 36, 465-480,
1969.
Chitale, S. V., River channel patterns, J. Hydraul. Div.} Am. Soc. Civ. Eng.,
96(1), 201-221.
Colombini, M., G. Seminara, and M. Tubino, Finite-amplitude alternate bars,
J. Fluid Mech.} 181, 213-232, 1987.
Crosato, A., Simulation model of meandering processes of rivers, extended
abstracts, Euromech 215 Conference, Sept. 15-19, Genova, Italy, 158-161,
1987.
Dietrich, W. E., and J. D. Smith, Influence of the point bar on flow through
curved channels, Water Resour. Res., 19(5), 1173-1192, 1983.
Engelund, F., Instability of flow in a curved alluvial channel, J. Fluid Mech.,
72} 145-160, 1975.
Engelund, F., Flow and bed topography in channel bends, J. Hydraul. Div.,
Am. Soc. Div. Eng.} 100(HYll), 1631-1648, 1974.
Engelund, F., and o. Skovgaard, On the origin of meandering and braiding in
alluvial streams, J. Fluid Mech., 57, 289-302, 1973.
Fredsoe, J., Meandering and braiding of rivers, J. Fluid Mech., 82, 609-624,
1978.
Hasegawa, K., Research on an equation of bank erosion considering
non-equilibrium conditions, Proc. Japan Soc. Civ. Eng., 316(12), 1981.
Hickin, E. J., and G. C. Nanson, The character of channel migration on the
Beatton River, Northeast British Columbia, Canada, Geol. Soc. Am. Bull.,
86, 487-494, 1975.
Ikeda, S., Lateral bed load transport on side slopes, J. Hydraul. Div.} Am.
Soc. Civ. Eng.} 108(HYll), 1369-1373, 1982.
Ikeda, S., G. Parker, and K. Sawai, Bend theory of river meanders.
1. Linear development, J. Fluid Mech.} 112, 363-377, 1981.
Ikeda, S., and T. Nishimura, Flow and bed profile in meandering sand-silt
rivers, J. Hydraul. Eng., Am. Soc. Civ. Eng., 112(7), 562-579, 1986.
Johannesson, H. and G. Parker, Linear theory of river meanders, this volume,
1989.
Johannesson, II. and G. Parker, Inertial effects on secondary and primary flow
in curved channels, External Memorandum No. 208, St. Anthony Falls
Hydraulic Laboratory, Univ. of Minnesota, 1988a
Johannesson, H., and G. Parker, Secondary flow in a mildly sinuous channel,
J. Hydraul. Eng.} Am. Soc. Giv. Eng., in press, 1988b.
Kalkwijk, J. P. Th., and H. J. De Vriend, Computation of flow in shallow
river bends, J. Hydraul. Res.} 18(4), 327-242, 1980.
Kikkawa, H., S. Ikeda, and A. Kitagawa, Flow and bed topography in curved
open channels, J. Hydraul. Div., Am. Soc. Civ. Eng., 102(HY9), 1327-1342,
1976.
Kinoshita, R., Investigation of channel deformation in Ishikari River, Report of
Bureau of Resources, Department of Science and Technology, Japan, 174 p.
(in Japanese), 1961.
Kinoshita, R., and H. Miwa, River channel formation which prevents
downstream translation of bars, Shinsabo, 94, 12-17, 1974.
Kuroki, M., and T. Kishi, Regime criteria on bars and braids, Hydraulics
Paper, Research Laboratory of Civil and Environmental Engineering,
Hokkaido University, Japan, 23 p., July 1985.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

415

Parker and Johannesson

Langbein, W. B. and Leopold, L. B., River meanders-theory of minimum


variance, U.S.G.S. Professional Paper 422H, 15 p., 1966.
Leopold, L. B., and M. G. Wolman, River channel patterns:
Braided
meandering and straight, U.S.G.S. Professional Paper 282-B, 85 p., 1957.
Nelson, J. M., Mechanics of flow and sediment transport over nonuniform
erodible beds, Ph.D. thesis, University of Washington, 227 p., 1988.
Odgaard, J. A., Transverse bed slope in alluvial river bends, J. Hydraul. Div.,
Am. Soc. Giv. Eng., 107(HYI2), 1677-1694, 1981.
Odgaard, J. A., Meander flow model.
I: Development, J. Hydraul. Div.,
Am. Soc. Giv. Eng., 112(12), 1117-1136, 1986a.
Odgaard, J. A., Meander flow model. II: Applications, J. Hydraul. Div.,
Am. Soc. Giv. Eng., 112(12), 1137-1150, 1986b.
Ozaki, S., and T. Hayashi, On the formation of alternate bars and braids and
the dominant meander length, Proc. Japan Soc. Giv. Eng., 333(12),
109-117, 1983.
Parker, G., On the cause and characteristic scales of meandering and braiding
in rivers, J. Fluid Mech., 76, 457-480, 1976.
Parker, G., Stability of the channel of the Minnesota River near State Bridge
No. 93, Minnesota, Project Report 205, St. Anthony Falls Hydraulic
Laboratory, University of Minnesota, U.S.A., 33 p., 1982.
Parker, G. Discussion of:
Lateral bedload transport on side slopes, by S.
Ikeda, J. Hydraul. Eng., Am. Soc. Giv. Eng., 110(2), 197-203, 1984.
Parker, G., and A. G. Anderson, Basic principles of river hydraulics, J.
Hydraul. Div., Am. Soc. Giv. Eng., 103(HY9), 1077-1087, 1977.
Parker, G., P. Diplas, and J. Akiyama, Meander bends of high amplitude, J.
Hydraul. Eng., Am. Soc. Giv. Eng., 109(10), 1323-1337, 1983.
Parker, G., and E. D. Andrews, On the time development of meander bends,
J. Fluid Mech., 162, 139-156, 1986.
Smith, J. D. and S. R. McLean, A model for meandering streams, Water
Resour. Res., 20(9), 1301-1315, 1984.
Struiksma, N., K. W. Olesen, C. Flokstra, and H. J. De Vriend, Bed
deformations in curved alluvial channels, J. Hydraul. Res., 23(1), 57-79,
1985.
Thorne, C. R., L. W. Zevenbergen, J. B. Bradley, and J. C. Pitlick,
Measurements of bend flow hydraulics on the Fall River at bankfull stage,
WRD Project Report No. 85-3, National Park Service, Colorado State
University, Fort Collins, Colorado, 70 p., 1985.
Wolman, M. G., The natural channel of Brandywine Creek, Pennsylvania,
U.S.G.S. Professional Paper 271, 56 p., 1955.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Copyright 1989 by the American Geophysical Union.

Bar and Channel Formation in Braided Streams


Yuichiro Fujita

Disaster Prevention Research Institute, Kyoto University


Gokanosho, Uji, Kyoto, 611, Japan
The present state of fluvial hydraulics of braided rivers is reviewed, with
the main emphasis on research in Japan but with brief citations of research in
other countries.
The research reviewed includes field observations, and
empirical and theoretical studies on formative conditions and processes of
stream braiding.
In order to increase the knowledge of the hydraulic characteristics of
braided rivers, hydraulic experiments on the behavior of multiple row bars
leading to braided streams were conducted in three flumes of different widths,
0.5 m, 1.8 m and 3.0 m.
Results are described on changes in bed
configuration and stream bed variation, as well as hydraulic conditions. The
higher modes of bars were gradually replaced with lo\ver ones, and braided
patterns appeared in cases with small depths. A criterion for braided streanlS
and for the co--existence of multi pIe row bars of several modes is proposed
This co--existence and the
from a modification of that for alternate bars.
development of wave height and length of multiple row bars are discussed via
comparison with a previous linear stability theory. The decrease in mode and
the occurrence of stream braiding is elucidated, and proved to be predictable
from the development of bar height and the ceasing of sediment transport on
large bars.
Introduction
Braided rivers constitute a typical fluvial morphology, and multiple-row, or
high-mode sand bars are regarded as a general form of incipient streanl
Ineandering. Their complicated features have attracted many geomorphologists,
\vho have attempted to find out the governing factors of their formation.
Braided rivers also present difficult problems for river regulation to hydraulic
engineers because of the rapid, radical change in stream patterns followed by
bed and bank scour even during small floods [Hashimoto, 1956].
Thorough investigations on braided channels have been carried out in
several fields concerned with understanding landforms.
The main fields in
science are geology, sedimentology, and geography (including geomorphology),
\vhereas in engineering, which tends to have a rnore practical orientation, it is
civil or hydraulic engineering [Lane, 1955]. Both fields employ field studies,
laboratory experiments and theoretical analysis, and because they often share
common research goals, there is a need for researchers to be aware of the
work done in other fields.
In geological and sedimentological research, the primary goal is to
understand the controls on sorting and stratification such that the deposits of

417
Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

418

ancestral braided rivers contained in the geologic record can be recognized and
the fluvial processes that generated them can be properly interpreted.
Theoretical models that focus on prediction of stratigraphy generated over
geologic time spans tend to be greatly simplified and are of less value to
engineering applications.
Also, most studies provide only descriptions of
topographic change without quantitative analysis of the processes controlling
channel topography in braided rivers.
Geographers have studied braided channels through both laboratory
experiments and field work. Their investigations are often aimed at classifying
the relative importance of natural versus artificial changes in landscapes and
their findings have been very useful to river engineers. In Japan, for instance,
almost all braided channels exist on alluvial fans formed by ancestral braided
channels and geographers have pointed out that changes in braided channel
courses can be inferred from historical loci of human occupation.
Studies by hydraulic engineers of braided channels are focused on more
specific and practical problems. For example, the behavior of double row bars
were studied by Miwa [1980] because of their influence on the location of
intake structures for agricultural water use.
Mosley [1982] documented the
effects on a braided river of an artificial flood produced by discharge froln a
dam to evaluate the effects of a proposed power station on spawning potential
of the river. In order to examine flood hazards, several researchers have built
physical models based on field data on hydrological and hydraulic
characteristics of braided rivers. Recent examples in Japan are the studies of
the Shinano River [Fakami and Baba, 1977]' the Qloi River [Shizuoka Local
Construction Office for Rivers, Ministry of Construction, 1979], the Hii River
[Sakano and Yamamoto, 1981] and the Kurobe River [Ishikawa, 1983]. Even in
these model studies, however, prediction of changes and migration of bars has
been qualitative, and observed bed variation in the model, such as local scour
This has
depth, has not compared well with that found in the field.
Inotivated considerable effort to document scour in river channels [Ministry of
Construction River Bureau and Public Works Res. Inst., 1982]. At the saIne
time, basic studies on the hydraulic characteristics and formative processes of
braided channels have been initiated. Development of double and multiple rovv
bars in braided channels can be viewed as part of the general process of bar
formation that relates to river meander initiation as well.
In the following, the present state of Japanese research on braided channels
is reviewed with some citations of work done in other countries included.
Results of experiments conducted by the author in three flumes with different
\vidths are then presented to describe the process of bar formation and the
hydraulic characteristics of multiple row bars.

Research on Braided Rivers, Especially in Japan

Several Aspects of Braided Streams in Japan


Braided rivers are found on flat plains, in glacial valleys, on outwash
deposits, and on alluvial fans, and appear to vary in morphology with valley
slope, stream power, and amount and grain size of sediment load, as proposed
by Schumm [1981]. Although considerable morphologic variation occurs, any
channel that has two or more main streams across a single cross-section may
be classed as braided. Braided channels seem to occur when some specific set
of conditions are satisfied rather than in places defined only by hydrology,
geomorphology, or geology.
In Japan, most braided rivers flow over alluvial fans, hence they are apt to
be considered to exist only on alluvial fans.
In braided rivers on steep

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

419

Fujita

Fig. 1.

Surface waves in a flood flow in the Joganji River on an alluvial fan.

alluvial fans, such as the Joganji River and the Kurobe River in the Hokuriku
District, violent flows appeared during floods with great standing waves (Figure
1) [Tsuritani and Igarashi, 1971]). Large differences in water surface elevation
along opposite banks, similar to superelevation observed in river meanders
(Figure 2) [DPRI Group, I(yoto Univ., 1970], were caused by flow
concentration corresponding to bar development.
Flow concentration caused
severe bank scour even during medium floods, such as the rainy season flood
in 1978 in the Joganji River [Toyama Local Construction Office, Ministry of
Construction, 1978]. Ishikawa [1983] conducted a physical model test on the
Kurobe River and suggested that the banks of the low water course should be
protected to prevent erosion during floods of small to middle magnitude.

5.0

,.----r---r----r---,---,.----,---r:l6~~..,.-----,---r--~---r--~--,r----"""""

~.O

3.0
(I

Bank failure.
e.
o .

.-

: ::~::~

'952.7

0 1

-3.0

0'.

eo

.Og

eo
o

00

8
Bank fa; 1ure ..:..st.

-~.O

~~
- 5.00~--1.....0 - - 2......0---013.-0- .......
O--5.....0 - - 6......0~-~1.~0--a .....
0 - - 9.....0 -----'0....0 -...-1-01
. .-0--1-2.....
0--.-3.6-.0----'14.0
x( km

Fig. 2.

Spatial changes in the difference in water surface elevation at each


bank, ~H (positive right hand side higher than left), during several floods in
the Kurobe River.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

420

Thus, braided rivers on alluvial fans are very active, and deposition of bars
as the river freely shifts causes fan growth. [(adomura [1971] explained that
these bars and streams have in the past formed micro topography on alluvial
fans, which in turn influenced historical land use.
He also described the
development process of the Tenryugawa alluvial fan and the O'oigawa alluvial
fan in the Chubu District. Similarly, [(ayane [1971] surveyed the I(urobegawa
alluvial fan to show a location map of old sand bars and abandoned strean1
channels as a part of a hydrological study of water circulation in the fan.
Large variations in braided rivers have been reported in other countries.
Among them, the 110 km channel shift of the I(osi River from the east side
of an inland delta to the west in India is well-known [Gole and Chitale, 1966].
Since 1963, this migration was controlled by a sediment barrage.
Strean1
channel variations became intense in the upper aggrading reach of the barrage,
while downstream only minor variations occurred [Strivastava, 1983].
Tendencies toward significant channel changes were analyzed probabilistically
by Graf [1981] for 112 years of stream channel pattern for a 56.4 km reach in
the Gila River in Arizona.
Tendencies toward decrease in bar mode and
fixing of channel location in degradational reaches have been stressed in Japan
in recent years because these are accompanied by severe local scour of the
river bed and bank, and because degrading reaches have increased in length in
many rivers [Ministry of Construction, River Bureau River and Public Works
Res. [nst., 1982; Suga, 1983a,b]. Channel degradation is due to a reduction of
sediment supply as a result of large dams and development of sabo (debris
control) works constructed in mountainous area.

Mode-I.5

Fig. 3.

Schematic sketches of bar modes and arrangement of scour holes.

Braided rivers present complicated features with streams being divided into
several channels with submerged and emerged bars, depending on flood stage
as demonstrated in a field experiment conducted by Mosley [1982]. [(inoshita
[1957] showed that bars play a significant role in stream channel processes in
alluvial rivers and that a fundamental unit is an array of alternate bars
described as a single row (mode 1) of bars (Fig. 3). He also pointed out that
bed forms similar to alternate bars are discernible even in braided rivers,
which are composed of lateral repetitions of the alternate bar form, i.e.
lTIultiple rows of bars, bounded by many small streams (Fig. 3). Ikeda [1975]
documented bars of various modes in the Omoi, I(urobe, Oloi and Joganji
R.ivers and related these bars to large floods.
Church and Jones [1982]
explained braided streams as extensive bar assemblages in a kind of
hierarchical structure.
They classified bars into two groups, "hydraulic
element" bars and "storage element II bars.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

421

Fujita

Brice [1964] tried to express states of braided rivers quantitatively by his


index;" Howard et ale [1970] presented a topological approach to this
Fujita et ale [1986] proposed two simple and objective methods to

"braiding
problem.
judge the
geometry

modes (number of rows of alternate bar array) of bars with uneven


and entangled edge, including an intermediate state of bars [Ogawa

and Fukami, 1979].

Because the number and arrangement of scour troughs correspond to the


bar mode (Fig. 3), the mode can be determined from longitudinal changes in
the number of troughs scoured in a cross section. Namely, the mode m is
m = 2p - 1 when the number of scour holes, p, is constant in the
longitudinal direction, and is m = 2p if the number alternates between p and
p + 1 (Fig. 3). Moreover, mode 1.5 is defined as a state where mode 1 bars
are accompanied by small bars at opposite sides alternatively.
The mode
assignment for the number of troughs varying in the downstream direction is
based upon the observation that cross-sections averaged longitudinally show a
bell-shape in channels with alternate bars. The height of the bell shape is
proportional to the height of the alternate bars [Fujita and Muramoto, 1985].
The averaged cross-section with multiple rows is expected to show the same
number of peaks as the order of the bar mode.

Observations on Stream Channel Processes in Braided Rivers


Field observations on braided rivers have been carried out by
sedimentologists rather than hydraulic engineers in other countries, hence
sedimentary processes, such as sorting and the formation of deposits in braided
rivers have mainly been described; fluvial processes have not been analyzed in
detail. The investigations by Williams and Rust [1969] on the Donjek River,
Yukon, Canada, and by Cant and Walker [1978] on the South Saskatchewan
River in Canada, were done from a sedimentological standpoint, but also
describe in detail bar and channel formation processes in braided rivers.
Similar studies in sedimentology are rare in Japan.
One of the few examples of field research on braided rivers from a
hydraulic engineering perspective is that by I(inoshita [1978] on the O'oi River.
He took pictures of the river continuously during floods from a hill about 90
m above the channel.
From these pictures, he mapped changes in bed
configuration and documented bar behavior during floods. This was done for
the purpose of designing physical model tests to assess the probable effects of
artificially widening a narrow reach upstream of the braided zone.
Fairly
regularly spaced bars were observed Inigrating downstream, deforming not only
their shapes but also their array pattern during floods. His observations were
based primarily on features that were visible in a small part of the channel
during the flood recession, and bed configurations preserved after each flood.
Sunada [1985a,b] surveyed bed variation in a double row bar reach in the
I(amanashi River, Yamanashi, Japan, during several floods; he compared the
results with those from an experimental flume.
He pointed out several
problems regarding the correspondence between the field observations and the
experiments. Most of these problems seem to stem from a lack of knowledge
about the fundamental characteristics of braided channels.

Formative Conditions of Braided Rivers and Double to Multiple Row Bars


Many theories for the formative conditions for braided rivers and
double/multiple row bars have been proposed, although often these theories are
not based on specific physical processes.
Such theories include not only
stability models of the bed variation in two dimensions, but also theories

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

422

which employ some generalized laws, such as minimum energy dissipation or


minimum stream power. The latter theories, however, fail to predict specific
channel pattern formations and their controls and, consequently, have seldom
been used in Japan. Also, most investigations express the formative conditions
using governing factors obtained either empirically or from similitude
(dimensional analysis).
Empirically-derived formative conditions are often expressed by regime-type
equations, such as (1), which are based on the empirical result that the
formation of channel pattern in braided rivers, is governed mainly by valley
slope, discharge and sediment material [Schumm, 1981].

I = f (d) Qa

(1)

Here I is the valley slope, d is representative grain size, Q is discharge and a


is an exponent.
Such equations have been reported for more than thirty
years, by such authors as Leopold and Wolman [1957], Lane [1957]' Henderson
l1961], Ackers and Charlton [1970], Chitale' [1973], and Muramoto [1976].
Except for Henderson, these authors included only valley slopes and discharge
in (1). Recently, similar investigations have been published which attempt to
support these regime equations [Osterkamp, 1978, Begin, 1981]; others have
disputed the validity of the regime equations [Carson, 1984].
Osterkamp examined f (d) statistically, grouping data from 76 rivers
according to sinuosity - degree of braiding, and grain size gradation. He could
not, however, express the channel pat tern conditions distinctively. Similarly,
but in more detail, Begin tried to confirm the physical meaning of (1) using
359 data points and a regime equation for depth. Braided streams occurred
under rather high shear stress conditions, and despite his intention, his
research proved the invalidity of regime expressions for defining the formative
conditions of braided channels. Carson pointed out that the regime equation
approach was not valid because the previous equations expressed only
relationships for incipient motion of the most frequent sediment size in the
data set used.
Expressions of regime type would be very useful if they were valid because
discharge can be estimated fairly easily from precipitation and slope, and bed
luaterial size can be measured in the field. Similar investigations have rarely
been reported in Japan because natural and historical circumstances
surrounding rivers differ from those in countries where the regime equations
were developed and, subsequently, Japanese researches recognize deficiencies as
suggested by Carson.
Formative conditions of braided streams in diagrams with non-dimensional
parameters were first proposed by Ikeda [19731. His criterion, equation (2) was
derived from the same similarity concept o( distorted models that Sukegawa
[1972] used to distinguish between governing conditions for micro- and mesoscale bedforms.

(2)
where B is the full channel width, h the mean depth, U* the shear velocity
and U*c the critical shear velocity.
[(uroki et al. l19751 modified Ikeda's
diagralu adding many flume data.
Tamai et al. 11978l, who conducted
experiments on the formation of braide~ streams and double row bars,
modified the parameters to include hydraulic resistance. Though the formative
conditions in cases of rather steep slope are expressed fairly well in these
diagrams, the basic concept of distorted models has little physical meaning.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

423

Fujita

Therefore, Muramoto and Fujita [1977, 1978] applied dimensional analysis to


this problem and extracted three governing parameters, so as to eliminate
overlap of channel geometry. The effect of slope I on formative conditions
proved to be small when I is less than 1/20.
By means of a detailed
examination of the BI d - hi d plane, the condition for braided stream
formation is found to be,

(hi d)

(BI d)

2/3

< 0.15 for

1 < U*

I U*c
2

< 12

(3)

Yamaguchi and Okabe [1981] inspected river data in Hokkaido Island and

proposed a tentative expression for hydraulic characteristics of braided rivers as


well as rivers with alternate bars.
By using Ikeda's diagram, Shizuoka Local Construction Office for Rivers,
Ministry of Construction l1980] carried out a physical model test and produced
bedforms very similar to those observed in the O'oi River. Such similarity
between models and field observation was also shown by Ikeda [1982], in
experilnental flumes with different scales. Miwa [1983] proposed a dimensional
analysis of this similarity, assuming there existed "a (some) force(s)" which
yields such a coincidence between prototypes and models. His result, however,
included the Froude number as a main parameter, although I(inoshita
discarded it in his analysis.
Moreover, his diagram predicted forlnative
conditions for braided streams that were opposite to the observed relationships
in the field data presented by Fukami [1979].
Recent stability theories on linearized river bed variations have been able
to produce definite criterion for bar formation.
In this context, the most
important ideas for braided stream and bar formation were those proposed by
Engelund and Skovgaard [1973], who emphasized the transverse slope effects on
bed trans~ort and development of the bar modes. Following in their footsteps,
Fredsoe l1978] showed SOine examples of formative criteria both with and
without suspended sediment.
Combining their ideas with contributions of
others, [(uroki and I(ishi [1984] succeeded in obtaining expressions for the
cri teria for formation of meso-scale bed configurations, including flat bed
braided streams, without suspended sedilnent and under steady flow.
More
recently, stability theories in non-linear form have been developed to predict
not only a criterion for bar formation similar to the earlier theories, but also
the final shapes of bars [Fukuoka and Yamasaka, 1985; Fukuoka et al., 1986;

Colombini et al., 1987].

In contrast. to the above theories for the initiation of bars, Bettes and
White [1983] elucidated stream braiding from the final state of the strean1

channel in equilibrium. According to them, if the valley slope is steeper than


that for v/hich the channel is just stable when it is carrying one-half the
water discharge, two channels can be presumed to appear. They calculated
hydraulic conditions for a criterion for braided streams involving the
appearance of three channels, and showed reasonable agreement between
observations in rivers and those in an experimental flume.
Though this
approach is very interesting, individual stream channels in braided rivers do
not seem to be identical to a stable single-thread channel, and meandering
channels with high sinuosity probably cannot be explained by their theory.

Formative Processes of Braided Streams


In order to grasp the fundamental characteristics of braided bar formation,
it is very useful to observe the process of development from a flat initial
state, under simplified conditions.
Because this process has rarely been

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

424

observed in rivers, it must inevitably be studied in flume experiments.


Experiments on braided streams have been conducted both in laterally erodible
channels and in channels with fixed sidewalls.
Leopold and Wolman [1957] showed that a central bar in a channel with
erodible banks would emerge causing local channel widening, and then the
Ikeda [1973] examined braided stream
evolution of a braided stream.
formation in channels with erodible banks. He observed that initially, channels
widened and alternate bars developed and migrated downstream. When the
channels became fairly wide, small-scale bars formed along the edges of the
alternate bars. These new bars grew, causing meandering and widening of the
channel. Eventually the bars emerged and the stream became braided. A
similar process was described by Ashmore [1982], who emphasized that the
same processes were observed by Williams and Rust [1969] and Cant and
IValker [1978]. Based on the results of large-scale experiments, Fujita and
Muramoto [1984] presented a criterion governing the formative process of
channel patterns.
Therein they proposed that a braided channel appeared
when channel widening caused water depths to become too small to maintain
Hong and Davies [1979] claimed that
the development of alternate bars.
topological quantities in their small experiments had the same characteristics
as those measured in the Rakaia River in New Zealand.
As noticed in field observations and in flume experiments, the development
of double to luultiple row bars is a distinctive feature of braided streams. It
is important to clarify the hydraulic characteristics of these bars in order to
predict the behavior of braided rivers.
Channels with fixed sidewalls are
suitable for the purpose of conducting experiments on bar formation.
Braided stream and/or bars in high modes were observed as a part of
experiments conducted on meso-scale bedform development [I(inoshita 1962,
[(ondo and I(omori, 1974].
In experiments by Tarnai et ale [1978],
double-row bars often turned into single-row bars for a while, and hysteresis
of bedform change was observed when the bed material was poorly sorted
sand.
The Shizuoka Local Construction Office for Rivers, Ministry of Construction
f1979l car~ied out m~ny preparatory ~xperir~ents and observed the be~~vior. of
hal's In hIgh luode, In order to obtaIn baSIC data for model tests [I\ znoshzta,
1980]. It was concluded that:
1) sand is better than crushed coal as regards
the experimental formation of bedforms, 2) bars developed in time and in the
downstream direction, 3) bar development became complicated in cases of
lnultiple bars because of changes in the transverse direction, and 4) bars
decreased in mode from 4-5 to 2 during the experiments. Such findings are
interesting and useful for future studies. An overlap of formative conditions
for bars in various modes was also observed by Miwa [1984], who examined
the characteristics of 20 separate runs under varying conditions in order to
calculate the bar mode first formed. In his investigation, the coexistence of
bars in various modes was noted, but their change in time was not described.
Such chanl$es in time of bars in various modes were studied in detail by
Fujita et ale l1986, 1987, 1988], using three flumes with different widths, as
reported below. They showed that individual bars of the double row mode
(mode 2) have the same geometrical characteristics and development titne
as those of alternate bars, and that development time Te and wave height, ZB
could be estimated by (4) and (5), respectively.

(4)
0.0051 (BB/h)

2/3

(h/d)

-1/3

/(1 - u~)

Copyright American Geophysical Union

(5)

Water Resources Monograph

River Meandering

Vol. 12

425

Fujita

Here lB and BB are the wavelength and width of individual bars, qB is the
bedload per unit width, and u' is a dimensionless grain velocity defined such
9
that value of (1 - u~) approach the value U*c/ U* when the effect of the
transverse bed slope is neglected. Fijita et al. also proposed a criterion for
the formation of neglected multiple row bars and braided streams which allov/s
for the coexistence of bars. They also elucidated the controls or the tilne
required to decrease bar mode, and to form the lowest mode of bars.
Finally, along with the development of numerical analyses, direct
calculations of two dimensional evolution of river bed pattern have been done
using the shallow water equations. In particular, Shimizu and Itakura [1986]
carried out a numerical analysis of the development of braided streams using
as an initial condition a small hump on the bed near a side wall. They also
performed calculations of bed evolution in the Ishikari River during an actual
flood.

Experiments on Formative Process of Multiple Row Bars


and Braided Streams

Apparatus and Procedures


Three series of experiments were carried out in three flumes of widths of
0.5, 1.8 and 3.0 m respectively in the Ujigawa Hydraulic Laboratory, Disaster
Prevention Research Institute, Kyoto University.
The experiments are
identified as Series A, B, and C, respectively.
The runs of series A were carried out in an 18 m reach in 50.2 em wide
steel flume 20.5 m in length with an adjustable slope. A motorized carriage
was used for grading and for holding instruments. Water was supplied to the
upstream end from a tail box through a tank used to measure discharge.
During the adjustment of discharge, water from this tank was bypassed
through a trough.
A reach 2 m upstream of the experimental reach was
drained by a small pUlllp as soon as the water supply was stopped. Sand was
not supplied at the inlet during the experiments because local deposits of this
sand inevitably have a great influence on the occurrence of sand bars in such
shallow flows. A servo-type water gauge was used for measurements of water
elevations along three longitudinal lines 15 em apart just after the beginning
and just before the termination of each run.
The bed topography was
measured along nine longitudinal lines 5 em apart with an optical sand surface
detector.
Continuous plan pictures of bedforms were taken from a camera
attached to the instrument carriage. The output from the measurements of
bed and water surface topography was recorded continuously on an X-V
recorder.
Calibrations for the absolute difference between water elevation and bed
level measurelllents were found to be insufficient. Therefore, almost all the
runs of series A were repeated in order to obtain accurate measurements of
hydraulic quantities, and to confirm the reproducibility of the experiments. In
this case, measurements of elevation were made with a point gauge at 1 III
intervals. Because the measurement of water surface elevations required the
duration of the run to complete, bed levels were interpolated to coincide with
the time of water level measurement by using measured values of the bed at
the initial and final state.
The runs of series B were carried out in a 3 m wide, 43 m long channel
partition within a 7.5 m wide, 243 m long concrete flume [Fujita and
Muramoto, 1982b]. Water was supplied through a measurement tank attached

Copyright American Geophysical Union

Water Resources Monograph

426

River Meandering

Vol. 12

Bar and Channel Formation

to a triangular weir situated in the center of a regulating basin for the


experimental reach.
At the downstream end of the experimental reach, an
abrupt drop was provided; a full-width steel wall prevented overflow of sand
from the reach. The stream bed was smoothed with a3 m blade attached to
a motorized sand scraper which traveled along rails laid on the concrete
sidewalls. Some sand was manually fed in run B-3 to prevent the formation
of an armor coat. In run B-4, the same amount of sand that was trapped in
an enclosed area just below the downstream end was fed into the upstream
end. The runs were terminated when overall bed change became small. The
measurements taken were similar to series A, but were done continuously in
the transverse direction, at 2 m intervals downstream for water stage, and at
1 m intervals for bed topography. As in series A, instruments were mounted
on a motorized measurement carriage.
In run B-4, where the prescribed
conditions were almost identical to run B-2, the measurement was programilled
to stop several times in order to observe changes in bed configuration and to
measure bed evolutions in detail.

Fig. 4.

A schematic sketch of the experimental flume used for runs C-l to 11


(1: measurement tank; 2: traction motor; 3: measurement carriage; 4: supply
pipe; 5: sand trap; 6: camera).

The runs of series C were conducted in a 1.8 m wide, 15 m long, 0.2 III
deep flume, which was made by widening the middle 15 m reach of the flume
used in series A (Figure 4).
A mobile carriage was used for lllounting
llleasurement devices, and a camera was mounted on the carriage. The control
and measurement of discharge were accomplished as in series A except for
addition of another pump. Sand was laid to a thickness of 6.5 em, between a
point 0.7 m from the upstream end and a point 0.2 m from the downstrealll
end. In a 0.9 m reach at the upstream end, a steel plate the full width of
the flume was buried and pulled up automatically in order to maintain a
constant water stage. The plate was returned to the channel bottom at the
end of each run after bed surface measurements; the hole was filled wi th sand.
When this sand supply was insufficient, a suitable amount of sediment was
injected manually. Control of the measurement carriage and data collection
system were automated as much as possible.
Water stages and bed
topography were measured continuously in the transverse direction at 1 m and
0.5 m intervals respectively according to prescribed programs.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

427

Fujita

100

t1~</r--

,80

~.
I

F (%)

,. ~RUn-C-'/............
I

Run-A

/
/

40

20

/"

o --- :~
0.1

Run-B

/1

60

Fig. 5.

II

J'

~r;
y

(mm)

10

Grain size distributions of the sands used in the experiments.

Experimental Conditions and Hydraulic Quantities Measured


The bed materials used in series A and C were selected so that grain
motion could be observed easily, and so that grains would move actively even
with small depth. A uniform sand with drn = 0.99 mm and d50 = 0.92 lTIn1
was used in series A; sand with drn = 1.05 and d50 = 0.92 mm was used in
series C. In series B, a rather widely graded sand was used such that drn and
d50 were 0.88 and 0.61 mm, respectively.
The grain size distributions are
shown in Figure 5.
These experimental conditions listed in Table 1 were
chosen to cover broadly the range of formative conditions for braided streams
predicted by (3).
In series A, the critical condition between alternate bars and braided
streams was studied using the parameter in (3) as a guide in order to observe
the overall aspects of the formative process of braided streams. In series B,
values of various parameters were selected to be almost identical to the mean
values obtained for groups of runs in series A where similar bed configurations
appeared. In series C, the experimental reach was short compared with the
width. In cases of bar formation at higher modes, this length was concluded
to be sufficient because wavelength was found to decrease inversely with bar
lTIode.
For series A and B, the experimental conditions were chosen to
generally cover a region where (hi d)(Bj d)2/3 < 0.07.
For a few values of slope I, and of depth h, as selected from values of
(hi d)j(BI d)2/3, the discharge and bedload transport rate "vere evaluated so as
to characterize flow resistance and bedload functions.
In all experiments, profiles of cross-sectionally averaged water surface and
bed elevations showed degradation in the extreme upstream reach in cases of
insufficient sand supply. Changes in the mean slope were, however, negligible
because of the generally steep slopes. Longitudinal changes in water surface
followed the bed profiles well, hence the changes in the mean depth were also
small, and hydraulic quantities averaged in the longitudinal direction could be
used to represent experimental conditions. These quantities are also listed in
Table 1. The depths are generally small, so as to produce the large ratios of
width to depth characteristics of braided streams. Due to the small depths,

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

~
~

00

Table 1.
Time
Run

No.
A-I
A-2
A-3
A-4
A-5
A--6
A-7
A-8
A-9
A-10
A-II

Flume
Width
B

(cm)

Prescribed conditions and hydraulic quantities measured in the experiments.


Bed Dis- Mean
Slope charge Depth

h
(l/sec) (cm)

Mean
Velocity
V

(em/sec)

Velocity
Froude Energy Shear
Slope Velocity Coeff.
No.

Fr

Ie
(10-

U*
3)

U/U*

Manning's
Roughness
n

(em/sec)

30'00"
45'00"
15'00"
18'00"
16'53"
15'00"
20'00"
13'30"
18'15"
8'02"
10'05"

50.2
50.2
50.2
50.2
50.2
50.2
50.2
50.2
50.2
50.2
50.2

1/50
1/50
1/50
1/50
1/50
1/50
1/50
1/50
1/25
1/25
1/25

0.65
0.45
0.45
0.30
1.00
1.20
0.75
2.50
0.50
0.30
0.75

0.37
0.30
0.54
0.72
0.44
1.37
0.35
0.28
0.37

24.23
19.92
36.89
33.20
33.96
36.35
28.46
21.34
40.38

1.27
1.16
1.60
1.25
1.64
0.99
1.54
1.29
2.12

19.87
19.78
19.61
19.61
19.39
19.45
37.85
37.83
37.76

2.67
3.40
3.19
3.67
2.87
4.98
3.58
3.21
3.61

9.09
8.32
11.57
9.05
11.85
7.31
7.96
6.66
10.99

0.0138
0.0145
0.0115
0.0154
0.0109
0.0212
0.0156
0.0178
0.0114

B-1 4hr40'
B-2 3hr40'
B-3 11hrOO'
B-4
24'
1hr26'
2hr23'
2hr51 '
4hr04'
4hr37'
5hr40'
6hr26'
8hr40'

301.
301.
301.
301.

1/200 9.90
1/200 30.75
1/200 14.95
1/200 30.20
31.50
30.00
30.30
30.30
29.70
29.70
29.10
28.50

1.24
2.62
1.41
2.58
2.30
2.54
2.52
2.40
2.67
2.59
2.55
2.57

26.52
38.99
35.23
39.34
45.71
39.41
40.18
40.47
37.12
38.40
38.36
37.40

0.83
0.77
0.95
0.79
0.97
0.79
0.81
0.89
0.73
0.77
0.77
0.76

5.01
4.90
4.79
5.12
5.17
5.26
5.32
5.25
5.19
5.29
5.30
5.02

2.43
3.52
3.56
3.54
3.37
3.58
3.58
3.46
3.64
3.62
3.59
3.48

10.90
13.74
11.09
11.14
13.58
11.07
11.30
12.32
10.29
10.66
10.77
10.83

0.0144
0.0157
0.0114
0.0157
0.0126
0.0158
0.0155
0.0141
0.0172
0.0165
0.0165
0.0164

Copyright American Geophysical Union

to

""'1

0(1

:::::r

~
~

~
~
0
""'1

S
~
c:"+-

o
~

Water Resources Monograph

C-1 1hr41'
2hr10'
C-2
31'
1hr07'
C-3
23'
1hr11 '
2hr14'
3hr15'
C-4
30'
1hr09'
C-5 1hr15'
1hr53'
C--6
36'
58'
C-7
11 '
1hr15'
C-8
46'
1hr20'
C-9
14'
45'
1hr12'
C-1
18'
40'

River Meandering

180

1/100

5.00

180

1/100

4.50

180

1/100

4.00

180

1/100

7.50

180

1/100 10.00

180

1/50

4.50

180

1/50

9.00

180
180
180

1/50
1/50
1/50

4.00
10.50

180

1/33

3.50

1.00
1.12
0.86
0.70
0.92
0.86
0.92
0.92
1.42
1.54
1.68
1.46
0.98
0.95
1.06
0.99
0.82
0.78
1.50
1.28
1.33
0.70
0.84

28.77
25.80
29.15
36.05
24.10
26.24
24.48
24.71
29.04
26.77
33.19
38.01
25.88
26.85
46.18
51.59
27.30
28.82
39.33
46.64
44.52
28.16
23.68

0.93
0.79
1.01
1.39
0.80
0.92
0.82
0.84
0.78
0.68
0.82
1.00
0.85
0.90
1.44
1.72
0.97
1.06
1.03
1.34
1.25
1.09
0.84

Vol. 12

9.55
9.64
10.27
9.90
10.15
10.48
9.61
10.18
10.43
9.93
10.04
9.14
19.37
19.64
24.34
19.82
19.54
18.97
19.53
18.26
17.92
30.92
30.30

3.00
3.20
2.91
2.57
2.99
2.93
2.91
2.97
3.76
3.82
3.99
3.57
4.28
4.22
4.90
4.25
3.94
3.79
5.30
4.70
4.76
4.58
4.94

9.94
8.35
10.18
14.37
8.16
9.10
8.51
8.51
7.84
7.15
8.47
10.72
6.12
6.42
9.86
13.20
6.98
7.73
7.47
10.16
9.53
6.23
4.87

0.0157
0.0188
0.0145
0.0100
0.0182
0.0162
0.0175
0.0178
0.0204
0.0228
0.0196
0.0149
0.0245
0.0234
0.0162
0.0125
0.0208
0.0188
0.0213
0.0157
0.0167
0.0229
0.0301

c.
:;:
~

~
~

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Forluation

430

the values of hydraulic resistance show some scatter. Deviation of the energy
slopes from the initial bed slope, however, is small.
In series A, hydraulic quantities were calculated from the repeated
experiments. Although the resistance in a few runs differed somewhat from
estimated values because of very small depths, all of the values of Froude
number, Fr, were greater than unity and were consistent with the observed
upper-regime flow. The smallest depth was 0.28 cm in run A-10; the mean
aspect ratio was about 180. In the other five runs, the depth was smaller
than 0.5 Clli, and the aspect ratio exceeded 100. Values of U* 2.7 to 3.7
cm/sec, Le. 15 to 60 percent greater than the critical value, of U*c = 2.33
cm/sec.
In the runs of series B, hydraulic quantities were evaluated at 2 m
intervals.
Except for the extreme upstream reach, water surface levels
deviated little from those at the initial stage, and neither wavy fluctuations
nor a trend of bed level variation are distinguished clearly. Deviations fro111
the initial profile were local and small, implying no systematic change due to
bar development.
Because the channel width of 3 m allowed the use of
relatively large depths, measurement errors became small, and calculated
hydraulic resistance was almost the same as the sand roughness. In run B-1,
the shear velocity was nearly equal to the critical value of U*c = 2.19 cm/sec,
whereas it was larger by a factor of 1.6 in run B-2 and 1.2 in run B-3. The
aspect ratios of 180 and 120 agreed with those in the corresponding runs of
series A. The Froude numbers of 0.7 to 0.9 were consistent with the observed
transition regime.
In series C, hydraulic quantities were calculated at the same cross-sections
as those where water elevation was measured, using continuous data records
stored on floppy disks. In some runs with small depth, periodic changes in
depth appeared in the late stages; plan pictures of bedforms were found to
reflect the existence of bars. Areas of small depth were located at the front
of emergent bars. Though scatter of the mean depth tended to increase with
slope, its magnitude was within 30% for mean depths of 0.6 to 1.5 cm. This
small scatter produced reasonable estimates of velocity and energy slope. Fr
varied between 0.8 and 1.0 for all experiments, corresponding to the transition
regime observed in runs with small discharge. In contrast, in runs C-7 and
C-9 with large discharge, variations of Fr became large due to the bar
development.
Fluctuations of U* also were small, Le. within 0.5 cm/sec.
Absolute values of U* ranged from 2.5 to 5.5 em/sec, and were one to two
times the critical value, U* = 2.4 em/sec. Flow resistance also varied within
a reasonable range in spite of the formation of large bars. Averaged quanti ties
also showed no trend in time.
No systematic changes in mean hydraulic quantities were discerned in
longitudinal direction or in time, but various bedforllis developed, and in SOllie
runs braided streams appeared with clearly--emerged bars. Specific features are
described in the following section.
Specific

Featur~

of Changes in Bed Configuration

Run A
In the initial runs A-I and A-2, the flow inlet device caused flow to enter
at an angle to the flume walls. This influenced the whole experimental reach;
in run A-I it produced a single-thread meandering stream.
In run A-2,
which carried reduced discharge, scour troughs appeared at the center and at
both sides repeatedly in the downstream reaches, where the influence of the
inlet condition was small.
Run A-3 was carried out under the same

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

431

Fujita

conditions as run A-2 until T=15 min (T being the experimental time
elapsed); very similar bedforms were observed.
In run A-4, with a small
discharge close to incipient sand motion, the sand surface changed very slowly.
First, two scour holes appeared in each cross section on almost the entire
reach, and then in some reaches the stream eroded the banks around the bars
to form a meander.
In runs A-5 to A-7, discharge was increased.
In run A-5, two bars
formed in every cross section. In some reaches this consisted of one large and
one small bar; in other reaches the bars were almost the same size. Similarly,
two large and small bars appeared in run A-v. There were one to three scour
holes and two to three bars, each different in size, in run A-7. The numbers
of these bars and scour holes did not change radically during the run.
In run A-8, bedform characteristics near the critical condition between
alternate and double row bar formation were explored. Following the initial
development of antidunes, double-row bars became clear at T = 3 min. At T
= 5 ll1in, 30 sec, oblique edges peculiar to alternate bars were discerned on
the water surface: they reached the downstream end at T = 8 111in, 30 sec.
At T=13 min, 29 sec, well-developed alternate bars with a wavelength of 1.7
m occupied the reach downstream of x = 8 m (x is the longitudinal
coordinate), but double row bars persisted upstream. At the final stage, the
edges of alternate bars were observed even in the most downstream reaches.
The channel slope was raised to 1/25 in runs A-9 to A-II. In run A-9,
the water surface presented a pattern corresponding to 2 to 3 rows of bars at
T = 3 min; subsequently the flow began to concentrate into deep areas, which
produced emergent bars at T = 9 min. The rather small discharge in run
A-I0 yielded a pattern of high mode bars of 4 to 5 rows at T = 1 min 45
sec, immediately after the beginning. These evolved into complicated braided
streams, with emergent bars in the upper reach. The downstream reach was
characterized by regularly ordered bars in 2 to 3 rows, with a wavelength of
about I m. In the final run, A-II, the mean depth was almost same as run
A-8; single or double rows of bars of small height and uncertain type
occurred.

Run B
The aspect ratio of run B-1 was intended to coincide with that of run A-4
and A-10, where the highest bar mode was observed. After the beginning of
the run, helicoidal flow with longitudinal streaks, discerned clearly by clouds of
fine suspended particles, immediately dominated the whole bed. The low shear
stress, almost critical, did not cause any change in the bed subsequently.
Because the bed became coarse at the upstream reach, the experiment was
terminated.
In run B-2 (Figure 6), the aspect ratio was similar to that of runs A-5
and A-6, which produced double row bars. The bed was initially covered
with streaks and helicoidal flow similar to those in run B-1. At T = 15 min,
the upstream reach was covered with small water surface waves for a 5 to 10
m length.
Below the waves, a feature that looked like a bar edge was
observed. The reach covered with surface waves migrated downstream, leaving
slight bar edges suggesting that the turbulence caused by the surface waves
produced double row bars. The first bar edge in the most upstream reach had
an alternate bar-like pattern from left to right, but it was divided into two
parts near the center, bifurcating a new edge to the left-hand side wall (Fig.
6). The edges of the double row bar then propagated downstream, repeatedly
forming bar edges from the center to both side walls, and then back to the
center (Fig. 6). At the same time, bars developed from upstream, approaching

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Forlnation

432
rOhr3Z-40'

:Ihl~
rh46'-52'

:~:

:1

2h42~4~

o
Fig. 6.

40m

Plans of formative process of double rows bars in run B-2.

a regular, symmetrical form with a wavelength of 6 to 8 m. At the final


state T = 3 hr, 30 min, the bed was scoured deeply along the edges, whereas
no bedforms were recognized at the upper reach from x = 15 m.
Longitudinal streaks also occurred clearly in run B-3, and the streamlines
meandered slightly by T = 1 hr. At T = 3 hr, 40 min, a bar edge pattern
appeared downstream from x = 8 m at an interval of about 10 m. The flow
began to form a main stream channel, by flowing around developed bars,
The
which were in turn deformed into complicated patterns by the flow.
edges of these emergent bars were divided into two or three portions by small
streams. The main flow formed a meandering stream of single thread by the
end of the run, at T = 9 hr, 40 min.
The formation of bars of higher mode than 3 could not be seen in run
B-2.
The reproducibility of experiments was examined in run B--4 by
interrupting the flow six times to inspect the bed In this run, similar surface
waves occurred in the upper reach, and very thin bars of mode 4 or so were
discerned at early interruptions, as shown in Figure 7. After that, wavelength
and height of bars as well as bar width increased, while bar mode decreased
to 2 at T = 4 hr, 15 min, which coincided with run B-2. However, uneven
development of both side bars produced irregular bars of mode 1 to 1.5 that
partly emerged, forming an almost single meandering thread of main flow.
These differences from B-2 might be partly ascribed to a larger amount of
sand supply at the inlet.
Run

In run C-1, many lines crossing the channel obliquely covered the whole
water surface at T = 5 min. These remained in the downstream reach for a
period, whereas in the upper reach from x = 5 m, they were deflected by bars

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Fujita

Vol. 12

433

Fig. 7.

Plans of formative process of bars of various modes in run B-4.

to form scale-like patterns (squamation) with round edges of mode 5. After T


= 20 min, bars of mode 3 to 4 predominated in the whole reach except for
the most upstream portion, where linguoid bars of mode 2 appeared. These
bars of mode 2 occupied the whole bed at T = 1 hr, 34 min.
In run C-2, bars in squamation of mode 8 to 10 that appeared just after
the beginning altered to those of mode 3 to 4 in irregular pattern at T = 30
min. Furthermore, they evolved into mode 2 from T = 45 min to 1 hr. At
T = 2 hr, 25 min, a bar pattern of mode 4 prevailed again; these bars were
compounded from bars of mode 2 forming braided patterns.
In run C-3 (Figure 8a), bars of mode 6 to 7 were arranged in neat
squalnation during an early stage. They changed to mode 3 and 4, but the
edges of the mode 3 bars became irregular because the flow was concentrated
into scour hollows developed along the faces of the bars. Once bars of mode
2 formed on the bed, a single, large thread of flow meander occurred, similar
to that in run B-3. Several small streams divided the emerged bars, and the'
channel pattern was concluded to be a braided stream.
In run C-4 (Figure 8b), the stream bed had been covered with bars of
mode 6 to 7 in squamation by T = 7 to 8 Inin. They altered to clearly
arranged bars of mode 5 with a wavelength of about 1 m at T = 30 n1in.
They persisted even at T = 1 hr at a reach upstream of x = 6 m, while
double row bars of mode 2 developed downstream. Because this state of bars
lasted to the end of run and no bar emerged, it was concluded that no
braided stream appeared in this run.
In run C-5 (Figures 8c and lOa) a discharge of 10 f/sec (maximum among
the experiments) was supplied with a slope of 1/100. The bedforms in the
early stage were the same as in run C-4, and they lasted throughout the run
in the most upstream reach.
Except for the upstream reach, symmetrical,
regular bars of mode 4 appeared in the whole reach at T = 30 min and
remained in the middle to lower streanl reach at T = 1 hr.
These bars
changed into bars of mode 3 from the most downstream reach at T = 47 min,

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

434
(a) Run C-3
27'

37'

3hr12'

(b) Run C-4


17 '

2,8 ,

1hr98'

(C)

Fig. 8.

Development of bed configuration in runs with a slope of 1/100.

and this change propagated up to near the middle reach by the end of the
run.
In run C-6 to C-9, flume slope was increased to 1/50. In the first run
with this slope, bars in squamation of mode 6 to 7 covered the whole reach,
but they soon became so irregular that the bar edges were not distinct at T

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

435

Fujita

== 15 min.
Simultaneously, the arrangement of scour holes and deposits
shifted to a pattern similar to mode 3. At T == 45 min, large bars emerged;
they were divided by a few small streams. The main flow cut around the
emerged bars, forming a channel pattern that looked like a sequence of the
number "8" lying end to end. Moreover, small bars formed along this main
flow.
Such a state of the stream bed was regarded as typical of braided
rivers.
In run C-7, which was characterized by a large discharge (Figure 9a),
bedforms changed progressively from bars in squamation of mode 6 to 8 at T
== 5 min, to uneven bars having various widths and a wavelength of about 1.5
m at T == 10 min. They thence evolved into regularly arranged bars of mode
4 with wavelengths of 1.6 m at T == 17 min. While maintaining this third
form, the stream started to concentrate into a pattern similar to mode 3;
double row bars appeared at T == 40 min. The bars did not emerge; instead
the bed was covered with a complicated mode 2 pattern overlapped by
small-scale bars.
In run C-8 (Figures 9b and lOb), the bed at T == 5 min was occupied by
bars in squamation of the highest mode observed, Le. mode 8 to 9; these sho\\1
a beautiful, geometrical pattern. Then the bar edges became rounded, as they
developed. Due to a decrease in bar mode to 5 to 7 at T == 10 min, and to
3 in the upstream reach and 4 in the downstream reach at T == 25 min, the
bars became large and irregular, concentrating the flow into scour pools. A
pattern of mode 3 occupied the whole reach at T == 40 min, and in parts the
bed was covered with a mode 2 pattern at T == 55 min. Small bars remained
on the large emergent bars, and the bed evolved into a braided stream.
Because bar mode decreased in all of the runs of series C, the maximull1
discharge of 10.5 i/ sec was supplied in run C-9 (Figure 9c) in order to
document the final or the lowest mode of bar metamorphosis. In this run,
bars of mode 4 to 5 appeared over the whole bed by T == 7 min. The lllode
at T == 14 min was 4 to 5 in the reach above x == 8 m, 3 to 4 in the middle
reach and 2 in the reach below x == 11 m. In the middle 10 m reach at T ==
25 min, the bed developed mode 2 to 3 bars, although small and large bars
were compounded. After that, flow was concentrated along the downstrealll
edges of bars, especially in the mode 3 reach, which did not maintain a mode
2 pattern. This reach rather developed a large bar near the center with a
form similar to an alternate bar where, in the lower reach below x == 7 m,
small bars remained on it. Furthermore, similar bars facing opposite to those
in the middle reach occurred in the upper reach at T == 1 hI', 10 min.
However, as these bars \vere nearly mode 1, they were unstable, so the
wavelength reduced by two-thirds after several minutes, and the bars seemed
to decompose into small bars of mode 4 to 5 again.
Thus, even under conditions of a fairly large ratio of width to depth, the
bar mode could diminish to unity only for an instant. If the channel was
long enough, it was not clear whether such bars of mode 1 would keep
developing. However, considering that the final heights of bars of mode 1 are
estimated to be very large according to equation (5) and that bars in run B-2
were similar to mode 1 at first, it may be that such bars of mode 1
eventually return to mode 2 or 3 after that.
In run C-10, conducted under a slope of 1/33 and a small discharge, bars
in squamation of mode 6 to 10 with a wavelength of 0.5 to 1 m formed in
the early stage. Their dimensions became uneven and their form irregular at
T == 6 to 7 min; their edges being elongated in the downstream direction. At
T == 25 min, the main streams were arranged in a pattern of mode 2 to 4
corresponding to large bars, and at the same time there appeared small but
distinctive scour pools along with small emergent bars. The main flow pattern

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

436
(a) Run C-7
17 '

39'

(b) Run C-8


5'

26'

56'

1hr14'

(e) Run C-9


8'

25'

1hr10'

Fig. 9.

Development of bed configuration in runs with a slope of 1/50.

at T = 40 min took the form of two number


emergent bars divided by small streams. The bed
be a typical braided stream.
After that this
downstream, and new bars formed within the main

8's, and there were many


condition was considered to
bed configuration migrated
stream.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Fujita

437

4,

10
,

''3;

If

(Ill)

'

.~,.

:T=30'

T=~h

T=)

'i

T=lh58'

(b)
Fig. 10.

Sketches of bedform development in run C-5, without emergent bars, and in


run C-8, characterized by stream braiding.

Summary of Specific Features of the Formation of Multiple Row Bars and


Stream Braiding in the Experiments
Series A and C runs were carried out at a fairly steep slope, using uniform
sand. Although bedform pattern development varied according to experimental
conditions, it was common to every experiment that regular, symmetrical bars
of very high mode appeared in the early stage, and that they evolved toward
a lower mode as time elapsed. This process was clearly documented in the
series C runs through observations and photography. Rapid changes in the
runs of series A, however, made it generally difficult to observe the process
and to take photographs.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

438

There were large differences in the final state of flow, bars and bar mode
among the runs, due to differing experimental conditions.
In cases with
relatively large depths and/or discharge, bars similar to alternate bars appeared
(runs A--6 and C-9). In cases with moderate depth, such as runs A-5, A-7,
C-4, C-5 and C-7, fairly symmetrical regular patterns of submerged double
row bars formed. In cases with small depths, such as run A-9, A-10, C-2,
C-3, C-8 and C-10, the occurrence of emerged bars divided by small streams,
and of a main flow pattern in the form of the number 8 caused the channel
pattern to appear braided when the bar mode was reduced to 2 or 3.
In contrast, the runs of series B, characterized by relatively mild slopes
and widely graded sand, had conspicuous longitudinal helicoidal flow in the
early stage of each experiment. Bar modes higher than 4 were difficult to see
without periodic interruptions of the flow. Double or triple row bars were
observed in the upstream reach in the very early stages; they systematically
propagated downstream. It is not clear at present why higher mode bars were
not formed in runs of series B.

Bed Evolution During the Process of Development of Bars


and Stream Braiding

Characteristics of Bed Variations


In run B-2, double row bars developed in neat order.
The final bed
profiles are depicted in Figure 11 along three longitudinal lines, Le. the
channel center and next to both sidewalls. The profiles along both sidewalls
reseluble each other very much, presenting a periodic variation of three
wavelengths.
Similarly, the center profile shows almost the same variation,
but with a phase lag of one bar length. It is suggested that double row bars
develop regularly not only in plan but also in vertical profile. Moreover, the
bed level on the bars aggraded slightly above the initial level, while scour
holes became deeper, coinciding with the bed evolution characteristic of
alternate bar development, implying a similar mechanism of development.
In the runs of series C, cross-sectional evolution of the bed could be
divided into two categories, that is, large-scale, simple changes and
small-scale, complicated changes. Although these change~ apparently seelned

70......-----------------,

z
(em)

Run B- 2
--Center
----Right
---Lett

90

lOa

o
Fig. 11.

40

Longitudinal bed profiles in run B-2, with double row bars.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

439

Fujita

to occur simultaneously in many cross sections because of bar migration, the


former changes were often found in runs with relatively large depths, in which
the bar surfaces were kept smooth because the flow was sufficient to move
sand grains over them, even at the final state. In contrast, the latter changes
corresponded to runs where braided streams appeared and large bars were
incised due to the fact that higher mode bars remained on them as relics from
the later stages of the experiments. Some large-scale simple cross-sectional
changes in runs C-1 and C-4 are shown in Figure 12; the complicated changes
of runs C-8 and C-10 are shown in Figure 13.
Cross-sectional changes at a reach between x = 9.5 m and 11.5 m in run
C-1 show that scour holes appear in the center and near both sidewalls, even
at the rather early stage of T = 28 min. Though the vertical magnitude of
bed variation increased as time elapsed, corresponding to the development of
double row bars, eventually exceeding 4 em, the bed variation itself was simple
enough to be traced easily in the figures. The smooth surfaces of large bars
showed that no influence remained of the antecedent bars of higher mode.
Though the number of measurement is few, Figure 12 also shows very clear
changes in cross-sectional shape at a middle reach (x = 4.5 to 9.0 m) in run
C-4. Scour pools decrease in number from 4 or 5 to 2 or 4 at T = 30 min,
and from 3 or 4 to 2 at T = 70 min. Almost all bars had smooth surfaces,
and intervals between the scoured trenches are nearly identical. Such features
imply that bars of fairly high mode form regularly in neat order and that
subsequent lower mode bars were not affected by previous bars of higher
mode.

I"

j'

i' i'

I' iii Iii iii

=--=.=
r: ~~:~~
_ - r=J031111n

Run (-1

(C~i~'l

H. .
F l
i

X=lom
(clm)" *tI-~,;
10

fX=
12"10.Sm<:: r

(em)

10 /

:::-

lf~_llm
<4

(em)

01

10 ,-/

41

~;Af rtn ~

'\1'"

X= It. 5: -

10~'

Z
(em)

,."".

Fig. 12.

'1

i"

I,,"""

200

iii ii"

- - T.

------. r.

--_. r

X= 4.Sm

Iii iii iii iii"

Olllin

ii' i i

Run (-4

30mln
70ml

~-.,""'....r::t,...-<::.:I;.=:

(em)

15

(em)

IS

X= 5m '~.-L~"
d

.. V' ;v

9l
1"lF

'<"

IF=s.smM
.... -:,...-\,

(em)

-..

.. -..... ~~---~,,~

"'-......

15

(em)

fX=~

:!!W'"==-%g--'~?\

l IS '~,'

=6.sm

\,1

"."

1
.

f9

IS ~'C:'~-'\~.~~--::~::;'-"''''''

lOa Y(cm)

ii"

Run (-4

l
(em)

5-

'"

I'i iii"

la-

"i"
o

I"

i'"

IDa

i""""
Y(cm) 200

"i"""
o

IDa

i"'"

Y (cm)

200

Changes in cross--sectional shapes in runs C-l and C-4, without emergent


bars.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

440
Iii' iii i i i iii' iii' , I

- - T-

Iii"

iii , iii

i , , iii

"

Ii

iii

Omin

==.~ l: 2S:~~

Run [-8

_ . - T= 46min

25~=
7m~,80min
Z~,~
(cm)

20

Z
X=_i'.5m_~_l -~
(cm)I~'
20
Ir

}J~

_,'<It l

X= 8m

2'

'"s'~

l
(Cz~:~~~'4~ l
~

8,5~

X=

Il)O[F?~r
15 1X= 9m

o
Fig. 13.

100

"Ii

(em)

Ii

200

100

r (em)

200

Ii

'Ii

100

r (cm)

200

Changes in cross-sectional shapes In runs C-8 and C-I0, characterized by


stream braiding.

Figure 13 shows very complicated changes in cross-sectional shapes during


braided stream formation at a reach from x = 7 to 11.5 m in run C-8, where
the highest mode of bars formed the most regular pattern. Changes in height
are small, within 2 cm, and only a few exceed 1 cm in individual differences
in height. Looking carefully at the bed changes during each interruption, it
can be noticed that crests and troughs corresponding to regular bars of higher
mode at T = 7 min decreased in their number and became of uneven interval
at T = 27 min. Along with this decrease in bar mode to 2 to 4, scour holes
became large. The small change in bed elevation on large bars, such as. the
left half of the cross section at x = 9 m, suggests that surfaces emerged
through the water surface at a very early stage. In cases with steep slope,
such as run C-8, even small streams on a large bar have fairly high shear
stress; they thus can erode the bar so as to render the cross sections clearly
notched.
Bed variations at a reach from x = 12 to 14 m in the steep run C-I0
show intricate features of small scale changes, in a fashion similar to run C-8
(Figure 13). In the early stage, scour holes were located at relatively uniforn1
intervals. From their number, the bar mode was regarded to be 8 to 10 at T
= 4 min, and to reduce for a short time to 4 to 6 at T = 18 min. This
mode continued until T = 40 min; the bed was notched both on the surfaces
of large bars and on the bottom of deeper flow areas. After T = 1 hI', no
substantial change appeared in the bed condition except that large bars became
more prominent.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

441

Fujita
(emJ

A-7

-0'2~,
.... __

-,

---2 taZ

A~~_~ 4::;Q/~
~ '. '.
003

A-9

-o./~ ~"
taZ
O

A-a

-8.z~
o

tal

0.3

A-9

-o~v ~,c;---..,
tal
0.2

0.2

A-IO

oJ;.~J
A-7
tal

_o.:~

6Z

oafc=-ct'Zl 6

4l

klO

0.2

0.2

-o~rA-1I

A-II

4l

0.2
0..,

10 20 30 40
Y femJ

Fig. 14.

'0

, "
,
,
20 30 40

{emJ

Cross-sectional shapes averaged longitudinally in runs A-2 to 11. Figures on


the right-hand side refer to repeated runs.
Numbers in left figures indicate
the time of flow stoppage.

Changes in Longitudinal Averaged Cross-Sections


As mentioned above, the development of bars can be examined using the
change in cross-sectional shapes averaged longitudinally.
Therefore, these
cross-sectional shapes are inspected here.
The averaged cross-sectional shapes in some runs of series A are depicted
in Figure 14. The cross sections in the first set of runs were averaged in a
middle 10-m reach by using data taken at a 10--em interval. Repeated run
cross sections were calculated from 1 m interval data; and widths and
differences in height were larger than the former because bed elevations were
measured closer to the sidewalls. In spite of these differences, especially in
data intervals, cross sections in both cases agree well with each other in the
number and height of peaks and their shapes, confirming the reproducibility of
the experiments. A comparison between the modes determined here and those
from the plan pictures results in the fairly good agreement shown in Table 2.
Major bars masking the existence of small bars make it difficult to judge
the intermediate configuration from the averaged cross section. In run A-7,
however, where bars of mode 2 and 1.5 were recorded with photographs at the
early and final stages respectively, the averaged cross sections show that a
regular shape with two peaks in the early stage turned into a mode 1 forIn
with a depressed crest, typical of intermediate configurations.
In cases of
mode 2 from photographs, the number of peaks are the same as determined
through direct surveys. In cases of a mode higher than 2, coincidences are
found only in run A-4 and the repeated run A-g. Paying attention to slight
zigzag changes in the figures, the mode agreed also in run A-10 and A-II.
Figure 15 shows a clear double bell-shape in run B-2, appearing as if two
rows of alternate bars were put in parallel. This coincides with the regular
double row bars formed in this run. The average cross section for run B-3
consists of a large bell-shape, with a depressed crest. The number of peaks

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

442

Comparison of bar modes determined


from plan pictures with those from
cross-sectional shapes averaged
longitudinally in runs of A series.

Table 2.

Mode by
photo

Run
No

Mode from longitudinal


averaged bed
First

A-2
A-3
A-4
A-5
A-6
A-7
A-8
A-9
A-I0
A-II

2
2
3
2
1.5
2
1
3
4
3

2
2
3
2
1
2
1
1
2
1

B-2
B-3

2
3

2
3

Reconducted

2N3
2
2
1

IN2
1
3

2N4
IN2

counted is three; this is consistent with the mode as determined from the
photograph. Such a cross-sectional shape was probably caused by a slightly
meandering main flow.
In series C, longitudinally averaged cross sections were calculated not only
for the whole reach, but also for three sub-reaches, upper (x = 2 to 6 m),
middle (x = 6 to 10 m) and lower (x = 10 to 14.5 m). This was done
because it was observed clearly in some runs that bedforms varied spatially,
reducing the bar mode in the downstream direction.
When large bars
appeared in low mode with wavelengths comparable to sub-reach lengths,
judgment becomes incorrect or impossible.
Figure 16 shows the averaged

_I r - - - - - - - - - - - - - - . ,

B-3

fjZ
(em) "

Fig. 15.

Y(m}

Cross-sectional shapes averaged longitudinally in runs B-2 and B-3.


broken lines show the initial bed.

Copyright American Geophysical Union

The

Water Resources Monograph

River Meandering

Vol. 12

Fujita

443

I i . , iii' iii iii i'

Run C-3

I' i"

'i I

........ T. 23",1"

- - - Til 71",ln
_ . - r.134",ln
_ .._- T= 195",1"

i""

I""

i"

Run [-8

o .: . ....

..,

.'

.,.,:::"

"-

"

~.O

-1
1

IH

-1
1

X = 6.0 TO 10.0 m

Zavr
~m)
aF-----,.,.,......,,~~lf"i'LI"'._.."""""""~~

(em!

OJ"

. .,.'- J \
v

'.

-I
I,

., i ' i '

Fig. 16.

I i'i"

100

i"

Y (em)

-1
i I

200

X = 10. a TO 14.5 m

r;~~~~1

"

r~.~.

.-..1......- : .

-1

-1

:: 10.0 TO 14.5 m

X = 2.0 TO 6.0 m

O~~\oo":=IoIII~"""'-'~~~,....;;i

bl~~":~~1
r'" . "\
Zavr

lcm)

-I

_00-

lavr

lcmb~

~vr

"I

4",1"
- - - T. 18",ln
_ . - T. 40",1"
T- 50.ln

~.

-I

I F"" X = 2.0 TO

ii"

. T-

r=)I~a f'"X~O:'rl:::

m~

I'"

i""

Run C-10

Zavr 1 r +
= 2.0 TOX
14.5
(em)

I""

"I

I"."

i'"

I"""", I
~m)
200

100

I""

'i"

I"

100

'(

I'll"

(em)

I
200

Cross-sectional shapes averaged longitudinally to characterize stream braiding


in runs C-3, C--8 and C-IO.

shapes in runs C-3, C-8, and C-10, as examples of braided stream forlnatiol1,
and Figure 17 shows those for runs C-4, C-5, and C-9 where no bars
emerged. The former shapes are notched, and the overall difference in height
is small. It is possible to read the bar mode from these figures, though they
are entangled. The averaged shapes in this case did not vary systematically,
implying that bed variations had almost same properties in the longitudinal
direction in the case of the braided stream formation.
In contrast, in the
latter case where the bars did not emerge, averaged shapes generally have
simple, smooth forms with large differences in height, and therefore, the bar
mode was judged more easily and agreed better with those determined fron1
pictures. The cross-sectional shapes averaged for the whole reach are quite
different from those for the sub-reaches. This suggests a spatial change in the
mechanism of bed evolution in cases without emergent bars.

Changes in Mode and Geometry of Bars


and Formative Conditions

Changes in Modes and Geometrical Characteristics of Bars


Characteristics of changes in the bar mode with elapsed time are depicted
in Figure 18. In the early stages, by using the mode judged from pictures in
the runs series C and run B-4, bars were seen to be in quite high mode:
more than 8 in runs of C-7 and C-8 with slope of 1/50, and about 5 in runs
C-3 and C-5 with slope of 1/100. These evolved to bars of a lower Inodes

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

444
I' i"

i',""

i'

i'"

I""

"

Run C-4

i'"

'I""

Run C-5

lcm6

,I

I""

'I""

i"

i'" I

Run C-9

. T- 30.ln
- - - T- 75mln
- . - T-113mln

~x= 2.0

Zo.",,1

i'"

TO 14.5 m

_"

.
-1

~x
=2
TO +
6.0 m 1 .

lo.",,1

(em)

""
I."

'.:

....

-1

~ := 6'::ak~~
1 . . '.

t;;)1
a f':Y:,;;,;

IO'OH

-I

La""
(em)

La""

j/

O~~~:----;O-~'--~"--:I

lcm)

I~X
:

TO

o {"

J''./-

14.5 m

.. ..;-::

-'

-1

-1

\
I.,.".

i"

Fig. 17.

I"

"'"

y (em)

100

"

I " " " " , I., i,',

200

100

.
9
6

+
1

2
T (hr)

10 ....6 ----R-u-n-C---e---------::
..

runs C-4, C-5 and C-9,

Run C-4
6 x-2-6m Including small bars
6-10 <>- Emerged bars,
10-14.5
Braided streams

f-

o'--_.. . . .'

6
6

6 "

5 .... "9

6~ -9'
.~...

o L...-_----L1

Run C-7
6 x-2-6m Including small bars
6-10 <>- Emerged bars,
10-14.5
Braided streams

6-10 <>-Emerged bars,


10-14.5
Braided streams

, ,

.. ,

f ~

..

..I....-1_ _11..-_---'-I_ _-01

0 1 2

..I....-1_ _'''--_.....L.1_ _.....I

0 1 2
T (hr)
10..--------------.

6 x-2-6m Including small bars

/:rJ

5-

200

Fig. 18.

"

200

10,........------------.

Run C-3
6 x-2-~m Including small bars
6-10 <>- Emerged bars
Braided streams
10-14.5

5 ....

"

Y (em)

Cross-sectional shapes averaged longitudinally in


which lacked energent bars.

10

T (hr)

o'--_.. . . .I

"'-'_ _I"--_......L.'_ _.....J

0 1 2

T (hr)

Changes in bar modes at upper, middle, and lower reaches in runs C-4 and
C-7 without emergent bars (right), and runs C-3 and C-8 characterized by
braiding (left).

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Fujita

445
Run

8-4
6

x a-10m

p x- 0-20m

Xa:>-30m

9 x 30- 42.8m

ott

T(hr)

Fig. 19.

Changes in bar modes at four reaches in runs B-4.

with time. Such changes were more rapid in runs with steeper slopes. In
cases without braided streams, the bar mode is clearly higher in the upstream
reach than in the downstream reach except for the early stage. On the other
hand, no systematic change in bar mode in the longitudinal direction was
discerned in cases with braided streams.
In cases with large depths, the
upstream boundary conditions of uniform levels of both the bed and water
surface seemed to propagate to the upper reach, such that the bed tended to
maintain the same conditions as the initial state there; bars became larger in
the downstream direction. In contrast, with small depths, flow states seemed
to be easily controlled by local conditions of the bed surface, which is
considered to result in the formation of similar bedforms for the whole reach.
Figure 19 depicts the change in bar mode in run B-4, characterized by
fairly large depths. Bar modes decreased gradually from 4 to less than 2; the
lowest modes were seen in the lower reach at every interruption. Although
this longitudinal tendency of mode decrease was silnilar to that of the runs of
series C with larger depths, it was not as clear as series C because bars of the
highest mode 4 were very thin, and disappeared by T = 3 hr. Distributions
of f B and ZB at each interruption in run B-4 are shown in Figure 20. At T
= 32 min, values of ZB were less than 1 em, whereas values of e varied
B
\videly from 0.5 to 6 m with a mode value of 1 to 2 m. After that, ZB
continued to increase in value and distributed range; the mode values of f B
also increased, but the range changed little until T = 4 hI', 15 min, ,vhen
both f B and ZB suddenly became large and a positive correlation developed
The
bet\veen them.
By T = 8 hI', this relation became more clear.
maximum and mean values (noted by the suffix "max" and an overbar,
These heights
respectively) of f and ZB changed as shown in Figure 21.
B
The
became large uniformly and increased gradually near certain values.
wavelengths were rather large from the early phase, and increased rapidly to
constant values after keeping these values for two hours.
Such phased
development of bar geometry was produced by changes in bar mode.
The

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

446
15.---------------------.,

15.----------------...

T=32'
10

10

la

la

(em)

(em)

o
o

o
o

la

o
o

o
0

15
(m)

15

15

T=5 hr46

T= , 33'
hr

10

10
0

la

la

(eml

(em)

5
o

o~

0
0

0
0

0>

~% 00

q]

10

la

15
(ml

15

la

(m)

la

(ml

15

15

T=8hrS2'

T= 2 hr 33'

10

10
la

Zs

(eml

(em)

0
0

0
0 0

8:'l80

0
0

q]
Fig. 20.

0
0

~ooo

OCb

0
0

10

la

15
(m)

15

Changes in geometrical characteristics of bars (bar wavelengths and heights)


in run B-4.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

447

Fujita
12

Is max

is

6
4

4
T(hr)

12
10

- - Zs 'mo~

------ Z;

8
Ze(cm)
6
4

2
0-----''"----..........-------------''-------'--2
6
8
4 T(hr)

Fig. 21.

Changes with time in the mean and maximum values of bar lengths and
heights in run B-4.

\vavelength, in particular, elongated suddenly due to the change in bar width


associated with appearance of bars of lower mode.

Formative Conditions for Bars in High Mode and Braided Streams


Formative conditions for bars of mode 2, 3, and higher are examined using
the data of series A and B, and data from previous studies. Predicted regions
of mode 2 and 3 bars were divided tentatively by the line (hid) B/d)2/3 =
0.07 shown in Figure 22. This diagram can give a measure of the final mode
of bars and stream braiding, but it was inadequate to predict the very high
lTIode of bars that appeared in the early stages of the series C runs.
The development of bars in various mode under the same hydraulic
conditions, as demonstrated in Figures 18 and 19, indicates an overlap of their
formative conditions. In addition, even in cases of high mode, individual bars
have a form similar to alternate bars, and the cross-sectional shapes averaged
longitudinally were confirmed to have peaks of the same number of modes of
It can be
regularly formed bars at each phase in the formative process.
concluded that individual bars in multiple rows have the same mechanism of

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

448

Mode

No Bar

<D 1.5

10

1------,
I
I

Double Row Bar(Braided)

()2

Mode 3

Mode 2

~3

-4

1/rc

lJ

o Authors

t1 ",-.~'
..

-E7Tamai

bMiwa

b,,1

Q Ogawa

6 Kinoshita

Fig. 22.

" ...I ...


lP~~ol

.'0

-q- ~6

OS

~7

,l ~B3

..

bib
'-OS2

Q))

gg ~I

10

Semi Bar

Ib

cl.

bb ~

19

tl

A1 ternating

Ba~

A previous criterion applied for bar formation in various modes.

bed evolution as that which leads to development of alternate bars.


This
conclusion is confirmed by an inspection of the bar characteristics in run B-2.
Accordingly, even for bars of such high 1110 des , formative conditions should be
predictable in terms of the unit bars which compose the scale-like pattern of
the alternate bar's formative condition.
Formative conditions for alternate bars are found within the ranges
[Muramoto and Fujita, 1977, 1978],
0.15 < (h/ d) / (B/ d)2/3 < 0.45;

(6)

Assuming that channel width, B, can be replaced by the mean bar width of
mode m, BB = B/m, in the main parameter, and by using the inverse of this
parameter to make values of B and bar modes correlate positively, the
formative condition for bars in mode m becomes, as shown in Figure 23,
2.2 m2/3

< (B/d)2/3/(h/d) < 6.7m2/3

(7)

Formative regions for each bar mode overlap in fairly wide ranges,
corresponding to the experimental observations. Bar modes are also shown in
Figure 24 as functions of the above parameter. Experimental data are plotted
for all runs in this figure.
The highest modes in the experiment do not
exceed the upper line, which expresses the higher boundary of inequality (7).
They are systematically below a broken line in Figure 24, expressed as:
2.8 m 2/3

(B/ d)2/3 /(h/ d)

Copyright American Geophysical Union

(8)

Water Resources Monograph

River Meandering

Vol. 12

449

Fujita

10
TITe

Fig. 23.

A criterion proposed for individual bar modes.

Since photographs were not necessarily taken immediately after the beginning
in all the experiments, it cannot be guaranteed that the highest modes that
appeared in the experiment are depicted in Figure 24. Even so, the bar mode
that becomes visible first seems to be still lower than the higher boundary of
the inequality of (7).
As for the final modes, in runs where bars were submerged until the end of
the experiment, all data are plotted above the lower line. In contrast, in
cases where bars emerged, data for the final phases plot below the lower line.
They correspond to the formation of braided streams, as shown by closed

10
m

Fig. 24.

Comparison of a criterion for bar mode with observed values (open circles:
submerged bars; closed circle: braided streams).

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

450

circles. Additionally, almost all open circles below the lower line correspond
to data for bars which emerged at later stages. Therefore, the lower line in
Figure 24 gives both the lowest modes of bars which are covered with
sufficient flow depth after a long flow duration such that the channel does not
appear braided, and a measure for the highest modes of braided streams.
Amplification (or growth rate) in a linear stability theory of various bar
modes were calculated for hydraulic conditions averaged with time in all of
runs. The stability theory of [(uroki and [(ishi [1984, 1985] was used because
that theory relies on appropriate and well-formulated assumptions. Analytical
results for runs B-4, C-4, and C-8 are listed in Table 3. Bars of mode 1 to
6, 1 to 10 and 1 to 18 have possibilities to appear in runs B-4, C-4, and
C-8, respectively, and all of the observed modes fall within these predicted
ranges.
In run B-4, bars of modes 2' and 3 have the same maximlun
amplification, and those of mode 4, which appeared first, have the third
greatest value. The greatest values in run C-4 and C-8 occur for modes 4
and 8, respectively, and coincide with the experimental observations. A few
modes near the maximum amplification have almost the same values as the
maximum, implying the co-existence of modes. The mode formed first can be
predicted from the theory, but the calculations are rather complicated and
tedious.
Thus, it is concluded that the formation of bars in various modes and
braided streams can be predicted by using diagrams such as Figure 24.
However, it is necessary to clarify the formative process quantitatively in order
to predict the time required for the changes in the bar mode and the
appearance of braided streams.

The Process by which Bar Mode Decreases


Differences in development time are regarded as the first cause of decrease
in bar mode from high order to low order. Assuming that the wavelength of
bars is I B = 5 BB and substituting B/m for BB' (4) and (5) yield

T = 0.57
e

0.0051 m- 8/3 B2(B/ d)2/3 /(h/ d)/(I-u l ) / qB


9

ZB = 0.0051 m- 5/3 B (B/ d)2/3 /(h/ d)/(1 -

u~)

(9)
(10)

According to (9) and (10), the development time and the wavelength vary in
proportion to the -8/3 and -5/3 powers of m respectively; hence bars in
higher mode are predicted to have very small heights and to develop very fast.
By using Ashida-Michiue's bed load function, relations between the bar
mode and the development time and wave height of bars can be evaluated
Because shear stresses evaluated from the measured
(Figures 25 and 26).
water stages and bed levels are quite small in runs C-2 and C-3, the
development time of bars in mode 5 is estimated to take more than two
hours, even though wave heights predicted by (10) are only 1.5 em. However,
this discrepancy results from low estimated values of bed load due to slnall
values of measured depths.
If bedload were measured more accurately, the
predictions could be improved.
In fact, bars in mode 5 are predicted to
develop fully within ten to twenty minutes in runs C-4, C-7, and C-8, which
agrees well with observations.
For bars in the highest mode which could be seen in the experiments, Le.
mode 8 in runs with slope of 1/100, predicted wave heights are quite small,
only 6 mm at most. Consequently, their development time was so short that

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Fujita

451
7

10

10

T.(sec)
105

104

let

-102

10'

Fig. 25.

/0

The development time of bars of individual modes estimated for the runs of
series C.

they were rarely found in the experiments, and few bars in modes higher than
5 were observed in experiments with mild slopes. In contrast, in runs C-8
and C-10, with slopes steeper than 1/50, bars in mode 8 are predicted to
develop higher than 1 cm; this corresponds well with the experiInental
observations.
Figures 27 and 28 depict comparisons of (9) and (10) with experimental
data from run B--4, respectively.
In Figure 27, a main body of observed
values plot below a line expressing (9), and values away from the line in this
area correspond to developing bars, whereas those along the line are developed
ones. Thus, the required time to full development of bars can be estimated
by (9). Although maximum bar heights (particularly those of mode 2) plot
above the line of (10) in Figure 28, the observed bar heights are generally
below those predicted from (10) because these heights were influenced by larger
heights of the other mode.
These figures indicate that the predicted
development time and the height of bars of mode 1 are equal to factors of 6
and 1.7 times experimental values, respectively, and that these bars keep
developing until they emerge so prominently as to produce a meandering n1ain
flow. Equations (9) and (10) give only average rates of bar development, and
it is difficult to predict when bars of a certain mode become visible.
Evaluation of bar heights during development is important to predict the
reduction of bar modes to lower values.
One method to estimate rates of development of wave heights is to use the
results of stability theories.
Assuming that initial amplitudes of bed
disturbances are the same for all modes, an order of appearance can be

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

452

100 r------,----,.--...,----,.----r---r-............,.-

Zs
(em)

Run C-IO

10

RunC-2

3
I

4
5

10

Fig. 26.

The bar heights of individual modes estimated for the runs of series C.

evaluated from amplifications as listed in Table 3. Similar amplifications of a


few modes close to that of the maximum indicate simultaneous development of
some modes, e.g. modes 7 to 9 in run C-8. Furthermore, if initial amplitudes
are known, the increase in bar heights and time required to reach certain
heights can be predicted for individual bar modes. This might enable the
estimation of the time when lower-mode bars become visible by comparison of
calculated heights with equilibrium values.

The Lowest Mode of Bars and Formation of Braided Streams


Although initial higher mode bars were often distinctive, they did not affect
the eventual formation of bars of lower modes; instead they left only
ephemeral traces on the bar surface of the lower modes. Therefore, bars of
lower mode are governed by more intensive forces. The decrease in bar mode,
however, stopped at some stage, except for the cases where formative
conditions for alternate bars were satisfied at the lowest mode. This stoppage
results from a reduced capacity for sediment transport on bars, caused by flow
concentration into the thalweg during bar development.
This arrests bar
growth at a lower mode. Simultaneously, as migration of bars developed at
that time is also impeded, the deeply scoured thalwegs are fixed and enlarged.
Flow concentration is accelerated and part of each bar begins to emerge. Thus
the stream channel evolves into a braided stream. This critical condition for
bar development is considered simply below.
Since flow cross-sectional areas change little in spite of bar development, a
simple cross-sectional shape can be used (Figure 29) to represent the loca.tion

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Fujita

453
10 7

Run 8-4

6 x: 0-10
e x 10-20

to'

=:

Te(sec)

x = 20-30

X =30-42.8 m

la'

L...--

Fig. 27.

-"--_----L.._---'------'----I-----i-.L.--'--"

3 m4

6 789 0

A comparison of the development time of bars of individual modes estimated


with those observed in run B-4.

where maximum discharge is concentrated in a deep part of the thalweg along


a developed bar edge. When averaged shear stress on the bar is reduced to
critical, the sediment transport which caused bar growth ceases. Defining this
critical depth as he and the thalweg width as BT , from the continuity of the
flow cross-sectional area,
(11 )
Supposing energy slopes on the bar and in the thalweg to be identical, the
relative wave height of bars at cessation of development is found to be,

(12)
Thus, relative height at the end of bar development depends on the shear
stress, and becomes large with increasing shear stress. Values of BBI B are
T
about 4 to 5 and values of U* 2c I U*2 are at least 114 in the present
experiments, and therefore, bars in an appropriate lTIode can continue their
In the case of bars of
development until ZBI h reaches values of 3 to 4.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

t+:-CJ1
t+:--

Table 3.

Bar
Mode
Run B-4
1
2
3
4
5
6
7
Run C-4
1
2
3
4
5
6
7

Amplifications of bars of individual modes calculated from a previous linear stability theory.

Lateral
Wave
Number

Maximum
AmpliBar
Width fication

Longitudinal Wave Number and Wave Length


Lowest
Number

(m)

(xl 0--:3)

0.026
0.053
0.079
0.105
0.132
0.158
0.184

3.01
1.51
1.00
0.75
0.60
0.50
0.43

0.0279
0.0408
0.0408
0.0351
0.0255
0.0129
-0.0026

0.0020
0.0041
0.0064
0.0092
0.0129
0.0187

0.026
0.052
0.077
0.103
0.129
0.155
0.181

1.80
0.90
0.60
0.45
0.36
0.30
0.26

0.0667
0.1756
0.2295
0.2468
0.2410
0.2186
0.1830

0.0025
0.0054
0.0084
0.0115
0.0148
0.0185
0.0227

Longest
Length

Max. Amplification
Number Length

18.61
8.58
5.56
4.06
3.14
2.52
2.05

0.014
0.022
0.027
0.031
0.033
0.035
0.036

5.74
3.57
2.90
2.58
2.40
2.29
2.21

0.0211
0.0375
0.0482
0.0544
0.0565
0.0533

0.014
0.027
0.035
0.040
0.044
0.047
0.050

3.22
1.74
1.35
1.16
1.05
0.98
0.93

0.0297
0.0394
0.0524
0.0616
0.0682
0.0727
0.0750

Copyright American Geophysical Union

Shortest
Length
(m)

(m)

(m)
40.49
19.37
12.30
8.59
6.14
4.23

Higest
Number

3.76
2.11
1.64
1.45
1.40
1.49
2.25
1.18
0.89
0.75
0.68
0.64
0.62

to
~
~

:::::l
0..

Q
=::T
~

~
t:;j

~
~

S
~

o
e'+

Water Resources Monograph

8
9
10
11
Run C-8
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19

River Meandering

Vol. 12

0.207
0.232
0.258
0.284

0.23
0.20
0.18
0.16

0.1359
0.0785
0.0115
-0.0646

0.0279
0.0347
0.0474

1.67
1.34
0.98

0.052
0.053
0.055
0.056

0.90
0.87
0.85
0.83

0.0751
0.0723
0.0624

0.62
0.64
0.75

0.014
0.028
0.042
0.056
0.070
0.084
0.098
0.112
0.126
0.140
0.154
0.168
0.182
0.195
0.209
0.223
0.237
0.251
0.265

1.80
0.90
0.60
0.45
0.36
0.30
0.26
0.23
0.20
0.18
0.16
0.15
0.14
0.13
0.12
0.11
0.11
0.10
0.09

0.0253
0.1055
0.2106
0.3008
0.3658
0.4078
0.4311
0.4394
0.4356
0.4215
0.3986
0.3678
0.3299
0.2855
0.2350
0.1786
0.1167
0.0495
-0.0230

0.0009
0.0024
0.0039
0.0053
0.0068
0.0083
0.0098
0.0114
0.0130
0.0147
0.0165
0.0185
0.0206
0.0229
0.0256
0.0286
0.0324
0.0377

27.08
10.44
6.49
4.72
3.71
3.04
2.57
2.21
1.93
1.71
1.52
1.36
1.22
1.10
0.98
0.88
0.78
0.67

0.005
0.012
0.018
0.023
0.027
0.030
0.033
0.035
0.037
0.039
0.040
0.042
0.043
0.044
0.045
0.045
0.046
0.047
0.047

4.91
2.10
1.40
1.10
0.94
0.83
0.76
0.71
0.68
0.65
0.62
0.60
0.59
0.58
0.56
0.55
0.55
0.54
0.53

0.0080
0.0170
0.0252
0.0324
0.0385
0.0438
0.0482
0.0520
0.0551
0.0576
0.0596
0.0611
0.0621
0.0625
0.0623
0.0614
0.0594
0.0557

3.15
1.48
1.00
0.78
0.65
0.57
0.52
0.48
0.46
0.44
0.42
0.41
0.40
0.40
0.40
0.41
0.42
0.45

~:
C"T"

C11
C11

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

456
100..------------Run 8-4

X:IO-20
0 T:032'
{)- X"'2Q-30
T-, hr33'
X-30-42.S m ~ Til r 33'
Q T-4hr I5'
() T-5 hr 46'
T- Shr52 ,

Za(cm)

$
e

-<D--tl}

-0

0./ _ _ _ _ _--'-_----''-------''-_....1.....-"""''''----.....1....--..............---'
I
2
4
8 10
6
m

Fig. 28.

A comparison of the bar heights of individual modes estimated with those


observed in run B-4.

high mode, (9) seems to give a simple measure for the final mode of bars.
Figure 30 depicts these relative bar heights for the cases of B/ h = 100 and
200, when u I is 0.4, that is, U* 2 / U*2 is 0.36.
9

In the former case, B/ h = 100,

bars in mode 2 hardly develop until ZB/ h exceeds 3, and alternate bars seem
In the latter case, bars of mode 2 grow more
to dominate on the bed.
rapidly than those of mode 1, and this mode will last until the final state.
This analysis indicates that the right-hand side boundaries in Figure 23 and

Fig. 29.

Definition sketch of cross-sectional shape of bars and thalweg.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Fujita

457

10

Alt. Bar

10

Ze/h

Ze/h
------

,,'

(>r.

uci =0.4

<Q

B/h=IOO

CO

(X

0.1

10
(8/d)2/j{h/d)
Fig. 30.

0.1

Ug~;:

0.4

B/h=200

10
(B/d}2/ {h/d)

Examples of the lowest bar modes.

the lower line in Figure 24 cannot be used to predict the formation of bars in
multiple modes in all cases because these boundaries were defined using data
from limited experimental conditions in which Blh did not exceed 100.

Conclusions
Although this paper does not review in detail all research on braided
streams, it does at least provide a list of investigations on braided rivers and
bars of high modes in Japan.
Detailed hydraulic experiments were carried out on the behavior of multiple
row bars and the formation of braided streams in three flumes, with widths of
The characteristic features of change in bed
0.5, 1.8, and 3.0 m.
configuration, and also stream bed variation, were described. Higher modes of
bars were observed to evolve into lower ones. Braided patterns appeared in
cases with small depths; during this transition to lower modes, bar geometry
changed suddenly, increasing in wavelength and heights. Bars in high modes
proved to have the same hydraulic properties as alternate bars. Several results
of previous studies on alternate bars could be applied to predict the behavior
of multiple-mode bars. It was concluded that the formation of bars of various
modes, and stream braiding, as well, can be predicted by a criterion proposed
in this paper. This criterion allows the co-existence of several modes, and
The times
corresponds to results of an existing linear stability theory.
required for the developluent and the reduction of bar mode are analyzed using
the bar heights of individual modes in equilibrium. Mechanisms controlling
the formation of the lowest mode and the development of braided streams are
also discussed.
However, it is necessary to examine the formative process
luore quantitatively from the viewpoint of sediment transport in order to
clarify these mechanisms.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Bar and Channel Formation

458
Acknowledgement

The author is very grateful to Mr. H. Akamatsu and Mr. N. Nagata for
their effort in the execution of all the experiments, and to Prof. Y. Muramoto
for his appropriate advice. A part of the experimental results was presented
in an international conference on "River Regime" held by Hydraulic Research
Ltd., England, on May 18-20, 1988.
Notation

an exponent

full width of the channel

BB
B

individual widths of bars in high modes


width of thalweg part of individual widths of bars
representative size of the bed material (the

diameter)
mean diameter of the bed material

mean diameter of the bed material

so

Fr

he
I

IB
m

Froude number
depth averaged for B
critical depth for sediment motions on bars
slopes of valleys or stream channels

wavelength of bars
bar mode
number of scour holes in each cross section

total discharge

qB

T
T

UI

U*
U* c

bedload per unit width


experimental time elapsed
developing time of bars
non-dimensional grain veloci ty of bed load.
mean shear stress for h

critical shear stress for d


longitudinal coordinate

transverse coordinate

z
ZB

mean

vertical coordinate
wave height of bars
References

Ackers, P., and F. G. Charlton, The shapes and resistance


meandering channels, Proc. ICE, 73628, 349-370, 1970.

Copyright American Geophysical Union

of small

Water Resources Monograph

River Meandering

Vol. 12

459

Fujita

Ashmore, P. E., Laboratory modeling of gravel braided stream morphology,


Earth Surface Processes and 'landforms, 7, 201-225, 1982.
Begin, Z. B., The relationship between flow-shear stress and stream pattern,
J. Hydrology, 52, 307-319, 1981.
Bettess, R., and W. R. White, Meandering and braiding of alluvial channels,
Proc. ICE, Part 2, 525-538, 1983.
Brice, J. C., Channel patterns and terrace of the Loup Rivers in Nebraska,
U. S. Geol. Survey, Prof. Paper 422-D, 1964.
Cant, D. J., and R. G. Walker, Fluvial processes and facies sequences in the
sandy braided South Saskatchewan River, Canada, Sedimentology, 25,
625-648, 1978.
Carson, M. A., The meandering - braiding river threshold: a reappraisal, J.
Hydrology, 73, 315-334, 1984.
Chitale, S. V., Tneories and ~lationships of river channel patterns, J.
Hydrology, 19, 285-308, 1973.
Church, M., and D. Jones, Channel bars in gravel bed rivers, Gravel Bed
Rivers, ed. Hey, R. D., J. C. Bathurst and C. R. Thorne, John Wiley &
Sons Ltd., 291-324, 1982.
Colombini, M., G. Seminara, and M. Tubino, Finite-amplitude alternate bars,
J. Fluid Mech., 181, 213-232, 1987.
Committee on Hydraulics, Japan Soc. Civ. Eng., Three-dimensional
characteristics of flood flow and stream channel patterns, Report of Task
Committee, 92 p. (in Japanese), 1982.
DPR1, Kyoto Univ. Research Group for Disaster in Toyama District, Synthetic
Report on Flood Disaster in August, 1969, 66-83 (in Japanese), 1970.
Engelund, F. and o. Skovgaard, On the origin of meandering and braiding,
J. Fluid Mech., 57 (2), 289-302, 1973.
Fredsoe, J., Meandering and braiding of rivers, J. Fluid Mech., 84 (4),
609-624, 1978.
Fujisawa, H., H. Takahide, and M. Onaka, On SOlne phase of river bed
variations and local scour, Proc. 37th Annual meeting, Japan Soc. Civ.
Eng., 11-270, 539-540 (in Japanese), 1982.
Fujita, Y., H. Akamatsu, and Y. Muramoto, Experiments on formative process
of double row bars and braided streams, Annual Disast. Prevo Res. Inst.,
Kyoto Univ., 29B-2, 451-472 (in Japanese), 1986.
Fujita, Y., H. Akamatsu, and Y. l\1uramoto, Formative process of braided
streams, Proc. 31st Japanese Conf. on Hydraulics, Japan Soc. Civ. Eng.,
695-700 (in Japanese), 1982.
Fujita, Y., and Y. Muramoto, Y., On the mechanisms of alternate bar
development, Proc. 26th Japanese Conf. on Hydraulics, Japan Soc. Civ.
Eng., 25-30, (in Japanese), 1982a.
Fujita, Y., and Y. Muramoto, Y., Experimental study on stream channel
processes in alluvial rivers, Bull. Diast. Prevo Res. [nst., Kyoto Univ.,
32(1), 49-96, 1982b.
Fujita, Y., and Y. Muramoto, On the formation of stream channel patterns, in
j.

Proc., 3rd Congr., Asian and Pacific Regn. Div., Int'l. .l4ssoc. Hydr. Res.,

Paper C, 276-287, 1982c.


Fujita, Y. and Y. Muramoto, Study on the process of development of
alternate bars, Bull. Diast. Prevo Res. [nst., Kyoto Univ., 35(3), 55-86,
1985.
Fujita, Y. and Y. Muramoto, Multiple bars and stream braiding, Proc. Int'l.
Conf. on "River Regime," Hydraulic Research Ltd., England (in press),
1988.
Fukami, T. and R. Baba, Design of a low water course on a model test - an
example of the Shinano River, Proc. 32nd Annual Meeting, Japan Soc. Civ.
Eng., 11-286, 552-553 (in Japanese), 1977.

Copyright American Geophysical Union

Water Resources Monograph

460

River Meandering

Vol. 12

Bar and Channel Formation

Fukami, T., Co-existence of sand bars and sand ripples and dunes, Technical
Rpt. for civil engineers, 21-10, 27-32 (in Japanese), 1977.
Fukuoka, S. and S. Yamasaka, Equilibrium height of alternate bars based on
non-linear relationships among bed profiles, flow and sediment discharge,
Proc., Japan Soc. Civ. Eng., 357, 11-3, 45-54, 1985.
Fukuoka, S., S. Yamasaka, and Y. Shimizu, Dominant wave length and
formation conditions of alternate bars based on non-linear analysis, Proc.,
Japan Soc. Civ. Eng., 363, 11-4, 115-124, 1985.
Gole, C. V. and S. V. Chitale, Inland delta building activity of Kosi River,
Jour. Hydr. Div., Am. Soc. Civ. Eng.,92(HY2) , 111-126, 1966.
Graf, W. 1., Channel instability in a braided sand bed river, Water Resour.
Res. 17(4), 1087-1094, 1981.
Henderson, R. M., Stability of alluvial channels, Jour. Hyd. Div., Am. Soc.
Civ. Eng., 87(HY6), 109-138, 1961.
Hashimoto, 1(., New Technology for River Channel Regulation, MorikitaShuppan (Publisher), 308p. (in Japanese), 1956.
Hong, L. B. and T. R. H. Davis, A study of a stream braiding, Summary,
Geol. Soc. Amer. Bull., 90(1), 1094-1095, 1979.
Howard, A. D., M. A. Keech, and C. L. Vincent, Topological and geometrical
properties of braided streams, Water Resour. Res., 6(6), 1674-1688, 1970.
Ikeda, H., A study on the formation of sand bars in an experiment flume,
Geographi. Rev. Japan, 46-7, 435-450 (in Japanese), 1973.
Ikeda, H., On the bed configuration in alluvial channels: Their types and
their condition of formation with reference to bars, Geographi. Rev. Japan,
48-10, 712-730 (in Japanese), 1975.
Ikeda, H., An experimental study of similitude law of alternating bars, Bul/.
Environ. Res. Center, Univ. Tsukuba, 6, 3-14 (in Japanese), 1982.
Ishikawa, T., Hydraulic model test on a flood flow in the I(urobe River, Proc.
27th Japanese Con! on Hydraulics, 753-760 (in Japanese), 1983.
I(adomura, H., Micro topography and its formation on alluvial fans, Alluvial
Fan, ed., Yazawa, D., H. Heya and S. I(aizuka, I(okin-Shoin, 55-96 (in
Japanese), 1971.
I(ayane, 1. and S. Yamamoto, Hydrological Cycle in an Alluvial Fan,
Kokin-Shoin, 12-17 (in Japanese), 1971.
Kinoshita, R., On the formation of sand bars on river bed - An observation
on
river meander, Proc. Japan Soc. Civ. Eng., 42, 1-21 (in Japanese),
1957.
I(inoshita, R., Comprehensive observations of flood flows and stream channel
patterns, Lecture Notes, 14th Summer Seminar on Hydraulics, Course A (in
Japanese), 1978.
I(inoshita, R., Historical survey of meander in the Ishikari River - Appendix,
Paper, National Resour. Bureau, Agency of Science and Technology, 36, 174
P (in Japanese), 1962.
Kondo, Y. and Y. Komori, A fundamental experiment on alternate bars,
Monthly Rept., Laboratory, Hokkaido Development Bureau, 51(2) 1-9 (in
Japanese), 1974.
Kuroki, M., and T. I(ishi, Regime criteria on bars and braids in alluvial
straight channels, Proc., Japan Soc. Civ. Eng., 342, 87-96 (in Japanese),
1984.
Kuroki, M., and T. Kishi, Regime criteria on bars and braids, Hydraulic
Paper 9, Res. Lab. Civil and Environ. Eng., Hokkaido Univ., 23p, 1985.
I(uroki, M., T. Kishi, and S. Imaizumi, A experimental study on constrained
meandering flow, PrJc., 30th Annual Meeting Japan Soc. Civ. Eng., //-169
(in Japanese), 1975.
Lane, E. W., The importance of fluvial morphology in hydraulic engineering,
Trans., Am Soc. Civ. Eng., 81, Paper 795, 1-17, 1955.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

461

Fujita

Lane, E. W., A study of the shapes of channels formed by natural streams


flowing in erodible material U. S. Army Corps of Eng, MRD. Sed. Sere 9,
106p. 1957.
,.
Leopold, L. B., and M. G. Wolman, River channel patterns:
Braided,
meandering and straight, U. S. Geol. Survey, Prof. Paper 282-B, 83-300,
1957.
Ministry of Constructions (River Bureau and Public Works Res. Inst.),
Research studies on local scour in rivers, Proc. 36th Study Meeting for
Technology, 582-619 (in Japanese), 1982.
Miwa, H., Diversion weir and imbricate sand-gravel bars in an alluvial fan
river, Trans JSIDRE, 90, 41-47, (in Japanese), Dec. 1987.
Miwa, H., Dynamic similarity of the pattern of model bars to prototype, Proc.
27th Japanese Congo on Hydraulics, 733-740 (in Japanese), 1983.
Miwa, H., Relations between serial and parallel alternating bars, Proc. 28th
Japanese Conf. on Hydraulics, 775-781 (in Japanese), 1984.
Mosley, M. P., Analysis of the effect of changing discharge on channel
morphology and instream uses in a braided river, Ohau River, New
Zealand, Water Resour. Res., 18(4), 800-812, 1982.
Muramoto, Y., Patterns and processes of alluvial stream channels, Lecture
Notes, 12th Summer seminar on Hydraulics, course A (in Japanese), 1976.
Muramoto, Y. and Y. Fujita, Study on meso scale river bed configuration,
Annuals, Disast. Prevo Res. Inst., Kyoto Univ., 20B-2, 243-258 (in
Japanese), 1977.
and Y. Fujita, A classification and formative conditions of
l\1uramoto, Y.
river bed configuration of meso scale, Proc. 22nd Japanese Conf. on
Hydraulics, Japan Soc. Civ. Eng., 275-282 (in Japanese), 1978.
Muramoto, Y., Y. Fujita, and N. Nagata, Changes with tilue in sand bars in
braided streams, Proc. Annual Meeting, [(ansai Ch., Japan Soc. Civ. Eng.,
11-35 (in Japanese), 1987.
.
Ogawa, R. and T. Fukami, Intermediate patterns between alternate and
double row bars, Proc. 34th Annual Meeting, Japan Soc. Civ. Eng., 11-95,
189-190, (in Japanese), 1979.
Osterkalnp, W. R., Gradient discharge, and particle-size relations of alluvial
channels in I{ansas, American J. Science, 278, 1253-1268, 1978.
Sakano, A. and I{. Yamamoto, Model test similarity of alluvial channels \\lith
coal powder, Technical Rpt. for Civil Engineers, 23-3, 141-146 (in
Japanese), 1981.
Schumm, S. A., Evolution and response of the fluvial system, sedimentologic
implications, SEPM Special Publication, 31, 19-29, Aug., 1981.
Shimizu, Y. and T. Itakura, Two-Dimensional Flow and Stream Bed
Variations in Rivers, Hokkaido Development Bureau, Laboratory for Rivers,
65p. (in Japanese), 1986.
Shizuoka Local Construction Office for Rivers, Ministry of Construction,

Working Report on Bed Configuration and Flood Flow in the O'oi River,

ed. Kinoshita, R., 159 p. (in Japanese), 1979.


Shizuoka Local Construction Office for Rivers,

Ministry of Construction,

Assessment by Hydraulic Model Tests for Wideninp of Ushio Narrows in the


O'oi River, ed. Kinoshita, R., 117 p. (in Japanese), 1980.

Srivastava, P. L., Effect of Bhimnagar barrage on the morphology of the


River Kosi, Proc. 26th Conf. Int'l. Assoc. Hydr. Res., Mosco\v, 116-125,
1983.
Suga, K., Shortening of bar length due to degradation, Proc. 27th Japanese
Conf. on Hydraulics, 761-766 {in Japanese), 1983a.
Suga, K., Recent abrupt change of river regime and its characteristics, Lecture
Notes, 19th Summer seminar on Hydraulics, course A (in Japanese), 1983b.

Copyright American Geophysical Union

Water Resources Monograph

462

River Meandering

Vol. 12

Bar and Channel Formation

Sukegawa, N., Hydraulic conditions for the formation of alternating bars in


experimental flumes, Proc., Japan Soc. Civ. Eng., 207, 42-47 (in Japanese),
1972.
Sunada, K., River bed variation due to a flood runoff in 1(amanashi River
(Fuji River), Proc., Japan Soc. Civ. Eng., 363, 11-4, 235-243 (in Japanese),
1985.
Sunada, K., River bed variation in the Kamanashi River (the Fuji River) -2,
40th Annual Meeting, Japan Soc. Civ. Eng., 11-239, 477-478 (in Japanese),
1985.
Tarnai, N., S. Nagao, and F. Mikuni, On the large scale bar patterns in a
straight channel, Proc. 22nd Japanese Con! on Hydraulics, 265-273 (in
Japanese), 1978.
Toyama Local Construction Office, Ministry of Construction, Flood Disaster in
the Joganji River in the Rainy Season (in Japanese), 1978.
Turitani, Y. and T. Igarashi, On stream-bed variations of the Joganji River,
Shin-Sabo, 78, 25-36 (in Japanese), 1971.
Williams, P. F. and B. R. Rust, The sedimentology of a braided river, J.
Sedimentary Petrology, 39(2), 649--679, 1969.
Yamaguchi, H. and K. Okabe, Sand bar patterns and the planform of rivers,
Proc. 35th Study meeting for Technology, 662--665 (in Japanese), 1981.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Published in 1989 by the American Geophysical Union.

Topographic Response of a Bar in the Green River, Utah to


Variation in Discharge
E. D. Andrews and J. M. Nelson

Lakewood, Colorado
Abstract
A fully nonlinear numerical model was used to analyze the characteristics of flow
and sediment transport in a reach of the Green River near Ouray, Utah. The study
reach has gentle, but complicated curvature and contains a prominent mid-channel
bar composed of medium to fine sand. Bankfull channel width and area vary
significantly through the study reach. Results of the model calculations were
compared with the observed lateral distribution of unit discharge and mean annual
bed-material transport and were found to be in excellent agreement.
Adjustment of channel topography over a period of time at a constant discharge
was calculated at three discharges, 50 m 3 /s (approximately 40 percent of the mean
annual flow), 275 m 3 /s, and 475 m3 /s (approximately the bankfull discharge).
Each simulation began with the same topography, which was surveyed at a
discharge of N275 m:J /s on July 15-16, 1986, 5 weeks after a peak discharge of N620
m3 /s.
Gradually decreasing flows following the flood peak and significant
bed-material transport rates tended to maintain a channel topography, in
equilibrium with the discharge. As expected, computed changes in bed elevation
were quite small everywhere during a discharge of 275 m 3 /s for 2 days. Conversely,
channel topography evolved rapidly over a period of 2 days at simulated discharges
of 50 m 3 /s and 475 m 3 /s. Bar topography was enhanced greatly at a discharge of
475 m 3 /s. Sediment accumulates on the bar surface and is eroded from the side
channels except in the downstream part of the secondary channel where some
sediment is deposited. At a discharge of 50 m3 /s, the bar crest stands nearly 1 m
above the water surface, and there was no flow through the secondary channel.
More than 1 m of material accumulates in the primary channel along most of the
length of the bar, over a period of two days.
A comparison of aerial photographs taken of the Ouray reach between 1963 and
1987 shows that the bar we investigated has changed very little during this period,
in spite of the fact that bar topography tends to adjust relatively quickly to changes
in river discharge. Neither established vegetation nor debris appear to have a
stabilizing effect upon the bar. Long-term stability of the bar appears to be due to a
constriction in channel width and area located slightly downstream of the bar apex.
Introduction
Most of the science of river mechanics has been discovered from studying rivers
carefully selected to ensure equilibrium between channel configuration and the water
and sediment discharge. Existing analytical approaches and methods typically are
463
Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Topographic Response of a Bar

464

based upon assumed equilibrium conditions. Even the analysis of river channel
change commonly focuses on prior and future equilibrium conditions, while
neglecting the actual course of channel adjustment.
Disequilibrium rivers are common; indeed, they may be the norm rather than
the exception. In many instances, human activities, e.g., agriculture, forest
clearing, grazing, urbanization, and the development of water supplies, have caused
large, long-term changes in the quantity of flow and sediment supplied to the
channel network. Natural events, also, can have a very significant impact on river
equilibrium. Moreover, adjustment of river channels to such long-term trends
frequently occur over a period of decades to centuries. Disequilibrium rivers are not
only common, they are persistent on a human time scale. As our understanding of
long-term trends in streamflow and sediment transport increases, assumptions of
equilibrium are becoming progressively less justifiable.
Realistic anClJysis of
long-term channel changes requires theoretical approaches which are not based upon
the existence of equilibrium conditions.
In recent years, physically-based numerical models which accurately describe
the flow structure and distribution of boundary shear stress throughout a
nonuniform channel with arbitrary topography and longitudinally varying radius of
curvature have been developed [see Smith and McLean, 1984; and Nelson and Smith,
1988a,b]. The divergence of the local bed-material transport field, which represents
the evolution of channel topography per unit time, can be computed from the
distribution of boundary shear stress. Nelson and Smith [1988b] describe the
simulated development of steady-state point-bars and alternating bars in an
initially flat-bedded channel. This analysis does not require one to assume that
equilibrium exists, and indicates the way to realistically analyze long-term channel
adjustment.
This chapter will describe the application of a fully nonlinear numerical model to
analyze the evolution of channel topography in a reach of the Green River near
Ouray, Utah, over relatively short periods of time at various discharges. The reach
has a gentle, but complicated curvature and highly nonuniform bankfull channel
width and area. The reach contains a prominent mid-channel bar which was the
primary focus of our investigation. The bar is composed of well-sorted sand and is
not stabilized by vegetation or debris.

Green River
The Green River drains approximately 116,000 square kilometers along the west
flank of the Rocky Mountains in Wyoming, Colorado, and Utah (Figure 1). It is
one of the principal tributaries in the Colorado River Basin. The main stem Green
River originates in the Wind River Mountains of Wyoming, and flows southerly to
its confluence with Colorado River near Moab, Utah. The principal tributaries of
the Green River are the Blacks Fork, Yampa, Duchesne, White, Price, and San
Rafael Rivers. Flow in the Green River has been regulated by Flaming Gorge
Reservoir since October 1962. Tributaries to the Green River have numerous small
impoundments, especially in their headwaters. Except for the Duchesne River,
however, the tributaries are generally free flowing and unregulated at present
(1988).
The location of the long-term gaging stations in the Green River basin are
shown in Figure 1. Water discharge has been recorded daily at most gages for
several decades. Extensive records of suspended-sediment concentration also have
been collected at the gages shown in Figure 1. The available historic information
concerning flow and sediment transport in Green River provides one of the longest
and most comprehensive descriptions for any river in the world. Andrews [19861
analyzed the water and sediment discharge records and determined mean annua

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Andrews and Nelson


112

465

Ill

110

109

lOS

lOr

106

44r----....,-r----_----~----......,----.....,...---__,

.J \
.~

t'

43

IDAHO

"~\

>\~
.

(
:

42'

-------1
I

WYOMING

-,

\~

,/

/ '

\ . . - ""\......I'

.8
aSln 8 oun dary

\.. . .,-.

I:\
,
L)I

-"-"-1\

'

4\'

UTAH

I.>~

~'~1-~

,r,J

40

39

38----_..Ioo-

Fig. 1.

........

...L.-

25

50

--1....

100 MILES
I

_ _- - -

Green River basin with location of principal water and sediment gaging stations.

water and sediment budgets for three reaches of the Green River downstream from
Flaming Gorge Reservoir before and after construction of the reservoir.
The present discussion will focus on fluvial conditions in the 105 kilometer reach
between the Jensen and Ouray gaging stations. In this reach, the Green River flows
through a succession of wide, alluvial valleys and narrow, bedrock gorges. Within
the alluvial valleys, the Green River has a meandering bankfull channel and an
extensive, well-vegetated floodplain. Vegetated channel islands, standing at about
the same elevation as the floodplain, are common. At discharges less than about

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

466

Vol. 12

Topographic Response of a Bar

150

300 METERS

Contour Interval 1 Meter

Fig. 2.

Channel topography of the Ouray reach.

60% of the bankfull value, unvegetated, sandy channel bars become emergent and
give the river a braided pattern within the bankfull channel.
In order to investigate the topographic response of channel bars to variations in
discharge, we selected a study reach about 10 channel widths in lengths, located
within the Ouray National Wildlife Refuge 6.5 km upstream from the mouth of the
Duchesne River. Ten cross sections were established within the study reach and
resurveyed 5 times between July 1986 and Sept. 25, 1987. In addition, the water
discharge, suspended sediment concentration, and size and bed-material size were
determined each time the channel cross sections were surveyed.
A contour map of the study reach is shown in Figure 2. This topography was
generated from the initial survey on July 15-17, 1986, at a discharge of ",280 m 3 /s,
on the descending limb of a hydrograph which peaked at 620 m 3 /s 38 days
previously. A larger, primary channel lies on the left side of the bar and a smaller,
secondary channel lies to the right of the bar. The secondary channel is shallower
than the primary channel. Flow through the secondary channel ceases at a
discharge of 200 m 3 /s, when the water surface is "'1.5 m below the bankfull stage.
The top of the bar lies about 1.2 m below the bankfull elevation and is free of any
vegetation. Furthermore, the bar surface is generally clear of large debris, such as
logs or tree stumps, as well as coarse bed-material. Thus, large obstructions in the
flow do not appear to influence the stability of the bar. A comparison of available
aerial photographs shows that this bar has existed since 1962 with roughly the same
configuration and location.
The channel of the Green River in the vicinity of the study reach appears to be
typical of alluvial reaches between the Jensen gage and the mouth of the Duchesne
River. Although the study reach is located 14 km upstream from the Ouray gage,
daily water discharge and suspended sediment concentration determined at the
Jensen gage, 90 km upstream, provide a much more accurate description of flow and
sediment transport in the study reach than does the record of the Ouray gage.
Tributary inflow to the reach between the Jensen gage and the study reach is very
small. The Duchesne River, drainage area ",11,000 km 2 , and White River, drainage
area '"13,300 km2 , represent 93 percent of the difference in drainage area between
the Jensen and Ouray gage, and join the Green River downstream from the study
reach. Therefore, the record of the Jensen gage will be used to describe the
variations of flow and sediment transport which have occurred in the study reach
over the past 40 years.
Andrews [1986] provided a detailed description of the effects of reservoir
regulations upon flow and sediment transport in the Green River basin, including

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

467

Andrews and Nelson

the Jensen gage. During the pre-reservoir period from 1948 to 1962, the mean
annual suspended sediment load at the Jensen gage was 6.3xl06 ton/yr. The Jensen
gage is located in a steep, narrow reach with coarse gravel and cobble bed-material
immediately downstream of the Split Mountain gorge. All sand-size material is
suspended at the sampled cross-section even at relatively common discharges. Thus,
the sampled suspended sediment loads at this gage are probably a very good
estimate of the total sediment load in the alluvial reaches downstream. The mean
annual load of sand-sized material (0.062-2.00 mm) at the Jensen gage from 1948 to
1962 was 2.3xl0 6 ton/yr.
During the period from 1964 to 1982, when daily sampling of suspended
sediment ceased at the Jensen gage, the mean annual suspended sediment load
decreased by 54 percent to 2.9xl0 6 ton/yr. The Inean annual load of sand-sized
material decreased by 65 percent to 0.84xl06 ton/yr.
Although the mean annual suspended sediment load transported past the Jensen
gage has decreased substantially as a result of reservoir regulation, an approximate
equilibrium persisted between quantity of sediment supplied to and transported out
of the reach of the Green River from the mouth of the Yampa River to the mouth of
the Duchesne River, a distance of N160 river km [Andrews, 1986]. The decrease in
the mean annual suspended sediment load at the Jensen gage is approximately equal
to the mean annual quantity of sediment deposited in Flaming Gorge Reservoir.
Consequently, there is no appreciable accumulation or depletion of bed-material in
the Green River downstream from its confluence with the Yampa River to its
confluence with the Duchesne River, including the study reach.
The quantity of sediment transported by a given discharge at the Jensen gage
does not appear to have been affected by Flaming Gorge Reservoir. The variation of
daily suspended sediment transport rate as a function of water discharge was
computed for each size fraction from <0.004 mm up to the 0.250-0.500 mm fraction
by a least-squares regression of the log-transform data during the pre- and
post-reservoir periods. No statistically significant change in the relations was
detected by the F-statistic at the 95th-percentile level of confidence.
One of the principal downstream effects of Flaming Gorge Reservoir is to
decrease the range of daily mean flows. The mean annual discharges of the Green
River during the pre- and post-reservoir periods are virtually identical at both the
Jensen and Green River, Utah, gages. The percentage of times that various daily
mean discharges are equalled or exceeded, however, is substantially different for
most flows.
The decrease in the mean annual sediment transport rate at the Jensen gage
between 1962 and 1982 compared to the pre-reservoir period is due entirely to a
decrease in the magnitude of river flow that are equalled or exceeded less than 30
percent of the time.
Daily mean water discharges with a duration of 5 percent or less have decreased
in magnitude by 25 and 35 percent during the post-reservoir period at both the
Jensen and Green River, Utah, gages. The magnitude of daily mean discharges with
a duration greater than 30 percent, however, have increased to the extent that the
mean annual runoff measured during the pre- and post-reservoir period is virtually
unchanged at both gages. Thus, the decrease in annual sediment transport results
from a more uniform annual hydrograph rather than a decrease in the annual runoff.
Runoff in the Green River during the period from 1983 to 1986 was quite
extraordinary and greatly complicates the analysis of channel adjustment. In spite
of the presence of Flaming Gorge Reservoir, the duration of relatively large
discharges was only slightly less than during the pre-reservoir period. For example,
a discharge of 510 m3 /s was equalled or exceeded 4.96 percent of the time between
1947-62, compared to 4.18 percent of the time between 1978 to 1986. Spring
snowmelt runoff during 1983 and 1984 was exceptionally large. Annual runoffs at
the Green River near Green River, Utah gage during the 1983 and 1984 water years
were the largest and second largest in the period of record, 1920-86. Approximately

Copyright American Geophysical Union

Water Resources Monograph

468

River Meandering

Vol. 12

Topographic Response of a Bar

70 percent of the days with mean daily discharge greater than 510 m 3 /s during the
period from 1978 to 1986 occurred in 1983 and 1984.
The mean annual load of sand transported past the Jensen gage during the
period 1983 to 1986 was 2.7xl0 6 ton/yr, or more than 3 times the mean annual
quantity transported during the period immediately after the reservoir was
constructed, 1963 to 1982. The relatively large discharges and sediment transport
rates during the period from 1982 to 1986 are comparable to the condition which
existed prior to the reservoir.
Changes in the Green River channel since the construction of Flaming Gorge
Reservoir must be deduced primarily from a comparison of large scale aerial
photographs (Andrews, 1986]. All of the Green River gages downstream from the
reservoir are ocated in stable reaches, where coarse gravel and cobble bed-material
as well as bedrock severely limit the adjustment of channel width and depth.
Pucherelli rwritten communication, 1988] analyzed recent changes in bankfull
channel width as well as the number and area of islands in several reaches of the
Green River, using aerial photographs taken in 1963, 1974, 1978, and 1986. One of
the reaches he investigated extends approximately 18 km upstream from the mouth
of the Duchesne River, and includes our study reach. Average bankfull channel
width in this part of the Green River decreased from ",215 m to 189 m during the
period from 1964-78. During the subsequent period, 1978-86, however, bankfull
channel width increased to 198 m.
Although gradual accretion of bank material was common in the Green River
between 1964 and 1978, the most significant process narrowing the channel was
aggradation of channel bars, resulting in their attachment to the bank and
incorporation into the floodplain. In nearly all instances, thick vegetation became
established before the bar became attached. This narrowing of the channel since the
beginning of flow regulation appears to involve the same processes responsible for
forming the pre-reservoir floodplain. Inspection of the floodplain adjacent to the
study reach shows that it has been constructed by the aggregation of islands as well
.
as the gradual accretion of scroll bars.
The remainder of this chapter will consider in detail the nature of flow and
sediment transport in the vicinity of a mid-channel bar. In addition, the evolution
of bar topography over time at various discharges will be described. Sand bars are
quite common in alluvial reaches of the Green River between the Jensen and Ouray
gages. The particular bar we have selected appears to be a typical example in every
way. The fluvial nature of this bar and associated reach will be investigated by
means of a fluid dynamic model, which will be described in the following section.

Model of Flow and Sediment Transport


The morphology of alluvial channels is ultimately the product of a strong
nonlinear coupling between channel geometry, the flow field in the channel, and the
transport of bed material by the flow. The first step in understanding the
development of channel topography and its response to variations in discharge is a
careful treatment of the various aspects of this coupling. The approach we have
taken treats the interactions between the flow, bed topography, and sediment
transport by combining three separate components of the full coupling, as follows:
(1) a flow algorithm capable of making accurate predictions of the velocity and
boundary shear stress fields in channels of known topography and planform, (2) a
method for determining sediment fluxes over the bed given the flow field, and (3) a
method for determining temporal changes in channel topography due to both
sediment transport and changes in stage. In this section, each of these three
components is briefly described, with emphasis on the physical effects which are
incorporated and, importantly, those which are neglected in our formulation.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

469

Andrews and Nelson

Finally, the interrelation of these three elements and their combination into a
topographic evolution model is briefly discussed.

The Flow Model


The fundamental requirement of a coupled flow-bed topography model is a
careful treatment of the fluid dynamic~. In particular, the response of the flow to
topographic nonuniformity must be treated accurately, because the forcing on the
flow field due to the nonuniformity of the bed has a significant influence upon the
topography. It has been demonstrated [Yen and Yen, 1971; Dietrich and Smith,
1983; Berg and Odgaard, 1988] that the convective accelerations due to planform
and topographic nonuniformity in channels are often of the same order of magnitude
as the pressure gradient and stress divergence in the local momentum balance.
Thus, a model that is based upon a perturbation expansion about a zero-order
steady uniform flow will not provide a satisfactory simulation of conditions in a
reach with complex topography. Accordingly, the non-linear treatment originally
developed by Smith and McLean [1984] and subsequently expanded upon and
applied to the mobile bed by Nelson and Smith [1989a, b, this volume] is employed
here. The mathematical development of this treatment is presented in detail in the
above work, and will not be repeated here. In order to interpret the results of the
evolution model presented below, however, it is useful to briefly restate the basic
assumptions used in developing the flow model. They are as follows:
1.
2.
3.
4.
5.

6.
7.

The flow is treated as quasi-steady.


The pressure distribution is assumed to be hydrostatic.
Lateral stresses and boundary layers are neglected.
The vertical velocity and length scales are an order of magnitude smaller
than the cross-stream velocity and length scales which are, in turn, an order
of magnitude smaller than the streamwise velocity and length scales.
The vertical structure of the zero-order streamwise velocity field is given by
a similarity structure (e.g., a logarithmic or quasi-logarithmic profile). One
of the consequences of this assumption is that, at lowest order, the
boundary shear stress is related to the vertically-averaged velocity field by
a drag coefficient (Note that this coefficient depends upon local roughness
and is not necessarily spatially constant).
Cross-stream secondary flows associated with both channel planform and
streamline curvature are evaluated.
Advection of streamwise and cross-stream momentum fluxes by the
secondary flows are neglected [although they may be included iteratively,
see Nelson and Smith, 1988b].

The inputs required for this computational model are the planform and
topography of the channel, the local roughness lengths in the channel, and the
discharge. The flow model computes three-dimensional velocity fields, downstream
and cross-stream boundary shear stresses, and water-surface elevation in the
channel. The predicted values can be expected to reproduce the actual values with
reasonable accuracy, provided that none of the assumptions above are violated. In
practice, the assumptions restrict the application of this model to channels where
depths are substantially smaller than widths and streamwise variations in width and
topography are gentle. Thus, the model is not valid for the case of rapid width
variations (Le., banks at angles greater than twenty or thirty degrees to the
downchannel direction), and cannot be expected to give good results if large-scale
bed slopes along the direction of fluid motion exceed ten degrees or so. These
situations, however, are relatively uncommon, especially in alluvial channels.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Topographic Response of a Bar

470

The assumption of quasi-steadiness in the flow model does not preclude


treatment of cases where discharge and bed topography vary in time. Rather, this
assumption limits the range of applicability of the algorithm to cases in which the
time scales associated with discharge variations and bed elevation changes are long
compared to the time scales of the local flow (e.g., the time required for a fluid
parcel to traverse the reach of interest). For the case of discharge variations, this
condition requires that the unsteadiness in the flow field makes a negligible
contribution to the local momentum balance or, equivalently, that the contribution
of the flood wave slope to the local surface slope is small. For temporal changes in
bed elevation, model validity is maintained as long as the rate of change in bed
elevation is much less than the flow velocity, a condition that is almost always
satisfied. Thus, although the discharge and bed elevation may vary slowly in time,
each may be treated as constant in the flow computation for a majority of natural
flows.
These approximations allow simplification of the equations expressing
momentum balance for the flow, and also lead to a straightforward iterative
technique for predicting bed evolution and the response of the bed to discharge
variation, as discussed below in greater detail.
As mentioned above, the flow model requires that the spatial distribution of
roughness length, zo, be specified as an input. The roughness lengths input to the
model must be representative of all sources of momentum extraction from the flow,
including grain saltation and form drag on bars and bedforms. When the sediment
size and bedform and bar geometries are known, the value of the overall roughness
parameter may be calculated using the method described by Nelson and Smith
l1988a]. Where some areas of the bed are much different than others (e.g., dunes
versus upper plane bed), this detailed approach should be used. When bedform
geometries are not well-known, but bed material and sediment transport rates are
somewhat similar over much of the bed, a simple alternative is to use the value of (0
( = zo/h) that yields the observed elevation drop over the reach of interest. This
approach distributes the roughness length in proportion to depth which, at least for
channels with duned beds, is a reasonable approximation. Choosing the value of (0
such that the observed elevation drop through the reach of interest matches that
predicted by the model ensures that the total channel drag is accounted for in an
integral sense. The potential error incurred using this technique is due to the
spatial distribution of roughness, rather than the magnitude. The simpler technique
was employed in all calculations presented herein.

Sediment Transport Model


The goal for the sediment transport component of a bed evolution model is an
equation or set of equations that enable one to calculate the sediment flux over each
point on the bed. In general, this requires the treatment of both suspended load
transport and bedload transport. In the case of bedload transport, any of several
different bedload equations may be employed. These equations predict the sediment
flux in the bedload layer given particle size, skin friction boundary shear stress, and
the critical shear stress necessary to initiate motion. These equations typically
contain empirical coefficients which are set using flume or field measurements. As a
result, each of the various equations typically works best over the range in which it
was calibrated, and the best bedload equation to employ in a given situation
depends upon the flow conditions and sediment size. Various bedload equations and
their ranges of applicability are discussed in detail by Wiberg and Smith l1989].
Where sediment is transported as both bedload and suspended load or primarily
suspended load, the fluxes of sediment over the bed depend upon the boundary shear
stress and sediment size, as well as the local flow velocities and turbulent
diffusivities. In general, the suspended sediment concentration in the flow depends
upon a balance between downward particle settling, upward turbulent diffusion, and

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

471

Andrews and Nelson

vertical and horizontal advection by the mean flow field. Although the full problem
is tractable, often the computations can be greatly simplified and treated in a less
computationally intensive manner. Where the suspended sediment concentration
profile responds to spatial variations in boundary shear stress over horizontal length
scales that are short compared to the spatial scales of the bed topography,
quasi-uniformity may be assumed.
This assumption implies that the local
concentration profile is in equilibrium with the local boundary stress. Under such
conditions, the analysis of Rouse J19371 is correct at the lowest order. This
approximation is similar to one use in the flow model described herein, where the
spatial distribution of acceleration and deceleration was assumed to have a weak
effect on the vertical structure of the primary flow.
Assuming a quasi-uniform vertical structure for the sediment concentration field
allows the application of a total sediment load equation, wherein both the bedload
and suspended load fluxes are calculated using a single expression. One of the
better known of these equations is the Engelund-Hansen [1967] equation, given by
q

.08

T*

5/2

(1)

where qs is the volumetric transport rate per unit width, Ps and P are the densities
of the sediment and fluid, respectively, d is the particle size in transport, T* is the
dimensionless shear stress, and f is the friction factor (drag coefficient).
Given the distribution of boundary shear stress and the particle size, one can
calculate the sediment flux at each location on the bed using the Engelund-Hansen
equation. A correction for the effect of form drag on the skin friction boundary
shear stress is an implicit part of this equation. As a result, this equation can only
be expected to retain validity where bedforms are similar to those in the
experiments used in calibrating the equation. In addition, this expression can be
expected to yield reasonable results only when the ratio of skin friction boundary
shear stress to the critical shear stress is large (large transport stage), because it
contains no critical shear stress dependence. Furthermore, the expression is
applicable only where the spatial scales over which bottom stress varies are long
compared to the distance a particle is advected while it settles.
The
Engelund-Hansen equation was calibrated using transport stages (ratio of boundary
to critical shear stress) and sediment sizes and size distributions similar to those
found in the Green River, and it can be expected to give good predictions of local
sediment fluxes in the case of interest.
Given the direction and magnitude of boundary shear stress at a point on the
bed, the local vector sediment flux was calculated using the Engelund-Hansen
equation. Then, the streamwise and cross-stream components of the local vector
sediment flux were determined. This approach neglects the relatively weak
redistribution of suspended sediment by secondary currents and is justified because
most of the suspended sediment is near the bed, where the flow direction is
approximately the same as the boundary shear stress. Because a vast majority of
the sediment flux is suspended, no gravitational correction was included.

Evolution of Channel Topography


Channel configuration may change with time as a result of two distinctly
different mechanisms. First, the channel may evolve due to erosion and deposition
of sediment which makes up the channel bed and banks. The so-called "erosion
equation" relates the flux of sediment to the rates of erosion and deposition of the
bed as follows:

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

'-fopographic Response of a Bar

472

(2)
where

1J

is the bed elevation, (

is the volumetric vector sediment flux and c

b
=N.65 is the percentage of the bed made up of sediment (Le., unity minus the

porosity).
Using computed sediment fluxes, the rate of erosion is easily calculated from
Equation (2). The erosion rate, in conjunction with an assumed time increment,
predicts the modification to the bed in that interval. .Thus, this equation may be
employed in an iterative manner to predict temporal adjustment of the bed to flow
conditions, as discussed below.
The other way in which the channel may be modified is through changes in
stage. Although this effect does not modify the actual bed topography (except
through erosion and deposition, as discussed in the previous paragraph), it does have
a profound effect on the planform and depth of the channel. For example, the
centerline trace of a typical meandering stream is typically more sinuous at lo\', flow
than at bankfull; the centerline is defined as the locus of points midway between the
two water edges. Similarly, the relative cross-sectional area at a given longitudinal
location in the reach may change dramatically from low to high flow. These effects
may seem obvious or even trivial, but they play a major role in determining the
difference in patterns of erosion and deposition at different stages in a given reach.
The method used in developing low flow initial channels from bankfull
topographic surveys is relatively straightforward.
First, a comprehensive
topographic map is constructed from surveyed points.
In order to obtain
approximate channel alignment and topography below bankfull .elevation, the
bankfull channel is "sliced" along a surface parallel to the bankfull free surface.
This surface is positioned at a distance below the bankfull free surface equivalent to
the desired stage drop. Points below the plane are interpolated using a tensioned
spline and resampled in a regular grid. The potential error in this approach arises
from assuming that the pattern of superelevation in the free surface is relatively
invariant with stage change. This approximation is never precisely correct, but the
error incurred in the depths is typically very small. Furthermore, the size of the
error can be calculated using the flow model, and a new free surface can be
calculated iteratively to improve the estimated depth. For the Green River reach
described herein, the error was typically less than two or three centimeters. This is
well within the uncertainty in the surveyed topography, and was assumed to be
negligible.
The technique described above also can be employed to predict variations in
channel form due to stage change while erosion and deposition are taking place on
the bed. In other words, the channel topography can be adjusted with each
increment of time to account for both erosion and deposition and stage changes in a
quasi-steady sense. This allows the .investigation of the response of the channel
topography in cases where the change in stage occurs over time scales which are
short compared to the bed modification and in cases where discharge and stage vary
slowly compared to the bed evolution.

The Evolution Model


The coupling of the various components of the full evolution model is shown in
Figure 3. As shown, the inputs to the evolution model consist of the following: the
initial planform and topography of the channel, the sediment size making up the
bed, and the stage and discharge as a function of time. Using the specified channel
geometry and initial discharge, the flow model described herein can be employed to
calculate the spatial distribution of boundary shear stress. These values are inserted

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Andrews and Nelson

473

MODEL INPUTS:
INITIAL CHANNEL PLANFORM
INITIAL CHANNEL TOPOGRAPHY
SEDIMENT SIZE
DISCHARGE OR STAGE/
LJISCHARGE AS A FUNCTION
OF TIME

FLUID
DYNAMICAL
MODEL

ENGELUND-HANSEN
TOTAL LOAD
EQUATION

CALCULATE

NO

CHOOSE l!:l t SUCH


THAT 6h h AND
CHANGE IN PLANFORM IS
SMALL. CALCULATE
NEW GEOMETRY.

Fig. 3.

QUASISTABILITY

YES

>--'-"'~~

II'

Iterative computation scheme.

in the Engelund-Hansen total load equation in order to obtain vector (i.e.,


downstream and cross-stream) sediment fluxes over the bed. The divergence of
these fluxes yields the rates o( erosion and deposition on the bed. If the stage and
discharge are constant, the erosion and deposition on the bed are the only temporal
changes. In this case, a time step is chosen such that the change in bed elevation is
a small fraction of the flow depth at each point. This specified. time step in
conjunction with the calculated erosion and deposition rates gives' the predicted
channel topography after one time step has elapsed. The new topography is then
input to the flow model, and the entire procedure is repeated in an iterative manner.
If the stage and discharge also vary in time, the changes to the bed are
calculated from the erosion equation as above, and then the new channel
morphology is calculated using the "slicing" technique described above. In this
case, the time step chosen is such that both the change in bed elevation and the
change in planform due to stage variation are relatively small.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Topographic Response of a Bar

474

Q=50
Maximum value

~.:::::..:::. ~

~ ~f~~
~

249.7 dy/cm 2

::

Q =275

Maximum value

92.5 dy/cm 2

f~~~%
~

-:;::..

___f_~

Q=475

Maximum value = 138 dy/cm 2

Fig. 4.

Distribution of boundary shear stress in the Ouray reach at discharges of 50 m 3 /s,


275 m 3 /s, and 475 m 3 /s.

The computation of topographic evolution process described above repeated


iteratively until either the desired time interval has passed or a quasi-steady state is
reached. (Quasi-steady state refers to the situation in which the bar form is
well-developed and has a constant shape, although it still may migrate
downstream.) For this investigation, the evolution models were used primarily to
determine the response of bed topography to various discharges, rather than to
calculate the steady-state topography.
Response of Flow to Channel Topography
Lateral and downstream variations in boundary shear stress and verticallyaveraged velocity are due to changes in channel area, channel curvature, and bed
topography. Channel area varies greatly through the Ouray reach, and has a
significant influence on the characteristics of flow and sediment transport. Bankfull
channel area varies from 580 m 2 near cross section 7 to 400 m 2 near cross section
36. Channel area decreases by 25 percent between cross section 7 and cross section
20, located at the bar crest. The distributions of boundary shear stress in the Ouray
reach computed for 3 discharges, 50 m3 /s, 275 m 3 /s, and 475 m 3 /s, are compared in
Figure 4. Similarly, the distributions of vertically-averaged velocity computed for

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

475

Andrews and Nelson

Maximum value

172 em/s

Maximum value = 104.6 em/s

Maximum value = 127.7 em/s

Fig. 5.

Distribution of vertically-averaged velocity in the Ouray reach at discharges of


50 m 3 /s, 275 m 3 /s, and 475 m 3 /s.

the same 3 discharges are compared in Figure 5. The bankfull discharge in this
reach is about 475 m3 /s, and the mean annual discharge is about 130 m3 /s. The
same bankfull channel topography, shown in Figure 2, was used for the calculations
compared in Figures 4 and 5.
Channel curvature through the Ouray reach, as indicated by the channel banks
or centerline, is moderate. Curvature of the streamlines as indicated by the
vertically-averaged velocity vectors shown in Figure 5, however, is appreciably
larger. The greater curvature of the streamlines compared to the banks is due to
the very substantial influence of channel topography. Steering of flow around the
bar is shown by the divergence of the boundary shear stress and vertically-averaged
velocity beginning near cross section 13. Topographic steering is also significant
further downstream where the point-bar on the left bank forces the high velocity
core to the right bank. Streamline curvature increases as discharge decreases,
especially in the vicinity of the bar. The apex of the bar is within 20 cm of the
water surface at a discharge of 275 m3 /s. Consequently, there is very little flow
over the top of the bar at this discharge, and most of the flow must go around the
bar. Lateral convective accelerations are especially large between cross sections 15
to 18.
Topographic steering of the flow has a substantial influence on the spatial
distribution of sediment transport rates. In the vicinity of cross section 9, the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Topographic Response of a Bar

476
5.0

0
0

4.0
3.0
2.0
1.0
0.0

5.0

4.0

Of

13
o

3.0

0::

<

:I::
U

2.0

rn

C
e-.

1.0

::::>

0.0

0 0
0

5.0

4.0
o

3.0

o
o

2.0
1.0

0.0
125.0

0.0

-125.0

DISTANCE FROM CENTERLINE [M]


Fig. 6.

Comparison of measured and computed lateral distribution of unit discharge at cross


sections 8, 13, 17, 23, 31 and 34 in the Ouray reach.

lateral distribution of sediment transport rate is nearly uniform at both 275 m3 /s


and 475 m3 /s. Over the apex of the bar near cross section 19, however, the lateral
distribution of sediment transport is strongly asymmetrical near the bankfull
discharge. As discharge decreases, the sediment transport field over the bar
becomes increasingly more asymmetrical and the cross channel component of the
sediment transport field increases.

Comparison of Measured and Computed Unit Discharge


The computed and measured unit discharges are in very good agreement through
the Ouray reach. This comparison is shown in figure 6, where the lateral

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

477

Andrews and Nelson


0

5.0

34

4.0
3.0
2.0
1.0
0.0

5.0

4.0

tJ

Of

tJ
d
0::

3.0

::t:

2.0

e-..

1.0

0.0

<

U
lfJ

31

5.0

23

4.0
3.0
2.0
1.0
0.0
125.0

0.0

-125.0

DISTANCE FROM CENTERLINE [M]


Fig. 6. cont.

distribution of unit discharge, determined from measurements made at six cross


sections using a temporary cable and boat, is compared to the lateral distribution of
unit discharge computed by the model. The largest errors, as one would expect,
occur at the upstream end of the study reach in the vicinity of cross section 5. This
result occurs because the lateral distribution of flow entering the study reach has
been assumed to be uniform, although this assumption is only an approximation.
In some instances, where the unit discharge is increasing or decreasing rapidly in
the cross stream direction, there appears to be discrepancies between the measured
and computed values. These apparent discrepancies occur because the model results

Copyright American Geophysical Union

Water Resources Monograph

478

River Meandering

Vol. 12

Topographic Response of a Bar

are based upon 13 verticals computed for each cross section, compared to the 20-30
measured verticals per cross section. Therefore, the computed lateral distribution of
unit discharge is somewhat smoother than the measured distribution.

Computed Transport Rate ofBed-Material through Ouray Reach


Similarly, the computed and measured rates of bed material transport through
the Ouray reach are in good agreement. The rate of bed-material transport through
each of the 41 cross sections in the Ouray reach was determined at a given discharge
by integrating the computed local values in a cross section. Except for discharges
around 275 m 3 / s, the cross-sectionally averaged sediment transport rates varied
significantly through the reach. An estimate of the amount of bed-material larger
than 0.0625 mm in diameter being transported through the Ouray reach at a given
discharge can be determined by averaging over the 41 cross sections. Then, the
relation between reach averaged bed-material transport and water discharge was
determined for 6 discharges between 100 m 3 /s and 475 m 3 /s. An estimate of the
mean annual bed-material load transported through the Ouray reach during the
period from 1982-86 was computed using the recorded water discharges at the
Jensen gage, and the locally determined bed-material transport relation. During
the period of relatively large discharges, 1983-86, the mean annual bed-material
load transported through the Ouray reach was computed to be approximately
3.1xl06 tons/yr. The mean annual load of bed-material sized sediment determined
by sampling at the Jensen gage and regression analysis was 2.7xl0 6 tonslyr during
the same period.
One would expect the mean annual load of sand-sized material transported by
the Green River through the Ouray reach to be somewhat larger than the quantity
transported at the Jensen gage because the concentration of suspended sediment in
the bottom 7.6 cm of the flow is not sampled by the equipment used at the Jensen
gage. Hence, the actual concentration of suspended sand at the Jensen gage is
slightly greater than the measured. This discrepancy is greatest at small discharges
and least at large discharges.
Evolution of Bar Topography in Response to Various Discharges
Adjustment of channel topography with time at a constant discharge was
investigated at 3 discharges, 50 m 3 Is, 275 m 3 /s, and 475 m 3 /s. Each calculation
began with the same bed topography, that which existed on July 15 and 16, 1986 at
a discharge of approximately 275 m3 Is. The techniques used to determine the
location of the banks and to compute erosion and deposition have been described
above. Each calculation covered 100 increments of 1728 seconds, or a total of 2
days.
Evolution of channel topography in the Ouray reach at a constant discharge of
275 m 3 /s is shown in figure 7 at 5 cross sections, numbers 16, 18, 20, 22, and 24.
These cross sections were selected to illustrate in detail changes in the bar
topography. During the spring of 1986, discharge through the Ouray reach peaked
at ",620 m 3 /s on June 8 and 9, and decreased gradually to ",280 m 3 Is on July 15
when the channel topography was surveyed.
As one would expect, the channel topography was nearly in equilibrium with a
discharge of 275 m 3 /s and computed changes in the bed elevation were everywhere
quite small. The model calculations indicate that sediment is deposited very slowly
on the bar surface and in the upstream portion of the primary channel. Conversely,
sediment is eroded very slowly from the downstream portion of the primary channel.
Predicted evolution of channel topography in the Ouray reach at a constant
discharge of 475 m 3 /s is shown in figure 8 at 5 cross sections, numbers 16, 18, 20,
22, and 24. Near the head of the bar (cross section 16), there is no significant

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Andrews and Nelson

479

DISTANCE FROM CENTERLINE [M]


125.0

0.0

-125.0

0.0
1.0
2.0
3.0
4.0

18

5.0
6.0
0.0
1.0
2.0
3.0
4.0

16

5.0
6.0

:r::
Eo-

0.0

0..

1.0

2.0
3.0
4.0

24

5.0
6.0
0.0
1.0
2.0
3.0
4.0

22

5.0
6.0
0.0
1.0
2.0
3.0
4.0
5.0
6.0

Fig. 7.

Predicted evolution of bar topography at 5 cross sections during a discharge of 275


m 3 /s for 2 days. Dashed line shows original topography and solid line shows final
topography.

change in bed elevation. Downstream from section 16, however, a complex pattern
of scour and fill develops. Significant erosion of the primary channel occurs along
the entire length of the bar. Between cross sections 18 to 20, the secondary channel
also erodes through a much slower rate than in the primary channel. Downstream
from cross section 20, located near the crest of the bars, sediment accumulates very

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Topographic Response of a Bar

480

DISTANCE FROM CENTERLINE [M]


-125.0

0.0

125.0

u.Q,.....-..,..--------L.....------------J
1.0
2.0
3.0

4.0

18

5.0
6.0

0.0 - + - - - r r - - - - - - - - - - l - - - - - - - - -

1.0
2.0
3.0

4.0
5.0

16

6.0
~

125.0
0.0 ;-_---.

..,...

1.0

E=
0..

2.0

0.0
.l--

-125.0
--,-_---...J

3.0

4.0
5.0
6.0

0.0 --t-----,,.....------.L....-------L--....,......------J
1.0
2.0
3.0

4.0
5.0

22

6.C
O.O+---~---------L-------r---.....J

1.0
2.0
3.0

4.0
5.0
6.0

Fig. 8.

Predicted evolution of bar topography at 5 cross sections during a discharge of 475


m 3 /s for 2 days. Dashed line shows original topography and solid line shows final
topography.

slowly in the secondary channel. The bar surface erodes slowly in the vicinity of
cross section 18. Further downstream, however, erosion of the bar surface ceases,
and a significant quantity of sediment accumulates over the apex of the bar.
Model calculations using a discharge of 475 m 3 /s indicate that the bar
topography is enhanced appreciably by flow near the bankfull stage. Sediment
accumulates on the bar surface and is eroded from the side channels, except in the

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

481

Andrews and Nelson

downstream part of the secondary channel where some sediment is deposited.


Calculations of bar evolution using a discharge of 550 m 3 /s reveal a similar pattern
of scour and fill, except the magnitude of bed elevation changes is greater and the
transition from scour to fill over the bar surface and in the secondary channel are
shifted downstream. Calculations of bar evolution using a discharge somewhat less
than 475 m 3 /s indicate that the magnitude of bed elevation changes would be
smaller everywhere and the transition from scour to fill over the bar surface and in
the secondary channel would be shifted upstream.
Predicted evolution of channel topography in the Ouray reach at a constant
discharge of 50 m 3 /s for a period of 2 days is illustrated in figure 9. At this
discharge, the bar apex stands nearly 1 m above the water surface and there is no
flow through the secondary channel. The entire river flow is confined to the
primary channel. A substantial quantity of sediment accumulates in cross section
16, located upstream of the bar apex. Progressing downstream to cross section 19
through 23, the thickness of accumulated sediment decreases. Comparison of
changes in bed elevation at these cross section after various intervals of time shows
that the zone of most rapid deposition is moving downstream.
At a discharge of 50 m3 / s, the reach averaged bed-material transport rate is
about one-tenth of the transport rate at a discharge of 475 m 3 /s. Nevertheless, the
rates of scour and fill are locally quite large compared to values computed at a
discharge of 475 m 3 /s. The local rate of scour and fill depends upon the divergence
of the sediment flux rather than its magnitude. As shown in figure 5, divergence of
the boundary shear stress, and, hence, the rate of scour and fill increases
significantly over the upstream face of the bar as discharge decreases. Strong local
divergence of the sediment flux is a result of a significant disequilibrium between
the flow and channel topography. The degree of disequilibrium is not unexpected
because initial bed topography was essentially in equilibrium with a discharge of
1\1275 m3 /s.

Discussion of Results
A comparison of aerial photographs taken of the Ouray reach between 1963 and
1987 shows that the bar we investigated has changed very little during this period.
The general size and shape of the bar which is exposed at low discharge, has
remained essentially the same. Although vegetated islands exist in this portion of
the Green River, vegetation has not become established on this bar. Finally, the
location of the bar has remained fixed, whereas channel bars are frequently observed
to migrate downstream in relatively straight reaches. As noted above, the Green
River floodplain in the vicinity of the Ouray reach has been constructed primarily
by the aggregation of vegetated islands. Therefore, some bars, although perhaps a
relatively small percentage, evolve from an unvegetated bar, such as the one we
investigated, to become a vegetated bar and finally, are incorporated into the
floodplain when the secondary channel aggrades to near the bankfull elevation.
Consequently, the stability of the bar we have investigated is perhaps greater than
one would expect.
The apparent stability of the bar indicates that the overall rate of bar evolution
is relatively slow compared to the length of time covered by the aerial photographs.
Calculated rates of bar evolution, however, indicate that the bar topography adjusts
rapidly over a period of several days to a few weeks when the discharge changes.
Maximum computed rates of scour and fill are on the order of a few tens of
centimeters per day. To some extent, the lack of long-term evolution in bar
topography is due to a balance between fill at one range of discharge against scour
at another range of discharge. For example, sediment was eroded from the primary
channel at flows greater than 275 m 3 /s and deposited in the primary channel at

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Topographic Response of a Bar

482

DISTANCE FROM CENTERLINE [M]


-125.0
0.0
0.0 -t-------~--""'"'"----.,....----------J
125.0

1.0
2.0

......... -.-

3.0
4.0

18

5.0
6.0

0.0 +1.0
2.0

~--""""--____:::I_---------J

....

3.0
4.0

16

5.0
6.0

:r:

125.0

b:

0.0

2.0
3.0

0.0

-125.0

-+-----~---...a....---__=r_--------l

1.0

4.0

24

5.0
6.0

O.O~----~---"""""----......,....----------I

1.0
2.0
3.0
4.0

22

5.0
6.0

0.0 -+---------,r-----...I...----...,.....---------J

1.0
2.0
3.0
4.0

20

5.0
6.0

Fig. 9.

Predicted evolution of bar topography at 5 cross sections during a discharge of 50


m 3 /s for 2 days. Dashed line shows original topography and solid line shows final
topography.

flows less than 275 m 3 Is. In contrast, however, sediment was deposited on the crest
of the bar at any discharge large enough to cover the entire bar to a depth of several
centimeters. Although the rate of sediment deposition on the bar crest is
substantially less than the maximum computed rate of scour and fill in the Ouray
reach, given sufficient time, about one week at a discharge of 475 m 3 I~, the crest of

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

483

Andrews and Nelson

the bar will grow to within ",10-15 centimeters of the water surface. In our
calculations, deposition on the bar crest ceases when the local boundary shear stress
is less than the critical value. In actuality, the mechanics of flow and sediment over
the bar crest become very complex when the local depth is less than 10-15
centimeters, and the model does not simulate locally important processes; advection
of suspended bed-material and wave-current interactions. Given appropriate flow
conditions and bar topography, suspended bed-material, especially the smallest bed
particle, can be transported from upstream areas with relatively large boundary
shear stress and deposited on the crest of the bar even when the local boundary
shear stress is less than critical.
As the bar crest approaches the water surface, wind-induced waves on the river
become an important factor limiting further growth of the bar. Waves breaking on
the bar crest resuspend the bed material, which then can be advected or diffused
away from the crest. Especially large waves are generated by interaction with the
river current when the wind is blowing upstream. Water waves with amplitudes up
to 20 centimeters are rather common in a channel as wide as the Green River
through the Ouray reach. Typically, water waves break when the depth is slightly
greater than the wave amplitude. Therefore, it appears likely that water waves
become a significant factor inhibiting further growth of the bar crest when the local
flow depth is less than roughly 20 centimeters. Furthermore, breaking waves may
erode the bar crest as discharge decreases after a peak, especially when the discharge
decreases slowly.
Wind erosion also may influence the elevation of the bar crest. During a
majority of the year, the bar crest is emergent and dry. Aerial photographs taken in
late summer and fall show that the bar crest is extensively modified by eolian
transport. The annual rate of eolian deflation, however, is unknown. A windstorm
with a velocity of 7 m 3 /s, a rather common occurrence in eastern Utah, would
deflate the dry bar surface at a rate of 2-3 em/day. Thus, the total eolian deflation
between successive spring floods may be as much as a few tens of centimeters. The
combined erosion of the bar surface by water waves and windstorms is probably on
the order of 0.5 meters per year. The predicted rate of sediment accumulation on
the bar crest at a discharge of 475 m3 /s is about 15 em per day. Between 1963 and
1987, a discharge of 475 m 3 /s occurred approximately 3 days per year. Therefore, it
is possible that the elevation of the bar crest is maintained approximately constant
over a period of years due to deposition from fluvial transport and erosion by water
waves and windstorms. The significance of bar erosion by water waves and
windstorms deserves further attention.
The downstream migration of channel bars in relatively straight channels with
uniform width and cross-sectional area is a commonly observed phenomena. As
noted above, channel width and cross-sectional area vary considerably through the
Ouray reach, this appears to explain the long-term stability of the location of the
bar. Predicted bar evolution at discharges of 275 m 3 /s and 475 m 3 /s indicate a
slight tendency towards downstream migration. At a discharge of 475 m 3 /s,
sediment is eroded very slowly from the upstream face of the bar and deposited very
slowly on the downstream face. At discharges less than 275 m 3 /s, the zones of both
scour and fill shift upstream and, thus, counteract the pattern at larger discharges.
When these calculations are repeated using a larger channel width and area in the
vicinity of cross sections 18-22, the magnitude of scour and fill over the bar surface
at 475 m3 /s is greatly enhanced, and the upstream shift in the pattern of scour and
fill at relatively small discharges is less pronounced. Therefore, the constriction of
channel width and area in the vicinity of cross sections 18-22 probably determines
the location of the bar.
The constriction of channel width and area in the vicinity of cross sections 18-22
also appears to explain the persistence of the secondary channel. As shown in
Figure 8, cross sections 16, 18, and 20, the secondary channel scoured rapidly at a
discharge of 475 m 3 /s. Conversely, at discharges less than 275 m 3 /s, model

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Vol. 12

Topographic Response of a Bar

484

calculations predict that the secondary channel will fill slowly. The rapid rate of
scour at a discharge of 475 m 3 /s apparently has been sufficient to maintain the
secondary channel. When the calculations of bar evolution at 475 m 3 /s is repeated
using a larger channel width and area. in the vicinity of the constriction, the rate of
scour in the secondary channel decreases. Thus, it appears that the constriction of
channel width and area is sufficient to prevent a net accumulation of sediment in
the secondary channel over a period of years.

Conclusions
The conclusions of our investigation are:
1. The fully nonlinear flow model described by Nelson and Smith rl988a,bl
predicts flow structure, sediment transport rates and temporal evolution ot channe
topography all of which are in good agreement with observed conditions in the
Green River near Ouray, Utah.
2. Over the past 30 years, construction and operation of a reservoir, as well as
natural variation in the snowmelt runoff, have caused substantial and persistent
year-to-year differences in the magnitude of discharge equalled or exceeded less
than 10 percent of the time in the Green River near Ouray, Utah. Annual sediment
loads have varied accordingly. The quantity of sediment transported by a given
discharge, and the approximate balance between sediment supplied and transported
within appreciable reaches, however, have not been affected. The bankfull channel
of the Green River has adjusted slowly to the large changes in annual flood
discharges.
3. The flow velocity field and distribution of boundary shear stress are
determined primarily by bed topography and longitudinal variations in channel
width. At all discharges investigated, streamline curvature exceeded channel
curvature. Stream line curvature increased significantly as discharge decreased.
4. Channel topography changed very slowly at a discharge of 275 m 3 /s, which
was approximately the discharge when the topography was surveyed. Hence, the
baseline topography appears to be nearly the steady state bed configuration for 275
m 3 /s.
5. Channel topography adjusts relatively quickly to discharges appreciably
different from 275 m 3 /s. Initially, divergence of the bed-material transport rate
were locally .large. Extensive scour and/or fill occurred at most cross sections at
discharges appreciably greater or less than 275 m 3 /s. Divergence of the sediment
transport field tended to decrease as bed topography evolved.
6. Compared to the baseline configuration, topography of the channel bar was
greatly enhanced following a simulated flow of 475 m 3 /s for 2 days. Sediment is
deposited on the bar surface and is eroded from the primary and secondary channels,
except in the downstream part of the secondary channel where some sediment is
deposited.
7. Compared to the bankfull topography, bar topography was diminished
following a flow of 50 m 3 /s for 2 days. The primary channel aggraded IVI m along
most of its length, while the bar crest and secondary channels were emergent.
8. Model calculations predict that sediment will be deposited on the bar crest at
all discharges large enough to cover the entire bar to a depth of several centimeters
or more. Erosion of the bar crest was not indicated at any discharge. Actual
deposition of sediment on the bar crest probably is slightly greater than predicted
by the model, because the advection of suspended sediment from areas of relatively
large boundary shear stress into areas of relatively small boundary shear stress is
neglected. The bar crest is most likely deflated by wind-induced waves and/or wind
transport of sand when the bar crest is emergent and dry.

Copyright American Geophysical Union

Water Resources Monograph

River Meandering

Andrews and Nelson

Vol. 12

485
Notations

cb
d

concentration of sediment in the stream bed


grain diameter

f
L
qs

friction factor in the Engelund-Hansen total load equation


flow depth
volumetric sediment transport rate per unit width

qs
t

vector volumetric sediment transport rate per unit width


time
bed elevation
fluid density

TJ

p
Ps
T

sediment density
dimensionless boundary shear stress

References
Andrews, E. D., Downstream effects of Flaming Gorge Reservoir on the Green
River, Colorado and Utah, Geol. Soc. Am. Bull., 97, 1012-1023, 1986.
Dietrich, W. E. and J. D. Smith, Influence of the point bar on flow through curved
channels, Water Resour. Res., 19, 1173-1192, 1983.
Engelund, F. and F. Hansen, A monograph on sediment transport in alluvial
streams, in Teknisk Verlag: Copenhagen, Denmark, Technical University of
Denmark, 63 pp., 1967.
Nelson, J. M. and J. D. Smith, Flow in meandering channels with natural
topography, in Parker, Gary (00.), this volume, Am Geophys. Union Monograph,
1989a.
Nelson, J. M. and J. D. Smith, Evolution and stability of erodible channel beds, in
Parker, Gary (ed.), this volume: Am. Geophys. Union Monograph, 1989b.
Parker, Gary and E. D. Andrews, On the time development of meander bends, J.
Fluid Mech., 162,139-156,1986.
Pucherelli, M. J., Green River channel mapping study, u.S. Bur. Reclamation,
Unpublished Letter Report, 13 pp., 1987.
Rouse, H., Modern conceptions of the mechanics of fluid turbulence, Trans. Am.

Soc. Civ. Eng., 102,463-543, 1937.

Smith, J. D. and S. R. McLean, Spatially averaged flow over a wavy surface, J.


Geophys. Res., 83, 1735-1746, 1977.
Smith, J. D. and S. R. McLean, A model for flow in meandering streams, Water
Resour. Res., 29(9), 1301-1315, 1984.
Wiberg, P. L. and J. D. Smith, A model for calculating bedload transport of
sediment, J. Hydraul. Div., Am. Soc. Civ. Eng. In press.
Yen, C. and B. C. Yen, Water surface configuration in channel bends, J. Hydraul.
Div. Am. Soc. Giv. Eng., 97(HY2), 303-321, 1971.

Copyright American Geophysical Union

The
Water Resources
Monograph
Series
Vol. 1

SYNTHETIC STREAM FLOWS


M. B. Fiering and B. B. Jackson (1971)
Available in microfiche only.
ISBN 0-87590-300-2.
Vol. 2

BENEFIT-COST ANALYSIS FOR


WATER SYSTEM PLANNING
C. W. Howe (1971), 144 pages
ISBN 0-87590-302-9.
Vol. 3

OUTDOOR RECREATION AND


WATER RESOURCES PLANNING
J. L Knetsch (1974)
Available in microfiche only.
ISBN 0-87590-304-5.

Vol. 7

URBAN STORMWATER
HYDROLOGY
D. F. Kibler (1982), 271 pages
ISBN 0-87590-308-8.
A practical guide 10 ament methods and models
used in analyzing di erenl types of stormwaler
management problems. Bridges !he gap between
current practices and new studies. A major reference wor1< for environmental researchers. praetidng
engineers. and urban planners.

Vol. 8

THE SCIENTIST AND ENGINEER


IN COURT
M. D. Bradley (1983), 111 pages
ISBN 0-87590-309-6.
To be an expert witness !he SCIentist or engineer
must have a wor1ong knowledge of !he judicial
process and courtroom procedures. ThlS volume offers a complete Introduction to !he role of an expert
wrtness In litigation proceedings.

Vol. 9

GROUNDWATER HYDRAULICS
J. S. Rosenshein and G. D. Bennett, (1984),
420 pages, ISBN 0-87590-310-X.
PrOVIdes stale-<lf-lhe-science insig I Inlo groundwater hydraulics and an applICation of hydraulics
loward solving groundwater problems. Principal areas covered are aquifer hydraulics. heat transport.
and modeling.

Vol. 4

MULTIOBJECTIVE WATER
RESOURCE PLANNING
D. C. Major (1977), 81 pages
ISBN 0-87590-305-3.
Vol. 5

GROUNDWATER MANAGEMENT:
The Use of Numerical Models
Y. Bachmat, J. Bredehoeft, B. Andrews, et
aJ (1985), 127 pages, 2nd ed.,
ISBN 0-87590-314-2.
Includes plannIng. ImplementalJOn. and adaptive
control of policles and programs relaled 10 !he exploralIOn. Inventory. development and operalJOn 01
water resources contatnlng groundwater

Vol. 6

METROPOLITAN WATER
MANAGEMENT
J. Gordon Milliken and G. Taytor (1981),
180 pages, ISBN 0-87590-307 -x.

Vol. 10

GROUNDWATER TRANSPORT: HANDBOOK OF MATHEMATICAL MODELS


I. Javandel, C. Doughty, and C. F. Tsang
(1984),228 pages, ISBN 0-87590-313-4.
ReViews, selects. and demonstrates !he best and
most practical mathemabCal models to predict !he
extenl 01 groundwater subsurface contamination.

Vol. 11

HILLSLOPE STABILITY AND LAND USE


R. C. Sidle, A. J. Pearce, and C. L.
O'Loughlin (1985) ISBN 0-87590-315-0.
This book emphasizes !he natural factors affect
Ing slope stability. including soils and geomorphic,
hydrologIC. vegetation. and seismic factors. II also
provides baSIC information on landslide classifica
tion. global damage. and analytical methods.
To purchase these volumes or to establish a stand/fig order 10f this series:
Write:

American Geophysical Union


2000 Ronda Avenue. .W.
Washington. DC 20009

Call:

800-4242488 toll-free
(202) 462-6900 (DC area or outside USA)
TWX (710) 8229300
FAX (202) 328-0566

Deals WTth desogn and ImplementalJOn of water


supply planning ,n areas suflenng from hmrted water resources. Social, environmental. and economic

costs are considered in this comprehensIVe evaJualIOn and analysIs of new and eXIsting alternative

waler strategies.

0-87590-316-9

Potrebbero piacerti anche