Sie sind auf Seite 1von 13

BOUNDARY SHEAR IN CURVED CHANNEL

WITH SIDE OVERFLOW


By Y. Ramsis Fares)
The characteristic changes in the boundary shear stress field of a channel
bend at the intersection with a side overflow are investigated in this paper. On the basis
of a field survey on bottom topography changes of the meandering river Allan Water at
the cut-off section, a detailed study of boundary shear stresses in an idealized rigid bed
model was undertaken. Both mathematical and experimental approaches were employed
in the idealized model study. The analysis of results showed that continual reductions in
shear stresses occurred in the bend at the side overflow region. The maximum reduction
of shear stresses was 37% in cases of low side overflows and 82% in cases of high side
overflows. These reductions were attributed to the development of stagnation and separation zones at the intersection associated with strong lateral outward currents. Based
on the idealized model study, the features observed in the bed topography of Allan Water
at the cutoff section are most likely to develop in cases of combined bend and high side
overflows.

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

ABSTRACT:

INTRODUCTION

Flood relief channel's schemes are commonly used to control high flows in natural river
systems. They develop naturally as well as artificially. Natural flood relief channels (known as
cutoff channels) develop as the river adjusts to its course during high flood periods. Along a
meandering river, zones of riffles and pools are continually formed at river banks as a result of
the helical motion induced in the flow by the channel curvature. If systematic variations in the
erosional and depositional activities occur, the river develops a highly irregular plan form.
Depending on the strength of the spiral flow and sediment properties of the channel structure,
the meander becomes overgrown and cutoff occurs. Thereafter, normal flows proceed along the
river's meander path and flood flows spill across the neck of the meander loop into the cutoff
channel. The time taken for the meander to reach this critical (cutoff) stage may vary from 2
to 10 yr (Gagliano and Howard 1983).
The second type of flood relief channel's scheme is artificial. These are usually constructed
at critical locations along the river to avoid long-term destabilization in the river regime. As
part of a fundamental study of flow mechanisms in meandering rivers during flood periods,
using flood relief channel's schemes, a field study was carried out on a meander loop/neck cutoff
intersection on the Allan Water, Perthshire, Scotland (Herbertson and Fares 1991). The cutoff
was constructed on the neck of the meander loop in September 1986. At the cutoff section of
the river, the Allan Water is 15 m wide, drains a catchment area of 210 km 2 , and has an average
annual flow of 6 mO/s. The cutoff channel is about 15 m wide, and 20 m long, and has a bed
level approximately 1 m above the river bed level. The position of the cutoff and the overall
geometry of the meander loop are shown in Fig. 1. Measurements of bed topography profiles
in the river were recorded at three sections namely (see Fig. 2): upstream (section A-A), the
middle (section B-B), and downstream (section C-C) of the upstream intersection between the
flood channel and the river, from 1987-1990, after the installation of the cutoff. [Figs. I and 2
represent profiles after the Herbertson and Fares (1991) study.]
The field investigation carried out on Allan Water showed the partial neck cutoff of a meander
loop results in a marked change in the bed topography from that usually found in a meandering
channel. The traditional riffle/pool bed profile was completely destroyed, and the following
features were observed (Fig. 1): (1) Development of a longitudinal bar at a distance of about
40% of the channel width from the front of the cutoff; (2) formation of a deep scour hole at
the downstream section of the river just beyond the cutoff intersection; and (3) migration of
the thalweg toward the inside of the bend. To understand the mechanisms by which the above
complex features are developed, an investigation was carried out on the characteristic changes
to bend flow in an idealized model situation (see Fig. 3). Based on the analysis of the local
changes observed in the secondary flow structure of the idealized bend model, an explanation
was provided for changes in bed topography due to the installation of the cutoff channel (Herbertson and Fares 1991). For the longitudinal bar to form, the typical one-cell structure of the
secondary flow had to be replaced by a two-cell structure rotating in an opposite direction, with
lLect., Dept. of Civ. Engrg., Univ. of Surrey, Guildford GU2 5XH, England.
Note. Discussion open until June I, 1995. To extend the closing date one month. a written request must be
filed with the ASCE Manager of Journals. The manuscript for this paper was submitted for review and possible
publication on July 30, 1993. This paper is part of the Journal of Hydraulic Engineering, Vol. 121. No. I. January.
1995. ASCE. ISSN 0733-9429/95/0001-0002-0014/$2.00 + $.25 per page. Paper No. 6664.
2

JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

circulation pattern

o. \

---

\ (

0.5

--

~--...
~

SECTION
A-A

-- -

\, \

to
,,

O.

: SECTION

...J

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

>
W
...J

0.5

W
In

Longitudinal Bar

to

circulation pattern

o.(__ ~_~
\

........"

"'-J9.~'O
->7,
......

/ I/

~ / . -- -

0.5 \ r-'.-1..J~."
\,

'-/

)
-

"

"~::::...~." -... - -/! /


~ '::'::':::"_'

._._J988

'-'-'

./'_._/

1.0

/ SECTION
I
CC

!/

'-,~'

'.,.. ~j

12

15

RIVER WIDTH(m)

- - - August 1987
_ . _ June 1988

_ .. _
June 1989
_______ April 1990

NOT TO SCALE
FIG. 1.

Allan Water-Meander Loop and Cutoff

FIG. 2.
tion

Bed Topography Profiles in Allan Water at Cutoff Sec-

one large cell between the bar and the inner bank and a smaller cell between the bar and the
outbank (see Fig. 2, section CoC)o This two-cell-type circulation in front of the cutoff was
responsible for the deposition of the loose material of the channel and for the initiation of the
longitudinal bar. The formation of the deep scour hole downstream of the intersection was
attributed to the strong downward currents of the outer bank cell, which caused the scouring
of bed material and generated local deepening close to the cutoff channel. The thalweg migration
was attributed to the inner cell of the secondary flow, which produced bank cutting on the inner
side of the bend and caused inward shifting of the thalweg. The latter conclusion was qualitatively
supported by measurements of flow patterns in meandering channels with overbank flows (McKeogh
and Kiely 1989).
However, the aforementioned complex flow mechanism does not directly explain the inevitable changes in the boundary shear stress field as a result of the intersection. This paper
investigates these changes in the idealized rigid bed model of the intersection (see Fig. 3) in
which entry to the flood relief (cutoff) channel is controlled by a side overflow. Both numerical
and experimental approaches are used in the study, which focuses on changes in the boundary
shear stress field for cases of combined bend and high (or low) side overflows. On the basis of
the rigid bed model study, an explanation for changes in the bed topography in Allan Water
and an assessment of the river regime is provided.

MATHEMATICAL FORMULATION OF BEND FLOW MODEL


The mathematical formulation is divided into two main stages. In the first stage, the formulation of flow around a gently curved channel bend in undertaken. This is followed, in the
second stage, by simulating the effects of the side overflow intersection on bend flow characteristics. For curved channel flows, the boundary shear stress field is resolved into longitudinal
and radial components because of the effects of secondary flow and boundary resistance. Consequently, two-dimensional (20) models are generally more conductive to the study of boundary
shear distribution in meandering channels. Uncertainties in shear stress calculations are found
to vary with flow depth, but are highly influenced by uncertainties in transverse velocities
(Siegenthaler and Shen 1983). Expressions for the boundary shear components are
( 1-3)
JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

Centre of
Curvature
FIG. 3.

Sketch for Idealized Rigid Bed Model of Intersection

where r, s, and z = cylindrical coordinates in the radial, longitudinal, and vertical directions
from an arbritrary origin at the bottom of the bend channel (see Fig. 3); dz = gradient operator
in z-direction; f.l = turbulent momentum exchange coefficient; U r and u, = velocity components
in r- and s-directions, respectively; and Tn T, = radial and longitudinal components of the total
shear stress To. In these expressions, shear stress components are explicitly introduced in terms
of me vertical gradients of longitudinal and radial flow momentum (by gradients of both U r and
uJ. In the formulation, the influence of these gradients on flow structure is introduced by a
depth-averaged model in which shear stresses can be calculated from the depth-averaged velocities (u",) rather than from local velocity components. By considering the momentum equation
in the longitudinal s-direction, the balance between the vertically integrated forces applied to
an elementary area (dr'ds) can be expressed as (see Fares et al. 1992) (where the main flow
momentum and the secondary flow momentum are in balance with the pressure gradient and
bed friction)

(4)
Where the functions <1>, and <1>2 are given by
(5. 6)

In (4) dr and d, = gradient operators in r- and s-directions, respectively; h = local flow depth;
R = local bend radius; Ss = longitudinal surface slope; U* = V (T)p) = shear velocity, with
p being flow density; and g = acceleration due to gravity. The functions <1>1 and <1>2 represent
the contribution of the longitudinal and secondary flow momentum to the motion. It is evident
from (4) that the convective transport of the longitudinal flow momentum by the secondary
flow is an important cause of redistribution of depth-averaged velocities; hence, of the shear
stresses. The evaluation of secondary flow contribution on velocity distributions in channel
bends was analytically investigated by Johannesson and Parker (1989). Coefficients (3 and CI. are
4

JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

the main and secondary flow convection factors for <P], <P 2, respectively. A combination of (5)
and (6) gives the following expression:

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

2
-<P =
<PI

R,,~

h=

R",

R,,~

(h)B (B)
R",
-

(7)

Where R,,~ = a/[3 is the ratio between secondary and main momentum coefficients. Eq. (7)
shows that <P 2/<P I decreases with increasing aspect ratio Blh and bend tightness R",/B, where R m
= bend mid radius; and B = width. Therefore, the more shallow the flow and the less tight
the bend, the smaller the contribution of the secondary flow for a given channel roughness. Fig.
4 gives the linear variation of the <P 2 /<P I ratio with the hiR m ratio for different bed roughness
coefficients, in which C is the Chezy coefficient of roughness. In both (7) and Fig. 4 the <P 2/<P I
ratio increases with increasing bed roughness and decreases with both the aspect and bend
tightness ratios. As expected, this indicates a greater dispersion of flow momentum in the
transverse direction for roughened channels and strongly curved flows.
Calculation Procedure

The solution for the distribution of depth-averaged velocity U m around the bend is numerically
obtained using a finite-difference scheme. The method was previously reported [see Fares (1992)
and Fares et at. (1992)]; therefore, the details are not repeated. By substituting (5) and (6) into
(4), the latter, after a slight transformation, results in
1
[
o,U", = -~ R,,~ O,(hu",)

h] +"2S, (l3u",gUm)
- h

+ 2r u'"

U'"
- 213h

(8)

C2

The equation is an explicit version of (4) for the streamwise variation of U m along the bend.
In the procedure, the channel bend is divided into a concentric mesh (see Fig. 5) in which at
each grid point [i, j] (where i = r-coordinate index and j = s-coordinate index) U m is computed.
The numerical calculations started at the inner bank and proceeded in radius increment, I1r,
across the channel to the outer bank. Then the computations proceeded longitudinally in steps
of 11s. For any finite-difference scheme, the consistency and accuracy of computations depend
on grid size, which in this case I1r and I1s. Preliminary test runs showed that 11 sections across
the channel width and 1.25 angle increments along the tested (60) bend were sufficient for the
C=20

.4

Side Overflow

Bend Strip

.3

Channel Bend

~b=60

~Ili/"-

=40
~

~tl FLOW
./

.2

.1

grtd structure for combined


bend + side overftow

grtd structure for bend


only conftguretlon

----.- dlreeUon

.0

.1

.2

.3

FIG. 4. Relation Between Flow Momentum Ratio $2/$, and hi


Rm Ratio for Different Values of Roughness Coefficient C

0'

computation
grid point
grid point In each strip

FIG. 5. Grid Structure Used in Numerical Calculations of TwoBend Situations


JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

computed results to approach consistency. The longitudinal surface slope S,[i, j] was evaluated
using the Chezy flow-resistance equation

u.,,[i, j]
gC 2h[i, j]

S,[i, j]

(9)

The initial value of S, at the bend entrance (i.e., at 8b


0) was assumed to match that of
the uniform flow at the upstream reach of the bend channel. Subsequent values of h[i, j + 1]
were then calculated from

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

h[i, j + 1]

h[i, j] - SJi, j]Li5

( 10)

Before proceeding towards the calculation of shear stresses, the predicted depth and velocity
profiles were checked against experimental data and found satisfactory [see Fares and Herbertson
(1990, 1991, 1992)]. The longitudinal shear stress component T, was calculated from

(11 )

The radial shear stress component Tr was calculated using an approach similar to that adopted
by Bouwmeester (1972) and Jansen et al. (1979). From the output of (8), and by using (1), (5)
and (6), a relationship between the depth-averaged velocity u m , modified by the radial flow
momentum, and T r was obtained. It reads

~p

= -2

~R'"
U C'2(1
2

C =

- C)
,

Ii
..j"kC

(l2a,b)

Where k = Von-Karman constant. Eq. (12) shows the lateral shear stress T r depends strongly
on U m , C, and the h/R ratio. For cases of bend flows of similar velocities and roughness
coefficients, T r increases as flow depth increases and as the degree of bend curvature decreases.
The preceding equation is only applicable to the interior region of the channel cross section,
away from the bank region. The effects of walls on bend flow behavior is usually limited to the
bank region and considered negligible for flows of aspect ratios B/h 2: 6, and for bend tightness
ratios R,,/B 2: 3. A relationship can be derived between the ratio of the radial to the longitudinal
shear stresses T)Ts and the corresponding ratio <1>2/<1>1 of the depth-averaged flow momentum.
By combining (11) and (12), with the use of (5) and (6), the following expression can be obtained:
<1>,

<I>~

(T)

k
= 2(C' - 1)

R,,~ ~

( 13)

This equation shows a direct relationship between the contribution of the secondary flow
momentum and that of the transverse boundary shear stress, normalized by their contribution
in the longitudinal direction. It can be deduced from (13) that the order of magnitude of the
transverse shear stress increases by increasing bed roughness and the intensity of secondary flow
in the motion. Finally, having determined the shear stress components T, [from (11)] and T r
[from (12)], the total shear stress To was calculated from (3).
Formulation of Bend Flow at the Side Overflow Region

In the case of flow in a combined channel bend and active side overflow, the discharge, dQfR,
over a step length, dCw, across the side overflow can be calculated from
dQfR _ A D (h
dC", I'
W

C )1.5
"

(14)

Where A w = weir coefficient, which includes velocity and discharge coefficients; D f = a coefficient, which simulates the drowning condition of flow over the weir (51); C" = weir crest
height; and h = water head in the bend channel. The streamwise variation of bend flow at the
intersection is simulated by the spatially varied flow equations, with decreasing discharge, along
the crest width of the overflow; that is
dh
dC",

[So - Sf - ef3g~ 1) (:~:) (u + d~fR) ]/[1 - f3(F2 + ~F)]

(15)

Where

(~r =~

(16)

S.. = bed slope of the channel; Sf = friction slope; u = mean velocity of the bend flow; F =
6

JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

Froude number ( = utv(gh; and A = area of bend flow (= Bh). Eq. (15) = basic equation
for modeling the lateral overflow effects on flow in the main channel. However, since the main
difference between flows in straight and curved channels is the effect of the superelevation and
secondary circulation, the direct application of (15) would lead to unrealistic solutions. Hence,
the equation, or rather the approach in which it is applied, has to be modified to include these
effects. To do this, the bend cross section is divided into a series of concentric strips (subchannels)
of equal width (see Fig. 5), each of which has a certain (mean) depth and a certain (mean)
velocity. Each strip can be treated separately and the model equation can be applied. The
advantage of using this approach is that each concentric strip contributes to the overflow according to its mean water depth. The contribution increases as water elevation increases and
strips nearest the bend's outer bank contribute most to the side overflow. This procedure's
validity depends mainly on the degree of bend curvature rm / B and the aspect ratio of the flow
B/h. The more shallow and less curved the bend, the more realistic the predictions obtained.
Calculation Procedure

The depth and velocity profiles at the side overflow were obtained by solving (14) and (15),
simultaneously, through a standard numerical method of integration [see Price (1977) and Chapra
and Canale (1985)]. For subcritical flow conditions in the bend, which are under downstream
control, computations should start at the downstream end of the intersection and work toward
the upstream end. However, in evaluating the depth and velocity profiles of the approach (bend)
flow, as described in the preceding section, the computations needed to be carried out in the
downstream direction. Hence, an assumption for the bend flow at the intersection was necessary
for the computations to proceed from the downstream end toward the upstream end. The
assumption in the approach was that the specific energy was constant for each bend strip. This
allowed the values of depth and velocity in each strip to be determined at the downstream end,
from which the computations start. This assumption does not imply a constant energy across
the bend width, since the radial variations of specific energy are implicitly introduced through
variations of both depth and velocity in each strip. The assumption of constant specific energy
was experimentally confirmed for the case in which the main channel is straight [Ranga Raju
et al. (1979) and Uyumaz and Muslu (1985)]. Thus, this study extends the use of this assumption
(in a modified form) to include the case of a gently curved channel.
At the overflow zone, the bend cross section is divided into a concentric mesh (see Fig. 5)
in which at each grid point [i, I], i = r-coordinate index and I = s-coordinate index along the
intersection. Depth and velocity profiles at each concentric strip were computed (the number
of strips is equal to the number of nodes in the radial direction minus 1). For the 60 bend, the
side overflow was located between 25 and 35 to allow significant development of superelevation
and secondary circulation in the flow. The initial values of depth and velocity, at each grid point
of the upstream end of the intersection, were fed from the output of the bend flow model [(8)
and (10)]. With the assumption of constant specific energy, computations were carried out in
each strip from the downstream end of the intersection toward the upstream end. The predicted
depth and velocity values in each strip were then projected on all grid points of the concentric
mesh at which values of shear stress were computed. Similar to the case of flow in the bendonly situation, the predicted depth and velocity profiles were checked against measurements
and were found satisfactory, particularly in the case of combined bend and low side overflows
(Qr :::; 0.4 and h r :::; 0.36, where Qr and h r are the ratios of discharge and depth between the
bend and side overflow) [see Fares and Herbertson (1990, 1992)]. By using the predicted values
of depth and velocity, the longitudinal T, and transverse T r shear stress components were calculated from (11) and (12), respectively; and the total shear stress To was finally calculated from
(3). Experience showed that 11 sections (i.e., 10 concentric strips) across the channel width and
0.2SO angle increments along the intersections were sufficient to maintain consistency in the
predictions.
LABORATORY INVESTIGATION OF PROBLEM

Experimental data collected from a laboratory model are used for quantitative comparisons
with the numerical predictions. Full details of the experimental apparatus and equipments used
in measurements are reported in Fares and Herbertson (1993); hence, only the principal dimensions of the experimental flume will be given. The flume consisted of a 60 bend working
section lying between two straight channel reaches, see Fig. 6. Both the straight and curved
sections were of rectangular cross sections, 0.5 m wide and 0.12 m deep, with Manning's n value
of 0.01. The bend was specified as a gentle bend (i.e., the mid bend radius to width ratio R",I
B was fixed at 3). A gap was left in the outer wall of the bend at the apex (see Fig. 6) between
bend angles 8b = 25 and 35 to accommodate the side overflow channel. Entry to the side
channel was controlled by a broad crested weir of varying crest levels to allow changes in water
elevation at the channel entry. The downstream condition of the side channel was uncontrolled
JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

SIDE WEIR

.. ,
, __~~---J
...L Ir---..---:r'-~--~
I I

...

~3O--J

.50

I.

varl.ble

~b= 60

.80m

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

Rm=1.5m

'irL. '--._.__ .+FLq.YL___.~

+__.

1-0-----

'.00

JlL-

0.5m

B
Rm =
~=

FIG. 6.

.~m

~i

1.5m

60'

Principal Dimensions of Laboratory Flume

FIG. 7.
TABLE 1.

Measurement Locations in Bend Section of Flume

Experimental Program

Channel Bend Data


Test run
(1)

ha

ua

(mm)
(2)

(lis)

hlRm
(5)

Blh

(3)

(m/s)
(4)

55.5
70.0
60.0
70.0
80.0

5.38
7.70
6.00
7.39
6.40

0.196
0.220
0.200
0.211
0.160

0.037
0.047
0.040
0.047
0.053

Side Overflow Data

QrR

Ch

Q,

(6)

F
(7)

R
(8)

(mm)
(9)

h,
(10)

(lis)
(11 )

(12)

9.010
7.143
8.333
7.143
6.250

0.266
0.266
0.261
0.255
0.181

9,889
14,000
10,909
13,427
11,636

25
25
45
45
65

0.55
0.64
0.25
0.36
0.19

2.24
4.08
0.76
1.68
0.64

0.42
0.53
0.13
0.23
(UO

BAI-BVI
BA2-BV2
AI-VI
A2-V2
A3-V3
A4-V4
A5-V5

"Mean value upstream of channel bend.

and outflow from the curved channel was controlled by an adjustable sluice gate. The width
ratio between the overflow and the bend Cw/B was fixed at 0.6.
The experimental program was divided into two sets of measurements - the first set was
taken with the side channel completely closed (i.e., a bend-only configuration) and the second
set with the bend plus active side overflow configuration. A total of 20 test runs were performed
for water surface profiles, and seven more for point velocities and deviation angles of the resultant
horizontal velocities [Fares and Herbertson (1993)]. The flow conditions used for shear stress
m,easurements were the same as those for velocities and deviation angles in the two bend
situations (see Table 1); two test runs for the bend-only situation and five test runs for the bend
with active side overflow. Fig. 7 shows measurement positions along the bend channel upstream
of, at, and downstream of the side overflow.
SHEAR STRESS IN BENDONLY CONFIGURATION

Figs. 8 and 9 give comparisons between the predicted and observed shear stress profiles along
the channel bend at bend radii R/R", = 0.9, 1.0, and 1.1, for test runs BA1-BV1 and BA2BV2, respectively. In the figures, the total shear stress To is normalized by the shear stress at
the upstream straight reach of the bend Tn, (= pghSo). The bands, R/R"" are chosen to correspond
with locations near the inner bank, at the centerline, and near the outer bank of the bend,
respectively. In general, the profiles confirm the previous findings reported by, for example,
Ippen and Drinker (1962), Yen (1970), Engelund (1974), Varshney and Garde (1975), Zimmermann and Kennedy (1978), Nouh and Townsend (1979), and Chen and Shen (1983). Both
predicted and observed shear stress profiles show a gradual increase across the width of the
channel, with maximum stresses lie along the inner side of the bend. This trend is induced by
the increase in velocity gradients near the inner bank relative to those near the outer bank. In
the second half of the bend, the shear stresses at the inner side of the bend (R/R", = 0.9)
8

JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

2
prediction -

1.8

experiment4

1.6

<Il

...

0.8

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

0.2

... ... ... ...

\0-

...

... ... ... ... ...

... ... ... ...

0.6

RiRm = 0.9

0.4

... RiRm = 1.0

Sb

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

2.------~,

1.2

~o

... ... ...

,
I

0.8

0.6

+
+

RiRm = 0.9

... RiRm = 1.0

5 10 15 20

~
~

1,

0.8

RiRm = 0.9

... RiRm = 1.0

Sb
FIG. 10. Shear Stress Distribution in Bend with Low Side Overflow Situation, Test Run A3-V3, Q, = 0.13 and h, = 0.25

...

:
:.
r
I

,
:,
,
,

0.2 ...

527.53032.535 40 45 50 55 60

experiment...

=_~:
t~

1.2

j,
limit of numerical
model application

prediction -

:I

---*--;lIr---'

0.6

ZONE

;::~
.----r-.
0.4

... RiRm = 1.1

,, ...

--i SIDE OVERFLOW :--

1.8

...

RiRm=1.1

FIG. 9. Shear Stress Distribution in Bend-Only Situation, Test


Run BA2-Bv2

experiment...

1.4

...

0..1-....--.-----,....
s 0 - ,....S-2Q....--2....S-30....--35....--40c--45r--SOr--SSc--:60't::;;l
0

prediction -

1.6

RiRm = 0.9

... RiRm = 1.0

0.2

... RiRm = 1.1

1.8

...

0.8

2...--------...------,-------,

0.4

'0

FIG. 8. Shear Stress Distribution in Bend-Only Situation, Test


Run BA1-BVl

0.2

...

0,J....,0'--""I5- ' T
5--,3,....0---r35-4""'0-4""5--::5!:0--:'5r:'5"""":'60~
10-1...5-2,...0-2....
Sb

'to'

experiment...

0.6
0.4

1.6

1.2

...

prediction-

1.4

... ... ... ... ... ...


...
... ... ...

1.2

~
\'

--

1.4

...

1.8

RiRm = 1.1

:
i

'

,,
,
'

i': limit of numerical

,,~

,,: model application

5 10 15 20 ;527.53032.5;5 40 45 50 55 60
Sb

FIG. 11. Shear Stress Distribution in Bend with Low Side Overflow Situation, Test Runs A5-V5, Q, = 0.10 and h, = 0.19

decrease, and those at the central (R/R m = 1.0) and outer (R/R m = 1.1) regions continue to
increase. This pattern can be attributed to the outward flow momentum, which causes outward
shifting to the maximum (inner) velocities toward the outerside of the bend [Kalkwijk and De
Vriend (1980) and De Vriend (1981)]. For the central 60% of the channel width, maximum
stresses were persistent along the inner bend side (R/R m = 0.9). This is because the tested
length of the bend (8 b = 10 ~ 51n was insufficient for the secondary flow strength to have
considerable effect on shear stress distributions. Nevertheless, some outward skewness in the
shear profiles can be seen along the inner region (R/R m = 0.9) in the second half of the bend
(from 8b = 30 onwards).
SHEAR STRESS AT BEND/SIDE OVERFLOW INTERSECTION

The discussion of boundary shear distributions in this bend configuration follow two distinct
lines: quantitative and qualitative analyses. Quantitative analysis deals with comparisons between
the predicted and observed profiles, and the qualitative analysis focuses on changes to the shear
stress distribution as a result of introducing the side overflow. Results from only 3 test runs,
namely A3-V3, A5-V5, and AI-VI, are presented to correspond with conditions of high. medium, and low overflow crest levels, respectively. These are plotted in Figs. 10, 11, and 12 for
bend radii R/R m = 0.9,1.0 and 1.1. The full set of results is reported in Fares et al. (1992). As
in the case of the bend-only configuration, the total shear stress To is normalized by the shear
stress To.- of the straight upstream reach of the channel bend. For the bend region downstream
of the overflow, only experimental results will be given since the region is outside the scope of
the numerical model.
JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

I
:
I

1.6

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

~
~

1.2

experiment ...

II

:,

: limit of numerical
: model application
I

I
I

I
I
I
I

: ...

0.6

= 0.9
0.4
... R/Rm = 1.0
0.2 + RJRm = 1.1

0.8

prediction -

:
I

... ... ...

ZONE

_----~

1.4

en

~ SIDE OVERFLOW :.-.

1.8

RJRm

... ...
+

o""""r--r----r---r---,r--;----r---r--r--T----T---'p.-.--....
--r--,I-!

o 5 10 15 20 2527.53032.535 40 45 50 55 60
eb

FIG. 12. Shear Stress Distribution in Bend with High Side Overflow Situation, Test Run A1-V1, Or
0.42 and hr = 0.55

Combined Bend and Low Side Overflows (Q r


and A5-V5)

:"S

0.4, h r

:"S

0.36) (Test Runs A3-V3

In the region upstream of the intersection (for 6" = 0 ~ 20, Figs. 10 and 11), similarity
between the overflow and nonoverflow situations is seen between the predicted and observed
shear stress profiles. The predicted profiles behave in a manner similar to those in the bendonly configuration. Progressive development of secondary currents associated with a superelevation at the water surface causes high velocities, and high stresses, to develop along the inner
side of the bend, as explained in the preceding section. In general, there is very little effect
produced by the side overflow on shear stress profiles in this region.
Along the intersection (61) = 25 ~ 35), there is gradual reduction in shear stresses, particularly in the region close to the overflow (R/R m = 1.1). High stresses continue to occupy the
inner side of the bend (R/R m = 0.9). This mechanism may be explained as follows: as a result
of introducing the side overflow to the curved channel, the outward cross currents combined
with the preexisting secondary flow in the bend enhance the outward shifting of momentum
and diminish the superelevation at water surface. This process increases across the bend width
toward the overflow and causes more advection of momentum into the side channel. Consequently, a reduction in bend velocities and shear stresses occurs, particularly along the inner
side of the bend (R/R m = 0.9). The reduction in stresses is manifested by the formation of a
stagnation zone along the inner bank of the bend, opposite the overflow, in addition to the
(already existing) separation zone at the upstream corner of the side overflow channel (6" =
25). The effect of the separation zone on the shear stresses at the outer region of the bend (at
RIRm = 1.1) can be detected from the noticeable departure of the measured profiles from the
predicted ones, particularly along the first half of the intersection (from bend angle 6" = 25
-> 30).
To evaluate the local reduction in stresses at the overflow section, the experimental and
theoretical values plotted in Figs. 10 and 11 are presented in Table 2. These are given as the
percent reduction in shear stresses normalized by those at the upstream section of the overflow,
i.e., % reduction of To/(-ro)6" = 25, where (To)6" = 25 is the shear stress To at 6" = 25, for
bend radii RIR", = 0.9, 1.0, and 1.1, respectively. It can be seen that reductions are always
greater at the outer region of the bend (RIR m = 1.1) than at the central (RIR", = 1.0) and
inner sides (RIR", = 0.9). For test run A3-V3, the maximum reduction in shear stresses is found
to be 32% (observed) and 37% (predicted). For test run A5-V5, the maximum reduction is
20% (observed) and 31 % (predicted). The closer the section to the outer bank, the more the
advection of momentum into the side channel. For test run A3-V3, the predicted values of Tol
(To)6" = 25" are greater than the observed ones by about 4-5% for bend radii RIRI", = 0.9
and 1.1, and by about 14% at RIR", = 1.0. For test run A5-V5, the average values of predicted
reduction in stresses are higher than those of the measured ones by about 12%.
For the bend region downstream of the overflow (6" = 40 -> 50), the maximum shear
10

JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

TABLE 2.

Percent Reduction of To/(T o)6b

Test Run
A1-V1

Bend angle

C)

0.9

(1 )

(2)

27.5
30.0
32.5
35.0

19.21
35.83
50.60
63.38

27.5
30.0
32.5
35.0

10.55
22.36
39.90
67.27

1.0
(3)

R/Rm

1.1

(4)

0.9
(5)

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

27.00
49.54
67.95
82.20

13.05
29.43
47.90
48.57

27.45
54.06
81.32
82.17

(6)
=

5.92
11.84
17.10
23.03

(b) Percent reduction of T"/(ToW,,

RiRm

1.0

(a) Percent reduction of To/(T")tI,,


23.17
42.92
59.67
73.50

Test Rune
A5-V5

Test Run b
A3-V3

R/Rm

tl b

= 25 at Side Overflow Region

0.9

(7)

(8)

1.0
(9)

1.1
(10)

25 (theoretical prediction)
8.26
15.70
23.22
30.50

1.1

9.65
18.99
28.04
36.68

4.94
9.88
14.20
19.14

6.35
12.70
19.05
25.m;

7.77
15.53
23.20
30.48

5.88
8.40
14.29
11.76

8.18
10.18
17.18
19.45

25 (experimental measurements)

5.15
9.56
16.18
19.12

6.02
11.94
15.75
16.55

8.72
21.15
29.99
31.78

0.76
5.34
7.63
7.63

"Side overflow condition for high overflow: h, = 0.55; Q, = 0.42.


"Side overflow condition for medium overflow: h, = 0.25; Q, = 0.13.
"Side overflow condition for low overflow: h, = 0.19; Q, = 0.10.

stresses at the inner side of the bend (R/R", = 0.9) decrease gradually, and those along the
center (R/R", = 1.0) and outer side (R/R", = 1.1) remain virtually constant. This pattern
continues steadily until a uniform value of stress, across the bend width, is eventually realized.
Maximum shear stresses would be expected along the outer bank for relatively longer channel
bends in a manner similar to that of the bend-only configuration.
Combined Bend and High Side Overflows (Q r > 0.4, h r > 0.36) (Test Run A1-V1)
Upstream of the side overflow (6 h = 0 ----> 20), it can be seen from Fig. 12 that, maximum
stresses lie along the inner side of the bend. There is a strong tendency toward a uniform stress
distribution across the width of the channel, which indicates the features developed in the flow
at the intersection have been propagated upstream. Similar to the pattern of flow along the
intersection in cases of low side overflows, the outward advection of velocities, combined with
a reduction in the superelevation of the water surface, results in a nearly uniform distribution
of shear stress across the width of the bend just upstream of the overflow region (61) = 20---->
25). This explains the poor performance of the numerical model in this region, since no account
was taken in the formulation of effects propagating upstream from the intersection.
Along the intersection (61) = 25 ----> 35), shear profiles are greatly influenced by the strong
cross currents generated in the flow and the development of stagnation and separation zones.
The decrease in stresses along the inner side is a direct result of the stagnation zone, and the
reduction of stresses along the outer side is caused by outward lateral currents. For this test
run, the separation zone is found to occupy almost half the width of the channel bend and about
halfthe width of the side overflow channel (Fares and Herbertson 1993). The combined reduction
in flow depth and velocities in the bend, at the side overflow region, causes a drop in shear
stresses, particularly in the second-half of the intersection beyond the separation zone (61) =
30 ----> 3SO). Table 2 gives percent values of these reductions along the intersection, expressed
by the shear stress ratio To(T o)61> = 2SO, at radii R/R", = 0.9, 1.D, and 1.1. As in the case of low
side overflow, the reduction in shear stresses is higher in the outer bend region (R/R", = 1.1)
than in the central (R/R", = 1.D) and inner (R/R", = D.9) regions. The observed reductions vary
between ~82% at R/R", = 1.1 and ~67% at R/R", = D.9 for 6h = 35. In general, the predictions
are overestimated at the central and outer regions of the bend and underestimated at the inner
region. The difference between the predicted and observed shear stress ratio TO< T o)6 h = 25,
for the inner and central regions at 61> = 3D.no and 32.SO, varies from ~ 11 % to 13%, and the
overall difference is small, ~4%.
Downstream of the overflow (i.e., at 6h = 40 ----> 50), stresses along the central (R/R", =
1.D) and outer (R/R", = 1.1) regions of the bend increase and those along the inner side (R/R",
= D.9) diminish substantially. The reduction in shear stresses along the inner side of the bend
is a consequence of the continual growth in the size of the stagnation zone downstream of the
overflow. As a result, high shear stresses become concentrated in the outer half of the bend
cross section. This concentration of stresses indicates that, in a natural river situation, severe
bed scour is likely to occur in this region. This result was experimentally justified by observing
the high velocities of high outward deviations near the channel bed [see Fares and Herbertson
( 1993)].
JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

11

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

BED TOPOGRAPHY CHANGES IN ALLAN WATER

On the basis of the investigation carried out on the rigid bed model, an attempt will be made
to provide an explanation for bed topography changes in the Allan Water river, at the cutoff
section. For combined bend and low side overflow conditions, the river is expected to be governed
by the bend flow regime with an occasional spillage of flood flow across the neck of the cutoff.
In contrast, changes in bed topography are most likely to develop in cases of bend flows combined
with high side overflows. The occurrence of both longitudinal bar and local scour hole in the
bed topography of the river (Figs. 1 and 2) may be qualitatively explained in terms of shear
stress distributions obtained from the idealized model. This can be achieved by matching locations of the features developed in the river, with those of shear stresses in the idealized model.
A comparison of the appropriate locations in Figs. 1 and 2 with those in Fig. 13 and Table 2
showed a reasonably good correlation between the location of the longitudinal bar and that of
minimum shear stresses along the intersection (at R/R m = 1.1 and (l" = 32.5" ----> 35). By
applying the same analogy for the scour hole location downstream of the intersection, a zone
of high shear stresses in the bend region (l" = 40 ----> 50, along radii R/Rm = 1. 0 and 1.1, was
detected.
Finally, with respect to the migration of the thalweg toward the inside of the bend, the idealized
model study showed that the shear stresses along the inner side of the bend (RfR m = 0.9)
continue to fall rapidly to insignificant values beyond the intersection (Fig. 12 (l" = 40 ----> 5(n.
This would indicate that the formation of a bar is likely in this region, which would eventually
lead to the complete blockage of the meander loop and hence, to the development of the oxbow
lake (Allen 1965). The observed inward migration of the thalweg may be caused by the change
in the secondary flow cell structure, from one cell to two cells, rotating in the opposite sense
(Fig. 2). The change in the cell structure of the secondary flow is initiated by the longitudinal
bar development in front of the cutoff (expressed by a zone of minimum shear stresses, at
R/R = 1.0 and (l" = 32.5 ----> 35 in Fig. 12). The bar, once it reaches a considerable length,
splits the bend flow and produces the two-cell pattern. This, in turn, forces the flow to move
away from the cutoff producing active cutting on the inner side of the bend. It is evident,
however, that the existence of zones of high and low shear stresses in the river, at the cutoff
section, causes reduction in the river width and encourages the flow to take the shorter route
across the cutoff. Further investigations into the time and length scales involved in the mechanism
by which the meander is adjusted to the cutoff situation, to provide more rigorous assessment
of the long-term stability of the river regime, are recommended.
I1l

CONCLUSIONS

The object of the paper was to determine the changes occurring in the boundary shear stress
field, in a curved channel, resulting from the introductions of a side overflow on the outer bank.
Despite the fact that the study focused only on the hydraulic behaviour in an idealized rigid
bed model situation, it provided a mechanism for qualitatively evaluating the features developed
in the bed topography of a river system at the meander loop/cutoff section. Both numerical and
experimental approaches were employed in the rigid bed model situation. The analysis showed
the shear stress field is strongly dependent on depth and discharge ratios between the two
channels. The following conclusions may be drawn from the rigid bed model study:
A numerical model was suggested for simulating the shear stress distribution in a channel
bend with a side overflow. The model was based on the combination of a depth-averaged bend
flow model and a spatially varied flow equation. The basic features of flows in curved channels,
the superelevation and secondary circulation, were incorporated in the model formulation. Based
on the analysis of results, the overall difference between predicted and measured shear stress
values at the side overflow region was found to be ~ 7%.
For all discharge Qr and depth h r ratios tested, the introduction of the side overflow on the
outer bank of the channel bend caused a reduction in the boundary shear stresses. A maximum
reduction of shear stresses was always found along the outer side of the bend (R/Rm = 1.1),
close to the overflow.
For combined bend and low side overflow conditions (Qr :S 0.4 and h r :S 0.36), little effect
was produced by the side overflow on shear stress profiles in the bend region upstream of the
intersection. Along the intersection, a gradual reduction was found in stresses, with high stresses
continuing to occupy the inner side of the bend (R/R = 0.9). The maximum reduction of shear
stresses in this region was 32% (measured) and 37% (predicted). In the bend region downstream
of the overflow, the stresses along the inner side of the bend (RfR = 0.9) decreased steadily,
and those at the central (R/R m = 1.0) and outer side (R/R m = 1.1) zones remained virtually
constant.
For combined bend and high side overflow conditions (Qr > 0.4 and h r > 0.36), the effect
of overflow on the shear stress field was significant in all bend regions of the intersection. A
tendency towards a uniform stress distribution was observed in the upstream region as a result
of diminishing the superelevation at the water surface and the outward shifting of bend velocities
I1l

I1l

12

JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

toward the overflow. Along the intersection, the strong lateral outward currents and the formation of a stagnation zone (along the inner bank of the bend) and a separation zone (at the
upstream corner of the side overflow) caused substantial reduction in shear stresses. The observed reduction of stresses was 82% in the outer side of the bend (R/R m = 1.1) and 67% in
the inner side (R/R m = 0.9). Downstream of the overflow region, the shear stresses continued
to diminish considerably along the inner side of the bend, with an accompanying increase in
those along the outer region of the bend cross section, as a result of the continual stagnation
zone growth beyond the intersection.

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

ACKNOWLEDGMENTS
The
service
for his
helpful

APPENDIX I.

numerical calculations of boundary shear stresses were carried out using the central HP-UNIX computer
at the University of Surrey. The writer wishes to thank Dr. J. G. Herbertson of the University of Glasgow
suggestions throughout the study. The writer also wishes to thank the anonymous reviewers for their
comments.

REFERENCES
Allen, J. R. L. (1965). "A review of the origin and characteristics of recent alluvial sediments." Sedimentology,
5(2), 89.
Bouwmeester, J. (1972). "Basic principles for the movement of water in natural and artificial water courses."
Internal Note, Delft Univ. of Technol., Delft, The Netherlands.
Chapra, S. C, and Canale, R. R. (1985). Numerical methods for engineers with personal computer applications.
McGraw-Hili Book Co., Inc., New York, N.Y.
Chen, G., and Shen H. W. (1983). "River curvature-width ratio effect on shear stress." Proc., Int. Conf. on
River Meandering, ASCE, New York, N.Y., 687-699.
De Vriend, H. J. (1981). "Velocity redistribution in curved rectangular channels." J. of Fluid Mech., Vol. 107,
423-439.
Engelund, F. (1974). "Flow and bed topography in channel bends." J. Hydr. Div., ASCE, 100(11), 1631-1648.
Fares, Y. R. (1992). Modelling of the cross flow strength in channel bends." Proc., Int. Conf. on Protection and
Development of the Nile and Other Major Rivers, Water Res. Ctr., Nile Res. Inst., Qanater, Egypt, Vol. 1/2,
3.2.1-3.2.15.
Fares, Y. R., and Herbertson, J. G. (1990). "Partial cut-off of meander loops-a comparison of mathematical
and physical model results." Proc., Int. Conf. on River Flood Hydr., Wiley and Sons, Chichester, England,
289-297.
Fares, Y. R., and Herbertson, J. G. (1991). "Formulation of the flow characteristics in a channel bend with a
side overflow." Proc., 24th Congr. of IAHR, IAHR Publ., Delft, The Netherlands, Vol. A, AI733-A740.
Fares, Y. R., and Herbertson, J. G. (1992). "Performance of side channel overflows at river bends. " Proc., Int.
Conf. on Protection and Development of the Nile and Other Major Rivers, Water Res. Ctr., Nile Res. Inst.,
Qanater, Egypt, Vol. 2/2, 8.3.1-8.3.15.
Fares, Y. R., and Herbertson, J. G. (1993). "Behaviour of flow in a channel bend with a side overflow (flood
relief) channel." J. Hydr. Res., 31(3), 383-402.
Fares, Y. R., Laufs, W., and Herbertson, J. G. (1992). "Boundary shear changes in a channel bend at flood relief
(cut-off) channel intersection." Rep. No. CE-FM/FWL/92.1, Dept. of Civ. Engrg., Univ. of Surrey, England.
Gagliano, S. M., and Howard, P. C (1983). "The neck cut-off oxbow lake cycle along the lower Mississippi
river." Proc., 1nt. Conf. on River Meandering, ASCE, New York, N.Y., 147-158.
Herbertson, J. G., and Fares, Y. R. (1991). "Bed topography changes produced by partial cut-off of a meander
loop." Proc., Eur. Conf. on Adv. in Water Resour. Technol., A. A. Balkema, Rotterdam, The Netherlands,
113-120.
Ippen, A. T., and Drinker, P. A. (1962). "Boundary shear stresses in curved trapezoidal channels." 1. Hydr.
Div., ASCE, 88(5), 143-179.
Jansen, P. P. et al. (1979). Principles of river engineering, the non-tidal alluvial river. Pitman Publishers Ltd.,
London, England.
Johannesson, H., and Parker, G. (1989). "Secondary flow in mildly sinuous channel." J. Hydr. Engrg., ASCE,
115(3), 289-308.
Kalkwijk, J. P. T., and De Vriend, H. J. (1980). "Computation of the flow in shallow river bends." 1. Hydr.
Res., 18(4),327-342.
McKeogh, E. J., and Kiely, G. K. (1989). "Experimental study of the mechanisms of flood flow in meandering
channels." 23rd Congr. of 1AHR, IAHR Publ., Delft, The Netherlands, Vol. B, B/491-B/498.
Nouh, M. A., and Townsend, R. D. (1979). "Shear stress distribution in stable channel bends." J. Hydr. Div..
ASCE, 105(10), 1233-1245.
Price, R. K. (1977). "A mathematical model for river flow - Theoretical development." Rep. No. INT 127,
Hydr. Res. Wallingford Ltd., Oxfordshire, England.
Ranga Raju, K. G., Prasad, B., and Gupta, S. K. (1979). "Side weir in rectangular channel." J. Hydr. Div.,
ASCE, 105(5),547-554.
Siegenthaler, M. C, and Shen, H. W. (1983). "Shear stress uncertainties in bends from equations." Proc., Int.
Conf. on River Meandering, ASCE, New York, N.Y., 662-674.
Uyumaz, A., and Muslu, Y. (1985). "Flow over side weirs in circular channels." J. Hydr. Engrg., ASCE, 111 (1),
144-160.
Varshney, D. V., and Garde, R. J. (1975). "Shear distribution in bends in rectangular channels." J. Hydr. Div.,
ASCE, 101(8), 1053-1066.
Yen, C. L. (1970). "Bed topography effect on flow in a meander." J. Hydr. Div., ASCE, 96(1), 57-73.
Zimmermann, C., and Kennedy, J. F. (1978). "Transverse bed slopes in curved alluvial streams." 1. Hydr. Div.,
ASCE, 104(1),33-48.
JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

13

APPENDIX II.

NOTATIONS

Downloaded from ascelibrary.org by GADJAH MADA UNIVERSITY on 03/03/16. Copyright ASCE. For personal use only; all rights reserved.

The following symbols are used in this paper:

area of bend flow ( = Bh);


weir coefficient [in (14)];
total width of bend;
Chezy coefficient of roughness;
nondimensional coefficient of roughness [ = y/gl( kC)];
crest height of side overflow;
total width of side overflow;
drowned flow reduction factor (s 1);
Froude number of bend flow [=u/Y/(gh)];
acceleration due to gravity;
mean and local flow depth;
water head ratio (=hw/h);
water head above weir crest ( = h - CI;
indices for r- and s-coordinates, respectively;
Von Karman constant (=0.41);
s-coordinate index at overflow region of bend;
n - Manning's coefficient of roughness (= 0.01);
discharge in bend channel;
Q
discharge in flood (side overflow) channel;
Q1R
discharge ratio (= QfR/Q);
Q,
local and mid bend radii, respectively;
R,R",
R
Reynolds number of bend flow;
ratio between secondary and main momentum coefficients ( = a/!3);
Ra~
r, s, Z
radial, tangential, and vertical cylindrical coordinates, respectively;
channel bed slope;
So
friction slope;
SI
longitudinal surface slope in channel bend;
S,
mean velocity in bend;
U
depth-averaged velocity in bend;
Um
velocity components in r- and s-directions, respectively;
U,., U.\
shear velocity (=Y/(T)p;
u*
coefficients for secondary and main flow momentum, respectively;
a, !3
step size along s-direction at overflow zone;
~C"
step sizes along r- and s-directions, respectively;
~r, ~s
gradient operators in r, s, and z directions;
an an az
local bend angle;
81>
total bend angle ( = 60);
61>
turbulent momentum exchange coefficient;
JJflow density;
P
radial, longitudinal, and total shear stress components;
T,., T.\., To
bed shear stress at 61> = 25;
(T o)61> = 25"
bed shear stress upstream of bend ( = pghSo); and
functions for main and secondary flow momentum.

Subscripts
b

fR
r
w

14

bend;
flood relief (side overflow) channel;
ratio; and
weir

JOURNAL OF HYDRAULIC ENGINEERING

J. Hydraul. Eng., 1995, 121(1): 2-14

Das könnte Ihnen auch gefallen