You are on page 1of 14

BROAD-CRESTED WEIR

By Willi H. Hager, 1 Member, ASCE, and Markus Schwalt 2


Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

ABSTRACT," The flow features over the broad-crested weir with vertical upstream
wall and sharp-crested corner are analyzed experimentally. Only the long-crested
weir is considered, for which the discharge coefficient remains practically constant.
For a relative overflow depth between 10% and 40%, the surface profile, the bottom
pressure profile, the boundary separation profile, and the velocity profiles close to
the upper corner are self-similar, provided effects of scale may be dropped. For
extremely long-crested weirs, undular flow occurs. The first wave profile is shown
to be identical with the solitary wave profile. The main properties of the undular
hydraulic jump are explored. The broad-crested weir is characterized by insensi-
tivity to tailwater submergence. The modular limit is "found practically constant at
75% of the tailwater level, independent of the relative head on the weir. The
discharge-head relation for submerged flow is analyzed under a novel approach.
Finally, recommendations are specified under which a broad-crested weir may be
used as a discharge measurement structure.

INTRODUCTION

The b r o a d - c r e s t e d w e i r is a simple discharge m e a s u r e m e n t device. Its


main features are:

9 G e o m e t r i c a l simplicity of s h a p e i n v o l v i n g a s q u a r e e l e m e n t of l e n g t h
L~ and h e i g h t w (Fig. 1).
9 Distinct types of flow d e p e n d i n g o n r e l a t i v e length of w e i r ~w =
Ho/Lw.
9 A l m o s t constant discharge coefficient Ca for the " b r o a d - c r e s t e d w e i r "
t y p e with 0.1 < ffw < 0.4.
9 E x t r e m e l y low sensitivity on t a i l w a t e r s u b m e r g e n c e .
9 L o w - c o s t o v e r f l o w structure.

Actually, the b r o a d - c r e s t e d w e i r is u s e d o n l y for less i m p o r t a n t and sec-


ondary works, as it is n o t c o n s i d e r e d v e r y a c c u r a t e in t e r m s of discharge
m e a s u r e m e n t and is hydraulically t o o p o o r as an o v e r f l o w structure. B a s e d
on a literature r e v i e w , H a g e r (1993) was able, h o w e v e r , to p r o p o s e this w e i r
type as an additional s t a n d a r d s t r u c t u r e p r o v i d e d the f o l l o w i n g c o n d i t i o n s
are satisfied:

9 S h a r p - c r e s t e d u p s t r e a m w e i r corner.
9 Vertical u p s t r e a m face.
9 S m o o t h and h o r i z o n t a l w e i r surface.
9 W e i r length Lw such that 0.1 < ~ < 0.4.
9 M i n i m u m o v e r f l o w d e p t h ho = 50 m m .
9 R e c t a n g u l a r and straight a p p r o a c h and t a i l w a t e r channels.

lSr. Res. Engr., V A W , ETH-Zentrum, CH-8092 Zurich, Switzerland.


2Res. Engr., VAW, ETH-Zentrum, CH-8092 Zurich, Switzerland.
Note. Discussion open until July 1, 1994. To extend the closing date one month,
a written request must be filed with the A S C E Manager of Journals. The manuscript
for this paper was submitted for review and possible publication on July 15, 1992.
This paper is part of the Journal of Irrigation and Drainage Engineering, Vol. 120,
No. 1, January/February, 1994. 9 ISSN 0733-9437/94/0001-0013/$1.00 + $.15
per page. Paper No. 4385.

13

J. Irrig. Drain Eng. 1994.120:13-26.


Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

ho i~ili~~iiiiiiii!iiii!jii~i!iiiiiii!iiii!iil
iiiiiiii iii~ ~_~

I~ LW ~1
FIG. 1. Broad-Crested Weir, Notation

9 Prismatic channel geometry.


9 Tailwater submergence below modular limit.

Compared to the sharp-crested fully aerated weir, the following advan-


tages prevail:

9 Structural stability up to 2- to 3-m head may easily be accomplished.


9 Concrete may be used as material with steel upstream corner for
abrasion protection.
9 Reduced tailwater submergence effect.
9 No nappe aeration must be provided.

As disadvantages, this overflow structure is less analyzed, its discharge


coefficient is some 30% lower, and more material is needed. However, as
the sharp-crested weir is used for heads up to 0.6 m, or maybe to 1 m as
upper limit, and the ogee-overflow structures are normally too costly for
overflow heads below 2 - 3 m, the broad-crested weir could fill in the gap.
Its application is not necessarily restricted to Third World countries, as the
simple structure may also be interesting in planes where submergence is a
concern.
The present paper aims at adding to the actual knowledge and specifies
the recommendations mainly in view of a variant to existing discharge mea-
surement structures. The review is rather short as a separate paper was
presented (Hager 1993).

REVIEW

The discharge equation of the broad-crested weir is


Q = Cab(2gH3) 1/2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)
where Q -- discharge, Ca = discharge coefficient, b = channel width, O
= gravitational acceleration and 11o = ho + Q:/(29b2[ho + w] 2) = approach
energy head. Note that Ca is related to Ho, and not to the approach flow
depth ho (Fig. 1). Thus, the effect of velocity of approach, that is ho/w with
w = weir height, is implicitly contained.
According to Horton (1907), the discharge coefficient depends exclusively
on the relative weir length ~ = Ho/L,,, provided effects of viscosity and
surface tension may be neglected. For water, the typical limit head HL is
some 30-50 ram. For Ho > IlL, the discharge coefficient may be expressed
~4

J. Irrig. Drain Eng. 1994.120:13-26.


for the so-called confined weir according to (5) presented later. For ~w
0, the asymptotic discharge coefficient is Cdo = 0.326.
A distinct feature of the broad-crested weir is the corner separation, which
was analyzed by Keutner (1934) and Moss (1972). Its length was found to
Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

be 0.77ho, and its maximum height is 0.15ho. Tracy (1957) was able to
generalize the surface profile using ho as normalizing parameter, provided
0,1 -< ~w -< 0.4.
Govinda Rao and Muralidhar (1963) classified the flow over the broad-
crested weir as:

9 0 < ~w -< 0.1 for long-crested weir.


9 0.1 < ~ -< 0.4 for broad-crested weir.
9 0.4 < ~ -< 1.5 for short-crested weir.
9 1.5 < ~w for sharp-crested weir.

Their data refer to unconfined weir flow and deviate from the usual discharge
equation, therefore. According to Singer (1964) the effect of weir height
Ho/w may be neglected if Ho < w/2. Further, a number of limits concerning
the approach flow depth, channel width, weir height, and crest length were
specified. Harrison (1964) pointed at the importance of sharp crested up-
stream corner and analyzed the effect of Reynolds number for the round-
crested weir (Harrison 1967).
Ranga Raju and Ahmad (1973) studied the broad-crested weir in both
the prismatic and the converging channel. Crabbe (1974) expanded the
application limits as proposed by Singer (1964) in terms of weir length and
weir height, and Sreetharan came up with limits as wide as 0.08 < ~w < 5.6
and 0.006 < Ho/w < 4. For large values of ~w, aeration of the lower nappe
is essential, and a significant advantage of the broad-crested weir is dropped.
According to Ramamurthy et al. (1987) the upstream corner of a broad-
crested weir may be considered sharp provided the radius of curvature is
smaller than R < 0.094w. Thus, extreme sharpness of corner radius on the
flow is not necessary.

EXPERIMENTS

The experiments were conducted in a horizontal rectangular channel


499-mm wide and 700-ram high. The total length of channel was some 7 m,
with the front side of glass, and the bottom and back side of black PVC.
The broad-crested weir of weir height w = 401 mm and length Lw = 500
mm was located at 4,828 mm downstream of the inlet, and various elements
to improve the approach flow were provided upstream of the weir. Thus,
an excellent, smooth, and waveless flow was obtained that could be analyzed
accurately. The tailwater submergence was adjusted by a flap gate located
at the channel end 7.5 m from the inlet. The discharge was measured with
a 90~ notch to the nearest 0.1 mm.
Surface profiles were observed with a precise point gauge ( + 0.1 mm).
Particular attention was paid to standing wave patterns at low overflow
depth, and it was important to collect accurate surface readings, as the
typical flow depth was 10-20 mm only. Overflow heads between 27 mm
and 205 mm were generated in order to study the main flow features.
Eighteen pressure taps were located along the centerline of weir, including
seven at the vertical front face and 11 at the horizontal weir crest. The
maximum distance between the taps was 100 mm, but the spacing was much
15

J. Irrig. Drain Eng. 1994.120:13-26.


closer at the upstream sharp corner to obtain insight in the separation
process. The upstream corner had the maximum sharpness that could be
finished with conventional PVC. An estimated radius of curvature was some
0.1 mm. The pressure taps were connected with a manometer battery where
Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

bottom pressure could be read to • mm. In general, the oscillations in


the manometer tubes were less than • 1 mm. The manometer was also used
to inject dye to the flow. From the point of view of visualization, fluorescine
was found optimal and was used at taps located close to the sharp corner.
Velocities were measured with a miniature propeller meter, at selected
sections and each 10 mm in height. As the probe was not turnable relative
to the horizontal, only the streamwise component was observed. N o dynamic
velocity and pressure readings were taken.

USUAL WEIR FLOW

In total, 14 extended series of observations were conducted. In all series,


the surface profile and the bottom pressure profile was measured. Table 1
gives a summary of the main flow characteristics. The Froude number Fo
is based on the approach velocity Vo = Q/(b[ho + w]) and the Reynolds
number Ro on the velocity (gHo) 1/2. The length scales are (ho + w) and Ho,
respectively, such that Fo = Q/[gb2(ho + w)3] '/2 and Ro = (gH3o) v2 1)-1
with v = 1.15.10-6m2s - 1 as kinematic viscosity for water of 15~ temper-
ature. The Froude numbers are very low, whereas turbulent flow occurs in
all cases. Series with < < a > > following the run number involve no surface
waves. A preliminary experiment indicated almost 2 - D - f l o w features such
that all remaining flow characteristics are related to the centerline.
Fig. 2 shows some typical free surface profiles for flows without surface
waves. The profiles for low discharge have a local minimum downstream
from the sharp corner, whereas for larger discharge, the flow gradually
decreases from the approach to the tailwater channel. All cases are referred
to typical or usual flow over broad-crested weir as 0.10 < ~w < 0.40.
The data in the vicinity of the upstream weir corner (x = 0) were further

TABLE 1. Main Characteristics of Extended Experiments


Q ho 1-Io
Run (Is-') (rnm) (mm) c~, Fo Ro
(1) (2) (3) (4) (5) (6) (7)
1 8.25 50.8 50.9 0.324 0.0174 4.13.104
2 4.55 34.2 34.2 0.325 0.0101 2.27-104
3 3.32 27.8 27.8 0.323 0.0076 1.67' 104
4 3.77 30.3 30.3 0.323 0.0085 1.90-104
5 6.26 42.1 42.1 0.327 0.0136 3.11-104
5a 6.26 42.1 42.1 0.327 0.0136 3.11" 104
6a 10.90 60.7 60.7 0.329 0.0222 5.39-104
7a 17.81 84.1 84.4 0.328 0.0337 8.83" 104
8a 25.98 107.9 108.4 0.329 0.0458 12.86.104
9a 37.59 138.2 139.2 0.327 0.0607 18.71" 104
lOa 54.83 176.2 178.0 0.330 0.0800 27.05.104
lla 68.07 202.1 204.7 0.332 0.0930 33.36" 104
12 3.60 29.3 29.5 0.322 0.0082 1.83-104
13 3.15 27.0 27.0 0.321 0.0072 1.60.104
14 4.19 32.2 32.3 0.326 0.0094 2.09' 104

16

J. Irrig. Drain Eng. 1994.120:13-26.


Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

- 0.5

- 0.4

l
414 416 4'.8 5 512 514m

FIG. 2. Free Surface Profiles for Various Runs (Table 1 ) (~r is Location of Pressure
Taps)

-1 0 I 2
1 . . . . x/Ho- ,

0 =======================================================
i!!i...........................................................
u ..............................................................
6 1

FIG. 3. Main Dimensionless Flow Features: Surface Profile y(X) for 11o (mm) =
(<a) 42.1, (~>) 84.1, (v) 138.2, and (A) 200.3; Bottom Pressure Profile yp(X) (Dark
Symbols); (..-) Boundary Separation Curve y,(X); and Velocity Profiles U for
Ho (mm) = (e) 121, (o) 181

analyzed by plotting the surface profile y ( X ) with y = h/Ho and X = x/Ho


where h --- flow depth with respect to the weir crest. According to Fig. 3,
the data describe a smooth curve starting at y ( X ~ - ~) = 1, passing the
origin at y(0) = 0.90 and tending to an asymptotic value y,,,(X > > 0)
0.46. The curve is almost symmetric a b o u t the average height yo = (1 +
0.46)/2 = 0.73 at location X , ~ 2/3, and m a y be a p p r o x i m a t e d as

Y - Y" - Tgh(Xo - X ) ................................. (2)


1 + c. - y.

Herein, c, = 0.03 is a correction factor for - 1 < X < + 2 and Tgh = the
tangent-hyperbolicus function. F o r larger X one may set co = 0, and the
correct asymptotic boundaries are reached.
The bottom pressure head curve yp(X) with yp = p/(pgHo) where p =
pressure and p = density shows some distinct features (Fig. 3), namely an
almost constant pressure p,,,/(pgHo) = 0.56 to 0.57 in the separation bubble
just beyond the upstream weir corner, a gradual rise to the m a x i m u m PM/
(pOlio) = 0.73 at XM = 1.05 and a subsequent decrease towards the free

17

J. Irrig. Drain Eng. 1994.120:13-26.


surface profile. This feature is in relation to streamline curvature and also
to the impact of primary current on the weir crest at X ----XM. The bubble
pressure is in agreement with Moss (1972), who found p,J(pOHo) = 0.58
Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

and pM/(pgHo) = 0.69.


The corner separation profile hs(x) or ys = hs/Ho of X between primary
and secondary flows was established with dye injected in various pressure
taps. The results of several experiments including two discharges and three
taps showed remarkable agreement (Fig. 3), The maximum separation co-
ordinates (X~M; YsM) = (0.44; 0.20) may vary by _+0.01. Following the
approach of Hager (1991) one may set X* = X(exp[-1]/XsM) = 0.836X
and y* = y , ( e x p [ - 1]/ysM) = 1.840y, such that the transformed separation
profile is given as
y~* = -x* In x* ...... , ..................................... (3)

This shape is in agreement/with the lower nappe of a sharp-crested, fully


aerated weir. The coordinates of separation crest are (0.44; 0.20) for the
present problem and (0.25; 0.112) for the sharp-crested weir. The ratios of
crest coordinates are 0.25/0.44 = 0.568 and 0.112/0.20 = 0.560 and thus
practically constant (when setting 0.446 instead of 0.440 as estimated, full
agreement may be obtained). Extrapolating with the value 0.66 for the
sharp-crested weir for the end of separation bubble yields 0.66/0.56 = 1.18
for the broad-crested weir (Fig. 3).
The horizontal velocity component u was plotted as U = u/(29Ho) v~ for
two experiments. The data reveal almost perfect similarity (Fig. 3) such that
a single line could easily be drawn through the data. At X = - 0 . 5 the
maximum velocity occurs close to the surface. This pattern is dramatically
modified over the weir crest due to streamline curvature. At the upstream
corner section X = 0 the velocity increases from the surface to the bottom,
except for a small boundary layer. At X -- + 0.5 the pattern is analogous
except for the reverse velocity in the boundary separation zone. The latter
could not be analyzed in detail as no appropriate velocity probe was avail-
able. At X = + 1 streamlines are curved differently with the center of
curvature above the flow. As a result, the velocity maximum is shifted to
the surface. At X = + 2, finally, streamlines are almost parallel and a fairly
uniform velocity distribution may establish.
The velocity distribution at the separation crest section was further an-
alyzed, in order to compare it with the velocity distribution over a sharp-
crested weir. The radius of crest curvature was estimated to R/Ho ~ 1.2.
Further, with a forward flow depth y(X = 0.5) = 0.78 the surface velocity
may be computed to U = (1 - 0.78) 1/2 = 0.469 using Bernoulli's equation,
as compared to 0.475 from observations. The maximum velocity occurs at
the interface between primary and secondary flow at an elevation Z =
z/Ho = Ys = 0.20 (Fig. 3). With a pressure head Ym = 0.565 one deduces
U, = (1 - y,)1/2 = 0.66 as compared to 0.60 at a slightly higher position
Z = 0.22. Clearly, the maximum computational value may not be reached
as viscous effects modify the theoretical profile.
Following again Hager (1991) and assuming a potential vortex for the
crest velocity distribution, one obtains relative to the surface layer 0.47(1.2
+ 0.58) = 0.83 = U(1.2 - 0.20 + Z) with y - y, = 0.78 - 0.20 = 0.58
as the forward flow depth and Z as height above the weir crest. Solving for
U yields
18

J. Irrig. Drain Eng. 1994.120:13-26.


0.83
U(Z) - 1.0 + Z fory~<Z<y ............................ (4)

For the separation height Z = 0.2, the velocity obtains U(0.2) = 0.69 as
Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

compared to 0.60 from observations. The velocity peak is of course not


reproduced. At Z = 0.4, the observation U = 0.56 and the computation
U = 0.59 compare already better. The velocity distribution thus follows the
potential vortex.
According to Hager (1993), the discharge coefficient Ca in (1) may
be expressed for Ho > HL only as a function of relative weir length ~ =
Ho/Lw as

Cd = Cdo(1 -- '5) -~ (1 1 + ~4 ............................ (5)

with + = 2/9. Thus, for very long weirs, ~w --~ 0 and Cd = (?do, whereas
for the sharp-crested, fully aerated weir Cd(~w --~ o~) = (1 -- +)-~ (?.do =
Cd=. Based on a literature survey Cdo = 0.326 was found. Deviations more
than 1% (Ca = 0.329) occur for ~w > 0.45. In the present experiments with
~ < 0.41, Cd/Cdo varied between _ 1.5%, and the coefficient of discharge
is surprisingly constant (Table 1). The minimum value Ca(~w = 0.05) =
0.321 was some 1.5% lower than Cdo due to scale effects. On larger models
with Ho > 50 mm, such effects do not occur, as is well known from com-
parable studies.

UNDULAR WEIR FLOW


A particular feature of broad-crested weirs is undular flow on the crest
as ~w < 0.1 (Govinda R a o and Muralidhar 1963). The origin of undular
flow is viscosity, as the flow is decelerated on long crests and acceleration
occurs only upstream of the drop into the tailwater channel. Fig. 4 shows
typical undular flow involving one to four wave crests. A n y other number
of wave crests up to 10 could be generated by modifying the approach energy
head Ho.
In all cases, a minimum flow depth Y,m occurred upstream from the first
wave crest at location Xum with index u relating to undular flow, and index
rn indicating the minimum. The data )turn and Yum for a particular value
Ho(cm) (Fig. 5) reveal a change at Ho ~ 3.5 cm. For Ho < 3.5 cm, the flow
is governed by a scale effect, whereas Yum = 0.46 remains almost constant
otherwise.
The flow was further studied in terms of number nw of wave crests for
the particular length of weir Lw = 500 mm. Fig. 6(a) shows the relation
nw(Ho) and indicates an increasing wave number with decreasing head Ho.
The domain of head for a particular wave number to occur increases with
decreasing n~, as is indicated by the dashed lines. The first wave maximum
hum increases as the crest number nw decreases [Fig. 6(b)].
The first wave profile was analyzed with the nondimensional coordinates

X, - xu - X,m . .......................................... (6a)


X u M -- Xu m

hu - hum
Yu = h u ~ - h,m " . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (6b)

19

J. Irrig. Drain Eng. 1994.120:13-26.


Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

(a) I

00
0 0 0 0 0
D 9 9 9 9 9 9 9 9

(b) } I t

ooo

O Q O g O 0 0 0 0 0 0 0 0
a g O 0 0

(c)
,cm, .

t
9 0 ~ ooo
o o o o ~ ~ 1 7 6 1 7 6

1~ 9 9 9 9

I
(d) 490 ' ' tcn l

FIG. 4. Undular Flow on Broad-Crested Weir: (a) One Wave Crest; (b) Two Wave
Crests; (c) Three Wave Crests; and (d) Four Wave Crests

4 0.6
Xurn\ Yurnl '
\
o\o\
I
I ~ 0.5

5
I , , Ho , f Ho
2 0A
0 4 8 12 b) o 4 8 i2
a)

FIG. 5. Minimum Flow Depth for Undular Weir Flow: (a) X.m; and (b) Y.m as Func-
tion of Approach Energy Head Ho (cm)

involving the minimum (x..,; h.m) and the maximum (x.~; h.M) wave heights
of the first wave (Table 2). Fig. 7 shows that the wave profiles y.(X.) have
a general trend, except at the wave rear side depending on whether addi-
tional (n~ > 1) or no more wave peaks follow.

20

J. Irrig. Drain Eng. 1994.120:13-26.


10 10
nw nw\i. '
),'t\ \x
Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

\t\., \\
\w,,
\ \
\*\
\ x

"~ H \.~ g 9 huff


0 I I I 0 1 , 1 ,

a) o 2 [cm] 4 _, 0 2 4 [cm} 6
b)
FIG. 6. (a) Number of Waves nw(Ho); and (b) First Wave Amplitude h,,M(n3

TABLE 2. Characteristics of First Undular Wave


Run 1 2 3 4 5 6 12 13 14
(1) (2) (3) (4) (5) (6) (7) (8) (9) (1 O)
x,,M - x,,,,, (mm) 160 120 50 60 140 200 65 50 85
h,,M - h,,,, (mm) 20.2 11.7 7.5 11.1 16.9 2.5 11.4 9.4 12.4
FI ( - - ) 1.44 1.42 1.21 1.285 1.425 1.47 1.32 1.26 1.375
1 + 11 ( - - ) 1.85 1.73 1.52 1.73 1.84 1.09 1.78 1.68 1.83
I4o (ram) 50.9 34.2 27.8 30.3 42.1 60.7 29.5 27.0 32.3
HI (turn) 48.3 '32.3 25.1 27.8 40.1 58.7 27.1 24.6 30.4
AHJHo (%) 5.1 5.5 9.7 8.4 4.7 3.3 8.0 9.0 5.9
Hz (mm) -- 31.0 25.0 27.0 38.7 -- 26.2 24.1 29.0
AHzlH~ (%) -- 4.0 0.40 2.9 3.5 -- 3.3 20.3 4.6

I g
FIG. 7. First Wave Profiley,(X,) for Ho (mm) = (&) 27, (A) 29.5; (*) 30.3, (0) 34.2;
(a) 42.1; and (n) 50.9 ( ) Average Symmetric Profile, ( . . . . ) Transitions to
Second Wave

The wave height and wave length were correlated to the wave approach
F r o u d e n u m b e r F1 = Q / ( g b 2 h 3 , , , ) v2. Fig. 8(a) s h o w s t h a t t h e d a t a f o l l o w
t h e a m p l i t u d e o f a solitary w a v e

I + ,q = F~ ................................................ (7)
up tO F1 = 1.41, w h e r e "q = ( h , M - h , , ~ ) / h , m = t h e r e l a t i v e w a v e c r e s t
h e i g h t ( T a b l e 2). F o r F1 > 1.41, s o l i t a r y w a v e s a r e k n o w n t o b r e a k , a n d
t h e f u n d a m e n t a l a s s u m p t i o n o f e n e r g y c o n s e r v a t i o n a c r o s s t h e w a v e is n o
m o r e satisfied.

21

J. Irrig. Drain Eng. 1994.120:13-26.


8 i "i

%.

/,
Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

J
/

a)
1
1.2
I

1.4
P I

1.6
3
1.2
-,-~, 1.4
I1=1
1,6
b)
FIG. 8. Scalings of First Wave Profile: (a) Wave Height I + ~; and (b) Wave Length
h, as a Function of FI [Notation Fig. 7 and 1to [ram] = (o) 27.8, (v) 32.3, and (e)
6o.7].

f
0.6m 9 + , , 0.99

0.97

0,4 - - - 0 9 3
- ~ " ~ " ~ ! ~0.77
~:~--'-~0.47

L
4A , 4.'6 . . . 48 . . 5 i
52 ' 5.4m

FIG. 9. Submerged Flow for Variable Tailwater Submergence

The wave length ~, = (x,M - Xum)/hum was also plotted against F1, and
compared to the theoretical result h, = 3(2[F 2 - 1]) 1/2. According to Fig.
8(b), the prediction compares well with observation for F1 < 1.41, as before.
Then, the solitary wave profile was compared to the first wave profile
according to Fig. 7, and reasonable agreement between the two was found.
Thus, the first wave on a broad-crested weir represents a standing solitary
wave relative to the scalings h,,,, F u m . Note the dissipation of energy from
the approach channel (11o) to the approach of solitary w a v e H 1 t o the second
wave minimum (H2) according to Table 2. The relative energy loss A H / H o
with AH1 = Ho - Ha increases as 1to decreases. Regarding the head loss
AHz = Ha - 112 across the first wave, the percentage loss decreases as 14o
or H1 decreases, however. Thus, higher waves have a larger energy loss
than small waves.

SUBMERGED FLOW

An important characteristic of discharge measurement structures is the


limit between submerged and unsubmerged (free) flow, called the modular
limit. Fig. 9 shows a number of profiles for a discharge Q = 17.301s- 1 and
various tailwater submergences h t , measured from the weir crest (Fig. 1),
or Yt = ht/Ho as the submergence index. Note that for a distinct discharge,
the scaling Ho varies for submerged flow.
For Yt = 0.47 the surface drop at the end of weir is much reduced, and
small tailwater waves occur. For Yt = 0.77 a direct hydraulic jump occurs

22
J. Irrig. Drain Eng. 1994.120:13-26.
on the weir crest with the toe at x = 505 cm. The approach Froude number
is F1 -- 1.25, such that the ratio of sequent depths is 1.34 as computed from
the conventional formula as compared to 1.33 from observation, when ig-
noring the standing wave. In the experiment, this jump was not undular,
Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

although F1 = 1.25. For y~ = 0.85, the hydraulic jump becomes undular


with a front bore and the toe located at x = 495 cm. Further increasing to
y, = 0.915 yields an undular jump without front bore. F o r y , = 0.93 a highly
unstable wave train with small capillary waves develops. Long smooth and
more stable waves occur for y, = 0.96 and total submergence with only
slight surface wave action may be noted for y, = 0.97. The flow is practically
at rest when y, = 0.99.
The modular limit (index L) was analyzed by varying the relative head
~w = Ho/Lw, as shown in Fig. 10(b). The data increase to some maximum
Y,L = 0.79 for ~w ~ 0.3. In the domain 0.1 < ~w < 0.4 of typical broad-
crested weir flow, the modular limit is always above 70%, typically some
75%. This is high as compared to other structures such as round-crested
weirs or sharp-crested weirs. For the latter, the tailwater must even be at
least 100 mm below the weir crest.
The flow features under submerged flow involve the submergence factor

Qs
r = -- . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (8)
Q
with Qs = discharge under submerged flow, and Q = the corresponding
flee flow discharge. An index for the tailwater flow depth h, is • = (hi -
hL)/(Ho - hL) with hL as limit flow depth. For modular limit flow h, = hL,
X = 0 and • -~ 1 occurs for deeply submerged flow. The function ~(X) is
shown for two heads Ho under submerged flow in Fig. 10(a). Both data
groups follow the same line, and the curve +(• was shown to read for all
heads 0.1 < ~w < 0.4 as
+ = (1 - • .............................................. (9)
In all experiments a hydraulic jump occurs for modular limit flow on the
weir apron (Fig. 9). Its distance from the front corner is roughly 2.5Ho, that
is downstream of the end of bottom separation (Fig. 3). As long as tailwater
waves occur, the limit submergence is not yet reached. The incipient hy-
draulic jump starts due to wedges at the channel sides. Once these shock
waves meet at the center of channel above the end of weir, one large standing
wave develops and eventually breaks when slightly increasing the tailwater

YtL

Z ~W
/
, i = i 0.5
a) 0 0,2 0.4 0.6 08 1 b)O 0.5

FIG. 10. Submerged Flow: (a) Submergence Characteristic for Discharge +(• 14o
[mini = (e) 84, (o) 142; (b) Modular Limit Y,L(~w)

23
J. Irrig. Drain Eng. 1994.120:13-26.
level. Moving the breaking front (toe of jump) to the position ~ 2.5Ho
influences the bottom separation with an adverse pressure gradient, and
subsequently the approach head 11o.
This mechanism may also be described with the choking flow phenomena.
Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

A hysteresis effect was found in the modular limit, depending whether the
tailwater is raised or lowered. The present data refer to raising tailwater.
For initially submerged flows getting nonsubmerged, the modular limit was
0.02 to 0.05 lower in y, than the data in Fig. 10(b).

DISCHARGE MEASUREMENT STRUCTURE

Based on a literature review (Hager 1993) and the present observations,


the broad-crested weir may be used as an accurate discharge measurement
structure, provided it has standard shape, i.e.:

~ A straight and prismatic rectangular channel of minimum width b


>- 0.30 m.
9 A weir body with vertical up- and downstream faces of minimum
height w = 0.15 m.
9 A sharp-crested upstream corner and a smooth horizontal weir crest.

From the hydraulic point of view, the weir should satisfy the conditions:

9 Approach overflow head 1to such that the weir is with "a broad
crest" (0.1 < Ho/w < 0.4).
9 The overflow head Ho such that the effect of approach velocity is
insignificant (Ho < w/2).
9 The overflow head such that no scale effects occur, typically 14o >--
40 mm to 50 mm.
9 The tailwater submergence y, < 0.75 in order to be in the free flow
domain.

No discharge measurement is recommended for submerged flow due to


tailwater waves and reduction in accuracy. The approach head Ho should
be observed some three maximum heads Ho~t upstream from the upstream
weir face, as also recommended for sharp-crested weirs. The maximum head
HoM on the weir ranges typically from 0.5 m to 2 m, corresponding to a
capacity between 0.5 m 3 s - 1 and 5 m3s - 1 per meter of weir width.

CONCLUSIONS

The broad-crested weir is recommended as a discharge measurement


structure, provided it satisfies some standard conditions. Based on a liter-
ature review and a detailed experimental investigation, the weir flow over
the "broad-crested" weir is found to be self-similar. The main features
including surface profile, separation profile, bottom pressure profile, and
velocity profile close to the upper crest corner are analyzed. Also the undular
weir type flow is considered but excluded for discharge measurement. In-
terestingly, the submergence range is large and the broad-crested weir is
suited for high tailwater submergence.
24

J. Irrig. Drain Eng. 1994.120:13-26.


APPENDIX I. REFERENCES
Crabbe, A. D. (1974). "Some hydraulic features of the square-edged broad-crested
weir." Water and Water Engrg., 78(10), 354-358.
Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

Govinda Rao, N. S., and Muralidhar, D. (1963). "Discharge characteristics of weirs


of finite-crest width." Houille Blanche, 18(5), 537- 545.
Hager, W. H. (1991). "Experiments on standard spillway." Proc. Inst. Cir. Engrs.,
90, 399- 416.
Hager, W. H. (1993). "Breitkroniger lJberfall." Wasser, Energie, Luft (in German).
Harrison, A. J. M. (1964). "Some comments on the square-edged broad-crested
weir." Water and Water Engrg., 68(11), 445-448.
Harrison, A. J. M. (1967). "The streamlined broad-crested weir." Proc. Inst. of Cir.
Eng. (London) Part 1, 38, 657-678; 40, 575- 599.
Horton, R. E. (1907). "Weir experiments, coefficients, and formulas." Dept. of the
Interior, U.S. Geological Survey, Water-Supply and Irrigation Paper 200. Gov-
ernment Printing Office, Washington, D.C.
Keutner, C. (1934). "Strrmungsvorg~inge an breitkronigen Wehrkrpern und an Ein-
laufbauwerken." Bauingenieur, 15(37/38), 366-371; 15(39/40), 389-392 (in Ger-
man).
Moss, W. D. (1972). "Flow separation at the upstream edge of a square-edged broad-
crested weir." J. Fluid Mech., 52(2), 307- 320.
Ramamurthy, A. S., Tim, U. S., and Rao, M. V. J. (1988). "Characteristics of
square-edged round-nosed broad-crested weirs." J. Irrig. and Drain. Engrg., ASCE,
114(1), 61-73; 115(4), 766.
Ranga Raju, K. G., and Ahmad, I. (1973). "Discharge characteristics of suppressed
and contacted broad-crested weirs." J. Irrigation and Power (India), 30(4), 157-
166.
Singer, J. (1964). "Square-edged broad-crested weir as a flow measurement device."
Water and Water Engrg., 68(6), 229-235.
Sreetharan, M. (1988). "Discharge characteristics of rectangular profiled weirs."
Proc. Nat. ASCE Conf. Hydr. Engrg., S. R. Abt. and J. Gessler, eds., ASCE,
969-978, New York, N.Y.
Tracy, H. J. (1957). "Discharge characteristics of broad-crested weirs." U.S. Dept.
of Interior, Geological Survey Circular 397, Washington, D.C.

APPENDIX II. NOTATION

The following symbols are used in this, paper:

b channel width;
G= discharge coefficient;
Cdo = basic discharge coefficient;
cd~= discharge coefficient for sharp-crested weir;
Ca correction coefficient;
Fo = approach Froude number;
9 = gravitational acceleration;
HL= limit energy head;
14o= approach overflow energy head;
h= flow depth;
ho = approach overflow head;
ks = separation height;
ht = height of tailwater submergence;
L w = weir length;
n w number of waves;
p = pressure;
Q= discharge;
25
J. Irrig. Drain Eng. 1994.120:13-26.
No = approach Reynolds number;
R = crest r a d i u s o f c u r v a t u r e ;
U = relative horizontal velocity;
Downloaded from ascelibrary.org by University of Aberdeen, Bedford Road on 07/13/13. Copyright ASCE. For personal use only; all rights reserved.

u = horizontal velocity component;


Vo = approach velocity;
w = weir height;
X = relative length coordinate;
x = longitudinal coordinate;
y = r e l a t i v e flow d e p t h ;
yp = pressure head;
y, = submergence ratio;
Z = zlHo;
z = vertical coordinate;
~w = relative weir length;
~1 = relative wave length;
K = tailwater index;
v = k i n e m a t i c viscosity;
P = density;
+ = constant; and
qJ = discharge reduction.

Subscripts
a = average;
L = limit;
M = maximum;
m = minimum;
o = approach;
s = separation;
u = undular; and
1 = first w a v e a p p r o a c h .

26

J. Irrig. Drain Eng. 1994.120:13-26.

You might also like