You are on page 1of 55

Source: HYDRAULIC DESIGN HANDBOOK

CHAPTER 18
HYDRAULIC DESIGN OF
STILLING BASINS AND
ENERGY DISSIPATORS

C. Y. Wei and James E. Lindell


Harza Engineering Company
Chicago, Illinois.

18.1 INTRODUCTION
Stilling basins and energy dissipators are usually provided in conjunction with develop-
ment of spillways, outlet works, and canal structures. It is often necessary to perform
hydraulic model studies of individual structures to be certain that these energy dissipating
devices will operate as anticipated. A relatively large volume of data is available from
many laboratory and field studies performed in the past (Blaisdell, 1948; Bowers and
Toso, 1988; Bowers and Tsai, 1969; Chadwick and Morfett, 1986; Chaudhry, 1993; Chow
1959; French, 1985; George, 1978; Hendreson, 1966; International Commission on Large
Dams (ICOLD), 1987; Novak et al., 1990; Peterka, 1964; Robert son et al., 1988; Senturk,
1994; Toso and Boweis, 1988; U.S. Bureau of Reclamation (USBR), 1974, 1987; Vischer
and Hager, 1995, 1998). Based on the results of many intensive studies, in 1958, A. J.
Peterka (1964) of the U.S. Bureau of Reclamation (USBR) published a summary report
of USBR’S studies entitled Hydraulic Design of Stilling Basins and Energy Dissipators,
Engineering Monograph No. 25. Since then, this publication has been referenced widely
within the hydraulic engineering community and still is one of the best references on this
subject available today. Energy dissipators are used to dissipate excess kinetic energy pos-
sessed by flowing water. An effective energy dissipator must be able to retard the flow of
fast moving water without damage to the structure or to the channel below the structure.
Vischer (1995) classified various types of energy dissipators by their features as: (1) by
sudden expansions, (2) by abrupt deflections, (3) by counterflows, (4) by rough walls, (5)
by vortex devices, and (6) by spray inducing devices. The stilling basins and energy dis-
sipators discussed in this chapter are related to energy dissipation by expansion and
deflection.
There are two basic types of energy dissipators. They are hydraulic jump-type dissi-
pators and impact-type dissipators. The hydraulic jump type energy dissipators dissipate
excess energy through formation of highly turbulent rollers within the jump. The impact-
type dissipators direct the water into an obstruction that diverts the flow in all directions
and generates high levels of turbulence and in this manner dissipates the energy in the
flow. In other cases, the flow is directed to plunge into a pool of water where the energy

18.1
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.2 Chapter Eighteen

(a)

(b)

Exhibit 18.1: Wanapum project, Washington


(a) General view of the spillway in operation.
(b) Layout of the spillway and stilling basin.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.3

is diffused and dissipated. The impact-type energy dissipators include check drops and
vertical drops, baffled outlets, baffled aprons, and vertical stilling wells. Generally, the use
of an impact-type energy dissipator results in smaller and more economical structures.

18.2 STILLING BASINS


Using six test flumes, USBR conducted model studies for five stilling basin designs. The
results are summarized and presented in the Engineering Monograph No. 25 mentioned
above (Peterka, 1964). In Basin I tests, all test flumes were used and the test data obtained
provides basic hydraulic information concerning hydraulic jumps on a horizontal apron.
The Type II basin was developed for high dam and earth dam spillways and large canal
structures where the approach velocity is high and the corresponding Froude number
exceeds 4.5. Type III stilling basin is suitable for general canal structures, small outlet
works, and small spillways where the approach velocity is moderate or low and does not
exceed 50–60 ft/s (15–18 m/s) and the unit discharge is less than 200 ft3/s/ft (18 m3/s). For
smaller canal structures, outlet works, and diversion dams where the approach Froude
number is relatively low (between 2.5 and 4.5) and the heads of the structures do not
exceed 50 ft (15 m), Type IV stilling basin may be used. However, the jumps in the basin
may be unstable and alternative design such as the modified Type IV basin may be con-
sidered. To achieve greater structure economy for high dam spillways, Type V stilling
basin with sloping apron may be considered. Photos of several stilling basin in operation
are given in Exhibits 17.2, 17.4, 18.1, and 18.2.

18.2.1 General Hydraulic Jump Basin (Basin I)

The basic elements and characteristics of a hydraulic jump on horizontal aprons (Fig.
18.1) is provided to aid designers in selecting more practical basins such as Basins II, III,
IV, V, and VI. Jump occurs on a flat floor with no chute blocks, baffled piers or end sill in
the basin. Usually, it is not a practical basin because of its excessive length. For a
high–velocity flow down a spillway chute with known terminal velocity (Fig. 18.1) and
depth entering the basin, the required tail water depth, the length of jump, and loss of ener-
gy can be determined based on the curves provided in Fig. 18.2a–e.

18.2.2 Stilling Basins for High Dam and Earth Dam Spillways
and Large Canal Structures (Basin II)

This stilling basin was developed for use on high spillways, large canal structures, and
so forth for approach Froude numbers above 4.5. With chute blocks and dentated end sill,
the jump and basin length can be reduced by about 33 percent. The basic design features
of Basin II stilling basin are given in Fig. 18.3a. For preliminary designs, the curves for
estimating required tailwater depth, and length of jump are given in Figures 18.3b and c.
The water surface and pressure profiles can be determined based on Fig. 18.3d and e.
The water surface profile in this basin can be closely approximated by a straight line
making an angle a (jump angle) with the horizontal. This line can also be considered as
a pressure profile. The USBR guidelines for designing this type of stilling basins are
given as follows:

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.4 Chapter Eighteen

(a)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.5

(b)

(c)

Exhibit 18.2 Mayfield hydroelectric project, Washinton


(a) Layout of the spillway including flip bucke.
(b) General view of the spillway and the stilling pool
(c) General view of the spillway and stilling pool.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.6 Chapter Eighteen

FIGURE 18.1 Curves for velocity entering stilling basins from 0.8:1 to 0.6:1 steep slopes.(From
Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.7

FIGURE 18.2 Basic hydraulic jump basins on horizontal aprons. (Basin I) (From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.8 Chapter Eighteen

FIGURE 18.3 Stilling basins for high dam and earth dam spillways and large canal structures. (Basin
II)(From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.9

1. Determine velocity V1 of flow entering the jump. Figure 18.1 may be used. This
chart represents a composite of experience, computation, and a limited amount of
experimental information obtained from prototype tests on Shasta and Grand Coulee
Dams. The chart provides a fair degree of accuracy for chute having slopes of 0.8:1
or steeper, where computation is a difficult and arduous procedure. The asymptot-
ic nature of the terminal velocity curves is also depicted in Fig. 18.1. For a constant
head of 2.5 ft (0.8m) on the spillway crest, the terminal velocity does not increase
significantly (from 51 ft/s or 15.5 m/s to 53 ft/s or 16.2 m/s) as the vertical distance
(fall) from the reservoir level to stilling basin floor increases from 200 to 600 ft
(61–183 m).
2. Set apron elevation to utilize full conjugate tail water depth. Add a factor of safety
if needed. A minimum margin of safety of 5 percent of tailwater depth (D2) is rec-
ommended.
3. Exercise caution with effectiveness of the basin at lower values of the Froude number
V1 / (gD1)1/2) of 4 or lower. D1 is the depth of the flow entering the basin.
(V
4. Determine the length of basin using the curve shown in Fig. 18.3c.
5. Use the depth of flow entering the basin, D1 as the height of chute blocks. The
width and spacing should be equal to approximately D1 but can be varied to avoid
fractional blocks. A space equal to D1/2 is preferable along each side of wall to
reduce spray and maintain desirable pressures.
6. As shown in Fig. 18.3a, set the height of the dentated sill equal to 0.2D2 and the
maximum spacing approximately 0.15D2. For narrow basins, the width and spac-
ing may be reduced but they should remain equal.
7. It is not necessary to stagger the chute blocks with respect to the sill dentates.
8. It is recommended that the sharp intersection between chute and basin apron be
replace with a curve of reasonable radius of at least 4D1 when the chute slope is 1:1
or greater. Chute blocks can be incorporated on the curve surface as readily as on
the plane surfaces. The chute slope (0.6:1–2:1) does not have significant effect on
the stilling basin action unless it is nearly horizontal.

Following the above rules should result in a safe, conservative stilling basin for
spillways up to 200 ft (60 ms) high and for flows up to about 500 (ft3/s/ft) [46.5
(m3/s/m)] basin width, provided that jet entering the basin is reasonably uniform
both as to velocity and depth. For greater falls, larger unit discharges, or possible
asymmetry, a model study of the specific design is recommended.

18.2.3 Short Stilling Basins for Canal Structures, Small Outlet Works,
and Small Spillways [Basin III and the St. Anthony (SAF) Basin]

For structures carrying relatively small discharges at moderate velocities, a shorter basin
having a simpler end sill may be used if baffled piers are placed downstream from the
chute blocks (Fig. 18.4). In this section, stilling basins for smaller structures in which
velocity at the entrance to the basin are moderate or low ( up to 50–60 ft/s or 15–18 m/s)
and discharges of up to 200 ft3/s/ft of width or 18 m3/s/m of width are discussed. The still-
ing basin action is very stable for this design. It has a large factor of safety against sweep-
out of the jump and operates equally well for all values of the Froude number above 4.0.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.10 Chapter Eighteen

FIGURE 18.4 Short stilling basins for canal structures, small outlet works and small spillways.(Basin
III) (From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.11

This basin should not be used for velocities above 50 ft/s or 15 m/s to avoid potential cav-
itation damages aginst baffle piers. Instead, Basin II type stilling basin should be consid-
ered or hydraulic model studies should be performed. The following USBR guidelines
pertain to the design of the Basin III type stilling basin:

1. The stilling basin operates best at full conjugate tail water depth, D2. A reasonable
factor of safety is inherent in the conjugate depth for all values of the Froude num-
ber and it is recommended that this margin of safety not be reduced.
2. Determine the length of basin using the design curve given in Fig. 18.4c. It is less
than one-half the length of the natural jump. It should be noted that an excess of tail
water depth does not substitute for basin length or vice versa.
3. Exercise caution with effectiveness of the basin at lower values of the approach
Froude number [VV1 / (gD1)1/2] of 4.5 or lower.
4. Height, width, and spacing of chute blocks should equal the average depth of flow
entering the basin, or D1. Width of blocks may be decreased, provide spacing is
reduced a like amount. Should D1 proved to be less than 8 in or 20 cm, the blocks
should be made 8 in or 20 cm high.
5. The height of the baffle piers (Fig. 18.4a) varies with the Froude number and is
given in Fig. 18.4d. In narrow structures, block width and spacing may be reduced,
provided both are reduced a like amount. A half space is recommended adjacent to
the walls.
6. The upstream face of the baffle piers should be set at a distance of 0.8D2 from the
downstream face of the chute blocks. This dimension is important.
7. The height of the solid end sill is given in Fig. 18.4d. The slope is 2:1 upward in
the direction of flow.
8. It is undesirable to round or streamline the edges of the chute blocks, end sill, or
baffle piers. It reduces the effectiveness of the energy dissipation. However, small
chamfers on the block edges to prevent chipping of the edges and to reduce cavita-
tion erosion may be used.
9. It is recommended that a radius of reasonable length greater than 4D1 be used at the
intersection of the chute and basin apron for slopes of 45º or greater.
10. As a general rule, the slope of the chute has little effect on the stilling basin action
unless long flat slopes are involved.
11. Experience indicates that the Type III basin works well for flow less than 200
ft3/s/ft or 18 m3/s/m based on basin width and approach velocity at the entrance of
up to 50–60 ft/s or 15–18 m/s.

The St. Anthony Falls (SAF) Hydraulic Laboratory of the University of Minnesota had
also developed a similar basin for small spillways, outlet works, and small canal structures
for approach Froude numbers ranging from 1.7 to 17 (Blaisdell, 1948; Chow, 1959). This
basin was developed to achieve about 70 to 90 percent reduction of the jump lengths.
This basin is commonly known as the SAF stilling basin (Fig. 18.5). Since the basin is
relatively short so that a significant amount of residual energy can still exist downstream
from the end sill, the channel reach downstream from the stilling basin should be allowed
to erode until a stable scour depth is reached. Otherwise riprap protection should be pro-
vided to minimize scour (Sec. 18.7). The guidelines for designing this basin are summa-
rized as follows (Blaisdell, 1948; Chow, 1959):

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.12 Chapter Eighteen

FIGURE 18.5 The SAF stilling basin. (From Blaisdell, 1948)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.13

1. The length L of the stilling basin is determined by the following equation.


4.5y2
L  0.
F1 76
where y2 is the theoretical sequent depth of the jump corresponding to the approach
flow depth y1 and F1 is the approach Froude number.
2. The height of the chute blocks and floor blocks is y1, and the width and spacing are
approximately 0.75y1.
3. The distance from the upstream end of the stilling basin to the floor blocks is L/3.
L
4. No floor blocks should be placed closer to the side-wall than 3y1/8.
5. The floor blocks should be placed downstream from openings between the chute
blocks.
6. The total width of the floor blocks should occupies about 40 to 55 percent of the
stilling basin width.
7. The widths and spacings of the floor blocks for diverging stilling basins should
be increased in proportion to the increase in stilling basin width at the floor block
location.
8. The height of end sill is given by c  0.07y2.
9. The depth of tailwater above the stilling basin floor is given by
 F12 
y'2  1.10   y for F1  1.7 to 5.5
 1 20  2

y'2  0.85y2 for F1  5.5 to 11.0

 F12 
y'2  1.00   y for F1  11 to 17
 8 00  2
10. The top of the side-wall above the maximum tailwater level to be expected during
the life of the structure is given by z  y2/3.
11. Wing-walls should be equal in height to the stilling basin side-walls. The top of the
wing-wall should have a slope of 1H:1V.V
12. The wing-wall should be placed at an angle of 45º to the outlet center line.
13. The stilling basin side-walls may be parallel for a rectangular stilling basin or they
may diverge as an extension of the transition side-walls for a trapezoidal stilling
basin as shown in Fig. 18.5.
14. A cutoff wall of nominal depth should be used at the end of the stilling basin.
15. The effect of entrained air should be neglected in the design of the stilling basin.

18.2.4 Low Froude Number Stilling Basins


(Basin IV and Modified Basin IV)

This stilling basin was developed for canal structures, outlet works, and diversion dams
where the approach Froude number of the basin is relatively low (between 2.5 and 4.5)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.14 Chapter Eighteen

FIGURE 18.6 Low Froude number (2.5–4.5) stilling basin design (Basin
IV).(From Peterka, 1964)

and the heads of the structures are about 50 ft (or 15 m). In this case, the jump is not fully
developed and unstable and the methods of design discussed previously do not apply.
Alternative design and/or wave suppressors or Basin VI type stilling basin with a hang-
ing baffle for energy dissipation may be considered. Guidelines for developing
a low Froude number stilling basin (Basin IV) as depicted in Fig. 18.6 are given as
follows (Peterka, 1964):

1. A model study of the stilling basin is imperative.


2. Reduction of excessive waves created in the unstable jump is the main problem
concerning the design of the stilling basin.
3. A tailwater depth of 10 percent greater than the conjugate depth is strongly recom-
mended.
4. Place as few appurtenances as possible in the path of the flow, as volume occupied
by appurtenances helps to create a backwater problem, thus requiring higher
training walls.
5. Use Fig. 18.6 to develop the design of the stilling basin. The number of deflector
blocks shown in the figure is a minimum requirement.
6. The length of basin can be obtained from Fig. 18.2c. No baffle piers are needed in
the basin.
7. The recommended maximum width of the deflector blocks is equal to D1 but
0.75D1 is preferable from a hydraulic standpoint. The ratio of block width to spac-
ing should be maintained as 1:2.5.
8. The extreme tops of the deflector blocks are 2D1 above the floor of the stilling
basin.
9. To accommodate the various slopes of chutes and ogee shapes encountered, the
horizontal top length of the deflector blocks should be at least 2D1. The upper sur-
face of each block is sloped at 5º in a downstream direction for better operation
especially at lower discharges.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.15

10. The addition of a small triangular sill placed at the end of the apron for scour con-
trol is desirable. An end sill of the type developed for short stilling basins (Basin
III) can be used. The slope of the upstream face of the sill is 2:1 and the height of
the sill can be determined based on Fig. 18.4d.
11. Basin IV stilling basin is applicable to rectangular cross sections only to minimize
potential wave–related problems.
Type IV stilling basin performs effectively in dissipating the energy at low Froude
number flows for small canals and for structures with small unit discharges. It is also
effective in minimizing wave problems. Based on additional model tests, the U.S. Bureau
of Reclamation (USBR) has developed a modified stilling basin for low Froude number
approach flows (George, 1978). This stilling basin is suitable for approach flows with
Froude numbers ranging from 2.5 to 5.0. The basin is relatively short and is provided with
chute blocks, baffle piers, and a dentated end sill as shown in Fig. 18.7a. The guidelines
for designing Modified Type IV stilling basin are given as follows (George, 1978):

FIGURE 18.7(a) Low Froude number stilling basin (Modified Basin IV).(From George, 1978)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.16 Chapter Eighteen

1. A hydraulic model study is recommended to confirm the design. Erosion tests


should be included. Such tests should be made over a full range of discharges to
determine erosion potential downstream from the basin and to determine the poten-
tial for the abrasive bed materials to move upstream into the basin.
2. Determine the theoretical D2 based on the known unit discharge and the approach
flow depth D1.

FIGURE 18.7(b) Design curves for modified Basin IV stilling basin.(From


Georges, 1978)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.17

3. Determine tailwater depth as TW  1.05 D2.


4. Set the length of the basin L  3D2 (approximately).
5. Use Fig. 18.7a to develop the basic dimensions of the basin.
6. Determine the distance X from the chute blocks to the baffle piers. X varies from 1.3
to 0.7 times D2 as the approach Froude number varies from 2.5 to 5.6 as shown in
Fig. 18.7b.
7. Determine the distance L1 from the toe of the chute to the upstream face of the end
sill from Fig. 18.7b.
8. If (L1  the length of the end sill) is longer than L then the stilling basin should be
extended to include the end sill.
9. Set the widths of the baffle piers equal to 0.7D1 and heights equal to 1.0D1.
10. Determine the number of chute blocks and baffle piers by the following equations.

The total number of chute blocks and spaces N = (width – 2kW)/


W W

where
k  fractional width of block equal to side clearance, 0.375  k  0.50
width  total width of stilling basin
W  0.70D1

The N value obtained should be rounded to the nearest odd number and then adjust
values of W and k should be adjusted.

11. Use 0.2D1 as the top length of the baffle piers.


12. Determine end sill dimensions.

height  0.2D2
width, W  0.15D2

top length of end sills  0.2  height

The number of blocks and spaces N  (basin width)/W

(N should be rounded to the nearest odd number and then the value of W
should be adjusted)

18.2.5 Stilling Basin with Sloping Apron

To achieve greater structural economy, a stilling basin with a sloping apron can be con-
sidered. This type of stilling basin is usually used on high dam spillways. It needs greater
tail water depth than horizontal apron. The energy dissipation is as effective as occurs in
the true hydraulic jump on a horizontal apron. The primary concern in sloping apron
design is the tail water depth which is required to move the front of the jump up the slope
to the location where the jump is expected to start. It may not be economically feasible to
design the basin to confine the entire jump, especially when sloping aprons are used in

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.18 Chapter Eighteen

FIGURE 18.8 Stilling basins with sloping aprons. (From Peterka, 1964)

conjunction with medium or high overfall spillways where the rock foundation is in fair-
ly good condition. When shorter aprons are used, the riverbed downstream must act as part
of the stilling basin. On the other hand, when the quality of foundation material is ques-
tionable, it is desirable to make the apron sufficiently long to confine the entire jump. The
total apron length may range from about 40 to 80 percent of the length of jump.
The hydraulic jump may occur in several ways on a sloping apron, as depicted in
Fig. 18.8. The jump may have its toe form on the slope and the jump itself ends over the
horizontal apron (Case B), or ends at the junction of the slope and the horizontal apron
(Case C), or the entire jumps forms on the slope (Case D). For practical purposes the action
in Cases C and D is the same. Guidelines for the design of sloping aprons are given below:

1. Determine an apron arrangement that will give the best economy for the maximum
discharge condition.
2. The first consideration should be to determine the apron slope that will require the
minimum amount of excavation, the minimum amount of concrete, or both, for the
maximum discharge and tailwater condition.
3. Position the slope so that the front of the jump will form at the upstream end of the
slope for the maximum discharge and tailwater condition (Fig. 18.9). It may be nec-
essary to raise or lower the apron, or change the slope entirely. Data obtained from
13 existing spillways are also shown in Fig. 18.9. Each point in the figure has been
connected with an arrow to the tan(θ) curve corresponding to the apron slope. The
adequacy of the tailwater depth of these spillways can then be evaluated.
4. Use Fig. 18.10 to determine the length of the jump for maximum or other
flows. Shorter basins may be used where a solid bed exists. For most installations,
an apron length of about 60 percent of the length of jump for the maximum dis-
charge condition should be sufficient. Longer basins are needed only when the
downstream riverbed is in very poor condition.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.19

5. Ascertain that the tailwater and length of basin available for energy dissipation
are sufficient for, say 1/4, 1/2, and 3/4 capacity. If the tailwater depth is deficient,
a different slope or a new position of the sloping portion of the apron should be
considered.
6. Horizontal and sloping aprons will perform equally well for high values of the
Froude number if the proper tail water depth is provided.
7. The slope of the chute upstream from a stilling basin has no significant effect on
the hydraulic jump when the velocity distribution and depth of flow are reasonably
uniform on entering the jump.
8. A small solid triangular sill should be provided at the end of the apron to lift the
flow as it leaves the apron for scour protection. The most effective height is
between 0.05D2 and 0.10D2 and a slope of 3:1–2:1. Several existing stilling basins
with sloping aprons are shown in Fig. 18.11. All stilling basins shown were
designed with the aid of model studies.
9. The stilling basin should be designed to operate with as nearly symmetrical flow in
the stilling basin as possible to avoid formation of large circulating eddies and
transport of riverbed material into the apron area, and the potential undermining of
the wing walls and riprap.
10. A model study is advisable where the discharge over high spillways exceeds 500
ft3/s/ft or 46.5 m3/s/m based on the apron width, where there is any form of asym-
metry involved, and for the high values of the Froude number where stilling basins
become more costly and the performance becomes less acceptable.

FIGURE 18.9 Comparison of existing sloping apron designs with experimental results.
(From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.20 Chapter Eighteen

FIGURE 18.10 Jump length in terms of conjugate depth, D2 for stilling basins with sloping aprons.(From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.21

FIGURE 18.11(a) Existing stilling basins with sloping aprons.(From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.22 Chapter Eighteen

FIGURE 18.11(b) Existing stilling basins with sloping aprons.(From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.23

18.2.6 Other Types of Stilling Basins

Other types of stilling basins that may be considered include: (1) positive step basin, (2)
negative step basin, (3) baffle–sill basin, (4) baffle-block basin, (5) expanding stilling
basin, and (6) bucket stilling basin. Detailed discussions of these basins have been pro-
vided by Vischer and Hager (1995). These basins are briefly discussed as follows:

1. Positive–step basin. An upward step of a given height is provided in a prismatic chan-


nel. No end sills are included. The required basin length is significantly longer than
that of a classical jump basin.
2 Negative–step basin. A downward step is provided. No end sills are included. It
requires a slightly longer basin length than the positive step basin. Basins with steps
have not been popular because it is easier to use sills or blocks in a horizontal apron
than to change the apron elevation at the step section.
3 Baffle–sill basin. A weir-type sill is provided to form a basin. The flow over the sill
may be submerged or free. The sill is capable of stabilizing the jump in a shorter
basin and with lower tailwater than is the classical jump basin. Sills can be econom-
ical and effective devices for energy dissipation even without additional appurte-
nance included.
4. Baffle–block basin. Baffle blocks are normally arranged in one or several rows that are
oriented perpendicular to the direction of approach flow. Standard baffle blocks such
as the USBR blocks should be used. Baffle blocks are prone to cavitation damage and
should not be used for approach velocities above 20 m/s. For velocities between 20 and
30 m/s, a chamfer on the block edges should be provided to reduce the cavitation
potential.
5. Expanding stilling basin. There are two types of expanding basins, namely gradually
expanding basin and abruptly expanding basin. The gradually expanding basin
requires less tailwater depth and can be used for highly variable tailwater. This type of
basin is suitable for approach flow with Froude numbers less than 4. Very few basins
of this type have been built. An abruptly expanding basin has been studied and report-
ed by Vischer and Hager (Novak et al., 1990). No practical applications have been
reported.

18.2.7 Fluctuating Pressures on Stilling Basin Floors

When designing a stilling basin to achieve highest possible hydraulic efficiency in terms of
energy dissipation, one should also consider the structural aspects of the stilling basin. The
effect of transient pressures caused by turbulence in the jump can be significant and should
be considered in the design of the structure. Extensive discussions of this subject have been
provided by International Commission on Large Dams (ICOLD, 1987), Toso and Bowers
(1988), and Visher and Hager (1995), and so on. The hydromechanic characteristics and the
turbulence level of the jump in a stilling basin depends not only on the relative tailwater
level but also on the geometry and the concrete finish conditions of the basin floor and
training walls. The pressure fluctuations resulting from intense macro-scale turbulence in
the jump must be carefully considered during the design of the structure. The pressure fluc-
tuations vary widely in amplitude at all locations within the jump. The maximum half-
amplitude of the fluctuation has been determined to be approximately 40 percent of the
mean approach velocity head with a frequency of about 1 Hz. The dominant pulsating
components have frequencies between 0 and 10 Hz. When the pressure becomes negative

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.24 Chapter Eighteen

at a point on the apron surface, a dangerous local instability may develop with respect to
the uplift pressure at the bottom of the concrete slab. Some projects have experienced high
uplift pressures under large areas of the basin floor and resulted in complete floor concrete
slabs being torn up (Bowers and Toso, 1988; ICOLD, 1987; Toso and Bowers, 1988). In
addition, cavitation, abrasion, and vibration due to intense turbulence and pressure fluctu-
ations can also contribute significantly to the damage of a stilling basin.
Based on the model studies of USBR Type II and Type III stilling basins, Toso and
Bowers (1988) obtained the following useful conclusions:
1. The pressure fluctuations in the jump tend to approach a definite limit, on the order of
80 to 100 percent of the approach velocity head. This is on the order of 10–20 times
the root-mean square (rms) of the pressure fluctuation.
2. Addition of chute blocks, intermediate blocks, and end sills did not result in signifi-
cantly higher maximum negative and positive deviations than those for basins without
blocks and sills. The energy dissipation was quicker.
3. Side-wall pressure fluctuations are very significant, and peak at one to two inflow
depths above the floor.
4. The longitudinal extent of extreme pressure pulsation in the zone of maximum turbulence
is on the order of eight times the inlet flow depth. The lateral extent of a characteristic
pulse is approximately 1.6 times the longitudinal extent or 13 times the inlet flow-depth.
ICOLD (1987) recommended, as a minimum precaution, that the following two con-
ditions be considered when designing the stilling basin apron:
1. Full downstream uplift pressure applied over the entire area of the floor with basin empty.
2. Full uplift pressure equals 12 percent of the approach velocity head applied under the
whole basin, with the basin full.
If necessary, the basin floor can be strengthened by providing anchors or using thick-
er slabs which may be held in place by the side walls.
ICOLD (1987) also recommended following structural arrangements to minimize
potential uplift damages due to undesirable turbulent flow induced pressure fluctuations.
1. All contraction joints should be fitted with properly located and embedded seals.
2. There should be no drain openings in the training wall inside the basin. However, drain
outlets in a dentated sill at the beginning of a stilling basin have performed satisfactorily.
3. Keep the areas of the floor slabs as large as possible.
4. Connect slabs by means of dowels, shear keys, and reinforcement across the joints.
5. Keep horizontal construction joints to a minimum, with dowels across them.
6. If drainage is necessary, keep it well away (1–1.5 m at least) from the wetted surfaces
so that abrasion or cavitation erosion will not make it accessible to the turbulent flow.

18.3 DROP-TYPE ENERGY DISSIPATORS


For small drops in canals with values of the Froude number between 2.5 and 4.5, a
drop–type energy dissipator which is in the form of a grating is particularly applicable to
reduce wave actions and dissipating energy. The device causes the over falling water jet
to separate into a number of long, thin sheets of water which falls nearly vertical into the
canal below. It has excellent capability in dissipating energy and eliminating wave prob-
lems. Guidelines for developing a drop-type energy dissipator are given as follows
(Peterka, 1964):

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.25

1. The device is highly recommended for approach flow with the Froude number below 3.0.
2. The Froude number is computed at the top of the drop.
3. The dissipator consists of a series of steel rails, channel irons, or timber beams in the
form of grating installed at the drop (Fig. 18.12).
4. The spacing beams may vary from 2/3 to the full width of the beams. The narrower
spacing is more effective. Use the following expression to compute the length of
beams:
Q
L  
CSN 兹兹2苶苶gy苶
where Qtotal discharge (ft3/s or m3/s, C  experimental coefficient (dimensionless),
S  width of a space in feet or meters, N  the number of spaces, g  the accelera-
tion of gravity (ft/s2 or m/sec2), and y  the depth of flow in the canal upstream (ft or
m). The value of C is about 0.245.
5. The length of the beams varies from about 2.9 to 3.6 times the depth of the approach
flow.
6. The rails or beams may be tilted downward at an angle of 3º or more to provide some
self-cleaning capability. It may also be made adjustable and tilted upward to act as a
check to maintain a certain level in the canal upstream. However, more frequent clean-
ing of the device may be required.

18.4 WAVE SUPPRESSORS


A wave suppressor is used to provide greater wave reduction to a proposed structure or an
existing waterway. Two types of wave suppressors may be considered. They are raft-type
and underpass-type wave suppressors. Both are applicable to most open-channel water-
ways having rectangular, trapezoidal, or other cross-sectional shapes. Both types may be
used without regard to the Froude number. The underpass-type suppressor provides
greater degrees of wave reduction but may be less economical than the raft-type.

FIGURE 18.12 Drop-type energy dissipator for small drop


canals.(From Pterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.26 Chapter Eighteen

FIGURE 18.13 Raft-type wave suppressor. (From Peterka, 1964)

18.4.1 Raft-Type Wave Suppressors

A number of rafts of different designs were tested by USBR (Peterka, 1964). The most
effective raft arrangement was found to consist of two rigid stationary rafts 20 ft (6.10
m) long by 8 ft (2.45 m) wide, made from 6- by 8-in timbers, placed in the canal down-
stream from the stilling basin as shown in Fig. 18.13. The arrangement is also applicable
for suppressing waves having a regular period such as wind waves or waves produced by
operation of pumps. Guidelines for designing a raft–type wave suppressor are provided
as follows:

1. A space should be left between timbers and lighter crosspieces are placed on the rafts
parallel to the flow. It creates many open spaces resembling rectangular holes.
2. The rafts should be perforated in a regular pattern and there should be some depth to
these holes.
3. The ratio of hole area to total area of the raft may vary from 1:6 to 1:8.
4. The 8 ft (2.5 m) width, W
W, as shown in Fig. 18.13, is a minimum dimension.

5. The raft must have sufficient thickness so that the troughs of the waves do not break
free from the underside.
6. At least two rafts should be used, and the rafts should be rigid and held stationary.
7. The top surfaces of the rafts are set at the mean water surface in a fixed position so that
they cannot move.
8. Spacing between rafts should be at least three times the raft dimension, measured par-
allel to the flow. Each raft can decrease the wave height about 50 percent.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.27

9. For suppressing waves having regular periods, the second raft should be placed down-
stream at some fraction of the wave length to maintain its effectiveness. It may be nec-
essary to make the second raft portable for narrow canals.

18.4.2 Underpass-Type Wave Suppressors

Based on numerous studies conducted, USBR determined that the most effective wave
dissipator to be located downstream from a stilling basin is the short-tube type underpass
wave suppressor (Peterka, 1964). When it becomes necessary to make the raft-type wave
suppressors adjustable or portable, or a moderate increase in depth in the stilling basin can
be tolerated, consideration should be given to the underpass-type wave suppressors. It
may be added to an existing structure or included in the original construction. It can be
used to prevent wave overtopping of the canal lining or bank erosion due to waves. The
structure consists of a horizontal roof placed in the flow channel with a headwall suffi-
ciently high to cause all flow to pass beneath the roof as shown in Fig. 18.14a. Three main
factors should be considered when designing an underpass-type suppressor. They are the
submergence of the roof, the length of the underpass, and the increase in flow depth
upstream of the underpass. The following guidelines may be used to design an underpass-
type suppressor:

1. The height of the roof above the channel floor may be set to reduce wave heights effec-
tively for a considerable range of flows or channel stages.
2. The maximum wave reduction occurs when the roof is set 33 percent of the flow depth
below the water surface for maximum discharge. The submergence and the percent
reduction in wave height becomes less, in general, for smaller-than-maximum dis-
charges.
3. Fig. 18.14c can be used to estimate the wave reduction. The upper curve shown in the
figure was obtained from the study conducted for the short tube underpass wave sup-
pressor of the Carter Lake Dam No.1 Outlet Works. The lower curve shows the model
test results of the Friant-Kern Canal (Fresno, California) underpass type suppressor for
less than maximum discharges with smaller wave heights and shorter periods. The
wave period greatly affects the performance of a given underpass. The suppressor pro-
vides a greater percentage reduction on short period waves. The wave action below a
stilling basin usually has no measurable period and the water surface is choppy and
consists of generated and reflected waves. The waves found downstream from
hydraulic jumps or energy dissipators usually have a period of not more than 5 s. There
is a tendency for the wave period to become less with decreasing discharge.
4. The underpass is most effective when the velocity beneath the underpass is less than
about 10 ft/s or 3 m/s and the channel length downstream from the underpass is three
to four times the length of the underpass.
5. The minimum length of underpass required depends on the amount of wave reduction
considered necessary. For nominal wave reduction to prevent canal lining overtopping
or bank erosion due to waves, a length 1.0–1.5D2 will provide about 60 to 75 percent
wave height reduction. For greater wave reduction, a longer underpass is necessary.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.28 Chapter Eighteen

FIGURE 18.14 Underpass-type wave suppressor (From Peterka, 1964)

For wave periods up to about 5 s, an underpass 2.0–2.5D2 long may provide up to 88


percent wave reduction. Up to about 93 percent of wave height reduction can be
achieved by using an underpass 3.5–4.0D2 long. This length includes a 4:1 sloping roof
extending from the underpass roof elevation to the tail water surface. The sloping por-
tion should not exceed one-quarter of the total underpass length and slopes flatter than
4:1 provide better draft tube action and are more desirable.
6. The greatest wave reduction occurs in the first D2 of underpass length, the construction
of two short underpasses rather than one may be considered. An additional wave reduc-

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.29

tion of 10 percent may be achieved but the extra cost of an additional headwall should
be considered.
7. The backwater effect of the underpass can be determined based on Fig. 18.14b.
8. For design purposes, pressures along the underside of the roof may be considered to
be atmospheric. The average pressures on the headwall and the downstream vertical
wall may be considered as hydrostatic.

18.5 IMPACT-TYPE STILLING BASIN FOR PIPE OR OPEN


CHANNEL OUTLETS
This is an impact-type energy dissipator equipped with a hanging-type ᑦ-shaped baffle,
contained in a relatively small boxlike structure, which requires no tail water for success-
ful performance (Fig. 18.15). The energy dissipation is accomplished by flow striking the
vertical hanging baffle and being turned upstream by the horizontal portion of the baffle
and by the floor, in vertical eddies, and is greater than in a hydraulic jump of the same
Froude number. It may be used to substitute Basin IV-type stilling basin for low Froude

FIGURE 18.15 Basic design of an impact-type stilling basin (From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.30 Chapter Eighteen

FIGURE 18.16 Selection of width for an impact-type stilling basin. (From Peterka, 1964)

FIGURE 18.17 Comparison of energy losses – impact basin and


hydraulic jump.(From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.31

number applications as discussed in Sec. 18.2.4. The impact-type stilling basin generally
provides greater efficiency than that of a jump on horizontal floor (Fig. 18.17). Based on
hydraulic model test results, generalized design rules and procedures have been developed
by USBR (Peterka, 1964) and are given below to allow determination of the proper basin
size and all critical dimensions for a range of discharges up to 339 ft3/s (9.6 m3/s) and
velocities up to about 30 ft/s (9.1 m/s).

1. The use of the impact-type stilling basin discussed in this section is limited to instal-
lation where the velocity at the entrance to the stilling basin does not greatly exceed
30 ft/s (9.1 m/s).
2. The basin operates as well whether a small pipe flows full or a larger pipe flows par-
tially full is used. An open channel having a width less than the basin width will per-
form equally well.
3. Determine the stilling basin dimensions using Figs. 18.15 and 18.16 and Table 18.1,
Columns 3–13 for the maximum expected discharge. For discharges exceeding 339
ft3/s (or 10 m3/s), it may be more economical to consider multiple units side by side.
4. Compute the necessary pipe area from the velocity and discharge. The values in
Table 18.1, Columns 1 and 2, are suggested sizes based on a velocity of 12 ft/s (3.7
m/s) and the desire that the pipe run full at the discharge given in Column 3. The
relationship between discharge and basin size given in the table should be main-
tained regardless of the pipe size chosen. An open–channel entrance may be used in
place of a pipe. The approach channel should be narrower than the basin with invert
elevation the same as the pipe.
5. A moderate depth of tail water will improve the performance although tail water is
not a key factor for successful operation. For best operation, set the basin so that
maximum tail water does not exceed d  g/2.
6. Recommended thickness of various parts of the basin are given in Columns 14-18,
Table 18.1.
7. Determine the minimum size of individual riprap protective stones which will resist
movement when critical velocity occurs over the end sill. Most of the riprap should
consist of the sizes given in Table 18.1, Column 19 or larger. The following empiri-
cal equation, which was developed based on studies performed by Marvis and
Laushey, and Berry as reported by USBR (Peterka, 1964), may also be used to deter-
mine the stone size with reasonable accuracy.

Vb  2.6 兹
兹d

where Vb  bottom velocity (ft/s), and d  diameter of rock (in). The rock is
assumed to have a specific gravity of about 2.65. The accuracy of the equation for
velocities above 16 ft/s (4.9 m/s) is not known.
8. The entrance pipe or channel may be tilted downward about 15º without affecting
performance adversely. For greater slopes use a horizontal or sloping pipe (up to 15º)
two or more diameters long just upstream from the stilling basin. Proper elevation of
the invert at entrance is maintained as shown in Fig. 18.15.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
18.32
TABLE 18.1 Impact–Type Stillling Basin Dimensions.

Suggested Max. Feet and Inches Inches


Pipe Size* dis–
charge
Dia Area Q W H K a b c d e f g tw tf tb tp K Suggested
(In) (ftt3) Riprap Size
(1) (2) (3) (4) (5) (6) (7) (8) (9) (10) (11) (12) (13) (14) (15) (16) (17) (18) (19)

18 1.77 21✝ 5–6 4–3 7–4 3–3 4–1 2–4 0–11 0–6 1–6 2–1 6 61/2 6 6 3 4.0
24 3.14 38 6–9 5–3 9–0 3–11 5–1 2–10 1–2 0–6 2–0 2–6 6 61/2 6 6 3 7.0
30 4.91 59 8–0 6–3 10–8 4–7 6–1 3–4 1–4 0–8 2–6 3–0 6 61/2 7 7 3 8.5
36 7.07 85 9–3 7–3 12–4 5–3 7–1 3–10 1–7 0–8 3–0 3–6 7 71/2 8 8 3 9.0
42 9.62 115 10–6 8–0 14–0 6–0 8–0 4–5 1–9 0–10 3–0 3–11 8 81/2 9 8 4 9.5
48 12.57 151 11–9 9–0 15–8 6–9 8–11 4–11 2–0 0–10 3–0 4–5 9 91/2 10 8 4 10.5
54 15.90 191 13–0 9–9 17–4 7–4 10–0 5–5 2–2 1–0 3–0 4–11 10 101/2 10 8 4 12.0
60 19.63 236 14–3 10–9 19–0 8–0 11–0 5–11 2–5 1–0 3–0 5–4 11 111/2 11 8 6 13.0
72 28.27 339 16–6 12–3 22–0 9–3 12–9 6–11 2–9 1–3` 3–0 6–2 12 121/2 121/2 8 6 14.0

Source: From Peterka (1964).


Abbreviation: a, ;b, ;c, ;d, ;e, ;;ff, ;g, ;k, ;ttb, ;ttf, ;tp, ;ttw,. (see Fig. 18.15)
*Suggested pipe will run full when velocity is 12 ft/sec or half full when velocity is 24 ft/s. Size be modified

Any use is subject to the Terms of Use as given at the website.


for other velocities by Q  AV, V but relation between Q and basin dimensions shown must be manteined.

Copyright © 2004 The McGraw-Hill Companies. All rights reserved.


✝For discharge less than 21 ft/s, obtain width from curve of Fig. 18.14. Other dimensions proportional to W;

H  3W/4,
W L  4W/3, W dW W/6, etc.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Use curve of Fig. 18.21 to determine riprap size.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.33

9. The invert of the entrance pipe, or open channel, should be held at the elevation in
line with the bottom of the baffle and the top of the end sill, regardless of the size of
the pipe selected.
10. If a hydraulic jump is expected to form in the downstream end of the pipe and the
pipe is sealed by the incoming flow, install a vent about one-sixth the pipe diameter
at any convenient location upstream from the jump.
11. For the best possible operation of basin, use an alternative end sill and 45º wall
design as shown in Fig. 18.15. Erosion tendencies will be reduced.

18.6 BAFFLED APRON FOR CANAL OR SPILLWAY


DROPS (BASIN IX)
Baffled aprons or chutes have been used in many irrigation projects for being practical and
economical. The chute is constructed on an excavated slope, 2:1 or flatter, extending to
below the channel bottom. The multiple rows of baffle piers on the chute prevent exces-
sive acceleration of the flow and provide a reasonable terminal velocity. Initial tailwater
is not a prerequisite for the structure to be effective. Backfill is placed over one or more
rows of baffles to restore the original streambed elevation. It prevents excessive accelera-
tion of the flow entering the channel when scour or downstream channel degradation
occur. Through extensive model studies, the hydraulic design of the energy dissipators

FIGURE 18.18 Basic design of a baffled chute (From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.34 Chapter Eighteen

FIGURE 18.19 Recommended baffle pier heights and allowable velocities for baffled chutes (From
Peterka, 1964)

with baffled aprons have been generalized (Peterka, 1964). Basic proportions of a baffled
chute are given in Fig. 18.18 and a simplified design procedure has been developed and is
outlined as follows:

1. The baffled apron should be designed for the maximum expected discharge, Q.
2. The unit discharge q  Q/W may be as high as 60 ft3/s/ft [or 5.6 m3/s/m] based on
chute width, W.
W
3. Approach velocity, V1, should as low as practical. Use recommended approach veloc-
ity (Curve D) shown in Fig. 18.19.
4. The vertical offsets between the approach channel floor and the chute is used to cre-
ate a stilling pool or desirable V1 and will vary in individual installations. See Fig.
18.20 for examples of approach pool arrangements. Use a short-radius curve to pro-
vide a crest on the sloping chute. Place the first row of baffle piers close to the top
of the chute no more than 12 inches or 30 cm in elevation below the crest.
5. Use the recommended height for baffled pier Curve B, Fig. 18.19.
6. Baffle pier widths and spaces should be equal and about 1.5 H but not less than H.
Partial blocks, width 1/3 H to 2/3 H, should be placed against the training walls in
Rows 1, 3, 5, 7, and so forth, alternating with spaces of the same width in Rows 2,
4, 6, and so on.
7. The slope distance (along a 2:1 slope) between rows of baffle piers should be 2H,
twice the baffle height H. When the baffle height is less than 3 ft (or 91.5 cm), the
row spacing may be greater than 2 H but should not exceed 6 ft or 183 cm.
8. The baffle piers may be constructed with their upstream faces normal to the
chute surface.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.35

FIGURE 18.20a Examples of existing baffled chute designs (From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.36 Chapter Eighteen

FIGURE 18.20a Examples of existing baffled chute designs (From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.37

FIGURE 18.20b Examples of existing baffled chute designs (From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.38 Chapter Eighteen

9. Four rows of baffle piers are required to establish full control of flow. The chute
should be extended to below the normal downstream channel elevation and at least
one row of baffles should be buried in the backfill.
10. The chute training wall should be three times as high as the baffle piers to contain
the main flow and splash.
11. Riprap consisting of (6 to 12-in) (15 to 38-cm) stone should be placed at the down-
stream ends of the training walls to prevent eddies from working behind the chute.
The riprap should not extend appreciably into the flow area.

18.7 RIPRAP FOR STILLING BASIN


DOWNSTREAM PROTECTIONS
Riprap stones are placed on the channel bottom and bank downstream of a stilling basin
to prevent bank erosion caused by surges and residual energy from the stilling basin to
reduce the possible undermining of the structure by the erosive currents. Factors affect-
ing design of the riprap include size or weight of the individual stones, the shape of the
large stones, the gradation of the entire mass of riprap, the thickness of the layer, the
type of filter or bedding material placed beneath the riprap, the slope of the riprap layer,
velocity and direction of currents, and eddy action and waves, etc. Based on published
material, laboratory observations and field experience, a design curve (Fig. 18.21) was
developed for the determination of the individual stone size to resist a range of veloci-
ties (Reference 3). Use the estimated bottom velocity or the average velocity at the end
sill of the stilling basin to find the maximum stone size in Fig. 18.21. Specify riprap so
that most of the graded mixture consists of this size. Place the riprap in a layer at least
1.5 times as thick as the maximum stone size. It is recommended that the riprap be
placed over a filter, or bedding, composed of gravel or graded gravel having the larger
particles on the surface.

18.8 SUBMERGED DEFLECTOR BUCKETS


There are occasions that it is desirable to deliver the spillway discharge directly to the
river without additional streambed protection works, the jet may be projected beyond the
structure by a deflector bucket which acts as an energy dissipator at the base of a steep
open chute spillway. USBR had developed both slotted and solid deflector buckets
(Fig. 18.25) for high, medium, and low dam spillways. Both types require a greater depth
of tailwater than a hydraulic jump stilling basin. However, the hydraulic action and the
resulting performance of the two buckets are different. In general, the slotted bucket is
an improvement over the solid type, particularly for lower ranges of tail water depths.
USBR (Peterka, 1964) developed a simplified seven-step design procedure for the slot-
ted bucket as follows:

1. Determine Q, q (per foot or meter of bucket width), V1, D1; compute Froude number
from F  V1/(g D1)1/2 for maximum flow and intermediate flows.
2. Enter Fig. 18.22 with F to find bucket radius parameter R/(D1 V12/2g) from which
minimum allowable bucket radius, R, may be computed.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.39

FIGURE 18.21 Curve to determine maximum stone size in riprap mixture. (From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.40 Chapter Eighteen

FIGURE 18.22 Minimum allowable bucket radius for slotted and solid buckets. .(From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.41

FIGURE 18.23 Minimum tail waterlimit for slotted and solid buckets..(From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.42 Chapter Eighteen

FIGURE 18.24 Maximum tail water limit for slotted and solid buckets. .(From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.43

FIGURE 18.25 Examples of submerged bucket designs.


.(From Peterka, 1964)

FIGURE 18.26 Average water surface profiles


for submerged buckets. (From Peterka, 1964)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.44 Chapter Eighteen

FIGURE 18.27 Water surface profile characteristics


for slotted buckets (From Peterka, 1964).

3. Enter Fig. 18.23 with R/(D1 V12/2g) and F to find Tmin/D


/ 1 from which minimum tail-
water depth limit Tmin, may be computed.
4. Enter Fig. 18.24 as in Step 3 above to find maximum tailwater depth limit, Tmax.
5. Set bucket invert elevation so that tail water curve elevations are between tailwater
depth limits determined by Tmin and Tmax. Keep apron lip and bucket invert above
riverbed, if possible. For best performance, set bucket so that the tailwater depth is near
Tmin. Check factor of safety against sweep out.
6. Complete the design of the bucket, using Fig. 18.25 to obtain tooth size, spacing,
dimensions, and so on.
7. Use Figs. 18.26 and 18.27 to estimate the water surface profile in and downstream
from the bucket.

18.9 FLIP BUCKETS


Flip bucket or ski-jump energy dissipators are often used in association with high over-
flow dams to reduce the project cost when spray from the jet can be tolerated and the ero-
sion by the plunging jet can be controlled. Most of the energy is dissipated when the jet
plunges into the tailwater. Factors affecting the horizontal throw distance from the bucket
lip to the point of jet impact are the exit velocity of the jet at the bucket lip, the bucket lip

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.45

angle, and the difference in elevation between the lip and the tailwater. With the origin of
the coordinates taken at the lip of the bucket, the trajectory of the jet may be expressed by
the following equation:

(a)

(b)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.46 Chapter Eighteen

(c)
Exhibit 18.3: Strontia springs project, Colorado (Courtesy Denver Water Department, Denver, Colorado)
(a) General view of the spillway with low-level-outlet-work in operation.
(b) General view of the spillway with low-level-outlet-work in operation.
(c) Close-up view of the spillway in operation showing free trajectory and impact at the
plunge pool.

(a)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.47

(b) Layout of the dam showing spillway, plunge pool, power intakes, power house and diversion tunnels.
(a) A view of the spillway in operation showing free tajectory and imact at the plunge pool.
(b)
Exhibit 18.4 Mossyrock Hydroelectric Project, Mossyrock, Washington

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.48 Chapter Eighteen

FIGURE 18.28 Flip bucket and toe curve pressures. (From USACE, 1998)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.49

x2
y  x tan θ  
K[4 (d  hv) cos2 θ]
where θ  angle of edge of lip with horizontal, K  factor usually assumed as 0.9 to com-
pensate for loss of energy, d  depth of water on bucket, hv  velocity head of jet at the
lip of the bucket

In general, the exit angle at the lip should not exceed 30º and the minimum radius of
curvature should not be less than 5 times the depth of water on the bucket. The pressure
distribution on spillway flip buckets associated with high–overflow dams can be estimat-
ed based on the Corps of Engineers test data (USACE, 1988) as shown in Fig. 18.28. For
design purposes, allowance for spillway energy losses should be included in the compu-
tation of the energy head, HT at the invert of the bucket. A discussion of the plunge
pool hydrulics including scour depth and jet diffusion is given in section 17.3. Photos
of several flip bucket type energy dissipators are given in Exhibits 17.2, 17.3, 17.11,
18.3, and 18.4

18.9.3 Gas Supersaturation

Gas supersaturation problems occur at dams with spillways designed with deep plunge
pools and with deep stilling basins that operate submerged hydraulic jumps. When spilled
water with entrained air plunges to depths where the pressures can significantly exceed
one atmosphere, the flow becomes supersaturated with gasses. Fish exposed to these gas
supersaturated conditions develop gas emboli in the tissues. This condition known as gas
bubble disease, cause injury to the fish, and can result in death..
When considering the selection and design of an energy dissipator for use in a dam
project, gas supersaturation must be considered. Deflectors that direct discharges along
the surface and energy dissipating devices that disperse the flow to reduce the depth of the
plunge, such as Howell-Bunger valves, are considerations.
Stilling basins that are designed for high unit discharges, but primarily operate for lower
discharges often have deep basins and excess tailwater depths for the lower discharges. In
this condition the hydraulic jump is submerged, with the flow plunging to the bottom of a
deep pool in the stilling basin. In large spillways these conditions can cause supersatura-
tion. In situations where it is not practical to use alternative energy dissipators or design the
spillway and stilling basin with lower unit discharges, it may be necessary to divide the
spillway and stilling basin with walls. This permits operation of a portion of the structure
at higher unit discharge for lower releases, thus effectively reducing the tailwater excess.

18.9.4 Abrasion in Stilling Basins

Many stilling basins are subject to at least some wear due to abrasion from material that
gets washed into the basin and circulates in contact with the concrete surfaces with the
flow. To minimize problems due to abrasion, stilling basins should be operated with
uniform discharge. Spillways with crest gates should be operated with all gates opened
equally to avoid recirculation in the stilling basin. When only one or a few gates are
opened on a spillway with a wide stilling basin, circulation patterns develop in the spill-
way which can transport streambed material from downstream into the stilling basin. It is
necessary for the designer to consider all conditions under which the spillway and energy

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.50 Chapter Eighteen

dissipator will operate. If nonuniform operation of gates is expected to be required to pass


low discharges, consideration should be given to dividing the spillway and stilling basin
with guide walls to provide a portion that could be used to pass low discharges without
creating recirculation patterns in the stilling basin.

18.10 STILLING BASIN DESIGN EXAMPLES

18.10.1 Design Example 1

The crest of an overfall spillway is 200 ft (61 m) above the horizontal floor of the still-
ing basin and the slope of the spillway chute is 0.7:1. The head (H)
H on the spillway crest
is 30 ft (9.14 m) and the maximum unit discharge (q) is 480 (ft3/s/ft) [44.6 (m3/s/m]
based on the the stilling basin width. Design a Type II stilling basin for these conditions
(Peterka, 1964).

Step 1. Determine approach conditions including velocity (V


V1) of flow entering the basin.

a. Compute the total distance from the reservoir level to the basin floor (total fall)
Z.
Z  head on the crest (H)
H + vertical distance from crest to basin floor
 30  200  230 ft (70.1 m)
b. Entering Fig. 18.1 with Z (  230 ft) and H (  30 ft) and determine the ratio of
VA) versus theoretical velocity (V
actual velocity (V VT) that is,
VA
  0.92
VT
c. Compute the theoretical velocity based on the equation given in Fig. 18.2.

VT  冪2莦g莦莦2莦3莦0莦莦莦3莦20莦  117.6 ft/s (35.8 m/s)







d. Compute the actual velocity VA ( V1 of the jump) and the corresponding depth
D1 and the approach Froude number F1.

V1  VA  117.6  0.92  108.2 ft/s (33.0 m/s)

q 480  4.44 ft (1.35 m)


D1     
V1 108.2
V1 108.2
F1     9.04
兹g苶苶

兹苶 D1苶 兹
兹3苶苶2苶
.2苶
苶4苶.44苶
Step 2. Set basin apron elevation:
a. Determine tailwater depths. Entering Fig. 18.3b with the Froude number (F1)
W/D2  1.0 gives
of 9.04, the heavy dashed line for TW/

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.51

W/D1  12.3
TW/

b. Compute D2.

D2  TW  12.3  D1  12.3  4.44  54.6 ft (16.6 m)

c. Check factor of safety (FS) with the minimum tailwater depth required as given
in Fig. 18.3b.

For F1  9.04, TW / 1  11.85, TW


Wmin/D Wmin  11.85  4.44  52.6 ft (16.0 m)
FS  (TW  TW / 2  (54.6  52.6) / 54.6
Wmin)/D
 4.0 percent  5 percent (recommended minimum margin of safety)

To satisfy 5 percent minimum margin of safety:


Wmin  0.05  D2  52.6  0.05  54.6  55.3 ft (16.9 m)
Use TW  TW
Reposition the stilling basin apron accordingly.
Step 3. Check the effectiveness of the stilling basin:

F1  9.04 4.0
The jump should be fully developed for effective energy dissipation.
Step 4. Determine the basin length:
a. Entering Fig. 18.3c with F1  9.04 and determine the corresponding value
of L/
L/D2.
L
  4.28
D2
b. LII  L  4.28  54.6  234 ft (71.3 m)
Step 5. Determine chute block height, width, and spacing.
Referring to Fig. 18.3a, the recommended height, width, and spacing of the
chute block is D1.
Height  width  spacing  D1  4.44 ft  4 ft 5.3 in (use 4 ft 6 in or 1 m
35 cm)
Step 6. Determine the height, width, and of the dentated sill based on the recommended
dimensions shown in Fig. 18.3a.
a. Height  0.2D2  0.2  54.6  10.92 ft (use 11 ft or 3 m 33 cm)
b. Width  spacing  0.15D2  0.15D2  0.15  54.6  8.19 ft (2.50 m) (use
8 ft 3 in or 2 m 50 cm)

18.10.2 Design Example 2

In this example (Peterka, 1964), the dimensions of the Type III stilling basin of a small
dam are to be determined. The width of the basin is 50 ft (15.24 m) and the flow is sym-
metrical. Based on the design of the spillway, values of V1 and D1 for the range of dis-
charges to be considered have been determined and are given as follows:

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.52 Chapter Eighteen

FIGURE 18.29 Tailwater and jump elevation curve for design example 2. (From peterka,
1964)

Q q V1 D1
ftt3/s (m3/s) ftt3/s/ft [(m3/s/m] ft/s (m/s) ft/s (m/s)

3,900 (110.5) 78.00 (7.25) 69.0 (21.0) 1.130 (0.344)


3,090 ( 87.5) 61.80 (5.74) 66.0 (20.1) 0.936 (0.285)
2,022 ( 57.3) 40.45 (3.76) 63.0 (19.2) 0.642 (0.196)
662 ( 18.7) 13.25 (0.87) 51.0 (15.5) 0.260 (0.079)

Resulting from a backwater analysis of the downstream channel, the tailwater rating
curve is also available as shown in Fig. 18.29. The tailwater elevation for 3900 ft3/s (110.5
m3/s) is at elevation 617.50 ft (188.22 m).

Step 1. Compute the “jump elevation curve.”


a. Compute F1 based on given V1 and D1 values.
(computed F1 values are shown in the table below)
b. Determine D2 by entering Fig. 18.4b with the computed values of F1.
c. Assume the most adverse operating condition occurs at the maximum discharge
of 3900 ft3/s (110.5 m3/s) and set the apron elevation accordingly.

D2  17.8 ft (5.43 m)

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.53

Apron elev.  617.5  17.8  El. 599.7 ft

 188.22  5.43  El. 182.8 m

d. Compute jump elevations for the remaining three discharges as shown in the table
below for “jump elevation Curve a.”

Q F1 D1 D2 / D1 D2 Jump Elev. Jump Elev.


ftt3/s ft ft Curve a, ft Curve a', ft

3900 11.42 1.130 15.75 17.80 617.5 615.0


3090 12.02 0.936 16.60 15.54 615.2 612.7
2022 13.85 0.642 19.20 12.33 612.0 609.5
662 17.62 0.260 24.50 6.37 606.1 603.6
or

Q F1 D1 D2 / D1 D2 Jump Elev. Jump Elev.


m3/s m m Curve a, m Curve a', m

110.5 11.42 0.344 15.75 5.43 188.22 187.45


87.5 12.02 0.285 16.60 4.74 187.52 186.75
57.3 13.85 0.196 19.20 3.76 186.54 185.78
18.7 17.62 0.079 24.50 1.94 184.74 183.98

Step 2. Compare the jump elevation curve with the tailwater rating curve as shown in
Fig. 18.29.

It indicates tailwater depth deficiency for smaller discharges especially at approx-


imately 2850 ft3/s (80.7 m3/s) where the curvature of the tailwater rating curve is
concave upward.

Step 3. Shift the apron elevation curve downward such that the full conjugate depth is
realized at the most adverse 2850 ft3/s (80.7 m3/s) tailwater condition. A down-
ward shift of 2.5 ft (0.76 m) is required as indicated by “jump elevation Curve a'”
in Fig. 18.29 and the accompanying table.

Step 4. Reset the apron elevation:

Apron elev.  599.7  2.5  El. 597.20ft

 182.79  0.76  El. 182.03 m

Step 5. Determine the remaining stilling basin details based on the maximum discharge
of 3900 ft3/s (110.5 m3/s).
Step 6. Determine basin length based on conjugate depth:
Entering Fig. 18.4c with F1  11.42.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

18.54 Chapter Eighteen


LIII
  2.75
D2
The basin length  LIII  2.75  17.80  48.95 ft (14.9 m)
Step 7. Determine the height, width, and spacing of chute blocks in accordance with
Fig. 18.4a.

h1  W1  S1  1.0D1 1.13 ft (use 13 or 14 in) (35 cm)

Step 8. Determine height of the baffle piers in accordance with Fig. 18.4d.

h3  2.5D1  2.5  1.13  2.825 ft (use 34 in) (86 cm)

Step 9. Compute the spacing of the baffle piers as 0.75h3.

Baffle pier spacing  0.75  34  25.5 in (65 cm)

Step 10.Compute the distance between the baffle piers and the chute blocks as 0.8D2.

Distance  0.8  17.8  14.24 ft (4.34 m)

Step 11.Compute the height of the solid end sill h4 based on Fig. 18.4d.

h4  1.60D1  1.60  1.13  1.81 ft (use 22 in) (55 cm)

The final dimensions of the Type III stilling basin are shown in Fig. 18.29.

REFERENCES
Blaisdell, F. W., “Develop and Hydraulic Design—Saint Anthony Falls Stilling Basin,”
Transactions, ASCE, 113, P.334 1948.
Bowers, C. E., and J. W. Toso, “Karnafuli Project, Model Studies of Spillway Damage,” Journal of
Hydraulic Engineering, ASCE, 114 (5), 1988.
Bowers, C. E., and F. Y. Tsai, “Fluctuating Pressures in Spillway Stilling Basins,” Journal of
Hydraulic Engineering, ASCE, 95 (HY6), 1969.
Chadwick, A. J., and J. C. Morfett, Hydraulics in Civil Engineering, Allen & Unwin, London,
1986.
Chaudhry, M. H., Open-Channel Flow, Prentice-Hall, Englewood Cliffs, NJ, 1993.
Chow, V. T., Open-Channel Hydraulics, McGraw-Hill, New York, 1959.
French, R. H., Open-Channel Hydraulics, McGraw-Hill, New York, 1985.
George, R. L., Low Froude Number Stilling Basin Design, REC-ERC-78-8, U.S. Bureau of
Reclamation, 1978.
Henderson, F. M., Open Channel Flow, Macmillan, New York, 1966.
International Commission on Large Dams (ICOLD), Spillways for Dams, Bulletin 58, ICOLD,
Paris, 1987.
Novak, P., A. I. B. Moffat, C. Nalluri, and R. Narayanan, Hydraulic Structures, Unwin Hyman,
London, 1990.
Peterka, A. J., Hydraulic Design of Stilling Basins and Energy Dissipators, Engineering

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
HYDRAULIC DESIGN OF STILLING BASINS AND ENERGY DISSIPATORS

Hydraulic Design of Stilling Basins and Energy Dissipators 18.55

Monograph No. 25, U.S. Bureau of Reclamation, Denver, Co, 1964.


Roberson, J. A., J. J. Cassidy, and M. H. Chaudhry, Hydraulic Engineering, Houghton Mifflin,
Boston, 1988.
Senturk, F., Hydraulics of Dams and Reservoirs, Water Resources Publications, Highlands Ranch,
COl 1994.
Toso, J. W., and C. E. Bowers, “Extreme Pressures in Hydraulic-Jump Stilling Basins,” Journal of
Hydraulic Engineering, ASCE, 114 (8), 1988.
U.S. Army Corps of Engineers (USACE), Hydraulic Design Criteria, U.S. Army Corps of
Engineers Waterways Experiment Station, Vicksburg, MS, 1988.
U.S. Bureau of Reclamation (USBR), Small Canal Structures, U.S. Bureau of Reclamation,
Denver, CO, 1974.
U.S. Bureau of Reclamation (USBR), Design of Small Dams, U.S. Bureau of Reclamation, Denver,
CO, 1987.
Vischer D. L., and W. H. Hager, Energy Dissipators—Hydraulic Design Considerations, IAHR
Hydraulic Structures Design Manual No. 9, A. A. Balkema, Rotterdam, Netherlands, 1995.
Vischer D. L., and W. H. Hager, Dam Hydraulics, John Wiley & Sons, New York, 1998.

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright © 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.

You might also like