Sunteți pe pagina 1din 128

W

ater resources engineering is con-


cerned with the protection, devel-
opment, and efficient management
of water resources for beneficial
purposes. It involves planning, design, and con-
struction of projects for supply of water for domes-
tic, commercial, public, and industrial purposes,
flood prevention, hydroelectric power, control of
rivers and water runoff, and conservation of water
resources, including prevention of pollution.
Water resources engineering primarily deals
with water sources, collection, flow control, trans-
mission, storage, and distribution. For efficient man-
agement of these aspects, water resources engineers
require a knowledge of fluid mechanics; hydraulics
of pipes, culverts, and open channels; hydrology;
water demand, quality requirements, and treat-
ment; production of water from wells, lakes, rivers,
and seas; transmission and distribution of water
supplies; design of reservoirs and dams; and pro-
duction of hydroelectric power. These subjects are
addressed in the following articles.
21.1 Dimensions and Units
A list of symbols and their dimensions used in this
section is given in Table 21.1. Table 21.2 lists conver-
sion factors for commonly used quantities, includ-
ing the basic equivalents between the English and
metric systems. For additional conversions to the
metric system (SI) of units, see the appendix.
Fluid Mechanics
Fluid mechanics describes the behavior of water
under various static and dynamic conditions. This
theory, in general, has been developed for an ideal
liquid, a frictionless, inelastic liquid whose parti-
cles follow smooth flow paths. Since water only
approaches an ideal liquid, empirical coefficients
and formulas are used to describe more accurately
the behavior of water. These empiricisms are
intended to compensate for all neglected and
unknown factors.
The relatively high degree of dependence on
empiricism, however, does not minimize the
importance of an understanding of the basic theo-
ry. Since major hydraulic problems are seldom
identical to the experiments from which the empir-
ical coefficients were derived, the application of
fundamentals is frequently the only means avail-
able for analysis and design.
21.2 Properties of Fluids
Specific weight or unit weight w is defined as
weight per unit volume. The specific weight of
water varies from 62.42 lb/ft
3
at 32 F to 62.22 lb/ft
3
at 80 F but is commonly taken as 62.4 lb/ft
3
for the
majority of engineering calculations. The specific
weight of sea water is about 64.0 lb/ft
3
.
Density is defined as mass per unit volume and
is significant in all flow problems where acceleration
21
WATER RESOURCES
ENGINEERING
*
M. Kent Loftin
Chief Civil Engineer
South Florida Water Management District
West Palm Beach, Florida
*Revised and updated from Sec. 21, Water Engineering, by Samuel B. Nelson, in the third edition.
21.1
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.2
I
Section Twenty-One
Symbol Terminology Dimensions Units
A Area L
2
ft
2
C Chezy roughness coefficient L
1/ 2
/ T ft
1/ 2
/s
C
1
Hazen-Williams roughness coefficient L
0.37
/ T ft
0.37
/s
d Depth L ft
d
c
Critical depth L ft
D Diameter L ft
E Modulus of elasticity F/ L
2
psi
F Force F lb
g Acceleration due to gravity L/ T
2
ft/s
2
H Total head, head on weir L ft
h Head or height L ft
h
f
Head loss due to friction L ft
L Length L ft
M Mass FT
2
/ L lbs
2
/ ft
n Manning's roughness coefficient T/ L
1/3
s/ ft
1/3
P Perimeter, weir height L ft
P Force due to pressure F lb
p Pressure F/ L
2
psf
Q Flow rate L
3
/ T ft
3
/ s
q Unit flow rate L
3
/ TL ft
3
/ (sft)
r Radius L ft
R Hydraulic radius L ft
T Time T s
t Time, thickness T, L s, ft
V Velocity L/ T ft / s
W Weight F lb
w Specific weight F/ L
3
lb/ft
3
y Depth in open channel, distance from solid boundary L ft
Z Height above datum L ft
Size of roughness L ft
Viscosity FT/ L
2
lbs/ ft
Kinematic viscosity L
2
/ T ft
2
/ s
Density FT
2
/ L
4
lbs
2
/ ft
4
Surface tension F/ L lb/ft
Shear stress F/ L
2
psi
Symbols for dimensionless quantities
Symbol Quantity
C Weir coefficient, coefficient of discharge
C
c
Coefficient of contraction
C

Coefficient of velocity
F Froude number
f Darcy-Weisbach friction factor
K Head-loss coefficient
R Reynolds number
S Friction slopeslope of energy grade line
S
c
Critical slope
Efficiency
Sp. gr. Specific gravity
Table 21.1 Symbols, Dimensions, and Units Used in Water Engineering
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.3
is important. It is obtained by dividing the specific
weight w by the acceleration due to gravity g. The
variation of g with latitude and altitude is small
enough to warrant the assumption that its value is
constant at 32.2 ft/s
2
in hydraulics computations.
The specific gravity of a substance is the ratio
of its density at some temperature to that of pure
water at 68.2 F (20 C).
Modulus of elasticity E of a fluid is defined as
the change in pressure intensity divided by the cor-
responding change in volume per unit volume. Its
value for water is about 300,000 psi, varying slight-
ly with temperature. The modulus of elasticity of
water is large enough to permit the assumption
that it is incompressible for all hydraulics problems
except those involving water hammer (Art. 21.13).
Surface tension and capillarity are a result of
the molecular forces of liquid molecules. Surface
tension is due to the cohesive forces between liq-
uid molecules. It shows up as the apparent skin
that forms when a free liquid surface is in contact
with another fluid. It is expressed as the force in
the liquid surface normal to a line of unit length
drawn in the surface. Surface tension decreases
with increasing temperature and is also dependent
on the fluid with which the liquid surface is in con-
tact. The surface tension of water at 70F in contact
with air is 0.00498 lb/ ft.
Capillarity is due to both the cohesive forces
between liquid molecules and adhesive forces of
liquid molecules. It shows up as the difference in
liquid surface elevations between the inside and
outside of a small tube that has one end sub-
merged in the liquid. Since the adhesive forces of
water molecules are greater than the cohesive
forces between water molecules, water wets a sur-
Area
1 acre = 43,560 ft
2
1 mi
2
= 640 acres
Volume
1 ft
3
= 7.4805 gal
1 acre-ft = 325,850 gal
1 MG = 3.0689 acre-ft
Power
1 hp = 550 ft lb/s
1 hp = 0.746 kW
1 hp = 6535 kWh/ year
Weight of water
1 ft
3
weighs 62.4 lb
1 gal weighs 8.34 lb
Table 21.2 Conversion Table for Commonly Used Quantities
Discharge
1 ft
3
/s = 449 gal/min = 646,000 gal/day
1 ft
3
/s = 1.98 acre-ft/day = 724 acre-ft/year
1 ft
3
/s = 50 miners inches in Idaho, Kansas,
Nebraska, New Mexico, North Dakota, and
South Dakota
1 ft
3
/s = 40 miners inches in Arizona, California,
Montana, and Oregon
1 MGD* = 3.07 acre-ft/day = 1120 acre-ft/year
1 MGD* = 1.55 ft
3
/s = 694 gal/min
1 million acre-ft/year = 1380 ft
3
/s
Pressure
1 psi = 2.31 ft of water
= 51.7 mm of mercury
1 in of mercury = 1.13 ft of water
1 ft of water = 0.433 psi
1 atm

= 29.9 in of mercury = 14.7 psi


Metric equivalents
Length: 1 ft = 0.3048 m
Area: 1 acre = 4046.9 m
2
Volume: 1 gal = 3.7854 L
1 m
3
= 264.17 gal
Weight: 1 lb = 0.4536 kg
*Prefix M indicates million; for example, MG = million gallons

atm indicates atmospheres.


Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.4
I
Section Twenty-One
face and rises in a small tube, as shown in Fig. 21.1.
Capillarity is commonly expressed as the height of
this rise. In equation form,
(21.1)
where h = capillary rise, ft
= surface tension, lb/ft
w
1
and w
2
= specific weights of fluids below and
above meniscus, respectively, lb/ft
= angle of contact
r = radius of capillary tube, ft
Capillarity, like surface tension, decreases with
increasing temperature. Its temperature variation,
however, is small and insignificant in most problems.
Surface tension and capillarity, although negli-
gible in many water engineering problems, are sig-
nificant in others, such as capillary rise and flow of
liquids in narrow spaces, formation of spray from
water jets, interpretation of the results obtained on
small models, and freezing damage to concrete.
Atmospheric pressure is the pressure due to the
weight of the air above the earths surface. Its value
at sea level is 2116 psf or 14.7 psi. The variation in
atmospheric pressure with elevation from sea level
to 10,000 ft is shown in Fig. 21.2. Gage pressure, psi,
is pressure above or below atmospheric. Absolute
pressure, psia, is the total pressure including
atmospheric pressure. Thus, at sea level, a gage
pressure of 10 psi is equivalent to 24.7 psia. Gage
pressure is positive when pressure is greater than
atmospheric and is negative when pressure is less
than atmospheric.
Vapor pressure is the partial pressure caused
by the formation of vapor at the free surface of a
liquid. When the liquid is in a closed container, the
partial pressure due to the molecules leaving the
surface increases until the rates at which the mole-
cules leave and reenter the liquid are equal. The
vapor pressure at this equilibrium condition is
called the saturation pressure. Vapor pressure
increases with increasing temperature, as shown in
Fig. 21.3.
Cavitation occurs in flowing liquids at pres-
sures below the vapor pressure of the liquid. Cavi-
tation is a major problem in the design of pumps
and turbines since it causes mechanical vibrations,
pitting, and loss of efficiency through gradual
destruction of the impeller. The cavitation phe-
nomenon may be described as follows:
Because of low pressures, portions of the liquid
vaporize, with subsequent formation of vapor cav-
ities. As these cavities are carried a short distance
downstream, abrupt pressure increases force them
Fig. 21.1 Capillary action raises water in a
small-diameter tube. Meniscus, or liquid surface, is
concave upward.
Fig. 21.2 Atmospheric pressure decreases with
elevation above mean sea level. The curve is based
on the ICAO standard atmosphere.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.5
to collapse, or implode. The implosion and ensu-
ing inrush of liquid produce regions of very high
pressure, which extend into the pores of the metal.
(Pressures as high as 350,000 psi have been mea-
sured in the collapse of vapor cavities by the Fluid
Mechanics Laboratory at Stanford University.)
Since these vapor cavities form and collapse at
very high frequencies, weakening of the metal
results as fatigue develops, and pitting appears.
Cavitation may be prevented by designing
pumps and turbines so that the pressure in the liq-
uid at all points is always above its vapor pressure.
Viscosity, of a fluid, also called the coefficient
of viscosity, absolute viscosity, or dynamic viscos-
ity, is a measure of its resistance to flow. It is
expressed as the ratio of the tangential shearing
stresses between flow layers to the rate of change
of velocity with depth:
(21.2)
where = shearing stress, lb/ft
2
V = velocity, ft/s
y = depth, ft
Viscosity decreases as temperature increases but
may be assumed independent of changes in pres-
sure for the majority of engineering problems.
Water at 70 F has a viscosity of 0.00002050 lbs/ft
2
.
Kinematic viscosity is defined as viscosity
divided by density . It is so named because its
units, ft
2
/s, are a combination of the kinematic units
of length and time. Water at 70 F has a kinematic
viscosity of 0.00001059 ft
2
/s.
In hydraulics, viscosity is most frequently
encountered in the calculation of Reynolds num-
ber (Art. 21.8) to determine whether laminar, tran-
sitional, or completely turbulent flow exists.
21.3 Fluid Pressures
Pressure or intensity of pressure p is the force per
unit area acting on any real or imaginary surface
within a fluid. Fluid pressure acts normal to the
surface at all points. At any depth, the pressure acts
equally in all directions. This results from the
inability of a fluid to transmit shear when at rest.
Liquid and gas pressures differ in that the varia-
tion of pressure with depth is linear for a liquid
and nonlinear for a gas.
Hydrostatic pressure is the pressure due to
depth. It may be derived by considering a sub-
merged rectangular prism of water of height h, ft,
and cross-sectional area A, ft
2
, as shown in Fig. 21.4.
The boundaries of this prism are imaginary. Since
the prism is at rest, the summation of all forces in
both the vertical and horizontal directions must be
zero. Let w equal the specific weight of the liquid,
lb/ft
3
. Then, the forces acting in the vertical direc-
tion are the weight of the prism wA h, the force
due to pressure p
1
, psf, on the top surface, and the
force due to pressure p
2
, psf, on the bottom surface.
Summing these vertical forces and setting the total
equal to zero yields
Fig. 21.3 Vapor pressure of water increases rapidly with temperature.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.6
I
Section Twenty-One
(21.3a)
Division of Eq. (21.3a) by A yields
(21.3b)
For the special case where the top of the prism
coincides with the water surface, p
1
is atmospheric
pressure. Since most hydraulics problems involve
gage pressure, p
1
is zero (gage pressure is zero at
atmospheric pressure). Taking h to be h, the depth
below the water surface, ft, then p
2
is p, the pres-
sure, psf, at depth h. Equation (21.3b) then becomes
(21.4)
Equation (21.4) gives the depth of water h of
specific weight w required to produce a gage pres-
sure p. By adding atmospheric pressure p
a
to Eq.
(21.4), absolute pressure p
ab
is obtained as shown in
Fig. 21.4. Thus,
(21.5)
21.3.1 Pressures on Submerged
Plane Surfaces
This is important in the design of weirs, dams,
tanks, and other water control structures. For hor-
izontal surfaces, the pressure-force determination
is a simple matter since the pressure is constant.
For determination of the pressure force on inclined
or vertical surfaces, however, the summation con-
cepts of integral calculus must be used.
Figure 21.5 represents any submerged plane
surface of negligible thickness inclined at an angle
with the horizontal. The resultant pressure force
P, lb, acting on the surface is equal to p dA. Since
p = wh and h = y sin , where w is the specific
weight of water, lb/ft
3
,
(21.6)
Equation (21.6) can be simplified by setting
ydA = y

A, where A is the area of the submerged


surface, ft
2
; and y

sin = h

, the depth of the cen-


troid, ft. Therefore,
(21.7)
Fig. 21.4 Hydrostatic pressure varies linearly with depth.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.7
where p
cg
is the pressure at the centroid, psf.
The point on the submerged surface at which
the resultant pressure force acts is called the center
of pressure (c.p.). It is below the center of gravity
because the pressure intensity increases with
depth. The location of the center of pressure, rep-
resented by the length y
p
, is calculated by summing
the moments of the incremental forces about an
axis in the water surface through point W (Fig.
21.5). Thus, Py
p
= y dP. Since dP = wy sin dA and
P = wy sin dA,
(21.8)
The quantity y
2
dA is the moment of inertia of
the area about the axis through W. It also equals
AK
2
+ Ay

2
, where K is the radius of gyration, ft, of
the surface about its centroidal axis. The denomi-
nator of Eq. (21.8) equals y.

A. Hence
(21.9)
and K
2
/y

is the distance between the centroid and


center of pressure.
Values of K
2
for some common shapes are given
in Fig. 21.6 (see also Fig. 6.29). For areas for which
radius of gyration has not been determined, y
p
may
be calculated directly from Eq. (21.8).
The horizontal location of the center of pres-
sure may be determined as follows: It lies on the
vertical axis of symmetry for surfaces symmetrical
about the vertical. It lies on the locus of the mid-
points of horizontal lines located on the sub-
merged surface, if that locus is a straight line. Oth-
erwise, the horizontal location may be found by
taking moments about an axis perpendicular to the
one through W in Fig. 21.5 and lying in the plane
of the submerged surface.
Example 21.1: Determine the magnitude and
point of action of the resultant pressure force on a
5-ft-square sluice gate inclined at an angle of 53.2
to the horizontal (Fig. 21.7).
From Eq. (21.7), the total force P = wh

A, with
Fig. 21.5 Total pressure on a submerged plane surface depends on pressure at the center of gravity
(c.g.) but acts at a point (c.p.) that is below the c.g.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.8
I
Section Twenty-One
Thus, P = 62.4 4 25 = 6240 lb. From Eq. (21.9),
its point of action is a distance y
p
= y

+ K
2
/y

from
point G, and y

= 2.5 +
1
/2(5.0) = 5.0 ft. Also, K
2
=
b
2
/12 = 5
2
/12 = 2.08. Therefore, y
p
= 5.0 + 2.08/5 =
5.0 + 0.42 = 5.42 ft.
21.3.2 Pressure on Submerged
Curved Surfaces
The resultant pressure force on submerged curved
surfaces cannot be calculated from the equations
developed for the pressure force on submerged
plane surfaces because of the variation in direction
of the pressure force. The resultant pressure force
can be calculated, however, by determining its hor-
izontal and vertical components and combining
them vectorially.
A typical configuration of pressure on a sub-
merged curved surface is shown in Fig. 21.8. Con-
sider ABC a 1-ft-thick prism and analyze it as a free
body by the principles of statics. Note:
1. The horizontal component P
H
of the resultant
pressure force has a magnitude equal to the
Fig. 21.6 Radius of gyration and location of centroid (c.g.) of common shapes.
Fig. 21.7 Sluice gate (crosshatched) is subjected to hydrostatic pressure. (See Example 21.1.)
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.9
pressure force on the vertical projection AC of
the curved surface and acts at the centroid of
pressure diagram ACDE.
2. The vertical component P
V
of the resultant pres-
sure force has a magnitude equal to the sum of
the pressure force on the horizontal projection
AB of the curved surface and the weight of the
water vertically above ABC. The horizontal
location of the vertical component is calculated
by taking moments of the two vertical forces
about point C.
When water is below the curved surface, such
as for a taintor gate (Fig. 21.9), the vertical compo-
nent P
V
of the resultant pressure force has a mag-
nitude equal to the weight of the imaginary vol-
ume of water vertically above the surface. P
V
acts
upward through the center of gravity of this imag-
inary volume.
Example 21.2: Calculate the magnitude and
direction of the resultant pressure on a 1-ft-wide
strip of the semicircular taintor gate in Fig. 21.9.
The magnitude of the horizontal component P
H
of the resultant pressure force equals the pressure
force on the vertical projection of the taintor gate.
From Eq. (21.7), P
H
= wh

A = 62.4 2.5 5 = 780 lb.


The magnitude of the vertical component of the
resultant pressure force equals the weight of the
imaginary volume of water in the prism ABC above
the curved surface. The volume of this prism is
R
2
/4 = 3.14 25/4 = 19.6 ft
3
, so the weight of the
water is 19.6w = 19.6 62.4 = 1220 lb = P
V
.
The magnitude of the resultant pressure force
equals
The tangent of the angle the resultant pressure
force makes with the horizontal = P
V
/P
H
=
1220/780 = 1.564. The corresponding angle is 57.4.
The positions of the horizontal and vertical
components of the resultant pressure force are not
required to find the point of action of the resultant.
Its angle with the horizontal is known, and for a
constant-radius surface, the resultant must act per-
pendicular to the surface.
Fig. 21.8 Hydrostatic pressure on a submerged
curved surface. (a) Pressure variation over the sur-
face. (b) Free-body diagram.
Fig. 21.9 Taintor gate has submerged curved
surface under pressure. Vertical component of
pressure acts upward. (See Example 21.2.)
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.10
I
Section Twenty-One
21.4 Submerged and Floating
Bodies
The principles of buoyancy govern the behavior of
submerged and floating bodies and are important in
determining the stability and draft of cargo vessels.
The buoyant force acting on a submerged body equals
the weight of the volume of liquid displaced.
A floating body displaces a volume of liquid equal to
its weight.
The buoyant force acts vertically through the center
of buoyancy c.b., which is located at the center of gravi-
ty of the volume of liquid displaced.
For a body to be in equilibrium, whether floating
or submerged, the center of buoyancy and center of
gravity must be on the same vertical line AB (Fig.
21.10a). The stability of a ship, its tendency not to
overturn when it is in a nonequilibrium position, is
indicated by the metacenter. It is the point at which a
vertical line through the center of buoyancy inter-
sects the rotated position of the line through the
centers of gravity and buoyancy for the equilibrium
condition AB (Fig. 21.10b). The ship is stable only if
the metacenter is above the center of gravity since
the resulting moment for this condition tends to
right the ship.
The distance between the ships metacenter
and center of gravity is called the metacentric height
and is designated by y
m
in Fig. 21.10b. Given in feet
by Eq. (21.10) y
m
is a measure of degree of stability
or instability of a ship since the magnitudes of
moments produced in a roll are directly propor-
tional to this distance.
(21.10)
where I = moment of inertia of ships cross
section at waterline about longitudi-
nal axis through 0, ft
4
V = volume of displaced liquid, ft
3
y
s
= distance, ft, between centers of
buoyancy and gravity when ship is
in equilibrium
The negative sign should be used when the center
of gravity is above the center of buoyancy.
21.5 Manometers
A manometer is a device for measuring pressure. It
consists of a tube containing a column of one or
two liquids that balances the unknown pressure.
The basis for the calculation of this unknown pres-
sure is provided by the height of the liquid col-
umn. All manometer problems may be solved with
Eq. (21.4), p = wh. Manometers indicate h, the pres-
sure head, or the difference in head.
The primary application of manometers is mea-
surement of relatively low pressures, for which
aneroid and Bourdon gages are not sufficiently
Fig. 21.10 Stability of a ship depends on the location of its metacenter relative to its center of gravity (c.g.).
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.11
accurate because of their inherent mechanical lim-
itations. However, manometers may also be used
in precise measurement of high pressures by
arranging several U-tube manometers in series
(Fig. 21.12c). Manometers are used for both static
and flow applications, although the latter is most
common.
Three basic types are used (shown in Fig.
21.11): piezometer, U-tube manometer, and differ-
ential manometer. Following is a brief discussion of
the basic types.
The piezometer (Fig. 21.11a) consists of a tube
with one end tapped flush with the wall of the con-
tainer in which the pressure is to be measured and
the other end open to the atmosphere. The only liq-
uid it contains is the one whose pressure is being
measured (the metered liquid). The piezometer is a
sensitive gage, but it is limited to the measurement
of relatively small pressures, usually heads of 5 ft of
water or less. Larger pressures would create an
impractically high column of liquid.
Example 21.3: The gage pressure p
c
in the pipe
of Fig. 21.11a is 2.17 psi. The liquid is water with w
= 62.4 lb/ft
3
. What is h
m
?
Fig. 21.11 Basic types of manometers. (a) Piezometers; (b) U-tube manometer; (c) differential
manometer.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.12
I
Section Twenty-One
For pressures greater than 5 ft of water, the U-
tube manometer (Fig. 21.11b) is used. It is similar to
the piezometer except that it contains an indicat-
ing liquid with a specific gravity usually much
larger than that of the metered liquid. The only
other criteria are that the indicating liquid should
have a good meniscus and be immiscible with the
metered liquid.
The U-tube manometer is used when pressures
are either too high or too low for the piezometer.
High pressures can be measured by arranging U-
tube manometers in series (Fig. 21.12c). Very low
pressures, including negative gage pressures, can
be measured if the bottom of the U tube extends
below the center line of the container of the
metered liquid. The most common use of the U-
tube manometer is measurement of the pressures
of flowing water. In this application, the usual indi-
cating liquid is mercury.
A movable scale, as opposed to a fixed scale,
facilitates reading the U-tube manometer. The
scale is positioned between the two vertical legs
and moved to adjust for the variation in distance
h
m
from the center line of the pressure vessel to the
Fig. 21.12 Manometer shapes: (a) Sump in manometer to damp flow disturbances. (b) Inverted U for mea-
suring pressures on liquids with low specific gravity. (c) Series arrangement for measuring high pressures.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.13
indicating liquid. This zero adjustment enables a
direct reading of the heights h
i
and h
m
of the liquid
columns. The scale may be calibrated in any con-
venient units, such as ft of water or psi.
The differential manometer (Fig. 21.11c) is
identical to the U-tube manometer but measures
the difference in pressure between two points. (It
does not indicate the pressure at either point.) The
differential manometer may have either the stan-
dard U-tube configuration or an inverted U-tube
configuration, depending on the comparative spe-
cific gravities of the indicating and metered liq-
uids. The inverted U-tube configuration (Fig.
21.12b) is used when the indicating liquid has a
lower specific gravity than the metered liquid.
Example 21.4: A differential manometer (Fig.
21.11c) is measuring the difference in pressure
between two water pipes. The indicating liquid is
mercury (specific gravity = 13.6), h
i
is 2.25 ft, h
m1
is
9 in, and z is 1.0 ft. What is the pressure differential
between the two pipes?
The pressure at B, psf, is
p
B
= p
c2
+ w
2
h
m2
= p
c2
+ 62.4 2.0 = p
c2
+ 125
The pressure at A, psf, is
p
A
= p
c1
+ w
1
h
m1
+ w
i
h
i
= p
c1
+ 62.4 0.75 + 13.6 62.4 2.25
= p
c1
+ 1957
Since the pressure at A must equal that at B,
p
c2
+ 125 = p
c1
+ 1957
Hence, the pressure differential between the pipes is
p
c2
p
cl
= 1832 psf = 12.7 psi
When small pressure differences in water are
measured, if the specific gravity of the indicating
liquid is between 1.0 and 2.0 and the points at
which the pressure is being measured are at the
same level, the actual pressure difference, when
expressed in feet of water, is magnified by the dif-
ferential manometer. For example, if the actual dif-
ference is 0.50 ft of water and the indicating liquid
has a specific gravity of 1.40, the magnification will
be 2.5; that is, the height of the liquid column h
i
will
be 1.25 ft of water. The closer the specific gravities
of the metered and indicating liquids, the greater
the magnification and sensitivity. This is true only
up to a magnification of about 5. Above 5, the
increased sensitivity may be deceptive because the
meniscus between the two liquids becomes poorly
defined and sluggish in movement.
Many factors affect the accuracy of manome-
ters. Most of them, however, may be neglected in
the majority of hydraulics applications since they
are significant only in precise reading of manome-
ters, such as might be required in laboratories. One
factor, however, is significant: the existence of
surges in the manometer caused by the pulsations
and disturbances in the flow of water resulting
from turbulence. These surges make reading of the
manometer difficult. They may be reduced or elim-
inated by installing a large-diameter section, or
sump, in the manometer, as shown in Fig. 21.12a.
This sump will damp the pulsations and keep the
distance from the center line of the conduit to the
indicating liquid essentially at a constant value.
21.6 Fundamentals of Fluid
Flow
For fluid energy, the law of conservation of energy
is represented by the Bernoulli equation:
(21.11)
where Z
1
= elevation, ft, at any point 1 of flow-
ing fluid above an arbitrary datum
Z
2
= elevation, ft, at downstream point in
fluid above same datum
p
1
= pressure at 1, psf
p
2
= pressure at 2, psf
w = specific weight of fluid, lb/ft
3
V
1
= velocity of fluid at 1, ft/s
V
2
= velocity of fluid at 2, ft/s
g = acceleration due to gravity, 32.2 ft/s
2
The left side of the equation sums the total ener-
gy per unit weight of fluid at 1, and the right side,
the total energy per unit weight at 2. Equation
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.14
I
Section Twenty-One
(21.11) applies only to an ideal fluid. Its practical use
requires a term to account for the decrease in total
head, ft. through friction. This term h
f
, when added
to the downstream side of Eq. (21.11), yields the
form of the Bernoulli equation most frequently used:
(21.12)
The energy contained in an elemental volume
of fluid thus is a function of its elevation, velocity,
and pressure (Fig. 21.13). The energy due to eleva-
tion is the potential energy and equals WZ
a
, where
W is the weight, lb, of the fluid in the elemental
volume and Z
a
is its elevation, ft, above some arbi-
trary datum. The energy due to velocity is the
kinetic energy. It equals WV
2
a
/ 2g, where V
a
is the
velocity, ft/s. The pressure energy equals Wp
a
/w,
where p
a
is the pressure lb/ft
2
, and w is the specific
weight of the fluid, lb/ft
3
. The total energy, in the
elemental volume of fluid is
(21.13)
Dividing both sides of the equation by Wyields the
energy per unit weight of flowing fluid, or the total
head ft:
(21.14)
p
a
/w is called pressure head; V
2
a
/2g, velocity head.
As indicated in Fig. 21.13, Z + p/w is constant
for any point in a cross section and normal to the
flow through a pipe or channel. Kinetic energy at
the section, however, varies with velocity. Usually,
Z + p/w at the midpoint and the average velocity
at a section are assumed when the Bernoulli equa-
tion is applied to flow across the section or when
total head is to be determined. Average velocity,
ft/s = Q/A, where Q is the quantity of flow, ft
3
/s,
across the area of the section A, ft
2
.
Example 21.5: Determine the energy loss
between points 1 and 2 in the 24-in-diameter pipe in
Fig. 21.14. The pipe carries water flowing at 31.4 ft
3
/s.
Fig. 21.13 Energy in a liquid depends on ele-
vation, velocity, and pressure.
Fig. 21.14 Flow from an elevated reservoirapplication of the Bernoulli equation. (See Example 21.5.)
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.15
Average velocity in the pipe = Q/A = 31.4/ 3.14
= 10 ft/s. Select point 1 far enough from the reser-
voir outlet that V
1
can be assumed to be 0. Since the
datum plane passes through point 2, Z
2
= 0. Also,
since the pipe has free discharge, p
2
= 0. Thus sub-
stitution in Eq. (21.12) yields
where h
f
is the friction loss, ft. Hence, h
f
= 50 1.55
= 48.45 ft.
Note that in this example h
f
includes minor losses
due to the pipe entrance, gate valve, and any bends.
The Bernoulli equation and the variation of
pressure may be represented graphically, respec-
tively, by energy and hydraulic grade lines (Fig.
21.15). The energy grade line, sometimes called the
total head line, shows the decrease in total energy
per unit weight Hin the direction of flow. The slope
of the energy grade line is called the energy gradi-
ent or friction slope. The hydraulic grade line lies a
distance V
2
/2g below the energy grade line and
shows the variation of velocity or pressure in the
direction of flow. The slope of the hydraulic grade
line is termed the hydraulic gradient. In open-
channel flow, the hydraulic grade line coincides
with the water surface, while in pressure flow, it
represents the height to which water would rise in
a piezometer (see also Example 21.7, Art. 21.9).
Momentumis a fundamental concept that must
be considered in the design of essentially all water-
works facilities involving flow. A change in momen-
tum, which may result from a change in either
velocity, direction, or magnitude of flow, is equal to
the impulse, the force F acting on the fluid times the
period of time dt over which it acts. Dividing the
total change in momentum by the time interval
over which the change occurs gives the momentum
equation, or impulse-momentum equation:
Fig. 21.15 Energy grade line and hydraulic grade line indicate variations in energy and pressure
head, respectively, in a liquid as it flows along a pipe or channel.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.16
I
Section Twenty-One
(21.15)
where F
x
= summation of all forces in X direc-
tion per unit time causing change in
momentum in X direction, lb
= density of flowing fluid, lbs
2
/ft
4
(specific weight divided by g)
Q = flow rate, ft
3
/s
V
x
= change in velocity in X direction, ft/s
Similar equations may be written for the Y and Z
directions. The impulse-momentum equation often
is used in conjunction with the Bernoulli equation
[Eq. (21.11) or (21.12)] but may be used separately.
Example 21.6: Calculate the resultant force on
the reducer elbow in Fig. 21.16. The pipe center
line lies in a horizontal plane. The pipe reduces
from 48 in in diameter to 16 in. The pressure at the
upstream side of the reducer bend (point 1) is 100
psi, and the water flow is 100 ft
3
/s. (Neglect friction
loss at the bend.)
Velocity at points 1 and 2 is found by dividing
Q = 100 ft
3
/s by the respective areas: V
1
= 100
4/4
2
= 7.96 ft/s and V
2
= 100 4/1.33
2
= 71.5 ft/s.
With p
1
known, the Bernoulli equation for the
flow in the elbow is:
Solution of the equation yields the pressure at 2:
p
2
= 9500 psf
The total pressure force at 1 is P
1
= p
1
A
1
= 181,000
lb, and at 2, P
2
= p
p
A
2
= 13,200 lb.
Let R be the force, lb, exerted by the pipe on the
fluid (equal and opposite in direction to the force
against the pipe, which is to be determined). Then,
the force F changing the momentum of the fluid
equals the vector sum P
1
P
2
+ R. To find F, apply
Eq. (21.15) first in the X direction, then in the Y
direction, and determine the resultant of the forces:
In the X direction, since V
x
= (7.96 sin 53.2
71.5) = 65.1 and the density = 62.4/ 32.2= 1.94,
F
x
= 181,000 cos 53.2 13,200 + R
x
= 1.94 100 65.1
R
x
= 82,600 lb
In the Y direction, since V
y
= (7.96 cos 53.2 0)
= 4.78,
F
y
= 181,000 sin 53.2 + R
y
= 1.94 100 4.78
R
y
= 145,700 lb
The resultant R =

R
x
2
+R
y
2

= 167,500 lb. It acts


at an angle with the horizontal such that tan =
145,700/82,600; so = 60.5. The force against the
pipe acts in the opposite direction.
Fig. 21.16 Flow induces forces in a pipe at bends and at changes in size of sectionapplication of
momentum equation. (See Example 21.6.)
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.17
21.7 Water Resources
Modeling
A model is a tool that can be used to determine the
likely response of a system to a given set of stimuli
without having to actually impose those stimuli on
the system. In water resources engineering, models
are used to determine the likely response of a sys-
tem, such as a river, aquifer, or drainage basin, to a
given set of stimuli, such as storm rainfall, droughts,
alternative management schemes, or proposed
works. Models are cost-effective and convenient for
such investigations. See also Art. 1.7.
Models can typically be categorized as one of
three major types:
Physical Models
I
The system (prototype) is
modeled with physical components that represent
components of the system. Usually, scale factors
are applied to set the model at only a fraction of
the size and cost of the prototype. Physical models
are expensive to build, operate, and maintain but
are especially useful in analyzing complex phe-
nomena that are not easy or presently possible to
express mathematically.
Analog Models
I
The system (prototype) is
modeled with electronic circuits that represent
components of the system. Some conveyance and
resistance phenomena such as those found in
transmission networks and groundwater analyses
are easily modeled with analog techniques inas-
much as electric current flow and water flow
behave similarly in certain instances. Analog mod-
els are an abstraction of the prototype. Popular
before the advent of digital computers, analog
models are now infrequently used in view of the
efficiency and portability of mathematical models.
Mathematical Models
I
The system (proto-
type) is modeled with sets of mathematical expres-
sions that represent components of the system.
Mathematical models are normally programmed in
an appropriate computer language, and through
execution of the computer program, simulations of
prototype behavior are possible. Mathematical
models are limited only by the model creators abil-
ity to describe the prototype mathematically, the
capability of the computing resources, or availabili-
ty of data to support the modeling effort. They can
be as simple or as complex as a given analysis
requires and are among the most cost-effective
means to perform certain analyses.
A fourth mode of modeling, hybrid modeling,
employs both physical and mathematical models.
It exploits the advantages of these types of models
while avoiding their limitations. For instance, com-
plex three-dimensional flow patterns, erosional
scour, and sediment deposition occurring in the
immediate vicinity of a bridge pier or water control
structure can be best modeled with a physical
model while the overall water surface, momen-
tum, and velocity profile over the encompassing
river reach can be best modeled by an appropriate
mathematical model.
With hybrid models, one model often provides
input to or verification of the other model. In the
preceding example, the mathematical model
would provide depth and velocity profile input to
the physical model, and the physical model may
be able to provide a more accurate estimate of local
head loss at the pier or structure. In this way, the
two models can be executed interactively until all
common boundary conditions synchronize. The
resulting hybrid model will consist of a mathemat-
ical model that properly accounts for overall
hydraulic effects and local head loss at the pier or
structure and a physical model that properly
accounts for localized forces affecting the stability
or performance of the pier or structure.
21.7.1 Similitude for Physical Models
A physical model is a system whose operation can
be used to predict the characteristics of a similar
system, or prototype, usually more complex or
built to a much larger scale. A knowledge of the
laws governing the phenomena under investiga-
tion is necessary if the model study is to yield accu-
rate quantitative results.
Forces acting on the model should be propor-
tional to forces on the prototype. The four forces
usually considered in hydraulic models are inertia,
gravity, viscosity, and surface tension. Because of
the laws governing these forces and because the
model and prototype are normally not the same
size, it is usually not possible to have all four forces
in the model in the same proportions as they are in
the prototype. It is, however, a simple procedure to
have two predominant forces in the same propor-
tion. In most models, the fact that two of the four
forces are not in the same proportion as they are in
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.18
I
Section Twenty-One
the prototype does not introduce serious error. The
inertial force, which is always a predominant force,
and one other force are made proportional.
Ratios of the forces of gravity, viscosity, and sur-
face tension to the force of inertia are designated,
respectively, Froude number, Reynolds number,
and Weber number. Equating the Froude number
of the model and the Froude number of the proto-
type ensures that the gravitational and inertial
forces are in the same proportion. Similarly, equat-
ing the Reynolds numbers of the model and proto-
type ensures that the viscous and inertial forces
will be in the same proportion. And equating the
Weber numbers ensures proportionality of surface
tension and inertial forces.
The Froude number is
(21.16)
where F = Froude number (dimensionless)
V = velocity of fluid, ft/s
L = linear dimension (characteristic, such
as depth or diameter), ft
g = acceleration due to gravity, 32.2 ft/s
2
For hydraulic structures, such as spillways and
weirs, where there is a rapidly changing water-sur-
face profile, the two predominant forces are inertia
and gravity. Therefore, the Froude numbers of the
model and prototype are equated:
(21.17a)
where subscript m applies to the model and p to
the prototype. Squaring both sides of Eq. (21.17a)
and grouping like terms yields
(21.17b)
Let V
r
= V
m
/V
p
and L
r
= L
m
/L
p
. Then
(21.18)
The subscript r indicates ratio of quantity in model
to that in prototype.
If the ratios of all the physical dimensions of a
model to all the corresponding physical dimen-
sions of the prototype are equal to the length ratio,
the model is termed a true model. In a true model
where the Froude number is the governing design
criterion, the length ratio is the only variable. Once
the length ratio has been set, all the physical
dimensions of the model are fixed. The discharge
ratio is determined as follows:
(21.19a)
Since V
r
= L
r
1/2
and A
r
= area ratio = L
2
r
,
(21.19b)
By this method all the necessary characteristics of a
spillway or weir model can be determined.
The Reynolds number is
(21.20)
R is dimensionless, and is the kinematic viscosity
of fluid, ft
2
/s. The Reynolds numbers of model and
prototype are equated when the viscous and inertial
forces are predominant. Viscous forces are usually
predominant when flow occurs in a closed system,
such as pipe flow where there is no free surface. The
following relations are obtained by equating
Reynolds numbers of the model and prototype:
(21.21a)
(21.21b)
The variable factors that fix the design of a true
model when the Reynolds number governs are the
length ratio and the viscosity ratio.
The Weber number is
(21.22)
where = density of fluid, lbs
2
/ft
4
(specific
weight divided by g)
= surface tension of fluid, psf
The Weber numbers of model and prototype
are equated in certain types of wave studies, the
formation of drops and air bubbles, entrainment of
air in flowing water, and other phenomena where
surface tension and inertial forces are predomi-
nant. The velocity ratio is determined as follows:
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.19
(21.23a)
(21.23b)
The fluid properties and the length ratio fix the
design of a model governed by the Weber number.
In some cases, such as a morning-glory spill-
way, inertial, viscous, and gravity forces all have an
important effect on the flow. In these cases it is
usually not possible to have both the Reynolds and
Froude numbers of the model and prototype
equal. The solution to this type of problem is most-
ly empirical and may consist of an attempt to eval-
uate the effects of viscosity and gravity separately.
For the flow of water in open channels and
rivers where the friction slope is relatively flat,
model designs are often based on the Manning
equation. The relations between the model and
prototype are determined as follows:
(21.24)
where n = Manning roughness coefficient (T/L
1/3
,
T representing time)
R = hydraulic radius (L)
S = loss of head due to friction per unit
length of conduit (dimensionless)
= slope of energy gradient
For true models, S
r
= 1, R
r
= L
r
. Hence,
(21.25)
In models of rivers and channels, it is necessary for
the flow to be turbulent. The U.S. Waterways
Experiment Station has determined that flow will
be turbulent if
(21.26)
where V = mean velocity, ft/s
R = hydraulic radius, ft
= kinematic viscosity, ft
2
/s
If the model is to be a true model, it may have to be
uneconomically large for the flow to be turbulent.
Another problem also encountered in true models
is surface tension. In a true model of a wide river
where the depth may be only a fraction of an inch,
the surface tension will distort the flow to such an
extent that the model may be useless. To overcome
the effect of surface tension and to get turbulent
flow, the depth scale is often made much larger
than the length scale. This type of model is called a
distorted model.
The relations between a distorted model of a
channel and a prototype are determined in the
same manner as was Eq. (21.24). The only differ-
ence is that the slope ratio S
r
equals the depth ratio
d
r
and the hydraulic-radius ratio is a function of the
width ratio and depth ratio.
One type of model, called a movable-bed
model, is used to study erosion and transportation
of silt in riverbeds. Because the laws governing the
transportation of material are not fully understood,
movable-bed models are built largely on the basis
of experience and give only qualitative results.
21.7.2 Types and Applications of
Mathematical Models
Used in many applications of water resources engi-
neering, mathematical models are, in particular,
applied in hydrologic and hydraulic investigations
of man-made and natural systems for both surface-
water and groundwater purposes. The system
(prototype) is modeled with sets of mathematical
expressions that represent components of the sys-
tem. These expressions, in turn, are linked togeth-
er to represent the system as a whole.
Mathematical models are used for both analysis
and design. They are normally programmed in an
appropriate computer language, and through exe-
cution of the computer program, simulations of
prototype behavior are possible. They may be sin-
gle-purpose (for a specific site) or general purpose
(applicable to a variety of sites).
Single-purpose models typically represent the
specific temporal and spatial descriptions of the
prototype directly in the computer code. For
instance, the logical representation of prototypes,
such as flow networks, catchment areas, and infil-
tration parameters, may be part of the source code
and is said to be hardwired into the computer pro-
gram. For such models, the software (the computer
program code) and the application input codes
(hydrologic and hydraulic parameters) are bound
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.20
I
Section Twenty-One
into one entity. This, however, usually has more dis-
advantages than advantages, especially when mod-
ifications of the model are required or when the
model has to be applied by engineers who were not
involved in the original program coding. The pre-
ferred approach in modeling is instead to develop
general-purpose models by writing software that is
essentially independent of application input code.
General-purpose models are used for specific
analytical tasks. These may be as simple as deter-
mination of excess rainfall, given rainfall and rain-
fall-loss parameters, or as complex as long-period
simulation of flow and pollutant transport in com-
bined groundwater and surface-water systems.
Advances are continually being made in com-
puter resources and use of models is becoming
more widespread. As a result, the desirability of
more uniformity of software packages and of
object-oriented software has become apparent. In
object-oriented software, every program compo-
nent is generalized as much as feasible and the
entire program is essentially a collection of modu-
lar software components. This approach, when
fully implemented, will provide complete compat-
ibility among all types of water resources software.
Also, this approach will provide nearly complete
compatibility of all databases, of all databases and
software, and among water resources modelers in
the government, academia, and private sectors.
The result will be a reduction in duplication of the
efforts of software developers and modelers and
an increase in the efficiency of water-resources
engineering investigations.
Typical applications of mathematical models
include the following: stochastic processes; evapo-
ration and irrigation; hydrodynamics; hydrologic
forecasting; watershed hydrology; design of
hydraulic structures; reservoir regulation; flood or
drought impacts; flow routing; channel and river
hydraulics; sediment or pollutant transport; quan-
tity and quality of water supply; ecosystem
impacts and restoration; impacts of dam breaks;
wave or tidal analyses; landfill leachate analyses;
and groundwater yield, seepage, or pollution.
Several different models varying in complexity
or sophistication, or both, and in application type
may be required in many types of investigations.
As a general rule, if comparisons of different plans
are required, the fewer the number of models
employed in a given study, the greater the chance
that meaningful results will be produced. The
availability and quality of data for calibration and
verification, the model output required for design
or evaluation, and the general acceptance by the
engineering community should be considered in
selection of a model or group of models for any
investigation.
Mathematical modeling is one of the fastest
changing fields in engineering. Applications
should be upgraded accordingly if their continued
use is expected.
(D. R. Maidment, Handbook of Hydrology, D.
H. Hoggan, Computer-Assisted Floodplain
Hydrology and Hydraulics, N. S. Grigg, Water
Resources Planning, V. J. Zipparo and H. Hasen,
Davis Handbook of Applied Hydraulics,
McGraw-Hill, New York.)
Pipe Flow
The term pipe flow as used in this section refers to
flow in a circular closed conduit entirely filled with
fluid. For closed conduits other than circular, rea-
sonably good results are obtained in the turbulent
range with standard pipe-flow formulas if the
diameter is replaced by four times the hydraulic
radius. But when there is severe deviation from a
circular cross section, as in annular passages, this
method gives flows significantly underestimated.
(J. F. Walker, G. A. Whan, and R. R. Rothfus, Fluid
Friction in Noncircular Ducts, Journal of the Ameri-
can Institute of Chemical Engineers, vol. 3, 1957.)
21.8 Laminar Flow
In laminar flow, fluid particles move in parallel lay-
ers in one direction. The parabolic velocity distrib-
ution in laminar flow, shown in Fig. 21.17, creates a
shearing stress = dV/dy, where dV/dy is the rate
of change of velocity with depth and is the coef-
ficient of viscosity (see Viscosity, Art. 21.2). As this
shearing stress increases, the viscous forces
become unable to damp out disturbances, and tur-
bulent flow results. The region of change is depen-
dent on the fluids velocity, density, and viscosity
and the size of the conduit.
A dimensionless parameter called the Reynolds
number has been found to be a reliable criterion
for the determination of laminar or turbulent flow.
It is the ratio of inertial forces to viscous forces, and
is given by
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.21
(21.27)
where V = fluid velocity, ft/s
D = pipe diameter, ft
= density of fluid, lbs
2
/ft
4
(specific
weight divided by g, 32.2 ft/s
2
)
= viscosity of fluid lbs/ft
2
= / = kinematic viscosity, ft
2
/s
For a Reynolds number less than 2000, flow is lam-
inar in circular pipes. When the Reynolds number
is greater than 2000, laminar flow is unstable; a dis-
turbance will probably be magnified, causing the
flow to become turbulent.
In laminar flow, the following equation for head
loss due to friction can be developed by consider-
ing the forces acting on a cylinder of fluid in a pipe:
(21.28)
where h
f
= head loss due to friction, ft
L = length of pipe section considered, ft
g = acceleration due to gravity, 32.2 ft/s
2
w = specific weight of fluid, lb/ft
3
Substitution of the Reynolds number yields
(21.29)
For laminar flow, Eq. (21.29) is identical to the
Darcy-Weisbach formula Eq. (21.30) since in lami-
nar flow the friction f = 64/ R.
(E. F. Brater, handbook of Hydraulics, 6th ed.,
McGraw-Hill Book Company, New York.)
21.9 Turbulent Flow
In turbulent flow, the inertial forces are so great
that viscous forces cannot dampen out distur-
bances caused primarily by the surface roughness.
These disturbances create eddies, which have both
a rotational and translational velocity. The transla-
tion of these eddies is a mixing action that affects
an interchange of momentum across the cross sec-
tion of the conduit. As a result, the velocity distrib-
ution is more uniform, as shown in Fig. 21.18, than
for laminar flow (Fig. 21.17).
For a Reynolds number greater than 2000 but to
the left of the dashed line in Fig. 21.l9, there is a
transition from laminar to turbulent flow. In this
region, there is a laminar film at the boundaries that
covers some of the smaller roughness projections.
This explains why the friction loss in this region has
both laminar and turbulent characteristics. As the
Reynolds number increases, this laminar boundary
layer decreases in thickness until, at completely tur-
bulent flow, it no longer covers any of the rough-
ness projections. To the right of the dashed line in
Fig. 21.19, the flow is completely turbulent, and vis-
cous forces do not affect the friction loss.
Because of the random nature of turbulent
flow, it is not practical to treat it analytically. There-
fore, formulas for head loss and flow in the turbu-
lent regions have been developed through experi-
mental and statistical means. Experimentation in
turbulent flow has shown that:
The head loss varies directly as the length of the pipe.
The head loss varies almost as the square of the
velocity.
Fig. 21.17 Velocity distribution for lamellar
flow in a circular pipe is parabolic. Maximum veloc-
ity is twice the average velocity.
Fig. 21.18 Velocity distribution for turbulent
flow in a circular pipe is more nearly uniform than
that for lamellar flow.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.22
I
Section Twenty-One
The head loss varies almost inversely as the diameter.
The head loss depends on the surface roughness of
the pipe wall.
The head loss depends on the fluids density and
viscosity.
The head loss is independent of the pressure.
21.9.1 Darcy-Weisbach Formula
One of the most widely used equations for pipe
flow, the Darcy-Weisbach formula satisfies the
above condition and is valid for laminar or turbu-
lent flow in all fluids.
(21.30)
where h
f
= head loss due to friction, ft
f = friction factor (see Fig. 21.19)
L = length of pipe, ft
D = diameter of pipe, ft
V = velocity of fluid, ft/s
g = acceleration due to gravity, 32.2 ft/s
2
It employs the Moody diagram (Fig. 21.19) for eval-
uating the friction factor f. (L. F. Moody, Friction
Factors for Pipe Flow, Transactions of the American
Society of Mechanical Engineers, November 1944.)
Because Eq. (21.30) is dimensionally homoge-
neous, it can be used with any consistent set of units
without changing the value of the friction factor.
Fig. 21.19 Chart relates friction forces for flow in pipe to Reynolds numbers and condition of pipes.
, ft
Steel pipe:
Severe tuberculation and incrustation 0.03 0.008
General tuberculation 0.008 0.003
Heavy brush-coat asphalts, enamels,
and tars 0.003 0.001
Light rust 0.001 0.0005
New smooth pipe, centrifugally
applied enamels 0.0002 0.00003
Hot-dipped asphalt; centrifugally
applied concrete linings 0.0005 0.0002
Steel-formed concrete pipe, good
workmanship 0.0005 0.0002
New cast-iron pipe 0.00085
Table 21.3 Typical Values of Roughness for Use
in the Moody Diagram (Fig. 21.19) to Determine f
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.23
Roughness values (ft) for use with the Moody
diagram to determine the Darcy-Weisbach friction
factor f are listed in Table 21.3.
The following formulas were derived for head
loss in waterworks design and give good results
for water-transmission and -distribution calcula-
tions. They contain a factor that depends on the
surface roughness of the pipe material. The accu-
racy of these formulas is greatly affected by the
selection of the roughness factor, which requires
experience in its choice.
21.9.2 Chezy Formula
This equation holds for head loss in conduits and
gives reasonably good results for high Reynolds
numbers:
(21.31)
where V = velocity, ft/s
C = coefficient, dependent on surface
roughness of conduit
S = slope of energy grade line or head
loss due to friction, ft/ft of conduit
R = hydraulic radius, ft
Hydraulic radius of a conduit is the cross-sec-
tional area of the fluid in it divided by the perime-
ter of the wetted section.
21.9.3 Mannings Formula
Through experimentation, Manning concluded
that the C in the Chezy equation [Eq. (21.31)]
should vary as R
1/6
(21.32)
where n = coefficient, dependent on surface rough-
ness. (Although based on surface roughness, n in
practice is sometimes treated as a lumped parameter
for all head losses.) Substitution into Eq. (21.31) gives
(21.33a)
Upon substitution of D/4, where D is the pipe
diameter, for the hydraulic radius of the pipe, the fol-
lowing equations are obtained for pipes flowing full:
(21.33b)
(21.33c)
(21.33d)
(21.33e)
where Q = flow, ft
3
/s.
Tables 21.4 and 21.11 (p. 21.47) give values of n
for the foot-pound-second system. See also Table
22.3 for velocity and flow at various slopes.
21.9.4 Hazen-Williams Formula
This is one of the most widely used formulas for
pipe-flow computations of water utilities, although
it was developed for both open channels and pipe
flow:
(21.34a)
For pipes flowing full:
(21.34b)
(21.34c)
(21.34d)
(21.34e)
where V = velocity, ft/s
C
1
= coefficient, dependent on surface
roughness
R = hydraulic radius, ft
S = head loss due to friction, ft/ft of pipe
D = diameter of pipe, ft
L = length of pipe, ft
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.24
I
Section Twenty-One
Q = discharge, ft
3
/s
h
f
= friction loss, ft
The C
1
terms in Table 21.5 are in the foot-pound-
second system.
Determination of flow in branching pipes illus-
trates the use of friction-loss equations and the
hydraulic-grade-line concept.
Example 21.7: Figure 21.20 shows a typical
three-reservoir problem. The elevations of the
hydraulic grade lines for the three pipes are equal
at point D. The Hazen-Williams equation for fric-
tion loss [Eq. (21.34d)] can be written for each pipe
meeting at D. With the continuity equation for
quantity of flow, there are as many equations as
there are unknowns:
(21.35a)
(21.35b)
(21.35c)
(21.36)
where p
D
= pressure at D
w = unit weight of liquid
With the elevations Z of the three reservoirs and
the pipe intersection known, the easiest way to
solve these equations is by trying different values
of p
D
/w in Eqs. (21.35) and substituting the values
obtained for Q into Eq. (21.36) for a check. If the
value of Z
d
+ p
D
/w becomes greater than Z
b
, the
sign of the friction-loss term is negative instead of
positive. This would indicate water is flowing from
reservoir A into reservoirs B and C. Flow in pipe
network is easily determined with available com-
puter programs, many of which are specialized to
solve specific pipe design problems efficiently.
21.10 Minor Losses in Pipes
Energy losses occur in pipe contractions, bends,
enlargements, and valves and other pipe fittings.
These losses can usually be neglected if the length
of the pipeline is greater than 1500 times the pipes
diameter. However, in short pipelines, because
Variation Use in designing
Material of pipe
From To From To
Clean cast iron 0.011 0.015 0.013 0.015
Dirty or tuberculated cast iron 0.015 0.035
Riveted steel or spiral steel 0.013 0.017 0.015 0.017
Welded steel 0.010 0.013 0.012 0.013
Galvanized iron 0.012 0.017 0.015 0.017
Wood stave 0.010 0.014 0.012 0.013
Concrete 0.010 0.017
Good workmanship 0.012 0.014
Poor workmanship 0.016 0.017
Table 21.4 Values of n for Pipes, to Be Used with the Manning Formula
Fig. 21.20 Flow between reservoirs. (See Exam-
ple 21.7.)
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.25
these losses may exceed the friction losses, minor
losses must be considered.
21.10.1 Sudden Enlargements
The following equation for the head loss, ft, across
a sudden enlargement of pipe diameter has been
determined analytically and agrees well with
experimental results:
(21.37)
where V
1
= velocity before enlargement, ft/s
V
2
= velocity after enlargement, ft/s
g = 32.2 ft/s
2
It was derived by applying the Bernoulli equation
and the momentum equation across an enlargement.
Another equation for the head loss caused by
sudden enlargements was determined experimental-
ly by Archer. This equation gives slightly better agree-
ment with experimental results than Eq. (21.37):
(21.38)
A special application of Eq. (21.37) or (21.38) is the dis-
charge from a pipe into a reservoir. The water in the
reservoir has no velocity, so a full velocity head is lost.
21.10.2 Gradual Enlargements
The equation for the head loss due to a gradual con-
ical enlargement of a pipe takes the following form:
(21.39)
where K = loss coefficient (see Fig. 21.21).
Since the experimental data available on grad-
ual enlargements are limited and inconclusive, the
values of K in Fig. 21.21 are approximate. (A. H.
Gibson, Hydraulics and Its Applications, Consta-
ble & Co., Ltd., London.)
21.10.3 Sudden Contraction
The following equation for the head loss across a
sudden contraction of a pipe was determined by
the same type of analytical studies as Eq. (21.37):
(21.40)
where C
c
= coefficient of contraction (see Table
21.6)
V = velocity in smaller-diameter pipe, ft/s
This equation gives best results when the head loss
is greater than 1 ft. Table 21.6 gives C
c
values for
sudden contractions, determined by Julius Weis-
bach (Die Experiments-Hydraulik).
Another formula for determining the loss of
head caused by a sudden contraction, determined
experimentally by Brightmore, is
(21.41)
This equation gives best results if the head loss is
less than 1 ft.
A special case of sudden contraction is the
entrance loss for pipes. Some typical values of
the loss coefficient K in h
L
=KV
2
/ 2g, where V is
the velocity in the pipe, are presented in Table 21.7.
Type of pipe C
1
Cast iron:
New All sizes, 130
5 years old All sizes up to 24 in, 120
24 in and over, 115
10 years old 12 in, 110
4 in, 105
30 in and over, 85
40 years old 16 in, 80
4 in, 65
Welded steel Values the same as for cast-iron
pipe, 5 years older
Riveted steel Values the same as for cast-iron
pipe, 10 years older
Wood stave Average value, regardless of
age, 120
Concrete or Large sizes, good workmanship,
concrete-lined steel forms, 140
Large sizes, good workmanship,
wood forms, 120
Centrifugally spun, 135
Vitrified In good condition, 110
Table 21.5 Values of C
1
in Hazen and Williams
Formula
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.26
I
Section Twenty-One
21.10.4 Bends and Standard Fitting
Losses
The head loss that occurs in pipe fittings, such as
valves and elbows, and at bends is given by
(21.42)
Table 21.8 gives some typical K values for these
losses.
The values in Table 21.8 are only approximate.
K values vary not only for different sizes of fitting
but with different manufacturers. For these rea-
Fig. 21.21 Head-loss coefficients for a pipe with diverging sides depend on the angle of divergence
of the sides.
A
2
/A
1
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
C
c
0.62 0.63 0.64 0.66 0.68 0.71 0.76 0.81 0.89 1.0
Table 21.6 C
c
for Contractions in Pipe Area from A
1
to A
2
Pipe projecting into reservoir K = 0.80
Sharp-cornered entrance K = 0.50
Bellmouth entrance K = 0.05
Slightly rounded entrance K = 0.25
Table 21.7 Coefficients for Entrance Losses
Fitting K
Globe valve, fully open 10.0
Angle valve, fully open 5.0
Swing check valve, fully open 2.5
Gate valve, fully open 0.2
Closed-return bend 2.2
Short-radius elbow (r/ D 1.0)* 0.9
Long-radius elbow (r/ D 1.5) 0.6
45 elbow 0.4
*r = radius of bend; D = pipe diameter.
Table 21.8 Coefficients for Fitting Losses and
Losses at Bends
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.27
sons, manufacturers data are the best source for
loss coefficients.
Experimental data available on bend losses
cover a rather narrow range of laboratory experi-
ments utilizing small-diameter pipes and do not
give conclusive results. The data indicate the loss-
es vary with surface roughness, Reynolds number,
ratio of radius of bend r to pipe diameter D, and
angle of bend. The data are in agreement that the
head loss, not including friction loss, decreases
sharply as the r/D ratio increases from zero to
around 4 or 5. When r/D increases above 4 or 5,
there is disagreement. Some experiments indicate
that the head loss, not including friction loss in the
bend, increases significantly with an increasing
r/D. Experiments on smooth pipes, indicate that
this increase is very slight and that above an r/D of
4, the bend loss essentially remains constant. (H.
Ito, Pressure Losses in Smooth Pipe Bends, Trans-
actions of the American Society of Civil Engineers,
series D, vol. 82, no. 1, 1960.)
Because experiments have produced such
widely varying data, bend-loss coefficients give
only an approximation of losses to be expected.
Figure 21.22 gives values of K for 90 bends for use
with Eq. (21.42). (K. H. Beij, Pressure Losses for
Fluid Flow in 90 Pipe Bends, Journal of Research,
National Bureau of Standards, vol. 21, July 1938.)
To obtain losses in bends other than 90, the fol-
lowing formula may be used to adjust the K values
given in Fig. 21.22:
(21.43)
where = deflection angle, deg
The K value may be used in place of K in Eq.
(21.42).
Minor losses are often given as the equivalent
length of pipe that has the same energy loss for the
same discharge. (V. J. Zipparo and H. Hasen,
Davis Handbook of Applied Hydraulics, 4th ed.,
McGraw-Hill, Inc., New York.)
21.11 Orifices
An orifice is an opening with a closed perimeter
through which water flows. Orifices may have any
shape, although they are usually round, square, or
rectangular.
21.11.1 Orifice Discharge into
Free Air
Discharge through a sharp-edged orifice may be
calculated from
(21.44)
where Q = discharge, ft
3
/s
C = coefficient of discharge
a = area of orifice, ft
2
g = acceleration due to gravity, ft/s
2
h = head on horizontal center line of ori-
fice, ft
Coefficients of discharge C are given in Table
21.9 for low velocity of approach. If this velocity is
significant, its effect should be taken into account.
Equation (21.44) is applicable for any head for
which the coefficient of discharge is known. For
low heads, measuring the head from the center
line of the orifice is not theoretically correct; how-
ever, this error is corrected by the C values.
The coefficient of discharge C is the product of
the coefficient of velocity C

and the coefficient of


contraction C
c
. The coefficient of velocity is the
ratio obtained by dividing the actual velocity at the
vena contracta (contraction of the jet discharged)
by the theoretical velocity. The theoretical velocity
may be calculated by writing Bernoullis equation
for points 1 and 2 in Fig. 21.23.
(21.45)
Fig. 21.22 Recommended values of head-loss
coefficients K for 90 bends in closed conduits.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.28
I
Section Twenty-One
With the reference plane through point 2, Z
1
= h,
V
1
= 0, p
1
/w = p
2
/w = 0, and Z
2
= 0, and Eq. (21.45)
becomes
(21.46)
The actual velocity, determined experimentally, is
less than the theoretical velocity because of the
energy loss from point 1 to point 2. Typical values
of C

range from 0.94 to 0.99.


The coefficient of contraction C
c
is the ratio of
the smallest area of the jet, the vena contracta, to
Dia. of circular orifices, ft Side of square orifices, ft
0.02 0.04 0.1 1.0 0.02 0.04 0.1 1.0
0.637 0.618 0.4 0.643 0.621
0.655 0.630 0.613 0.6 0.660 0.636 0.617
0.648 0.626 0 610 0.590 0 8 0.652 0.631 0.615 0.597
0.644 0.623 0.608 0.591 1 0.648 0.628 0.613 0.599
0.637 0.618 0.605 0.593 1.5 0.641 0.622 0.610 0.601
0.632 0.614 0.604 0.595 2 0.637 0.619 0.608 0.602
0.629 0.612 0.603 0.596 2.5 0.634 0.617 0.607 0.602
0.627 0.611 0.603 0.597 3 0.632 0.616 0.607 0.603
0.623 0.609 0.602 0.596 4 0.628 0.614 0.606 0.602
0.618 0.607 0.600 0.596 6 0.623 0.612 0.605 0.602
0.614 0.605 0.600 0.596 8 0.619 0.610 0.605 0.602
0.611 0.603 0.598 0.595 10 0.616 0.608 0.604 0.601
0.601 0.599 0.596 0.594 20 0.606 0.604 0.602 0.600
0.596 0.595 0.594 0.593 50 0.602 0.601 0.600 0.599
0.593 0.592 0.592 0.592 100 0.599 0.598 0.598 0.598
*Hamilton Smith, Jr., Hydraulics, 1886.
Table 21.9 Smiths Coefficients of Discharge for Circular and Square Orifices with Full Contraction*
Head,
ft
Fig. 21.23 Fluid jet takes a parabolic path.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.29
the area of the orifice. Contraction of a fluid jet will
occur if the orifice is square-edged and so located
that some of the fluid approaches the orifice at an
angle to the direction of flow through the orifice.
This fluid has a momentum component perpen-
dicular to the axis of the jet which causes the jet to
contract. Typical values of the coefficient of con-
traction range from 0.61 to 0.67.
If the water entering the orifice does not have
this momentum, the contraction is completely sup-
pressed. Figure 21.24a is an example of a partly
suppressed contraction; no contraction occurs at
the bottom of the jet. In Fig. 21.24b, the edges of the
orifice have been rounded to reduce or eliminate
the contraction. With a partly suppressed orifice,
the increased area of jet caused by suppressing the
contraction on one side is partly offset because
more water at a higher velocity enters on the other
sides. The result is a slightly greater coefficient of
contraction.
21.11.2 Submerged Orifices
Flow through a submerged orifice may be comput-
ed by applying Bernoullis equation to points 1 and
2 in Fig. 21.25.
(21.47)
where h
L
= losses in head, ft, between 1 and 2.
Assuming V
1
0, setting h
1
h
2
= h, and using
a coefficient of discharge C to account for losses,
Eq. (21.48) is obtained.
(21.48)
Values of C for submerged orifices do not differ
greatly from those for nonsubmerged orifices. (For
table of values of coefficients of discharge for sub-
merged orifices, see E. F. Brater, Handbook of
Hydraulics, 6th ed., McGraw-Hill Book Company,
New York.)
21.11.3 Discharge under Falling
Head
The flow from a reservoir or vessel when the
inflow is less than the outflow represents a condi-
tion of falling head. The time required for a certain
quantity of water to flow from a reservoir can be
calculated by equating the volume of water that
flows through the orifice or pipe in time dt to the
Fig. 21.24 Types of orifices: (a) Sharp-edged with partly suppressed contraction. (b) Round-edged with
no contraction.
Fig. 21.25 Discharge through a submerged
orifice.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.30
I
Section Twenty-One
volume decrease in the reservoir (Fig. 21.26):
(21.49)
Solving for dt yields
(21.50)
where a = area of orifice, ft
2
A = area of reservoir, ft
2
y = head on orifice at time t, ft
C = coefficient of discharge
g = acceleration due to gravity, 32.2 ft/s
2
Expressing the area as a function of y[A = F(y)] and
summing from time zero, when y = h
1
, to time t,
when y = h
2
, Eq. (21.50) becomes
(21.51)
If the area of the reservoir is constant as y varies,
Eq. (21.51) upon integration becomes
(21.52)
where h
1
= head at the start, ft
h
2
= head at the end, ft
t = time interval for head to fall from h
1
to h
2
, s
21.11.4 Fluid Jets
Where the effect of air resistance is small, a fluid
discharged through an orifice into the air will fol-
low the path of a projectile. The initial velocity of
the jet is
(21.53)
where h = head on center line of orifice, ft
C

= coefficient of velocity
The direction of the initial velocity depends on
the orientation of the surface in which the orifice is
located. For simplicity, the following equations
were determined assuming the orifice is located in
a vertical surface (Fig. 21.23). The velocity of the jet
in the X direction (horizontal) remains constant.
(21.54)
The velocity in the Y direction is initially zero and
thereafter a function of time and the acceleration
of gravity:
(21.55)
The X coordinate at time t is
(21.56)
The Y coordinate is
(21.57)
where V
avg
= average velocity over period of time
t. The equation for the path of the jet [Eq. (21.58)],
obtained by solving Eq. (21.57) for t and substitut-
ing in Eq. (21.56), is that for a parabola:
(21.58)
Equation (21.58) can be used to determine C

experimentally. Rearranging Eq. (21.58) gives


(21.59)
The X and Y coordinates can be measured in a lab-
oratory and C

calculated from Eq. (21.59).


Fig. 21.26 Discharge from a reservoir with
dropping water level.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.31
21.11.5 Orifice Discharge into
Short Tubes
When water flows from a reservoir into a pipe or
tube with a sharp leading edge, the same type of
contraction occurs as for a sharp-edged orifice. In
the tube or pipe, however, the water contracts and
then expands to fill the tube. If the tube is dis-
charging at atmospheric pressure, a partial vacu-
um is created at the contraction, as can be seen by
applying the Bernoulli equation across points 1
and 2 in Fig. 21.27. This reduced pressure causes
the flow through a short tube to be greater than
that through a sharp-edged orifice of the same
dimensions. If the head on the tube is greater than
50 ft and the tube is short, the water will shoot
through the tube without filling it. When this hap-
pens, the tube acts as a sharp-edged orifice.
For a short tube flowing full, the coefficient of
contraction C
c
= 1.00 and the coefficient of veloci-
ty C

= 0.82. Therefore, the coefficient of discharge


C = 0.82. Solving for head loss as a proportion of
final velocity head, a K value for Eq. (21.42) of 0.5 is
obtained as follows: The theoretical velocity head
with no loss is V
2
T
/ 2g. Actual velocity head is V
2
a
/2g
= (0.82 V
T
)
2
/2g = 0.67 V
2
T
/2g. The head loss h
L
=
1.00 V
2
T
/ 2g 0.67V
2
T
/ 2g = 0.33V
2
T
/ 2g. From h
L
=
KV
2
a
/ 2g, where V
2
a
/ 2g is the actual velocity head,
K = 2gh
L
/V
2
a
= (0.33 V
2
T
2g)/(2g 0.67 V
2
T
) = 0.5
For a reentrant tube projecting into a reservoir
(Fig. 21.28), the coefficients of velocity and discharge
equal 0.75, and the loss coefficient K equals 0.80.
21.11.6 Orifice Discharge into
Diverging Conical Tubes
This type of tube can greatly increase the flow
through an orifice by reducing the pressure at the
orifice below atmospheric. Equation (21.60) for the
pressure at the entrance to the tube is obtained by
writing the Bernoulli equation for points 1 and 3
and points 1 and 2 in Fig. 21.29.
(21.60)
where p
2
= gage pressure at tube entrance, psf
w = unit weight of water, lb/ft
3
h = head on center line of orifice, ft
a
2
= area of smallest part of jet (vena con-
tracta, if one exists), ft
2
a
3
= area of discharge end of tube, ft
2
Fig. 21.27 Flow from a reservoir through a
tube with a sharp-edged inlet.
Fig. 21.28 Flow from a reservoir through a
reentrant tube resembles that through a flush tube
(Fig. 21.27) but the head loss is larger.
Fig. 21.29 Diverging conical tube increases
flow from a reservoir through an orifice by reduc-
ing the pressure below atmospheric.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.32
I
Section Twenty-One
Discharge is also calculated by writing the
Bernoulli equation for points 1 and 3 in Fig. 21.29.
For this analysis to be valid, the tube must flow
full, and the pressure in the throat of the tube must
not fall to the vapor pressure of water. Experiments
by Venturi show the most efficient angle to be
around 5.
21.12 Siphons
A siphon is a closed conduit that rises above the
hydraulic grade line and in which the pressure at
some point is below atmospheric (Fig. 21.30). The
most common use of a siphon is the siphon spillway.
Flow through a siphon can be calculated by
writing the Bernoulli equation for the entrance
and exit. But the pressure in the siphon must be
checked to be sure it does not fall to the vapor pres-
sure of water. This is accomplished by writing the
Bernoulli equation across a point of known pres-
sure and a point where the elevation head or the
velocity head is a maximum in the conduit. If the
pressure were to fall to the vapor pressure, vapor-
ization would decrease or totally stop the flow.
The pipe shown in Fig. 21.31 is also commonly
called a siphon or inverted siphon. This is a mis-
nomer since the pressure at all points in the pipe is
above atmospheric. The American Society of Civil
Engineers recommends that the inverted siphon
be called a sag pipe to avoid the false impression
that it acts as a siphon.
21.13 Water Hammer
Water hammer is a change in pressure, either above
or below the normal pressure, caused by a variation
of the flow rate in a pipe. Every time the flow rate is
changed, either increased or decreased, it causes
water hammer. However, the stresses are not critical
in small-diameter pipes with flows at low velocities.
The water flowing in a pipe has momentum
equal to the mass of the water times its velocity.
When a valve is closed, this momentum drops to
zero. The change causes a pressure rise, which
begins at the valve and is transmitted up the pipe.
The pressure at the valve will rise until it is high
enough to overcome the momentum of the water
and bring the water to a stop. This pressure
buildup travels the full length of the pipe to the
reservoir (Fig. 21.32).
At the instant the pressure wave reaches the
reservoir, the water in the pipe is motionless, but at
a pressure much higher than normal. The differen-
tial pressure between the pipe and the reservoir
then causes the water in the pipe to rush back into
the reservoir. As the water flows into the reservoir,
the pressure in the pipe falls.
At the instant the pressure at the valve reaches
normal, the water has attained considerable momen-
tum up the pipe. As the water flows away from the
closed valve, the pressure at the valve drops until dif-
ferential pressure again brings the water to a stop.
Fig. 21.30 Siphon between reservoirs rises
above hydraulic grade line yet permits flow of
water between them.
Fig. 21.31 Sag pipe permits flow between two reservoirs despite a dip and a rise.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.33
This pressure drop begins at the valve and continues
up the pipe until it reaches the reservoir.
The pressure in the pipe is now below normal,
so water from the reservoir rushes into the pipe.
This cycle repeats over and over until friction
damps these oscillations. Because of the high
velocity of the pressure waves, each cycle may take
only a fraction of a second.
The equation for the velocity of a wave in a
pipe is
(21.61)
where U = velocity of pressure wave along pipe,
ft/s
E = modulus of elasticity of water, 43.2
10
6
psf
= density of water, 1.94 lbs/ft
4
(specific
weight divided by acceleration due to
gravity)
D = diameter of pipe, ft
E
p
= modulus of elasticity of pipe material,
psf
t = thickness of pipe wall, ft
21.13.1 Instantaneous Closure
The magnitude of the pressure change that results
when flow is varied depends on the rate of change
of flow and the length of the pipeline. Any gradual
movement of a valve that is made in less time than
it takes for a pressure wave to travel from the valve
to the reservoir and be reflected back to the valve
produces the same pressure change as an instanta-
neous movement. For instantaneous closure:
(21.62)
where L = length of pipe from reservoir to valve,
ft
T = time required to change setting of
valve, s
A plot of pressure vs. time for various points
along a pipe is shown in Fig. 21.32 for the instanta-
neous closure of a valve. Equation (21.63a) for the
pressure rise or fall caused by adjusting a valve
was derived by equating the momentum of the
water in the pipe to the force impulse required to
bring the water to a stop.
(21.63a)
In terms of pressure head, Eq. (21.63a) becomes
(21.63b)
where p = pressure change from normal due to
instantaneous change of valve set-
ting, psf
h = head change from normal due to
instantaneous change of valve set-
ting, ft
V= change in the velocity of water
caused by adjusting valve, ft/s
If the closing or opening of a valve is instanta-
neous, the pressure change can be calculated in
one step from Eq. (21.63).
21.13.2 Gradual Closure
The following method of determining the pressure
change due to gradual closure of a valve gives a
quick, approximate solution. The pressure rise or
head change is assumed to be in direct proportion
to the closure time:
(21.64)
Fig. 21.32 Variation with time of pressure at
three points in a penstock, for water hammer from
instantaneous closure of a valve.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.34
I
Section Twenty-One
where h
g
= head change due to gradual closure, ft
t
i
= time for wave to travel from the
valve to the reservoir and be reflect-
ed back to valve, s
T = actual closure time of valve, s
h = head rise due to instantaneous clo-
sure, ft
L = length of pipeline, ft
V = change in velocity of water due to
instantaneous closure, ft/s
g = acceleration due to gravity, 32.2 ft/s
2
Arithmetic integration is a more exact method
for finding the pressure change due to gradual
movement of a valve. The calculations can be read-
ily programmed for a computer and are available
in software packages. Integration is a direct means
of studying every physical element of the process
of water hammer. The valve is assumed to close in
a series of small movements, each causing an indi-
vidual pressure wave. The magnitude of these
pressure waves is given by Eq. (21.63). The indi-
vidual pressure waves are totaled to give the pres-
sure at any desired point for a certain time.
The first step in this method is to choose the
time interval for each incremental movement of
the valve. (It is convenient to make the time inter-
val some submultiple of L/U, such as L/aU, where a
equals any integer, so that the pressure waves
reflected at the reservoir will be superimposed
upon the new waves being formed at the valve.
The wave formed at the valve will be opposite in
sign to the water reflected from the reservoir, so
there will be a tendency for the waves to cancel
out.) Assuming a valve is fully open and requires T
seconds for closing, the number of incremental
closing movements required is T/t, where t, the
increment of time, equals L/aU.
Once the time interval has been determined, an
estimate of the velocity change Vduring each time
interval must be made, to apply Eq. (21.63). A rough
estimate for the velocity following the incremental
change is V
n
= V
o
(A
n
/A
o
), where V
n
is the velocity
following a certain incremental movement, V
o
the
original velocity, A
n
the area of the valve opening
after the corresponding incremental movement,
and A
o
the original area of the valve opening.
The change in head can now be calculated
with Eq. (21.63). With the head known, the esti-
mated velocity V
n
can be checked by the following
equation:
(21.65)
where H
o
= head at valve before any movement
of valve, ft
H
o
+ h = total pressure at valve after particu-
lar movement; this includes pres-
sure change caused by valve move-
ment plus effect of waves reflected
from reservoir, ft
A
n
= area of valve opening after n incre-
mental closings; this area can be
determined from closure character-
istics of valve or by assuming its
characteristics, ft
2
If the velocity obtained from Eq. (21.65) differs
greatly from the estimated velocity, then that
obtained from Eq. (21.65) should be used to recal-
culate h.
(V. J. Zipparo and H. Hasen, Davis Handbook
of Applied Hydraulics, 4th ed., McGraw-Hill, Inc.,
New York.)
Example 21.8: The following problem illustrates
the use of the preceding methods and compares
the results: Steel penstock, length = 3000 ft, diam-
eter = 10 ft, area = 78.5 ft
2
, initial velocity = 10 ft/s,
penstock thickness = 1 in, head at turbine with
valve open = 1000 ft, and modulus of elasticity of
steel = 43.2 10
8
psf.
(For penstocks as shown in Fig. 21.32, thickness
and diameter normally vary with head. Thus, the
velocity of the pressure waves is different in each
section of the penstock. Separate calculations for
the velocity of the pressure wave should be made
for each thickness and diameter of penstock to
obtain the time required for a wave to travel to the
reservoir and back to the valve.)
Velocity of pressure wave, from Eq. (21.61), is
= 3180 ft/s
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.35
The time required for the wave to travel to the
reservoir and be reflected back to the valve = 2L/U
= 6000/3180 = 1.90 s.
If closure time T of the valve is less than 1.90 s,
the closure is instantaneous, and the pressure rise,
from Eq. (21.63), is
Assuming T = 4.75 s, approximate equation
(21.64) gives the following result:
21.13.3 Surge Tanks
It is uneconomical to design long pipelines for
pressures created by water hammer or to operate a
valve slowly enough to reduce these pressures.
Usually, to prevent water hammer, a surge tank is
installed close to valves at the end of long conduits.
A surge tank is a tank containing water and con-
nected to the conduit. The water column, in effect,
floats on the line.
When a valve is suddenly closed, the water in
the line rushes into the surge tank. The water level
in the tank rises until the increased pressure in the
surge tank overcomes the momentum of the water.
When a valve is suddenly opened, the surge tank
supplies water to the line when the pressure
drops. The section of pipe between the surge tank
and the valve (Fig. 21.33) must still be designed for
water hammer; but the closure time to reduce the
pressures for this section will be only a fraction of
the time required without the surge tank.
Although a surge tank is one of the most com-
monly used devices to prevent water hammer, it is
by no means the only one. Various types of relief
valves and air chambers are widely used on small-
diameter lines, where the pressure of water ham-
mer may be relieved by the release of a relatively
small quantity of water.
Pipe Stresses
21.14 Pipe Stresses
Perpendicular to the
Longitudinal Axis
The stresses acting perpendicular to the longitudi-
nal axis of a pipe are caused by either internal or
external pressures on the pipe walls.
Internal pressure creates a stress commonly
called hoop tension. It may be calculated by taking a
free-body diagram of a 1-in-long strip of pipe cut by
a vertical plane through the longitudinal axis (Fig.
21.34). The forces in the vertical direction cancel out.
The sum of the forces in the horizontal direction is
(21.66)
where p = internal pressure, psi
D = outside diameter of pipe, in
F = force acting on each cut of edge of
pipe, lb
Hence, the stress, psi, on the pipe material is
(21.67)
where A = area of cut edge of pipe, ft
2
t = thickness of pipe wall, in
Fig. 21.33 Surge tank is placed near a valve on
a penstock to prevent water hammer.
Fig. 21.34 Internal pipe pressure produces hoop
tension.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.36
I
Section Twenty-One
From the derivation of Eq. (21.67), it would
appear that the diameter used for calculations
should be the inside diameter. However, Eq. (21.67)
is not theoretically exact and gives stresses slightly
lower than those actually developed. For this reason
the outside diameter often is used (see also Art. 6.10).
Equation (21.67) is exact for all practical purpos-
es when D/t is equal to or greater than 50. If D/t is
less than 10, this equation will usually be quite con-
servative and therefore will yield an uneconomical
design. For steel pipes, Eq. (21.67) gives directly the
thickness required to resist internal pressure.
For concrete pipes, this analysis is approximate,
however, since concrete cannot resist large tensile
stresses. The force F must be carried by steel rein-
forcing. The internal diameter is used in Eq. (21.67)
for concrete pipe.
When a pipe has external pressure acting on it,
the analysis is much more complex because the
pipe material no longer acts in direct tension. The
external pressure creates bending and compressive
stresses that cause buckling.
(S. P. Timoshenko and J. M. Gere, Theory of
elastic Stability, 2nd ed., McGraw-Hill Book Com-
pany, New York.)
21.15 Pipe Stresses Parallel
to the Longitudinal Axis
If a pipe is supported on piers, it acts like a beam. The
stresses created can be calculated from the bending
moment and shear equations for a continuous circu-
lar hollow beam. This stress is usually not critical in
high-head pipes. However, thin-walled pipes usual-
ly require stiffening to prevent buckling and exces-
sive deflection from the concentrated loads.
21.16 Temperature Expansion
of Pipe
If a pipe is subject to a wide range of temperatures,
the pipe should be the stress due to temperature
variation designed for or expansion joints should
be provided. The stress, psi, due to a temperature
change is
(21.68)
where E = modulus of elasticity of pipe material,
psi
T = temperature change from installation
temperature
c = coefficient of thermal expansion of
pipe material
The movement that should be allowed for, if
expansion joints are to be used, is
(21.69)
where L = movement in length L of pipe
L = length between expansion joints
21.17 Forces Due to Pipe
Bends
It is common practice to use thrust blocks in pipe
bends to take the forces on the pipe caused by the
momentum change and the unbalanced internal
pressure of the water.
In all bends, there will be a slight loss of head
due to turbulence and friction. This loss will cause
a pressure change across the bend, but it is usually
small enough to be neglected. When there is a
change in the cross-sectional area of the pipe, there
will be an additional pressure change that can be
calculated with the Bernoulli equation (see Exam-
ple 6, Art. 21.6). In this case, the pressure differen-
tial may be large and must be considered.
The force diagram in Fig. 21.35 is a convenient
method for finding the resultant force on a bend.
The forces can be resolved into X and Y compo-
nents to find the magnitude and direction of the
resultant force on the pipe. In Fig. 21.35:
V
1
= velocity before change in size of
pipe, ft/s
V
2
= velocity after change in size of pipe,
ft/s
p
1
= pressure before bend or size change
in pipe, psf
p
2
= pressure after bend or size change in
pipe, psf
A
1
= area before size change in pipe, ft
2
A
2
= area after size change in pipe, ft
2
F
2m
= force due to momentum of water in
section 2 = V
2
Qw/g
F
1m
= force due to momentum of water in
section 1 = V
1
Qw/g
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.37
P
2
= pressure of water in section 2 times
area of section 2 = p
2
A
2
P
1
= pressure of water in section 1 times
area of section 1 = p
1
A
1
w = unit weight of liquid, lb/ft
3
Q = discharge, ft
3
/s
If the pressure loss in the bend is neglected and there
is no change in magnitude of velocity around the
bend, Eqs. (21.70) and (21.71) give a quick solution.
(21.70)
(21.71)
where R = resultant force on bend, lb
= angle R makes with F
1m
p = pressure, psf
w = unit weight of water, 62.4 lb/ft
3
V = velocity of flow, ft/s
g = acceleration due to gravity, 32.2 ft/s
2
A = area of pipe, ft
2
= angle between pipes (0 < < 180)
Although thrust blocks are normally used to
take the force on bends, in many cases the pipe
material takes this force. The stress caused by this
force is directly additive to other stresses along
the longitudinal axis of the pipe. In small pipes,
the force caused by bends can easily be carried by
the pipe material; however, the joints must also
be able to take these forces.
Culverts
A culvert is a closed conduit for the passage of sur-
face drainage under a highway, a railroad, canal, or
other embankment. The slope of a culvert and its
inlet and outlet conditions are usually determined
by the topography of the site. Because of the many
combinations obtained by varying the entrance con-
ditions, exit conditions, and slope, no single formula
can be given that will apply to all culvert problems.
The basic method for determining discharge
through a culvert requires application of the
Bernoulli equation between a point just outside
the entrance and a point somewhere downstream.
An understanding of uniform and nonuniform
flow is necessary to understand culvert flow fully.
However, an exact theoretical analysis, involving
detailed calculation of drawdown and backwater
curves, is usually unwarranted because of the rela-
Fig. 21.35 Forces produced by flow at a pipe bend and change in diameter.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.38
I
Section Twenty-One
tively low accuracy attainable in determining
runoff. Neglecting drawdown and backwater
curves does not seriously affect the accuracy but
greatly simplifies the calculations.
21.18 Culverts on Critical
Slopes or Steeper
In a culvert with a critical slope, the normal depth
(Art. 21.22) is equal to the critical depth (Art. 21.23).
Entrance Submerged or Unsubmerged
but Free Exit
I
If a culvert is on critical slope or
steeper, that is, the normal depth is equal to or less
than the critical depth, the discharge will be entire-
ly dependent on the entrance conditions (Fig.
21.36). Increasing the slope of the culvert past crit-
ical slope (the slope just sufficient to maintain flow
at critical depth) will decrease the depth of flow
downstream from the entrance. But the increased
slope will not increase the amount of water enter-
ing the culvert because the entrance depth will
remain at critical.
The discharge is given by the equation for flow
through an orifice if the entrance is submerged, or
by the equation for flow over a weir if the entrance
is not submerged. Coefficients of discharge for weirs
and orifices give good results, but they do not cover
the entire range of entry conditions encountered in
culvert problems. For this reason, computer soft-
ware, charts, and nomographs have been devel-
oped and are used almost exclusively in design.
(Handbook of Concrete Culvert Pipe Hydraulics,
EB058W, Portland Cement Association.)
Entrance Unsubmerged but Exit Sub-
merged
I
In this case, the submergence of the exit
will cause a hydraulic jump to occur in the culvert
(Fig. 21.37). The jump will not affect the culvert dis-
charge, and the control will still be at the inlet.
Entrance and Exit Submerged
I
When
both the exit and entrance are submerged (Fig.
21.38), the culvert flows full, and the discharge is
independent of the slope. This is normal pipe flow
and is easily solved by using the Manning or
Darcy-Weisbach formula for friction loss [Eq.
(21.33d) or (21.30)]. From the Bernoulli equation for
the entrance and exit, and the Manning equation
for friction loss, the following equation is obtained:
(21.72)
Solution for the velocity of flow yields
(21.73)
Fig. 21.36 Flow through a culvert with free discharge. Normal depth d
n
is less than critical depth d
c
;
slope is greater than the critical slope. Discharge depends on the type of inlet and the head H.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.39
where H = elevation difference between head-
water and tailwater, ft
V = velocity in culvert, ft/s
g = acceleration due to gravity, 32.2 ft/s
2
K
e
= entrance-loss coefficient (Art. 21.20)
n = Manning s roughness coefficient
L = length of culvert, ft
R = hydraulic radius of culvert, ft
Equation (21.72) can be solved directly since the
velocity is the only unknown.
21.19 Culverts on Subcritical
Slopes
Critical slope is the slope just sufficient to maintain
flow at critical depth. When the slope is less than
critical, the flow is considered subcritical (Art. 21.23).
Entrance Submerged or Unsubmerged
but Free Exit
I
For these conditions, depending
on the head, the flow can be either pressure or
open-channel.
The discharge, for the open-channel condition
(Fig. 21.39), is obtained by writing the Bernoulli
equation for a point just outside the entrance and
a point a short distance downstream from the
entrance. Thus,
(21.74)
The velocity can be determined from the Man-
ning equation:
(21.75)
Substituting this into Eq. (21.74) yields
(21.76)
where H = head on entrance measured from bot-
tom of culvert, ft
K
e
= entrance-loss coefficient (Art. 21.20)
Fig. 21.37 Flow through a culvert with entrance
unsubmerged but exit submerged. When slope is
less than critical, open-channel flow takes place, and
d
n
> d
c
. When slope exceeds critical, flow depends
on inlet condition, and d
n
< d
c
.
Fig. 21.38 With entrance and exit of a culvert
submerged, normal pipe flow occurs. Discharge is
independent of slope. The fluid flows under pres-
sure. Discharge may be determined from Bernoulli
and Manning equations.
Fig. 21.39 Open-channel flow occurs in a cul-
vert with free discharge and normal depth d
n
greater than the critical depth d
c
when the entrance
is unsubmerged or slightly submerged. Discharge
depends on head H, loss at entrance, and slope of
culvert.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.40
I
Section Twenty-One
S = slope of energy grade line, which for
culverts is assumed to equal slope of
bottom of culvert
R = hydraulic radius of culvert, ft
d
n
= normal depth of flow, ft
To solve Eq. (21.76), it is necessary to try differ-
ent values of d
n
and corresponding values of R
until a value is found that satisfies the equation. If
the head on a culvert is high, a value of d
n
less than
the culvert diameter will not satisfy Eq. (21.76).
This means the flow is under pressure (Fig. 21.40),
and discharge is given by Eq. (21.72).
When the depth of the water is slightly below
the top of the culvert, there is a range of unstable
flow fluctuating between pressure and open chan-
nel. If this condition exists, it is good practice to
check the discharge for both pressure flow and
open-channel flow. The condition that gives the
lesser discharge should be assumed to exist.
Short Culvert with Free Exit
I
When a
culvert on a slope less than critical has a free exit,
there will be a drawdown of the water surface at
the exit and for some distance upstream. The mag-
nitude of the drawdown depends on the friction
slope of the culvert and the difference between the
critical and normal depths. If the friction slope
approaches critical, the difference between normal
depth and critical depth is small (Fig. 21.39), and
the drawdown will not extend for any significant
distance upstream. When the friction slope is flat,
there will be a large difference between normal
and critical depth. The effect of the drawdown will
extend a greater distance upstream and may reach
the entrance of a short culvert (Fig. 21.41). This
drawdown of the water level in the entrance of the
culvert will increase the discharge, causing it to be
about the same as for a culvert on a slope steeper
than critical (Art. 21.18). Most culverts, however,
are on too steep a slope for the backwater to have
any effect for an appreciable distance upstream.
Entrance Unsubmerged but Exit Sub-
merged
I
If the level of submergence of the exit is
well below the bottom of the entrance (Fig. 21.37),
the backwater from the submergence will not
extend to the entrance. The discharge for this case
will be given by Eq. (21.76).
If the level of submergence of the exit is close to
the level of the entrance, it may be assumed that
the backwater will cause the culvert to flow full and
Fig. 21.40 Culvert with free discharge and normal depth d
n
greater than critical depth d
c
flows full when
the entrance is deeply submerged. Discharge is given by equations for pipe flow.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.41
a pipe flow condition will result. The discharge for
this case is given by Eqs. (21.72) and (21.73).
When the level of submergence falls between
these two cases and the project does not warrant a
trial approach with backwater curves, it is good
practice to assume the condition that gives the less-
er discharge.
21.20 Entrance Losses for
Culverts
Flow in a culvert may be significantly affected by loss
in head because of conditions at the entrance (Arts.
21.18 and 21.19). Table 21.10 lists coefficients of
entrance loss K
e
for some typical entrance conditions.
These values are for culverts flowing full. When
the entrance is not submerged, the coefficients are
usually somewhat lower. But because of the many
unknowns entering into determination of culvert
flow, the values tabulated can be used for sub-
merged or unsubmerged cases without much loss
of accuracy.
Example 21.9: Given: Maximum head above the
top of the culvert = 5 ft, slope = 0.01, length = 300
ft, discharge Q = 40 ft
3
/s, n = 0.013, and free exit.
Find: size of culvert.
Procedure: First assume a trial culvert; then
investigate the assumed section to find its dis-
charge. Assume a 2 2 ft concrete box section. Cal-
culate Q assuming entrance control, with Eq.
(21.44) for discharge through an orifice. The coeffi-
cient of discharge C for a 2-ft-square orifice is about
0.6. Head h on center line of entrance = 5 +
1
/2 2
= 6 ft. Entrance area a = 2 2 = 4 ft
2
.
For entrance control, the flow must be supercritical
and d
n
must be less than 2 ft. First find d
n
.
To calculate the hydraulic radius, assume the
depth is slightly less than 2 ft. since this will give
the maximum possible value of the hydraulic
radius for this culvert.
Application of Eq. (21.33a) gives
Since d
n
is greater than the culvert depth, the flow
is under pressure, and the entrance will not control.
Since the culvert is under pressure, Eq. (21.72)
applies. But
H = 5 + 0.01 300 = 8 ft
(see Fig. 21.40). The hydraulic radius for pipe flow
is R = 2
2
/8 =
1
/2. Substitution in Eq. (21.72) yields
Q =Va = 9.95 4 = 39.8 ft
3
/s
Fig. 21.41 Drawdown of water surface at a free
exit of a short culvert with slope less than critical
affects depth at entrance and controls discharge.
Inlet condition K
e
Sharp-edged projecting inlet 0.9
Flush inlet, square edge 0.5
Concrete pipe, groove or bell, projecting 0.15
Concrete pipe, groove or bell, flush 0.10
Well-rounded entrance 0.08
Table 21.10 Entrance Loss Coefficients for
Culverts
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.42
I
Section Twenty-One
Since the discharge of tile assumed culvert sec-
tion under the allowable head equals the maxi-
mum expected runoff, the assumed culvert would
be satisfactory.
Open-Channel Flow
Free surface flow, or open-channel flow, includes
all cases of flow in which the liquid surface is open
to the atmosphere. Thus, flow in a pipe is open-
channel flow if the pipe is only partly full.
21.21 Basic Elements of Open
Channels
A uniform channel is one of constant cross section.
It has uniform flow if the grade, or slope, of the
water surface is the same as that of the channel.
Hence, depth of flow is constant throughout.
Steady flowin a channel occurs if the depth at any
location remains constant with time.
The discharge Qat any section is defined as the
volume of water passing that section per unit of
time. It is expressed in cubic feet per second, ft
3
/s,
and is given by
(21.77)
where V = average velocity, ft/s
A = cross-sectional area of flow, ft
2
When the discharge is constant, the flow is said to
be continuous and therefore
(21.78)
where the subscripts designate different channel
sections. Equation (21.78) is known as the continu-
ity equation for continuous steady flow.
In a uniform channel, varied flow occurs if the
longitudinal water-surface profile is not parallel with
the channel bottom. Varied flow exists within the
limits of backwater curves, within a hydraulic jump,
and within a channel of changing slope or discharge.
Depth of flow d is taken as the vertical distance,
ft, from the bottom of a channel to the water sur-
face. The wetted perimeter is the length, ft, of a line
bounding the cross-sectional area of flow, minus
the free surface width. The hydraulic radius R
equals the area of flow divided by its wetted
perimeter. The average velocity of flow V is defined
as the discharge divided by the area of flow,
(21.79)
The velocity head H
V
, ft, is generally given by
(21.80)
where V = average velocity from Eq. (21.79), ft/s
g = acceleration due to gravity, 32.2 ft/s
2
Velocity heads of individual filaments of flow vary
considerably above and below the velocity head
based on the average velocity. Since these veloci-
ties are squared in head and energy computations,
the average of the velocity heads will be greater
than the average-velocity head. The true velocity
head may be expressed as
(21.81)
where is an empirical coefficient that represents
the degree of turbulence. Experimental data indi-
cate that may vary from about 1.03 to 1.36 for
prismatic channels. It is, however, normally taken
as 1.00 for practical hydraulic work and is evaluat-
ed only for precise investigations of energy loss.
The total energy per pound of water relative to
the bottom of the channel at a vertical section is
called the specific energy head H
e
. It is composed
of the depth of flow at any point, plus the velocity
head at the point. It is expressed in feet as
(21.82)
A longitudinal profile of the elevation of the spe-
cific energy head is called the energy grade line, or
the total-head line. A longitudinal profile of the
water surface is called the hydraulic grade line.
The vertical distance between these profiles at any
point equals the velocity head at that point.
Figure 21.42 shows a section of uniform open
channel for which the slopes of the water surface
S
w
and the energy grade line S equal the slope of
the channel bottom S
o
.
Loss of head due to friction h
f
in channel length
L equals the drop in elevation of the channel Z in
the same distance.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.43
21.22 Normal Depth of Flow
The depth of equilibrium flow that exists in the
channel of Fig. 21.42 is called the normal depth d
n
.
This depth is unique for specific discharge and
channel conditions. It may be computed by a trial-
and-error process when the channel shape, slope,
roughness, and discharge are known. A form of the
Manning equation has been suggested for this cal-
culation. (V. T. Chow. Open-Channel Hydraulics,
McGraw-Hill Book Company, New York.)
(21.83)
where A = area of flow, ft
2
R = hydraulic radius, ft
Q = amount of flow or discharge, ft
3
/s
n = Mannings roughness coefficient
S = slope of energy grade line or loss of
head, ft, due to friction per lin ft of
channel
AR
2/3
is referred to as a section factor. Depth d
n
for
uniform channels may be computed with computer
software or for manual computations simplified by
use of tables that relate d
n
to the bottom width of a
rectangular or trapezoidal channel, or to the diame-
ter of a circular channel. (See, for example, E. F.
Brater, Handbook of Hydraulics, 6th ed., McGraw-
Hill Book Company, New York.)
In a prismatic channel of gradually increasing
slope, normal depth decreases downstream, as
shown in Fig. 21.43, and specific energy first
decreases and then increases as shown in Fig. 21.44.
The specific energy is high initially where the
channel is relatively flat because of the large nor-
mal depth (Fig. 21.43). As the depth decreases
downstream, the specific energy also decreases. It
reaches a minimum at the point where the flow
satisfies the equation
(21.84)
in which T is the top width of the channel, ft. For a
rectangular channel, Eq. (21.84) reduces to
Fig. 21.42 Characteristics of uniform open-channel flow.
Fig. 21.43 Prismatic channel with gradually
increasing bottom slope. Normal depth increases
downstream as slope increases.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.44
I
Section Twenty-One
(21.85)
where V = Q/A = mean velocity of flow, ft
3
/s
d = depth of flow, ft
This indicates that the specific energy is a minimum
where the normal depth equals twice the velocity
head. As the depth continues to decrease in the
downstream direction, the specific energy increases
again because of the higher velocity head (Fig. 21.44).
21.23 Critical Depth of Open-
Channel Flow
The depth of flow that satisfies Eq. (21.84) is called
the critical depth d
c
. For a given value of specific
energy, the critical depth gives the greatest dis-
charge, or conversely, for a given discharge, the
specific energy is a minimum for the critical depth
(Fig. 21.44).
In the section of mild slope upstream from the
critical-depth point in Fig 21.43, the depth is
greater than critical. The flow there is called sub-
critical flow, indicating that the velocity is less
than that at critical depth. In the section of steeper
slope below the critical-depth point, the depth is
below critical. The velocity there exceeds that at
critical depth, and flow is supercritical.
Critical depth may be computed for a uniform
channel once the discharge is known. Determina-
tion of this depth is independent of the channel
slope and roughness since critical depth simply rep-
resents a depth for which the specific energy head
is a minimum. Critical depth may be calculated by
trial and error with Eq. (21.84), or it may be found
directly from tables (E. F. Brater, Handbook of
Hydraulics, 6th ed., McGraw-Hill Book Company,
New York). For rectangular channels, Eq. (21.84)
may be reduced to
(21.86)
Fig. 21.44 Specific energy head H
e
changes with depth for constant discharge in a rectangular channel
of changing slope. H
e
is a minimum for flow with critical depth.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.45
where d
c
= critical depth, ft
Q = quantity of flow or discharge, ft
3
/s
b = width of channel, ft
Critical slope is the slope of the channel bed
that will maintain flow at critical depth. Such slopes
should be avoided in channel design because flow
near critical depth tends to be unstable and exhibits
turbulence and water-surface undulations.
Critical depth, once calculated, should be plot-
ted for the full length of a uniform channel,
regardless of slope, to determine whether the nor-
mal depth at any section is subcritical or supercrit-
ical. [As indicated by Eq. (21.85), if the velocity
head is less than half the depth in a rectangular
channel, flow is subcritical, but if velocity head
exceeds half the depth, flow is supercritical.] If
channel configuration is such that the normal
depth must go from below to above critical, a
hydraulic jump will occur, along with a high loss of
energy. Critical depth will change if the channel
cross section changes, so the possibility of a
hydraulic jump in the vicinity of a transition
should be investigated.
For every depth greater than critical depth,
there is a corresponding depth less than critical that
has an identical value of specific energy (Fig. 21.44).
These depths of equal energy are called alternate
depths. The fact that the energy is the same for
alternate depths does not mean that the flow may
switch from one alternate depth to the other and
back again; flow will always seek to attain the nor-
mal depth in a uniform channel and will maintain
that depth unless an obstruction is met.
It can be seen from Fig. 21.44 that any obstruc-
tion to flow that causes a reduction in total head
causes subcritical flow to experience a drop in
depth and supercritical flow to undergo an
increase in depth.
If supercritical flow exists momentarily on a flat
slope because of a sudden grade change in the
channel (Fig. 21.52b, p. 21.57), depth increases sud-
denly from the depth below critical to a depth
above critical in a hydraulic jump. The depth fol-
lowing the jump will not be the alternate depth,
however. There has been a loss of energy in mak-
ing the jump. The new depth is said to be sequent
to the initial depth, indicating an irreversible
occurrence. There is no similar phenomenon that
allows a sudden change in depth from subcritical
flow to supercritical flow with a corresponding
gain in energy. Such a change occurs gradually,
without turbulence, as indicated in Fig. 21.45.
21.24 Mannings Equation for
Open Channels
One of the more popular of the numerous equa-
tions developed for determination of flow in an
open channel is Mannings variation of the Chezy
formula,
(21.87)
Fig. 21.45 Change in flow stage from subcritical to supercritical occurs gradually.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.46
I
Section Twenty-One
where R = hydraulic radius, ft
V = mean velocity of flow, ft/s
S = slope of energy grade line or loss of
head due to friction, ft/lin ft of channel
C = Chezy roughness coefficient
Manning proposed
(21.88)
where n is the coefficient of roughness in the earli-
er Ganguillet-Kutter formula (see also Art. 21.25).
When Mannings C is used in the Chezy formula,
the Manning equation for flow velocity in an open
channel results:
(21.89)
Since the discharge Q = VA, Eq. (21.89) may be
written
(21.90)
where A = area of flow, ft
2
Q = quantity of flow, ft
3
/s
Roughness Coefficient for Open Channels.
Values of the roughness coefficient n for Man-
nings equation have been determined for a wide
range of natural and artificial channel construction
materials. Excerpts from a table of these coeffi-
cients taken from V. T. Chow, Open-Channel
Hydraulics, McGraw-Hill Book Company, New
York, are in Table 21.11. Dr. Chow compiled data
for his table from work by R. E. Horton and from
technical bulletins published by the U.S. Depart-
ment of Agriculture.
Channel roughness does not remain constant
with time or even depth of flow. An unlined channel
excavated in earth may have one n value when first
put in service and another when overgrown with
weeds and brush. If an unlined channel is to have a
reasonably constant n value over its useful lifetime,
there must be a continuing maintenance program.
Shallow flow in an unlined channel will result
in an increase in the effective n value if the channel
bottom is covered with large boulders or ridges of
silt since these projections would then have a larg-
er influence on the flow than for deep flow. A deep-
er-than-normal flow will also result in an increase
in the effective n value if there is a dense growth of
brush along the banks within the path of flow.
When channel banks are overtopped during a
flood, the effective n value increases as the flow
spills into heavy growth bordering the channel.
(Although based on surface roughness, n in prac-
tice is sometimes treated as a lumped parameter for
all head losses.) The roughness of a lined channel
experiences change with age because of both dete-
rioration of the surface and accumulation of foreign
matter; therefore, the average n values given in
Table 21.11 are recommended only for well-main-
tained channels. (See also Art. 21.9 and Table 21.4.)
21.25 Water-Surface Profiles
for Gradually Varied
Flow
Examples of various surface curves possible with
gradually varied flow are shown in Fig. 21.46.
These surface profiles represent backwater curves
that form under the conditions illustrated in exam-
ples (a) through (r).
These curves are divided into five groups,
according to the slope of the channel in which they
appear (Art. 21.23). Each group is labeled with a let-
ter descriptive of the slope: M for mild (subcritical),
S for steep (supercritical), C for critical, H for hori-
zontal, and A for adverse. The two dashed lines in
the left-hand figure for each class are the normal-
depth line N.D.L. and the critical-depth line C.D.L.
The N.D.L. and C.D.L. are identical for a channel of
critical slope, and the N.D.L. is replaced by a hori-
zontal line, at an arbitrary elevation, for the chan-
nels of horizontal or adverse slope.
There are three types of surface-profile curves
possible in channels of mild or steep slope, and
two types for channels of critical, horizontal, and
adverse slope.
The M1 curve is the familiar surface profile from
which all backwater curves derive their name and is
the most important from a practical point of view. It
forms above the normal-depth line and occurs when
water is backed up a stream by high water in the
downstream channel, as shown in Fig. 21.46a and b.
The M2 curve forms between the normal- and
critical-depth lines. It occurs under conditions
shown in Fig. 21.46c and d, corresponding to an
increase in channel width or slope.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.47
The M3 curve forms between the channel bot-
tom and critical-depth line. It terminates in a
hydraulic jump, except where a drop-off in the
channel occurs before a jump can form. Examples
of the M3 curve are in Fig. 21.46e and f (a partly
opened sluice gate and a decrease in channel
slope, respectively).
The S1 curve begins at a hydraulic jump and
extends downstream, becoming tangent to a hori-
zontal line (Fig. 21.46g and h) under channel condi-
tions corresponding to those for Fig. 21.46a and b.
The S2 curve, commonly called a drawdown
curve, extends downstream from the critical depth
and becomes tangent to the normal-depth line
under conditions corresponding to those for Fig.
21.46i and j.
The S3 curve is of the transitional type. It forms
between two normal depths of less than critical
Min Avg Max
A. Open-channel flow in closed conduits
1. Corrugated-metal storm drain 0.021 0.024 0.030
2. Cement-mortar surface 0.011 0.013 0.015
3. Concrete (unfinished)
a. Steel form 0.012 0.013 0.014
b. Smooth wood form 0.012 0.014 0.016
c. Rough wood form 0.015 0.017 0.020
B. Lined channels
1. Metal
a. Smooth steel (unpainted) 0.011 0.012 0.014
b. Corrugated 0.021 0.025 0.030
2. Wood
a. Planed, untreated 0.010 0.012 0.014
3. Concrete
a. Float finish 0.013 0.015 0.016
b. Gunite, good section 0.016 0.019 0.023
c. Gunite, wavy section 0.018 0.022 0.025
4. Masonry
a. Cemented rubble 0.017 0.025 0.030
b. Dry rubble 0.023 0.032 0.035
5. Asphalt
a. Smooth 0.013 0.013
b. Rough 0.016 0.016
C. Unlined channels
1. Excavated earth, straight and uniform
a. Clean, after weathering 0.018 0.022 0.025
b. With short grass, few weeds 0.022 0.027 0.033
c. Dense weeds, high as flow depth 0.050 0.080 0.120
d. Dense brush, high stage 0.080 0.100 0.140
2. Dredged earth
a. No vegetation 0.025 0.028 0.033
b. Light brush on banks 0.035 0.050 0.060
3. Rock cuts
a. Smooth and uniform 0.025 0.035 0.040
b. Jagged and irregular 0.035 0.040 0.050
Table 21.11 Values of the Roughness Coefficient n for Use in the Manning Equation
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.48
I
Section Twenty-One
depth under conditions corresponding to those for
Fig. 21.46k and l.
Examples in Fig. 21.46m through r show condi-
tions for the formation of C, H, and A profiles.
The curves in Fig. 21.46 approach the normal-
depth line asymptotically and terminate abruptly
in a vertical line as they approach the critical depth.
The curves that approach the bottom intersect it at
Fig. 21.46 Typical flow profiles for channels with various slopes. N.D.L. indicates normal-depth line;
C.D.L., critical-depth line.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.49
a definite angle but are imaginary near the bottom
since velocity would have to be infinite to satisfy
Eq. (21.77) if the depth were zero. The curves are
shown dotted near the critical-depth line as a
reminder that this portion of the curve does not
possess the same degree of accuracy as the rest of
the curve because of neglect of vertical components
of velocity in the calculations. These curves either
start or end at what is called a point of control.
A point of control is a physical location in a
prismatic channel at which the depth of steady
flow may readily be determined. This depth is usu-
ally different from the normal depth for the chan-
nel because of a grade change, gate, weir, dam, free
overfall, or other feature at that location that caus-
es a backwater curve to form. Calculations for the
length and shape of the surface profile of a back-
water curve start at this known depth and location
and proceed either up or downstream, depending
on the type of flow. For subcritical flow conditions,
the curve proceeds upstream from the point of
control in a true backwater curve. The surface
curve that occurs under supercritical flow condi-
tions proceeds downstream from the point of con-
trol and might better be called a downwater curve.
The point of control is always at the down-
stream end of a backwater curve in subcritical flow
and at the upstream end for supercritical flow. This
is explained as follows: A backwater curve may be
thought of as being the result of some disruption
of uniform flow that causes a wave of disturbance
in the channel. The wave travels at a speed, known
as its celerity, which always equals the critical
velocity for the channel. If a disturbance wave
attempts to move upstream against supercritical
flow (flow moving at a speed greater than critical),
it will be swept downstream by the flow and have
no effect on conditions upstream. A disturbance
wave is held steady by critical flow and moves
upstream in subcritical flow.
When a hydraulic jump occurs on a mild slope
and is followed by a free overfall (Fig. 21.51), back-
water curves form both before and after the jump.
The point of control for the curve in the supercrit-
ical region above the jump will be located at the
vena contracta that forms just below the sluice
gate. The point of control for the backwater curve
in the subcritical region below the jump is at the
free overfall where critical depth occurs. Computa-
tions for these backwater curves are carried toward
the jump from their respective points of control
and are extended across the jump to help deter-
mine its exact location. But a backwater curve can-
not be calculated through a hydraulic jump from
either direction. The surface profiles involved ter-
minate abruptly in a vertical line as they approach
the critical depth, and a hydraulic jump always
occurs across critical depth. See Art. 21.27.5.
(R. H. French, Open-Channel Hydraulics,
McGraw-Hill, Inc., New York.)
21.26 Backwater-Curve
Computations
The solution of a backwater curve involves compu-
tation of a gradually varied flow profile. Solutions
available include the graphical-integration, direc-
tion-integration, and step methods. Explanations of
both the graphical- and direct-integration methods
are in V. T. Chow, Open-Channel Hydraulics,
McGraw-Hill Book Company, New York.
Two variations of the step method include the
direct or uniform method and the standard
method. They are simple and widely used and are
available in many software packages.
For step-method computations, the channel is
divided into short lengths, or reaches, with rela-
tively small variation. In a series of steps starting
from a point of control, each reach is solved in suc-
cession. Step methods have been developed for
channels with uniform or varying cross sections.
Direct step method of backwater computation
involves solving for an unknown length of channel
between two known depths. The procedure is
applicable only to uniform prismatic channels with
gradually varying area of flow.
For the section of channel in Fig. 21.47, Bernoullis
equation for the reach between sections 1 and 2 is
(21.91)
where V
1
and V
2
= mean velocities of flow at
sections 1 and 2, ft/s
d
1
and d
2
= depths of flow at sections 1
and 2, ft
g = acceleration due to gravity,
32.2 ft/s
2
S

= average head loss due to


friction, ft/ft of channel
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.50
I
Section Twenty-One
S
o
= slope of channel bottom
L = length of channel between
sections 1 and 2, ft
Note that S
o
L = z, the change in elevation, ft, of
the channel bottom between sections 1 and 2, and
S

L = h
f
, the head loss, ft, due to friction in the same
reach. (For uniform, prismatic channels, h
i
, the
eddy loss, is negligible and can be ignored.) S

equals the slope calculated for the average depth


in the reach but may be approximated by the aver-
age of the values of friction slope S for the depths
at sections 1 and 2.
Solving Eq. (21.91) for L gives
(21.92)
where H
e1
and H
e2
are the specific energy heads for
sections 1 and 2, respectively, as given by Eq.
(21.82). The friction slope S at any point may be
computed by the Manning equation, rearranged as
follows:
(21.93)
where R = hydraulic radius, ft
n = roughness coefficient (Art. 21.24)
Note that the slope S used in the Manning
equation is the slope of the energy grade line, not
the channel bottom. Note also that the roughness
coefficient n is squared in Eq. (21.93), and its value
must therefore be chosen with special care to avoid
an exaggerated error in the computed friction
slope. The smaller the value of n, the longer the
backwater curve profile, and vice versa. Therefore,
the smallest n possible for the prevailing condi-
tions should be selected for computation of a back-
water curve if knowledge of the longest possible
flow profile is required.
The first step in the direct step method involves
choosing a series of depths for the end points of
each reach. These depths will range from the
depth at the point of control to the ending depth
for the backwater curve. This ending depth is often
the normal depth for the channel (Art. 21.22) but
Fig. 21.47 Channel with constant discharge and gradually varying cross section.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.51
may be some intermediate depth, such as for a
curve preceding a hydraulic jump. Depths should
be chosen so that the velocity change across a
reach does not exceed 20% of the velocity at the
beginning of the reach. Also the change in depth
between sections should never exceed 1 ft.
The specific energy head H
e
should be comput-
ed for the chosen depth at each of the various sec-
tions and the change in specific energy between
sections determined. Next, the friction slope S
should be computed at each section from Eq.
(21.93). The average of two sections gives the fric-
tion slope S

between sections. Finally, the differ-


ence between S

and slope of channel bottom S


o
should be computed and the length of reach deter-
mined from Eq. (21.92).
Standard step method allows computation of
backwater curves in both nonprismatic natural
channels and nonuniform artificial channels as
well as in uniform channels. This method involves
solving for the depth of flow at various locations
along a channel with Bernoullis energy equation
and a known length of reach.
A surface profile is determined in the following
manner: The channel is examined for changes in
cross section, grade, or roughness, and the loca-
tions of these changes are given station numbers.
Stations are also established between these loca-
tions such that the velocity change between any
two consecutive stations is not greater than 20% of
the velocity at the former station. Data concerning
the hydraulic elements of the channel are collected
at each station. Computation of the surface curve is
then made in steps, starting from the point of con-
trol and progressing from station to stationin an
upstream direction for subcritical flow and down-
stream for supercritical flow. The length of reach in
each step is given by the stationing, and the depth
of flow is determined by trial and error.
Nonprismatic channels do not have well-
defined points of control to aid in determining the
starting depth for a backwater curve. Therefore,
the water-surface elevation at the beginning must
be determined as follows:
The step computations are started at a point in
the channel some distance upstream or downstream
from the desired starting point, depending on
whether flow is supercritical or subcritical, respec-
tively. Then, computations progress toward the ini-
tial section. Since this step method is a converging
process, this procedure produces the true depth for
the initial section within a relatively few steps.
The energy balance used in the standard step
method is shown graphically in Fig. 21.47, in which
the position of the water surface at section 1 is Z
1
and at section 2, Z
2
, referred to a horizontal datum.
Writing Bernoullis equation [Eq. (21.11)] for sec-
tions 1 and 2 Yields
(21.94)
where V
1
and V
2
are the mean velocities, ft/s, at sec-
tions 1 and 2; the friction loss, ft, in the reach (S

L)
is denoted by h
f
; and the term h
i
is added to
account for eddy loss, ft.
Eddy loss, sometimes called impact loss, is a
head loss caused by flow running contrary to the
main current because of irregularities in the chan-
nel. No rational method is available for determina-
tion of eddy loss, and it is therefore often accounted
for, in natural channels, by a slight increase in Man-
nings n. Eddy loss depends mainly on a change in
velocity head. For lined channels, it has been
expressed as a coefficient k to be applied as follows:
(21.95)
The coefficient k is 0.2 for diverging reaches, from
0 to 0.1 for converging reaches, and about 0.5 for
abrupt expansions and contractions.
The total head at any section of the channel is
(21.96)
where Z equals the elevation of the channel bot-
tom above the given datum plus the depth of flow
d at that section. Friction slope S is computed from
Eq. (21.93). Then, S

, the average friction slope for


the reach, is calculated as the mean of the slope for
the section and the preceding section. Friction loss
h
f
is the product of S

and the length of the reach L.


Eddy loss h
i
is found from Eq. (21.95). Next, total
head H, ft, is obtained from Eq. (21.94), which, after
substitution of H from Eq. (21.96), becomes
(21.97)
where H
1
and H
2
equal the total head of sections 1
and 2, respectively. The value of total head com-
puted from Eq. (21.97) must agree with the value of
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.52
I
Section Twenty-One
total head calculated previously for the section or
the assumed water-surface elevation Z
1
is incorrect.
Agreement is assumed if the two values of total
head are within 0.1 ft in elevation. If the two values
of total head do not agree, a new water-surface ele-
vation must be assumed for Z
1
and the computa-
tions repeated until agreement is obtained. The
value that finally leads to agreement gives the cor-
rect water-surface elevation.
Backwater curves for natural river or stream
channels (irregularly shaped channels) are calcu-
lated in a manner similar to that described for reg-
ularly shaped channels. However, some account
must be taken of the varying channel roughness
and the differences in velocity and capacity in the
main channel and the overbank or floodplain por-
tions of the stream channel. The most expeditious
way of determining the backwater curves is to plot
the channel cross section to a scale convenient for
measurement of lengths and areas; subdivide the
cross section into main channels and floodplain
areas; and determine the discharge, velocity, and
friction slope for each subarea at selected water-
surface elevations. Utilizing the above data, deter-
mine the total discharge (the sum of the subarea
discharges), the mean velocity (the total discharge
divided by the total area), and (the energy coef-
ficient or coriolis coefficient to be applied to the
velocity head). Many of the available computer
software packages that compute backwater pro-
files are applicable to irregular channels and flood-
ed overbank areas.
The backwater curve is usually started by
assuming normal depth at a point some distance
downstream from the start of the reach under
analysis. Several intermediate cross sections
should be taken between the point where normal
depth is assumed and the start of the reach for
which a detailed water-surface profile is required.
This allows the intermediate sections to dampen
out any minor errors in the assumed starting
water-surface elevation.
The accuracy or validity of the water-surface
profile is contingent on an accurate evaluation of
the channel roughness and judicious selection of
cross-section location. A greater number of cross
sections generally enhances the validity of the
water-surface profile; however, because of the
extensive calculations involved with each cross
section, their number should be limited to as few
as accuracy permits.
The effect of bridges, approach roadways, bridge
piers, and culverts can be determined using proce-
dures outlined in R. H. French, Open-Channel
Hydraulics, McGraw-Hill Book Company, New
York, and J. N. Bradley, Hydraulics of Bridge Water-
ways, Hydraulics Design Series no. 1, 2nd ed., U.S.
Department of Transportation, Federal Highway
Administration, Bureau of Public Roads, 1970.
21.27 Hydraulic Jump
This is an abrupt increase in depth of rapidly flow-
ing water (Fig. 21.48). Flow at the jump changes
from a supercritical to a subcritical stage with an
accompanying loss of kinetic energy (Art. 21.23).
A hydraulic jump is the only means by which
the depth of flow can change from less than critical
to greater than critical in a uniform channel. A jump
will occur either where supercritical flow exists in a
channel of subcritical slope, as shown in Figs. 21.51
and 21.52b, or where a steep channel enters a reser-
voir. The first condition is met in a mild channel
downstream from a sluice gate or ogee overflow
spillway, or at an abrupt change in channel slope
from steep to mild. The second condition occurs
where flow in a steep channel is blocked by an over-
flow weir, a gate, or other obstruction.
A hydraulic jump can be either stationary or
moving, depending on whether the flow is steady
or unsteady, respectively.
21.27.1 Depth and Head Loss in a
Hydraulic Jump
Depth at the jump is not discontinuous. The change
in depth occurs over a finite distance, known as the
length of jump. The upstream surface of the jump,
known as the roller, is a turbulent mass of water,
Fig. 21.48 Hydraulic jump.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.53
which is continually tumbling erratically against the
rapidly flowing sheet below.
The depth before a jump is the initial depth,
and the depth after a jump is the sequent depth.
The specific energy for the sequent depth is less
than that for the initial depth because of the ener-
gy dissipation within the jump. (Initial and
sequent depths should not be confused with the
depths of equal energy, or alternate depths.)
According to Newtons second law of motion,
the rate of loss of momentum at the jump must
equal the unbalanced pressure force acting on the
moving water and tending to retard its motion.
This unbalanced force equals the difference
between the hydrostatic forces corresponding to
the depths before and after the jump. For rectan-
gular channels, this resultant pressure force is
(21.98)
where d
1
= depth before jump, ft
d
2
= depth after jump, ft
w = unit weight of water, lb/ft
3
The rate of change of momentum at the jump per
foot width of channel equals
(21.99)
where M = mass of water, lbs
2
/ft
V
1
= velocity at depth d
1
, ft/s
V
2
= velocity at depth d
2
, ft/s
q = discharge per foot width of rectan-
gular channel, ft
3
/s
t = unit of time, s
g = acceleration due to gravity, 32.2 ft/s
2
Equating the values of F in Eqs. (21.98) and (21.99),
and substituting V
1
d
1
for q and V
1
d
1
/d
2
for V
2
, the
reduced equation for rectangular channels becomes
(21.100)
Equation (21.100) may then be solved for the
sequent depth:
(21.101)
If V
2
d
2
/d
1
is substituted for V
1
, in Eq. (21.100),
(21.102)
Equation (21.102) may be used in determining the
position of the jump where V
2
and d
2
are known.
Relationships may be derived similarly for chan-
nels of any cross section.
The head loss in a jump equals the difference in
specific-energy head before and after the jump.
This difference (Fig. 21.49) is given by
(21.103)
where H
e1
= specific-energy head of stream before
jump, ft
H
e2
= specific-energy head of stream after
jump, ft
The specific energy for free-surface flow is given
by Eq. (21.82).
The depths before and after a hydraulic jump
may be related to the critical depth by the equation
(21.104)
where q = discharge, ft
3
/s per ft of channel
width
d
c
= critical depth for the channel, ft
It may be seen from this equation that if d
1
= d
c
, d
2
must also equal d
c
.
21.27.2 Jump in Horizontal
Rectangular Channels
The form of a hydraulic jump in a horizontal rec-
tangular channel may be of several distinct types,
depending on the Froude number of the incoming
flow F = V/(gL)
1/2
[Eq. (21.16)], where L is a charac-
teristic length, ft; V is the mean velocity, ft/s; and g
= acceleration due to gravity, ft/s
2
. For open-chan-
nel flow, the characteristic length for the Froude
number is made equal to the hydraulic depth d
h
.
Hydraulic depth is defined as
(21.105)
where A = area of flow, ft
2
T = width of free surface, ft
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.54
I
Section Twenty-One
For rectangular channels, hydraulic depth equals
depth of flow.
Various forms of hydraulic jump, and their rela-
tion to the Froude number of the approaching flow
F
1
, were classified by the U.S. Bureau of Reclama-
tion and are presented in Fig. 21.49.
For F
1
= 1, the flow is critical and there is no
jump.
For F
1
= 1 to 1.7, there are undulations on the
surface. The jump is called an undular jump.
For F
1
= 1.7 to 2.5, a series of small rollers devel-
op on the surface of the jump, but the downstream
water surface remains smooth. The velocity
throughout is fairly uniform and the energy loss is
low. This jump may be called a weak jump.
For F
1
= 2.5 to 4.5, an oscillating jet is entering
the jump. The jet moves from the channel bottom
to the surface and back again with no set period.
Each oscillation produces a large wave of irregular
period, which, very commonly in canals, can trav-
el for miles, doing extensive damage to earth banks
and riprap surfaces. This jump may be called an
oscillating jump.
For F
1
= 4.5 to 9.0, the downstream extremity of
the surface roller and the point at which the high-
velocity jet tends to leave the flow occur at practi-
cally the same vertical section. The action and posi-
tion of this jump are least sensitive to variation in
tailwater depth. The jump is well-balanced, and
the performance is at its best. The energy dissipa-
tion ranges from 45 to 70%. This jump may be
called a steady jump.
For F
1
= 9.0 and larger, the high-velocity jet grabs
intermittent slugs of water rolling down the front
face of the jump, generating waves downstream and
causing a rough surface. The jump action is rough
but effective, and energy dissipation may reach 85%.
This jump may be called a strong jump.
Note that the ranges of the Froude number
given for the various types of jump are not clear-
cut but overlap to a certain extent, depending on
local conditions.
21.27.3 Hydraulic Jump as an
Energy Dissipator
A hydraulic jump is a useful means for dissipating
excess energy in supercritical flow (Art. 21.23). A
jump may be used to prevent erosion below an
overflow spillway, chute, or sluice gate by quickly
reducing the velocity of the flow over a paved
apron. A special section of channel built to contain
a hydraulic jump is known as a stilling basin.
If a hydraulic jump is to function ideally as an
energy dissipator, below a spillway, for example,
the elevation of the water surface after the jump
must coincide with the normal tailwater elevation
for every discharge. If the tailwater is too low, the
high-velocity flow will continue downstream for
some distance before the jump can occur. If the tail-
water is too high, the jump will be drowned out,
and there will be a much smaller dissipation of total
head. In either case, dangerous erosion is likely to
occur for a considerable distance downstream.
The ideal condition is to have the sequent-depth
curve, which gives discharge vs. depth after the
jump, coincide exactly with the tailwater-rating
curve. The tailwater-rating curve gives normal
depths in the discharge channel for the range of
flows to be expected. Changes in the spillway
design that can be made to alter the tailwater-rating
Fig. 21.49 Type of hydraulic jump depends on
Froude number.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.55
curve involve changing the crest length, changing
the apron elevation, and sloping the apron.
Accessories, such as chute blocks and baffle
blocks are usually installed in a stilling basin to
control the jump. The main purpose of these acces-
sories is to shorten the range within which the
jump will take place, not only to force the jump to
occur within the basin but to reduce the size and
therefore the cost of the basin. Controls within a
stilling basin have additional advantages in that
they improve the dissipation function of the basin
and stabilize the jump action.
21.27.4 Length of Hydraulic Jump
The length of a hydraulic jump L may be defined as
the horizontal distance from the upstream edge of
the roller to a point on the raised surface immedi-
ately downstream from cessation of the violent tur-
bulence. This length (Fig. 21.48) defies accurate
mathematical expression, partly because of the
nonuniform velocity distribution within the jump.
But it has been determined experimentally. The
experimental results may be summarized conve-
niently by plotting the Froude number of the
upstream flow F
1
against a dimensionless ratio of
jump length to downstream depth L/d
2
. The result-
ing curve (Fig. 21.50) has a flat portion in the range
of steady jumps. The curve thus minimizes the
effect of any errors made in calculation of the
Froude number in the range where this information
is most frequently needed. The curve, prepared by
V. T. Chow from data gathered by the U.S. Bureau of
Reclamation, was developed for jumps in rectangu-
lar channels, but it will give approximate results for
jumps formed in trapezoidal channels.
For other than rectangular channels the depth
d
1
used in the equation for Froude number is the
hydraulic depth given by Eq. (21.105).
21.27.5 Location of a Hydraulic
Jump
It is important to know where a hydraulic jump
will form since the turbulent energy released in a
jump can extensively scour an unlined channel or
destroy paving in a thinly lined channel. Special
reinforced sections of channel must be built to
withstand the pounding and vibration of a jump
and to provide extra freeboard for the added depth
at the jump. These features are expensive to build;
therefore, a great savings can be realized if their
use is restricted to a limited area through a knowl-
edge of the jump location.
The precision with which the location is pre-
dicted depends on the accuracy with which the
friction losses and length of jump are estimated
and on whether the discharge is as assumed. The
method of prediction used for rectangular chan-
nels is illustrated for a sluice gate in Fig. 21.51.
Fig. 21.50 Length of hydraulic jump in a horizontal channel depends on sequent depth d
2
and the
Froude number of the approaching flow.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.56
I
Section Twenty-One
The water-surface profiles of the flow approach-
ing and leaving the jump, curves AB and ED in Fig.
21.51, are type M3 and M2 backwater curves,
respectively (Fig. 21.46e and c).
Backwater curve ED has as its point of control
the critical depth d
c
, which occurs near the channel
drop-off. Critical depth does not exist exactly at the
edge, as theory would indicate, but instead occurs
a short distance upstream. The distance is small
(from three to four times d
c
) and can be ignored for
most problems. The actual depth at the brink is
71.5% of critical depth, but it is normally assumed
to be 0.7d
c
for simplicity.
The point of control for backwater curve AB is
taken as the depth at the vena contracta, which
forms just downstream from the sluice gate. The
distance from the gate to the vena contracta L
e
is
nearly equal to the size of gate opening h. The
amount of contraction varies with both the head
on the gate and the gate opening. Depth at the
contraction ranges from 50 to over 90% of h. The
depth of flow at the vena contracta may be taken
as 0.75h in the absence of better information.
Jump location is determined as follows: The
backwater curves AB and ED are computed in
their respective directions until they overlap, using
the step methods of Art. 21.26. With values of d
2
obtained from Eq. (21.101), CB, the curve of depths
sequent to curve AB, is plotted through the area
where it crosses curve ED. A horizontal intercept
FG, equal in length to L, the computed length of
jump, is then fitted between the curves CB and ED.
The jump may be expected to form between the
points H and G since all requirements for the for-
mation of a jump are satisfied at this location.
If the downstream depth is increased because
of an obstruction, the jump moves upstream and
may eventually be drowned out in front of the
sluice gate. Conversely, if the downstream depth
is lowered, the jump moves to a new location
downstream.
When the slope of a channel has an abrupt
change from steeper than critical (Art. 21.23) to
mild, a jump forms that may be located either
above or below the grade change. The position of
the jump depends on whether the downstream
depth d
2
is greater than, less than, or equal to the
depth d
1
sequent to the upstream depth d
1
. Two
possible positions are shown in Fig. 21.52.
It is assumed, for simplicity, that flow is uni-
form, except in the reach between the jump and
the grade break. If the downstream depth d
2
is
greater than the upstream sequent depth d
1
, com-
puted from Eq. (21.101) with d
1
given, the jump
occurs in the steep region, as shown in Fig. 21.52a.
The surface curve EO is of the S1 type (Fig. 21.46)
and is asymptotic to a horizontal line at O. Line CB
is a plot of the depth d
1
sequent to the depth of
approach line AB. The jump location is found by
producing a horizontal intercept FG, equal to the
computed length of the jump, between lines
CBand EO. A jump will form between H and G
since all requirements are satisfied for this location.
As depth d
2
is lowered, the jump moves down-
stream to a new position, as shown in Fig. 21.52b. If
d
2
is less than d
1
, computed from Eq. (21.102), the
Fig. 21.51 Graphical method for locating hydraulic jump beyond a sluice gate.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.57
jump will form in the mild channel and can be
located as described for Fig. 21.51.
(R. H. French, Open-Channel Hydraulics,
McGraw-Hill, Inc., New York.)
21.28 Flow at Entrance to a
Steep Channel
The discharge Q, ft
3
/s, in a channel leaving a reser-
voir is a function of the total head H, ft, on the
channel entrance, the entrance loss, ft. and the
slope of the channel. If the channel has a slope
steeper than the critical slope (Art. 21.23), the flow
passes through critical depth at the entrance, and
discharge is at a maximum. If the channel entrance
is rectangular in cross section, the critical depth d
c
=
2
/3H
e
[according to Eqs. (21.82) and (21.85)],
where H
e
is the specific energy head, ft, in the
reservoir and datum is the elevation of the lip of
the channel (Fig. 21.53a).
From Q = AV, with the area of flow A = bd
c
=
2
/3bH
e
and the velocity
the discharge for rectangular channels, ignoring
entrance loss, is
(21.106)
where b is the channel width, ft.
If the entrance loss must be considered, or if the
channel entrance is other than rectangular, the
inlet depth must be solved for by trial and error
since the discharge is unknown. The procedure for
finding the correct discharge is as follows:
A trial discharge is chosen. Then, the critical
depth for the given shape of channel entrance is
determined (see those in E. F. Brater, Handbook of
Hydraulics, 6th ed., McGraw-Hill Book Company,
New York.) Adding d
c
to its associated velocity
head gives the specific energy in the channel
entrance, to which the resulting entrance loss is
added. This sum then is compared with the specif-
ic energy of the reservoir water, which equals the
depth of water above datum plus the velocity head
of flow toward the channel. (This velocity head is
normally so small that it may be taken as zero in
most calculations.) If the specific energy computed
for the depth of water in the reservoir equals the
sum of specific energy and entrance loss deter-
mined for the channel entrance, then the assumed
discharge is correct; if not, a new discharge is
assumed, and the computations continued until a
balance is reached.
A first trial discharge may be found from Q =
A

2g(H
e
d)

, where (H
e
d) gives actual head pro-
ducing flow (Fig. 21.53). A reasonable value for the
depth d would be
2
/
3
H
e
for steep channels and an
even greater percentage of H
e
for mild channels.
The entrance loss equals the product of an
empirical constant k and the change in velocity
head H

at the entrance. If the velocity in the


reservoir is assumed to be zero, then the entrance
loss is k(V
2
1
/ 2g), where V
1
is the velocity computed
for the channel entrance. Safe design values for
the coefficient vary from about 0.1 for a well-
rounded entrance to slightly over 0.3 for one with
squared ends.
Fig. 21.52 Hydraulic jump may occur at a change
in bottom slope, or (a) above it, or (b) below it.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.58
I
Section Twenty-One
21.29 Flow at Entrance to a
Channel of Mild Slope
When water flows from a reservoir into a channel
with slope less than the critical slope (Art. 21.23),
the depth of flow at the channel entrance equals
the normal depth for the channel (Art. 21.22). The
entrance depth and discharge are dependent on
each other. The discharge that results from a given
head is that for which flow enters the channel
without forming either a backwater or drawdown
curve within the entrance. This requirement neces-
sitates the formation of normal depth d since only
at this equilibrium depth is there no tendency to
change the discharge or to form backwater curves.
(In Fig. 21.53b, d is normal depth.)
A solution for discharge at entrance to a chan-
nel of mild slope is found as follows: A trial dis-
charge, ft
3
/s, is estimated from Q = A

2g(H
e
d)

,
where H
e
d is the actual head, ft, producing flow.
H
e
is the specific energy head, ft, of the reservoir
water relative to datum at lip of channel; A is the
cross-sectional area of flow, ft
2
; and g is acceleration
due to gravity, 32.2 ft/s
2
. The normal depth of the
channel is determined for this discharge from Eq.
(21.83). The velocity head is computed for this
depth-discharge combination, and an entrance-loss
calculation is made (see Art. 21.33). The sum of the
specific energy of flow in the channel entrance and
the entrance loss must equal the specific energy of
the water in the reservoir for an energy balance to
exist between those points (Fig. 21.53b). If the trial
discharge gives this balance of energy, then the dis-
charge is correct; if not, a new discharge is chosen,
and the calculations continued until a satisfactory
balance is obtained.
Fig. 21.53 Flow at entrance to (a) steep channel; (b) mild-slope channel.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.59
21.30 Channel Section of
Greatest Efficiency
If a channel of any shape is to reach its greatest
hydraulic efficiency, it must have the shortest possi-
ble wetted perimeter for a given cross-sectional area.
The resulting shape gives the greatest hydraulic
radius and therefore the greatest capacity for that
area. This can be seen from the Manning equation
for discharge [Eq. (21.83)], in which Q is a direct
function of hydraulic radius to the two-thirds power.
The most efficient of all possible open-channel
cross sections is the semicircle. There are practical
objections to the use of this shape because of the
difficulty of construction, but it finds some use in
metal flumes where sections can be preformed.
The most efficient of all trapezoidal sections is the
half hexagon, which is used extensively for large
water-supply channels. The rectangular section
with the greatest efficiency has a depth of flow
equal to one-half the width. This shape is often
used for box culverts and small drainage ditches.
21.31 Subcritical Flow around
Bends in Channels
Because of the inability of liquids to resist shearing
stress, the free surface of steady uniform flow is
always normal to the resultant of the forces acting
on the water. Water in a reservoir has a horizontal
surface since the only force acting on it is the force
of gravity.
Water reacts in accordance with Newtons first
law of motion: It flows in a straight line unless
deflected from its path by an outside force. When
water is forced to flow in a curved path, its surface
assumes a position normal to the resultant of the
forces of gravity and radial acceleration. The force
due to radial acceleration equals the force required
to turn the water from a straight-line path, or
mV
2
/ r
c
for m, a unit mass of water, where V is its
average velocity, ft / s, and r
c
the radius of curva-
ture, ft, of the center line of the channel.
The water surface makes an angle with the
horizontal such that
(21.107)
The theoretical difference y, ft. in water-surface
level between the inside and outside banks of a
curve (Fig. 21.54) is found by multiplying tan by
the top width of the channel T, ft. Thus,
(21.108)
where the radius of curvature r
c
of the center of the
channel is assumed to represent the average cur-
vature of flow. This equation gives values of y
smaller than those actually encountered because of
the use of average values of velocity and radius,
rather than empirically derived values more repre-
sentative of actual conditions. The error will not be
great, however, if the depth of flow is well above
critical (Art. 21.23). In this range, the true value of
y would be only a few inches.
The difference in surface elevation found from
Eq. (21.108), although it involves some drop in sur-
face elevation on the inside of the curve, does not
allow a savings of freeboard height on the inside
bank. The water surface there is wavy and thus
needs a freeboard height at least equal to that of a
straight channel.
The top layer of flow in a channel has a higher
velocity than flow near the bottom because of the
retarding effect of friction along the floor of the chan-
nel. A greater force is required to deflect the high-
velocity flow. Therefore, when a stream enters a
curve, the higher-velocity flow moves to the outside
of the bend. If the bend continues long enough, all
the high-velocity water will move against the outer
bank and may cause extensive scour unless special
bank protection is provided.
Fig. 21.54 Water-surface profile at a bend in a
channel with subcritical flow.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.60
I
Section Twenty-One
Since the higher-velocity flow is pressed direct-
ly against the bank, an increase in friction loss
results. This increased loss may be accounted for in
calculations by assuming an increased value of the
roughness coefficient n within the curve. Scobey
suggests that the value of n be increased by 0.001
for each 20 of curvature in 100 ft of flume. His val-
ues have not been evaluated completely, however,
and should be used with discretion. (F. C. Scobey,
The Flow of Water in Flumes, U.S. Department of
Agriculture, Technical Bulletin 393.)
21.32 Supercritical Flow
around Bends in
Channels
When water, traveling at a velocity greater than
critical (Art. 21.23), flows around a bend in a chan-
nel, a series of standing waves are produced. Two
waves form at the start of the curve. One is a posi-
tive wave, of greater-than-average surface eleva-
tion, which starts at the outside wall and extends
across the channel on the line AME (Fig. 21.55).
The second is a negative wave, with a surface ele-
vation of less-than-average height, which starts at
the inside wall and extends across the channel on
the line BMD. These waves cross at M, are reflect-
ed from opposite channel walls at D and E, recross
as shown, and continue crossing and recrossing.
The two waves at the entrance form at an angle
with the approach channel known as the wave
angle
o
. This angle may be determined from the
equation
(21.109)
where F
1
represents the Froude number of flow in
the approach channel [Eq. (21.16)] .
The distance from the beginning of the curve to
the first wave peak on the outside bank is deter-
mined by the central angle
o
. This angle may be
found from
(21.110)
where T is the normal top width of channel and r
c
is the radius of curvature of the center of channel.
The depths along the banks at an angle <
o
are
given by
(21.111)
where the positive sign gives depths along the
outside wall and the negative sign, depths along
the inside wall. The depth of maximum height for
the first positive wave is obtained by substituting
the value of
o
found from Eq. (21.110) for in Eq.
(21.111).
Standing waves in existing rectangular chan-
nels may be prevented by installing diagonal sills
at the beginning and end of the curve. The sills
introduce a counterdisturbance of the right magni-
tude, phase, and shape to neutralize the undesir-
able oscillations that normally form at the change
of curvature. The details of sill design have been
determined experimentally.
Good flow conditions may be ensured in new
projects with supercritical flow in rectangular chan-
nels by providing transition curves or by banking
the channel bottom. Circular transition curves aid
in wave control by setting up counterdisturbances
in the flow similar to those provided by diagonal
sills. A transition curve should have a radius of cur-
vature twice the radius of the central curve. It
should curve in the same direction and have a cen-
tral angle given, with sufficient accuracy, by
(21.112)
Transition curves should be used at both the begin-
ning and end of a curve to prevent disturbances
downstream.
Banking the channel bottom is the most effec-
tive method of wave control. It permits equilibri-
um conditions to be set up without introduction of
a counterdisturbance. The cross slope required for
Fig. 21.55 Plan view of supercritical flow
around a bend in an open channel.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.61
equilibrium is the same as the surface slope found
for subcritical flow around a bend (Fig. 21.54). The
angle the bottom makes with the horizontal is
found from the equation
(21.113)
21.33 Transitions in Open
Channels
A transition is a structure placed between two
open channels of different shape or cross-sectional
area to produce a smooth, low-head-loss transfer
of flow. The major problems associated with
design of a transition lie in locating the invert and
determining the various cross-sectional areas so
that the flow is in accord with the assumptions
made in locating the invert. Many variables, such
as flow-rate changes, wall roughness, and channel
shape and slope, must be taken into account in
design of a smooth-flow transition.
When proceeding downstream through a tran-
sition, the flow may remain subcritical or super-
critical (Art. 21.23), change from subcritical to
supercritical, or change from supercritical to sub-
critical. The latter flow possibility may produce a
hydraulic jump.
Special care must be exercised in the design if
the depth in either of the two channels connected
is near the critical depth. In this range, a small
change in energy head within the transition may
cause the depth of flow to change to its alternate
depth. A flow that switches to its subcritical alter-
nate depth may overflow the channel. A flow that
changes to its supercritical alternate depth may
cause excessive channel scour. The relationship of
flow depth to energy head can be shown on a plot
such as Fig. 21.44, p. 21.44.
To place a transition properly between two
open channels, it is necessary to determine the
design flow and calculate normal and critical
depths for each channel section. Maximum flow is
usually selected as the design flow. Normal depth
for each section is used for the design depth. After
the design has been completed for maximum flow,
hydraulic calculations should be made to check the
suitability of the structure for lower flows.
The transition length that produces a smooth-
flowing, low-head-loss structure is obtained for an
angle of about 12.5 between the channel axis and
the lines of intersection of the water surface with
the channel sides, as shown in Fig. 21.56. The
length of the transition L
t
is then given by
(21.114)
where T
2
and T
1
are the top widths of sections 2
and 1, respectively.
In design of an inlet-type transition structure,
the water-surface level of the downstream channel
must be set below the water-surface level of the
upstream channel by at least the sum of the
increase in velocity head, plus any transition and
friction losses. The transition loss, ft, is given by
K(V
2
/2g), where K, the loss factor, equals about 0.1
for an inlet-type structure; V is the velocity
change, ft/s; and g = 32.2 ft/s
2
. The total drop in
water surface y
d
across the inlet-type transition is
then 1.1 [(V
2
/2g)], if friction is ignored.
For outlet-type structures, the average velocity
decreases, and part of the loss in velocity head is
recovered as added depth. The rise of the water
surface for an outlet structure equals the decrease
in velocity head minus the outlet and friction loss-
es. The outlet loss factor is normally 0.2 for well-
designed transitions. If friction is ignored, the total
rise in water surface y
r
across the outlet structure is
0.8[(V
2
/2g)].
Many well-designed transitions have a reverse
parabolic water-surface curve tangent to the water
surfaces in each channel (Fig. 21.57). After such a
water-surface profile is chosen, depth and cross-
sectional areas are selected at points along the
transition to produce this smooth curve. Straight,
angular walls usually will not produce a smooth
parabolic water surface; therefore, a transition
with a curved bottom or sides has to be designed.
Fig. 21.56 Plan view of a transition between
two open channels with different widths.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.62
I
Section Twenty-One
The total transition length L
t
is split into an even
number of sections of equal length x. For Fig. 21.57,
six equal lengths of 10 ft each are used, for an
assumed drop in water surface y
d
of 1 ft. It is
assumed that the water surface will follow parabola
AC for the length L
t
/ 2 to produce a water-surface
drop of y
d
/ 2 and that the other half of the surface
drop takes place along the parabola CB. The water-
surface profile can be determined from the general
equation for a parabola, y = ax
2
, where y is the verti-
cal drop in the distance x, measured from A or B.
The surface drops at sections 1 and 2 are found
as follows: At the midpoint of the transition, y
3
=
ax
2
= y
d
/ 2 = 0.5 = a(30)
2
, from which a = 0.000556.
Then y
1
= ax
2
1
= 0.000556(10)
2
= 0.056 ft and y
2
=
ax
2
2
= 0.000556(20)
2
= 0.222 ft.
21.34 Weirs
A weir is a barrier in an open channel over which
water flows. The edge or surface over which the
water flows is called the crest. The overflowing
sheet of water is the nappe.
If the nappe discharges into the air, the weir has
free discharge. If the discharge is partly under water,
the weir is submerged or drowned.
21.34.1 Types of Weirs
A weir with a sharp upstream corner or edge such
that the water springs clear of the crest is a sharp-
crested weir (Fig. 21.58). All other weirs are classed
as weirs not sharp-crested. Sharp-crested weirs are
classified according to the shape of the weir open-
ing, such as rectangular weirs, triangular or V-
notch weirs, trapezoidal weirs, and parabolic
weirs. Weirs not sharp-crested are classified
according to the shape of their cross section, such
as broad-crested weirs, triangular weirs, and, as
shown in Fig. 21.59, trapezoidal weirs.
The channel leading up to a weir is the channel
of approach. The mean velocity in this channel is
the velocity of approach. The depth of water pro-
ducing the discharge is the head.
Sharp-crested weirs are useful only as a means
of measuring flowing water. In contrast, weirs not
sharp-crested are commonly incorporated into
Fig. 21.57 Profile of reverse parabolic water-surface curve for well-designed transitions.
Fig. 21.58 Sharp-crested weir. Fig. 21.59 Weir not sharp-crested.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.63
hydraulic structures as control or regulation
devices, with measurement of flow as their sec-
ondary function.
21.34.2 Rectangular Sharp-Crested
Weirs
Discharge over a rectangular sharp-crested weir is
given by
(21.115)
where Q = discharge, ft
3
/s
C = discharge coefficient
L = effective length of crest, ft
H = measured head = depth of flow
above elevation of crest, ft
The head should be measured at least 2.5Hupstream
from the weir, to be beyond the drop in the water
surface (surface contraction) near the weir.
Numerous equations have been developed for
finding the discharge coefficient C. One such equa-
tion, which applies only when the nappe is fully
ventilated, was developed by Rehbock and simpli-
fied by Chow:
(21.116)
where P is the height of the weir above the chan-
nel bottom (Fig. 21.58) (V. T. Chow, Open-Chan-
nel Hydraulics, McGraw-Hill Book Company,
New York).
The height of weir P must be at least 2.5H for a
complete crest contraction to form. If P is less than
2.5H, the crest contraction is reduced and said to be
partly suppressed. Equation (21.116) corrects for
the effects of friction, contraction of the nappe,
unequal velocities in the channel of approach, and
partial suppression of the crest contraction and
includes a correction for the velocity of approach
and the associated velocity head.
To be fully ventilated, a nappe must have its
lower surface subjected to full atmospheric pres-
sure. A partial vacuum below the nappe can result
through removal of air by the overflowing jet if
there is restricted ventilation at the sides of the
weir. This lack of ventilation causes increased dis-
charge and a fluctuation and shape change of the
nappe. The resulting unsteady condition is very
objectionable when the weir is used as a measur-
ing device.
At very low heads, the nappe has a tendency to
adhere to the downstream face of a rectangular
weir even when means for ventilation are provid-
ed. A weir operating under such conditions could
not be expected to have the same relationship
between head and discharge as would a fully ven-
tilated nappe.
A V-notch weir (Fig. 21.60) should be used for
measurement of flow at very low heads if accuracy
of measurement is required.
End contractions occur when the weir open-
ing does not extend the full width of the
approach channel. Water flowing near the walls
must move toward the center of the channel to
pass over the weir, thus causing a contraction of
the flow. The nappe continues to contract as it
passes over the crest. Hence, below the crest, the
nappe has a minimum width less than the crest
length.
Fig. 21.60 V-notch weir.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.64
I
Section Twenty-One
The effective length L, ft, of a contracted-width
weir is given by
(21.117)
where L = measured length of crest, ft
N = number of end contractions
H = measured head, ft
If flow contraction occurs at both ends of a weir,
there are two end contractions and N = 2. If the
weir crest extends to one channel wall but not the
other, there is one end contraction and N = 1. The
effective crest length of a full-width weir is taken
as its measured length. Such a weir is said to have
its contractions suppressed.
21.34.3 Triangular or V-Notch
Sharp-Crested Weirs
The triangular or V-notch weir (Fig. 21.60) has a dis-
tinct advantage over a rectangular sharp-crested
weir (Art. 21.34.2) when low discharges are to be
measured. Flow over a V-notch weir starts at a
point, and both discharge and width of flow
increase as a function of depth. This has the effect of
spreading out the low-discharge end of the depth-
discharge curve and therefore allows more accurate
determination of discharge in this region.
Discharge is given by
(21.118)
where = notch angle
H = measured head, ft
C
1
= discharge coefficient
The head H is measured from the notch elevation
to the water-surface elevation at a distance 2.5H
upstream from the weir. Values of the discharge
coefficient were derived experimentally by Lenz,
who developed a procedure for including the
effect of viscosity and surface tension as well as the
effect of contraction and velocity of approach (A. T.
Lenz, Viscosity and Surface Tension Effects on V-
Notch Weir Coefficients, Transactions of the Ameri-
can Society of Civil Engineers, vol. 69, 1943). His val-
ues were summarized by Brater, who presented
the data in the form of curves (Fig. 21.61) (E. F.
Brater Handbook of Hydraulics, 6th ed.,
McGraw-Hill Book Company, New York).
A V-notch weir tends to concentrate or focus
the overflowing nappe, causing it to spring clear of
the downstream face for even the smallest flows.
This characteristic prevents a change in the head-
discharge relationship at low flows and adds mate-
rially to the reliability of the weir.
21.34.4 Trapezoidal Sharp-
Crested Weirs
The discharge from a trapezoidal weir (Fig. 21.62)
is assumed the same as that from a rectangular
weir and a triangular weir in combination.
(21.119)
where Q = discharge, ft
3
/s
L = length of notch at bottom, ft
H = head, measured from notch bottom, ft
Z = b/H [substituted for tan (/2) in Eq.
(21.118)]
Fig. 21.61 Chart gives discharge coefficients
for sharp-crested V-notch weirs. The coefficients
depend on head and notch angle.
Fig. 21.62 Trapezoidal sharp-crested weir.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.65
b = half the difference between lengths
of notch at top and bottom, ft
No data are available for determination of coeffi-
cients C
2
and C
3
. They must be determined experi-
mentally for each installation.
21.34.5 Submerged Sharp-Crested
Weirs
The discharge over a submerged sharp-crested
weir (Fig. 21.63) is affected not only by the head on
the upstream side H
1
but by the head downstream
H
2
. Discharge also is influenced to some extent by
the height P of the weir crest above the floor of the
channel.
The discharge Q
s
, ft
3
/s, for a submerged weir is
related to the free or unsubmerged discharge Q,
ft
3
/s, for that weir by a function of H
2
/H
1
. Ville-
monte expressed this relationship by the equation
(21.120)
where n is the exponent of H in the equation for
free discharge for the shape of weir used. (The
value of n is
3
/2 for a rectangular sharp-crested weir
and
5
/2 for a triangular weir.) To use the Villemonte
equation, first compute the rate of flow Q for the
weir when not submerged, and then, using this
rate and the required depths, solve for the sub-
merged discharge Q
s
. (J. R. Villemonte, Sub-
merged-Weir Discharge Studies, Engineering
News-Record, Dec. 25, 1947, p. 866.)
Equation (21.120) may be used to compute the
discharge for a submerged sharp-crested weir of
any shape simply by changing the value of n. The
maximum deviation from the Villemonte equation
for all test results was found to be 5%. Where great
accuracy is essential, it is recommended that the
weir be tested in a laboratory under conditions
comparable with those at its point of intended use.
21.34.6 Weirs Not Sharp-Crested
These are sturdy, heavily constructed devices, nor-
mally an integral part of hydraulic projects (Fig.
21.59). Typically, a weir not sharp-crested serves as
the crest section for an overflow dam or the
entrance section for a spillway or channel. Such a
weir can be used for discharge measurement, but
its purpose is normally one of control and regula-
tion of water levels or discharge, or both.
The discharge over a weir not sharp-crested is
given by
(21.121)
where Q = discharge, ft
3
/s
C = coefficient of discharge
L = effective length of crest, ft
H
t
= total head on crest including veloci-
ty head of approach, ft
The head of water producing discharge over a
weir is the total of measured head H and velocity
head of approach H

. The velocity head of


approach is accounted for by the discharge coeffi-
cient for sharp-crested weirs but must be consid-
ered separately for weirs not sharp-crested. Thus,
for such weirs, Eq. (21.115) is rewritten in the form
(21.122)
where H = measured head, ft
V = velocity of approach, ft/s
V
2
/2g = H

, velocity head of approach, ft,


neglecting degree of turbulence
given by Eq. (21.81)
g = acceleration due to gravity, 32.2 ft/s
2
Since velocity and discharge are dependent on
each other in this equation and both are unknown,
discharge must be found by a series of approxima-
tions, which may be done as follows: First, com-
pute a trial discharge from the measured head,
neglecting the velocity head. Then, using this dis-
charge, compute the velocity of approach, velocity
head, and finally total head. From this total head,
Fig. 21.63 Submerged sharp-crested weir.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.66
I
Section Twenty-One
compute the first corrected discharge. This correct-
ed discharge will be sufficiently accurate if the
velocity of approach is small. But the process
should be repeated, starting with the corrected dis-
charge, where approach velocities are high.
The discharge coefficient C must be determined
empirically for weirs not sharp-crested. If a weir of
untested shape is to be constructed, it must be cal-
ibrated in place or a model study made to deter-
mine its head-discharge relationship. The problem
of establishing a fixed relation between head and
discharge is complicated by the fact that the nappe
may assume a variety of shapes in passing over the
weir. For each change of nappe shape, there is a
corresponding change in the relation between
head and discharge. The effect is most critical for
low heads. A nappe undergoes several changes in
succession as the head varies, and the successive
shapes that appear with an increasing stage may
differ from those pertaining to similar stages with
decreasing head. Therefore, care must be exercised
when using these weirs for flow measurement to
ensure that the conditions are similar to those at
the time of calibration.
Large weirs not sharp-crested often have piers
on their crest to support control gates or a road-
way. These piers reduce the effective length of
crest by more than the sum of their individual
widths because of the formation of flow contrac-
tions at each pier. The effective crest length for a
weir not sharp-crested is given by
(21.123)
where L = effective crest length, ft
L = net crest lengths, ft = measured
length minus width of all piers
N = number of piers
K
p
= pier-contraction coefficient
K
a
= abutment-contraction coefficient
H
t
= total head on crest including veloci-
ty head of approach, ft
(U.S. Department of the Interior, Design of Small
Dams, Government Printing Office, Washington,
DC 20402.)
The pier-contraction coefficient K
p
is affected by
the shape and location of the pier nose, thickness
of pier, head in relation to design heads, and
approach velocity. For conditions of design head
H, the average pier-contraction coefficients are as
shown in Table 21.12.
The abutment-contraction coefficient K
a
is
affected by the shape of the abutment, the angle
between the upstream approach wall and the axis
of flow, the head in relation to the design head,
and the approach velocity. For conditions of design
head H
d
, average coefficients may be assumed as
shown in Table 21.13.
21.34.7 Submergence of Weirs Not
Sharp-Crested
Spillways and other weirs not sharp-crested are
submerged when their tailwater level is high
enough to affect their discharge. Because of the
surface disturbance produced in the vicinity of the
crest, such a spillway or weir is unsatisfactory for
accurate flow measurement.
Approximate values of discharge may be found
by applying the following rules proposed by E. F.
Brater: (1) If the depth of submergence is not
greater than 0.2 of the head, ignore the submer-
gence and treat the weir as though it had free dis-
charge. (2) For narrow weirs having a sharp
upstream leading edge, use a submerged-weir for-
mula for sharp-crested weirs. (3) Broad-crested
Condition K
p
Square-nosed piers with corners rounded on a
radius equal to about 0.1 of the pier thickness 0.02
Rounded-nosed piers 0.01
Pointed-nosed piers 0
Table 21.12 Pier-Contraction Coefficients
Condition K
a
Square abutment with headwall at 90 to
direction of flow 0.20
Rounded abutments with headwall at 90 to
direction of flow when 0.5H
d
> r* > 0.15H
d
0.10
Rounded abutments where r* > 0.5H
d
and
headwall is placed not more than 45 to
direction of flow 0
*r = radius of abutment rounding.
Table 21.13 Abutment-Contraction Coefficients
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.67
weirs are not affected by submergence up to
approximately 0.66 of the head. (4) For weirs with
narrow rounded crests, increase discharge
obtained by a formula for submerged sharp-crest-
ed weirs by 10% or more. Of the above rules, 1, 2,
and 3 probably apply quite accurately, while 4 is
simply a rough approximation.
21.34.8 The Ogee-Crested Weir
The ogee-crested weir was developed in an attempt
to produce a weir that would not have the undesir-
able nappe variation normally associated with weirs
not sharp-crested. A shape was needed that would
force the nappe to assume a single path for any dis-
charge, thus making the weir consistent for flow
measurement. The ogee-crested weir (Fig. 21.64) has
such a shape. Its crest profile conforms closely to the
profile of the lower surface of a ventilated nappe
flowing over a rectangular sharp-crested weir.
The shape of this nappe, and therefore of an
ogee crest, depends on the head producing the dis-
charge. Consequently, an ogee crest is designed for
a single total head, called the design head H
d
.
When an ogee weir is discharging at the design
head, the flow glides over the crest with no inter-
ference from the boundary surface and attains
near-maximum discharge efficiency.
For flow at heads lower than the design head,
the nappe is supported by the crest and pressure
develops on the crest that is above atmospheric but
less than hydrostatic. This crest pressure reduces
the discharge below that for ideal flow. (Ideal flow
is flow over a fully ventilated sharp-crested weir
under the same head H.)
When the weir is discharging at heads greater
than the design head, the pressure on the crest is
less than atmospheric, and the discharge increases
over that for ideal flow. The pressure may become
so low that separation in flow will occur. According
to Chow, however, the design head may be safely
exceeded by at least 50% before harmful cavitation
develops (V. T. Chow, Open-Channel Hydraulics,
McGraw-Hill Book Company, New York).
The measured head H on an ogee-crested weir
is taken as the distance from the highest point of
the crest to the level of the water surface at a dis-
tance 2.5H upstream. This depth coincides with
the depth measured between the upstream water
level and the bottom of the nappe, at the point of
maximum contraction, for a sharp-crested weir.
This relationship is shown in Fig. 21.65.
Discharge coefficients for ogee-crested weirs are
therefore determined from sharp-crested-weir coef-
ficients after an adjustment for this difference in
head. These coefficients are a function of the
approach velocity, which varies with the ratio of
height of weir P to actual total head H
t
, where dis-
charge is given by Eq. (21.122). Figure 21.66 for an
ogee weir with a vertical upstream face gives coeffi-
cient C
d
for discharge at design head H
d
. (U.S.
Department of the Interior, Design of Small Dams,
Government Printing Office, Washington, DC
20402. This manual and V. T. Chow, Open-Channel
Fig. 21.64 Ogee-crested weir with vertical
upstream face.
Fig. 21.65 Location of origin of coordinates for
sharp-crested and ogee-crested weirs.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.68
I
Section Twenty-One
Hydraulics, McGraw-Hill Book Company, New
York, present methods for determining the shape of
an ogee crest profile.) When the weir is discharging
at other than the design head, the flow differs from
ideal, and the discharge coefficient changes from
the discharge coefficient given in Fig. 21.66.
Figure 21.67 gives values of the discharge coef-
ficient C as a function of the ratio H
t
/ H
d
, where H
t
is the actual head being considered and H
d
is the
design head.
Fig. 21.66 Chart gives discharge coefficients at design head H
d
for vertical-faced ogee-crested weirs.
(From Design of Small Dams, U.S. Bureau of Reclamation.)
Fig. 21.67 Chart gives discharge coefficients for vertical-faced ogee-crested weirs at heads H
t
other
than design head H
d
. (From Design of Small Dams, U.S. Bureau of Reclamation.)
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.69
If an ogee weir has a sloping upstream face,
there is a tendency for an increase in discharge over
that for a weir with a vertical face. Figure 21.68
shows the ratio of the coefficient for an ogee weir
with a sloping face to the coefficient for a weir with
a vertical upstream face. The coefficient of discharge
for an ogee weir with a sloping upstream face, if
flow is at other than the design head, is determined
from Fig. 21.66 and is then corrected for head and
slope with Figs. 21.67 and 21.68.
21.34.9 Broad-Crested Weir
This is a weir with a horizontal or nearly horizontal
crest. The crest must be sufficiently long in the direc-
tion of flow that the nappe is supported and hydro-
static pressure developed on the crest for at least a
short distance. A broad-crested weir is nearly rectan-
gular in cross section. Unless otherwise noted, it will
be assumed to have vertical faces, a plane horizontal
crest, and sharp right-angled edges.
Figure 21.69 shows a broad-crested weir that,
because of its sharp upstream edge, has contrac-
tion of the nappe. This causes a zone of reduced
pressure at the leading edge. When the head H on
a broad-crested weir reaches one to two times its
breadth b, the nappe springs free, and the weir acts
as a sharp-crested weir.
Discharge over a broad-crested weir is given by
Eq. (21.115) since the velocity of approach was
ignored in experiments performed to determine
the coefficient of discharge. These coefficients prob-
ably apply more accurately, therefore, where the
velocity of approach is not high. Values of the dis-
charge coefficient, compiled by King, appear in
Table 21.14. (E. F. Brater, Handbook of Hydraulics,
6th ed., McGraw-Hill Book Company, New York.)
21.34.10 Weirs of Irregular Section
This group includes those weirs whose cross sec-
tion deviates from typical broad-crested or ogee-
crested weirs. Weirs of irregular section, fairly com-
mon in waterworks projects, are used as spillways
and control structures. Experimental data are
available on the more common shapes. (See, for
example, E. F. Brater, Handbook of Hydraulics,
6th ed., McGraw-Hill Book Company, New York.)
Fig. 21.68 Chart gives design coefficients at design head H
d
for ogee-crested weirs with sloping
upstream face. (From Design of Small Dams, U.S. Bureau of Reclamation.)
Fig. 21.69 Broad-crested weir.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.70
I
Section Twenty-One
21.35 Sediment Transfer and
Deposition in Open
Channels
Sediment from open channels has many undesir-
able effects: Reservoirs have a reduced useful life
because of loss of storage through the accumula-
tion of silt. Sediment causes a hazard in navigable
channels and harbors and an increase in frequency
of flooding due to aggravation of rivers and flood
channels. Silting of arable land by flooding rivers
destroys fertility when the silt originates from
bank or gully erosion rather than from surface, or
soil, erosion. The cost of operating irrigation sys-
tems is increased by the need for frequent dredg-
ing. Water-supply facilities have increased costs
because of the necessity of providing desilting
works and because of the wear on mechanical
equipment, such as gates, valves, and turbines.
21.35.1 Sediment Deposition in
Reservoirs
Deposition of silt results when the transporting
forces of a river are dissipated as the river enters a
body of still water, such as a reservoir. Heavier silt
sizes, those forming the bed load, are deposited in
a delta as the river enters calm water. The smaller
silt sizes, those carried in suspension, travel farther
into the reservoir before deposition.
This incoming water, with its load of suspended
silt, has a specific gravity greater than that of the
clear water in the reservoir and may form a density
current, rather than mixing immediately with the
clear water. A density current, once formed, quickly
moves to the bottom and flows in a dense cloud
down the slopes of the reservoir until it is blocked
by a dam. The dense flow then spreads out in this
deeper area, where the stilling effect of the basin
eventually causes deposition of the sediment.
Deposits of fine sediment form about one-third of
the volume of silt deposits in a reservoir. Much of all
of this fine sediment is transported to its final loca-
tion by density currents. The visible delta formed by
the coarse sediments frequently distracts attention
from the unseen bottom deposits of fine sediment,
which are often of equal consequence.
Most reservoirs trap from 70 to almost 100% of
the incoming sediment, depending on whether
the reservoir is used for flood control or storage.
Breadth of crest of weir, ft
0.50 0.75 1.00 1.50 2.00 2.50 3.00 4.00 5.00 10.00 15.00
0.2 2.80 2.75 2.69 2.62 2.54 2.48 2.44 2.38 2.34 2.49 2.68
0.4 2.92 2.80 2.72 2.64 2.61 2.60 2.58 2.54 2.50 2.56 2.70
0.6 3.08 2.89 2.75 2.64 2.61 2.60 2.68 2.69 2.70 2.70 2.70
0.8 3.30 3.04 2.85 2.68 2.60 2.60 2.67 2.68 2.68 2.69 2.64
1.0 3.32 3.14 2.98 2.75 2.66 2.64 2.65 2.67 2.68 2.68 2.63
1.2 3.32 3.20 3.08 2.86 2.70 2.65 2.64 2.67 2.66 2.69 2.64
1.4 3.32 3.26 3.20 2.92 2.77 2.68 2.64 2.65 2.65 2.67 2.64
1.6 3.32 3.29 3.28 3.07 2.89 2.75 2.68 2.66 2.65 2.64 2.63
1.8 3.32 3.32 3.31 3.07 2.88 2.74 2.68 2.66 2.65 2.64 2.63
2.0 3.32 3.31 3.30 3.03 2.85 2.76 2.72 2.68 2.65 2.64 2.63
2.5 3.32 3.32 3.31 3.28 3.07 2.89 2.81 2.72 2.67 2.64 2.63
3.0 3.32 3.32 3.32 3.32 3.20 3.05 2.92 2.73 2.66 2.64 2.63
3.5 3.32 3.32 3.32 3.32 3.32 3.19 2.97 2.76 2.68 2.64 2.63
4.0 3.32 3.32 3.32 3.32 3.32 3.32 3.07 2.79 2.70 2.64 2.63
4.5 3.32 3.32 3.32 3.32 3.32 3.32 3.32 1.88 2.74 2.64 2.63
5.0 3.32 3.32 3.32 3.32 3.32 3.32 3.32 3.07 2.79 2.64 2.63
5.5 3.32 3.32 3.32 3.32 3.32 3.32 3.32 3.32 2.88 2.64 2.63
Table 21.14 Values of C in Q = CLH
3/2
for Broad-Crested Weirs
Measured
head
H, ft
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.71
Flood-control reservoirs are normally emptied
shortly after a storm, so the suspended materials
are carried out with the water before settling can
occur. This procedure reduces new deposits by
almost 30% after each storm. Storage reservoirs
used for water supply or power generation pur-
poses, on the other hand, normally retain any
inflow long enough for settlement of all suspend-
ed matter to occur. Their discharges are regulated
to allow generation of power or to produce a uni-
form flow downstream with no thought to the
venting of silt-laden storm flows.
The greater part of the annual suspended silt
load in a stream may be carried in a relatively short
time. The stream runs comparatively clear during
the remainder of the year.
Venting of much of the annual suspended silt
load is feasible through the use of density currents.
These currents are stable, once formed, and often
extend to the reservoir outlet. If density currents
are observed and their time of arrival at the outlet
determined, appropriate gates can be opened and
much of the fine sediment entering a storage reser-
voir can be vented before it has time to form per-
manent deposits. This venting operation can
extend the life of a reservoir by many years.
Numerous phenomena can destroy a reservoir,
such as loss of storage capacity by landslide and
loss of the dam by earthquake, landslide, overtop-
ping, or failure of materials. The most common
manner of destruction, however, is through loss of
storage by deposition of silt. Redemption of reser-
voir capacity lost through silting is almost always
economically unfeasible because of the wide distri-
bution of deposits in a reservoir and the large
quantity present. The most practicable means of
avoiding a loss in reservoir capacity are to prevent
formation of permanent deposits by taking advan-
tage of density currents and to control rate of sed-
iment production from eroding areas. When nei-
ther can be done, sufficient storage space must be
provided in the design of the reservoir to compen-
sate for depletion by silting during a reasonable
economic lifetime.
Sediment production and its transportation to
reservoirs or navigable waters cannot be prevent-
ed at costs proportionate to the resulting benefits.
However, nature may be economically improved
at times through a program of erosion control,
reducing sediment production to less than that
normally found under virgin conditions.
The deposits of silt that form in a storage reser-
voir are categorized into two distinct types: Delta
deposits, formed from the bed load, are coarse-
grained, with an in-place weight of about 80 lb/ft
3
.
Deposits produced from the suspended load are
fine-grained, with an average weight of about 30
lb/ft
3
. They constitute about one-sixth of the total
weight of sediment delivered but account for
about one-third of the volume of all deposits in a
storage reservoir because of their low density. If
sediment deposits are periodically above water,
because of fluctuations in the reservoir water level,
their density increases and the volume ratios given
above for continued submergence no longer apply.
21.35.2 Prediction of Sediment-
Delivery Rate
Two methods of approach are available for pre-
dicting the rate of sediment accumulation in a
reservoir; both involve predicting the rate of sedi-
ment delivery.
One approach depends on historical records of
the silting rate for existing reservoirs and is purely
empirical. By this method, the silting records of a
reservoir may be used to predict either the silting
rate for that reservoir or the probable pattern of
silt accumulation for a proposed reservoir in a
similar area. This method allows transposition of
data from one watershed to another because the
measured annual sediment accumulation of a
reservoir is expressed as a rate of sediment deliv-
ery per unit area of its watershed. Of course, the
rate is not uniform during the year, or from year to
year, because of variations in rainfall, but it should
average the computed annual amount over the
life of the project. The annual silt accumulation in
a reservoir is determined by surveying exposed
deltas and taking depth soundings. The resulting
volume is adjusted to account for any silt loss
through sluice gates or over the spillway and is
then expressed as silt delivery per square mile of
drainage area. This silt-delivery figure is further
adjusted for rainfall and runoff conditions, to give
a figure that could reasonably be expected during
a year of average rainfall. If this adjusted figure is
to be transposed to a neighboring drainage basin,
adjustments should be made to account for both
soil cover and rainfall differences between the
basins. (For a discussion of the factors upon which
this adjustment is based, see Art. 21.39.)
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.72
I
Section Twenty-One
Silt-delivery measurements or estimates do not
give total silt production for an area because part of
the silt produced in a basin is deposited on flood-
plains and in channels before it reaches a reservoir.
The difference between the amount of silt pro-
duced and that delivered increases as the size of the
drainage area increases because of the increased
chance that the silt will be deposited before it
reaches the reservoir. Therefore, if a silt-delivery
measurement or estimate is to be transposed to a
basin of different size, an adjustment should be
made to account for this discrepancy as well. The
information for this adjustment can come only
from field reconnaissance of the two areas to deter-
mine differences that might account for a variation
in deposition of silt along the water courses.
The second general method of calculating sedi-
ment-delivery rate involves determining the rate
of sediment transport as a function of stream dis-
charge and density of suspended silt. The total
sediment inflow for the year is then computed
from these relationships and the recorded stream-
discharge data.
The total quantity of sediment carried by a river
is assumed transported either as suspended load or
as bed load (Art. 21.35.1). The division is based on
particle size but depends on velocity of flow as well.
There is no sharp line of demarcation between the
two classes. According to Witzig, about 80% of the
volume of all sediment is produced by streambank
erosion; the remaining 20% is produced by land-
surface erosion. Constant erosion of the stream-
banks keeps the streambed well supplied with the
coarse silt that travels as bed load. The fine silt that
travels in suspension is produced in small amounts
by streambank erosion. But for the most part, this
silt comes from land-surface erosion, which gener-
ally occurs only during a storm.
The total quantity of sediment in suspension is
not necessarily related directly to discharge at all
times. The quantity is affected by seasonal varia-
tions in the supply and source of fine sediment
and by distribution of rainfall and runoff from the
watershed. Therefore, measurement of the sedi-
ment load for a given discharge does not necessar-
ily indicate the amount that may be carried by an
equal discharge at another time.
The bed load consists of the silt particles too
large to be held in suspension. This size range
includes particles of coarse sand, gravel, and boul-
ders. The bed-load particles are moved by rolling
along the bed of the stream. Some of the finer bed-
load particles are moved in a series of steps or
jumps representing a transition between trans-
portation as bed load and suspended load.
The quantity of bed load is considered a con-
stant function of the discharge because the sedi-
ment supply for the bed-load forces is always
available in all but lined channels. An accepted for-
mula for the quantity of sediment transported as
bed load is the Schoklitsch formula:
(21.124)
where G
b
= total bed load, lb/s
D
g
= effective grain diameter, in
S = slope of energy gradient
Q
i
= total instantaneous discharge, ft
3
/s
b = width of river, ft
q
o
= critical discharge, ft
3
/s per ft of river
width
= (0.00532/S
4/3
)D
g
An approximate solution for bed load by the
Schoklitsch formula can be made by determining
or assuming mean values of slope, discharge, and
a single grain size representative of the bed-load
sediment. A mean grain size of 0.04 in in diameter
(about 1 mm) is reasonable for a river with a slope
of about 1.0 ft/mi.
The size of grains moving on the bed of a river
depends on velocity of flow, which varies with
both slope and discharge. Therefore, the mean
grain size changes as the flow increases during a
storm or as the river changes slope along its course.
It is obvious that considerable error could result
from the use of Eq. (21.124) if it is necessary to
guess at a mean grain diameter in the absence of
carefully collected field data. Frequently, however,
if insufficient data or lack of money prevent more
thorough investigations, this shortcut can give
results of sufficient accuracy.
Numerous formulas have been developed to
represent the condition of flow involved in trans-
portation of suspended sediment. These formulas
express the degree of turbulent energy involved in
suspension of the sediment and the mode of transfer
of this energy to the silt and other fluid particles. The
formulas require a number of empirical constants
but are based on a sound physical and rational foun-
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.73
dation. They require information as to the sediment
composition by grain size, the actual quantity of silt
in suspension at a given depth, and the stream
velocity. (See H. A. Einstein, The Bed-Load Func-
tion for Sediment Transportation in Open-Channel
Flows, U.S. Department of Agriculture.)
An approximate determination of suspended
load may be made without using these complicat-
ed formulas. The weight of suspended sediment
transported by a river in an average year normally
equals about 20% of the weight transported as bed
load. The total weight of material annually moved
by a river is therefore equal to 120% of the weight
of material transported as bed load during the year
as computed from Eq. (21-124).
(W. H. Graf, Hydraulics of Sediment Trans-
port, McGraw-Hill Book Company, New York.)
21.36 Erosion Control
The various methods used in erosion control are
collectively called upstream engineering. They con-
sist of soil conservation measures such as refor-
estation, check-dam construction, planting of
burned-over areas, contour plowing, and regula-
tion of crop and grazing practices. Also included
are measures for proper treatment of high
embankments and cuts and stabilization of stream-
banks by planting or by revetment construction.
One phase of reforestation that may be applied
near a reservoir is planting of vegetation screens.
Such screens, planted on the flats adjacent to the
normal stream channel at the head of a reservoir,
reduce the velocity of silt-laden storm inflows that
inundate these areas. This stilling action causes
extensive deposition to occur before the silt reach-
es the main cavity of the reservoir. Use of vegeta-
tion screens, debris barriers, or desilting basins
above a reservoir should be planned with future
development in mind. For instance, if the dam is
raised at a later date, the accumulated silt in this
area would detract from the added storage that
might otherwise have been obtained.
Hydrology
Hydrology is the study of the waters of the earth,
their occurrence, circulation, and distribution,
their chemical and physical properties, and their
reaction with their environment, including their
relation to living things. A major concern is the cir-
culation, on or near the land surface, of water and
its constituents throughout the hydrologic cycle. In
this cycle, water evaporation from oceans, rivers,
lakes, and other sources is carried over the earth
and precipitated as rain or snow. The precipitation
forms runoff on the land, infiltrates into the soil,
recharges groundwater, discharges into streams,
and then flows into the oceans and lakes, from
which evaporation restarts the cycle. Thus hydrol-
ogy deals with precipitation, evaporation, infiltra-
tion, groundwater flow, runoff, and stream flow
21.37 Precipitation
The primary concern with precipitation in water
resources engineering is forecasting it. The means
for doing so are based on either current or past
data, or a combination of the two.
Current data, in the form of synoptic weather
charts, are published daily by the U.S. Weather
Bureau. These charts summarize the various mete-
orological factors, such as wind, temperature, and
pressure, through whose interaction precipitation
is produced.
Past data are primarily in the form of rainfall
records for a standard period, such as an hour, day,
or year. They are the major source of data for deter-
mination of the recurrence interval for storms of a
definite magnitude and the magnitude of storms
in a definite recurrence interval.
Rainfall records are obtained from rain gages,
which are of two types. The first type is a record-
ing or automatic gage. It continually records, by
ink pen and revolving drum, or digital microchip
technology, the variation in rainfall intensity as
well as the total rainfall volume. The second type is a
nonrecording gage; it measures only the total rain
volume that fell during the period between observa-
tions. The standard observation time for nonrecord-
ing gages for the U.S. Weather Bureau is 24 h.
Corrections must be made to rain-gage records
to account for the mean precipitation over the
entire drainage basin, for hourly rainfall rates when
only daily volumes are given, and for errors arising
out of the location of the gage. Most methods used
in runoff determinations are based on the assump-
tion that rainfall is uniform over the entire drainage
basin. This necessitates development of a correc-
tion factor to balance out the rainfall variation
caused by various topographical features in the
watershed. Rain gages tend to give rainfall volumes
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.74
I
Section Twenty-One
that are too small. This error is caused by the move-
ment of wind around the gage, and it increases as
wind velocity increases. This windage error is
much more pronounced when the rain gage is near
the top or bottom of a cliff or near other big obstruc-
tions. Care must be exercised in placement of rain
gages to ensure accuracy.
The probable maximum precipitation is the
greatest rainfall intensity or volume that could
ever be expected to occur in a specific drainage
basin. This rainfall magnitude is frequently used as
the design storm for major hydraulic structures to
serve the basin when the rainfall records are short
and extrapolation to the desired design-storm fre-
quency could be grossly inaccurate. The magni-
tude of probable maximum precipitation is based
on simultaneous occurrence of the maximum val-
ues of the meteorological factors that combine to
form precipitation. The two most important factors
are wind and air-mass moisture content. An idea of
the magnitude of the probable maximum precipi-
tation can also be obtained by transposing the
greatest rainfall that has occurred in a meteorolog-
ically homogeneous region. For methods for deter-
mining the probable maximum precipitation, see
D. R. Maidment, Handbook of Hydrology,
McGraw-Hill, Inc., New York.
Not all rain reaches the ground. A portion may
evaporate as it falls, while another portion may be
caught on leaves, branches, and other vegetation
surfaces. This phenomenon, called interception, is
a loss from a runoff standpoint since the rain
evaporates and never reaches the ground. Inter-
ception may be significant for small-intensity
storms occurring with little or no wind over an
area with heavy vegetation growth.
21.38 Evaporation and
Transpiration
These are processes by which moisture is returned
to the atmosphere. In evaporation, water changes
from liquid to gaseous form. In transpiration,
plants give off water vapor during synthesis of
plant tissue.
Evapotranspiration, commonly termed con-
sumptive use, refers to the total evaporation from
all sources such as free water, ground, and plant-
leaf surfaces. On an annual basis, the consumptive
use may vary from 15 in/year for barren land to 35
in/year for heavily forested areas and 40 in/year in
tropical and subtropical regions. Evapotranspira-
tion is important because, on a long-term basis, pre-
cipitation minus evapotranspiration equals runoff.
Evaporation may occur from free-water, plant,
or ground surfaces. Of the three, free-water sur-
face evaporation is usually the most important. It
must be considered in the design of a reservoir,
especially if the reservoir is shallow, has a relative-
ly large surface area, and is located in a semiarid or
arid region. Evaporation is a direct function of the
wind and temperature and an inverse function of
atmospheric pressure and amount of soluble solids
in the water.
The rate of evaporation is dependent on the
vapor-pressure gradient between the water sur-
face and the air above it. This relation is known as
Daltons law. The Meyer equation [Eq. (21.125)],
developed from Daltons law, is one of many evap-
oration formulas and is popular for making evapo-
ration-rate calculations.
(21.125)
(21.126)
where E = evaporation rate, in 30-day month
C = empirical coefficient, equal to 15 for
small, shallow pools and 11 for
large, deep reservoirs
e
w
= saturation vapor pressure, in of mer-
cury, corresponding to monthly
mean air temperature observed at
nearby stations for small bodies of
shallow water or corresponding to
water temperature instead of air
temperature for large bodies of deep
water
e
a
= actual vapor pressure, in of mercury,
in air based on monthly mean air
temperature and relative humidity
at nearby stations for small bodies of
shallow water or based on informa-
tion obtained about 30 ft above the
water surface for large bodies of
deep water
w = monthly mean wind velocity, mi/h
at about 30 ft above ground
= wind factor
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.75
As an example of the evaporation that may occur
from a large reservoir, the mean annual evapora-
tion from Lake Mead is 6 ft.
Evaporation from free-water surfaces is usually
measured with an evaporation pan. This pan is a
standard size and is located on the ground near
the body of water whose evaporation is to be
determined. The depth of water in this pan is
checked periodically and corrections made for fac-
tors other than evaporation that may have raised
or lowered the water surface. A pan coefficient is
then applied to the measured pan evaporation to
get the reservoir evaporation.
The standard evaporation pan of the National
Weather Service, called a Class A Level Pan, is in
widespread use. It is 4 ft in diameter and 10 in
deep. It is positioned 6 in above the ground. Its pan
coefficient is commonly taken as 0.70, although it
may vary between 0.60 and 0.80, depending on the
geographical region. Annual evaporation from the
pan ranges from 25 in in Maine and Washington to
120 in along the Texas-Mexico and California-Ari-
zona borders.
Evaporation rates from reservoirs may be
reduced by spreading thin molecular films on the
water surface. Hexadeconal, or cetyl alcohol, is one
such film that has been effective on small reser-
voirs where there is little wind. On large reser-
voirs, wind tends to push the film to the shore.
Since hexadeconal is removed by wind, birds,
insects, aquatic life, and biologic attrition, it must
be applied periodically for maximum effectiveness.
Hexadeconal appears to have no adverse effects on
either humans or wildlife.
Evaporation from ground surfaces is usually of
minor importance, except in arid, tropical, and
subtropical regions having high water tables and
where it pertains to the determination of initial
soil-moisture conditions in a runoff analysis.
(D. R. Maidment, Handbook of Hydrology,
McGraw-Hill, Inc., New York.)
21.39 Runoff
This is the residual precipitation remaining after
interception and evapotranspiration losses have
been deducted. It appears in surface channels,
natural or manmade, whose flow is perennial or
intermittent. Classified by the path taken to a
channel, runoff may be surface, subsurface, or
groundwater flow.
Surface flow moves across the land as overland
flow until it reaches a channel, where it continues
as channel or stream flow. After joining stream flow,
it combines with the other runoff components in
the channel to form total runoff.
Subsurface flow, also known as interflow, subsur-
face runoff, subsurface storm flow, and storm seepage,
infiltrates only the upper soil layers without joining
the main groundwater body. Moving laterally, it
may continue underground until it reaches a chan-
nel or returns to the surface and continues as over-
land flow. The time for subsurface flow to reach a
channel depends on the geology of the area. Com-
monly, it is assumed that subsurface flow reaches a
channel during or shortly after a storm. Subsurface
flow may be the major portion of total runoff for
moderate or light rains in arid regions since surface
flow under those conditions is reduced by unusual-
ly high evaporation and infiltration.
Groundwater flow, or groundwater runoff, is
that flow supplied by deep percolation. It is the
flow of the main groundwater body and requires
long periods, perhaps several years, to reach a
channel. Groundwater flow is responsible for the
dry-weather flow of streams and remains practi-
cally constant during a storm. Groundwater flow is
primarily the concern of water-supply engineers.
Surface and subsurface flow are of interest to
flood-control engineers.
In practice, direct runoff and base flow are the
only two divisions of runoff used. The basis for this
classification is travel time rather than path. Direct
runoff leaves the basin during or shortly after a
storm, whereas base flow from the storm may not
leave the basin for months or even years.
Runoff is supplied by precipitation. The portion
of precipitation that contributes entirely to direct
runoff is called effective precipitation, or effective rain
if the precipitation is rain. That portion of the pre-
cipitation which contributes entirely to surface
runoff is called excess precipitation, or excess rain.
Thus, effective rain includes subsurface flow,
whereas excess rain is only surface flow.
The two major characteristics that affect runoff
are climatic and drainage-basin factors. The num-
ber of factors is an indication of the complexity of
accurately determining runoff:
1. Climatic characteristics
a. Precipitationform (rain, hail, snow, frost,
dew), intensity, duration, time distribution,
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.76
I
Section Twenty-One
seasonal distribution, areal distribution,
recurrence interval, antecedent precipitation,
soil moisture, direction of storm movement
b. Temperaturevariation, snow storage,
frozen ground during storms, extremes dur-
ing precipitation
c. Windvelocity, direction, duration
d. Humidity
e. Atmospheric pressure
f. Solar radiation
2. Drainage-basin characteristics
a. Topographicsize, shape, slope, elevation,
drainage net, general location, land use and
cover, lakes and other bodies of water, artifi-
cial drainage, orientation, channels (size,
shape of cross section, slope, roughness,
length)
b. Geologicsoil type, permeability, ground-
water formations, stratification
21.40 Sources of Hydrologic
Data
The importance of exhausting all possible sources
of hydrologic data, both published and unpub-
lished, as the first step in design of a hydraulic
project cannot be overemphasized. The majority
of hydrologic data is collected and published by
government agencies, those of the Federal gov-
ernment being the largest and most important.
The principal source of precipitation data is the
U.S. Weather Bureau. Its extensive system of gages
supplies complete precipitation data as well as all
other types of hydrologic data. These data are
compiled and presented in monthly and yearly
summaries in the Bureaus Climatological Data.
In addition to the monthly and yearly summaries,
special-interest items, such as rainfall intensity for
various durations and recurrence intervals, are
published in Weather Bureau technical papers.
Other sources are Water Bulletins of the Interna-
tional Boundary Commission, the U.S. Agricultur-
al Research Service, and various state and local
agencies.
The principal source of runoff data is the Water
Supply Papers of the U.S. Geological Survey. These
papers contain records of daily flow, mean flow,
yearly flow volume, extremes of flow, and statisti-
cal data pertaining to the entire record. Also
included in the Papers are lists of reports covering
unusually large floods and records of discharge
collected by agencies other than the U.S. Geologi-
cal Survey. The Water Supply Papers are published
yearly in 14 parts; each part is for an area whose
boundaries coincide with natural-drainage fea-
tures, as shown in Fig. 21.70.
Other agencies that collect and publish stream-
flow and flood records are the Corps of Engineers,
TVA, International Boundary Commission, and
Weather Bureau. The Corps of Engineers publishes
data on floods in which loss of life and extensive
property damage occurred. Less obvious sources
of stream-flow data are water-right decrees by dis-
trict courts, county records of water-right filings
and State Engineer permits, and annual reports of
various interstate-compact commissions.
21.41 Methods for Runoff
Determinations
The method selected to determine runoff depends
on its applicability to the area of concern, the quan-
tity and type of data available, the detail required
in the final answer, and the accuracy desired.
Applicability depends on the characteristics of the
particular area and the assumptions from which
the method was developed. Quantity and type of
data available refer to the length, detail, and com-
pleteness of the hydrologic records, which may be
either precipitation or stream flow. An example of
the variation of detail in the final result may be
Fig. 21.70 Drainage subdivisions of the United
States for stream-flow records published in Water
Supply Papers, U.S. Geological Survey.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.77
found in the determination of flood runoff. Sever-
al methods yield only peak discharge; others give
the complete hydrograph. Accuracy is limited by
the cost of performing analyses and assumptions
made in the development of a method.
The methods that follow are a convenient
means for solving typical runoff problems encoun-
tered in water resources engineering. One method
pertains to minor hydraulic structures, the second
to major hydraulic structures. A minor structure is
one of low cost and of relatively minor importance
and presents small downstream damage potential.
Typical examples are small highway and railroad
culverts and low-capacity storm drains. Major
hydraulic structures are characterized by their
high cost, great importance, and large downstream
damage potential. Typical examples of major
hydraulic structures are large reservoirs, deep cul-
verts under vital highways and railways, and high-
capacity storm drains and flood-control channels.
21.41.1 Method for Determining
Runoff for Minor Hydraulic
Structures
The most common means for determining runoff for
minor hydraulic structures is the rational formula
(21.127)
where Q = peak discharge, ft
3
/s
C = runoff coefficient = percentage of
rain that appears as direct runoff
I = rainfall intensity, in/h
A = drainage area, acres
The assumptions inherent in the rational formula
are:
1. The maximum rate of runoff for a particular
rainfall intensity occurs if the duration of rain-
fall is equal to or greater than the time of con-
centration. The time of concentration is com-
monly defined as the time required for water to
flow from the most distant point of a drainage
basin to the point of flow measurement.
2. The maximum rate of runoff from a specific
rainfall intensity whose duration is equal to or
greater than the time of concentration is direct-
ly proportional to the rainfall intensity.
3. The frequency of occurrence of the peak dis-
charge is the same as that of the rainfall intensi-
ty from which it was calculated.
4. The peak discharge per unit area decreases as
the drainage area increases, and the intensity of
rainfall decreases as its duration increases.
5. The coefficient of runoff remains constant for
all storms on a given watershed.
Since these assumptions apply reasonably well
for urbanized areas with simple drainage facilities
of fixed dimensions and hydraulic characteristics,
the rational formula has gained widespread use in
the design of drainage systems for these areas. Its
simplicity and ease of application have resulted in
its being used for more complex urban systems
and rural areas where the assumptions are not so
applicable.
The rational formula is criticized for expressing
runoff as a fraction of rainfall rather than as rainfall
minus losses and for combining all the complex
factors that affect runoff into a single coefficient.
Although these and similar criticisms are valid, use
of a more complicated formula is not justified
because the time and money spent to obtain the
necessary data would not be warranted for minor
hydraulic structures.
Numerous refinements have been developed
for the runoff coefficient. As an example, the Los
Angeles County Flood Control District gives
runoff coefficients as a function of the soil and area
type and of the rainfall intensity for the time of
concentration. Other similar refinements are possi-
ble if the resources are available. Careful selection
of the runoff coefficient C will give values of peak
runoff consistent with project significance. The
values of C in Table 21.15 for urban areas are com-
monly recommended design values (V. T. Chow,
Hydrologic Determination of Waterway Areas for
the Design of Drainage Structures in Small
Drainage Basins, University of Illinois Engineering
Experimental Station Bulletin 426, 1962).
After selection of the design-storm frequency of
occurrence, for example, a 50- or 100-year-frequency
storm, the rainfall intensity I may be determined
from any of a number of formulas or from a statisti-
cal analysis of rainfall data if enough are available.
Chow lists 24 rainfall-intensity formulas of the form
(21.128)
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.78
I
Section Twenty-One
where I = rainfall intensity, in/h
K, b, n, and n
1
= respectively, coefficient, fac-
tor, and exponents depend-
ing on conditions that affect
rainfall intensity
F = frequency of occurrence of
rainfall, years
t = duration of storm, min
= time of concentration
Perhaps the most useful of these formulas is the
Steel formula:
(21.129)
where K and b are dependent on the storm fre-
quency and region of the United States (Fig. 21.71
and Table 21.16).
Equation (21.129) gives the average maximum
precipitation rates for durations up to 2 h.
The time of concentration T
c
at any point in a
drainage system is the sum of the overland flow
time; the flow time in streets, gutters, or ditches;
and the flow time in conduits. Overland flow time
may be determined from any number of formulas
developed for the purpose. (See D. R. Maidment,
Handbook of Hydrology, McGraw-Hill, Inc.,
New York.) The flow time in gutters, streets, ditch-
es, and conduits can be determined from a calcula-
tion of the average velocity using the Manning
equation [Eq. (21.89)] . The time of concentration is
usually expressed in minutes.
After determining the time of concentration,
calculate the corresponding rainfall intensity from
either Eq. (21.128) or Eq. (21.129), or any equivalent
method. Then select the runoff coefficient from
Table 21.15 and determine the peak discharge from
Eq. (21.127).
Since the rational formula assumes a constant
uniform rainfall for the time of concentration over
the entire area, the area A must be selected so that
this assumption applies with reasonable accuracy.
Adhering to this assumption may necessitate sub-
dividing the drainage area.
21.41.2 Method for Determining
Runoff for Major Hydraulic
Structures
The unit-hydrograph method, pioneered in 1932 by
LeRoy K. Sherman, is a convenient, widely accept-
Type of Runoff
Drainage Area Coefficient C
Business:
Downtown areas 0.70 0.95
Neighborhood areas 0.50 0.70
Residential:
Single-family areas 0.30 0.50
Multiunits, detached 0.40 0.60
Multiunits, attached 0.60 0.75
Suburban 0.25 0.40
Apartment dwelling areas 0.50 0.70
Industrial:
Light areas 0.50 0.80
Heavy areas 0.60 0.90
Parks, cemeteries 0.10 0.25
Playgrounds 0.20 0.35
Railroad-yard areas 0.20 0.40
Unimproved areas 0.10 0.30
Streets:
Asphaltic 0.70 0.95
Concrete 0.80 0.95
Brick 0.70 0.85
Drives and walks 0.75 0.85
Roofs 0.75 0.95
Lawns:
Sandy soil, flat, 2% 0.05 0.10
Sandy soil, avg, 27% 0.10 0.15
Sandy soil, steep, 7% 0.15 0.20
Heavy soil, flat, 2% 0.13 0.17
Heavy soil, avg, 27% 0.18 0.22
Heavy soil, steep, 7% 0.25 0.35
Table 21.15 Common Runoff Coefficients
Fig. 21.71 Regions of the United States for use
with the Steel formula.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.79
ed procedure for determining runoff for major
hydraulic structures. (Leroy K. Sherman, Stream-
flow from Rainfall by Unit-Graph Method, Engi-
neering News-Record, vol. 108, pp. 501-505, January-
June 1932.) It permits calculation of the complete
runoff hydrograph from any rainfall after the unit
hydrograph has been established for the particular
area of concern.
The unit hydrograph is defined as a runoff
hydrograph resulting from a unit storm. A unit
storm has practically constant rainfall intensity for
its duration, termed a unit period, and a runoff vol-
ume of 1 in (water with a depth of 1 in over a unit
area, usually 1 acre). Thus, a unit storm may have a
2-in/h effective intensity lasting
1
/2 h or a 0.2-in/h
effective intensity lasting 5 h. The significant part of
the definition is not the volume but the constancy
of intensity. Adjustments can be made within unit-
hydrograph theory for situations where the runoff
volume is different from 1 in, but corrections for
highly variable rainfall rates cannot be made.
The unit hydrograph is similar in concept to
determining a set of factors for a specific drainage
basin. The set consists of one factor for each vari-
able that affects runoff. The unit hydrograph is
much quicker, easier, and more accurate than any
such set of factors. The method is summarized by
the formula
(21.130)
The unit hydrograph thus is the link between rain-
fall and runoff. It may be thought of as an integral
of the many complex factors that affect runoff. The
unit hydrograph can be derived from rainfall and
stream-flow data for a particular storm or from
stream-flow data alone.
Assumptions made in the development of the
unit-hydrograph theory are:
1. Rainfall intensity is constant for its duration or
a specified period of time. This requires that a
storm of short duration, termed a unit storm, be
used for the derivation of the unit hydrograph.
2. The effective rainfall is uniformly distributed
over the drainage basin. This specifies that the
drainage area be small enough for the rainfall to
be essentially constant over the entire area. If
the watershed is very large, subdivision may be
required; the unit-hydrograph theory is then
applied to each subarea.
3. The base of the hydrograph of direct runoff is
constant for any effective rainfall of unit dura-
tion. This needs no clarification except that the
base of a hydrograph, that is, the time of storm
runoff, is largely arbitrary since it depends on
the method of base-flow separation.
4. The ordinates of the direct runoff hydrographs
of a common base time are directly proportional
to the total amount of direct runoff represented
Region
Coefficients
1 2 3 4 5 6 7
2 K 206 140 106 70 70 68 32
b 30 21 17 13 16 14 11
4 K 247 190 131 97 81 75 48
b 29 25 19 16 13 12 12
10 K 300 230 170 111 111 122 60
b 36 29 23 16 17 23 13
25 K 327 260 230 170 130 155 67
b 33 32 30 27 17 26 10
50 K 315 350 250 187 187 160 65
b 28 38 27 24 25 21 8
100 K 367 375 290 220 240 210 77
b 33 36 31 28 29 26 10
Table 21.16 Coefficients for Steel Formula
Frequency,
years
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.80
I
Section Twenty-One
by each hydrograph. Illustrated in Fig. 21.72,
this is basically the principle of superposition or
proportionality. It enables calculation of the
runoff for a storm of any intensity or duration
from a unit storm, which is of fixed intensity and
duration. A given storm may be resolved into a
number of unit storms. Then, the runoff may be
calculated by superimposing that number of
unit hydrographs.
5. The hydrograph of direct runoff for a given
period of rainfall reflects all the combined phys-
ical characteristics of the basin (commonly
referred to as the principle of time invariance).
This assumption implies that the characteristics
of the drainage basin have not changed since
the unit hydrograph was derived. Because this
applies with varying degrees of accuracy to
watersheds, the characteristics of the drainage
basin must be fixed or specified. Daily and
weekly variations in initial soil moisture are
probably the greatest source of error in this
method since they are largely unknown. Man-
made alterations and stream-flow conditions
can be accounted for much more easily.
For ease of manipulation, the unit hydrograph
is frequently expressed in histogram form as a dis-
tribution graph (Fig. 21.73), which illustrates the
percentages of total runoff that occur during suc-
cessive unit periods. The ordinate for each unit
period is the mean value of runoff for that period.
Since the unit hydrograph is derived for a unit
storm of specific duration, it may be used only for
storms divided into unit periods of that length.
Usually, because of storm variations, the unit peri-
od must be different from that for which the unit
hydrograph was derived. This requires the recal-
culation of the unit hydrograph for the new unit
Fig. 21.72 Unit hydrograph (a) prepared for a unit storm is used to develop a composite hydrograph
(b) for any storm.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.81
period. This is accomplished by offsetting two S
hydrographs by a time equal to the duration of the
desired unit period (Fig. 21.74). An S hydrograph is
a representation of the cumulative percentages of
runoff that occur during a storm which has a con-
tinuous constant rainfall. It is calculated by cumu-
latively plotting the distribution percentages that
make up the distribution graph. The distribution
percentages for the new unit hydrograph are
determined by taking the difference between
mean ordinates for the two offset S hydrographs
and dividing by the new unit period.
Transposition of a unit hydrograph from one
basin to another similar basin may be made by cor-
relating their respective shape and slope factors.
This method was developed by Franklin F. Snyder
(Transactions of the American Geophysical Union, vol.
19, pt. I, pp. 447454). Also, since S hydrographs
are a characteristic of a drainage basin, those from
various basins may be compared to obtain an idea
of the variations that might exist when transposing
data from one basin to another.
In the application of the unit-hydrograph
method, a loss rate must be established to deter-
mine effective rain. This loss, during heavy storms,
is usually considered to be entirely infiltration. The
infiltration capacity of a soil may be determined
experimentally by lysimeter or infiltrometer tests.
(R. K. Linsley et al., Hydrology for Engineers,
3rd ed., McGraw-Hill, Inc., New York.)
21.42 Groundwater
Groundwater is subsurface water in porous strata
within a zone of saturation. It supplies about 20%
of the United States water demand. Where
groundwater is to be used as a water-supply
source, the extent of the groundwater basin and
the rate at which continuing extractions may be
made should be determined.
Aquifers are groundwater formations capable
of furnishing an economical water supply. Those
formations from which extractions cannot be made
economically are called aquicludes.
Permeability indicates the ease with which
water moves through a soil and determines
whether a groundwater formation is an aquifer or
aquiclude.
The rate of movement of groundwater is given
by Darcys law:
(21.131)
where Q = flow rate, gal/day
K = hydraulic conductivity, ft/day or
m/day
I = hydraulic gradient, ft/ft or m/m
A = cross-sectional area, perpendicular
to direction of flow, ft
2
or m
2
Hydraulic conductivity is a measure of the ability
of a soil to transmit water. It is a nonlinear function
of volumetric soil water content and varies with
soil texture. Many methods are available for deter-
mining hydraulic conductivity. (See D. R. Maid-
ment, Handbook of Hydrology, McGraw-Hill,
Inc., New York.)
Fig. 21.73 Distribution graph represents a unit
hydrograph as a histogram.
Fig. 21.74 Distribution percentages are deter-
mined from an offset S hydrograph.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.82
I
Section Twenty-One
Transmissibility is another index for the rate of
groundwater movement and equals the product of
hydraulic conductivity and the thickness of the
aquifer. Transmissibility indicates for the aquifer as
a whole what hydraulic conductivity indicates for
the soil.
An aquifer whose water surface is subjected to
atmospheric pressure and may rise and fall with
changes in volume is a free or unconfined aquifer. An
aquifer that contains water under hydrostatic pres-
sure, because of impermeable layers above and
below it, is a confined or artesian aquifer. If a well is
drilled into an artesian aquifer, the water in this
well will rise to a height corresponding to the
hydrostatic pressure within the aquifer. Frequent-
ly, this hydrostatic pressure is sufficient to cause
the water to jet beyond the ground surface into the
atmosphere. An artesian aquifer is analogous to a
large-capacity conduit with full flow in that extrac-
tions from it cause a decrease in pressure, rather
than a change in volume. This is in contrast to a
free aquifer, where extractions cause a decrease in
the elevation of the groundwater table.
Groundwater Management
I
With increas-
ing use being made of groundwater resources,
effective groundwater management is an absolute
necessity. Adequate management should include
not only quantity but quality. Quantity manage-
ment consists of effective control over extractions
and replenishment. Quality management consists
of effective control over groundwater pollution
resulting from waste disposal, recycling, poor-qual-
ity replenishment waters, or other causes.
Several steps or investigations are necessary for
developing an effective management program. First
is a comprehensive geologic investigation of the
groundwater basin to determine the characteristics of
the aquifers. Second is a qualitative and quantitative
hydrologic study of both surface water and ground-
waters to determine historical surpluses and defi-
ciencies, safe yield, and overdraft. (Safe yield is the
magnitude of the annual extractions from an aquifer
that can continue indefinitely without bringing some
undesirable result. Deteriorating water quality, need
for excessive pumping lifts, or infringement on the
water rights of others are examples of undesirable
results that could define safe yield. Regardless of
how it is defined, safe yield applies only to a specific
set of conditions based largely on judgment as to
what is desirable. Extractions in excess of the safe
yield are termed overdrafts.) In conjunction with the
hydrologic study, present and future water demands
should be determined. A detailed water-quality
study should be made not only of the groundwater
within the basin but also of all surface waters, waste-
waters, and other waters that replenish the ground-
water basin. Undesirable water-quality and -quantity
conditions should be identified.
Following the preceding preliminary work,
alternative management plans should be formulat-
ed. These management plans should consider vari-
ations in the quantity of extractions; groundwater
levels; quality, quantity, and location of artificial
replenishment; source, quantity, and quality of
water supply; and methods of wastewater dispos-
al. All alternative plans must recognize all legal
and jurisdictional constraints.
The final step is the operational-economic evalu-
ation of the alternatives and the selection of a rec-
ommended groundwater management plan. Oper-
ations and economic studies are normally
conducted by superimposing present and future
conditions in each alternative plan on historical
hydrologic conditions that occurred during a base
period. (A base period is a period of time, usually a
number of years, specifically chosen for detailed
hydrologic analysis because conditions of water
supply and climate during the period are equivalent
to a mean of long-term conditions and adequate
data for such hydrologic analysis are available.)
Economic evaluation of alternative plans should
consider cost of water-supply facilities, cost of
replenishment water, cost of wastewater-disposal
facilities, cost of pumping groundwater at the vari-
ous operational levels considered, and indirect
water-quality use costs, among others. (Indirect
water-quality use costs are those indirect costs
incurred by water distributors and consumers as a
result of using water of different qualities. These
costs include increased soap costs, water softening
costs, and costs associated with the more rapid dete-
rioration of plumbing and waterworks equip-
mentall of which increase as the hardness and
salinity of the water increase.)
Operational studies should determine the most
efficient manner of joint operation of surface and
groundwater systems (conjunctive use).
Use of computers and the development of a
mathematical model for the groundwater basin are
almost essential because of the number of repeti-
tive calculations involved.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.83
Upon completion of the operational and eco-
nomic studies, the most favorable management
scheme should be selected as the recommended
plan. This selection should be based not only on
economic and operational considerations but on
social, institutional, legal, and environmental fac-
tors. The plan should be capable of being readily
implemented, flexible enough to accommodate
different growth rates, financially feasible, and
generally acceptable to the water and wastewater
agencies operating in the basin.
An operating agency should be designated or
formed to implement the recommended plan. The
agency should have adequate powers to control or
cooperate in the control of surface-water supplies
groundwater recharge sites, surface-water deliv-
ery facilities, amount and location of groundwater
extractions, and wastewater treatment and dispos-
al facilities. The operating agency should develop a
comprehensive monitoring network and a data
collection and evaluation program to determine
the effectiveness of the management plan and to
implement any changes in the plan deemed neces-
sary. This monitoring network may consist of
selected wells where groundwater levels and
chemical characteristics are measured and certain
surface-water sampling locations where both
quantitative and qualitative factors are measured.
The program should also include quantitative
evaluation of extractions, water used, wastewater
disposed, and natural and artificial replenishment.
Integration of the above data with the computer
model of the groundwater basin is an efficient
method of evaluating the groundwater manage-
ment scheme.
(Ground Water Management, Manual and
Report on Engineering Practice, no. 40, American
Society of Civil Engineers, 1987; J. Bear, Hydraulics
of Ground Water, N. S. Grigg, Water Resources
Planning, A. I. Kashef, Groundwater Engineering,
R. K. Linsley et al., Hydrology for Engineers, 3rd
ed., McGraw-Hill Book Company, New York.)
Water Supply
A waterworks system is created or expanded to
supply a sufficient volume of water at adequate
pressure from the supply source to consumers for
domestic, irrigation, industrial, fire-fighting, and
sanitary purposes. A primary concern of the engi-
neer is estimation of the quantity of potable water
to be consumed by the community since the engi-
neer must design adequately sized components of
the water-supply system. Water-supply facilities
consist of collection, storage, transmission, pump-
ing, distribution, and treatment works.
To assure continuous service to the consumer for
fire-fighting and sanitary purposes in the event of an
earthquake, fire, flood, or other unforeseen emer-
gency, careful consideration must be given to the
selection of standby equipment and alternative sup-
plies of water. Maximum protection must be given to
power sources and pumps that must be available to
operate continuously during emergency conditions.
A dependable supply with sufficient pressure for
fighting fires considerably increases capital expendi-
tures for system construction. The smaller the sys-
tem, the larger the percentage of the total cost
chargeable to dependable fire flow.
21.43 Water Consumption
The size of a proposed water-supply project is usu-
ally based on an average annual per capita con-
sumption rate. Therefore, forecasts of population
for the design period are of the greatest impor-
tance and must be made with care to ensure that
components for the project are of adequate size.
Estimation of future population, however, is a very
difficult task.
Several mathematical methods are available for
use in predicting populations of cities. Some meth-
ods commonly used are arithmetical increase, per-
centage increase, decreasing percentage increase,
graphical comparison with other cities, and the
ratio method of comparing a community with a
state or country of which the community is a part.
Great care and judgment must be exercised in pop-
ulation prediction since many factors, such as
industrial development, land speculation, geo-
graphical boundaries, and age of the city, may
drastically alter mathematical estimates.
The total water supply of a city is usually dis-
tributed among the following four major classes of
consumers: domestic, industrial, commercial, and
public.
Domestic use consists of water furnished to
houses, apartments, motels, and hotels for drink-
ing, bathing, washing, sanitary, culinary, and
lawn-sprinkling purposes. Domestic use accounts
for between 30 and 60% (50 to 60 gal per capita per
day) of total water consumption in an average city.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.84
I
Section Twenty-One
Commercial water is used in stores and office
buildings for sanitary, janitorial, and air condition-
ing purposes. Commercial use of water amounts to
about 10 to 30% of total consumption.
Industrial uses of water are diverse but consist
mainly of heat exchange, cooling, and cleaning. No
direct relationship exists between the amount of
industrial water used and the population of the
community, but 20 to 50% of the total quantity of
water used per capita per day is normally charged
to industrial usage. Usually the larger-sized cities
have a high degree of industrialization and show a
correspondingly greater percentage of total con-
sumption as industrial water.
Public use of water for parks, public buildings,
and streets contributes to the total amount of water
consumed per capita. Fire demands are usually
included in this class of water use. The total quan-
tity of water used for fire fighting may not be large,
but because of the high rate at which it is required,
it may control the design of the facilities. About 5
to 10% of all water used is for public uses.
Waste and miscellaneous usage of water include
that lost because of leakage in mains, meter mal-
functions, reservoir evaporation, and unauthorized
uses. About 10 to 15% of total consumption may be
charged to waste and miscellaneous uses.
Water Demand Rate
I
Many factors, such
as the climate, size of the city, standard of living,
degree of industrialization, type of service
(metered or unmetered), lawn sprinkling, air con-
ditioning, cost, pressure, and quality of the water,
influence the demand rate for water.
Presence of industries usually increases the total
per capita use of water but decreases the demand
fluctuation. A good estimate of the potential indus-
trial water demand can be made by relating demand
to the percent of land zoned for industrial use.
Small cities frequently have a low per capita
demand for water, especially if portions of the city
are unsewered. Fluctuations in demand are greater
in small cities, mainly because of the lack of large
industries. High standards of living increase water
demand and fluctuations in rate of use.
Warm and dry climates have a higher rate of
water consumption because of sprinkling and air
conditioning. Cold weather sometimes increases
consumption because water is allowed to run to
prevent pipes from freezing.
Demand for water is related to water-service
meters, cost, quality, and pressure. Metering water
reduces the quantity of water consumed by 10 to
25% because of the usual increase in total cost of
consumers if they continue to use water at the
unmetered rate. High water pressures increase
demand because of greater losses at leaking
mains, valves, and faucets. Normally, if the cost of
water increases, the demand for it decreases.
Demand for water usually increases with an
improvement in quality.
Demand rates vary with time of day, month,
and year. Table 21.17 is a comparison between
water-demand rates for the city of Los Angeles
and a national average calculated from data in a
U.S. Public Health Service Report. The national
demand-rate data, as presented in Table 21.17, are
the average of a range of values, including some
very high and very low rates due to variations in
climatic conditions, degree of industrialization,
and time of day. Examples of divergent average
daily demand rates for various United States cities
are: 230 gal per capita per day for Chicago, 210 gal
per capita per day for Denver, 150 gal per capita
per day for Baltimore, and 135 gal per capita per
day for Kansas City, Mo.
The California Water Atlas, 1979, State of Cal-
ifornia Office of Planning and Research, presents
National avg Los Angeles, Calif.
Gal per capita % of avg Gal per capita % of avg
per day annual rate per day annual rate
Avg day 160 100 175 100
Max day 265 165 280 160
Max h 400 250 420 240
Table 21.17 Water-Demand Rates
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.85
average monthly demand rates for water use in
four coastal and four inland cities for the period
1966 to 1970. In the atlas, the effect of warm, dry
climatic conditions is indicated for each location by
the ratio of average monthly use to the annual
average gallon per capita per day. The maximum
monthly water-demand rates ranged from 119 to
141% of the annual rate for the coastal cities and
from 144 to 187% of the annual rate for the dry,
inland, valley cities. Moreover, the annual demand
rates for the inland areas averaged 78% higher
than those for the coastal cities. The difference is
due primarily to the great amount of lawn sprin-
kling in Los Angeles. Past water-demand records
of both the city being considered and other cities of
similar size, industrialization, climate, and so on
should be considered and incorporated in
demand-rate projections for water systems.
The total quantity of water used for fighting
fires is normally quite small, but the demand rate is
high. The fire demand as established by the Amer-
ican Insurance Association is
(21.132)
where G = fire-demand rate, gal/min
P = population, thousands
The required fire flows computed from this formula
are listed in Table 21.18. When calculating the total
flow to be used in design, fire flow should be added
to the average consumption for the maximum day.
21.44 Water-Supply Sources
The major sources of a water supply are surface
water and groundwater. In the past, surface sources
have included only the commonly occurring natur-
al fresh waters, such as lakes, rivers, and streams,
but with rapid population expansion and increased
per capita water use associated with a higher stan-
dard of living, consideration must be given to
desalination and waste-water reclamation as well.
In selection of a source of supply, the various
factors to be considered are adequacy and reliabil-
ity, quality, cost, legality, and politics. The criteria
are not listed in any special order since they are, to
a large extent, interdependent. Cost, however, is
probably the most important because almost any
source could be used if consumers are willing to
pay a high enough price. In some local areas, as
increasing demands exceed the capacity of existing
sources, the increasing cost of each new supply
focuses attention on reclamation of local supplies
of wastewater and desalination.
Adequacy of supply requires that the source be
large enough to meet the entire water demand.
Total dependence on a single source, however, is
frequently undesirable, and in some cases, diversi-
fication is essential for reliability. The source must
Avg area served per hydrant
in high-value districts, ft*
gal/min MGD

Direct streams Engine streams


1,000 1,000 1.4 4 0.3 100,000 120,000
2,000 1,500 2.2 6 0.6 90,000
4,000 2,000 2.9 8 1.0 85,000 110,000
10,000 3,000 4.3 10 1.8 70,000 100,000
17,000 4,000 5.8 10 2.4 55,000 90,000
28,000 5,000 7.2 10 3.0 40,000 85,000
40,000 6,000 8.6 10 3.6 40,000 80,000
80,000 8,000 11.5 10 4.8 40,000 60,000
125,000 10,000 14.4 10 6.0 40,000 48,000
200,000 12,000 17.3 10 7.2 40,000 40,000
* American Insurance Association.

MGD = million gallons per day; MG = million gallons.


Table 21.18 Required Fire Flow, Hydrant Spacing, and Fire Reserve Storage*
Population
Duration,
h
Reserve
storage,
MG

Fire Flow
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.86
I
Section Twenty-One
also be capable of meeting demands during power
outages and natural or created disasters. The most
desirable supplies from a reliability standpoint, in
order, are (1) an inexhaustible supply, whether
from surface or groundwater, which flows by grav-
ity through the distribution system; (2) a gravity
source supplemented by storage reservoirs; (3) an
inexhaustible source that requires pumping; and
(4) sources that require both storage and pumping.
As demand increases and supplies become over-
taxed, conservation practices in everyday use
become a valuable management tool.
Quality of the source determines both accept-
ability and cost; it varies considerably between
regions. Preliminary estimates of quality can be
made by examining the source, geology, and cul-
ture of the area.
Legality of supply is determined by doctrines
and principles of water rights, such as appropria-
tion, riparian, and ownership rights. Appropria-
tion right gives the first right priority over later
rights: first in time means first in right. Riparian
right permits owners of land adjacent to a stream
or lake to take water from that stream or lake for
use on their land. Ownership right gives a
landowner possession of everything below and
above the land. Legality is especially important for
groundwater supplies or where there is transfer of
water from one watershed to another.
A political problem with water supply exists
because political boundaries seldom conform to
natural-drainage boundaries. This problem is espe-
cially acute in extensive water-importation plans,
but it even exists in varying forms for wastewater
reclamation and desalination projects.
Desalination processes are of two fundamental
types: those that extract salt from the water, such as
electrodialysis and ion exchange, and those that
extract water from the salt, such as distillation,
freezing, and reverse osmosis. The energy cost of
the former processes is dependent on the salt con-
centration. Hence, they are used mainly for brack-
ish water. The energy costs for the water-extraction
processes are essentially independent of salinity.
These processes are used for seawater conversion.
Very large dual-purpose nuclear power and desali-
nation plants, which take advantage of the
economies realized by enormous facilities, have
been proposed, but such plants are feasible only
for those large urban areas located on coasts. Trans-
mission and pumping costs make inland use
uneconomical. Although desalination may have
advantages as a local source, it is not at present a
panacea that will irrigate the deserts.
Acceptance of wastewater reclamation as a
water source for direct domestic use is hindered by
public opinion and uncertainty regarding viruses.
Much effort has been expended to solve these
problems. But until such time as they are solved,
wastewater reclamation will have only limited use
for water supply. In the meantime, reclaimed
water is being used for irrigation in agricultural
and landscaping applications.
(D. W. Prasifka, Current Trends in Water-Sup-
ply Planning, Van Nostrand Reinhold, New York.)
21.45 Quality Standards for
Water
The Safe Drinking Water Act of 1974 mandated
that nationwide standards be established to help
ensure that the public receives safe water through-
out the United States. National Interim Primary
Drinking Water Standards were adopted in 1975,
based largely on the 1962 U.S. Public Health Ser-
vice Standards (Publication no. 956), which were
used for control of water quality for interstate car-
riers. These earlier standards had been widely
adopted voluntarily by both public and private
utilities and received the immediate endorsement
of the American Water Works Association as a min-
imum standard for all public water supplies in the
United States. Similar standards were developed
by the World Health Organization as standards for
drinking-water quality at international ports
(International Standards for Drinking Water,
World Health Organization, Geneva, Switzerland).
Heightened concern over our changing environ-
ment and its health effect on water supplies was a
major cause of the change from voluntary to
mandatory water-quality standards.
The Safe Drinking Water Act defines contami-
nant as any physical, chemical, biological, or radio-
logical substance or matter in water. Maximum
contaminant level (MCL) indicates the maximum
permissible level of a contaminant in water that is
delivered to any user of a public water system. The
act clearly delineates between health-related qual-
ity contaminants and aesthetic-related contami-
nants by classifying the former as primary and the
latter as secondary contaminants.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.87
Primary Standards
I
Tables 21.19 to 21.21 list
tests and maximum contaminant levels required by
the National Interim Primary Drinking Water Regu-
lations (Federal Register 40, no. 248, 5956659588,
Dec. 24, 1975). Following are explanatory material
and testing frequency for compliance with the reg-
ulations. Enforcement responsibility rests with the
U.S. Environmental Protection Agency or with
those states electing to take primary responsibility
for ensuring compliance with the regulations. The
EPA updates standards periodically.
Microbiological Quality
I
The major dan-
ger associated with drinking water is the possibili-
ty of its recent contamination by wastewater con-
taining human excrement. Such wastewater may
contain pathogenic bacteria capable of producing
typhoid fever, cholera, or other enteric diseases.
The organisms that have been most commonly
employed as indicators of fecal pollution are
Escherichia coli and the coliform group as a whole.
Table 21.19 outlines the coliform test results
required to meet the MCL for bacteriological qual-
ity. When organisms of the coliform group occur in
three or more of the 10-mL portions of a single
standard sample, in all five of the 100-mL portions
of a single standard sample, or exceed the given
values for a standard sample with the membrane-
filter test, remedial measures should be undertak-
en until daily samples from the same sampling
point show at least two consecutive samples to be
of satisfactory quality.
The minimum number of samples to be taken
from the distribution system and examined each
month should be in accordance with the population
served. A minimum of 1 sample should be taken in
any case, with 11 samples taken for 10,000 popula-
tion, 100 for 100,000 population, 300 for 1,000,000
population, and 500 for 5,000,000 and over. For
details of methods, see Standard Methods for the
Examination of Water and Wastewater, American
Public Health Association, American Water Works
Association, Water Pollution Control Federation.
Turbidity
I
A limit on turbidity has been set as
a primary contaminant because high turbidity may
interfere with disinfection efficiency, especially in
virus inactivation, and excessive particulates may
Type of contaminant Maximum contaminant levels (MCL)

Microbiological contaminants in When using the membrane filter test:


all water systems

1 colony/100 mL for the average of all monthly samples


Or 4 colonies/100 mL in more than one sample if less than 20
samples are collected per month
Or 4 colonies/100 mL in more than 5% of the samples if 20 or
more samples are examined per month
When using the multiple-tube fermentation test: (10-mL portions)
Coliform shall not be present in more than 10% of the portions
per month
Not more than one sample may have three or more portions positive
when less than 20 samples are examined per month
Or not more than 5% of the samples may have three or more portions
positive when 20 or more samples are examined per month
Turbidity in surface water 1 TU monthly average (5 TU monthly average may apply at state option)
systems only
Or 5 TU average of two consecutive days
* From Safe Water: A Fact Book on the Safe Drinking Water Act for Non-Community Water Systems, American Water Works Asso-
ciation, 1979.

TU = turbidity unit.

Systems using surface water or groundwater.


Table 21.19 Primary Drinking Water StandardsMicrobiological and Turbidity*
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.88
I
Section Twenty-One
stimulate growth of microorganisms in a distribu-
tion system. Daily turbidity sampling of surface
water as it enters the distribution system is
required, with certain exceptions in systems that
practice disinfection and maintain an active resid-
ual disinfectant in the system.
Chemical Substances
I
The MCL for inor-
ganic and organic chemicals are listed in Table
21.20. Testing for these substances to determine
compliance must be performed yearly for commu-
nity water systems utilizing surface-water sources
and every 3 years for systems using groundwater.
Noncommunity water systems supplied by surface
or groundwater must repeat the tests every 5 years.
If the routine test results indicate that the level
of any substance listed exceeds the MCL, addition-
al check samples are required. For the inorganic
and organic chemicals, except nitrates, if one or
more MCL are exceeded, the data are reported to
the state within 7 days and three additional sam-
ples are taken at the same sampling point within 1
month. If the average value of the original and
three check samples exceeds the MCL, this is
reported to the state within 48 h, the public is noti-
fied, and then a monitoring frequency designated
by the state should continue until the MCL has not
been exceeded in two successive samples or until a
monitoring schedule is set up as a condition to a
variance, exemption, or enforcement action.
Table 21.20 Primary Drinking Water StandardsChemicals and Radioactivity*
Maximum contaminant
Type of contaminant levels (MCL)

Inorganic chemicals in all water systems

Arsenic 0.05 mg / L
Barium 1 mg / L
Cadmium 0.010 mg / L
Chromium 0.05 mg / L
Lead 0.05 mg / L
Mercury 0.002 mg / L
Selenium 0.01 mg / L
Silver 0.05 mg / L
Nitrate (as N) 10 mg / L
Organic chemicals in surface water systems only
Endrin 0.0002mg / L
Lindane 0.004 mg / L
Methoxychlor 0.1 mg / L
Toxaphene 0.005 mg / L
2, 4-D 0.1 mg / L
2, 4, 5-TP (Silvex) 0.01 mg / L
Radiological contaminants (natural) in all water systems

Gross alpha 15 pCi / L


Combined Ra-226 and Ra-228 5 pCi / L
Radiological contaminants (synthetic) in surface-water systems for
populations of 100,000 or more
Gross Beta 50 pCi / L
Tritium 20,000 pCi / L
Strontium-90 8 pCi / L
* From Safe Water: A Fact Book on the Safe Drinking Water Act for Non-Community Water Systems, American Water Works Asso-
ciation, 1979.

mg/ L = milligrams per liter; pCi / L = picocuries per liter.

Systems using surface or groundwater.


Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.89
When the nitrate test results indicate that the
MCL has been exceeded, one additional check
sample must be taken within 24 h. If the average of
the original and check sample exceeds the MCL,
the water supplier should report this to the state
within 48 h and notify the public. Continued mon-
itoring follows the same rules as indicated for the
other chemical substances.
Trihalomethanes
I
Amendments to the
interim primary regulations in 1979 set a limit for
chloroform and three related organic chemicals of
the trihalomethane group. The MCL for total tri-
halomethanes, including chloroform, bromo-
dichloromethane, dibromochloromethane, and
bromoform, is 0.10 mg/L. Monitoring and compli-
ance are required of community water systems
serving populations greater than 10,000 that add
disinfectant to the treatment process of surface and
groundwaters. For each treatment plant in the sys-
tem, a minimum of four samples must be taken for
each quarter of a year. All samples must be taken
on the same day: 25% of the samples must be taken
at the extreme ends of the distribution system; the
remainder may be taken from the central portion
of the distribution system. To determine compli-
ance with the MCL, the total trihalomethane con-
centrations of all samples taken for the quarter are
averaged. Next, the average concentration for the
current quarter and for the three previous quarters
are averaged, yielding the running annual average. If
this average is less than 0.10 mg/L, the water sys-
tem is in compliance.
For more information on monitoring require-
ments and modification of treatment techniques to
lower the trihalomethane concentration, if that is
required, see Trihalomethanes in Drinking Water,
American Water Works Association; Federal Regis-
ter 44, no. 281, 68624-68707, Nov. 29, 1979; and R. L.
Jolley, Water Chlorination, Environmental Impact,
and Health Effects, vols. 1 and 2, Ann Arbor Sci-
ence, Ann Arbor, Mich.
Fluoride Limits
I
Fluoride is considered an
essential constituent of drinking water for preven-
tion of tooth decay in children. Conversely, excess
fluorides may give rise to dental fluorosis (spotting
of the teeth) in children. The recommended lower,
optimum, and upper control limits for fluoride
concentrations, taken from the 1962 Drinking
Water Standards, are shown in Table 21.21. They
also recommended that fluoride in average con-
centrations greater than twice the optimum values
shall constitute grounds for rejection of the supply.
The latter concentrations, based on average air
temperature (Table 21.21, MCL column), were used
to set the MCL for the primary drinking water reg-
ulations.
Water suppliers may continue to use these
guidelines for optimum levels of fluoridation to
control dental caries in children at the discretion of
the state because the Safe Drinking Water Act pre-
cludes Federal regulations that may require the
addition of any substance for preventive health
care purposes unrelated to contamination of
drinking water.
Recommended control limits,
fluoride concentrations, mg/ L
Annual avg of max or ppm* Maximum
daily air contaminant
temperatures

Lower Optimum Upper level


53.7 or lower 0.9 1.2 1.7 2.4
53.8 58.3 0.8 1.1 1.5 2.2
58.4 63.8 0.8 1.0 1.3 2.0
63.9 70.6 0.7 0.9 1.2 1.8
70.7 79.2 0.7 0.8 1.0 1.6
79.3 90.5 0.6 0.7 0.8 1.4
* From Drinking Water Standards, U.S. Public Health Service, no. 956, 1962.

Based on temperature data obtained for a minimum of 5 years.


Table 21.21 Allowable Fluoride Concentration
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.90
I
Section Twenty-One
When fluoridation (supplementation of fluoride
in drinking water) is practiced, the average fluoride
concentration should be kept within the upper and
lower limits in Table 21.21. In addition, fluoridated
and defluoridated supplies should be sampled with
sufficient frequency to determine that the desired
fluoride concentration is maintained.
Radioactivity
I
The limiting values in Table
21.20 for radioactive substance apply to the aver-
age result obtained from analysis of four quarterly
samples or to one composite test sample formed
from four quarterly samples. All water systems
serving surface water or groundwater should be
tested for the natural radiological contaminants.
But only those systems serving surface water to
populations exceeding 100,000 are required to test
for synthetic contaminants.
When the average gross alpha activity is greater
than 5 pCi/L, additional tests should be run to
determine the levels of radium 226 and 228 sepa-
rately. If the combined alpha activity exceeds 5
pCi/L, the data should be reported to the state
within 48 h and the public notified. Monitoring
should be continued at quarterly intervals until the
annual average concentration no longer exceeds
the maximum contaminant level.
When the gross beta activity is greater than 50
pCi/L, an analysis should be performed to identify
the major radioactive constituents present. The
appropriate organ and total body doses should be
calculated to determine whether the maximum con-
taminant level of 4 millirem/year has been exceeded.
This calculation is required when tritium and stron-
tium-90 are both present in any concentration.
Special Monitoring
I
Among amendments
to the interim primary drinking water regulations
in 1980 were requirements for monitoring of sodi-
um concentration levels and corrosivity character-
istics. Samples for sodium analysis should be col-
lected annually for systems using surface waters
and every 3 years for systems supplying ground-
water exclusively.
The corrosivity sampling program requires that
two samples per plant be collected annually for
systems using surface-water sourcesone during
midwinter and one during midsummer. Only one
sample is needed for groundwater sources. The
measurements should include pH, calcium hard-
ness, alkalinity, temperature, total dissolved solids,
and calculation of the Langelier index. (See also
Art. 21.50 and Standard Methods for the Exami-
nation of Water and Wastewater, American Public
Health Association, American Water Works Associ-
ation, and Water Pollution Control Federation.) At
the discretion of the state, monitoring for addition-
al corrosivity characteristics, such as sulfates and
chlorides, and determination of the Aggressive
index may be required.
No maximum contaminant levels have been pro-
mulgated with respect to any of these parameters.
Secondary Standards
I
The aesthetic con-
taminants are covered by the secondary drinking
water regulations. The limits are called secondary
maximum contaminant levels (SMCL) and are listed
in Table 21.22. These levels represent reasonable
goals for drinking-water quality but are not Federal-
ly enforceable. The states may use these SMCL as
guidelines and establish higher or lower levels that
may be appropriate, dependent on local conditions,
such as unavailability of alternate source waters or
other compelling factors, if public health and wel-
fare are not adversely affected. (See National Sec-
ondary Drinking Water Regulations, U.S. Environ-
mental Protection Agency 570/9-76-000.)
Source Protection
I
The U.S. Public Health
Service Drinking Water Standards recognized the
need for protecting the source of water supplies, as
indicated by the following extract:
Maximum contaminant
Type of contaminant levels
Chloride 250 mg/L
Color 15 color units
Copper 1 mg/L
Corrosivity Noncorrosive
Foaming agents 0.5 mg / L
Iron 0.3 mg / L
Manganese 0.05 mg / L
Odor 3 threshold odor number
pH 6.5 8.5
Sulfate 250 mg / L
Total dissolved solids (TDS) 500 mg / L
Zinc 5 mg / L
Table 21.22 Secondary Drinking-Water
Standards
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.91
The water supply should be obtained from the most
desirable source feasible, and effort should be made to
prevent or control pollution of the source. If the
source is not adequately protected against pollution
by natural means, the supply shall be adequately pro-
tected by treatment.
Sanitary surveys shall be made of the water-supply
system, from the source of supply to the connection of
the customers service piping, to locate and correct
any health hazards that might exist. The frequency of
these surveys shall depend upon the historical need.
Adequate capacity shall be provided to meet peak
demands without development of low pressures and
the possibility of backflow of polluted water from cus-
tomer piping.
Case histories and monitoring programs have
been reported indicating that active source protec-
tion can enhance water quality with minimal extra
expense. (See R. B. Pojasek, Drinking-Water Qual-
ity Enhancement through Source Protection, Ann
Arbor Science Publishers, Inc., Ann Arbor, Mich.)
Water Treatment
Water is treated to remove disease-producing bac-
teria, unpleasant tastes and odors, particulate and
colored matter, and hardness and to lower the lev-
els of any contaminants when necessary to meet
water-quality standards. Some of the more com-
mon methods of treatment are plain sedimentation
and storage, coagulation-sedimentation, slow and
rapid sand filtration, disinfection, and softening
(see also Art. 21.51).
Long-term storage of water reduces the
amount of disease-producing bacteria and particu-
late matter. But economic conditions usually com-
pel water purveyors to use more efficient methods
of treatment, such as those mentioned above.
21.46 Sedimentation
Processes
Sedimentation or clarification is a process of
removing particulate matter from water through
gravity settlement in a basin by reducing the flow-
through velocity. Factors that affect the settling
rate of particulate matter suspended in water are
size, shape, and specific gravity of the suspended
particles; temperature and viscosity of the water;
and size and shape of the settling basin.
The settling velocity
s
of spherically shaped
particles in a viscous liquid can be found by use of
Stokes law if the Reynolds number R = d/, cal-
culated with =
s
, is equal to or less than 1.
(21.133)
where
s
= settling velocity of particle, m/s
g = acceleration due to gravity, m/s
2
= absolute viscosity of the fluid, Pas

1
= density of particle, g/mm
3
= density of fluid, g/mm
3
d = particle diameter, mm
If R > 2000, Newtons law applies:
(21.134)
where C
D
is the drag coefficient. Figure 21.75
shows a plot of C
D
values vs. Reynolds numbers, to
be used in Eq. (21.134).
In the region where 1.0 < R < 2000, there is a
transition from Stokes law to Newtons. The set-
tling velocity in this region is somewhere between
the values given by Newtons law and those given
by Stokes law; however, no exact expression has
been developed to give the velocity.
Figure 21.76 shows the relationship of settling
velocity to diameter of spherical particles with spe-
cific gravity S between 1.001 and 5.0.
21.46.1 Plain Sedimentation
The ideal settling basin (Fig. 21.77) is a sedimentation
tank in which flow is horizontal, velocity is constant,
and concentration of particles of each size is the
same at all points of the vertical cross section at the
inlet end. The basin has a volumetric capacity C,
depth h
o
, and width B. The surface loading rate or
overflow velocity
o
, equal to the settling velocity of
the smallest particle to be completely removed, can
be determined by dividing the flow rate Qby the set-
tling surface area A. For this ideal basin, the overflow
velocity therefore is
o
= Q/A = Q/BL
o
, where Q =
Bh
o
V and L
o
is the length of settling zone, V the flow-
through velocity. (Usually,
o
is expressed in gallons
per day per square foot of surface area.) The deten-
tion time t = h
o
/
o
= L
o
/V also equals the volumetric
capacity C divided by the rate of flow Q.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.92
I
Section Twenty-One
Fig. 21.75 Newton drag coefficients for spheres in fluids. (Observed curves, after Camp, Transactions of
the American Society of Civil Engineers, vol. 103, p. 897, 1946.)
Fig. 21.76 Chart gives settling velocities of spherical particles with specific gravities S, at 10 C.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.93
Particles with a settling velocity
s
>
o
, and
those that enter the settling zone between f and j
(at left in Fig. 21.77) with a settling velocity
s
larg-
er than (
1
= h
1
V/L
o
) but less than
o
, are removed
in this basin. The particles with a settling velocity

s
<
1
that enter the settling zone between f and e
are not removed in this basin.
The efficiency of a sedimentation basin is the
ratio of the flow-through period to the detention
time. The flow-through period is the time required
for a dye, salt, or other indicator to pass through
the basin. Settling-basin efficiencies are reduced by
many factors such as cross currents, short circuit-
ing, and eddy currents. A well-designed tank
should have an efficiency of 30 to 50%.
Some design criteria for sedimentation tanks are:
Period of detention2 to 8 h
Length-to-width ratio of flow-through channel
3:1 to 5:1
Depth of basin10 to 25 ft (15 ft average)
Width of flow-through channelnot over 40 ft (30
ft most common)
Diameter of circular tank35 to 200 ft (most com-
mon, 100 ft)
Flow-through velocitynot to exceed 1.5 ft/min
(most common velocity, 1.0 ft/min)
Surface loading or overflow velocity, gal per day
per ft
2
of surface areabetween 500 and 2000 for
most settling basins
Sedimentation tanks may be built in any of a
variety of shapes, for example, rectangular (Fig.
21.78a) or circular (Fig. 21.78b). Multistory tanks,
such as the two-story basin with a single tray in Fig.
27.8c, occupy less site area than the single-story
basin. The tubular settler (Fig. 21.78d) with parallel
flow upward provides very high surface areas.
(American Water Works Association and Amer-
ican Society of Civil Engineers, Water Treatment
Plant Design, McGraw-Hill, Inc., New York; G. M.
Fair, J. C. Geyer, and D. A. Okun, Water and
Wastewater Engineering, John Wiley & Sons, Inc.,
New York.)
21.46.2 Coagulation-Sedimentation
To increase the settling rate and remove finely
divided particles in suspension, coagulants are
added to the water. Without coagulants, finely
Fig. 21.77 Longitudinal section through an ideal settling basin.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.94
I
Section Twenty-One
Fig. 21.78 Types of sedimentation tanks: (a) Rectangular settling basin. (b) Circular clarifier. (c) Two-
story sedimentation basin. (d) Tubular settler.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.95
divided particles do not settle out because of their
high ratio of surface area to mass and the presence
of negative charges on them. The velocity at which
drag and gravitational forces are equal is very low,
and the negative charges on the particles produce
electrostatic forces of repulsion that tend to keep
the particles separated and prevent agglomeration.
When coagulating chemicals are mixed with water,
however, they introduce highly charged positive
nuclei that attract and neutralize the negatively
charged suspended matter.
Iron and aluminum compounds are commonly
used as coagulants because of their high positive
ionic charge. The alkalinity of the water being treat-
ed must be high enough for an insoluble hydroxide
or hydrate of these metals to form. These insoluble
flocs of iron and aluminum, which combine with
themselves and other suspended particles, precipi-
tate out when a floc of sufficient size is formed.
The more common coagulants are aluminum
sulfate, commonly known as alum [Al
2
(SO
4
)
3
.
18H
2
O]; ferrous sulfate (FeSO
4
7H
2
O); ferric chlo-
ride (FeCl
3
); and chlorinated copperas (a mixture of
ferric chloride and ferrous sulfate). The type and
amount of coagulant necessary to clarify a specified
water depend on the qualities of water to be treat-
ed, such as pH, temperature, turbidity, color, and
hardness. Jar tests are usually made in a laboratory
to determine the optimum amount of coagulant.
Some organic polymers are alternatives to the
metallic coagulants. Polymers are long-chain, high-
molecular-weight, organic polyelectrolytes. They
are available in three types: cationic, or positively
charged; anionic, or negatively charged; and non-
ionic, or neutral in charge. Cationic polymers are
generally the most suitable for use as primary
coagulants. Anionic polymers, however, are often
used as flocculant aids in conjunction with an iron
or aluminum salt to cause the formation of larger
floc particles. Thereby, lesser amounts of metallic
salt are needed to effect good coagulation.
Because of differences in the characteristics of
the suspended matter found in natural waters, not
all waters can be treated with equal success with
the same polymer or the same dosages. Jar tests
should be run with several dosages of the various
polymers available to aid in selecting the material
best suited for each water supply, considering both
cost and performance.
There are several reasons for considering the
use of polymers: increased settling rate and
improved filtrability of the floc, production of a
smaller volume of sludge, and easier dewatering.
Also, polymers have a minor effect on pH; conse-
quently, the need for final pH adjustment in the
finished water may be reduced.
Process Steps
I
The complete clarification
process is usually divided into three stages: (1)
rapid chemical mixing; (2) flocculation or slow stir-
ring, to get the small floc to agglomerate; and (3)
coagulation-sedimentation in low-flow-velocity
settling basins. Rapid chemical mixing may be
accomplished with many devices, such as mechan-
ical stirrers, centrifugal pumps, and air jets. The
time necessary for mixing ranges from a few sec-
onds to 20 min. Flocculation or slow stirring
increases floc size and speeds up settling. The
speed of the agitators must be great enough, how-
ever, to cause contact between the small floc but
not so great that the larger floc is broken up. Floc-
culator detention time should be in the 20- to 60-
min range. The coagulation-sedimentation process
takes place in a clarifier basin nearly identical to a
plain sedimentation basin. The detention period
for a clarifier should be between 2 and 8 h.
(G. L. Culp and R. L. Culp, New Concepts in
Water Purification, Van Nostrand Reinhold Com-
pany, New York; American Water Works Associa-
tion, Water Quality and Treatment, 4th ed., T. J.
McGhee, Water Supply and Sewerage, R. A. Cor-
bitt, Standard Handbook of Environmental Engi-
neering, McGraw-Hill, Inc., New York.)
21.47 Filtration Processes
Passing water through a layer of sand removes
much of the finely divided particulate matter and
some of the larger bacteria. The filtering process has
many components, such as physical straining,
chemical and biological reactions, settling, and neu-
tralization of electrostatic charges.
Direct Filtration
I
It is possible by use of
direct filtration to eliminate the sedimentation
step, in some instances, for treatment of raw
waters that are low in turbidity, color, coliform
organisms, plankton, and suspended solids, such
as paper fiber. Direct filtration is a water-treatment
process in which raw water is not settled prior to
the filtration step. It usually includes addition of a
coagulant to destabilize the colloidal particles and
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.96
I
Section Twenty-One
a polymer as a flocculant aid. The process requires
rapid mixing, agitation in a well-designed floccula-
tor for 10 to 30 min, addition of a polymer as a fil-
ter aid, and dual- or mixed-media filtration.
Pilot plant tests are essential for selecting the
best combination of coagulant and flocculant aid to
obtain a strong floc and to provide criteria for
design of the filtration units.
The principal advantages of direct filtration are
its lower capital and operation costs. Elimination of
settling basins can result in capital cost savings of
20 to 30%, and operational cost may be cut 10 to
30% by reduced chemical doses. Direct filtration
merits investigation before construction of new
facilities if the turbidity of the source water aver-
ages less than 25 TU.
Slow Sand Filters
I
These consist of an
underdrained, watertight container containing a 2-
to 4-ft layer of sand supported by a 6- to 12-in layer
of gravel. The effective size of the sand should be
in the 0.25- to 0.35-mm range. (The effective size is
the size of a sieve, in millimeters, that will pass
10%, by weight, of the sand. The uniformity coef-
ficient is the ratio of the size of a sieve that will
pass 60% of the sample to the effective size.) The
uniformity coefficient of the sand should be less
than 3. The sand is normally submerged under 4 or
5 ft of water. The water passes through the filter at
a rate of 3 to 6 MG per acre per day, depending on
the turbidity. The slow filter is not as versatile or as
efficient as rapid sand filters.
Rapid Sand Filtration
I
This is normally
preceded by chemical treatment, such as floccula-
tion-coagulation and disinfection, so the water can
be passed through the sand at a higher rate. Usu-
ally, the effluent from a rapid filter needs further
disinfection or chlorination because the bacteria
are not completely removed in this process. A dia-
gram of a typical gravity-type rapid sand filter is
shown in Fig. 21.79.
The normal order of flow through the varying
components of the filter is from the clarifiers (set-
tling tanks) to the top of the sand layer, through the
sand and gravel layers, through the underdrain lat-
erals to the main drain, and then through the con-
troller to the clear well for storage. Wash (cleaning)
water flow takes place in a reverse direction after
the filter effluent line has been closed. The wash-
Fig. 21.79 Gravity-type rapid sand filter.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.97
water flow is through the main drain to the laterals,
from the laterals upward through the gravel and
sand to the wash-water troughs. The troughs carry
the water to the gullet, which is drained to waste.
Some common design factors for rapid sand fil-
ters are:
Effective grain size0.35 to 0.55 mm
Uniformity coefficient1.20 to 1.75
Thickness of sand layer24 to 30 in depending on
grain size
Thickness of gravel layer15 to 24 in
Gravel sizefrom
1
/8 to 1
1
/2 in
Filtration rate2 to 4 gal/minft
2
(125 to 250 MG
per acre per day)
Total depth of each basin8 to 10 ft
Maximum head loss allowed before washing
sand8 to 10 ft
Sand expansion during washing25 to 50%
Wash-water rate15 to 20 gal/minft
2
Distance from top edge of wash-water trough to
top of unexpanded sand24 to 30 in
Length of filter runs between washings12 to 72 h
Spacing between wash-water troughs4 to 6 ft
Ratio of length to width of each basin1.25 to 1.35
Rapid sand filters are operated until the partic-
ulate matter and unsettled floc cover the openings
between the sand grains, creating a high head loss
across the filter. This high head loss slows down the
flow rate and may force some of the particulate
matter through the sand and gravel layers. Filters
are usually backwashed when the particulate-mat-
ter concentration increases in the filter effluent or
when the head loss reaches 8 to 10 ft. Backwashing
a filter consists of forcing filtered water through the
filter from the drains upward to the wash-water
troughs. The lightweight sediment is washed from
the sand grains by the moving water and some-
times by other agitating devices, such as rakes,
water sprays, and air jets. Filters must be washed
thoroughly or difficulties with mud balls, bed
cracking, or sand incrustation will be encountered.
Immediately after washing, filters pass water at
a high rate, which produces an undertreated efflu-
ent. Either manual or automatic rate control must
be used to prevent such an occurrence. Many
treatment plants control the rate of filtration by
using venturi tube devices, which throttle the filter
effluent line when there is high-velocity flow. As
clogging begins to occur in the filter, the velocity of
flow in the effluent line decreases, and the rate
controller then opens to increase the velocity.
A negative head is produced on the filter when
the head loss across the filter is greater than the
depth of water on the sand. Negative heads can
produce a condition known as air binding, which
is caused by removal of dissolved gases from the
water and formation between sand grains of bub-
bles that decrease filter capacity.
The underdrains of a filter are commonly made
of perforated pipe or porous plates. The under-
drains should be arranged so that each area filters
and distributes its proportionate share of water.
The ratio of total area of perforations to the total fil-
ter-bed area is normally in the 0.002:1 to 0.005:1
range. The diameter of the perforations varies
between
1
/4 and
3
/4 in.
Wash-water troughs should be evenly spaced,
and water should not have to travel more than 3 ft
horizontally to get to a wash-water gutter. The
depth of water flow in a horizontal gutter may be
calculated from
(21.135)
where Q = total flow received by trough,
gal/min
b = width of trough, in
y = water depth at upstream end of
trough, in
The total gutter depth can be found by adding 2 or
3 in of freeboard to the calculated depth y.
Other Processes
I
Anthracite coal may be
used in place of sand in gravity-type filters. The effec-
tive grain size is greater than that of sand, thus per-
mitting higher filtration rates and longer filter runs.
Dual-media, mixed-media, or deep coarse-media filters,
however, may be more advantageous. They operate
at the higher filtration rates of 4 to 8 gal/minft
2
.
A pressure filter is composed of a gravity-filter
medium enclosed in a watertight vessel. The filter-
ing medium may be sand, diatomaceous earth, or
anthracite coal. Pressure filters are primarily sup-
plemental and are used for specialized industrial
uses and for clarifying swimming pool water.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.98
I
Section Twenty-One
Filter galleries are made up of horizontal, perfo-
rated, or open-joint pipes, placed in shallow sand or
gravel aquifers. Galleries typically are fed by diver-
sion or pumping from streams into spreading basins
with gravel or sand bottoms. Some, however, may be
located in aquifers with high groundwater table. The
filtered water may be pumped from the gallery or
allowed to flow out one end by gravity.
(G. L. Culp and R. L. Culp, New Concepts in
Water Purification, Van Nostrand Reinhold Com-
pany, New York; American Water Works Associa-
tion, Water Quality and Treatment, 4th ed., and
American Society of Civil Engineers, Water Treat-
ment Plant Design, and T. J. McGhee, Water Sup-
ply and Sewerage, 6th ed., McGraw-Hill Book
Company, New York; G. M. Fair, J. C. Geyer, and D.
A. Okun, Water and Wastewater Engineering,
John Wiley & Sons, Inc., New York.)
21.48 Water Softening
Presence of the bicarbonates, carbonates, sulfates,
and chlorides of calcium and magnesium in water
causes hardness. Three major classifications of hard-
ness are: (1) carbonate (temporary) hardness caused
by bicarbonates, (2) noncarbonate (permanent)
hardness, and (3) total hardness. Municipal treat-
ment plants generally use either the lime-soda (pre-
cipitation) process or the base-exchange (zeolite)
process to reduce the hardness of the water to below
100 mg/L (about 100 ppm) of CaCO
3
equivalence.
In the lime-soda process, lime (CaO), hydrated
lime [Ca(OH)
2
], and soda ash (Na
2
CO
3
) are added
to the water in sufficient quantities to reduce the
hardness to an acceptable level. The amounts of
lime and soda ash required for softening to a resid-
ual hardness can be determined by use of chemical-
equivalent weights, taking into account that com-
mercial grades of lime and hydrated lime are about
90 and 68% CaO, respectively. Residual hardness of
50 to 100 mg/L as CaCO
3
remains in the treated
water because of the very slight solubility of both
CaCO
3
and Mg(OH)
2
. Hardness of water is normal-
ly expressed in grains per gallon (gpg) or mg/L of
CaCO
3
, where 1 gpg = 17.1 mg/L.
Chemical equations for the common lime-soda
softening processes are
(21.136)
(21.137)
(21.138)
(21.139)
Since the carbonate and magnesium hydroxide
particles settle out in sedimentation basins, facili-
ties must be provided for particle removal and dis-
posal. Deposition of CaCO
3
and Mg(OH)
2
on sand
grains, in clear wells, and in distribution pipes can
be prevented by recarbonation with CO
2
before
sand-filter treatment.
Hardness in water can be reduced to zero by
passing the water through a base-exchange or zeo-
lite material. These materials remove cations, such
as calcium and magnesium, from water and
replace them with soluble sodium and hydrogen
cations. Calcium can be removed from water as
shown by the following reaction:
(21.140)
where Ca
2+
is the calcium hardness ion removed,
Na
+
is the sodium ion replacing the Ca
2+
in water,
and R is the zeolite material. The reaction can be
reversed (from right to left) by increasing the Na
+
concentration to a high value, as generally is done
in regeneration of the softening unit.
Sodium chloride (table salt) is commonly used
to regenerate the unit. Regeneration requires
between 0.3 and 0.5 lb of salt per 1000 grains of
hardness removed.
Some hardness-removal capacities per cubic
foot of base-exchange material are: natural zeo-
lite2500 to 3000 grains, synthetic zeolite5000 to
30,000 grains (1 lb = 7000 grains).
(American Water Works Association, Water
Quality and Treatment, 4th ed., and American
Society of Civil Engineers, Water Treatment Plant
Design, McGraw-Hill Book Company, New York.)
21.49 Disinfection with
Chlorine
Chlorine in either the liquid, gas, or hypochlorite
form is frequently used for destroying bacteria in
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.99
water supplies. Other disinfectants are iodine,
bromine, ozone, chlorine dioxide, ultraviolet light,
and lime.
The reaction of chlorine with water is
(21.141)
The hypochlorous acid (HOCI) reacts with the
organic matter in bacteria to form a chlorinated com-
plex that destroys living cells. The amount of chlo-
rine (chlorine dose) added to the water depends on
the amount of impurities to be removed and the
desired residual of chlorine in the water. Chlorine
residuals of 0.1 or 0.2 mg/L are normally maintained
in water-treatment-plant effluent streams as a factor
of safety for the water as it travels to the consumer.
The concern over trihalomethane formation fol-
lowing chlorination of waters containing apprecia-
ble amounts of natural organic materials (Art. 21.62)
has led to use of alternate disinfectants. The prime
candidates are ozone and chlorine dioxide. The
benefits of ozone should be investigated for new or
modified treatment plants, particularly if there are
color or taste and odor problems in the raw water.
(American Water Works Association and Amer-
ican Society of Civil Engineers, Water Treatment
Plant Design, and T. J. McGhee, Water Supply
and Sewerage, McGraw-Hill, Inc., New York.)
21.50 Carbonate Stability
Water may either corrode or place a protective car-
bonate film on the interior surfaces of pipes. Which
it does depends on the nature and amount of
chemicals dissolved in the water.
An approximation of the stability of a water sup-
ply can be obtained by adding an excess of calcium
carbonate powder to one-half of a water sample. Stir
or shake each half sample at 5-min intervals for about
1 h. Filter both solutions; then, either take the pH or
determine the methyl orange alkalinity of each sam-
ple. If the untreated water has a higher alkalinity or
pH than the CaCO
3
-treated water, the water is satu-
rated with carbonate and may deposit protective
films in pipes. If the untreated water has a lower pH
or alkalinity value than the treated water, the water is
unsaturated with carbonate and may be corrosive. If
the pH or alkalinity is the same in both samples, the
water is in equilibrium in regard to carbonates.
The greater the difference in either alkalinity or
pH between the two samples, the greater the
amount of either unsaturation or saturation with
respect to carbonates. If the untreated water has a
much higher pH or alkalinity than the treated
water, the water is highly saturated with carbon-
ates. It can cause a problem with heavy carbonate
deposits in pipes and appurtenances of the pur-
veyor and consumer.
(G. M. Fair, J. C. Geyer, and D. A. Okun, Water
and Wastewater Engineering, John Wiley & Sons,
Inc., New York.)
21.51 Miscellaneous
Treatments
Many different methods of treatment are used to
remove such undesirable elements as color, taste,
odor, excessive fluorides, detergents, iron, man-
ganese, and substances exceeding the water-quali-
ty maximum contaminant levels (Art. 21.45).
Activated carbon is commonly used for taste and
odor removal. The carbon can be applied as a pow-
der to the water and later removed by a sand filter,
or the water can be passed through a bed of carbon
to remove natural and synthetic organic chemicals.
Treatment techniques for removal of inorganic
contaminants include conventional coagulation,
lime softening, cation exchange, anion exchange,
activated carbon, reverse osmosis, and electrodial-
ysis. Concerns over the potential for lead poison-
ing from lead in drinking water passing through
lead pipes installed long ago but still in use or from
leaded solder used for pipe joints have encouraged
abandonment of such practices. Where the pres-
ence of lead is detected in a water supply, despite
its low solubility, its concentration can be nearly
completely removed with lime softening or alum
and ferric sulfate coagulation.
(American Water Works Association and Amer-
ican Society of Civil Engineers, Water Treatment
Plant Design, McGraw-Hill, Inc., New York.)
Water Collection Storage and
Distribution
21.52 Reservoirs
The basic purpose of impounding reservoirs is to
hold runoff during periods of high runoff and
release it during periods of low runoff. The specif-
ic functions of reservoirs are hydroelectric, flood
control, irrigation, water supply, and recreation
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.100
I
Section Twenty-One
(see also Art. 21.52.1). Many large reservoirs are
multipurpose, as a consequence of which the spe-
cific functions may dictate conflicting design and
operating criteria. Also, equitable cost allocation is
more difficult.
Sizing of a reservoir for a project where the
demand for water is much greater than the mean
stream flow is an economic balance between bene-
fits and costs. A preliminary study of available
reservoir sites should be made to obtain the relative
costs for various size reservoirs. The dependable
flow that can be obtained from various-size reser-
voirs can be determined from the mass diagram for
stream flow. An economic comparison should then
be made of the benefits of various flows and the
costs of various reservoirs. The reservoir size that
will give the maximum benefit should be selected.
When the demand rate is known, as is the case
for many water-supply projects, the required size
of the reservoir can be determined directly from a
mass diagram of stream flow.
The mass diagram(Fig. 21.80) is a graphical plot
of total stream-flow volume against time. The
slope of the curve is the rate of flow.
Selection of the critical period of years for a mass
curve depends on the function of the reservoir. For
a water-supply or hydroelectric development, min-
imum flows will be critical, whereas for flood-con-
trol reservoirs, maximum flows will govern.
Reservoir capacity for a certain demand can be
obtained by drawing a line with a slope equal to
the demand tangent to the mass curve at the
beginning of a selected dry period, as shown by
lines AB and AC in Fig. 21.80. The ordinates d and
e represent the storage required to maintain
demands AB and AC.
Once a reservoir site has been selected, area-
volume curves (Fig. 21.81) are drawn to give the
characteristics of the site. The plot of volume vs.
water elevation is determined by planimetering
the area of selected contours within the reservoir
site and multiplying by the contour interval. Aeri-
Fig. 21.80 Mass diagram of stream flow.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.101
al mapping has made it possible to obtain accurate
contour maps at only a fraction of the costs of older
methods.
Another important consideration in the design
of reservoirs is deposition of sediment (see Arts.
21.35 and 21.52.2).
In selection of a site for a water-supply reser-
voir, give special attention to water quality. If pos-
sible, the watershed should be relatively uninhab-
ited to reduce the amount of treatment required.
(Water from practically all sources should be disin-
fected in the distribution system to ensure against
pollution and contamination.) Shallow reservoirs
usually give more problems with color, odor, and
turbidity than deep reservoirs, particularly in
warm climates or during warm seasons of the year.
Runoff heavily laden with silt and debris should be
diverted from the reservoir or treated before it is
mixed with the water supply. Alum is mixed into
reservoirs to reduce turbidity, and copper sulfate is
used to kill vegetation.
In deep reservoirs, during the summer months
the upper part of the reservoir will be warmed,
while below a certain level the temperature may be
many degrees cooler. The zone where the abrupt
temperature change takes place, which may be only
a few feet thick, is called the thermocline. The
waters above and below the thermocline circulate,
but there is no circulation across this zone. The water
in the lower level becomes low in dissolved oxygen
and develops bad tastes and odors. When the tem-
perature drops in the fall, the water at the upper
level becomes heavier than the water at the lower
level and the two levels become intermixed, causing
bad tastes and odors in the entire reservoir. To oxi-
dize organic matter and prevent poor water quality
in lower levels of reservoirs during summer months,
chlorine or compressed air should be released at var-
ious points on the bottom of the reservoirs.
21.52.1 Distribution Reservoirs
The two main functions of distribution reservoirs
are to equalize supply and demand over periods of
varying consumption and to supply water during
equipment failure or for fire demand. Major
sources of supply for some cities, such as New
York, San Francisco, and Los Angeles, are large dis-
tances from the city. Because of the large cost of
aqueducts, it is usually economical to size them for
the mean annual flow and provide terminal stor-
age for daily and seasonal fluctuations of demand.
Terminal storage is also necessary because of the
possibility of a failure along an aqueduct.
It is usually economical to have equalizing
reservoirs at various points in the distribution sys-
tem so that main supply lines, pumping plants, and
treatment plants can be sized for maximum daily
instead of maximum hourly demand. During hours
of maximum demand, water flows from these
reservoirs to the consumers. When the demand
drops off, the flow refills the reservoir. A mass dia-
Fig. 21.81 Area-capacity curves for a reservoir.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.102
I
Section Twenty-One
gram (Fig. 21.80) can be used to determine the
required capacity of the reservoir.
Equalizing reservoirs are usually built at the
opposite end of the system from the source of sup-
ply, so that during peak flows the maximum dis-
tance from the supply to the consumer is cut in half.
It is necessary for an equalizing reservoir to have an
elevation high enough to provide adequate pres-
sure throughout the system served. For the correct
hydraulic grade, it is necessary to build the reser-
voir above the area it serves. If the topography will
not allow a surface reservoir, a standpipe or an ele-
vated tank must be constructed. Standard elevated
tanks are available in capacities up to 2 MG.
21.52.2 Reservoir Trap Efficiency
The methods of Art. 21.35.2 for determining the
quantities of sediment delivered to a reservoir
require knowledge of the trap efficiency of the reser-
voir before the percentage of the incoming silt that
will remain to reduce storage can be determined.
Studies of trap efficiency were made by G. M. Brune,
who developed a curve to express the relationship
between trap efficiency and what he called the capac-
ity-inflow ratio for a reservoir (Fig. 21.82) (G. M. Brune,
Trap Efficiency of Reservoirs, Transactions of the
American Geophysical Union, vol. 34, no. 3, June 1953).
The higher the capacity-inflow ratio, acre-feet
of storage per acre-foot of annual inflow, the
greater the percentage of sediment trapped in a
reservoir. For any given storage reservoir, the trap
efficiency decreases with time since the capacity-
inflow ratio decreases as sediment builds up. The
rate of silting of a storage reservoir decreases when
the capacity is reduced to an amount such that
some spillage of silt-laden water occurs with each
major storm. This rate decrease occurs because an
increasing percentage of the annual suspended silt
load is vented before sedimentation can occur.
21.53 Wells
A gravity well is a vertical hole penetrating an
aquifer that has a free-water surface at atmospher-
ic pressure (Fig. 21.83). A pressure or artesian well
passes through an impervious stratum into a con-
fined aquifer containing water at a pressure greater
than atmospheric (Fig. 21.84). A flowing artesian
Fig. 21.82 Chart indicates percentage of incoming sediment trapped in reservoirs.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.103
Fig. 21.83 Gravity well in a free aquifer.
Fig. 21.84 Artesian well in a pressure aquifer.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.104
I
Section Twenty-One
well is an artesian well extending into a confined
aquifer that is under sufficient pressure to cause
water to flow above the casing head. A gallery or
horizontal well is a horizontal or nearly horizontal
tunnel, ditch, or pipe placed normal to groundwa-
ter flow in an aquifer.
21.53.1 Drawdown
When water is pumped from a well, the water
level around the well draws down and forms a
cone of depression (Fig. 21.83). The line of intersec-
tion between the cone of depression and the origi-
nal water surface is called the circle of influence.
Interference between two or more wells is
caused by the overlapping of circles of influence.
Drawdown for each interfering well is increased
and the rate of water flow is decreased for each
well in proportion to the degree of interference.
Interference between two or more closely spaced
wells may increase to the extent that the system of
wells produces one large cone of depression.
Since nearly all soils are heterogeneous, pump-
ing tests should be made in the field to determine
the value of the hydraulic conductivity K. A per-
meability analysis of a soil sample that is not rep-
resentative of the soil throughout the aquifer
would produce an unreliable value for K.
21.53.2 Flow From Wells
The steady flow rate Q can be found for a gravity
well by using the Dupuit formula:
(21.142)
where Q = flow, gal/day
K = hydraulic conductivity, ft/day under
1:1 hydraulic gradient
H = total depth of water from bottom of
well to free-water surface before
pumping, ft
h = H minus drawdown, ft
D = diameter of circle of influence, ft
d = diameter of well, ft
The steady flow, gal/day, from an artesian well is
given by
(21.143)
where t is the thickness of confined aquifer, ft (Fig.
21.84).
A long time elapses between the beginning of
pumping and establishment of a steady-flow con-
dition (a circle of influence with constant diame-
ter). Hence, correct values for drawdown and the
circle of influence can be obtained only after long
periods of continuous pumping.
A nonequilibrium formula developed by Theis
and a modified nonequilibrium formula produced
by Jacob are used in analyzing well flow conditions
where equilibrium has not been established. Both
methods utilize a storage coefficient S and the coef-
ficient of transmissibility T to eliminate complica-
tions due to the time lag before reaching steady
flow. (C. V. Theis, The Significance of the Cone of
Depression in Groundwater Bodies, Economic
Geology, vol. 33, p. 889, December 1938; C. E. Jacob,
Drawdown Test to Determine Effective Radius of
Artesian Well, Proceedings of the American Society of
Civil Engineers, vol. 72, no. 5, p. 629, 1940.) Comput-
er software packages are available for analysis of
groundwater flow with finite-element models.
21.53.3 Excavation of Wells
Wells may be classed by the method by which
they are constructed and their depth. Shallow
wells (less than 100 ft deep) are usually dug,
bored, or driven. Deep wells (depth greater than
100 ft) are usually drilled by either the standard
cable-tool, waterjet, hollow-core, or hydraulic
rotary methods.
21.53.4 Well Equipment
Essential well equipment consists of casing, screen,
eductor or riser pipe, pump (Art. 21.57), and motor.
The casing keeps the wall material and polluted
water from entering the well and prevents the
leakage of good water from the well.
The screen is placed below the casing to con-
tain the walls of the aquifer, to allow water to pass
from the aquifer into the well, and to stop move-
ment of the larger sand particles into the well. The
pump, motor, and eductor pipe are utilized to
move the water from the aquifer to the collecting
lines at the ground surface.
(G. M. Fair, J. C. Geyer, and D. A. Okun, Water
and Wastewater Engineering, John Wiley & Sons,
Inc., New York; T. J. McGhee, Water Supply and
Sewerage, 6th ed., McGraw-Hill, Inc., New York.)
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.105
21.54 Water Distribution
Piping
A water-distribution system should reliably provide
potable water in sufficient quantity and at adequate
pressure for domestic and fire-protection purposes.
To provide adequate domestic service, the pressure
in the main at house service connections usually
should not be below 45 psi. But if oversized plumb-
ing is provided, 35 psi is adequate. In steep hillside
areas, the system is usually divided into several dif-
ferent pressure zones, interconnected with pumps
and pressure regulators. Since each additional zone
causes increased expenses and decreased reliability,
it is desirable to keep their number to a minimum.
The American Water Works Association has recom-
mended 60 to 75 psi as a desirable range for pres-
sures; however, in areas of steep topography where
local elevation differences may be over 1000 ft, such
a narrow range is not practical.
House plumbing is designed to withstand a
maximum pressure of between 100 and 125 psi.
When the pressure in distribution lines is above
125 psi, it is necessary to install pressure regulators
at each house to prevent damage to appliances,
such as water heaters and dishwashers.
21.54.1 Water for Fire Fighting
Pressure requirements for fire fighting depend on
the technique and equipment used. Four methods
of supplying fire protection are:
1. Use of mobile pumpers which take water from
a hydrant. This method is used in most large
communities that have full-time, well-trained
fire departments. The required pressure in the
immediate area of the fire is 20 psi.
2. Maintenance of adequate pressure at all times
in the distribution system to allow direct con-
nection of fire hoses to hydrants. This technique
is commonly used in small communities that do
not have a full-time fire department and mobile
pumpers. The pressure in the distribution sys-
tem in the vicinity of a fire should be between
50 and 75 psi.
3. Use of stationary fire pumps located at various
points in the distribution system, to boost the
pressure during a fire and allow direct connec-
tion of hoses to hydrants. This method is not so
reliable or so widely used as the first two.
4. Use of a separate high-pressure distribution sys-
tem for fire protection only. There are only rare
instances in high-value districts of large cities
where this method is used because the cost of a
dual distribution system is usually prohibitive.
21.54.2 Hydraulic Analysis of
Distribution Piping
Distribution systems are usually laid out on a grid-
iron system with cross connections at various
intervals. Dead-end pipes should be avoided
because they cause water-quality problems.
Economic velocities are usually around 3 to 4 ft/s,
although during fires they can be much higher. Two-
and four-inch-diameter pipe can be used for short
lengths in residential areas; however, the American
Insurance Association (AIA) requires 6-in pipe for
fire service in residential areas. Also, maximum
length between cross connections is limited to 600 ft.
In high value districts, the AIA requires an 8-in pipe,
with cross connection at all intersecting streets. The
AIA standards also require that gate valves be locat-
ed so that no single case of pipe breakage, outside
main arteries, requires shutting off from service an
artery or more than 500 ft of pipe in high valued dis-
tricts, or more than 800 ft in any area. All small dis-
tribution lines branching from main arteries should
be equipped with valves. (Standard Schedule for
Grading Cities and Towns of the United States with
Reference to Their Fire Defenses and Physical Con-
ditions, American Insurance Association.)
Adequate service requires a knowledge of the
hydraulic grade at many points in a distribution
system for various flows. Several methods, based
on the following rules, have been developed for
analysis of complex networks:
1. The head loss in a conduit varies as a power of
the flow rate.
2. The algebraic sum of all flows into and out of
any pipe junction equals zero.
3. The algebraic sum of all head losses between
any two points is the same by any route, and
the algebraic sum of all head losses around a
loop equals zero.
A convenient device for simplifying complex
networks of various size pipes is the equivalent pipe.
For a series of different size pipes or several parallel
pipes, one pipe of any desired diameter and one
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.106
I
Section Twenty-One
specific length or any desired length and one spe-
cific diameter can be substituted; this will give the
same head loss as the original for all flow rates if
there are no take-outs or inputs between the two
end points. In complex networks, the equivalent
pipe is used mainly to simplify calculation.
Example 21.10: Determine the equivalent 8-in-
diameter pipe that will have the same loss of head
as the sections of pipe from A to D in Fig. 21.85a.
First, transform pipes CD, AB, and BD into
equivalent lengths of 8-in pipe; then, transform the
resulting sections into a single 8-in pipe with the
same head loss. The head loss may be calculated
from Eqs. (21.34d).
Assume any convenient flow through CD, say
500 gal/min. Equation (21.34d) indicates that loss of
head in 1000 ft of 6-in pipe is 32 ft and in 1000 ft of
8-in pipe, 7.8 ft. Then, the equivalent length of 8-in
pipe for CD is 500 32/7.8 = 2050 ft. Similarly, the
equivalent pipe for AB should be 165 ft long, and
for BD, 420 ft long. The network of 8-in pipe is
shown in Fig. 21.85b. It consists of pipe 1, 3000 +
2050 = 5050 ft long, connected in parallel to pipe 2,
165 + 420 = 585 ft long.
To reduce the parallel pipes to an equivalent 8-in
pipe, assume a flow of 1000 gal/min through pipe 2.
For this flow, the head loss in an 8-in pipe per 1000 ft
is 29 ft. Hence the head loss in pipe 2 would be 29
585/ 1000 = 17 ft. Since the pipes are connected in
parallel, the head loss in pipe 1 also must be 17 ft, or
3.37 ft /1000 ft. The flow that will produce this head
loss in an 8-in pipe is 310 gal/min [Eq. (21.34c)]. The
equivalent pipe, therefore, must carry 1000 + 310 =
1310 gal/min with a head loss of 17 ft. For a flow of
1310 gal/min, an 8-in pipe would have a head loss of
48 ft in 1000 ft, according to Eq. 21.34d. For a loss of
17 ft, an 8-in pipe would have to be 1000 17/ 48 =
350 ft long. So the pipes between A and D in Fig.
21.85a are equivalent to a single 8-in pipe 350 ft long.
Pipe Network Equations
I
For hydraulic
analysis of a water distribution system, it is conve-
nient to represent the network by a mathematical
model. Generally, it is useful to include in the
model only the major elements needed for a math-
ematical description of the basic network. (For
models that are to be used for such conditions as
low pressures in a small service region, however, it
may be necessary to include all the distribution
mains in the system.) The three analysis rules on p.
21.105 can then be used to develop a system of
simultaneous equations that can be solved for flow
and pressure in the network.
Typically, either the Darcy-Weisbach or the
Hazen-Williams formula is used to relate the char-
acteristics of each pipe in the system. Consequent-
ly, the equations for each pipe are nonlinear. As a
result, a direct solution generally is not available.
In practice, the equations are solved by an iteration
process, in which the values of some variables are
assumed to make the equations linear and then the
equations are solved for the other variables. The
initial assumptions are corrected and used to
develop new linear equations, which are solved to
obtain more accurate values of the variables.
One example of this technique is the Hardy
Cross method, a controlled trial-and-error method,
which was widely used before the advent of com-
puters. Flows are first assumed; then consecutive
adjustments are computed to correct these assumed
values. In most cases, sufficient accuracy can be
obtained with three adjustments; however, there
are rare cases where the computed adjustments do
not approach zero. In these cases, an approximate
method must be used.
Fig. 21.85 Distribution loop (a) may be
replaced by equivalent loop (b).
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.107
Assumed flows in a loop are adjusted in accor-
dance with the following equation:
(21.144)
where KQ
n
= h
f
= loss of head due to friction.
When the Hazen-Williams equation, used in Exam-
ple 21.10, is put in the form h
f
= KQ
n
then K =
4.727L/D
4.87
C
1
1.85
and n = 1.85. The expression
nKQ
n-1
equals (nKQ
n
/Q). In the Hazen-Williams
formula n = 1.85 for all pipes and can therefore be
taken outside the summation sign. Hence, the
adjustment equation becomes
(21.145)
It is important that a consistent set of signs be
used. The sign convention chosen for the follow-
ing example makes clockwise flows and the losses
from these flows positive; counterclockwise flows
and their losses are negative.
Approaches generally used for formulating the
equations for analysis of a water distribution net-
work include the following:
Flow method, in which pipe flows are the
unknowns.
Node method, in which pressure heads at the
pipe end points are the unknowns.
Loop method, in which the energy in each
independent loop is expressed in terms of the
flows in each pipe in the loop. In turn, the actual
flow in each pipe is expressed in terms of an
assumed flow and a flow correction factor for
each loop.
Computer software packages are available for
analysis of networks by such methods. They can
perform not only steady-state analyses of pres-
sures and flows in pipe networks but also time-
dependent analyses of pressure and flow under
changing system demands and of flow patterns
and basic water quality.
(V. J. Zipparo and H. Hasen, Davis Handbook
of Applied Hydraulics, McGraw-Hill, Inc., New
York; AWWA, Distribution Network Analysis for
Water Utilities, Manual of Water Supply Practices
M32, American Water Works Association, Denver,
Colo.; T. M. Walski, Analysis of Water Distribution
Systems, Van Nostrand Reinhold, New York.)
21.54.3 Cover over Buried Pipes
The cover required over distribution pipes
depends on the climate, size of main, and traffic. In
northern areas, frost penetration, which may be as
deep as 7 ft. is usually the governing factor. In
frost-free areas, a minimum of 24 in is required by
the AIA. If large mains are placed under heavy
traffic, the stress produced by wheel loads should
be investigated.
21.54.4 Maintenance of Water
Pipes
Maintenance of distribution systems involves keep-
ing records, cleaning and lining pipe, finding and
repairing leaks, inspecting hydrants and valves,
and many other functions necessary to eliminate
problems in operation. Valves should be inspected
annually and fire hydrants semiannually. Records
of all inspections and repairs should be kept.
Unlined distribution pipes, after years of usage,
lose much of their capacity because of corrosion
and incrustations. Cleaning and lining with
cement mortar restores the original capacity. Dead-
end pipes should be flushed periodically to reduce
the accumulation of rust and organic matter.
21.54.5 Economic Sizing of
Distribution Piping
When designing any major project, the designer
should choose the most economical of numerous
alternatives. Most of these alternatives can be sep-
arated and studied individually. An example of
two alternatives for a distribution system is one
serving peak hourly demands totally by pumps
and one doing it by pumps and equalizing reser-
voirs. The total costs of each plan should be com-
pared by an annual or present-worth cost analysis.
A method of determining minimum cost that
can readily be adapted to many conditions is set-
ting the first derivative of the total cost, taken with
respect to the variable in question, equal to zero. In
the sizing of pipes in a distribution system supplied
by pumps, the total costs of the pipes, pumping
plant, and energy may be expressed as an equation.
To find the most economical diameter of pipe, the
first derivative of the total cost, taken with respect
to the pipe diameter, should be set equal to zero.
The following equation for the most economical
pipe diameter was derived in that manner:
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.108
I
Section Twenty-One
(21.146)
where D = pipe diameter, ft
f = Darcy-Weisbach friction factor
b = value of power, dollars/hp per year
Q
a
= average discharge, ft
3
/s
S = allowable unit stress in pipe, psi
a = in-place cost of pipe, dollars/pound
i = yearly fixed charges for pipeline
(expressed as a fraction of total capi-
tal cost)
H
a
= average head on pipe, ft
21.54.6 Pipe Materials
Cast iron, steel, concrete, and plastics, such as
polyvinyl chloride, polyethylene, polybutylene, and
glass-fiber-reinforced thermosetting resins, are the
most common materials used in distribution pipes.
Wood pipelines are still in existence, but wood is
rarely used in new installations. Copper, lead, zinc,
brass, bronze, and plastic are materials used in small
pipes, valves, pumps, and other appurtenances.
Common pipe-joint materials are: cement, sand;
rubber, plastic, and sulfur compounds.
Cast iron is the most common material for city
water mains. Standard sizes range from 2 to 24 in
in diameter. Cast iron is resistant to corrosion and
usually has good hydraulic characteristics. If it is
cement-lined, the Hazen-Williams C value may be
as high as 145. In unlined pipes, however, iron
tubercles may form and seriously affect flow
capacity. Tuberculation can be prevented by lining
with cement or tar materials. The relatively high
cost of cast-iron pipe is only a slight disadvantage,
largely offset by the long average life of trouble-
free service. Bell-and-spigot and flange are the
most common joints in cast-iron pipe.
Steel is commonly used for large pipelines and
trunk mains but rarely for distribution mains. Steel
pipes with either longitudinal or spiral joints are
formed at steel mills from flat sheets. The tranverse
joints between pipe sections are usually made by
welding, riveting, bell-and-spigot with rubber gas-
ket, sealed flanges, or Dresser-type couplings.
Since steel is stronger than iron, thinner and
lighter pipes can be used for the same pressures.
Some disadvantages of thin steel pipe are inability
to carry high external loads, possibility of collapse
due to negative gage pressures, and high mainte-
nance costs due to higher corrosion rates and thin-
ner pipe walls. Steel pipes are usually corrosion-
protected on both the outside and inside with coal
tar or cement mortar. Under favorable conditions,
the life of steel pipe is between 50 and 75 years.
Concrete pipe may be precast in sections and
assembled on the job or cast in place. A machine
that produces a monolithic, jointless concrete pipe
without formwork has been developed for gravity-
flow and low-pressure applications. Most of the
precast-concrete pipe is reinforced or prestressed
with steel. Concrete pipe may be made watertight
by insertion of a thin steel cylinder in the pipe
walls. High-strength wire is frequently wound
around the thin steel cylinder for reinforcement.
Concrete is placed inside and outside the steel
cylinder to prevent corrosion and strengthen the
pipe. Some advantages of concrete pipe are low
maintenance cost, resistance to corrosion under
normal conditions, low transportation costs for
materials if water and aggregate are available local-
ly, and ability to withstand external loads. Some
disadvantages to be considered are leaching of free
lime from the concrete, the tendency to leak under
pressure due to the cracking and permeability of
concrete, and corrosion in strong acids or alkalies.
21.55 Corrosion in Water
Distribution Systems
Many millions of dollars are expended every year
to replace pipes, valves, hydrants, tanks, and
meters destroyed by corrosion. Some causes of cor-
rosion are the contact of two dissimilar metals with
water or soil, stray electric currents, impurities and
strains in metals, contact between acids and met-
als, bacteria in water, or soil-producing compounds
that react with metals.
Electrochemical corrosion of a metal occurs
when an electrolyte and two electrodes, an anode
and a cathode, are present. (Water may serve as an
electrolyte.) At the anode, the metal in contact with
the electrolyte changes into a positively charged par-
ticle, which goes into solution or forms an oxide film.
(The ease with which a metal changes to a metallic
ion when it is in contact with water depends on its
oxidation potential or solution pressure. Metals can
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.109
be arranged in an electromotive series of decreasing
oxidation potentials. Metals high in the electromo-
tive series corrode more readily than metals located
in a lower position.) For an iron pipe exposed to
water, for example, the anode reaction is Fe (metal)
Fe
2+
+ 2e, where e is an electron. At the cathode,
the metal having the excess electrons gives them up
to a charged particle, such as hydrogen in solution:
2H
+
+ 2e H
2
(gas). If the hydrogen gas produced
at the cathode is removed from the cathode by reac-
tion with oxygen to produce water molecules or by
water movement (depolarization), the corrosion
process continues (Fig. 21.86). Indications of corro-
sion in an inaccessible iron or steel pipeline are dis-
charges of rusty-colored water (due to the loosening
of rust and scale) and metallic-tasting water.
A marked decrease in capacity and pressure in
a pipe section usually indicates tuberculation
inside the line. Tuberculation is caused by the
deposition and growth of insoluble iron com-
pounds inside a pipe. Iron-consuming bacteria in
water can produce ferrous oxide directly if the iron
concentration is about 2 ppm. A continuous supply
of soluble iron in the presence of iron-consuming
bacteria or dissolved oxygen and basic substances
in the water increases the size of the tubercles.
Tubercles may become so large and decrease the
capacity in the pipe to such an extent that it has to
be cleaned or replaced.
Several factors influence the type and quantity
of metallic corrosion:
Presence of protective films. Some metals form
oxide films that act as protective layers for the
metal. Aluminum, zinc, and chromium are exam-
ples of this type of metal.
Strains, cracks, and undissolved impurities in a
metal act as sites for corrosion.
Agitation or movement of water increases the
corrosion rate of a metal because the oxygen sup-
ply rate to the cathode and the removal rate of
metal ions from the anode are increased. The pres-
ence of ionic compounds in the water speeds up
corrosion because the ions act as conductors of
electricity, and the more ions, the faster electrons
can move through the water.
Alternate wetting and drying tends to break up
the rust or oxide film, thus facilitating penetration
of the film by oxygen and water and lead to
increased corrosion.
High hydrogen-ion concentrations increase
corrosion rates because of the greater accessibility
of the hydrogen ions to the cathode.
Corrosion may be prevented or retarded by
proper selection of materials, use of protective coat-
ings, and treatment of the water. When selecting
materials, the engineer should take into account the
characteristics of the water and soil conditions
encountered. Protective coatings for metals may be
metallic or nonmetallic and applied on both the
inside and outside surfaces of the pipe. Common
nonmetallic coatings are cement and asphalt. Zinc is
an example of metallic coating materials used. Steel
pipe dipped in zinc (galvanized) or copper tubing is
commonly used for small service lines.
Also, to prevent corrosion, water may be treat-
ed with bases, such as soda ash, caustic soda, and
Fig. 21.86 Electrochemical corrosion of iron in low-pH water.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.110
I
Section Twenty-One
lime, to reduce hydrogen-ion concentration and to
induce precipitation of thin films of carbonates,
hydroxides, oxides, and so on on the walls of the
pipes. These thin films reduce the ability of water
to corrode otherwise unprotected metal surfaces.
Corrosion, however, normally precedes deposition
of scale because iron must be in solution to react
with the basic substances and dissolved oxygen in
the water to form scale.
Electrochemical corrosion of external surfaces of
pipelines and water tanks can be retarded by appli-
cation of a direct current to the metal to be protected
and to another metal that acts as a sacrificial anode
(Fig. 21.87). The potential applied to or produced by
the two metal surfaces must be large enough to
make the protected metal act as a cathode. The sacri-
ficial anode corrodes and must be replaced periodi-
cally. Zinc, magnesium, graphite, and aluminum
alloys are commonly used for anode materials.
(American Water Works Association, Water
Quality and Treatment, 4th ed., McGraw-Hill,
Inc., New York.)
21.56 Centrifugal Pumps
The purpose of any pump is to transform mechan-
ical or electrical energy into pressure energy. The
centrifugal pump, the most common waterworks
pump, accomplishes that in two steps. The first
transforms the mechanical or electrical energy into
kinetic energy with a spinning element, or impeller.
The kinetic energy is then converted to pressure
energy by diffuser vanes or a gradually diverging
discharge tube, called a volute (Fig. 21.88).
Water enters at the center, or eye, of the
impeller and is forced outward toward the casing
by centrifugal force. The discharge head of a cen-
trifugal pump is a function of the impeller diame-
ter and speed of rotation.
Design factors requiring consideration in the
selection of a centrifugal pump are net positive
suction head required, efficiency, horsepower, and
the head-discharge relationship.
Net positive suction head (NPSH) is the ener-
gy in the liquid at the center line of the pump. To
have practical meaning, it must be referred to as
either the required or available NPSH. Required
NPSH is a characteristic of the pump and is given
by the manufacturer. Available NPSH is a charac-
teristic of the system and is determined by the
engineer. It is the pressure in the liquid over and
above its vapor pressure at the suction flange of
the pump and is given, in feet, by
(21.147)
where p
a
= pressure, psia, on free-water surface
or at center line of closed conduit
p

= vapor pressure, psia, of water at its


pumping temperature
h
f
= friction loss in suction line, ft of
water
z = elevation difference, ft, between
pump center line and water surface
w = unit weight of liquid, lb/ft
3
If the suction water surface is below the pump cen-
ter line, z is negative. To prevent cavitation, it is nec-
essary to have the available NPSH always greater
Fig. 21.87 Cathodic protection of a metal.
Fig. 21.88 Volute-type centrifugal pump.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.111
than the required NPSH. For that reason, it is cus-
tomary to analyze a required NPSH vs. discharge
curve with the brake horsepower, head, and effi-
ciency curves when selecting a pump.
The operating point of a centrifugal pump is
determined by the intersection of the pumps head-
capacity curve and the system head curve, as shown
in Fig. 21.89. (Also included in Fig. 21.89 are the
other curves used in pump selection.) A system
head curve is a plot of the head losses in the system
vs. pump discharge. This curve shows the head dif-
ferential that must be supplied by the pump. In a
typical water-system analysis, there may be three or
four pertinent system head curves corresponding
to various consumption rates. The intersection of
these curves with the head vs. Q curve define a
range of operation rather than a single point.
Selection of a centrifugal pump is largely a mat-
ter of matching one of the many pumps available
to the system characteristics. An important consid-
eration is that the point of maximum efficiency
should be at or near the operating point. Centrifu-
gal pumps are available in almost any capacity
desired, with lifts of up to 700 ft per stage. Efficien-
cies may be as high as 93% for large pumps.
See also Art. 21.57 and check valves in Art. 21.58.
(I. J. Karassik et al., Pump Handbook, 2nd ed.,
McGraw-Hill Book Company, New York.)
21.57 Well Pumps
These are classified as centrifugal, propeller, jet,
helical, rotary, reciprocating, and air lift. Although
centrifugal pumps (Art. 21.56) are the most com-
mon for both shallow-well and deep-well pumps,
circumstances may dictate one of the other types.
Centrifugal pumps are used in wells over 6 in in
diameter. They have capacities up to 4000 or 5000
gal/min and heads up to 1200 ft, depending on the
number of stages. Efficiencies may be as high as
90% for the larger capacities; however, below 200
gal/min, the maximum efficiency is 75 to 80%.
Fig. 21.89 Curves used in selection of a centrifugal pump.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.112
I
Section Twenty-One
Propeller pumps are an axial-flow type. They
are used in high-capacity low-head applications.
Jet pumps (Fig. 21.90) operate by discharging
water through a nozzle and diverging conical tube,
which are located at the well bottom. The diverg-
ing conical tube creates lift by converting the high-
velocity head to pressure head. The suction con-
nection is made between the nozzle and entrance
to the diverging tube. Jet pumps have low efficien-
cies. They are used in small-capacity low-lift appli-
cations, especially where the water contains sand
or other impurities.
Helical pumps are a positive-displacement
type with a metal helical rotor rotating inside a
rubber helical stator. The screw action of the rotor
forces water through the pump and up the dis-
charge pipe. Helical pumps are small-capacity
high-lift pumps. They may be used in wells over 4
in in inside diameter.
Rotary pumps are also of the displacement
type. They have a fixed chamber in which gears,
vanes, cams, or pistons rotate with very close tol-
erances. These pumps have relatively constant
partial-load efficiencies. Full-load efficiencies range
from 50 to 85%. Because of the close tolerances,
they can be used only for sediment-free water.
Reciprocating pumps, either hand- or motor-
driven, utilize piston action to move water. Their
present-day use is primarily for small-capacity
low-lift private applications.
Air-lift pumps generate lift by using air bubbles
to decrease the specific weight of the column of
water in the discharge pipe below that of the sur-
rounding water in the well and create a pressure dif-
ferential that forces the water out of the well. Air-lift
pumps are the simplest and most foolproof of well
pumps since they have no submerged moving parts.
They can be used in any well but have the disad-
vantage of efficiencies below 50%.
Specific speed N
s
is a widely used criterion for
pump selection. It is the impeller speed corre-
sponding to a discharge of 1 gal/min at 1 ft of head
for the most efficient design.
(21.148)
where n = impeller speed, r/min
Q = discharge, gal/min
H = head, ft
The favorable design range of N
s
for radial-flow
(centrifugal) pumps is from 1500 to 4100. For N
s
between 4100 and 7500, mixed-flow pumps having
both radial and axial characteristics should be
used, and for N
s
above 7500, axial-flow (propeller)
pumps should be used.
Shallow-well pumps have their motors and
impellers at ground level, so that the entire lift is
suction. Since excessive suction lifts cause cavita-
tion, the lift is limited by atmospheric pressure and
the velocity head at the impeller, which is a func-
tion of specific speed. At sea level, the maximum
practical lift for a shallow-well pump is about 25 ft.
Deep-well pumps have their impellers close
enough to the water surface to eliminate cavita-
Fig. 21.90 Section through a jet pump (simpli-
fied).
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.113
tion. The motor may be at ground level with a long
shaft connecting it to the impellers, or it may be at
the bottom of the well, below and directly adjacent
to the impellers. The former type is called a deep-
well turbine pump and the latter a submersible pump.
Deep-well turbine pumps can be used only for
straight wells. The pump shaft is supported at
intervals of about 10 ft by rubber or metal bearings,
which are water- or oil-lubricated, respectively. If
sand is carried out with the water, an enclosed-
shaft or submersible pump must be used to pre-
vent bearing damage. Submersible pumps may be
used in crooked wells. Other advantages include
ease of increasing the well depth or lift and silent
operation. One disadvantage is that the motors are
difficult to reach for repairs.
(I. J. Karassik et al., Pump Handbook, 2nd ed.,
McGraw-Hill Book Company, New York.)
21.58 Valves
Water facilities use many different types of valves.
These are generally classified according to the
function they perform. The two major water-valve
classifications are isolating and controlling.
Isolating valves are used for separating or shut-
ting off sections of pipe, pumps, and control
devices from the rest of the system for inspection
and repair purposes. The major types of isolating
valves are gate, plug, sluice gate, and butterfly.
A control valve is normally used for continu-
ously controlling pressures and flow rates. Check,
needle, globe, air-relief, pressure-regulating, pres-
sure-relief, and altitude valves are usually consid-
ered as control valves.
Gate valves are the isolating valves used most
often in distribution systems, primarily because of
their low cost, availability, and low head loss when
fully open. They have limited value as control or
throttling devices because of seat wear and the
downstream deflection and chatter of the gate disk.
Also, the open area and rate of flow through the
valve are not proportional to the percentage open-
ing of the valve when partly open. Corrosion, solids
deposition, tubercle formation, large pressure dif-
ferences, and thermal expansion produce difficul-
ties in opening normally closed gate valves or in
closing normally open gate valves. Periodic inspec-
tion and operation of valves that are infrequently
operated will prevent many operational difficulties.
Some of the larger gate valves have gear-reduction
drives to permit manual operation. Very large
valves are operated by hydraulic and electric power.
A plug valve may be used for both control and
isolation purposes. It consists of a cylindrically
shaped plug (with a rectangular slot or circular ori-
fice) placed in a close-fitting cylindrical seat perpen-
dicular to the direction of flow. Cone and spherical
valves are special types of plug valves. Plug, cone,
and spherical valves can all be fully closed or
opened by a 90 rotation of the plug. The valves
may or may not be lubricated (large iron valves usu-
ally are). Hydraulic or electric power is commonly
utilized for operating the larger valves. Small plug
valves are commonly used for isolation purposes on
domestic and commercial service connections and
are known as service, curb, or corporation cocks.
Usually, because the meter is not directly adjacent to
the distribution pipe, three valves must be used, one
at the service connection, one just upstream of the
meter, and one between the meter and the cus-
tomers service line. Plug and cone valves are also
used for throttling and remote-control shutoff. Low
head loss, in-service lubrication features, and easy,
fast operation, even in the presence of unequal
pressures across the valve, are the major advantages
of plug-type valves. But these valves cost more than
gate, globe, and butterfly valves.
Butterfly valves can be used for throttling and
isolation purposes. The butterfly-valve mechanism
consists of a relatively thin circular disk pivoted on
a horizontal shaft. Hand or motor power, applied
through a gear-reduction device, rotates the disk.
Simplicity of construction and quick, easy opera-
tion are reasons why these valves are replacing
sluice gates and gate valves in many locations. Los
Angeles has replaced many sluice gates in reser-
voir towers with butterfly valves having seats of
corrosion-resistant metal, rubber, or Neoprene. A
disadvantage of butterfly valves is the higher cost
relative to sluice gates or gate valves.
Sluice gates are mainly used on the sides of
reservoir control towers and in open-channel
structures where pressure on one side of the gate
helps to seat it and prevent water leakage. Diffi-
culties with leakage and corrosion of gate frames
and stems are the main disadvantages of sluice
gates. Low cost and ease of operation in open-
channel flow conditions are the major advantages.
A needle valve is made of a streamlined plug or
needle that fits into a small orifice with a carefully
machined seat. Needle valves are used for accurate
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.114
I
Section Twenty-One
control of water flow because a large movement of
the needle is necessary before any measurable
change of flow rate takes place. Needle valves are
not normally used for isolating purposes because
of the high head losses produced by water flow
through the small orifices. Large-sized needle
valves are used for flow regulation under high
heads, such as for free discharge from reservoirs.
Interior-differential, tube, and hollow-jet are three
of the most common types of large needle valves.
Globe valves are commonly used in smaller sizes
for domestic purposes. The valve mechanism con-
sists of a screw-operated disk that is forced down on
a circular seat. Because of high head losses, globe
valves are rarely used for isolation purposes, but
they are commonly used for pressure regulation in
water-distribution systems. Many automatic control
valves, such as pressure regulators and altitude,
check, and relief valves, have globe-valve bodies
with various types of control mechanisms.
Pressure-regulating valves are used to reduce
pressures automatically. An air-relief and inlet valve
serves the dual purpose of allowing air to either
escape or enter a pipeline. Air that accumulates at
high points in a pipe impedes water flow and should
be allowed to escape through an air-relief valve
placed at this location. Furthermore, draining water
from low elevations in a pipeline may cause negative
pressures at higher elevations and collapse a pipe.
Air should be allowed to enter through air-relief and
inlet valves at the high points to prevent this.
Pressure-relief valves are used to release excess
pressure in an enclosure. Often, these excess pres-
sures are caused by sudden closure of a valve.
Altitude valves are used to control the water
level of elevated reservoirs. A pressure-activated
control closes the altitude valve when the tank is full
and opens the valve to allow water to flow from the
tank when pressure below the valve decreases.
Check valves are used in pipelines to allow for
one-directional flow only. Check valves placed in
centrifugal-pump suction lines are called foot
valves. These valves hold water in the suction line
and pump case so that the pump will not need
manual priming when started. The most common
check valve is the swing type.
21.59 Fire Hydrants
A fire hydrant normally consists of a cast-iron bar-
rel and a gate or compression-type shutoff valve,
which connects the barrel to the main. Two or more
hose outlets are normally located in the barrel
above the ground surface. Usually, an additional
gate valve is required between the hydrant and the
main to allow for shutoff and repair of the hydrant.
The number of 2
1
/2-in-diameter hose outlets on a
hydrant determines its class. For example, a hydrant
with two hose outlets is called a two-way hydrant.
Fire-hydrant construction standards have been
established by the American Water Works Associa-
tion and the American Insurance Association.
These standards relate the diameter of the barrel to
the size of the main-valve opening. A barrel diame-
ter of at least 4 in is required for a two-way hydrant,
5 in for a three-way hydrant, and 6 in for a four-
way hydrant. A minimum of two hose outlets is
required on a fire hydrant. Where pumper service
is necessary for adequate water pressure, a large
pumper outlet must be furnished. This may take
the place of one of the smaller 2
1
/2-in hose outlets.
The minimum allowable diameter for the pipe con-
nection between the main and the hydrant is 6 in.
Fire hydrants usually are either dry or wet bar-
rel, depending on the location of the main valve in
the hydrant. The main valve in the dry-barrel type
should be located below the frost line. When the
valve is in a closed position, a drain should be open
to prevent freezing of water in the barrel. The wet-
barrel, or California type, hydrants have the main
valve located near the hose outlets. Many fire
hydrants have a safety joint above the ground sur-
face to permit removal of the upper part of the bar-
rel with a minimum loss of water.
Hose connections 3
1
/16 in in diameter with 7
1
/2
threads per inch have been selected by the American
Insurance Association as standard to allow for inter-
change of fire-fighting equipment between cities.
Friction losses should not exceed 2
1
/2 psi in a
hydrant and 5 psi between the main and outlet
when flow is 600 gal/min.
21.60 Metering Devices
Metering devices are classified as either velocity or
displacement types. Velocity types measure the
velocity of flow either directly by current-measur-
ing devices or indirectly by venturi-principle
devices and are usually calibrated to indicate the
flow rate directly. The velocity-type metering
devices are applied to measurement of flows in
streams, rivers, and large pipes, such as trunk lines
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.115
of distribution systems. Displacement-type meter-
ing devices indicate flow rate directly, by recording
and integrating the rate at which their measuring
chambers are filled and emptied. Weighing meters
are also displacement-type metering devices, but
they are used primarily in laboratories. Displace-
ment types are used for the smaller flows in distri-
bution systems, such as meters for individual cus-
tomer connections.
Criteria for selection of a type of water meter
include accuracy and range of measurement,
amount of head loss through the meter, durability,
simplicity and ease of repairs, and cost.
Velocity-Type Metering Devices
I
Ven-
turi meters, or modifications thereof, are the most
common velocity-type devices. These meters pro-
duce a regular and predictable fall in the hydraulic
grade line that is related to flow rate. Three devices
that operate on this principle are the venturi, noz-
zle, and orifice plate meters shown in Fig. 21.91.
Straightening vanes are installed upstream from
these and other velocity-type meters if the pipe is of
insufficient length to eliminate helical flow compo-
nents caused by bends or other fittings.
The standard venturi meter (Fig. 21.91a) was
developed to provide a device with minimum
head loss. Since most of the loss is associated with
the diffuser section, its angle is the major factor in
determining the head loss.
Flow through a venturi meter is given by
(21.149)
(21.150)
where Q = flow rate, ft
3
/s
c = empirical discharge coefficient
dependent on throat velocity and
diameter
d
1
= diameter of main section, ft
d
2
= diameter of throat, ft
h
1
= pressure in main section, ft of water
h
2
= pressure in throat section, ft of
water
(For values of c and K for various throat diameters
and velocities, see E. F. Brater, Handbook of
Hydraulics, 6th ed., McGraw-Hill Book Company,
New York.)
As in venturi meters, flows through nozzle and
orifice-plate meters are calculated from the pres-
sure difference across the meters. Nozzle and ori-
fice-plate meters are used where conservation of
head is not the prime concern or where head dissi-
pation is desired.
Current meters consist of either a propeller or a
series of cups or vanes mounted on a shaft free to
rotate under the action of the flowing water. The
propeller type has its axis of rotation horizontal
and will not give accurate measurement unless the
current velocity is parallel to the axis of rotation.
The cup-type meter, called a Price meter, has a ver-
tical axis of rotation and measures currents whose
velocity is in any direction in a horizontal plane.
However, vertical velocity components, which do
not affect propeller meters, cause the Price meter
to indicate greater-than-actual velocities. A clicking
noise, made by the making and breaking of an
electrical contact and picked up by a set of ear-
phones, indicates the speed of rotation of the
meter. The clicking noise occurs either once each
revolution or once each five revolutions. Current
meters are used almost exclusively for stream flow,
although the propeller type is occasionally substi-
tuted for a venturi meter in pipe flow.
Displacement-Type Meters
I
These may
be piston, rotary, or nutating-disk types. The nutat-
ing disk is used, almost to the exclusion of the two
other types, for metering domestic-service connec-
tions. Its widespread use stems from its simplicity
of construction and long-term accuracy. The nutat-
ing-disk meter derives its name from the disks
nodding motion, which is similar to that of a top
before it stops. The disk is kept in motion by suc-
cessive volumes of water which enter above and
below it. A hard rubber that softens at high tem-
perature is usually used for the disk, so a backflow-
prevention device is required between a nutating-
disk meter and a water heater. Error of
nutating-disk meters is about 1.5% within the nor-
mal test-flow limits.
Compound meters contain separate measuring
devices for both low and high flows. They are usu-
ally a nutating-disk meter and a propeller-type
current meter, respectively. An automatic pressure-
sensing device directs the flow through the appro-
priate meter.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.116
I
Section Twenty-One
21.61 Water Rates
The interests of the public and individual customers
of water-supply systems can best be served by self-
sustained, utility-type enterprises. Rates charged to
finance these systems should be based on sound
engineering and economic principles and designed
to avoid discrimination between classes of cus-
tomers. Gross revenue should cover operating and
maintenance expenses, fixed charges on capital
investment, and development of the system. Billings
for water should be based on metered use and such
fixed charges as are required. Rate structures are typ-
ically based on demand, load factors, fire use, peak
rates of use, seasonal use, and similar items. The sys-
tem of accounting should conform to the legally
established system of accounting prescribed for the
utility, if any, or to some other recognized system.
Rates most commonly used today are flat rate,
step rate, and block rate.
Flat rate is a monthly or quarterly charge that
does not vary with the amount of water used. This
type of charge tends to encourage waste. Although
it has been commonly employed in small commu-
nities where water is not metered, flat rate is falling
into disuse.
Fig. 21.91 Venturi-type metering devices: (a) Standard venturi meter. (b) Nozzle meter. (c) Orifice-
plate meter.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.117
With step rate, customers are charged at one rate
per 1000 gal for all water used. The rate a customer
pays decreases as the total quantity used increases.
The major objection to this method is that a cus-
tomer who uses a quantity slightly less than the
point of rate change will pay more than the cus-
tomer who uses a little more.
The block rate schedule consists of one price per
1000 gal for the first volume or block of water used
per billing period and lesser rates for additional
blocks. This type of pricing tends to discourage
waste but does not restrict usage unnecessarily.
Both the step and block rates can have a monthly
service charge.
When fixing a system of rates, the supplier
should consider the following factors: (1) cost of
collection facilities, treatment chemicals, pumping
energy, and, where applicable, buying water from a
wholesale supplier; (2) cost of distribution and
treatment facilities; and (3) cost, including metering
and billing, of serving an individual customer. Cost
component 1, called the commodity component, is
directly dependent on total usage and therefore
should be distributed equally to all water sold. Cost
component 2, called the demand component,
depends on the peak usage of a customer. If a cus-
tomers usage is zero during peak hour, it will not
appreciably affect the cost or design of distribution
facilities. Since peak-hour demands usually govern
the design of a distribution system, this is a good
criterion for allocating distribution costs. It is gener-
ally recognized that residential areas, where the
majority of small users are, have very high ratios of
peak demand to total usage and should therefore
pay a major share of the demand component. Both
the step and block rates attempt to allocate this cost
to the small user by charging a higher rate for the
first water sold to a customer and charging decreas-
ing rates with increased usage. For most distribu-
tion systems, a large share of the demand compo-
nent also should be allocated to fire service. The
portion attributed to fire service is usually paid by
taxes. Cost component 3, called the customer com-
ponent, is usually distributed to the customer by a
monthly service charge that depends only on the
size of service. This charge is usually small.
Hydroelectric Power and Dams
Hydroelectric plants, which generate electric
power from water dropping a sufficient vertical
distance to drive large hydraulic turbines, supply
an appreciable portion of the electric power con-
sumed in the U. S. Hydroelectric generation is an
attractive power source because it is a renewable
resource and a nonconsumptive use of water. A
typical hydroelectric plant consists of a dam to
divert or store water from a river or stream; canals,
tunnels, penstocks, and a forebay to convey water
to turbines; draft tube, tunnel, or tailrace to return
water downstream to the river or stream; turbines
and governors; generators and exciters; equipment
such as protective devices and regulators; a build-
ing to house the machinery and equipment; and
transformers, switching equipment, and power
transmission lines to deliver the power produced
to a load center for distribution to consumers.
21.62 Hydroelectric-Power
Generation
Hydroelectric power is electrical power obtained
from conversion of potential and kinetic energy of
water. The potential energy of a volume of water is
the product of its weight and the vertical distance
it can fall:
(21.151)
where PE = potential energy
W = total weight of the water
Z = vertical distance water can fall
Because the kinetic energy of the supply source is
very small or zero in most hydropower (hydroelec-
tric power) developments, the kinetic-energy term
does not appear in power formulas.
Power is the rate at which energy is produced
or utilized.
1 horsepower (hp) = 550 ftlb/s
1 kilowatt (kW) = 738 ftlb/s
1 hp = 0.746 kW
1 kW = 1.341 hp
Power obtained from water flow may be computed
from
(21.152a)
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.118
I
Section Twenty-One
(21.152b)
where kW= kilowatts
hp = horsepower
Q = flow rate, ft
3
/s
w = unit weight of water = 62.4 lb/ft
3
h = effective head = total elevation dif-
ference minus line losses due to fric-
tion and turbulence, ft
= efficiency of turbine and generator
Hydroplants can be classified on a basis of
reservoir capacity and use as run-of-river hydro
without storage, base-load plants, run-of-river
plants with storage, and peak-load plants.
Run-of-River Hydro without Storage
I
This type of plant has no storage facilities. Power
generation is totally dependent on the flow of the
river. A development of this type is usually built for
some other purpose, such as navigation, power
production being only incidental.
The economics of a run-of-river hydroplant
depend on the minimum flow of the river. If the
minimum flow is very low, it will be necessary to
invest money in steam-generation facilities to pro-
vide supplemental power during low-flow peri-
ods. Therefore, the value of the plant will be only
the fuel saved that would otherwise be required
for steam generation.
Base-Load Hydro Plants
I
This type is also
a run-of-river hydroplant without storage, but it is
located on a river that provides a minimum flow
capable of serving the power demand without
supplementary steam-generating facilities. The
reliable plant capacity is set below the expected
minimum flow in the river. This type of run-of-
river hydroplant utilizes only a small proportion of
the flow of a river. It must pass not only high sea-
sonal flows but also the water it cannot utilize dur-
ing hours of low power demand.
Cost of a base-load plant can be compared with
the cost of the steam capacity that would be neces-
sary to serve power demands if hydrogeneration
were not developed.
Run-of-River Plants with Storage
I
A
small amount of storage can greatly increase the
reliable capacity of a hydroplant. The water not
required for generation during hours of low power
demand can be stored and used for generation
during periods of peak demand.
Storage can be provided for a daily, weekly, or
seasonal cycle. On a daily cycle, the required reser-
voir capacity is less than the rivers daily flow vol-
ume. On a weekly cycle, the flow during the periods
of low power demand on weekends is also stored to
give additional capacity for peak periods during the
week. On a seasonal cycle, the high flood flows are
stored to be used during periods of low flow. The
seasonal operation requires many times the storage
necessary for weekly or daily operation and there-
fore may be uneconomical unless the reservoir is
multipurpose. Then, part of its cost can be under-
written by flood-control or irrigation projects.
Peak-Load Plants
I
The power demand on
an electrical system fluctuates from a daily high to
a nightly low. Depending on the size of the utility
and type of customers served, peak demands may
be several times the magnitude of the low
demands encountered at night. These fluctuations
in demand necessitate generation facilities whose
full capacity is used only a few hours a day, during
periods of peak power demand (Fig. 21.92).
Capacity factor is the percentage of the time
the full capacity of a plant is used or the ratio of the
average power the plant produces to the plants
capacity. It can be computed on a daily, weekly, or
yearly basis.
Hydroplants that are used mainly to supply
power for the periods of peak demand are gener-
ally called peak-load plants. The main classes of
peak-load plants are pumped-storage plants and
run-of-river plants with storage.
If sufficient generating capacity and reservoir
storage are planned for a run-of-river hydroplant,
only a relatively small supply of water is needed to
produce a high generation capacity for a few hours
duration. This enables a large utility to use steam
generation at a high capacity factor where it is
most efficient and to supply peak demands from
hydroplants.
Pumped Storage
I
This is a means of storing
large quantities of energy, generated during peri-
ods when excess generating capacity is available, to
be used at some future time. Water is pumped from
a low reservoir to a higher one by energy from
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.119
steam or base-load hydro when power demand is
low. When needed, the water generates power by
flowing through a turbine back into the low reser-
voir. Because of friction loss in the penstock and
losses due to the imperfect efficiencies of pumps
and turbines, only two-thirds of the energy
required to pump the water is recovered.
The balance of energy between pumping and
generating can be on a daily or weekly basis. But
because the weekly cycle requires several times
more reservoir storage than the daily cycle, it usu-
ally is not as economical.
When pumped storage is operated at a high
capacity factor to transfer large quantities of elec-
tric energy from off-peak to peak, the energy loss
may make it uneconomical. This undesirable ener-
gy-loss feature of pumped storage is overcome
when it is used as reserve capacity.
Electrical systems require what is called spin-
ning reserve, which is capacity above that necessary
to serve the expected maximum load, ready
instantly to generate power in case of failure of
generating equipment or an unanticipated high
power demand. Many utilities keep a spinning
reserve capacity equal to the size of their largest
single generating unit, or 15% of their maximum
demand (Fig. 21.92).
(V. J. Zipparo and H. Hasen, Davis Handbook
of Applied Hydraulics, 4th ed., McGraw-Hill Book
Company, New York.)
21.63 Dams
Dams are usually classified on the basis of the type
of construction material or the method used to
resist water pressure. The main classifications are
gravity, arch, buttress, earth, and rock-fill.
Gravity dams are concrete or masonry dams
that resist the forces acting on them entirely by
their weight. Figure 21.93 shows the forces that act
on a typical gravity dam. The largest force is usual-
ly the hydrostatic force of the water F
1
. Its distribu-
tion is triangular, varying from zero at the top to
full hydrostatic at the bottom. Force F
2
represents
silt pressure, which results from deposition of silt
at the base of the dam. This silt pressure can be cal-
Fig. 21.92 Daily load curves for generating plants. (Department of Water and Power, Los Angeles, Calif.)
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.120
I
Section Twenty-One
culated by Rankines theory for earth pressure
using the submerged weight of the silt.
Force F
3
represents ice pressure against the face
of the dam. In cold climates, ice, which forms on
the reservoir surface, expands when the tempera-
ture rises and exerts a force on the top of a dam. In
the past, ice pressures as high as 50,000 psf have
been used for the design of dams in the north;
however, today it is realized these values are much
too high. A method of calculating these forces, pre-
sented by Edwin Rose, gives values ranging from
2000 to 10,000 psf, depending on the rate of tem-
perature rise and restraining conditions at the
edges of the reservoir. (E. Rose, Thrust Exerted by
Expanding Ice, Proceedings of the American Society
of Civil Engineers, May 1946.)
Practically all regions in the United States are
subject to earthquakes of varying intensity. Earth-
quakes cause vertical and horizontal accelerations of
the earth, which create forces on any object resting
on it. The magnitude of these forces equals the mass
of the object times the acceleration from the earth-
quake. These accelerations occur in every direction,
so the effect of the forces must be analyzed for all
directions. Most dams in seismically active regions in
the United States have been designed for an acceler-
ation equal to 0.1 g, where g is the acceleration due
to gravity. The effect of accelerations on the dam is
represented in Fig. 21.93 by forces F
4
and F
5
. Force F
6
represents the inertial force of the water on the face
of the dam. A close approximation of the force, given
by Eq. (21.153), was developed by von Karman.
(Pressure on Dams During Earthquakes, discus-
sion by von Karman, Transactions of the American Soci-
ety of Civil Engineers, vol. 98, p. 434, 1933.)
(21.153)
where w = unit weight of water, lb/ft
3
a = acceleration due to earthquake, ft/s
2
h = depth of water behind dam, ft
The force F
6
acts at a point 0.425h above the base.
Force F
7
is due to the weight of water on an
inclined face. Gravity dams usually have an
inclined upstream face to facilitate construction.
Force F
8
represents an uplift force that acts on
the undersurface of any section taken through the
dam or under the base of the dam. This uplift is
caused by the seepage of water through pores or
Fig. 21.93 Forces acting on a concrete gravity dam.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.121
imperfections in the foundations or through imper-
fectly bonded construction joints in the masonry. In
the past, engineers assumed that, because of bear-
ing contact, this pressure acted only on some per-
centage of the total area. Recent belief, however, is
that uplift acts on 100% of the area of the base.
A process used to reduce uplift pressures calls
for grouting along the heel and use of drains
behind the grout. When the base is not drained,
the uplift pressure is assumed to vary linearly from
between full and one-half hydrostatic pressure at
the heel to the full tailwater pressure at the toe.
Force F
9
represents the weight of the dam. It
acts at the centroid of the cross-sectional area of
the dam.
Summation of the vertical forces and of
moments about any point yields the foundation
pressure. The foundation pressure at the heel of
the dam should be compressive. Hence, the resul-
tant of all forces acting on the dam should fall
within the middle third of the base of the dam.
The basic modes of failure possible for a gravity
dam are by sliding along a horizontal plane, over-
turning by rotating about the toe, or failure of the
foundation material. The first two modes depend
mainly on the cross-sectional shape of the dam,
whereas the third depends on both the cross-sec-
tional shape and the foundation material.
Gravity dams can be built on earth foundations,
but their height in these cases has been limited to
around 65 ft. The main reason gravity dams are used
is that they can pass large flood flows over their crest
without damage. Their first cost and maintenance
cost are usually greater than those of earth or rock-
fill dams of comparable height and crest length.
Arch dams are concrete dams that carry the
force of the water through arch action. Stresses in
an arch dam may be determined with computers
by the finite-element method or by an approxi-
mate method in which the water force is divided
between elements: a series of horizontal arches
that span between the abutments and a series of
vertical cantilevers fixed at the foundation. The
distribution of load between the arches and can-
tilevers is determined by the trial-load method.
First, a division of the load is assumed and the
deflections in the arches and cantilevers are com-
puted. The deflection of an arch at any point must
equal the deflection of the cantilever at the same
point. If the deflections are not equal, a new divi-
sion of the load is assumed and the deflections
recalculated. This process is continued until equal
deflections are obtained.
The external forces an arch dam must resist are
basically the same as those on a gravity dam; how-
ever, their relative importance is much different.
On arch dams, uplift is not so important, but ice
loads and temperature stress are much more criti-
cal. Arch dams require much less concrete than
gravity dams and usually have a much lower first
cost. They are not suited to most sites, however,
since they must be located in a relatively narrow
canyon supported by good rock abutments.
Buttress dams consist of a watertight membrane
supported by a series of buttresses at right angles to
the axis of the dam. Although there are many types
of buttress dams, those widely used are the flat-slab
and the multiple-arch. These differ in that the water-
supporting membrane for the flat-slab type is a con-
tinuous concrete slab spanning the buttresses. In the
multiple-arch, the membrane is a series of concrete
arches. The multiple-arch requires less reinforcing
steel and can span longer distances between but-
tresses, but its formwork is more expensive.
The upstream face of a buttress dam is usually
inclined at about 45. The weight of the water on
the face is necessary to increase the dams resis-
tance to sliding and overturning.
The forces acting on a buttress dam are exactly
the same as those that act on a gravity dam. How-
ever, the vertical load of the water is much greater
on a buttress dam, and uplift forces are smaller.
The modes of failure are also the same, but the
structural design is much more critical.
Although buttress dams usually require less
than half the volume of concrete required by grav-
ity dams, they are not necessarily less expensive
because of the large amount of formwork and rein-
forcing steel required. With the rapidly increasing
cost of labor over the past several decades, the but-
tress dam has lost much of its earlier popularity.
Earth dams are designed to utilize materials
available at the dam site. They can be constructed
of almost any material with very primitive con-
struction equipment. Successful earth dams have
been built of gravel, sand, silt, rock flour, and clay.
If a large quantity of pervious material, such as
sand and gravel, is available and clayey materials
must be imported, the dam would have a small
impervious clay core, the material available locally
making up the bulk of the dam. Concrete has been
used for an impervious core, but it does not pro-
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.122
I
Section Twenty-One
vide the flexibility of clay materials. If pervious
material is not available, the dam can be construct-
ed of clayey materials with underdrains of import-
ed sand or gravel under the downstream toe to col-
lect seepage and relieve pore pressures.
Slopes of an earth dam are rarely greater than 2
horizontal to 1 vertical and are usually about 3 to 1.
The governing criterion is usually the stability of
the slopes against slide-out failure. Stability under
the action of seismic forces is especially critical. For
soils in which pore pressure changes develop as a
result of shear strain induced by an earthquake,
determination of appropriate values for yield
acceleration is very difficult. For some types of soil,
no well-defined yield acceleration exists; displace-
ments occur over a wide range of accelerations.
Another factor that sometimes determines the
steepness of the slopes is the amount of seepage
that can be tolerated. If the dam is on a pervious
foundation, it may be necessary to increase the
base width to reduce seepage. The seepage may
also be reduced by placing an impervious blanket
on the upstream side of the dam to increase the
seepage path or by using a cutoff wall in the foun-
dation, such as sheetpiling or a clay-filled trench.
Earth dams can be built to almost any height
and on foundations not strong enough for con-
crete dams. Improvements in earth-moving equip-
ment have resulted in a decreased cost for earth
dams, and rising labor costs have increased the
cost for concrete dams.
Rock-fill dams usually consist of a dumped rock
fill, a rubble cushion of laid-up stone on the upstream
face, bonding into the dumped rock, and an
upstream impervious facing, bearing on the rubble
cushion, with a cutoff wall extending into the foun-
dation. The dumped rock fill may consist of rocks
varying in size from small fragments to boulders
weighing as much as 25 tons. The fill is usually com-
pacted by dropping the rock, sometimes from as high
as 175 ft. onto the fill. Sluicing of the fill with high-
pressure hoses is also used to wash fines from
between contact points of the rock and reduce settle-
ment. The rubble cushion consists of rocks individu-
ally placed to reduce the voids and provide support
for the impervious facing. The facing is usually con-
crete, or wood over concrete, although steel has been
used occasionally. The cutoff wall is usually concrete.
Rock-fill dams are generally designed empiri-
cally. Low rock-fill dams may have an upstream
face as steep as
1
/2 horizontal on 1 vertical. The
downstream face is usually 1.3 on 1, the natural
angle of repose of rock. For dams over 200 ft high,
both the upstream and downstream faces are usu-
ally on a slope of 1.3 on 1.
The major problem encountered in rock-fill dams
is large settlements that occur after construction when
the reservoir is first filled. Vertical settlements and
horizontal displacements in excess of 5% of the height
of the dam have occurred; therefore, the impervious
facing must be very flexible or damage will occur dur-
ing settlement. One solution to this problem has been
to put a temporary facing on the dam and to replace
it with a permanent facing after settlement has taken
place. Temporary facings are usually of wood.
Rock-fill dams are used extensively in remote
locations where cement is expensive and the mate-
rials for an earth dam are not available. Their cost
compares favorably with that of concrete dams.
Leakage should be expected, but rock-fill dams are
very stable and have been overtopped without
suffering major damage.
(V. J. Zipparo and H. Hasen, Davis Handbook
of Applied Hydraulics, 4th ed., McGraw-Hill Book
Company, New York; Design of Small Dams and
Embankment Dams, U. S. Bureau of Relamation;
Earth and Rockfill Dams: General Design and
Construction Considerations, EM 1110-2-2300, U.
S. Army Corps of Engineers.)
21.64 Hydraulic Turbines
In the past, hydraulic power-generating machines
meant a large number of different types of equip-
ment. Today, however, the turbine is the only type
of importance in hydraulic power generation. Its
function is transformation of the kinetic and
potential energy of water into useful work.
Turbines are classified as impulse turbines and
reaction turbines.
Impulse turbines utilize the energy of water
by first transforming it into kinetic energy, by free
discharge of the water through a nozzle. The noz-
zle is directed at buckets positioned along the
perimeter of a water wheel. The force of the water
striking these buckets causes the wheel to rotate,
providing power.
The only type of water wheel used today in
impulse turbines was developed in 1880 by Pelton
the Pelton wheel (Fig. 21.94). The wheel is covered
by a housing to prevent splashing and to guide the
discharge after the water strikes the wheel.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.123
In most impulse turbines, the water wheel
rotates on a horizontal shaft and is acted on by the
discharge from one or two nozzles. But vertical
shafts may be used with as many as six nozzles, to
obtain a high efficiency for very low loads. In such
installations, efficiencies of 92% for full load and
slightly below 90% for loads as low as one-quarter
of full load have been obtained.
Impulse turbines are commonly used for heads
greater than 1000 ft. (An impulse turbine at the Reis-
seck Power Plant in Austria operates under a net
effective head of 5800 ft.) There is no lower limit of
head for impulse turbines. They have been used for
heads as low as 50 ft; however, the reaction turbine
is usually better suited to low heads at large flows.
Reaction Turbines
I
Types of reaction tur-
bines include the Francis (Fig. 21.95a), the pro-
peller-type (Fig. 21.95b) and the axial flow (Fig.
21.95c). In these, the flow from the headwater to
the tailwater is in a closed conduit system.
The Francis turbine usually consists of four
essential parts: scroll case, wicket gates, runner,
and draft tube.
The scroll case transfers the water from the pen-
stock (supply pipe) to the wicket gates and runner. It
distributes the water so that all points on the perime-
ter of the runner receive the same quantity of water.
The wicket gates, located just outside the
perimeter of the runner, control the amount of
water that enters the turbine. When the power
demand on the turbine changes, a governor actu-
ates a mechanism that opens or closes the gates.
The runner is the part of the turbine that trans-
forms the pressure and kinetic energy of the water
into useful work. As the water flows through the tur-
bine, it changes direction. This creates a force on the
runner, causing it to rotate and turn the generator.
The draft tube is a conical tube with diverging
sides. It decelerates the flow discharged from the
runner, so that the remaining kinetic energy may
be regained by conversion into suction head.
Francis turbines have a maximum efficiency of
about 94% when operated at or close to full load.
However, if the load drops below 50%, their effi-
ciency decreases rapidly. Francis turbines are com-
monly used for heads between 100 and 1000 ft. At
heads above 1000 ft. problems are encountered in
controlling cavitation and in building a scroll case
to take the high pressures. At heads below 100 ft.
the propeller-type turbine is usually more efficient.
Propeller Turbines
I
There are two types of
propeller turbines: the movable-blade type, such as
the Kaplan turbine, and the fixed-blade type. The
only difference between the two is that the pitch of
the propeller blades is adjustable in a Kaplan turbine.
The propeller turbine (Fig. 21.95) has the same
basic parts as the Francis turbine: scroll case, wick-
et gates, runner, and draft tube. The basic differ-
ence between the Francis turbine and the propeller
turbine is in the shape of the runner. The runner of
a propeller-type turbine operates in the same man-
ner as a fan or a ships propeller: The water mov-
ing past the blades creates a force that causes the
runner to rotate.
Propeller-type turbines are used for heads
ranging from a few feet to about 100 ft. The Kaplan
turbine has an efficiency of about 94% for full load
and drops only to 92% for 40% load. The fixed-
blade-type turbine also has an efficiency of about
94% for full load; however, its efficiency drops off
rapidly below full load.
Axial-Flow Turbines
I
These provide
enhanced performance for operation under low-
head and large capacity.
(V. J. Zipparo and H. Hasen, Davis Handbook
of Applied Hydraulics, 4th ed., McGraw-Hill Book
Company, New York.)
21.65 Methods for Control of
Flows from Reservoirs
Any reservoir with an appreciable drainage area
must have a spillway to discharge flood flows with-
Fig. 21.94 Impulse (Pelton) type of hydraulic
turbine.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.124
I
Section Twenty-One
out damage to the dam and to keep the reservoir
water surface below some predetermined level.
21.65.1 Spillways
An overflow spillway allows water to pass over the
crest of a section of the dam. This type of spillway is
widely used for concrete dams because, if designed
correctly, the dam will not be damaged by the water.
To use an overflow spillway for earth or rock-fill
dams, it is necessary to make the spillway a concrete
gravity section. This may not be possible for high
earth dams because the foundation may not be able
to support a high concrete gravity section.
Fig. 21.95 Reaction types of hydraulic turbines: (a) Francis; (b) Kaplan; (c) axial flow.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.125
The discharge over an overflow spillway is
given by the equation for discharge over a weir
(Art. 21.34). Since the discharge varies as the head
to the
3
/2 power, overflow spillways keep the water
level within close limits even when there is a large
variation in flows.
It is desirable for an overflow spillway to have
the form of the underside of the nappe of a sharp-
crested weir. This type of spillway, called an ogee
spillway, should be designedas should all spill-
waysso that separation of the water from the
face of the spillway will not occur. Thus, the dan-
ger of cavitation will be eliminated.
In a chute spillway, water flows over a crest
into a steeply sloping, lined, open channel. The
flow is made supercritical to keep the size and
length of the chute to a minimum. Gradual vertical
curves should be used in the chute to avoid sepa-
ration of the flow from the channel bottom.
Chute spillways are commonly used for earth
and rock-fill dams where the topography allows a
chute to carry the water away from the toe to elim-
inate the danger of undermining. The discharge
over the crest is given by the equations for discharge
over a weir or the entrance to an open channel.
In a side-channel spillway, the flow passes
over a crest into a channel parallel to this crest. The
crest is usually a concrete gravity section, although
it can be concrete laid on the natural embankment.
Side-channel spillways are often used in narrow
canyons where it is not possible to obtain sufficient
crest length for overflow or chute spillways. The
flow in the channel parallel to the crest is deter-
mined by applying the momentum principle in the
direction of flow and assuming the energy of the
water flowing over the crest is completely dissipat-
ed (U.S. Bureau of Reclamation, Design of Small
Dams, Government Printing Office, Washington,
DC 20402).
In a shaft spillway, sometimes called a morn-
ing-glory spillway, the water flows over a circular
weir into a vertical shaft. The shaft terminates in a
horizontal conduit that carries the water past the
dam. The weir can be sharp-crested, flared, or
ogee in cross section. (This type of spillway should
not be constructed over or through earth dams.) If
the topography is not suitable for a chute or side-
channel spillway, a shaft spillway may be the best
alternative.
There are two conditions of discharge for a
shaft spillway, both depending on the head on the
weir. When the head is relatively low, the dis-
charge is governed by the flow over the weir,
which is directly proportional to the
3
/2 power of
the head on the weir. As the head increases, at
some point the discharge will no longer be con-
trolled by the amount of water that can flow over
the weir but by the amount of water that can flow
through the conduit. The discharge for this condi-
tion is directly proportional to the
1
/2 power of the
elevation difference between the reservoir water
level and the level of discharge of the spillway con-
duit. Once this second condition is reached, a large
increase in head will cause only a small increase in
flow. Since analytical analysis of discharge does
not give good results on this type of spillway,
model tests are usually employed.
A siphon spillway (Fig. 21.96) is a closed conduit
for discharging water over or through a dam. The
entrance to a siphon spillway is usually submerged
below the normal water level so that it will not clog
with debris or ice. The discharge end of the siphon
is usually sealed by deflecting the flow across the
barrel or by submerging it so that air cannot enter.
The air vent shown in Fig. 21.96 determines the
reservoir level at which the siphon flow begins.
When the reservoir water level rises above the vent,
the siphons intake is sealed. Water flowing over the
crest of the siphon removes the air in the siphon and
full flow begins. Because the flow depends on the
siphoning action, siphon spillways hold the water
Fig. 21.96 Siphon spillway.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
21.126
I
Section Twenty-One
level of a reservoir within close limits. But they are
not good for handling large variations in flows
because their discharge is directly proportional to
the square root of the head. They are relatively
expensive because of the cost of forming the barrel.
21.65.2 Intake Structures
The various functions an intake structure may serve
include permitting withdrawal of water from vari-
ous levels of a reservoir, controlling flow, excluding
debris and ice from a conduit, and providing sup-
port for the conduit. The type of intake structure
required depends on the functions and characteris-
tics of the reservoir. The simplest type of intake is a
block of concrete supporting the end of a conduit
equipped with a bar screen to exclude foreign mat-
ter. In contrast, the intake towers at Hoover Dam,
which serve 30-ft-diameter penstocks, are 395-ft-
high concrete towers, with two 32-ft-diameter cylin-
der gates under a maximum head of over 300 ft.
Intake towers are commonly used where there
is a large fluctuation in the water level of a reser-
voir or where it is necessary to control the quality
of water used for a domestic supply. They are usu-
ally made of concrete and have ports at various
levels to permit selection of water from different
elevations. The ports are usually provided with
gates or valves and some type of trash rack.
The main hydraulic consideration in the design
of an intake is to keep losses to a minimum. To do
this, the velocities through the trash racks should
be kept less than 0.5 ft/s, and the standard rules for
reducing hydraulic losses should be observed.
21.65.3 Crest Gates
These include a number of different types of per-
manent and temporary devices that operate on the
crest of spillways to increase the storage of a reser-
voir temporarily while control of spillway flows is
retained. During periods of low flow when the full
spillway capacity is not required, the additional
head and storage gained with crest gates may be
very valuable.
Flashboards and stop logs are the most com-
mon types of crest gates used for small installations
under low head. Flashboards are usually wood
planks that span between vertical pipes that can-
tilever above the spillway crest. When the reser-
voir water surface reaches some predetermined
level, the pipes fail, allowing the full capacity of the
spillway to be utilized. Stop logs are wood planks
that span between slotted vertical piers which can-
tilever above the spillway crest.
On large stop-log installations, the hydrostatic
force creates large frictional forces between the
sliding element and the vertical guide, making
removal difficult. These frictional forces make it
necessary to use a type of gate that depends on
rolling rather than sliding friction and operates
freely under hydrostatic pressure.
Taintor gates and sliding gates mounted on low-
friction roller bearings are the most widely used
types of crest gates on major installations. In a tain-
tor gate (Fig. 21.97), the friction is concentrated in the
trunnion and does not affect the operation. Since
flow passes under taintor and slide gates, there is a
tendency for ice and trash to pile up against them,
causing damage and hampering operation.
Fig. 21.97 Taintor gate.
Fig. 21.98 Bear-trap gate.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Water Resources Engineering
I
21.127
Bear-trap and drum gates allow the flow to
pass over the top. The bear-trap gate consists of
two leaves hinged, as shown in Fig. 21.98. To raise
a bear-trap gate, water is admitted to the space
under the leaves to force the leaves up. The drum
gate (Fig. 21.99) consists of a segment of a cylinder
that is lowered into a recess in the crest when not
in use. Because of the large recess required in the
dam, drum gates are not suited to small dams.
(V. J. Zipparo and H. Hasen, Davis Handbook of
Applied Hydraulics, 4th ed., and H. E. Babbitt, J. J.
Doland, and J. L. Cleasby, Water Supply Engineer-
ing, McGraw-Hill Book Company, New York.)
Fig. 21.99 Drum gate.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
Copyright (C) 1999 by The McGraw-Hill Companies, Inc. All rights reserved. Use of
this product is subject to the terms of its License Agreement. Click here to view.
blank page 21.128

S-ar putea să vă placă și