Download as pdf or txt
Download as pdf or txt
You are on page 1of 118

FACULTY OF ENGINEERING

LECTURE NOTES FOR


WAR2204 & AMI3203 ENGINEERING HYDRAULICS

Prepared by
OKETCHO YORONIMO
oketchoyoronimo@gmail.com 0392934088

1
Short Description
The course is intended to provide to the students an Introduction to the fundamental principles of
Hydraulics, including fluid flows, Flow through Pipelines, Open Channel Hydraulics, Application of
Hydraulic principles in the design of Hydraulic structures and selection of Hydraulic Machinery.
Course Objectives
• To improve students understanding of the principles of fluid mechanics towards its application in
hydraulic analyses, and design of hydraulic structures.
• To make students understand the hydraulics concepts and their application in hydraulic designs and
engineering applications.
Learning outcomes
• Students should be able to apply the principles of fluid mechanics in hydraulic analysis
• Students should be able to apply to apply the hydraulic concepts in the design of hydraulic structures
and machines
Teaching and Learning Pattern
The teaching of students will be conducted through lectures, tutorials, short classroom exercises and group
discussions and laboratories/field activates.
Assessment Method
Assessment will be done through:
• Coursework which includes assignments, tests and practical work. Course work will carry a total
of 40%. Coursework marks will be divided into; Assignments 10%, Tests 20% and
laboratories/field activates 10%
• Written examination. Written examination will carry 60%.
Detailed course content
Topics Duration(hrs)
1. Open channel flow 6
1.1.Differences between open channel and pipe flow
1.2.Characteristics of open channel flow
1.3.Engineering applications of open channel flow
1.4.Classification of open channel flow
1.5.Open channel flow analysis
1.6.Steady flow analysis
1.7.Varied flow analysis
1.8. Specific energy and hydraulic jump
2. Flow in pipes 7
2.1. Laminar flow

2
2.2. Losses in pipes and fittings
2.5. expansion and contraction losses
2.6. surface roughness and the frictional formulae
2.7. Analysis of simple pipe networks
2.8. complex pipe network analysis
3. Hydraulic machines 15
3.1. Turbines
3.2. Pumps
3.3. Cavitation and water hammer effects
3.4. Pump selection
4. Hydraulic structures 15
4.1. Types of dams: Gravity, earth and rock fill dams.
4.2. Design and construction of dams
4.3.Weirs
4.4.Culverts
4.6. Spillways
4.7.Intakes
4.8.Gates
4.9.Stilling basins
5.0Laboratory /fieldwork 30 hours 30
Tutorials 15
Reading List
1. Les Hamill, (2002): Understanding Hydraulics. Palgrave Macmillan. .
2. Jacob Bear (2007): Hydraulics of Groundwater. Dover Publications.
3. B.S.Massey, (1988), Mechanics of fluids, 6th.Edition
4. B.S.Massey ,(1998), Mechanics of fluids, 7th.Edition
Andrew Chadwick, John Morfett and Martin Borthwick (2013), HYDRAULICS IN CIVIL AND
ENVIRONMENTAL ENGINEERING, 5th Edition.

3
Contents
Short Description ......................................................................................................................................................... 2
CHAPTER ONE: OPEN CHANNEL FLOW ........................................................................................................................ 7
Open channel flow (Flow with a Free Surface) ........................................................................................................ 7
1.1 Flow Classification .............................................................................................................................................. 8
1.2 Natural and Artificial Channels and Their Properties ........................................................................................ 9
1.3 Geometric properties of natural and artificial channels .................................................................................. 10
1.4 Optimum cross-section ....................................................................................................................................11
1.5 Velocity Distributions, Energy and Momentum Coefficients...........................................................................12
1.6 Laminar and Turbulent flow............................................................................................................................. 13
1.7 Uniform Flow ................................................................................................................................................... 14
1.7.1 Development of Friction Formulae ...........................................................................................................14
1.7.3 Manning Equation .....................................................................................................................................15
1.7.4 Uniform Flow Computations ..................................................................................................................... 17
1.7.5 Channel conveyance .................................................................................................................................18
1.7.6 Compound Channels .................................................................................................................................19
1.8 Rapidly Varied Flow: The Use of Energy Principles .................................................................................... 23
1.8.1 Energy Equation in Open Channels ...........................................................................................................23
1.8.2 Application of the Energy Equation ..........................................................................................................24
1.9 Subcritical, Critical and Supercritical Flow .......................................................................................................27
1.9.1 General Equation of Critical Flow .............................................................................................................28
1.9.3 Critical Depth and Critical Velocity (for a Rectangular Channel) .............................................................. 29
1.9.4 Froude Number .........................................................................................................................................29
1.10 Rapidly Varied Flow: The Use of Momentum Principle .................................................................................32
1.10.1 Hydraulic Jump........................................................................................................................................32
1.10.2 Solution of the Momentum Equation for a Rectangular Channel .......................................................... 33
1.10.3 Energy Dissipation in a Hydraulic Jump ..................................................................................................34
1.11 Gradually Varied Flow ........................................................................................................................................35
1.11.1 Significance of Bed Slope and Channel Friction .......................................................................................... 35
1.11.2 Flow Transitions ..........................................................................................................................................38
CHAPTER TWO: FLOW IN PIPES .................................................................................................................................41
2.1 Introduction ..................................................................................................................................................... 41
2.2 Fundamental Concepts of Pipe Flow ...............................................................................................................41

4
2.3 Analysis of pipelines. ........................................................................................................................................42
2.3.1 Pressure loss due to friction in a pipeline. ................................................................................................ 43
2.3.2 Pressure loss during laminar flow in a pipe .............................................................................................. 45
2.3.3 Pressure loss during turbulent flow in a pipe ........................................................................................... 45
2.3.4 Choice of friction factor f .......................................................................................................................... 47
2.3.5 The value of f for Laminar flow .................................................................................................................47
2.4 Local Head Losses ............................................................................................................................................55
2.4.1 Head loss due to a sudden expansion.......................................................................................................55
2.4.2 Losses at Sudden Contraction ................................................................................................................... 57
2.4.3 Other Local Losses.....................................................................................................................................57
2.5 Pipeline Analysis...............................................................................................................................................59
2.5.1 Flow in pipes with losses due to friction. ..................................................................................................62
2.5.2 Pipes in Series Example ............................................................................................................................. 64
2.5.3 Pipes in parallel .........................................................................................................................................65
2.5.4 Branched pipes..........................................................................................................................................68
2.6 Complex pipe network analysis ....................................................................................................................... 72
2.6.1 Loop (head balance) Method .................................................................................................................... 73
2.6.2 Nodal Method ...........................................................................................................................................76
2.7 Complex Networks ...........................................................................................................................................77
3.1 Introduction ..................................................................................................................................................... 78
3.2 Energy Transfer in Pumps and Turbines ..........................................................................................................78
3.3 Types of Pumps and Turbines .......................................................................................................................... 79
3.4 Pump and System Characteristics .................................................................................................................... 82
3.4.1 Pump characteristics .................................................................................................................................82
3.4.2 System characteristics............................................................................................................................... 82
3.4.3 Finding the Duty Point .............................................................................................................................. 83
3.5 Pumps in Parallel and in Series ........................................................................................................................ 84
3.6 Specific speed................................................................................................................................................... 88
3.6.1 Specific speed(pump) ................................................................................................................................ 88
3.6.2 Specific Speed for Turbines ....................................................................................................................... 89
3.7 Cavitation ......................................................................................................................................................... 90
CHAPTER FOUR: HYDRAULIC STRUCTURES ................................................................................................................91

5
4.1 Introduction ..................................................................................................................................................... 91
4.2 Dams ................................................................................................................................................................ 91
4.2.1 Gravity dams .............................................................................................................................................92
4.2.2 Arch dams .................................................................................................................................................95
4.2.3 Buttress dams ...........................................................................................................................................95
4.3 Spillways ........................................................................................................................................................... 96
4.3.1Types of spillways ......................................................................................................................................97
4.4 Culverts .......................................................................................................................................................... 100
4.4.1 Types of flow in culverts ......................................................................................................................... 101
4.4.2 Culvert analysis and design ..................................................................................................................... 104
4.5 Flow measuring structures............................................................................................................................. 107
4.5.1 Weirs ....................................................................................................................................................... 107
4.5.2 Venturi Flume..........................................................................................................................................113
4.6 Sluice gates and other control gates..............................................................................................................114

6
CHAPTER ONE: OPEN CHANNEL FLOW

Open channel flow (Flow with a Free Surface)


Open channel flow is characterized by the existence of a free surface (the water surface). In contrast to pipe
flow, this constitutes a boundary at which the pressure is atmospheric and across which the shear forces
are negligible. The longitudinal profile of the free surface defines the hydraulic gradient and determines
the cross-sectional area of flow, as shown in Figure.1. It also necessitates the introduction of a new
variable—the stage (see Figure.1)—to define the position of the free surface at any point in the channel. In
consequence, problems in open channel flow are more complex than those of pipe flow, and the solutions
are more varied, making the study of such problems both interesting and challenging.

7
Figure 1:Cmparison between open channel pipe flow
1.1 Flow Classification
Recalling that flow may be steady or unsteady and uniform or nonuniform, the major classifications
applied to open channels are as follows: Steady uniform flow, in which the depth is constant, both with
time and distance. This constitutes the fundamental type of flow in an open channel in which the gravity
forces are in equilibrium with the resistance forces. Steady non-uniform flow, in which the depth varies
with distance, but not with time. The flow may be either (a) gradually varied or (b) rapidly varied. Type
(a) requires the joint application of energy and frictional resistance equations. Type (b) requires the
application of the conservation equation and either energy or momentum principles. Unsteady flow, in
which the depth varies with both time and distance (unsteady uniform flow is very rare). This is the most
complex flow type, requiring the solution of conservation, energy, momentum and friction equations
through time. The various flow types are all shown in Figure 2.

8
Figure 2:Classification of open channel flow

1.2 Natural and Artificial Channels and Their Properties


Artificial channels comprise all man-made channels, including irrigation and navigation canals, spillway
channels, sewers, culverts and drainage ditches. They are normally of regular cross-sectional shape and
bed slope, and as such are examples of prismatic channels (which do not change their cross section in the
streamwise direction). Their construction materials are varied, but commonly used materials include
concrete, steel and earth. The surface roughness characteristics of these materials are normally well defined
within engineering tolerances. In consequence, the application of hydraulic theories to flow in artificial
channels will normally yield reasonably accurate results.

9
In contrast, natural channels are normally very irregular in shape, and their materials are diverse. The
surface roughness of natural channels changes with time, distance and water surface elevation. Therefore,
it is more difficult to apply hydraulic theory to natural channels and obtain satisfactory results. Many
applications involve man-made alterations to natural channels (e.g., river control structures and flood
alleviation measures). Such applications require an understanding not only of hydraulic theory but also of
the associated disciplines of sediment transport, hydrology and river morphology.
1.3 Geometric properties of natural and artificial channels
In the case of artificial channels, these may all be expressed algebraically in terms of the depth (y), as
shown in Table 5.1. This is not possible for natural channels, so graphs or tables relating them to stage (h)
must be used.

The commonly used geometric properties are defined as follows:


Depth (y)—the vertical distance of the lowest point of a channel section from the free surface Stage
(h)—the vertical distance of the free surface from an arbitrary datum
Area (A)—the cross-sectional area of flow normal to the direction of flow
Wetted perimeter (P)—the length of the wetted surface measured normal to the direction of flow
Surface width (B)—the width of the channel section at the free surface
Hydraulic radius (R)—the ratio of area to wetted perimeter (A/P)
Hydraulic mean depth (Dm)—the ratio of area to surface width (A/B)

10
EXAMPLE 1.1
A trapezoidal channel has a bottom width of 5.0m and its sides slope at an angle of 45°. If the depth of
flow is 2.0m, calculate the area of flow A, the wetted perimeter P, and the hydraulic radius R.
Soln:

tan 450 = 𝑥/2.0, 𝑠𝑜 𝑥 = 2 tan 450 =


2.0𝑚

Hence 𝐵𝑠 = 2.0 + 5.0 + 2.0 = 9.0𝑚


1
𝐴= (5.0 + 9.9)2.0 = 14.0𝑚2
2
Let length of wetted side slopes =y
cos 450 = 2.0/𝑦
So, cos 450 = 2.0/𝑦
𝑦 = 2.0/𝑐𝑜𝑠450 = 2.828𝑚
𝑃 = 2.828 + 5.000 + 2.828 = 10.656𝑚, 𝑅 = 𝐴/𝑃 = 14.0/10.656 = 1.314𝑚
1.4 Optimum cross-section
The most efficient cross-section of an open channel, from hydraulic view point, is the one which for a given
slope, roughness coefficient and the area of flow, carries the maximum flow rate.
From among all open channel cross-sections, for a given area of flow, the semi-circular channel has the
smallest wetted perimeter and is therefore the section of highest efficiency.
Hydraulically efficient rectangular channel;
Consider the rectangular channel below.

Area of flow, 𝐴 = 𝑏𝑦
And wetted perimeter, 𝑃 = 𝑏 + 2𝑦 = 𝐴/𝑦 + 2𝑦
For p to be minimum, 𝑑𝑝/𝑑𝑦 should be zero; differentiating p wrt y, keeping A as a constant,

11
𝑑𝑝 𝐴
=− 2+2
𝑑𝑦 𝑦
𝒅𝒑
= 0, we obtain 𝐴 = 2𝑦 2
𝒅𝒚

But 𝐴 = 𝑏𝑦 and therefore 𝑏 = 2𝑦


also
𝐴 2𝑦 2
𝑅= = = 𝑦/2
𝑃 4𝑦
Thus, for a hydraulically most efficient rectangular channel, the width should be twice the depth of flow,
so that the hydraulic radius is half the depth of flow.
Task:
Derive the expressions for the dimensions for a hydraulically most efficient trapezoidal channel.
1.5 Velocity Distributions, Energy and Momentum Coefficients
The point velocity in an open channel varies continuously across the cross section because of friction along
the boundary. However, the velocity distribution is not axisymmetric (as in pipe flow) due to the presence
of the free surface. One might expect to find the maximum velocity at the free surface, where the shear
stress is negligible, but the maximum velocity normally occurs below the free surface. Typical velocity
distributions are shown in Figure.3 for various channel shapes. The depression of the point of maximum
velocity below the free surface may be explained by the presence of secondary currents which circulate
from the boundaries towards the channel center.
The energy and momentum coefficients (α and β) defined in can only be evaluated for a channel if the
velocity distribution has been measured. For turbulent flow in regular channels, α rarely exceeds 1.15 and
β rarely exceeds 1.05. In consequence, these coefficients are normally assumed to be unity. However, in
irregular channels where the flow may divide into distinct regions, α may exceed 2 and should therefore be
included in flow computations. Referring to Figure 3, which shows a natural channel with two flood banks,
the flow may be divided into three regions. By making the assumption that α = 1 for each region, the value
of α for the whole channel may be found as follows:
𝑢3 𝑑𝐴 𝑉1 3 𝐴1 +𝑉2 3 𝐴2 +𝑉3 3 𝐴3
𝛼= ̅3𝐴
= ̅ 3 (𝐴1 +𝐴2 +𝐴3 )
………….1.1
𝑉 𝑉

12
Figure 3 Velocity distribution in open channels (Contour numbers are expressed as a percentage of maximum velocity)

Figure 4 Division of a channel into a main channel and flood banks.


𝑄 𝑉 𝐴 +𝑉 𝐴 +𝑉 𝐴
𝑉̅ = = 1 1 2 2 3 3…………………1.2
𝐴 𝐴1 +𝐴2 +𝐴3

1.6 Laminar and Turbulent flow


Open channel flow may be either laminar or turbulent, as in pipe flow. The criterion for determining
whether the flow is laminar or turbulent is the Reynolds number (Re).

These results can be applied to open channel flow if a suitable form of the Reynolds number can be found.
This requires that the characteristic length dimension, the diameter (for pipes), be replaced by an equivalent
characteristic length for channels. The one adopted is termed the hydraulic radius (R). Hence, the Reynolds
number for channels may be written as
𝜌𝑅𝑉
𝑅𝑒(𝑐ℎ𝑎𝑛𝑛𝑒𝑙) = … … … … … … … … … 1.3
𝜇
𝑹𝒆(𝒑𝒊𝒑𝒆)
For a pipe flowing ful, R=D/4, so 𝑹𝒆(𝒄𝒉𝒂𝒏𝒏𝒆𝒍) =
𝟒

and for laminar flow, Re(channel)<500


and for turbulent channel flow Re(channel)> 1000
In practice, the upper limit of Re is not so well defined for channels as it is for pipes and is normally taken
to be 2000.
13
The Darcy–Weisbach formula for pipe friction was introduced, and the relationship between laminar,
transitional and turbulent flow was depicted on the Moody diagram. A similar diagram for channels has
been developed. Starting from the Darcy–Weisbach formula,
𝜆𝐿𝑉 2
ℎ𝑓 = … … … … … … . .1.4
2𝑔𝐷
and making the substitutions R = D/4 and hf/L = S0 (where S0 = bed slope), then for uniform flow in an
open channel S V g R 0 2 2 4 = λ
𝜆𝑉 2
𝑠0 = … … … … … … … 1.5
2𝑔4𝑅
8𝑔𝑅𝑠0
𝜆= … … … … .1.6
𝑉2
The λ–Re relationship for pipes is given by the Colebrook–White transition law and by substituting R =
D/4 the equivalent formula for channels is
1 𝑘𝑠 0.6275
= −2𝑙𝑜𝑔 ( + ) … … … … . .1.7
√𝜆 14.8𝑅 𝑅𝑒 √𝜆
combination of 1.6 and 1.7 yields
𝑘𝑠 0.6275𝜈
𝑉 = −2√8𝑔𝑅𝑆0 log ( + ) … … … . .1.8
14.8𝑅 𝑅√8𝑔𝑅𝑆0

In practice, the flow in open channels is normally in the rough turbulent zone, and consequently it is
possible to use simpler formulae to relate frictional losses to velocity and channel shape, as discussed in
the next section.
1.7 Uniform Flow
1.7.1 Development of Friction Formulae
For uniform flow to occur, the gravity forces must exactly balance the frictional resistance forces which
constitute the boundary shear force. Figure 4 shows a small longitudinal section in which uniform flow
exists.
The gravity force resolved in the direction of flow = 𝝆𝒈𝑨𝑳 sin𝜽 and the shear force resolved in the direction
of flow = 𝝉𝟎 𝑷𝑳, where 𝜏0 is the mean boundary shear stress. Hence
𝜏0 𝑃𝐿 = 𝜌 𝑔 𝑠 𝐴𝐿 𝑖𝑛 𝜃 … … … … . .1.9
Considering channels of small slope only, then
𝑠𝒊𝒏𝜽 ≅ 𝒕𝒂𝒏𝜽 = 𝑺𝟎
hence,

14
𝜌𝑔𝐴𝑆0
𝜏0 = … … … … … … .1.10
𝑃
1.7.2 Chézy Equation
To interpret the above equation, an estimate of the magnitude of τ0 is required. Assuming a state of rough
turbulent flow, then
τ0 ∝ 𝑉 2 𝑜𝑟 τ0 = 𝐾𝑉 2 … … … 1.11

Figure 5 derivation of uniform flow equations

Substituting into the above equation for 𝝉𝟎


𝜌𝑔
𝑉 = √ 𝑅𝑆0 … … … … … … … 1.12,
𝐾

which may be written as


𝑉 = 𝐶 √𝑅𝑆0 … … … … … … … .1.13
This is known as the Chézy equation. It is named after the French engineer who developed the formula
when designing a canal for the Paris water supply in 1768. The Chézy coefficient C is not, in fact, constant
but depends on the Reynolds number and boundary roughness, as can be readily appreciated from the
previous discussions of the λ–Re diagram.
A direct comparison between C and λ can be found by substituting (1.6) into (1.13) to yield
𝑪 = √𝟖𝒈𝒇𝝀
1.7.3 Manning Equation
One of the most widely used discharge equations is that attributed to Manning in 1889. This was developed
from empirical observations. Manning found that the Chezy coefficient, C, could be expressed as
1
𝑅6
𝐶= … … … . .1.14
𝑛
This formula was developed from seven different formulae and was further verified by 170 observations.
Other research workers in the field derived similar formulae independently of Manning, including Hagen
in 1876, Gauckler in 1868 and Strickler in 1923. In consequence, there is some confusion as to whom the
equation should be attributed to, but it is generally known as the Manning equation

15
1
𝑉 = ( ) 𝑅 2/3 𝑆01/2
𝑛

and the equivalent formula for discharge is


1 𝐴5/3
𝑄= 𝑆 1/2 … … … . .1.15
𝑛 𝑃2/3 0

where n is a constant known as Manning’s n (it is numerically equivalent to Kutter’s n). It should be noted
that n is taken to be a constant, independent of units.
The value of the roughness coefficient n determines the frictional resistance of a given channel. It can be
evaluated directly by discharge and stage measurements for a known cross section and slope. However, for
design purposes, this information is rarely available, and it is necessary to rely on documented values
obtained from similar channels. For the case of artificially lined channels, n may be estimated with
reasonable accuracy. For natural channels, the estimates are likely to be rather less accurate.

EXAMPLE 1.2. A rectangular river channel 4.6m wide carries water at a depth of 0.6m. The slope of the
channel is 1 in 400. The channel has a poor alignment and the bed is covered with stones about 75mm to
150mm in size. Using the range of n values in Table 8.1, investigate the range of discharge that results
from the application of the Manning equation to the channel. What is the percentage difference in the flows,
calculated as a proportion of the smaller value?
Solution:
𝐴 = 4.6 × 0.6 = 2.76𝑚2
𝑃 = 4.6 + (2 × 0.6) = 5.8𝑚
𝑅 = 𝐴/𝑃 = 2.76/5.8 = 0.476𝑚

16
𝑆0 = 1/400 = 0.0025
Form the table of n values, n=0.040 to 0.08
With n=0.040
2
𝑉 = (1/𝑛)𝑅 3 𝑆01/2
2
𝑉 = (1/0.040)0.476 3 0.00251/2
= 0.76𝑚/𝑠
𝑄 = 𝐴𝑉 = 2.76 × 0.76 = 2.10𝑚3 /𝑠
With n=0.08, Q=1.05m3/s
Therefore, the difference between the two values is 100%, that is the larger value is twice the smaller value.
This illustrates that while the Manning equation may be reasonably accurate, it may be difficult to estimate
the values of the variables, so the answer obtained is nothing more than an estimate.
1.7.4 Uniform Flow Computations
Manning’s formula may be used to determine steady uniform flow. There are two types of commonly
occurring problems to solve. The first is to determine the discharge given the depth, and the second is to
determine the depth given the discharge. The depth is referred to as the normal depth, which is synonymous
with steady uniform flow. As uniform flow can only occur in a channel of constant cross section, natural
channels should be excluded. However, in solving the equations of gradually varied flow applicable to
natural channels, it is still necessary to solve Manning’s equation. Therefore, it is useful to consider the
application of Manning’s equation to irregular channels in this section. The following examples illustrate
the application of the relevant principles.
Example 1.3. Discharge from Depth for a Trapezoidal Channel
The normal depth of flow in a trapezoidal concrete lined channel is 2 m. The channel base width is 5 m
and has side slopes of 1:2. Manning’s n is 0.015 and the bed slope, S0, is 0.001. Determine the discharge
(Q), mean velocity (V) and the Reynolds number (Re).
𝑨 = (𝟓 + 𝟐𝒚)𝒚, 𝑷 = 𝟓 + 𝟐𝒚√𝟏 + 𝟐𝟐 , hence applying equation 1.15 for y=2m
𝟏 [(𝟓 + 𝟒)𝟐]𝟓/𝟑
𝑸= × × 𝟎. 𝟎𝟎𝟏𝟏/𝟐
𝟎. 𝟎𝟏𝟓 [𝟓 + (𝟐 × 𝟐√𝟓]𝟐/𝟑

= 𝟒𝟓𝒎𝟑 /𝒔
To find the mean velocity, simply apply the continuity equation:
𝑸 𝟒𝟓
𝑽= = = 𝟐. 𝟓𝒎/𝒔
𝑨 (𝟓 + 𝟒)𝟐

17
𝝆𝑹𝑽
The Reynolds number is given by, 𝑹𝒆 =
𝝁
𝑨
Where 𝑹=
𝑷
In this case,
(𝟓 + 𝟒)𝟐
𝑹= = 𝟏. 𝟐𝟗𝒎
[𝟓 = (𝟐 × 𝟐√𝟓)
𝟏𝟎𝟑 ×𝟏.𝟐𝟗×𝟐.𝟓
and 𝑹𝒆 = = 𝟐. 𝟖𝟑 × 𝟏𝟎𝟔
𝟏.𝟏𝟒×𝟏𝟎−𝟑

Note: Re is very high and corresponds to the rough turbulent zone. Therefore, Manning’s equation is
applicable. The interested reader may care to check the validity of this statement by applying the
Colebrook–White equation. First, calculate a kS value equivalent to n = 0015 for y = 2m [kS = 2.225 mm].
Then using these values of kS and n compare the discharges as calculated using the Manning and
Colebrook–White equations for a range of depths. Provided the channel is operating in the rough turbulent
zone, the results are very similar

Example 1.4. Depth from Discharge for a Trapezoidal Channel


If the discharge in the channel given in Example 1.3 were 30 m3/s, find the normal depth of flow.
Solution
𝟏 [(𝟓 + 𝟐𝒚)𝒚]𝟓/𝟑
𝑸= × × 𝟎. 𝟎𝟎𝟏𝟏/𝟐
𝟎. 𝟎𝟏𝟓 [𝟓 + 𝟐√𝟓𝒚]𝟐/𝟑

[(𝟓+𝟐𝒚)𝒚]𝟓/𝟑
Or 𝑸 = 𝟐. 𝟏𝟎𝟖 ×
[𝟓+𝟐√𝟓𝒚]𝟐/𝟑

At first sight this may appear to be an intractable equation, and it will also be different for different channel
shapes. The simplest method of solution is to adopt a trial-and-error procedure. Various values of y are
tried, and the resultant Q is compared with that required. Iteration ceases when reasonable agreement is
found. In this case y < 2 as Q < 45, so an initial value of 1.7 is tried.
Trial depth(m) Resultant Discharge(m3/s)
1.7 32.7
1.6 29.1
Hence, the solution is 1.63 30.1 y=1.63 for Q=30m3/s
1.7.5 Channel conveyance
Channel conveyance (K) is a measure of the discharge carrying capacity of a channel, defined by the
equation
𝑄 = 𝐾𝑆01/2 … … … … … … … … … … … 1.16

18
For any given water depth (or stage), its value may be found by equating this with Manning’s equation to
give
𝐴5/3
𝐾=
𝑛𝑃2/3
Its principal use is in determining the discharge and the energy and momentum coefficients in compound
channels. It is also a convenient parameter to use in the computational procedures for evaluating gradually
varying (steady and unsteady) flow problems in compound channels. The equation for the energy
coefficient α in a compound channel may be expressed in general terms as
3
∑𝑁 𝑖=1 𝑉𝑖 𝐴𝑖
𝛼= … … .1.17
𝑉̅ 3 ∑𝑁𝑖=1 𝐴𝑖
where N is the number of subsections
Similarly, it may be shown that

∑𝑁 𝑁 𝐾2
𝑖=1 𝐴𝑖 𝑖
𝛽= 2∑
… … . .1.18
( ∑𝑁 𝐾
𝑖=1 𝑖 ) 𝑖=1 𝐴 𝑖

Thus, both α and β can be evaluated for any given stage without explicitly determining Q i
1.7.6 Compound Channels
So far it has been assumed that there is only one main channel, be it rectangular, trapezoidal or circular.
However, natural rivers in flood usually comprise a main channel with a floodplain on each side. This is
called a compound or two stage channels. Within a compound channel it is unlikely that the roughness will
be uniform around the entire wetted perimeter. Thus, it is necessary to deal with both the compound nature
of the channel and the variable roughness. Although there are simple techniques that can be employed, it
should be appreciated that the answers obtained are often relatively inaccurate. This can easily lead to the
over-design or under-design of flood alleviation works, for example.

Figure 6 A compound channel with three subsections, each having a constant n

19
The compound channel in 3 has been split into three subsections, each subsection having the same
Manning’s n value. Subsection 1 is the left floodplain, and this has the same n as the main channel
(subsection 2). The right floodplain is subsection 3. The assumption is that each subsection can be analyzed
separately, then:

Total discharge = ∑ subsection discharges

i.e. Q = Q1 + Q2 + Q3
where the subscripts refer to the subsections. When conducting the calculations, only the real perimeters
are assigned a roughness and included in the calculation of P and R, the imaginary vertical dividing lines
between subsections (shown dashed) are not. It is assumed that the main channel and floodplains all have
the same longitudinal bed slope, SO = 1 in 600 in this example. Proceeding on this basis the calculations
for each of the subsections in the diagram are as follows
subsection1:
𝐴1 = 1.5 × 15.0 = 22.5𝑚2 , 𝑃1 = 1.5 + 15 = 16.5𝑚, 𝑅1 = 𝐴1 /𝑃1 = 22.5/16.5 = 1.364𝑚
𝐴1 2/3 1/2 22.5 1 1/2
𝑄1 = 𝑅1 𝑆0 = × 1.3642/3 × ( ) = 32.3𝑚3 /𝑠
𝑛1 0.035 600
Subsection2:
𝐴2 = 4 × 10.0 = 40.0𝑚2 , 𝑃2 = 2.5 + 10.0 + 1.7 = 14.2𝑚, 𝑅2 = 𝐴2 /𝑃2 = 40.0/14.2 = 2.817𝑚
𝐴2 2/3 1/2 40.0 1 1/2
𝑄2 = 𝑅2 𝑆0 = × 2.8172/3 × ( ) = 93.1𝑚3 /𝑠
𝑛2 0.035 600
Subsection3:
𝐴3 = 2.3 × 20.0 = 46.0𝑚2 , 𝑃3 = 2.3 + 20 = 22.3𝑚, 𝑅3 = 𝐴3 /𝑃3 = 46.0/22.3 = 2.063𝑚
𝐴3 2/3 1/2 46.0 1 1/2
𝑄3 = 𝑅3 𝑆0 = × 2.0632/3 × ( ) = 53.4𝑚3 /𝑠
𝑛3 0.057 600
Total discharge through the main channel and flood plains, 𝑄 = 32.3 + 93.1 + 53.4 = 176.8𝑚3 /𝑠

20
Figure 7 A compound channel where n (shown adjacent to the boundary) is different
Figure 4 illustrates a situation involving a compound channel where the calculations are complicated by
the fact that every part of the perimeter has a different n value. This may be encountered in both natural
and laboratory channels; in the latter case the channel bottom is often roughened artificially to produce
Froude numbers similar to those experienced in rivers. In such cases we need to calculate the average
(composite) roughness of the entire cross-section and then use this in the Manning equation to obtain the
total discharge. This average roughness (nAV) cannot be a straight-forward mathematical average of the n
values since the various parts of the subsection have different lengths and the effect of the roughness is
non-linear. However, the following equations can be used to determine the average roughness of a cross-
section that has a total overall wetted perimeter and hydraulic radius of P and R, but which comprises N
different lengths, any one of which is denoted by P i, Ri and ni;
𝑃𝑅5/3
𝑛𝐴𝑉 = 5 … … … 1.19
𝑃𝑖 𝑅𝑖 3
∑𝑁
𝑖=1 ( 𝑛 )
𝑖

2/3
∑𝑁
𝑖=1(𝑃𝑖 𝑛𝑖
3/2
)
𝑛𝐴𝑉 =[ ]
𝑃
1/2
∑𝑁 2
𝑖=1(𝑃𝑖 𝑛𝑖 )
𝑛𝐴𝑉 = [ ]
𝑃
Example 4. A compound channel has the cross-section shown in Fig. 4. The values next to the perimeter
boundaries are the Manning roughness values (ns/m1/3). Use the above equation to calculate nAV. The
calculations are conducted in the table below and the total values substituted into the equation.

21
1/2
∑𝑁 2
𝑖=1(𝑃𝑖 𝑛𝑖 ) 0.2607 1/2
𝑛𝐴𝑉 = [ ] =[ ] = 0.065
𝑃 61.246
Example5: During large floods, the water level in the channel given in Example 5.1 exceeds the bank-full
level of 2.5 m. The flood banks are 10 m wide and are grassed with side slopes of 1:3. The estimated
Manning’s n for these flood banks is 0.035. Estimate the discharge for a maximum flood level of 4 m and
the value of the velocity coefficient α.

Figure 8 compound channel

In this case, it is necessary to split the section into subsections (1), (2), (3) as shown in Figure 5. Manning’s
formula may be applied to each one in turn, and the discharges can be summed. The division of the section
into subsections is a little arbitrary. If the shear stress across the arbitrary divisions is small compared with
the bed shear stresses, it may be ignored to obtain an approximate solution. For section (1)
5 + 15
𝐴1 = ( ) 2.5 + (15 × 1.5) = 47.5𝑚2
2
& 𝑃1 = 5 + 2√5 × 2.5 = 16.18𝑚
47.55/3
hence 𝐾1 = = 6492.5
(16.182/3 ×0.015)

section (2) and (3) have the same dimensions, hence


10+14.5
𝐴2 = 𝐴3 = (
2
) × 1.5 = 18.38𝑚2

𝑃2 = 𝑃3 = 10 + (√10 × 1.5) = 14.74𝑚

22
18.385/3
and 𝐾2 = 𝐾3 = = 608.4
14.74 2/3 ×0.035
1 47.55/3
hence 𝑄1 = × × 0.0011/2
0.015 16.182/3

or 𝑄1 = 𝐾1 × 0.0011/2 = 205.3𝑚3 /𝑠
1 18.385/3
𝑄2 = 𝑄3 = × × 0.0011/2
0.035 14.74 2/3

or 𝑄2 = 𝑄3 = 𝐾2 × 0.0011/2 = 19.2𝑚3 /𝑠
hence 𝑄 = 𝑄1 + 𝑄2 + 𝑄3 = (𝐾1 + 𝐾2 + 𝐾3 ) × 0.0011/2 = 243.7𝑚3 /𝑠
the velocity coefficient may be found directly from equation 1.1
𝑉1 3 𝐴1 + 𝑉2 3 𝐴2 + 𝑉3 3 𝐴3
𝛼=
𝑉̅ 3 𝐴
𝑄1 205.3
𝑉1 == = 4.32𝑚/𝑠
𝐴1 47.5
𝑄2 19.2
𝑉2 = 𝑉3 = = = 1.04𝑚/𝑠
𝐴2 18.38
𝑄 243.7
𝑉̅ = = = 2.89𝑚/𝑠
𝐴 47.5 + 18.38 + 18.38
4.323 ×47.5+[2(1.043 ×18.38)]
Hence 𝛼= = 1.9
2.893 ×84.26
1.8 Rapidly Varied Flow: The Use of Energy Principles
1.8.1 Energy Equation in Open Channels
Bernoulli’s equation may be applied to any streamline. If the streamlines are parallel, then the pressure
distribution is hydrostatic. Referring to Figure 6, which shows uniform flow in a steep channel, consider
point A on a streamline. The pressure force at point A balances the component of weight normal to the bed,
i.e.,
𝑝𝐴 ∆𝑆 = 𝜌𝑔𝑑∆𝑆𝑐𝑜𝑠𝜃
𝑝𝐴 = 𝜌𝑔𝑑𝑐𝑜𝑠𝜃 = 𝜌𝑔𝑦2

23
Figure 9 Application of Bernoulli’s equation to uniform flow in open channels.

It is more convenient to express this pressure force in terms of y1 (the vertical distance from the streamline
to the free surface). Since
𝑦2
𝑑=− 𝑎𝑛𝑑 𝑑 = 𝑦1 cos 𝜃
cos 𝜃
then 𝑦2 = 𝑦1 cos 2 𝜃
and hence 𝑝𝐴 = 𝜌𝑔𝑦1 cos 2 𝜃
or 𝑝𝐴/𝜌𝑔 = 𝑦1 cos 2 𝜃
Most channels have very small bed slopes (e.g., less than 1:100, corresponding to θ less than 0.57° or cos2
θ greater than 0.9999) and therefore for such channels
cos 2 𝜃 ≈ 1
and 𝑃𝐴 /𝜌𝑔 ≈ 1
Hence, Bernoulli’s equation becomes
𝛼𝑉 2
𝐻=𝑦+ + 𝑧 … … .1.20
2𝑔
1.8.2 Application of the Energy Equation
Consider the problem shown in Figure 9. Steady uniform flow is interrupted by the presence of a hump in
the streambed. The upstream depth and the discharge are known, and it simply remains to find the depth
of flow at position (2). Applying the energy equation (1.20) and assuming frictional energy losses between
(1) and (2) are negligible, then
𝑉1 2 𝑉2 2
𝑦1 + = 𝑦2 + + Δ𝑧, (𝑡𝑎𝑘𝑖𝑛𝑔 𝛼 = 1)
2𝑔 2𝑔

24
There are two principal unknown quantities − V2 and y2 − and hence to solve this problem a second equation
is required – the continuity equation:
𝑉1 𝑦1 = 𝑉2 𝑦2 = 𝑞
Where 𝑞is the discharge per unit width. Combining these equations,
𝑞2 𝑞2
𝑦1 + = 𝑦1 + + 𝛥𝑧
2𝑔𝑦1 2 2𝑔𝑦2 2
and rearranging,
𝑞2
2𝑔𝑦2 3 + 𝑦2 2 (2𝑔Δ𝑧 − 2𝑔𝑦1 − ) + 𝑞 2 = 0 … … … … … … … … … … … … .1.21
𝑦1 2
This is a cubic equation, in which y2 is the only unknown, and to which there are three possible
mathematical solutions. As far as the fluid flow is concerned, however, only one solution is possible. To
determine which of the solutions for y2 is correct requires a more detailed knowledge of the flow.
Specific Energy To solve the problem mentioned earlier, the concept of specific energy was introduced by
Bakhmeteff in 1912. Specific energy (Es) is defined as the energy of the flow referred to the channel bed
as datum:
𝛼𝑉 2
𝐸𝑠 = 𝑦 + … … … … … … … … … … … … … … … .1.22
2𝑔
Application of the specific energy equation provides the solution to many rapidly varied flow problems.
For steady flow, the above equation may be rewritten as
𝛼(𝑄/𝐴)2
𝐸𝑠 = 𝑦 +
2𝑔
Now consider a rectangular channel:

25
𝑄 𝑏𝑞 𝑞
= =
𝐴 𝑏𝑦 𝑦
where b is the width of the channel. Hence,
𝛼𝑞 2
𝐸𝑠 = 𝑦 +
2𝑔𝑦 2
As q is constant,
𝛼𝑞 2
(𝐸𝑠 − 𝑦)𝑦 2 = = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡
2𝑔
or
(𝐸𝑠 − 𝑦) = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡/𝑦 2
Again, this is a cubic equation for the depth y for a given ES. Considering only positive solutions, then the
equation is a curve with two asymptotes:
as𝑦 → 0, 𝐸𝑠 → ∞
𝑦 → ∞, 𝐸𝑠 → 𝑦
This curve is now used to solve the problem given in Figure.9. In Figure 10, the problem has been redrawn
alongside a graph of depth versus specific energy. Equation 5.19 may be written as
𝐸𝑠1 = 𝐸𝑠2 + ∆𝑧 … … … … … … … … … … . .1.22

Figure 10 Use of specific energy to solve flow transition problems

intermediate point ES1 − ES2 > Δz, which is physically impossible. Hence, the flow depth at (2) must
correspond to point B on the specific energy curve. This is a very significant result, so an example of such
a flow transition follows.
Example 7

26
The discharge in a rectangular channel of width 5 m and maximum depth 2 m is 10 m3/s. The normal depth
of flow is 1.25m. Determine the depth of flow downstream of a section in which the bed rises by 0.2 m
over a distance of 1 m.
Solution The solution is shown graphically in Figure 7. Assuming frictional losses are negligible, then
(1.22) applies, i.e.,
𝐸𝑠1 = 𝐸𝑠2 + ∆𝑧
In this case,
𝐸𝑠1 = 𝑦1 + 𝑉2 2 /2𝑔 = 1.25 + [10/(5 × 1.25)]2 /2𝑔 = 1.38𝑚
𝐸𝑠2 = 𝑦2 + [10/(5𝑦2 )]2 /2𝑔 = 𝑦2 + 22 /2𝑔𝑦2 2
∆𝑧 = 0.2
Hence, 1.38 = 𝑦2 + (22 /2𝑔𝑦2 2 ) + 0.2
Or
1.18 = 𝑦2 + 2/𝑔𝑦2 2
This is a cubic equation for y2, but the correct solution is that given by point B in Figure 7, which in this
case is about 0.9 m. This is used as the initial estimate in a trial-and-error solution, as follows:

This result is often a source of puzzlement, and a simple physical explanation is called for. The answer lies
in realizing that the fluid is accelerating. Consider, initially, that the water surface remains at the same level
over the bed rise. As the water depth is less (over the bed rise) and the discharge constant, then the velocity
must have increased; i.e., the fluid must have accelerated. However, to accelerate the fluid, a force is
required. This is provided by a fall in the water surface elevation, which implies that the hydrostatic force
acting in the downstream direction is greater than that acting in the upstream direction.
1.9 Subcritical, Critical and Supercritical Flow
In Figure 8, the specific energy curve (for constant discharge) has been redrawn alongside a second curve
of depth against discharge for constant Es. The figure is now used to illustrate several important principles
of rapidly varied flow.
i. For a given constant discharge:

27
a) The specific energy curve has a minimum value ESc at point c with a corresponding depth yc–known as
the critical depth.
b. For any other value of Es, there are two possible depths of flow known at alternate depths, one of which
is termed subcritical and the other supercritical:
For supercritical flow, y < yc.
For subcritical flow, y > yc
2. For a given constant specific energy:
a. The depth–discharge curve shows that discharge is a maximum at the critical depth.
b. For all other discharges, there are two possible depths of flow (sub- and supercritical) for any particular
value of Es.

Figure 11 Variation of specific energy and discharge with depth

1.9.1 General Equation of Critical Flow


Referring to Figure 8, the general equation for critical flow may be derived by determining ES c and Qmax,
independently, from the specific energy equation. It will be seen that the two methods both result in the
same solution.

28
𝜶𝑸𝟐 𝒎𝒂𝒙 𝑩𝒄
Comparing, = 𝟏 … … … … … … … … … … 1.23
𝒈𝑨𝒄 𝟑

in other words, at the critical depth the discharge is a maximum and the specific energy is a minimum.
1.9.3 Critical Depth and Critical Velocity (for a Rectangular Channel)
Determination of the critical depth (yc) in a channel is necessary for both rapidly and gradually varied flow
problems. The associated critical velocity (Vc) will be used in the explanation of the significance of the
Froude number. Both of these parameters may be derived directly from equation (1.23) for any shape of
channel. For illustrative purposes, the simplest artificial channel shape is used here, i.e., rectangular. For
critical flow,
For a rectangular channel, Q=qb, B=b, and A=by. Substituting in (1.23 ) and taking 𝛼 = 1
1/3
𝑞2
𝑦𝑐 = ( ) … … … … … … … … … 𝟏. 𝟐𝟒
𝑔
and as Vcyc=q
then 𝑉𝑐 = √𝑔𝑦𝑐 … … … … … … … . 𝟏. 𝟐𝟓
𝑦𝑐 2
Also, as 𝐸𝑠𝑐 = 𝑦𝑐 + or 𝑦𝑐 = 𝐸𝑠𝑐 … … … … . . 𝟏. 𝟐𝟔
2 3

1.9.4 Froude Number


The Froude number (pronounced as Frood) is defined as
𝑽
𝑭𝒓 =
√𝒈𝑳
29
where L is a characteristic dimension. It is attributed to William Froude (1810–1897), who used such a
relationship in model studies for ships. If L is replaced by D m, the hydraulic mean depth, then the resulting
𝑽
dimensionless parameter 𝑭𝒓 =
√𝒈𝑫𝒎

is applicable to open channel flow. This is extremely useful, as it defines the regime of flow, and as many
of the energy and momentum equations may be written in terms of the Froude number. The physical
significance of Fr may be understood in two different ways. First, from dimensional analysis
𝒊𝒏𝒆𝒓𝒕𝒊𝒂𝒍 𝒇𝒐𝒓𝒄𝒆 𝝆𝑳𝟐 𝑳𝑽𝟐 𝑽𝟐
𝑭𝒓 𝟐 = = =
𝒈𝒓𝒂𝒗𝒊𝒕𝒂𝒕𝒊𝒐𝒏𝒂𝒍 𝒇𝒐𝒓𝒄𝒆 𝝆𝒈𝑳𝟑 𝒈𝑳
Secondly, by considering the speed of propagation c of a wave of low amplitude and long wavelength, a
fresh insight is gained. Such waves may be generated in a channel as oscillatory waves or surge waves.
Oscillatory waves (e.g., ocean waves), and surge waves. Both types of wave lead to the result that 𝒄 =

√𝒈𝒚. For a rectangular channel Dm = y, and hence


𝑉 𝑤𝑎𝑡𝑒𝑟 𝑣𝑒𝑙𝑜𝑐𝑖𝑡𝑦
𝐹𝑟 = = … … … … 𝟏. 𝟐𝟕
√𝑔𝑦 𝑤𝑎𝑣𝑒 𝑣𝑒𝑙𝑜𝑐𝑖𝑡𝑦

Also, for a rectangular channel 𝑉𝑐 = √𝑔𝑦𝑐 , and hence, for critical flow,

𝑉 √𝑔𝑦𝑐
𝐹𝑟 = = =1
√𝑔𝑦 √𝑔𝑦𝑐
This is a general result for all channels (as it can be shown that for non-rectangular channels
(𝑉𝑐 = √𝑔𝐷𝑚 and 𝑐 = √𝑔𝐷𝑚 )
𝑉 < 𝑉𝑐 and Fr<1
And for supercritical flow
V>Vc and Fr>1
The Froude number therefore defines the regime of flow. There is a second consequence of major
significance. Flow disturbances are propagated at a velocity of 𝒄 = √𝒈𝒚. Hence, if the flow is supercritical,
any flow disturbance can only travel downstream as the water velocity exceeds the wave velocity. By
contrast, for subcritical flow, flow disturbances can travel both upstream and
downstream. Such flow disturbances are introduced by all channel controls and local features. For example,
if a pebble is dropped into a still lake, then small waves are generated in all directions at equal speeds,
resulting in concentric wave fronts. Now imagine dropping the pebble into a river. The wave fronts will
move upstream slower than they do downstream, due to the current. If the water velocity exceeds the wave
velocity, the wave will not move upstream at all.

30
Figure 12 Flow disturbance. (a) Caused by the weir is transmitted upstream as Fr < 1 and yn > yc. (b) Caused by a change of slope is no t

transmitted upstream as Fr > 1 and yn < yc.

An example of how channel controls transmit disturbances upstream and downstream is shown in Figure
10. To summarize:

When the Froude number is close to one in a channel reach, flow conditions tend to become unstable,
resulting in wave formations. If the channel is a compound channel, for example, flood flows in a main
channel and its flood plains, then some very interesting and little investigated phenomena can occur.
Example 11:
Water flows down a rectangular channel that is 4.32m wide at a depth of 2.34m with a corresponding
velocity of 0.97m/s. Calculate the critical depth in the channel and determine whether the flow is subcritical
or supercritical.
Correct Solution:
𝑦𝐶 = (𝑄 2 /𝑔𝐵 2 )1/3
𝑄 = 𝐴𝑉 = 4.32 × 2.34 × 0.97 = 9.806𝑚3 /𝑠
𝑦𝐶 = (9.8062 /9.81 × 4.322 )1/3 = 0.807𝑚(𝐶𝑂𝑅𝑅𝐸𝐶𝑇)

31
Incorrect solution
𝐸 = 𝑉 2 /2𝑔 + 𝑦
𝐸 = (0.972 /2 × 9.81) + 2.34
𝐸 = 2.388𝑚
2 2
𝑦𝐶 = 𝐸 = × 2.388 = 1.592
3 𝐶 3
The flow in the channel is subcritical because the actual depth of flow (2.34m) is greater than the critical
depth (0.81m). No other calculation is needed to prove the flow is subcritical, although this can be
confirmed by calculating the Froude number.
𝐹 = 𝑉/(𝑔𝑦)1/2 = 0.97/(9.81 × 2.34)1/2 = 0.20 (< 1 𝑠𝑜 𝑠𝑢𝑏𝑐𝑟𝑖𝑡𝑖𝑐𝑎𝑙)
The incorrect solution is wrong because yc=2/3Ec, only applies to critical flow, and the conditions are not
critical in this example. Take care not to mis-apply equations!
1.10 Rapidly Varied Flow: The Use of Momentum Principle
1.10.1 Hydraulic Jump
A hydraulic jump occurs when a supercritical flow meets/changes to a subcritical flow. Situations where
the flow changes from supercritical to subcritical include the conditions downstream of a sluice gate and
where water flows over a steep dam spillway into a stilling basin. In all such cases a fast, shallow flow has
to change to a slow, deep flow. Unlike the transition described above, this cannot happen smoothly and
gradually. Instead there is a sudden increase in depth in the form of a hydraulic jump or standing wave.
This zone of rapidly varying flow consists of highly turbulent water that froths and boils. The water on the
wave front tends to have a motion that is directed downwards and back upstream, while the underlying
flow is expanding upwards in a downstream direction. Hence there is great turbulence, a lot of air
entrainment, and a considerable loss of energy. Downstream of the jump the water surface quickly becomes
calm, and the flow continues smoothly at the subcritical normal depth. It should be noted that the type of
jump occurs when the upstream flow has a Froude number of 2 or more. At lower F values the jump tends
to be either undular in nature or weakly developed so that it is not always easy to determine by visual
observation that there is a flow transition taking place. This is particularly true in natural rivers where the
flow can be turbulent and the water surface undulates anyway.
The resulting flow transition is rapid and involves a large energy loss due to turbulence. Under these
circumstances, a solution to the hydraulic jump problem cannot be found using a specific energy diagram.
Instead, the momentum equation is used. Figures13 a and b depict a hydraulic jump and the associated
specific energy and force–momentum diagrams. Initially, ΔEs is unknown, as only the discharge and
upstream depth are given. By using the momentum principle, the sequent depth y 2 may be found in terms

32
of the initial depth y1 and the upstream Froude number (Fr1). It may be noted that these two depths are also
often referred to as conjugate depths.

Figure 13Hydraulic jump. (a) with associated specific energy diagram (b). with associated force-momentum diagram

The momentum equation derived may be written as F M = ∆m where M is the momentum per unit time.
In this case, the forces acting are the hydrostatic pressure forces upstream (F1) and downstream (F2) of the
jump. Their directions of action are as shown in Figure 13 b. This is a point which often causes difficulty.
It may be understood most easily by considering the jump to be inside a control volume, and to consider
the external forces acting on the control volume
Net force in the x-direction=F1+F2
Momentum change (per unit time) =M1-M2
Hence, F1-F2=M2-M1
or F1+M1=F2+M2=constant (for constant discharge)
If depth (y)is plotted against force + momentum change(F+M) for a constant discharge (known as the
specific force diagram), then for a stable jump
𝐹 + 𝑀 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡 … … … … … … … … … 𝟏. 𝟐𝟖
Therefore, for any given initial depth, the sequent depth is the corresponding depthon the force-momentum
diagram.
1.10.2 Solution of the Momentum Equation for a Rectangular Channel
For a rectangular channel, (1.28) may be evaluated as follows:
𝐹1 = 𝜌𝑔(𝑦1 /2)𝑦1 𝑏 𝐹2 = 𝜌𝑔(𝑦2 /2)𝑦2 𝑏

33
𝑀1 = 𝜌𝑄𝑉1 𝑀2 = 𝜌𝑄𝑉2
𝑄1 𝑄2
= 𝜌𝑄 = 𝜌𝑄
𝑦1 𝑏 𝑦2 𝑏
Substituting and rearranging,
𝜌𝑔𝑏 2 𝜌𝑄 2 1 1
(𝑦1 − 𝑦22 ) = ( − )
2 𝑏 𝑦2 𝑦1
Substituting q=Q/b and simplifying,
1 2 𝑞2 1 1
(𝑦1 − 𝑦22 ) = ( − )
2 𝑔 𝑦2 𝑦1
1 𝑞 2 𝑦1 − 𝑦2
(𝑦1 − 𝑦2 )(𝑦1 + 𝑦2 ) = ( )
2 𝑔 𝑦2 𝑦1
1 𝑞2 1
(𝑦1 + 𝑦2 ) = ( )
2 𝑔 𝑦2 𝑦1
Substituting q=V1y1 and dividing by 𝑦12 ,
1 𝑦2 𝑦2 𝑉1 2
(1 + ) = = 𝐹𝑟12
2 𝑦1 𝑦1 𝑔𝑦1
This is a quadratic equation in y2/y1, whose solution is
𝑦1
𝑦2 = ( ) (√1 + 8𝐹𝑟12 − 1) … … . . 𝟏. 𝟐𝟗(𝒂)
2
Alternatively, it can be shown that
𝑦2
𝑦1 = ( ) (√1 + 8𝐹𝑟22 − 1) … … … … … 𝟏. 𝟐𝟗(𝒃)
2
1.10.3 Energy Dissipation in a Hydraulic Jump
Using (1.29a), y2 may be evaluated in terms of y1 and hence the energy loss through the jump determined:
Equation 1.29a may be written as
1 𝑦2 𝑦2
𝐹𝑟12 = ( + 1) … . . 𝟏. 𝟑𝟎
2 𝑦1 𝑦1
The upstream Froude number is related to q as follows:
𝑉1 𝑞/𝑦1
𝐹𝑟12 = =
√𝑔𝑦1 √𝑔𝑦1
𝑞2
𝐹𝑟12 =
𝑔𝑦13
More manipulation of equation 1.21 yields

34
(𝑦2 − 𝑦1 )3
∆𝐸 = … … . . 𝟏. 𝟑𝟏
4𝑦2 𝑦1
It should be noted that the energy loss increases very sharply with the relative height of the jump.
Example 10
Water flows down a steep concrete lined rectangular channel 5.0m wide at a depth of 0.65m when the
discharge is 19.0m3/s. At the bottom of the slope the channel becomes horizontal, but is otherwise
unchanged. Determine whether a hydraulic jump will form, the energy loss, and the approximate
dimensions of the jump.
Soln:
Considering the flow in the steep upstream channel, y1 = 0.65m. The upstream velocity, V1 = Q/A1 =
1
19.0/(5.0 × 0.65) = 5.846m/s. Upstream Froude number, 𝐹1 = 𝑉1/(𝑔𝑦1)2 = 5.846/(9.81 ×
0.65)1/2 = 2.32 Therefore the flow is supercritical in the upstream channel. It is probable that a weak
jump would form in the horizontal part of the channel. Assuming that the initial depth before the jump
is 𝑦1 = 0.65𝑚 then equation from
𝑦2
𝑦1 = ( ) (√1 + 8𝐹𝑟22 − 1)
2
0.65
𝑦2 = ( ) (√1 + 8 × 2.322 − 1) = 1.83𝑚
2
Therefore, the downstream depth would be 𝑦2 = 1.83𝑚, giving
𝑉2 = 𝑄/𝐵𝑦2 = 19.0/(5.0 × 1.83) = 2.077𝑚/𝑠
Thus, the height of the jump is 1.83-0.65=1.18m
1 1
𝐹𝑟2 = 𝑉2 /(𝑔𝑦2 )2 = 2.077/(9.81 × 1.83)2 = 0.49
Therefore, the downstream depth would be y2 = 1.83m, giving
The energy loss is defined by equation 1.31
(𝑦2 − 𝑦1 )3 (1.83 − 0.65)3
∆𝐸 = = × 0.65 × 1.83 = 0.35𝑚
4𝑦2 𝑦1 4

1.11 Gradually Varied Flow


1.11.1 Significance of Bed Slope and Channel Friction
It was assumed that frictional effects may be ignored in a region of rapidly varied flow. This is a reasonable
assumption, since the changes take place over a very short distance. However, bed slope and channel
friction are very important because they determine the flow regime under gradually varied flow conditions.
The discussion of the specific energy curve and the criteria for maximum discharge indicated that, for a
given specific energy or discharge, there are two possible flow depths at any point in a channel. The
solution of Manning’s equation results in only one possible flow depth (the normal depth). This apparent

35
paradox is resolved by noting the influence of the bed slope and channel friction. These parameters
determine which of the two possible flow depths will occur at any given point for uniform flow. In other
words, for uniform flow, the bed slope and channel friction determine whether the flow regime is sub- or
supercritical. The normal depth of flow may be less than, equal to, or greater than, the critical depth. For a
given channel shape and roughness, only one value of slope will produce the critical depth, and this is
known as the critical slope (Sc). If the slope is steeper than Sc, then the flow will be supercritical and the
slope is termed a steep slope. Conversely, if the slope is less steep than Sc, then the flow will be subcritical
and the slope is termed a mild slope.
Critical Bed Slope in a Wide Rectangular Channel
To illustrate this concept, the equation for the critical bed slope is now derived for the case of a wide
rectangular channel. For uniform flow,
1 𝐴5/3 1/2
𝑄= 𝑆
𝑛 𝑝5/3 0
And for critical flow
𝑄2𝐵
= 1(taking α = 1
𝑔𝐴3
Hence, combining the above equations to eliminate Q,

1 𝐴5/3 1/2 𝑔𝐴3


5/3
𝑆𝑐 = √
𝑛𝑝 𝐵

Where sc is the critical bed slope.


For a wide rectangular channel of width b, 𝐵 = 𝑏 𝑎𝑛𝑑 𝑃 ≈ 𝑏.
Making these substitutions for a wide rectangular channel,
1
1 (𝑏𝑦𝑐 ) 1/2 𝑔2 (𝑏𝑦𝑐 )3/2
𝑆 =
𝑛 𝑏 2/3 𝑐 𝑏1/2
Where 𝑦𝑐 is the critical depth. Hence,
𝑔𝑛2
𝑆𝑐 = 1/3
𝑦𝑐
Example: Determination of Critical Bed Slope
Given a wide rectangular channel of width 20 m, determine the critical bed slope and discharge for
critical depths of 0.2, 0.5, and 1.0 m. Assume that n = 0035.
Solution.
Using

36
𝑔𝑛2
𝑆𝑐 = 1/3
1.32
𝑦𝑐
and
1 𝐴5/3 1/2
𝑆
𝑛 𝑝5/3 0
For a particular yc,substitute into (1.32) to fin SC, and then substitute into mannings equation to find Q,
hence

The results of the following example demonstrate that the critical bed slope is dependent on discharge. In
other words, for a given channel with a given slope, it is the discharge which determines whether that
slope is mild or steep.
Example: Critical Depth and Slope in a Natural Channel
The data given next were derived from the measured cross section of a natural stream channel. Using the
data, determine the critical stage and associated critical bed slope for a discharge of 60 m 3/s assuming n =
0.04.

Solution: the general equation for critical flow – must be used, i.e.,
𝛼𝑄 2 𝐵𝑐
=1
𝑔𝐴3𝐶
Both A and B are functions of stage(h), as given in the previous table. In this case, the best method of
solution is a graphical one. Rearranging
𝛼𝑄 2 𝐴3𝐶
=
𝑔 𝐵𝑐

37
Hence if h is plotted against 𝐴3 /𝐵, the for h=hc, 𝐴3𝐶 /𝐵𝑐 = 𝛼𝑄 2 /𝑔 on the A3/B axis. From the given data,
values of A3/B for various h values may be calculated, i.e.,
And for critical flow 𝛼𝑄 2 /𝑔 = 602 /𝑔 = 367(𝑎𝑠𝑠𝑢𝑚𝑖𝑛𝑔 𝑡ℎ𝑎𝑡 𝛼 = 1).

By inspection, the critical stage must be between 1.5 and 2.0. Using linear interpolation
(ℎ𝑐 − 1.5) (367 − 273.1)
=
(2 − 1.5) (813.2 − 273.1)
ℎ𝑐 = 1.59𝑚
To find the critical bed slope, apply manningsa equation with h=hc=1.59, Q=60
1 𝐴5/3 1/2
𝑄= 𝑆
𝑛 𝑝𝑐 2/3 𝑐
or in this case
2/3 2
𝑄𝑛𝑃𝐶
𝑆𝐶 = ( 5/3 )
𝐴𝐶
For ℎ𝐶 = 1.59, again using linear interpolation,
0.09
𝑃𝐶 = 16.7 + (19.5 − 16.7) = 17.22
0.5

0.09
𝐴𝐶 = 16 + (24 − 16) = 17.44
0.5
2
60×0.04×17.22/3
Hence 𝑆𝐶 = ( ) = 0.0186
17.44 5/3

1.11.2 Flow Transitions


Figure 14 shows two types of transition, due to changes of bed slope. In Figure 14a it is presupposed that
the channel is of mild slope upstream and steep slope downstream. The critical depth (for a given discharge)
is constant. Upstream, the flow is subcritical and the depth is greater than the critical depth. Downstream,
the converse is true. In the vicinity of the intersection of the mild and steep slopes, gradually varied flow
is taking place and the flow regime is in transition from sub to supercritical. At the intersection the flow is
critical. In Figure 5.18b, the slopes have been reversed and the resulting flow transition is both more
spectacular and more complex. Upstream, the flow is supercritical, and downstream the flow is subcritical.

38
This type of transition is only possible through the mechanism of the hydraulic jump. Gradually varied
flow takes place between the intersection of the slopes and the upstream end of the jump. An explanation
as to why these two types of transition exist may be found in terms of the Froude number. Consider what
would happen if a flow disturbance was introduced in the transition region shown in Figure 14a. On the
upstream (mild) slope Fr < 1, and the disturbance would propagate both upstream and downstream. On the
downstream (steep) slope the disturbance would propagate downstream. The net result is that all flow
disturbances are swept away from the transition region, resulting in the smooth flow transition shown.

Figure 14 Flow transitions. (a) subcritical to supercritical. (b) supercritical to subcritical

Conversely, for the transition shown in Figure 5.18b, flow disturbances introduced upstream (on the steep
slope) propagate downstream only. Those introduced downstream propagate both upstream and
downstream. The net result in this case is that the disturbances are concentrated into a small region, which
is the hydraulic jump.
General Equation of Gradually Varied Flow
To determine the flow profile through a region of gradually varied flow, due to changes of slope or cross
section, the general equation of gradually varied flow must first be derived. The equation is derived by
assuming that for gradually varied flow the change in energy with distance is equal to the frictional losses
(e.g., ignoring any flow accelerations). Hence,
𝑑𝐻 𝑑 𝛼𝑉 2
= (𝑦 + + 𝑧) = −𝑆𝑓
𝑑𝑥 𝑑𝑥 2𝑔
Where sf is frictional slope. Rewriting.
𝑑 𝛼𝑉 2 𝑑𝑧
(𝑦 + + 𝑧) = − − 𝑆𝑓
𝑑𝑥 2𝑔 𝑑𝑥
or

39
𝑑𝐸𝑠
= 𝑆0 − 𝑆𝑓 (1.32)
𝑑𝑥
where so is the bed slope. From equation (5.7),

𝑑𝐸𝑠 𝑄2𝐵
=1− (𝑡𝑎𝑘𝑖𝑛 𝛼 = 1)
𝑑𝑦 𝑔𝐴3
and as
𝑉2 𝑄2𝐵
𝐹𝑟2 = =
𝑔𝐴/𝐵 𝑔𝐴3
Then
𝑑𝐸𝑠
= 1 − 𝐹𝑟2 … … . .1.33
𝑑𝑦
Combining this with (1.32) gives
𝑑𝑦 𝑠0 − 𝑠𝑓
= … … … . .1.34
𝑑𝑥 1 − 𝐹𝑟2
Which is the general equation of gradually varied flow.
Sf represents the slope of the total energy line 𝑑𝐻/𝑑𝑥 . Since the bed slope (S0) and the friction slope (Sf)
are coincident for uniform flow, the friction slope (S f) may be evaluated using Manning’s equation or the
Colebrook–White equation. Equations 1.33 and 1.34 are differential equations relating depth to distance.
There is no general explicit solution (although particular solutions are available for prismatic channels).
Numerical methods of solution are normally used in practice. These methods are considered in a later
section.

40
CHAPTER TWO: FLOW IN PIPES

2.1 Introduction
The flow of water, oil and gas in pipes is of immense practical significance in engineering. Water is
conveyed from its source, normally in pressure pipelines (Figure 4.1), to water treatment plants where
it enters the distribution system and finally arrives at the consumer. Oil and gas are often transferred
from their source by pressure pipelines to refineries (oil) or into a distribution network for supply
(gas).
This chapter describes the theories of pipe flow, beginning with a review of the historical context and
ending with the practical applications. The treatment is limited to steady flow in pressurized and non-
pressurized pipes, which is the principal means by which individual pipelines are designed
2.2 Fundamental Concepts of Pipe Flow
We will be looking here at the flow of real fluid in pipes – real meaning a fluid that possesses viscosity
hence loses energy due to friction as fluid particles interact with one another and the pipe wall. Recall
from Level 1 that the shear stress induced in a fluid flowing near a boundary is given by Newton's law
of viscosity:
𝒅𝒖
𝝉∝
𝒅𝒚
This tells us that the shear stress, 𝝉, in a fluid is proportional to the velocity gradient - the rate of
change of velocity across the fluid path. For a “Newtonian” fluid we can write:
𝒅𝒖
𝝉=𝝁
𝒅𝒚

41
where the constant of proportionality, μ, is known as the coefficient of viscosity (or simply viscosity).
Recall also that flow can be classified into one of two types, laminar or turbulent flow (with a small
transitional region between these two). The non-dimensional number, the Reynolds number, Re, is
used to determine which type of flow occurs:
𝝆𝒖𝒅
𝑹𝒆 =
𝝁
For a pipe:
Laminar flow: Re < 2000, Transitional flow: 2000 < Re < 4000 and Turbulent flow: Re > 4000
It is important to determine the flow type as this governs how the amount of energy lost to friction
relates to the velocity of the flow. And hence how much energy must be used to move the fluid. Flow
in pipes is usually turbulent some common exceptions are oils of high viscosity and blood flow.
Random fluctuating movements of the fluid particles are superimposed on the main flow – these
movements are unpredictable – no complete theory is available to analyze turbulent flow as it is
essentially a stochastic process (unlike laminar flow where good theory exists.) Most of what is known
about turbulent flow has been obtained from experiments with pipes. It is convenient to study it in this
form and also the pipe flow problem has significant commercial importance. We shall cover sufficient
to be able to predict the energy degradation (loss) is a pipe line. Any more than this and a detailed
knowledge and investigation of boundary layers is required. Note that pipes which are not completely
full and under pressure e.g. sewers are not treated by the theory presented here. They are essentially
the same as open channels which will be covered elsewhere in this module.
2.3 Analysis of pipelines.
To analyze the flow in a pipe line we will use Bernoulli’s equation. The Bernoulli equation was
introduced in the Level 1 module, and as a reminder it is presented again here:
𝒑𝟏 𝒖 𝟏 𝟐 𝒑𝟐 𝒖 𝟐 𝟐
+ + 𝒛𝟏 = + + 𝒛𝟐 = 𝑯 = 𝑪𝒐𝒏𝒔𝒕𝒂𝒏𝒕
𝝆𝒈 𝟐𝒈 𝝆𝒈 𝟐𝒈
Which is written linking conditions at point 1 to conditions at point 1 in a flow. H is the total head
which does not change. When applied to a pipeline we must also take into account any losses (or gains)
in energy along the flow length. Consider a pipeline as shown below linking two reservoirs A and B
with a pump followed by a pipe that expands before reaching the downstream reservoir.

42
Examining the energy (head) losses, there will be a loss as the fluid flows into the pipe, the Entry Loss
(hL entry), then the pump put energy into the fluid in terms of increasing the pressure head (h pump). As
the pipe expands the is an expansion loss (h L expansion), then a second expansion loss, labeled Exit loss
(hL exit), as the fluid leaves the pipe into the reservoir. Along the whole length of the pipe there is a loss
due to pipe friction (hf). The Bernoulli equation linking reservoir A with B would be written thus:
𝒑𝑨 𝒖 𝑨 𝟐 𝒑𝑩 𝒖 𝑩 𝟐
+ + 𝒛𝑨 + 𝒉𝒑𝒖𝒎𝒑 = + + 𝒛𝑩 + 𝒉𝑳 𝒆𝒏𝒕𝒓𝒚 + 𝒉𝑳 𝒆𝒙𝒑𝒂𝒏𝒔𝒊𝒐𝒏 + 𝒉𝑳 𝒆𝒙𝒊𝒕 + 𝒉𝒇
𝝆𝒈 𝟐𝒈 𝝆𝒈 𝟐𝒈
Where pA, uA, zA are the pressure, velocity and height of the surface of reservoir A. The corresponding
terms are the same for B) This is the general equation we use to solve for flow in a pipeline. The
difficult part is the determination of the head loss terms in this equation. The following sections
describe how these are quantified. Before continuing it is useful to note that the above general equation
can be quickly simplified to leave on expressions for head. Remember that the points A and B are
surfaces of reservoirs – they move very slowly compared to the flow in the pipe so we can say uA =
uB = 0. Also, the pressure is atmospheric, pA = pB = pAtmospheric. zA - zB is the height difference
between the two reservoir surfaces. So
(𝒛𝑨 − 𝒛𝑩 ) + 𝒉𝒑𝒖𝒎𝒑 + 𝒉𝑳 𝒆𝒏𝒕𝒓𝒚 + 𝒉𝑳 𝒆𝒙𝒑𝒂𝒏𝒔𝒊𝒐𝒏 + 𝒉𝑳 𝒆𝒙𝒊𝒕 + 𝒉𝒇
Which is the usual form we end up solving. Commented [D1]:

2.3.1 Pressure loss due to friction in a pipeline.


Consider a cylindrical element of incompressible fluid flowing in the pipe, as shown

43
The pressure at the upstream end, 1, is p, and at the downstream end, 2, the pressure has fallen by Δp
to (p-Δp). The driving force due to pressure (F = Pressure x Area) can then be written driving force =
Pressure force at 1 - pressure force at 2.
𝝅𝒅𝟐
𝒑𝑨 − (𝒑 − 𝚫𝒑)𝑨 = 𝚫𝒑𝑨 = 𝚫𝒑
𝟒
The retarding force is that due to the shear stress by the walls
= shear stress × area over which it acts
= τ𝑤 × area of pipewall
= τ𝑤 𝜋𝑑𝐿
As the flow is in equilibrium, driving force = retarding force
𝝅𝒅𝟐
∆𝒑 = 𝝉𝒘 𝝅𝒅𝑳
𝟒
𝝉𝒘 𝟒𝑳
∆𝒑 = (𝟐. 𝟏)
𝒅
Giving an expression for pressure loss in a pipe in terms of the pipe diameter and the shear stress at
the wall on the pipe. The shear stress will vary with velocity of flow and hence with Re. Many
experiments have been done with various fluids measuring the pressure loss at various Reynolds
numbers. These results plotted to show a graph of the relationship between pressure loss and Re look
similar to the figure below:

This graph shows that the relationship between pressure loss and Re can be expressed as
laminar ∆𝑃 ∝ 𝑢
𝑡𝑢𝑟𝑏𝑢𝑙𝑒𝑛𝑡 ∆𝑃 ∝ 𝑢𝑎

44
where 1.7 < a < 2.0
As these are empirical relationships, they help in determining the pressure loss but not in finding the magnitude
of the shear stress at the wall τw on a particular fluid. If we knew τw we could then use it to give a general
equation to predict the pressure loss.
2.3.2 Pressure loss during laminar flow in a pipe
In general, the shear stress τw. is almost impossible to measure. But for laminar flow it is possible to
calculate a theoretical value for a given velocity, fluid and pipe dimension. The pressure loss in a pipe
with laminar flow is given by the Hagen-Poiseuille equation:

𝟑𝟐𝝁𝑳𝒖
∆𝒑 =
𝒅𝟐
𝟑𝟐𝝁𝑳𝒖
Or in terms of head, 𝒉𝒇 = … … … … … . (𝟐. 𝟐)
𝝆𝒈𝒅𝟐

Where hf is known as the head-loss due to friction (Remember the velocity, u, is mean velocity – and
is sometimes written 𝑢̅.)
2.3.3 Pressure loss during turbulent flow in a pipe
In this derivation we will consider a general bounded flow - fluid flowing in a channel - we will then
apply this to pipe flow. In general, it is most common in engineering to have Re > 2000 i.e. turbulent
flow – in both closed (pipes and ducts) and open (rivers and channels). However analytical expressions
are not available so empirical relationships are required (those derived from experimental
measurements). Consider the element of fluid, shown in figure 3 below, flowing in a channel, it has
length L and with wetted perimeter P. The flow is steady and uniform so that acceleration is zero and
the flow area at sections 1 and 2 is equal to A.

𝒑𝟏 𝑨 − 𝒑𝟐 𝑨 − 𝝉𝒘 𝑳𝑷 + 𝑾 𝐬𝐢𝐧 𝜽 = 𝟎
writing the weight term as 𝜌𝑔𝐴𝐿 and sin θ = −Δz/L gives

45
𝐴(𝑝1 − 𝑝2 ) − 𝜏 𝑤 𝐿𝑃 − 𝜌𝑔𝐴𝛥𝑧 = 0
This can be rearranged to give
[ (𝑝1 − 𝑝2 ) − 𝜌𝑔∆𝑧] 𝑝
− 𝜏0 = 0
𝐿 𝐴
where the first term represents the piezometric head loss of the length L or (writing piezometric head
p*)
𝒅𝒑∗
𝝉𝟎 = 𝒎 … … . . (𝟐. 𝟑)
𝒅𝒙
where m = A/P is known as the hydraulic mean depth. Writing piezometric head loss as
𝑝 ∗ = 𝜌𝑔ℎ𝑓, then shear stress per unit length is expressed as
𝒅𝒑∗ 𝝆𝒈𝒉𝒇
𝝉𝟎 = 𝒎 =𝒎
𝒅𝒙 𝑳
So, we now have a relationship of shear stress at the wall to the rate of change in piezometric pressure.
To make use of this equation an empirical factor must be introduced. This is usually in the form of a
friction factor f, and written
𝝆𝒖𝟐
𝝉𝟎 = 𝒇
𝟐
where u is the mean flow velocity.
Hence
𝒅𝒑∗ 𝝆𝒖𝟐 𝝆𝒈𝒉𝒇
=𝒇 =
𝒅𝒙 𝟐𝒎 𝑳
So, for a general bounded flow, head loss due to friction can be written
𝒇𝑳𝒖𝟐
𝒉𝒇 = … … … … … … … … (𝟐. 𝟒)
𝟐𝒎
More specifically, for a circular pipe, 𝑚 = 𝐴/𝑃 = 𝜋𝑑2 /4𝜋𝑑 = 𝑑/4 giving
𝟒𝒇𝑳𝒖𝟐
𝒉𝒇 = … … … … … . . (𝟐. 𝟓)
𝟐𝒈𝒅
This is known as the Darcy-Weisbach equation for head loss in circular pipes (Often referred to as the
Darcy equation)
This equation is equivalent to the Hagen-Poiseuille equation for laminar flow with the exception of
the empirical friction factor f introduced. It is sometimes useful to write the Darcy equation in terms
𝟒𝑸
of discharge Q, (using Q = Au), 𝒖 =
𝝅𝒅𝟐

𝟔𝟒𝒇𝑳𝑸𝟐 𝒇𝑳𝑸𝟐
𝒉𝒇 = = … … … … … … … . (𝟐. 𝟔)
𝟐𝒈𝝅𝟐 𝒅𝟓 𝟑. 𝟎𝟑𝒅𝟓

46
𝒇𝑳𝑸𝟐
Or with a 1% error, = 𝟓 … … … … … … … . (𝟐. 𝟕)
𝟑𝒅

NOTE On Friction Factor Value. The 𝑓 value shown above is different to that used in American
practice. Their relationship is 𝑓𝐴𝑚𝑒𝑟𝑖𝑐𝑎𝑛 = 4 𝑓 Sometimes the f is replaced by the Greek letter λ.
where 𝜆 = 𝑓𝐴𝑚𝑒𝑟𝑖𝑐𝑎𝑛 = 4 𝑓 .Consequently great care must be taken when choosing the value of f
with attention taken to the source of that value.

2.3.4 Choice of friction factor f


The value of f must be chosen with care or else the head loss will not be correct. Assessment of the
physics governing the value of friction in a fluid has led to the following relationships.
1. ℎ𝑓 ∝ 𝐿
2. ℎ𝑓 ∝ 𝑣 2
1
3. ℎ𝑓 ∝ 𝑑

4. ℎ𝑓 𝑑𝑒𝑝𝑒𝑛𝑑𝑠 𝑜𝑛 𝑠𝑢𝑟𝑓𝑎𝑐𝑒 𝑟𝑜𝑢𝑔ℎ𝑛𝑒𝑠𝑠 𝑜𝑓 𝑡ℎ𝑒 𝑝𝑖𝑝𝑒𝑠


5. ℎ𝑓 𝑑𝑒𝑝𝑒𝑛𝑑𝑠 𝑜𝑛 𝑓𝑢𝑖𝑑 𝑑𝑒𝑛𝑠𝑖𝑡𝑦 𝑎𝑛𝑑 𝑣𝑖𝑠𝑐𝑜𝑠𝑖𝑡𝑦
6. ℎ𝑓 𝑖𝑠 𝑖𝑛𝑑𝑒𝑝𝑒𝑛𝑑𝑒𝑛𝑡 𝑜𝑓 𝑝𝑟𝑒𝑠𝑠𝑢𝑟𝑒
Consequently, f cannot be a constant if it is to give correct head loss values from the Darcy
equation. An expression that gives f based on fluid properties and the flow conditions is required.
2.3.5 The value of f for Laminar flow
As mentioned above the equation derived for head loss in turbulent flow is equivalent to that derived
for laminar flow – the only difference being the empirical f. Equation the two equations for head loss
allows us to derive an expression of f that allows the Darcy equation to be applied to laminar flow.
Equating the Hagen-Poiseuille and Darcy-Weisbach equations gives:
32𝜇𝐿𝑢 4𝑓𝐿𝑢2
=
𝜌𝑔𝑑2 2𝑔𝑑
16𝜇
𝑓=
𝜌𝑣𝑑
𝟏𝟔
𝒇= … … … … … … … … … … … … (𝟐. 𝟖)
𝐑𝐞
Blasius equation for f
Blasius, in 1913, was the first to give an accurate empirical expression for f for turbulent flow in
smooth pipes, that is:
𝟎. 𝟎𝟕𝟗
𝒇= … … … … … … … (𝟐. 𝟗)
𝑹𝒆𝟎.𝟐𝟓

47
This expression is fairly accurate, giving head losses +/- 5% of actual values for Re up to 100000
Nikuradse
Nikuradse made a great contribution to the theory of pipe flow by differentiating between rough and
smooth pipes. A rough pipe is one where the mean height of roughness is greater than the thickness of
the laminar sub-layer. Nikuradse artificially roughened pipe by coating them with sand. He defined a
relative roughness value ks/d (mean height of roughness over pipe diameter) and produced graphs of
f against Re for a range of relative roughness 1/30 to 1/1014.

A number of distinct regions can be identified on the diagram.


The regions which can be identified are:
1. Laminar flow (f = 16/Re)
2. Transition from laminar to turbulent.
An unstable region between Re = 2000 and 4000. Pipe flow normally lies outside this region
𝟎.𝟎𝟕𝟗
3. Smooth turbulent (𝒇 = )
𝑹𝒆𝟎.𝟐𝟓

The limiting line of turbulent flow. All values of relative roughness tend toward this as Re
decreases.
4. Transitional turbulent
The region which f varies with both Re and relative roughness. Most pipes lie in this region.
5. Rough turbulent. f remains constant for a given relative roughness. It is independent of Re.
The reasons why these regions exist:
Laminar flow: Surface roughness has no influence on the shear stress in the fluid. Smooth and
Transitional Turbulence: The laminar sub-layer covers and ‘smooths’ the rough surface with a thin
laminar region. This means that the main body of the turbulent flow is unaffected by the roughness.

48
Rough turbulence: The laminar sub-layer is much less than the height of the roughness so the
boundary affects the whole of the turbulent flow.
Hydraulically rough and smooth pipes
a. In the short entry length of the pipe the flow will be laminar but this will, a short distance
downstream, give way to fully developed turbulent flow and a laminar sub-layer.
b. In the laminar sub-layer is thick enough it will protect the turbulent flow from the roughness
of the boundary and the pipe would be hydraulically smooth.
c. If the laminar sub-layer is thinner than the height of roughness, then the roughness protrudes
through and the pipe is hydraulically rough.
d. The laminar sub-layer decreases in thickness with increasing Re. Therefore, surface may be
hydraulically smooth for low flows but hydraulically rough at high flows.
e. If the height of roughness is large the flow will be completely turbulent and f will be unaffected
by Re. i.e. if k/d is large then f remains constant.
Colebrook-White equation for f
Colebrook and White did a large number of experiments on commercial pipes and they also brought
together some important theoretical work by von Karman and Prandtl. This work resulted in an
equation attributed to them as the Colebrook-White equation:
𝟏 𝒌𝒔 𝟏. 𝟐𝟔
= −𝟒 𝐥𝐨𝐠 𝟏𝟎 ( + ) … … … … . (𝟐. 𝟏𝟎)
√𝒇 𝟑. 𝟕𝟏𝒅 𝑹𝒆√𝒇

It is applicable to the whole of the turbulent region for commercial pipes and uses an effective
roughness value (ks) obtained experimentally for all commercial pipes. Note a particular difficulty
with this equation. f appears on both sides in a square root term and so cannot be calculated easily.
Trial and error methods must be used to get f once ks¸ Re and d are known. (In the 1940s when
calculations were done by slide rule this was a time-consuming task.) Nowadays it is relatively trivial
to solve the equation on a programmable calculator or spreadsheet. Moody made a useful contribution
to help, he plotted f against Re for commercial pipes – see the figure below. This figure has become
known as the Moody Diagram (or sometimes the Stanton Diagram). [Note that the version of the
Moody diagram shown uses λ (= 4f) for friction factor rather than f. The shape of the diagram will not
change if f were used instead.]

49
He also developed an equation based on the Colebrook-White equation that made it simpler to
calculate f:
𝟏/𝟑
𝟐𝟎𝟎𝒌𝒔 𝟏𝟎𝟔
𝒇 = 𝟎. 𝟎𝟎𝟏𝟑𝟕𝟓 [𝟏 + ( + ) ] … … … … … (𝟐. 𝟏𝟏)
𝒅 𝑹

This equation of Moody gives f correct to +/- 5% for 4 × 103 < Re < 1 × 107 and for ks/d < 0.01. Barr
presented an alternative explicit equation for f in 1975
𝟏 𝒌𝒔 𝟏. 𝟐𝟔
= −𝟒 𝐥𝐨𝐠 𝟏𝟎 ( + ) … … … … (𝟐. 𝟏𝟐)
√𝒇 𝟑. 𝟕𝟏𝒅 𝑹𝒆√𝒇
𝟐
𝒌𝒔 𝟏.𝟐𝟔
or 𝒇 = 𝟏/ [−𝟒 𝐥𝐨𝐠 𝟏𝟎 ( + )] … … … … … … . (𝟐. 𝟏𝟑)
𝟑.𝟕𝟏𝒅 𝑹𝒆√𝒇

Here the last term of the Colebrook-White equation has been replaced with 5.1286/Re0.89 which
provides more accurate results for Re > 10 5. The problem with these formulas still remains that these

50
contain a dependence on ks. What value of ks should be used for any particular pipe? Fortunately, pipe
manufacturers provide values and typical values can often be taken similar to those in table 1 below.
Table 1 typical ks values

Pipe material ks (mm)


Brass, copper, glass Perspex 0.003
Asbestos cement 0.03
Wrought iron 0.06
Galvanized iron 0.15
Plastic 0.03
Bitumen-lined ductile iron 0.03
Spun concrete-lined ductile iron 0.03
Slimed concrete sewer 6.0

Example: A 1.2m diameter pipeline must discharge 2.672m3 /s when flowing full. If the viscosity is
1.005 × 10 − 6𝑚2 /𝑠 and the pipeline is constructed from smooth concrete pipes for which k s=
0.60mm., calculate the friction factor, 𝜆 and the required hydraulic gradient.
Soln: 𝑉 = 2.672 × 4(𝜋 × 1. 22 ) = 2.363𝑚/𝑠
𝑉𝐷
𝑅𝑒 = = 2.363 × 1.2/1.005 × 106 = 2.821 × 106
𝜈
𝑘/𝐷 = 0.00060/1.2 = 0.000500
𝟏 𝐬𝐤 𝟐.𝟓𝟏
The friction factor can either be found by solving equation = −𝟐 𝐥𝐨𝐠 𝟏𝟎 (𝟑.𝟕𝟏𝐃 + 𝐑𝐞 𝛌)by trial and
√𝛌 √
1/3
ks 106
error or by solving equation, λ = 0.0055 [(20000 𝐷 + Re
) ] directly.
1/3
k s 106
λ = 0.0055 [(20000 + ) ]
𝐷 Re
1/3
106
λ = 0.0055 [1 + (20000 × 0.000500 + ) ]
2.821 × 106

λ = 0.0055[1 + (10.0 + 0.354)1/3 ]


λ = 0.0175
the Darcy equation gives ℎ𝑓 = λLV 2 /2𝑔𝐷, so 𝑆𝑓 = λLV 2 /2𝑔𝐷 (since Sf=hf/L)
𝑆𝑓 = 0.0175 × 2.3633 /19.62 × 12
𝑆𝑓 = 0.0042 or 1 in 241

51
Hydraulic Research Station
The basic engineering objections to the use of the Colebrook–White equation were not overcome until
the publication of Charts for the Hydraulic Design of Channels and Pipes in 1958 by the Hydraulics
Research Station. In this publication, the three dependent engineering variables (Q, D and Sf) were
presented in the form of a series of charts for various kS values, as shown in Figure 4.6. Additional
information regarding suitable design values for kS and other matters were also included. Table 4.2
lists typical values for various materials. These charts are based on the combination of the Colebrook–
White equation (2.10) with the Darcy–Weisbach formula (2.5), to give
𝒌𝒔 𝟐. 𝟓𝟏𝒗
𝑽 = −𝟐√𝟐𝒈𝑫𝑺𝒇 𝐥𝐨𝐠 ( + ) … . (𝟐. 𝟏𝟒)
𝟑. 𝟕𝑫 𝑫√𝟐𝒈𝑫𝑺𝒇

where Sf = hf/L, the hydraulic gradient.


In this equation the velocity (and hence discharge) can be computed directly for a known diameter and
frictional head loss. Subsequently, the Hydraulics Research Station also produced Tables for the
Hydraulic Design of Pipes. These tables are currently published by HR Wallingford, the successor to
the Hydraulics Research Station. In practice, any two of the three variables (Q, D and Sf) may be
known, and therefore the most appropriate solution technique depends on circumstances. For instance,
in the case of an existing pipeline, the diameter and available head are known and hence the discharge
may be found directly from (2.14). For the case of a new installation, the available head and required
discharge are known and the requisite diameter must be found. This will involve a trial-and error
procedure unless the HRS charts or tables are used. Finally, in the case of analysis of pipe networks,
the required discharges and pipe diameters are known and the head loss must be computed. This
problem may be most easily solved using an explicit formula for λ or the HRS charts.

52
Examples illustrating the application of the various methods to the solution of a simple pipe friction
problem now follow.
A pipeline 10 km long, 300 mm in diameter and with roughness size 0.03 mm, conveys water from a
reservoir (top water level 850 m above datum) to a water treatment plant (inlet water level 700 m
above datum). Assuming that the reservoir remains full, estimate the discharge, using the following
methods:
(a) The Colebrook–White formula
(b) The Moody diagrams

53
(c) The HRS charts
Assume ν = 1.13 × 10−6 m2/s.
(a) Using (2.14),
𝐷 = 0.3𝑚, 𝑘𝑠 = 0.03𝑚𝑚
𝑠𝑓 = (850 − 700)/10000 = 0.015
hence
0.03 × 10−3 2.51 × 1.13 × 10−6
𝑉 = −2√2𝑔 × 0.3 × 0.15𝑙𝑜𝑔 ( + )
3.7 × 0.3 0.3√2𝑔 × 0.3 × 0.015
= 2.514𝑚/𝑠
2.514 × 𝜋 × 0.32
𝑄 = 𝑉𝐴 = = 0.178𝑚3 /𝑠
4
(b) The same solution should be obtainable using the Moody diagram; however, it is less accurate
since it involves interpolation from a graph. The solution method is as follows:
(1) Calculate kS/D
(2) Guess a value for V
(3) Calculate Re
(4) Estimate λ using the Moody diagram
(5) Calculate hf
(6) Compare hf with the available head (H)
(7) If H ≠ hf, then repeat from step 2 This is a tedious solution technique, but it shows why the
HRS charts were produced!
(1) kS/D = 0.03 × 10−3/0.3 = 0.0001. (2) As the solution for V has already been found in part
(a) take V = 2.5 m/s.
𝑽𝑫 𝟎.𝟑×𝟐.𝟓
(3) 𝑹𝒆 = = = 𝟎. 𝟔𝟔𝟒 × 𝟏𝟎𝟔
𝝂 𝟏.𝟏𝟑×𝟏𝟎𝟔

(4) Referring to Figure 4.5, Re = 0.664 × 106 and kS/D = 0.0001 confirms that the flow is in
the transitional turbulent region. Following the k S/D curve until it intersects with Re yields λ
0 0. 14 (Note: Interpolation is difficult due to the logarithmic scale.)

𝝀𝑳𝑽𝟐 𝟎.𝟎𝟏𝟒×𝟏𝟎𝟒 ×𝟐.𝟓𝟐


(5) 𝒉𝒇 = = = 𝟏𝟒𝟖. 𝟕𝒎
𝟐𝒈𝑫 𝟐𝒈×𝟎.𝟑

(6) 𝐻 = (850 − 700) = 150 ≠ 148.7

54
(7) A better guess for V is obtained by increasing V slightly. This will not significantly
alter λ, but will increase hf. In this instance, convergence to the solution is rapid because the correct
solution for V was assumed initially!
(c) If the HRS chart is used, then the solution of the equation lies at the intersection of the hydraulic
gradient line (sloping downwards right to left) with the diameter (vertical), reading off the
corresponding discharge (line sloping downwards left to right).
Sf = 0.015 100Sf = 1.5 and D = 300 mm giving Q = 180 L/s = 0.18 m3/s

2.4 Local Head Losses


In addition to head loss due to friction there are always head losses in pipe lines due to bends, junctions,
valves etc. For completeness of analysis these should be taken into account. In practice, in long pipe
lines of several kilometres the effect of local head losses may be negligible. For short pipeline the
losses may be greater than those for friction. A general theory for local losses is not possible; however
rough turbulent flow is usually assumed which gives the simple formula
𝒖𝟐
𝒉𝑳 = 𝑲 𝑳 … … … … … … … . . (𝟐. 𝟏𝟒)
𝟐𝒈
Where hL is the local head loss and kL is a constant for a particular fitting (valve or junction etc.) For
the cases of sudden contraction (e.g. flowing out of a tank into a pipe), of a sudden enlargement (e.g.
flowing from a pipe into a tank) then a theoretical value of kL can be derived. For junctions bend etc.
kL must be obtained experimentally.
2.4.1 Head loss due to a sudden expansion
Expansions and diverging flow are usually associated with an energy loss: the more sudden the expansion, the
greater the loss. Therefore, a very sudden expansion would certainly result in a significant loss of energy. With
short pipelines in particular where minor losses can be significant, this loss must be evaluated and included in
the Bernoulli equation for the system.
Consider water flowing through the expansion in Fig. 6.17. The flow in the smaller pipe will have a
relatively high velocity. On emerging into the larger cross-section, the stream will slow (in accordance
with the continuity equation (A1V1 = A2V2) and gradually expand to fill the larger pipe, as shown by
the streamlines in the diagram. In accordance with the Bernoulli equation, as the velocity decreases
the pressure will increase so as to keep the total energy roughly constant. Thus, a higher pressure exists
in the larger pipe than in the smaller, so the adverse pressure gradient is trying to push the water back
into the smaller pipe. This causes the reduction in velocity required by the continuity equation, and
results in turbulence and eddying of the water occupying the corners of the expansion. Between the

55
live stream of water which is travelling through the larger pipe and the body of eddying water in the
corners of the pipe, a separation boundary will form coinciding approximately with the two outer
streamlines.

Making the assumption that the pressure at the annular area A2 – A1 is equal to the pressure in the
smaller pipe p1. If we apply the momentum equation between positions 1 (just inside the larger
pipe) and 2 to give:
𝒑𝟏 𝑨𝟐 − 𝒑𝟐 𝑨𝟐 = 𝝆𝑸(𝒖𝟐 − 𝒖𝟏 )
Now use the continuity equation to remove Q. (i.e. substitute 𝑸 = 𝑨𝟐 𝒖𝟐)
𝒑𝟏 𝑨𝟐 − 𝒑𝟐 𝑨𝟐 = 𝝆𝑨𝟐 𝒖𝟐 (𝒖𝟐 − 𝒖𝟏 )
Rearranging and dividing by g give
𝒑𝟐 − 𝒑𝟏 𝒖 𝟐
= (𝒖 − 𝒖𝟐 ) … … … … … … … … . . (𝟐. 𝟏𝟕)
𝝆𝒈 𝒈 𝟏
Now apply the Bernoulli equation from point 1 to 2, with the head loss term hL
𝒑𝟏 𝒖 𝟏 𝟐 𝒑𝟐 𝒖 𝟐 𝟐
+ = + + 𝒉𝑳
𝝆𝒈 𝟐𝒈 𝝆𝒈 𝟐𝒈
and rearranging gives,
𝒖𝟏 𝟐 −𝒖𝟐 𝟐 𝒑𝟐 −𝒑𝟏
𝒉𝑳 = − … … … . . (𝟐. 𝟏𝟖)
𝟐𝒈 𝝆𝒈

Combining Equations 2.17 and 2.18 gives


𝒖 𝟏 𝟐 − 𝒖𝟐 𝟐 𝒖𝟐
𝒉𝑳 = − (𝒖𝟏 − 𝒖𝟐 )
𝟐𝒈 𝒈
(𝒖𝟏 − 𝒖𝟐 )𝟐
𝒉𝑳 = … … … … … … … … … … . (𝟐. 𝟏𝟗)
𝟐𝒈
Substituting again for the continuity equation to get an expression involving the two areas, (i.e.
u2=u1A1/A2) gives

56
𝑨𝟏 𝟐 𝒖 𝟏 𝟐
𝒉𝑳 = (𝟏 − ) … … … … … … … … … . . (𝟐. 𝟐𝟎)
𝑨𝟏𝟐 𝟐𝒈
Comparing this with Equation 2.14 gives kL
𝑨𝟏 𝟐
𝒌𝑳 = (𝟏 − ) … … … … … … (𝟐. 𝟐𝟏)
𝑨𝟏𝟐
When a pipe expands in to a large tank A1 << A2 i.e. A1/A2 = 0 so kL = 1. That is, the head loss is equal
to the velocity head just before the expansion into the tank.
2.4.2 Losses at Sudden Contraction

In a sudden contraction, flow contracts to point 1, forming a vena contraction. From experiment it has
been shown that this contraction is commonly about 40% (i.e. A1 = 0.6 A2). It is possible to assume
that energy losses from up to to 1 are negligible (no separation occurs in contracting flow) but that
major losses occur between 1 and 2 as the flow expands again. In this case Equation 2.20 can be used
from point 1 to 2 to give: (by continuity u 1 = A2u2/A1 = A2u2/0.6A2 = u2/0.6).
𝟎. 𝟔𝑨𝟐 (𝒖𝟐 /𝟎. 𝟔)𝟐
𝒉𝑳 = (𝟏 − )
𝑨𝟐 𝟐𝒈
𝒖𝟐 𝟐
𝒉𝑳 = 𝟎. 𝟒𝟒 … … … … … … … … (𝟐. 𝟐𝟐)
𝟐𝒈
i.e. At a sudden contraction kL = 0.44.
As the difference in pipe diameters gets large (A1/A2) then this value of kL will tend towards 0.5
which is equal to the value for entry loss from a reservoir into a pipe.
2.4.3 Other Local Losses
Large losses in energy in energy usually occur only where flow expands. The mechanism at work in
these situations is that as velocity decreases (by continuity) so pressure must increase (by Bernoulli).
When the pressure increases in the direction of fluid outside the boundary layer has enough momentum
to overcome this pressure that is trying to push it backwards. The fluid within the boundary layer has

57
so little momentum that it will very quickly be brought to rest, and possibly reversed in direction. If
this reversal occurs it lifts the boundary layer away from the surface as shown in Figure 8. This
phenomenon is known as boundary layer separation.

At the edge of the separated boundary layer, where the velocities change direction, a line of vortices
occur (known as a vortex sheet). This happens because fluid to either side is moving in the opposite
direction. This boundary layer separation and increase in the turbulence because of the vortices results
in very large energy losses in the flow. These separating / divergent flows are inherently unstable and
far more energy is lost than in parallel or convergent flow. Some common situation where significant
head losses occur in pipe are shown in figure 9

The values of kL for these common situations are shown in Table 2. It gives value that are used in
practice.

58
2.5 Pipeline Analysis
As discussed at the start of this chapter, for analysis of flow in pipelines we will use the Bernoulli
equation. Bernoulli’s equation is a statement of conservation of energy along a streamline, by this
principle the total energy in the system does not change, Thus the total head does not change. So the
Bernoulli equation can be written
𝒑 𝒖𝟐
+ + 𝒛 = 𝑯 = 𝑪𝒐𝒏𝒔𝒕𝒂𝒏𝒕
𝝆𝒈 𝟐𝒈

As all of these elements of the equation have units of length, they are often referred to as the following:
p u2
pressure head = ρg, velocity head = 2g, potential head = z and total head = H

59
In this form Bernoulli’s equation has some restrictions in its applicability, they are:
• Flow is steady;
• Density is constant (i.e. fluid is incompressible);
• Friction losses are negligible.
• The equation relates the states at two points along a single streamline.
Applying the equation between two points an including, entry, expansion, exit and friction losses, we
have
𝒑𝟏 𝒖 𝟏 𝟐 𝒑𝟐 𝒖 𝟐 𝟐
+ + 𝒛𝟏 = + + 𝒛𝟐 + 𝒉𝑳 𝒆𝒏𝒕𝒓𝒚 + 𝒉𝑳 𝒆𝒙𝒑𝒂𝒏𝒔𝒊𝒐𝒏 + 𝒉𝑳 𝒆𝒙𝒊𝒕 + 𝒉𝒇
𝝆𝒈 𝟐𝒈 𝝆𝒈 𝟐𝒈
Below we will see how these can be viewed graphically, then we will solve some typical problems for
pipelines and their various losses.
2.6. Pressure Head, Velocity Head, Potential Head and Total Head in a Pipeline.
By looking at the example of the reservoir with which feeds a pipe we will see how these different
heads relate to each other. Consider the reservoir below feeding a pipe of constant diameter that rises
(in reality it may have to pass over a hill) before falling to its final level.

To analyze the flow in the pipe we apply the Bernoulli equation along a streamline from point 1 on
the surface of the reservoir to point 2 at the outlet nozzle of the pipe. And we know that the total energy
per unit weight or the total head does not change - it is constant - along a streamline. But what is this
value of this constant? We have the Bernoulli equation;
𝒑𝟏 𝒖 𝟏 𝟐 𝒑𝟐 𝒖 𝟐 𝟐
+ + 𝒛𝟏 = 𝑯 = + + 𝒛𝟐
𝝆𝒈 𝟐𝒈 𝝆𝒈 𝟐𝒈
We can calculate the total head, H, at the reservoir, p1 = 0 as this is atmospheric and atmospheric
gauge pressure is zero, the surface is moving very slowly compared to that in the pipe so u1 = 0 , so
all we are left with is total head = H = z1 the elevation of the reservoir. A useful method of analyzing

60
the flow is to show the pressures graphically on the same diagram as the pipe and reservoir. In the
figure above the total head line is shown. If we attached piezometers at points along the pipe, what
would be their levels when the pipe nozzle was closed? (Piezometers, as you will remember, are simply
open-ended vertical tubes filled with the same liquid whose pressure they are measuring)

As you can see in the above figure, with zero velocity all of the levels in the piezometers are equal and
the same as the total. At each point on the line, when u=0, the Bernoulli equation says
𝒑
+ 𝒛𝟏 = 𝑯
𝝆𝒈
𝒑
The level in the piezometer is the pressure head and its value is given by .
𝝆𝒈

What would happen to the levels in the piezometers (pressure heads) if the water was flowing with
velocity = u? We know from earlier examples that as velocity increases so pressure falls.

𝒖𝟐
We see in this figure that the levels have reduced by an amount equal to the velocity head, 𝟐𝒈
. Now

as the pipe is of constant diameter, we know that the velocity is constant along the pipe so the velocity

61
head is constant and represented graphically by the horizontal line shown. (this line is known as the
hydraulic grade line). What would happen if the pipe were not of constant diameter? Look at the figure
below where the pipe from the example above is replaced by a pipe of three sections with the middle
section of larger diameter.

The velocity head at each point is now different. This is because the velocity is different at each point.
By considering continuity we know that the velocity is different because the diameter of the pipe is
different. Which point has the greatest diameter?
Pipe 2, because the velocity, and hence the velocity head, is the smallest. This graphical representation
has the advantage that we can see at a glance the pressures in the system. For example, where along
the whole line is the lowest pressure head? It is where the hydraulic grade line is nearest to the pipe
elevation i.e. at the highest point of the pipe.
2.5.1 Flow in pipes with losses due to friction.
In a real pipe line there are energy losses due to friction - these must be taken into account as they can
be very significant. How would the pressure and hydraulic grade lines change with friction? Going
back to the constant diameter pipe, we would have a pressure situation like this shown below.

How can the total head be changing? We have said that the total head - or total energy per unit weight
- is constant. We are considering energy conservation, so if we allow for an amount of energy to be
lost due to friction the total head will change. Equation 19 is the Bernoulli equation as applied to a

62
pipe line with the energy loss due to friction written as a head and given the symbol h f (the head loss
due to friction) and the local energy losses written as a head, h L (the local head loss).
𝒑𝟏 𝒖 𝟏 𝟐 𝒑𝟐 𝒖 𝟐 𝟐
+ + 𝒛𝟏 = 𝑯 = + + 𝒛𝟐 + 𝒉𝒇 + 𝒉𝑳 … … … . (𝟐. 𝟐𝟑)
𝝆𝒈 𝟐𝒈 𝝆𝒈 𝟐𝒈
10 Reservoir and Pipe Example
Consider the example of a reservoir feeding a pipe, as shown in the figure below

The pipe diameter is 100mm and has length 15m and feeds directly into the atmosphere at point C 4m
below the surface of the reservoir (i.e. za – zc = 4.0m). The highest point on the pipe is a B which is
1.5m above the surface of the reservoir (i.e. zb – za = 1.5m) and 5 m along the pipe measured from the
reservoir. Assume the entrance and exit to the pipe to be sharp and the value of friction factor f to be
0.08. Calculate a) velocity of water leaving the pipe at point C, b) pressure in the pipe at point B.
a) We use the Bernoulli equation with appropriate losses from point A to C and for entry loss k L = 0.5
and exit loss kL = 1.0.
For the local losses from Table 2 for a sharp entry kL = 0.5 and for the sharp exit as it opens in to the
atmosphere with no contraction there are no losses, so
𝒖𝟐
𝒉𝑳 = 𝟎. 𝟓
𝟐𝒈
Friction losses are given by the Darcy equation
4𝑓𝐿𝑢2
ℎ𝑓 =
2𝑔𝑑
Pressure at A and C are both atmospheric, uA is very small so can be set to zero, giving
𝑢2 4𝑓𝐿𝑢2 𝑢2
𝑧𝐴 = + 𝑧𝐶 + + 0.5
2𝑔 2𝑔𝑑 2𝑔

63
𝑢2 4𝑓𝐿
𝑧𝐴 − 𝑧𝐶 = (1 + 0.5 + )
2𝑔 𝑑
Substitute in the numbers from the question
𝑢2 4 × 0.08 × 15
4= (1.5 + ) , 𝑢 = 1.26𝑚/𝑠
2 × 9.81 0.1
(a) To find the pressure at B apply Bernoulli from point A to B using the velocity calculated above.
The length of the pipe is L1 = 5m:
𝑝𝐵 𝑢2 4𝑓𝐿1 𝑢2 𝑢2
𝑧𝐴 = + + 𝑧𝐵 + + 0.5
𝜌𝑔 2𝑔 2𝑔𝑑 2𝑔
𝑝𝐵 𝑢2 4𝑓𝐿1
𝑧𝐴 − 𝑧𝐵 = + (1 + 0.5 + )
𝜌𝑔 2𝑔 𝑑
𝑝𝐵 1.262 4 × 0.08 × 5.0
−1.5 = + (1.5 + )
1000 × 9.81 2 × 9.81 0.1

𝑝𝐵 = −28.58𝑘𝑁/𝑚2
That is 28.58 kN/m2 below atmospheric.
2.5.2 Pipes in Series Example
Consider the two reservoirs shown in figure 16, connected by a single pipe that changes diameter over
its length. The surfaces of the two reservoirs have a difference in level of 9m. The pipe has a diameter
of 200mm for the first 15m (from A to C) then a diameter of 250mm for the remaining 45m (from C
to B).

For the entrance use kL = 0.5 and the exit kL = 1.0. The join at C is sudden. For both pipes use f = 0.01.
Total head loss for the system H = height difference of reservoirs
hf1 = head loss for 200mm diameter section of pipe
hf2 = head loss for 250mm diameter section of pipe
hL entry = head loss at entry point
hL join = head loss at join of the two pipes hL exit = head loss at exit point

64
So
H = hf1 + hf2 + hL entry + hL join + hL exit = 9m
All losses are, in terms of Q:
𝑓𝐿1 𝑄 2
ℎ𝑓1 =
3𝑑1 5
𝑓𝐿2 𝑄 2
ℎ𝑓2 =
3𝑑2 5
2
𝑢1 2 1 4𝑄 𝑄2 𝑄2
ℎ𝐿 𝑒𝑛𝑡𝑟𝑦 = 0.5 = 0.5 ( 2 ) = 0.5 × 0.0826 4 = 0.0413
2𝑔 2𝑔 𝜋𝑑1 𝑑1 𝑑1 4
𝑢2 2 𝑄2 𝑄2
ℎ𝐿 𝑒𝑥𝑖𝑡 = 1.0 = 1.0 × 0.0826 4 = 0.0826 4
2𝑔 𝑑2 𝑑2
2
1 1
2( − ) 2
(𝑢1 − 𝑢2 )2
4𝑄 𝑑1 2 𝑑2 2 1 1
ℎ𝐿 𝑗𝑜𝑖𝑛𝑡 = =( ) = 0.0826𝑄 2 ( 2 − 2 )
2𝑔 𝜋 2𝑔 𝑑1 𝑑2
Substitute these into
hf1 + hf2 + hL entry + hL join + hL exit = 9 and
solve for Q, to give Q = 0.158 m3 /s
2.5.3 Pipes in parallel
When two or more pipes in parallel connect two reservoirs, as shown in Figure 17, for example, then
the fluid may flow down any of the available pipes at possible different rates. But the head difference
over each pipe will always be the same. The total volume flow rate will be the sum of the flow in each
pipe. The analysis can be carried out by simply treating each pipe individually and summing flow rates
at the end.

Pipes in Parallel Example Two pipes connect two reservoirs (A and B) which have a height difference
of 10m. Pipe 1 has diameter 50mm and length 100m. Pipe 2 has diameter 100mm and length 100m.
Both have entry loss kL = 0.5 and exit loss kL=1.0 and Darcy f of 0.008. Calculate: a) rate of flow for
each pipe b) the diameter D of a pipe 100m long that could replace the two pipes and provide the same
flow.

65
(a) Apply Bernoulli to each pipe separately. For pipe 1:
𝒑𝑨 𝒖 𝑨 𝟐 𝒑𝑩 𝒖 𝑩 𝟐 𝒖𝟏 𝟐 𝟒𝒇𝑳𝟐 𝒖𝟏 𝟐 𝒖𝟏 𝟐
+ + 𝒛𝑨 = + + 𝒛𝑩 + 𝟎. 𝟓 + + 𝟏. 𝟎
𝝆𝒈 𝟐𝒈 𝝆𝒈 𝟐𝒈 𝟐𝒈 𝟐𝒈𝒅𝟏 𝟐𝒈
pA and pB are atmospheric, and as the reservoir surface move s slowly uA and uB are negligible,
so
4𝑓𝐿 𝑢1 2
𝑧𝐴 − 𝑧𝐵 = (0.5 + + 1.0)
𝑑1 2𝑔
4 × 0.008 × 100 𝑢1 2
10 = (0.5 + + 1.0)
0.05 2 × 9.81
𝑢1 = 1.731𝑚/𝑠
𝜋 𝒅𝟐𝟏
And flow rate is given by 𝑄1 = 𝒖𝟏 = 0.0034𝑚3 /𝑠
4

For pipe 2:
𝒑𝑨 𝒖 𝑨 𝟐 𝒑𝑩 𝒖 𝑩 𝟐 𝒖𝟏 𝟐 𝟒𝒇𝑳𝟐 𝒖𝟏 𝟐 𝒖𝟏 𝟐
+ + 𝒛𝑨 = + + 𝒛𝑩 + 𝟎. 𝟓 + + 𝟏. 𝟎
𝝆𝒈 𝟐𝒈 𝝆𝒈 𝟐𝒈 𝟐𝒈 𝟐𝒈𝒅𝟏 𝟐𝒈

Again, pA and pB are atmospheric, and as the reservoir surface move slowly uA and uB are
negligible, so
4𝑓𝐿 𝑢2 2
𝑧𝐴 − 𝑧𝐵 = (0.5 + + 1.0)
𝑑2 2𝑔
4 × 0.008 × 100 𝑢2 2
10 = (1.5 + )
0.05 2 × 9.81
𝑢2 = 2.42𝑚/𝑠
And flow rate is given by
𝜋 𝒅𝟐𝟐
𝑄1 = 𝒖𝟐 = 0.0190𝑚3 /𝑠
4
(b) Replacing the pipe, we need Q=Q1+Q2=0.0034+0.0190=0.0224m3/s
For this pipe, diameter D, velocity u, and making the same assumptions about entry/exit losses,
we have:
𝒑𝑨 𝒖 𝑨 𝟐 𝒑𝑩 𝒖 𝑩 𝟐 𝒖𝟐 𝟒𝒇𝑳𝟐 𝒖𝟐 𝒖𝟐
+ + 𝒛𝑨 = + + 𝒛𝑩 + 𝟎. 𝟓 + + 𝟏. 𝟎
𝝆𝒈 𝟐𝒈 𝝆𝒈 𝟐𝒈 𝟐𝒈 𝟐𝒈𝒅𝟏 𝟐𝒈
4𝑓𝐿 𝑢2
𝑧𝐴 − 𝑧𝐵 = (0.5 + + 1.0)
𝐷 2𝑔

66
4 × 0.008 × 100 𝑢2
10 = (1.5 + )
𝐷 2 × 9.81
3.2 2
196.2 = (1.5 + )𝑢
𝐷
The velocity can be obtained from Q i.e.
𝜋𝐷2
𝑄 = 𝐴𝑢 = 𝑢
4
4𝑄 0.02852
𝑢= 2
=
𝜋𝐷 𝐷2
3.2 0.02852 2
So 196.2 = (1.5 + )( )
𝐷 𝐷2

0 = 241212𝐷5 − 1.5𝐷 − 3.2


which must be solved iteratively.
An approximate answer can be obtained by dropping the second term:
0 = 241212𝐷5 − 3.2
5
𝐷 = √3.2/241212 = 0.1058𝑚
Writing the function f(D)=241212D5-1.5D-3.2
f (01058) =-0.161
So, increase D slightly, try 0.107m f (0.107) = 0.022 i.e. the solution is between 0.107m and 0.1058m
but 0.107 if sufficiently accurate.
An alternative method (although based on the same theory) is shown below using the Darcy equation
in terms of Q.
𝒇𝑳𝑸𝟐
𝒉𝒇 =
𝟑𝒅𝟓
And the loss equations in terms of Q:
𝑢2 𝑄2 42 𝑄 2 𝑄2
ℎ𝐿 = 𝑘 =𝑘 2
=𝑘 2 4
= 0.0826𝑘 4
2𝑔 2𝑔𝐴 2𝑔𝜋 𝑑 𝑑
For Pipe 1 10=hLentry +hf + hLexit
𝑄2 0.008 × 100𝑄 2 𝑄2
10 = 0.0826 × 0.5 4
+ 5
+ 0.0826𝑘
0.05 3 × 0.05 0.054
𝑄 = 0.0034𝑚3 /𝑠 = 3.4𝑙𝑖𝑡𝑟𝑒𝑠/𝑠
For Pipe 1 10=hLentry +hf + hLexit
𝑄2 0.008 × 100𝑄 2 𝑄2
10 = 0.0826 × 0.5 4
+ 5
+ 0.0826𝑘
0.1 3 × 0.1 𝑑0.14

67
0.0188𝑚3
𝑄= = 18.8𝑙𝑖𝑡𝑟𝑒𝑠/𝑠
𝑠
2.5.4 Branched pipes
If pipes connect three reservoirs, as shown in Figure 17, then the problem becomes more complex.
One of the problems is that it is sometimes difficult to decide which direction fluid will flow. In
practice solutions are now done by computer techniques that can determine flow direction, however it
is useful to examine the techniques necessary to solve this problem.

For these problems it is best to use the Darcy equation expressed in terms of discharge – i.e. equation
2.7.
𝒇𝑳𝑸𝟐
𝒉𝒇 =
𝟑𝒅𝟓
When three or more pipes meet at a junction then the following basic principles apply:
1. The continuity equation must be obeyed i.e. total flow into the junction must equal total flow
out of the junction;
2. at any one point there can only be one value of head, and
3. Darcy’s equation must be satisfied for each pipe. It is usual to ignore minor losses (entry and
exit losses) as practical hand calculations become impossible – fortunately they are often
negligible.
One problem still to be resolved is that however we calculate friction it will always produce a
positive drop – when in reality head loss is in the direction of flow. The direction of flow is often
obvious, but when it is not a direction has to be assumed. If the wrong assumption is made then no
physically possible solution will be obtained. In the figure above the heads at the reservoir are
known but the head at the junction D is not. Neither are any of the pipe flows known. The flow in

68
pipes 1 and 2 are obviously from A to D and D to C respectively. If one assumes that the flow in
pipe 2 is from D to B then the following relationships could be written:
𝒁𝒂 − 𝒉𝑫 = 𝒉𝒇𝟏
𝒉𝑫 − 𝒁𝒃 = 𝒉𝒇𝟐
𝒉𝑫 − 𝒁𝒄 = 𝒉𝒇𝟑
𝑸𝟏 = 𝑸𝟐 + 𝑸𝟑
The hf expressions are functions of Q, so we have 4 equations with four unknowns, h D, Q1, Q2 and Q3
which we must solve simultaneously. The algebraic solution is rather tedious so a trial and error
method is usually recommended. For example, this procedure usually converges to a solution quickly:
1. estimate a value of the head at the junction, hD
2. substitute this into the first three equations to get an estimate for Q for each pipe.
3. check to see if continuity is (or is not) satisfied from the fourth equation
4. if the flow into the junction is too high choose a larger h D and vice versa.
5. return to step 2 If the direction of the flow in pipe 2 was wrongly assumed then no solution will be
found. If you have made this mistake then switch the direction to obtain these four equations
𝒁𝒂 − 𝒉𝑫 = 𝒉𝒇𝟏
𝒁𝒃 − 𝒉𝑫 = 𝒉𝒇𝟐
𝒉𝑫 − 𝒁𝒄 = 𝒉𝒇𝟑
𝑸𝟏 + 𝑸𝟐 = 𝑸𝟑
Looking at these two sets of equations we can see that they are identical if hD = zb. This suggests that
a good starting value for the iteration is zb then the direction of flow will become clear at the first
iteration.
Example of Branched Pipe –
The Three Reservoir Problem Water flows from reservoir A through pipe 1, diameter d1 = 120mm,
length L1=120m, to junction D from which the two pipes leave, pipe 2, diameter d2=75mm, length
L2=60m goes to reservoir B, and pipe 3, diameter d3=60mm, length L3=40m goes to reservoir C.
Reservoir B is 16m below reservoir A, and reservoir C is 24m below reservoir A. All pipes have f=
0.01. (Ignore and entry and exit losses.) We know the flow is from A to D and from D to C but are
never quite sure which way the flow is along the other pipe – either D to B or B to D. We first must
assume one direction. If that is not correct there will not be a sensible solution. To keep the notation
from above we can write za = 24, zb = 16 and zc = 0. For flow A to D

69
𝑧𝑎 − ℎ𝐷 = ℎ𝑓1
𝑓1 𝐿1 𝑄1 2
24 − ℎ𝐷 = = 16075𝑄1 2
3𝑑1 5
Assume flow is D to B ℎ𝐷 − 𝑧𝑏 = ℎ𝑓2
𝑓2 𝐿2 𝑄2 2
ℎ𝐷 − 8 = = 84280𝑄2 2
3𝑑2 5

Flow is from D to C
ℎ𝐷 − 𝑧𝑐 = ℎ𝑓3
𝑓3 𝐿3 𝑄3 2
ℎ𝐷 − 0 = = 171468𝑄3 2
3𝑑3 5
The final equation is continuity, which for this chosen direction D to B is
𝑄1 = 𝑄2 + 𝑄3
Now it is a matter of systematically questing values of hD until continuity is satisfied. This is best done
in a table. And it is usually best to initially guess hD = za then reduce its value (until the error in
continuity is small):
hj Q1 Q2 Q3 Q1=Q2+Q3 err
24.00 0.0000 0.01378 0.01183 0.02561 0.02561
20.00 0.01577 0.01193 0.01080 0.02273 0.00696
17.00 0.02087 0.01033 0.00996 0.02029 -0.00058
17.10 0.02072 0.01039 0.00999 0.02038 -0.00034
17.20 0.02057 0.01045 0.01002 0.02046 -0.00010
17.30 0.02042 0.01050 0.01004 0.02055 0.00013
17.25 0.02049 0.01048 0.01003 0.02051 0.00001
17.24 0.02051 0.01047 0.01003 0.02050 -0.00001

So, the solution is that the head at the junction is 17.24 m, which gives Q1 = 0.0205m3 /s, Q1 =
0.01047m3 /s and Q1 = 0.01003m3 /s.
Had we guessed that the flow was from B to D, the second equation would have been
𝒛𝒃 − 𝒉𝑫 = 𝒉𝒇𝟐

𝒇𝟐 𝑳𝟐 𝑸𝟐 𝟐
𝟖 − 𝒉𝑫 = = 𝟖𝟒𝟐𝟖𝟎𝑸𝟐 𝟐
𝟑𝒅𝟐 𝟓
and continuity would have been Q1 + Q2 = Q3. If you then attempted to solve this you would soon see
that there is no solution. An alternative method to solve the above problem is shown below. It does

70
not solve the head at the junction, instead directly solves for a velocity (it may be easily amended to
solve for discharge Q) [For this particular question the method shown above is easier to apply – but
the method shown below could be seen as more general as it produces a function that could be solved
by a numerical method and so may prove more convenient for other similar situations.] Again, for this
we will assume the flow will be from reservoir A to junction D then from D to reservoirs B and C.
There are three unknowns u1, u2 and u3 the three equations we need to solve are obtained from A to
B then A to C and from continuity at the junction D.
Flow from A to B
𝒑𝑨 𝒖 𝑨 𝟐 𝒑𝑩 𝒖 𝑩 𝟐 𝟒𝒇𝑳𝟐 𝒖𝟏 𝟐 𝟒𝒇𝑳𝟐 𝒖𝟐 𝟐
+ + 𝒛𝑨 = + + 𝒛𝑩 + +
𝝆𝒈 𝟐𝒈 𝝆𝒈 𝟐𝒈 𝟐𝒈𝒅𝟏 𝟐𝒈𝒅𝟐
Putting pA = pB and taking uA and uB as negligible, gives
𝟒𝒇𝑳𝟐 𝒖𝟏 𝟐 𝟒𝒇𝑳𝟐 𝒖𝟐 𝟐
𝒛𝑨 − 𝒛𝑩 = +
𝟐𝒈𝒅𝟏 𝟐𝒈𝒅𝟐
Put in the numbers from the question
𝟒 × 𝟎. 𝟎𝟏 × 𝟏𝟐𝟎𝒖𝟏 𝟐 𝟒 × 𝟎. 𝟎𝟏 × 𝟔𝟎𝒖𝟐 𝟐
𝟏𝟔 = +
𝟐𝒈 × 𝟎. 𝟎𝟏𝟐 𝟐𝒈 × 𝟎. 𝟎𝟕𝟓
𝟏𝟔 = 𝟐. 𝟎𝟑𝟖𝟕𝒖𝟏 𝟐 + 𝟏. 𝟔𝟑𝟏𝟎𝒖𝟐 𝟐 (𝒆𝒒𝒖𝒂𝒕𝒊𝒐𝒏 𝒊𝒊)
From continuity at the junction
Flow A to D = Flow D to B + Flow D to C
𝑄1 = 𝑄2 + 𝑄3
2
𝜋𝑑1 𝜋𝑑2 2 𝜋𝑑3 2
𝑢1 = 𝑢2 + 𝑢3
4 4 4
𝑑2 2 𝑑3 2
𝑢1 = ( ) 𝑢2 + ( ) 𝑢3
𝑑1 𝑑1
𝑢1 − 0.3906𝑢2 − 0.25𝑢3 = 0 (𝑒𝑞𝑢𝑎𝑡𝑖𝑜𝑛 𝑖𝑖𝑖)
the values of u1, u2 and u3 must be found by solving the simultaneous equation i, ii and iii. The
technique to do this is to substitute for equations i, and ii in to equation iii, then solve this expression.
It is usually done by a trial and error approach. i.e. from i,
𝒖𝟐 = √𝟗. 𝟖𝟏 − 𝟏. 𝟐𝟓𝒖𝟏 𝟐

from ii, 𝒖𝟑 = √𝟏𝟕. 𝟔𝟓𝟕 − 𝟏. 𝟓𝒖𝟏 𝟐


substituting in iii gives

𝒖𝟏 = 𝟎. 𝟑𝟗𝟎𝟔√𝟗. 𝟖𝟏 − 𝟏. 𝟐𝟓𝒖𝟏 𝟐 − 𝟎. 𝟐𝟓√𝟏𝟕. 𝟔𝟓𝟕 − 𝟏. 𝟓𝒖𝟏 𝟐 = 𝟎 = 𝒇(𝒖𝟏 )

71
This table shows some trial and error solutions
u f(u)
1 -1.14769
2 0.289789
1.8 -0.03176
1.85 0.046606
1.83 0.015107
1.82 -0.00057
Giving u1 = 1.82 m/s, so u2 = 2.38 m/s, u3 = 12.69 m/s
𝜋𝑑1 2
Flow rates are 𝑄1 = 4
𝑢1 = 0.0206𝑚3 /𝑠

𝜋𝑑2 2
𝑄2 = 𝑢2 = 0.0101𝑚3 /𝑠
4
𝜋𝑑3 2
𝑄3 = 𝑢3 = 0.0206𝑚3 /𝑠
4

Check for continuity at the junction


𝑄1 = 𝑄2 + 𝑄3
0.0206 = 0.0105 + 0.0101
2.6 Complex pipe network analysis
The solution methods described for the analysis of series, parallel and branched pipes are not very
suitable for the more complex case of networks. A network consists of loops and nodes as shown in
Figure 12.4.

Applying the continuity equation to a node, the sum of the inflows towards the node must equal the
sum of the outflows, therefore, treating outflows as negative and inflows as positive:
𝒏

∑ 𝒒𝒊 = 𝟎 … … … … … … … … … (𝟐. 𝟐𝟎)
𝒊=𝟐

where n is the number of pipes joined at the node.

72
Turning to the diagram of the loop in Figure 12.4, we can track the energy losses by starting at a node
and travelling around the loop calculating the loss for each pipe in turn. Here, it will be assumed that
the direction of travel is clockwise and that flows and head losses are positive clockwise (so in Figure
12.4, there are two positive flows and two negative). Therefore, applying the energy equation,
𝒎

∑ 𝒉𝒇𝒊 = 𝟎 … … … … … … … … … … . (𝟐𝟎. 𝟐𝟏)


𝒊=𝟐

where m is the number of pipes in a loop. The sign convention sets flow and head loss as positive
clockwise. In addition,
𝒉𝒇𝒊 = 𝒇(𝒒𝒊 ) … … … … . (𝟐. 𝟐𝟐)
where f(qi ) represents the Darcy–Weisbach/Colebrook–White equation. Equations 2.20 through 20.22
comprise a set of simultaneous non-linear equations, and an iterative solution is generally adopted.
The two standard solution techniques, the loop method and the nodal method, are now discussed.
2.6.1 Loop (head balance) Method
This method, originally proposed by Hardy-Cross in 1936, essentially consists of eliminating the head
losses from (2.20) and (2.22) to give a set of equations in discharge only. It may be applied to loops
where the external discharges are known and the flows within the loop are required. The basis of the
method is as follows:
1. Assume values for qi to satisfy Σqi = 0
2. Calculate hfi from qi
3. If Σhfi = 0, then the solution is correct
4. If Σhfi ≠ 0, then apply a correction factor δq to all qi and return to (2) A reasonably efficient
value of δq for rapid convergence is given by
∑ 𝒉𝒇𝒊
𝜹𝒒 = − …………………(2.23)
𝟐 ∑ 𝒉𝒇𝒊/ 𝒒𝒊
Step 2 may be carried out using the HRS charts or tables (for hand calculations) or using Barr’s explicit
formula for λ (for computer solution). Due account must be taken of the sign of qi and hfi, and of the
use of an appropriate convention (i.e., clockwise positive). Equation 2.23 may be derived as follows.
Using the Darcy–Weisbach formula
𝝀𝑳𝑽𝟐
𝒉𝒇 =
𝟐𝒈𝑫
or, for a given pipe and assuming that λ is constant
𝒉𝒇 = 𝑲𝑸𝟐

73
Taking the true flow to be Q, then
𝑸 = (𝒒𝒊 + 𝜹𝒒)
and taking the true head loss to be Hf
𝐻𝑓 = 𝑘(𝑞𝑖 + 𝛿𝑞)2
Expanding by the binomial theorem,
𝛿𝑞 2(2 − 1) 𝛿𝑞 2
𝐻𝑓 = 𝑘𝑞𝑖 2 [1 + 2 + ( ) + …………]
𝑞𝑖 2! 𝑞𝑖
ignoring second-order terms and above (for δq << qi)
𝐻𝑓 = 𝑘𝑞𝑖 2 [1 + (2𝛿𝑞/𝑞𝑖 )]
For a loop

∑ 𝐻𝑓𝑖 = 0 = ∑ 𝑘𝑞𝑖 2 + 2𝛿𝑞 ∑ 𝑘𝑞𝑖 2 / 𝑞𝑖

or

0 = ∑ ℎ𝑓𝑖 + 2𝛿𝑞 ∑ ℎ𝑓𝑖 /𝑞𝑖

Hence
∑ ℎ𝑓𝑖
𝛿𝑞 = −
2 ∑ ℎ𝑓𝑖 /𝑞𝑖
Example:
For the square pipe loop shown in the figure below, find the following: (a) The discharges in the loop
(b) The pressure heads at points B, C and D, if the pressure head at A is 70 m and A, B, C and D have
the same elevations
All pipes are 1 km long and 300 mm in diameter, with roughness 0.03 mm.

74
75
2.6.2 Nodal Method
This method, originally proposed by Cornish in 1939, consists of eliminating the discharges from
(2.21) and (2.22) to give a set of equations in head losses only. It may be applied to loops or branches
where the external heads are known and the heads within the networks are required. The basis of the
method is as follows:
1. Assume values for the head (HJ) at each junction
2. Calculate qi from HJ
3. If Σqi = 0, then the solution is correct
4. If Σqi ≠ 0, then apply a correction factor δH to HJ and return to 2, where
𝟐 ∑ 𝒒𝒊
𝜹𝑯 = ………………….(2.25)
∑ 𝒒𝒊 /𝒉𝒇𝒊

The derivation of (2.25) is similar to that for (2.24). Step 2 may be carried out using the HRS charts
or tables (for hand calculations) or using the Colebrook–White/Darcy–Weisbach equation for q in
terms of hf (for computer solution). An appropriate sign convention must be used for qi and h fi , e.g.,
qi and hfi positive entering a node.
Example
Resolve the preceding example using the nodal method.
Solution Assume a trial value of HJ = 750 m

76
2.7 Complex Networks
For more complex networks (e.g., more than one loop or junction), the loop and nodal methods may
both be applied with minor modifications. In the case of the loop method, the correction factor δq at
each iteration must be carried over from one loop to the next through any common pipes. For the nodal
method, the correction factor δH is applied to successive nodes through the network at each iteration.
The choice of loop or nodal method for the analysis of complex networks depends on the available
data. In terms of efficiency of solution and required computer storage space, the loop method requires
more storage but converges more quickly, whereas the nodal method requires less storage but
convergence is slower and not always achieved. Some computer packages use a hybrid (loop-nodal)
method.
A computational model, EPNET is also available for analysis and design of complex hydraulic
networks (Tutorial attached).

77
CHAPTER THREE: HYDRAULIC MACHINES

3.1 Introduction
The two types of hydraulic machinery of interest to the water resources Engineer are pumps and
turbines.
A pump converts mechanical energy into hydraulic energy while a turbine serves the opposite purpose.
Though most machines are single purpose, the reversible pump-turbine operates as either a pump or
turbine depending on the direction of rotation. There are many types of pumps and turbines, each type
has its own characteristics and is suitable for specific operating conditions. The design of hydraulic
machinery is a highly specialized field, and the emphasis of this chapter will be on selection and
application of the machinery.
3.2 Energy Transfer in Pumps and Turbines
Pumps and turbines are energy conversion devices:
pumps turn electrical or mechanical energy into fluid energy;
turbines turn fluid energy into electrical or mechanical energy
The energy per unit weight is the head, H:
𝒑 𝑽𝟐
𝑯= +𝒛+
𝝆𝒈 𝟐𝒈
The first two terms on the RHS comprise the piezometric head. The last term is the dynamic head.
Power
Power = rate of conversion of energy. If a mass m is raised through a height H it gains energy mgH.
If it does so in time t then the rate of conversion is mgH/t. For a fluid in motion the mass flow rate
(m/t) is ρQ. The rate of conversion to or from fluid energy when the total head is changed by H is,
therefore, 𝜌𝑄𝑔𝐻, or
𝑷𝒐𝒘𝒆𝒓 = 𝝆𝒈𝑸𝑯
Efficiency
Efficiency, 𝜂, is given by

78
𝒑𝒐𝒘𝒆𝒓𝒐𝒖𝒕
𝜼=
𝒑𝒐𝒘𝒆𝒓𝒊𝒏
where “powerout” refers to the useful power; i.e. excluding losses.
𝒑𝒐𝒘𝒆𝒓𝒐𝒖𝒕
For turbines: 𝜼=
𝝆𝒈𝑸𝑯

𝜼=
𝝆𝒈𝑸𝑯
For pumps:
𝒑𝒐𝒘𝒆𝒓𝒊𝒏
–1
Example. A pump lifts water from a large tank at a rate of 30 Ls . If the input power is 10 kW and
the pump is operating at an efficiency of 40%, find:
(a) the head developed across the pump;
(b) the maximum height to which it can raise water if the delivery pipe is vertical, with diameter 100
mm and friction factor λ = 0.015.
Solution:
Given; Q= (30/1000) =0.03m3/s
Powerin = 103W
𝜂 = 0.4
𝝆𝒈𝑸𝑯
(a) Using the equation 𝜼 =
𝒑𝒐𝒘𝒆𝒓𝒊𝒏
1000×9.8×0.03×𝐻
• 0.4 =
10×103
From which H=13.6m
(b) From the equation of head loss in pipes ie
𝒇𝑳𝑸𝟐
𝒉𝒇 = , f=0.015/4=0.00375
𝟑𝒅𝟓

Now the head developed by the pump should be able to overcome the frictional losses
through the pipe height while lifting the water through the same height i.e.
𝑯 = 𝒉𝒔 + 𝒉𝒇
where hs is the static lift which equals to pipe length (L) since the pipe is vertical.
Hence,
𝑳 + 𝟎. 𝟎𝟎𝟑𝟕𝟓𝑳 × (𝟎. 𝟎𝟑)𝟐
𝟏𝟑. 𝟔 =
𝟑 × (𝟎. 𝟏)𝟓
From which L=hs=12.2m
3.3 Types of Pumps and Turbines
Impulse and reaction turbines
𝒑 𝑽𝟐
In a pump or turbine, a change in fluid head (𝑯 = +𝒛+ ) may be brought about by a change in
𝝆𝒈 𝟐𝒈

pressure or velocity or both.

79
• An impulse turbine (e.g. Pelton wheel; water wheel) is one where the change in head is brought
about primarily by a change in velocity. This usually involves unconfined free jets of water (at
atmospheric pressure) impinging on moving vanes.
• A reaction turbine (e.g. Francis turbine; Kaplan turbine; windmill) is one where the change in
head is brought about primarily by a change in pressure.
Positive-Displacement and Dynamic pumps
Postive-displacement pumps operate by a change in volume; energy conversion is intermittent.
Examples in the human body include the heart (diaphragm pump) and the intestines (peristaltic pump).
In a reciprocating pump (e.g. a bicycle pump) fluid is sucked in on one part of the cycle and expelled
(at higher pressure) in another.
In dynamic pumps there is no change in volume and energy conversion is continuous. Most pumps are
rotodynamic devices where fluid energy is exchanged with the mechanical energy of a rotating element
(called a runner in turbines and an impeller in pumps), with a further conversion to or from electrical
energy. This course focuses entirely on rotary devices. Note that, for gases, pumps are usually referred
to as fans (for low pressures), blowers or compressors (for high pressures).
Radial, Axial and Mixed-Flow Devices
The terms radial and axial refer to the change in direction of flow through a rotodynamic device (pump
or turbine):

In a centrifugal pump flow enters along the axis and is expelled


radially.
(The reverse is true for a turbine.)

An axial-flow pump is like a propeller; the direction


of the flow is unchanged after passing through the
device.

80
A mixed-flow device is a hybrid device, used for
intermediate heads.
In many cases – notably in pumped-storage
power stations – a device can be run as either a
pump or a turbine.
Inward-flow reaction turbine  centrifugal
pump (high head / low discharge) (e.g. Francis
turbine) Propeller turbine  axial-flow pump
(low head / high discharge) (e.g. Kaplan turbine;
windmill)

Common Types of Turbine


Pelton wheels are impulse turbines
used in hydroelectric plant where there is a very high head of
water. Typically, 1 – 6 high-velocity jets of water impinge on
buckets mounted around the circumference of a runner.

Francis turbines are used in many large hydropower


projects (e.g the Hoover Dam), with an efficiency in
excess of 90%. Such moderate- to high-head turbines
are also used in pumped-storage power stations (e.g.
Dinorwig and Ffestiniog in Wales; Foyers in Scotland),
which pump water uphill during periods of low energy
demand and then run the system in reverse to generate
power during the day. This smooths the power demands
on fossil-fuelled and nuclear power stations which are
not easily brought in and out of operation. Francis turbines are like centrifugal pumps in reverse.
Kaplan turbines are axial-flow (propeller) turbines. In the
Kaplan design the blade angles are adjustable to ensure
efficient operation for a range of discharges.

Wells turbines were specifically developed for wave-


energy applications. They have the property that they
rotate in the same direction irrespective of the flow
direction.

81
Bulb generators are large-diameter variants of the Kaplan propeller turbine, which are suitable for the
low-head, high-discharge applications in tidal barrages (e.g. La Rance in France). Flow passes around
the bulb, which contains the alternator.
The Archimedes screw has been used since
ancient times to raise water. It is widely used in
water treatment plants because it can
accommodate submerged debris. Recently,
several devices have been installed beside weirs in
the north of England to run in reverse and generate
power.

3.4 Pump and System Characteristics


3.4.1 Pump characteristics
Pump characteristics are the head (H), input power (I) and efficiency (η) as functions of discharge (Q).
The most important is the H vs Q relationship. Typical shapes of these characteristics are sketched
below for centrifugal and axial-flow pumps.

Given the pump characteristics at one rotation rate (N), those at different rotation rates may be
determined using the hydraulic scaling laws (Section 4). Ideally, one would like to operate the pump:
• as close as possible to the design point (point of maximum efficiency);
• in a region where the H-Q relationship is steep; (otherwise there are significant fluctuations in
discharge for small changes in head).
3.4.2 System characteristics
In general, the pump has to supply enough energy to:
• lift water through a certain height – the static lift Hs;

82
• overcome losses dependent on the discharge, Q. Thus, the system head is H = Hs + hlosses
Typically, losses (whether frictional or due to pipe fittings) are proportional to Q 2, so that the system
characteristic is often quadratic: H = Hs + aQ2
The static lift is often decomposed into the rise
from sump to the level of the pump (the suction
head, Hs1) and that between the pump and the
delivery point (Hs2). The first of these is limited by
the maximum suction height (approximately 10 m,
corresponding to 1 atmosphere) and will be
discussed later in the context of cavitation.

3.4.3 Finding the Duty Point


The pump operates at a duty point where the head supplied by
the pump precisely matches the head requirements of the
system at the same discharge; i.e. where the pump and system
characteristics intersect.
In practice, it is desirable to run the pump at a speed where the
duty point is close to that of maximum efficiency. To do this
we need to determine how the pump characteristic varies with
rotation rate N.

Example.
A water pump was tested at a rotation rate of 1500 rpm. The following data was obtained. (Q is quantity
of flow, H is head of water, η is efficiency).

It is proposed to use this pump to draw water from an open sump to an elevation 5.5 m above. The
delivery pipe is 20.0 m long and 100 mm diameter, and has a friction factor of 0.005. If operating at
1500 rpm, find: (a) the maximum discharge that the pump can provide; (b) the pump efficiency at this
discharge; (c) the input power required.

Solution:
The method pf solution is to generate the system characteristic equation from H = Hs + aQ2, and then
plot it together with the pump characteristics on the same axes.
For this case,
0.005(20)
𝑎= = 3,333. ̅̅̅̅̅̅̅
3333
3(0.1)^5
Thus, the system equation is

83
̅̅̅̅̅̅̅𝑸𝟐
𝑯 = 𝟓. 𝟓 + 𝟑, 𝟑𝟑𝟑. 𝟑𝟑𝟑𝟑
pump characteristics
Q(L/s) 0 10 20 30 40 50
H(m) 10 10.5 10 8.5 6 2.5
ⴄ 0 0.4 0.64 0.72 0.64 0.4
system Characteristics, H=Hs+3,333.3333Q^2
Q(L/s) 0 10 20 30 40 50
H(m) 5.5 5.83 6.83 8.5 10.83 13.83

The plot is as shown below


pump x-tics system x-tics efficiency

16 0.8

14 0.7

12 0.6

10 0.5
H(m)

8 0.4

6 0.3

4 0.2

2 0.1

0 0
0 10 20 30 40 50 60
Q(L/s)

From the plot above, the operating point is where the system curve and the pump curve intersect from
which Q=30L/s and efficiency of 0.72
The input power is calculated from the equation
𝝆𝒈𝑸𝑯
𝜼=
𝒑𝒐𝒘𝒆𝒓𝒊𝒏
𝟏𝟎𝟎𝟎×𝟗.𝟖𝟏×𝟎.𝟎𝟑×𝟖.𝟓
𝟎. 𝟕𝟐 =
𝒑𝒐𝒘𝒆𝒓 𝒊𝒏
𝟏𝟎𝟎𝟎×𝟗.𝟖𝟏×𝟎.𝟎𝟑×𝟖.𝟓
𝒑𝒐𝒘𝒆𝒓 𝒊𝒏 = = 𝟑, 𝟒𝟕𝟒. 𝟑𝟕𝟓𝑾 ≅ 𝟑. 𝟓𝒌𝑾
𝟎.𝟕𝟐
3.5 Pumps in Parallel and in Series
Pumps in Parallel
Same head: H
Add the discharges: Q1 + Q2

84
Advantages of pumps in parallel are:
• high capacity: permits a large total discharge;
• flexibility: pumps can be brought in and out of
service if the required discharge varies;
• redundancy: pumping can continue if one is not
operating due to failure or planned maintenance.

Pumps in Series
Same discharge: Q
Add the heads: H1 + H2
Pumps in series may be necessary to generate high heads, or
provide regular “boosts” along long pipelines without large
pressures at any particular point.

Example.
A rotodynamic pump, having the characteristics
tabulated below, delivers water from a river at elevation
102 m to a reservoir with a water level of 135 m, through a 350 mm diameter cast-iron pipe. The
frictional head loss in the pipeline is given by hf = 550 Q2, where hf is the head loss in m and Q is the
discharge in m3 s –1. Minor head losses from valves and fittings amount to 50Q2 in the same units.

Pump characteristics: Q is discharge, H, is head, η is efficiency.


(a) Calculate the discharge and head in the pipeline (at the duty point). If the discharge is to be
increased by the installation of a second identical pump:
(b) determine the unregulated discharge and head produced by connecting the pump: (i) in parallel;
(ii) in series; (c) determine the power demand at the duty point in the case of parallel operation.
Solution:
for two identical pumps in parallel
Q(m3/S) 0 0.1 0.2 0.3 0.4
H(m) 60 58 52 41 25
ⴄ(%) 0 44 65 64 48
system curve, H=33+600Q^2
Q(m3/S) 0 0.1 0.2 0.3 0.4
H(m) 33 39 57 87 129

85
single pump parallel identical pumps system curve for parallel operation efficency

140 80

129
70
120
65 64
60
100

87 50
48
80
44
head(m)

40

60 60
58 57
52 30

40 41
39
20
33

25
20
10

0 0 0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45
Discharege(m^3/s)

Figure 15: System X-tics for parallel pump combination

86
series connection series system curve efficenecy single pump

140 80

70
120 120
116
65
64

104 60
100

50
head(m), efficiency(%)

48
82
80
44
40

60

30
50

40
20

20
10

0 0 0
0 0.05 0.1 0.15 0.2 0.25
Q(m3/s)

Figure 16: System x-tics for series pump combination

87
3.6 Specific speed
The specific speed (or type number) is a guide to the type of pump or turbine required for a particular
role.
Specific Speed for Pumps The specific speed, Ns, is the rotational speed needed to discharge 1 unit of
flow against 1 unit of head. (For what “unit” means in this instance, see below.) For a given pump, the
hydraulic scaling laws give
𝑄 𝑔𝐻
Π1 = 𝑁𝐷3 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡, Π2 = 𝑁2𝐷2 = 𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡
Eliminating D(and choosing an exponent that will make the combination proportional to N)
1/4
Π12 𝑄1/2 𝑁
( 3) =
Π2 (𝑔𝐻)3/4
Or since g is constant, then at any given (e.g. maximum) efficiency:
𝑄1/2 𝑁
= (𝑑𝑖𝑚𝑒𝑛𝑠𝑖𝑜𝑛𝑎𝑙)𝑐𝑜𝑛𝑠𝑡𝑎𝑛𝑡
𝐻 3/4
The constant is the specific speed, Ns, when Q and H in specific units(see below) are numerically equal to
1
3.6.1 Specific speed(pump)

𝑄1/2 𝑁
𝑁𝑆 =
𝐻 3/4

Notes.
• The specific speed is a single value calculated at the normal operating point (i.e. Q and H at the
maximum efficiency point for the anticipated rotation rate N).
• With the commonest definition (in the UK and Europe), N is in rpm, Q in m3 s –1. H in m, but
this is far from universal, so be careful.
• In principle, the units of Ns are the same as those of N, which doesn’t look correct from the
definition but only because that has been shortened from
(1 𝑚3 𝑠 −1 )1/2 𝑄1/2 𝑁
= 3/4
(1 𝑚)3/4 𝐻
• Because of the omission of g the definition of Ns depends on the units of Q and H. A less-
common (but, IMHO, more mathematically correct) quantity is the dimensionless specific speed
Kn given by
𝑄1/2 𝑁
𝐾𝑛 =
(𝑔𝐻)3/4
• If the time units are consistent then Kn has the same angular units as N (rev or rad).
• High specific speed ↔ large discharge / small head (axial-flow device).
Low specific speed ↔ small discharge / large head (centrifugal device). Approximate ranges of Ns are
(from Hamill, 2011):

88
Type Ns
Radial (centrifugal) head 10 – 70 large
Mixed flow 70 – 170 Medium head
Axial > 110 small head
Example.
A pump is needed to operate at 3000 rpm (i.e. 50 Hz) with a head of 6 m and a discharge of 0.2m 3 s –1.
By calculating the specific speed, determine what sort of pump is required.
3.6.2 Specific Speed for Turbines
For turbines the output power, P, is more important than the discharge, Q. The relevant dimensionless
groups are
𝑔𝐻 𝑃
Π2 = , Π3 =
𝑁 2 𝐷2 𝜌𝑁3 𝐷5
Eliminating D,
1/4
Π32 (𝑝/𝜌)1/2 𝑁
( 5) =
Π2 (𝑔𝐻)5/3
or, since 𝜌 and g are usually taken as constant (there does seem to be a presumption that turbines are
always operating in fresh water) then at any given efficiency:
𝑷𝟏/𝟐 𝑵
= (𝒅𝒊𝒎𝒆𝒏𝒔𝒊𝒐𝒏𝒂𝒍) 𝒄𝒐𝒏𝒔𝒕𝒂𝒏𝒕
𝑯𝟓/𝟒
The specific speed of a turbine, Ns, is the rotational speed needed to develop 1 unit of power for a head
of 1 unit. (For what “unit” means in this instance, see below.)
Specific speed(turbine)

𝑷𝟏/𝟐 𝑵
𝑵𝑺 =
𝑯𝟓/𝟒
Notes.
• With the commonest definition (in the UK and Europe), N is in rpm, P in kW (note), H in m,
but, again, this is not a universal convention. As with pumps, the units of Ns are the same as
those of N.
• As with pumps, a less-commonly-used but mathematically more acceptable quantity is the
dimensionless specific speed Kn, which retains the 𝜌 and g dependence:
(𝒑 /𝝆)𝟏/𝟐 𝑵
𝑲𝒏 =
(𝒈𝑯)𝟓/𝟒
Kn has the angular units of N (revs or radians) – see Massey (2005).
• High specific speed ↔ small head (axial-flow device)
• Low specific speed ↔ large head (centrifugal or impulse device)
Approximate ranges are (from Hamill, 2001):
Type Ns Head range
Pelton wheel (impulse) very 12 – 60 large head
Francis turbine (radial-flow) 60 – 500 large head
Kaplan turbine (axial-flow) 280 – 800 small head

89
3.7 Cavitation
Cavitation is the formation, growth and rapid collapse of vapour bubbles in flowing liquids. Bubbles
form at low pressures when the absolute pressure drops to the vapour pressure and the liquid
spontaneously boils. (Bubbles may also arise from dissolved gases coming out of solution). When the
bubbles are swept into higher-pressure regions they collapse very rapidly, with large radial velocities
and enormous short-term pressures. The problem is particularly acute at solid surfaces. Cavitation may
cause performance loss, vibration, noise, surface pitting and, occasionally, major structural damage.
Besides the inlet to pumps the phenomenon is prevalent in marine current turbines, ship and submarine
propellers and on reservoir spillways. The best way of preventing cavitation in a pump is to ensure that
the inlet (suction) pressure is not too low. The net positive suction head (NPSH) is the difference
between the pressure head at inlet and that corresponding to the vapour pressure:
𝑝𝑖𝑛𝑙𝑒𝑡 − 𝑝𝑣𝑎𝑝
𝑁𝑃𝑆𝐻 =
𝜌𝑔
The net positive suction head must be kept well above zero to allow for further pressure loss in the impeller. The
inlet pressure may be determined from Bernoulli:
𝑯𝒑𝒖𝒎𝒑 𝒊𝒏𝒍𝒆𝒕 = 𝑯𝒔𝒖𝒎𝒑 − 𝒉𝒆𝒂𝒅𝒍𝒐𝒔𝒔
𝒑𝒊𝒏𝒍𝒆𝒕 𝑽𝟐 𝒑𝒂𝒕𝒎
 𝝆𝒈
+ 𝒛𝒊𝒏𝒍𝒆𝒕 + 𝟐𝒈 = 𝝆𝒈
− 𝒉𝒇

Hence,
𝒑𝒊𝒏𝒍𝒆𝒕 𝒑𝒂𝒕𝒎 𝑽𝟐
= − 𝒛𝒊𝒏𝒍𝒆𝒕 − − 𝒉𝒇
𝝆𝒈 𝝆𝒈 𝟐𝒈

To avoid cavitation, one should aim to keep pinlet as large as possible by:
• keeping zinlet small or, better still, negative (i.e. below the level of water in the sump;
• keeping V small (large-diameter pipes);
• keeping hf small (short, large-diameter pipes).
The first also assists in pump priming.
Example:

Cavitation problems have been encountered in a mixed flow pump. The pump is sited with its intake 3
m above the water level in the reservoir and delivers 0.05 m3/s. The total head at outlet is 31.7 m, and at
inlet is −7m. Atmospheric pressure = pA = 101.4 kN/m2 and vapour pressure of water = pvap = 1.82
kN/m2. Determine the Thoma cavitation number. A similar pump is to operate at the same discharge and
head, but with pA = 93.4 kN/m2 and pvap = 1.2 kN/m2. What is the maximum height of the pump intake
above reservoir level if cavitation must be avoided? [0.081, 2.25 m]

90
CHAPTER FOUR: HYDRAULIC STRUCTURES

4.1 Introduction
Hydraulic structures are engineering structures constructed or installed for the purpose of harnessing
water resources (groundwater, surface water, and runoff). Hydraulic structures form part of most major
water engineering schemes e.g. irrigation, water supply, drainage, sewage networks and treatment,
hydropower etc. Hydraulic structures are classified as follows:
Classification purpose structure
1. Storage structures To retain or impound water Dams, reservoirs
2. Conveyance structures To transmit flow from 1 point to Penstocks, culverts, flume, open
another. channels, canals, spillways
3. Flow diversion For redirecting flow from one Weirs, coffer dams.
structure direction to another.
4. Erosion control To protect the area against Check dams, stilling basins etc.
structures/energy erosion and score.
dissipation structures.
5. Flow control/regulation To control the rate of flow from Gates, valves,
structures a given location past a given
point.

4.2 Dams
Dams may be constructed to store water for domestic and industrial use, for irrigation, to generate hydro-
electricity or to prevent flooding. Dam design and construction is a highly specialized branch of civil
engineering, necessarily so since a failure could cost thousands of lives. The type of dam constructed in

91
a particular location must be determined after considering all relevant factors such as the local geology,
shape of the valley, environmental concerns, climate, and the availability of local materials, expertise,
manpower and plant.
Types of Dams
There are only really three basic types of dam: gravity, arch, and buttress, which are further classified
according to the material from which they are constructed: concrete, masonry, rock, earth e.t.c, the
purpose.

Briefly the characteristics of these dams are as follows;


4.2.1 Gravity dams
Essentially, gravity dams are relatively simple. They are usually straight or slightly curved in plan and
rely on their own weight to resist the hydrostatic force that is trying to overturn the dam and push it

92
forward. Gravity dams can be constructed using concrete (Figs 9.1a and 9.2), masonry, or embankments
of rock or earth (earth here meaning any clay–silt–sand–gravel–rock mixture).

Elementary concrete gravity dam design


The design of a dam is a complex process, but some of the basic principles can be illustrated using a
simple rectangular concrete gravity dam. The two unavoidable forces acting on the dam are the
hydrostatic force (F) and the weight of the dam (W). If the line of the resultant of these two forces (R)
lies outside the downstream toe of the base then the dam would tip or overturn about this point. To
prevent tension cracks developing at the upstream face and potentially damaging water penetration of
the concrete, the ‘middle-third rule’ states that R should pass through the middle-third of the dam’s
base. Normally a factor of safety (about 1.3 to 4.0 according to circumstances) would be adopted so that
R lies well inside the middle-third, but Fig. 9.5a shows the limiting condition where R passes through
the very edge of the middle-third. Example 9.1 illustrates how the minimum base width is calculated.
To increase stability, generally the width of a dam increases towards its base (Fig. 9.5b). With this more
complex cross-section the middle third-rule can be applied at different levels: R1, R2 and R3 should all
fall within the middle-third of their base. The calculations are slightly more difficult, but the procedure
is similar. The uplift force on the base of the dam was ignored in Example 9.1; it can significantly reduce
the effective weight (W) of the dam.
Uplift occurs because water under pressure exerts a force at right angles to any surface it comes into
contact with, including upwards (see section 1.7). If the water underneath the dam at the upstream face
is at the full hydrostatic pressure (h) while at the downstream face the pressure is atmospheric, the
pressure distribution is as shown by the solid line in Fig. 9.5c; a high tailwater increases the pressure on
the base (dashed line). Usually the uplift pressure can be reduced to around 33–66% of the maximum
upstream value by constructing a relatively impermeable grout curtain or cut-off underneath the
upstream part of the dam. As the water seeps through the curtain there is a large head loss, so reducing
the head and uplift pressure on the base, which can be further reduced by drains drilled between the
curtain and downstream toe. These measures also reduce water seepage under the dam. Seepage from
the high to low pressure areas can cause erosion of the foundation. To help prevent a gravity dam failing
through sliding, it is usually keyed into the foundation to increase resistance. A sliding failure can occur
at foundation level, or at any level where the net horizontal force exceeds the shear resistance. We are
assuming here that the net horizontal force is due to hydrostatic pressure, but in practice horizontal
earthquake forces, ice forces and wave forces also have to be evaluated. For a horizontal plane, the shear

93
resistance is the product of the vertical force (weight) acting on the shear plane and the coefficient of
friction (f) between its two surfaces (Table 9.2).

Example:
A concrete gravity dam of rectangular cross-section is to be constructed (Fig. 9.5a). The density of the
concrete is 2350kg/m3 and that of the water 1000kg/m3. When the reservoir is full the maximum depth
of water equals the dam height of 30m. By considering a 1m length and using the middle-third rule,
determine the minimum dam width.

The hydrostatic force due to the water 𝐹 = 𝜌𝑔ℎ𝑐 𝐴


= 1000 × 9.81 × (30/2) × 30 × 1
= 4414.50 × 103 𝑁

94
Assume the width of the dam base = B, as shown in Fig. 9.5a.
The weight of the dam 𝑊 = 𝑤𝑒𝑖𝑔ℎ𝑡 𝑑𝑒𝑛𝑠𝑖𝑡𝑦 𝑜𝑓 𝑐𝑜𝑛𝑐𝑟𝑒𝑡𝑒 × 𝑣𝑜𝑙𝑢𝑚𝑒 𝑜𝑓 𝑑𝑎𝑚 𝑝𝑒𝑟 𝑚 𝑙𝑒𝑛𝑔𝑡ℎ
= 2350 × 9.81 × (30 × 𝐵 × 1)
= 691.61𝐵 × 103 𝑁
By similar triangles (𝑜𝑟 𝑡𝑎𝑛 𝜃 =) it is apparent that:
𝑊 (ℎ/3)
= where h/3=10m. W and F have the values above.
𝐹 (𝐵/6)

691.61𝐵 × 103 60
So =
4414.50 × 103 𝐵
2
691.61𝐵 = 264870
𝐵 = 19.57𝑚
Note: often B is about 2/3 of the water depth (h), so this is a good starting point for calculations.
Task: Repeat the calculations above using masonry of density 2700kg/m3 instead of concrete. By how
much does this change the minimum thickness of the dam?
4.2.2 Arch dams
Arch dams are almost always constructed from reinforced concrete, and use perhaps only 20% of the
concrete required for a gravity dam. An arch dam is rather like an arch bridge lying on its side with the
crown upstream. The strength of the arch is used to transmit the hydrostatic force to the foundations,
which must consist of rock, strong enough to withstand the high loads. Arch dams require a narrow,
steep sided valley or gorge. Typically, the length of the dam’s crest is limited to about 10 times its height.
Arch dams are technically complex, particularly thin double curvature (cupola) dams which are curved
both in plan and section. Most of the hydrostatic force is resisted by the arching action between the
abutments, and the remainder by cantilever action at the base. If an abutment fails, as at Malpasset
where the left abutment moved by as much as 2m, then the consequences can be disastrous.

4.2.3 Buttress dams


These are a hybrid of an arch and concrete gravity dam. The flat slab variety consists of a continuous
upstream face slab, either vertical or angled to increase stability, with downstream buttresses to provide
strength and support. The multiple arch type consists of a series of arches, and is used when a valley is
too wide for a single arch dam.
Typical heights being 30–90m with a flat upstream slab and 40–220m with a multiple arch.
Buttress dams use only about 60% of the concrete required for a gravity dam, but may not be any cheaper
because of the need to reinforce the concrete and to use more complex formwork. However, their
reduced weight means they can be built on weaker foundations.

95
Note: detailed design of dams and appurtenant structures will be handled under Design of Hydro-power
structures and soil and water engineering, with application of soil mechanics already covered.
4.3 Spillways
The majority of impounding reservoirs are formed as a result of construction of a dam. There are times
when the reservoir is full and the stream flow exceeds the demand. The excess water must therefore be
discharged safely from the reservoir. In most cases, to allow the water to simply overtop the dam would
result in failure of the structure. For this reason, carefully designed overflow passages are incorporated
as part of the design. The spillway must be sufficient to accommodate large flood discharges likely to
occur in the life of a dam. Because of the high velocities of flow often attained on spillways, some form
of energy dissipation and a scour prevention system at the base of the spillway must be provided for.
This often takes the form of a stilling basin. As indicated above, poor spillway design and operational
problems account for around 30% of all dam failures, so clearly this aspect of the design is crucial.

96
4.3.1Types of spillways
There are many types of spillway: overflow(ogee), chute, side-overflow (side channel), shaft, crest gates
and syphon. Which is used in any location will depend upon the type of dam (e.g. concrete or earth), its
size, the topography of the area and operational considerations.
An overflow/ogee spillway is simply part of the dam that is designed to allow water to flow over it. This
is quite common with concrete and masonry dams. With earth or rock dams, which would erode and
possibly fail if water flowed over them a special concrete spillway section must be constructed.
The discharge over the spillway can be related to the head (H) over the crest using a weir equation like
equation…, where Cw has a value between 1.7 and 2.3 depending upon H and the spillway geometry.
Sometimes overflow spillways end in a ski jump that throws the water upwards so that it lands far
enough downstream not to cause scour at the base of the dam.

The discharge relationship for a spillway is of the same form as other weirs.
3
𝑉𝑜2 2
𝑄 = 𝐶𝑤 𝐿 (ℎ + )
2𝑔
Where;
Q = discharge
Cw = coefficient
L = Length of crest
h = head on spillway (vertical distance from crest of spillway to reservoir level)
Vo = approach velocity
Hd = height of spillway related to the downstream floor level

The ideal overflow spillway should guide the water over the crest as smoothly as possible, and in cross-
section should resemble the underside of the aerated nappe. If the water lifts from the spillway a vacuum
can form resulting in cavitation damage.
A chute spillway is a steep concrete channel to take water from reservoir level down to river level.
Where the topography permits, they are often built on natural ground at the end of an earth or a rock fill
dam, since it is undesirable to locate the spillway on the dam itself, but chutes may be adopted with any
type of dam. The flow in the chute is supercritical, the high velocity resulting in the entrainment of air
and bulking (i.e. expansion vertically) of the water. Either to avoid building on earth or rock dams or for

97
other reasons such as topography, sometimes the water may enter a chute spillway via a side-overflow
weir located perpendicular to the upstream face of the dam. In narrow valleys, a side overflow weir may
be used to allow a chute spillway to be constructed along the downstream face of the dam.

A bell mouth overflow and shaft spillway consist essentially of a funnel entrance of fixed elevation
and a vertical shaft down which the overflow from the reservoir falls. The vertical shaft leads to a gradual
90° bend and a horizontal tunnel that passes around or under the dam and hence to the downstream river
channel. These spillways can be used with any type of dam, being particularly useful where space is
restricted and with earth or rock dams that require a separate spillway structure. They are also called
drop inlets and morning glory spillways. Provided the shaft diameter is large enough to take the flow,
the discharge over the bell mouth lip can be calculated by assuming a straight weir of the same length.
At larger discharges the tunnel itself may provide the flow control, as in pipe flow. Problems can arise
due to cavitation, while the bell mouth must be provided with some sort of screen to prevent logs or
other debris becoming wedged in the shaft.

98
Syphons spillways are useful where automatic action is desirable, the required discharge is not large,
space is restricted, it is necessary to increase the capacity of an existing spillway, or where it is necessary
to keep the reservoir level within a narrow range. To increase the range, syphons may be set with crests
at different levels. Syphons are not suitable where they are likely to be blocked by ice in winter. Design
guidelines are vague, so they are often the subject of model tests.

Once primed, the discharge can be calculated using the small orifice equation, where CD has a value of
about 0.9 and h is the vertical difference between the reservoir level and tailwater level. Since the
discharge depends upon h, which may vary relatively little, syphons tend to give a fairly constant
discharge regardless of reservoir level. As with all syphons, flow will cease if the pressure at the crest
falls below about -7.5m of water (or 2.8m of water absolute pressure). Thus 7.5m is the maximum
elevation of the syphon crest above the hydraulic grade line (i.e. total head line minus the velocity head).

99
4.4 Culverts
A culvert is a hydraulically short conduit which conveys stream flow through an embankment or past
some other type of flow obstruction. Culverts are constructed from a variety of materials and are
available in many different shapes and configurations.
Numerous cross-sectional shapes are available. The most commonly used shapes, depicted in Figure 5-
3, include circular, box (rectangular), elliptical, pipe-arch, and arch. The shape selection is based on the
cost of construction, the limitation on upstream water surface elevation, roadway embankment height,
and hydraulic performance.
The essential features of a culvert are the barrel which passes under the fill, the headwalls and wing
walls and the end walls or other devices at exit which improve flow conditions and prevent embankment
score; and in some cases, debris protection to prevent entrance of debris which might clog the barrel.

Culverts are made of a variety of materials, the main bases of selection being; the cost of installation
and maintenance, durability of the material and the size of the culvert
Small culverts may be made from pre-cast concrete, vitrified clay, cast iron or corrugated steel pipe. In
the larger size, multi-plate corrugated steel arches, reinforced concrete arches, or concrete box culverts
are usually selected.

100
Minimum diameter for culverts should be between 45cm to 60cm, though 30cm pipe may also be used
for small flows carrying no debris or trash. Whenever possible, the culvert barrel should follow the line
and grade of the natural channel.
It is common to place the culvert axis normal to the embankment centerline even though this requires
some changes in the natural channel.
The alternative is skew culvert which will be longer than the normal/straight culvert and will require
more complex construction of headwalls and end walls.

However, when the skew of the stream is large, a culvert normal to the alignment of the embankment
centerline will create bends in the channel which will be points of erosion.
Moreover, the hydraulic capacity of the culvert will be decreased by the poor entrance and exit
conditions. As a general rule, it is best that the culvert alignment conforms to the natural stream
alignment.
4.4.1 Types of flow in culverts

101
There are many ways to classify the flow through culverts. Several or all of the flow conditions may be
experienced at one culvert as the discharge and water level increases during a flood, and then falls. There
are many different flow classifications and design guidelines, such as those given by Chow (1981),
American Iron and Steel Institute (1984), French (1986) and CIRIA (1997), but in broad terms the types
of flow above can be listed in four categories:
1. channel flow with unsubmerged inlet (types 1, 3 and 4);
2. submerged inlet, but barrel only part full (type 2);
3. submerged inlet, barrel full, but free discharge at the outlet (type 5);
4. submerged inlet and outlet (type 6).

102
Culverts are often described as operating under inlet control or outlet control. The control point (CP)
has the lowest discharge capacity. Thus, the term ‘inlet control’ means simply that the inlet has a
lower discharge capacity than either the culvert barrel or the outlet, so the inlet is limiting the discharge
through the culvert (flow types 1 and 2). Inlet control tends to occur with larger culverts on relatively
steep gradients where generally the outlet will flow freely and will not submerge. Often the control is
where the velocity increases and critical depth occurs. Remember that DC increases inside a width

103
constriction (see Fig. 8.18 and equation (8.32)) so if the culvert is narrower than the upstream channel
it is possible that the flow will pass through critical depth near the entrance establishing inlet control.
With critical flow established, the characteristics of the channel downstream of the control do not affect
the flow upstream. Thus, an example of inlet control is when the flow passes through critical depth as it
enters the barrel, and then for the remainder of the culvert the flow is supercritical so that the partially-
full barrel can cope comfortably with any discharge that passes through the inlet. A variation on type 1
and 2 flow is that a hydraulic jump occurs in the barrel and the flow returns to subcritical before the exit.
When operating in inlet control the headwater depth is determined by the flow rate (Qm3 /s), the cross-
sectional area of the barrel (ABm2) and the shape (i.e. hydraulic efficiency) of the inlet. In this condition
the length and roughness of the barrel and the outlet conditions do not matter. When Q is small and the
entrance unsubmerged the culvert behaves like an open channel, and can be analyzed as such. Culverts
operating with inlet control are also sometimes described as being hydraulically short or hydraulically
long. With the inlet submerged, the flow contracts as it enters the culvert so that the barrel is initially
running part full, and then gradually expands again as friction slows the flow. The culvert is
hydraulically short if the flow exits the culvert before having expanded to fill the barrel; such a culvert
will never flow full like a pipe. The culvert is hydraulically long if the barrel is full when the exit is
reached. Generally, culverts operating in outlet control will be found where the bed slope is relatively
flat. In outlet control, it is the outlet which restricts the discharge through the culvert. The critical factors
are the tailwater level in the outlet channel (H2), and the slope, roughness and length of the barrel. If H2
is high as a result of an obstruction in the river channel further downstream, this affects the discharge
through the barrel. Usually the barrel will tend to run full over part or all of its length, and at the design
discharge will probably have a fully submerged inlet and quite possibly a submerged outlet. In this
condition an increase in tailwater level will produce a corresponding increase in headwater level in order
to maintain the required differential head. This condition is undesirable for several reasons: it results in
a relatively large head loss and increases the risk of upstream flooding and property damage (as for a
bridge, Fig. 9.16); in severe storms the level and quantity of water upstream of the inlet could threaten
the safety of the embankment; blockage by floating debris significantly exacerbates these problems;
extensive downstream damage may result if the barrel at exit is flowing full and blasting water into the
downstream channel. In the latter case an impact stilling basin (or similar) may be advisable.
4.4.2 Culvert analysis and design
When operating as an open channel, the techniques relating to uniform and non-uniform flow can be
used to determine the profile of the water surface in the culvert. If the flow is subcritical, then the

104
downstream conditions would determine the water level upstream. If the flow passes through critical,
then the conditions downstream of this point do not matter. The entrance to a culvert may become
submerged when the headwater depth H1 = 1.2Y, but this cannot be guaranteed until H1 > 1.5Y. When
the inlet is submerged but the exit is free (type 2 flow) this resembles a free small orifice so:
𝑸 = 𝑪𝒅 𝑨𝑩 [𝟐𝒈𝑯𝟏 ]𝟏/𝟐
where Cd is a dimensionless coefficient of discharge and AB is the cross-sectional area of the barrel (m2).
For pipe or box culverts set flush in a vertical headwall, the value of Cd is about 0.44 when H1/Y = 1.4
rising to 0.51 when H1/Y = 2.0 and 0.59 when H1/Y = 5.0. These values are applicable to pipes with or
without wingwalls. For box culverts with 45° wingwalls and a square edged soffit, Cd is about 0.44 when
H1/Y = 1.3 rising to 0.53 when H1/Y = 2.0 and 0.62 when H1/Y = 5.0 (French, 1986). Alternatively:
𝑸 = 𝑪𝒅 𝑨𝑩 [𝟐𝒈𝑯𝟏 − 𝒀/𝟐]𝟏/𝟐
where (H1 - Y/2) is the headwater depth measured above the center of the orifice. For circular and pipe-
arch culverts CD has a value ranging from 0.62 for square edged inlet structures to 1.0 for well- rounded
ones (ARMCO, undated). If the inlet is submerged and the barrel is full (type 5) or the outlet is
submerged (type 6) then the flow can be analyzed using the methods for reservoir–pipeline problems
i.e. discharge to the atmosphere or flow between two reservoirs respectively. In such cases the headwater
depth (H1) is determined by the tailwater level (H2) and the head loss through the culvert as shown in
Fig. 9.24. They can be analyzed by applying the energy equation to an upstream and downstream section:
𝟒/𝟑
𝑺𝑩 𝑳 + 𝑯𝟏 + 𝑽𝟐𝟏 /𝟐𝒈 = 𝑯𝟐 + 𝑲𝑬 𝑽𝟐𝟐 /𝟐𝒈 + 𝑽𝟐𝟐 𝒏𝟐𝑩 /𝑹𝑩 + 𝑽𝟐𝟐 /𝟐𝒈
If the approach velocity heads 𝑉12 /2𝑔 = 0 due to the ponding of water upstream, then SBL + H1 is the
headwater level (HWL) above the mean bed level at the culvert outlet (Fig. 9.24). If 𝑉2 = 𝑉2 then
equation (9.13) becomes:
𝟒/𝟑
𝑯𝑾𝑳 = 𝑯𝟐 + 𝑽𝟐𝑩 /𝟐𝒈 + 𝑲𝑬 𝑽𝟐𝑩 /𝟐𝒈 + 𝑽𝟐𝑩 𝒏𝟐𝑩 /𝑹𝑩
With a bevelled ring entrance, KE = 0.25; if the barrel projects from the fill with no headwalls, KE can
be as high as 0.90, but typically if the end of the barrel matches the fill slope or has square edged
wingwalls or headwalls then KE = 0.50 so:
𝟒/𝟑
𝑯𝑾𝑳 = 𝑯𝟐 + 𝟏. 𝟓𝑽𝟐𝑩 /𝟐𝒈 + 𝑽𝟐𝑩 𝒏𝟐𝑩 /𝑹𝑩
so headwater elevation above datum = z2 + HWL
and H1 = HWL - SBL (9.17). These equations can be applied in both submerged and open channel flow.
However, if the tailwater level (H2) is below the top of the culvert outlet, when calculating the headwater
level, the total head loss should be added to the larger of H2 or 0.5 (yC + Y) where yC is the critical depth
at the flow rate in question. Note that the headwater elevation can be minimized by using a rounded
entrance, a good alignment, a smooth barrel and effective exit.
Example:

105
A single barrel, rectangular culvert has to be designed for a river that has a 1 in 10year flood flow of
12.90m3 /s and a 1 in 100year discharge of 23.00m3/s. The maximum permissible upstream flood level
is 78.60m above Ordnance Datum (mOD). The length of the culvert barrel is 45m and its design
freeboard is 0.60m. The mean bed level at the outlet is 74.80mOD, and 60m upstream it is 74.90mOD.
During a 1 in 10year flood the downstream channel has a tailwater depth of 2.10m and a surface width
(BS) of 5.40m. During a 100-year flood the tailwater depth is 3.06m. The river has a bed of gravel and
some stones averaging around 120mm diameter. Determine a suitable size and slope for the culvert.
Solution
The approach adopted will be to make the culvert match as closely as possible the size, shape and slope
of the natural channel during the 1 in 10year flood and to design the culvert for type 3 subcritical channel
flow. Then a check of what happens during the 1 in 100year event will be undertaken.
1 in 10year flood, Q = 12.90 m3/s
Say maximum permissible upstream flood level = 78.60 - 0.30m freeboard = 78.30mOD. Natural slope
of stream, SO = (74.90 - 74.80)/60 = 1 in 600.
Make slope of culvert barrel SB = SO = 1 in 600.
Length of culvert = 45m. Say invert level of entrance = 74.80 + 45 × (1/600) = 74.88𝑚𝑂𝐷. Say
width of culvert barrel = bankful width BS = 5.40m. Assume depth of flow in culvert barrel, DB = 2.10m.
Velocity of flow in barrel, VB = 12.90/ (5.40 × 2.10) = 1.14m/s (0.75 < 1.14 < 2.00m/s, so OK).
Make culvert height Y = tailwater depth + culvert freeboard = 2.10 + 0.60 = 2.70m.
Use a rectangular concrete box section 5.40m wide × 2.70m high.
From equation of critical flow, critical depth in the rectangular culvert is: 𝑫𝑪 =
𝟐 𝟐 𝟏/𝟑
(𝑸 /𝒈𝑩 ) = (𝟏𝟐. 𝟗𝟎𝟐 /𝟗. 𝟖𝟏 × 𝟓. 𝟒𝟐 ) 𝟏/𝟑 = 𝟎. 𝟖𝟑𝒎.
Check tailwater level: 𝟎. 𝟓(𝑫𝑪 + 𝒀) = 𝟎. 𝟓(𝟎. 𝟖𝟑 + 𝟐. 𝟕𝟎) = 𝟏. 𝟕𝟕𝒎 (i.e. < 2.10m), so use 2.10m.
Allow for the possibility of giving the culvert a bed of gravel and stones or for the natural transport of
such material into the culvert, so say the composite bed/concrete roughness of the barrel is nB = 0.030
𝑹𝑩 = 𝑨𝑩 /𝑷𝑩 = (𝟐. 𝟏𝟎 × 𝟓. 𝟒𝟎)/(𝟓. 𝟒𝟎 + 𝟐 × 𝟐. 𝟏𝟎) = 𝟏. 𝟏𝟖𝒎.
Estimate head water level from
𝟒/𝟑
𝑯𝑾𝑳 = 𝑯𝟐 + 𝟏. 𝟓𝑽𝟐𝑩 /𝟐𝒈 + 𝑽𝟐𝑩 𝒏𝟐𝑩 /𝑹𝑩
𝟏𝟒𝟐
= 𝟐. 𝟏𝟎 + 𝟏. 𝟓 × 𝟏. 𝟏𝟗 . 𝟔𝟐 + 𝟏. 𝟏𝟒𝟐 × 𝟎. 𝟎𝟑𝟐 × 𝟒𝟓/𝟏. 𝟏𝟖𝟑/𝟒
= 𝟐. 𝟏𝟎 + 𝟎. 𝟏𝟎 + 𝟎. 𝟎𝟒
= 𝟐. 𝟐𝟒𝒎
Therefore, headwater elevation = 74.80 + 2.24 = 77.04mOD (< 78.30mOD maximum, so OK). This
method tends to overdesign, and a smaller culvert could be possible. It also assumes that the depth of

106
flow in the culvert is the same as in downstream channel. Check the actual flow depth in the barrel
(assuming uniform flow) using the Manning equation:
𝑨𝑩 𝟐/𝟑 𝟏/𝟐
𝑸= 𝑹 𝑺
𝒏𝑩 𝑩 𝑩
𝟓. 𝟒𝟎𝑫𝑩 𝟓. 𝟒𝟎𝑫𝑩 𝟐/𝟑 𝟏 𝟏/𝟐
𝟏𝟐. 𝟗𝟎 = [ ] ( )
𝟎. 𝟎𝟑 𝟓. 𝟒𝟎 + 𝟐𝑫𝑩 𝟔𝟎𝟎
𝟐/𝟑
𝟓. 𝟒𝟎𝑫𝑩
𝟏. 𝟕𝟔 = 𝑫𝑩 [ ]
𝟓. 𝟒𝟎 + 𝟐𝑫𝑩
By trial and error DB = 1.71m (note that this would give a higher velocity and higher losses). Thus DB
= 1.71m > DC (0.83m) so type 4 subcritical flow occurs, probably with a depth DB between 1.71 and
2.10m. It is unlikely that the culvert’s inlet will become submerged (see below)
1 in 100year flood, Q = 23.00 m3 /s
The tailwater depth = 3.06m > 2.70m height of the barrel (Y), so it is possible the barrel will be full with
type 6 flow and outlet control (i.e. the control is in the downstream channel). With a full barrel VB =
23.00/(5.40× 2.70) = 1.58m/s and RB = (5.40 × 2.70)/(2 × 5.40 + 2 × 2.70) = 0.90m.
𝟒/𝟑
𝑯𝑾𝑳 = 𝑯𝟐 + 𝟏. 𝟓𝑽𝟐𝑩 /𝟐𝒈 + 𝑽𝟐𝑩 𝒏𝟐𝑩 /𝑹𝑩
𝟓𝟖𝟐
= 𝟑. 𝟎𝟔 + 𝟏. 𝟓 × 𝟏. + 𝟏. 𝟓𝟖𝟐 × 𝟎. 𝟎𝟑𝟐 × 𝟒𝟓/𝟎. 𝟗𝟎𝟒/𝟑
𝟐𝒈
= 𝟑. 𝟎𝟔 + 𝟎. 𝟏𝟗 + 𝟎. 𝟏𝟐
= 𝟑. 𝟑𝟕𝒎
Therefore, headwater elevation = 74.80 + 3.37 = 78.17mOD (< 78.30mOD maximum, so OK). From
equation (9.17), H1 = HWL - SBL = 3.37 - (1/600) ×45 = 3.23m. The upstream depth ratio H1/Y =
3.23/2.70 = 1.20, so it is just possible that the entrance may submerge. Check the discharge capacity of
the entrance using an orifice equation. Assuming the entrance to the box culvert has 45° wingwalls and
a square top edge, for equation (9.10) the value of Cd is about 0.44 when H1/Y = 1.3 so assuming
submergence has occurred:
𝑸 = 𝑪𝒅 𝑨𝑩 [𝟐𝒈𝑯𝟏 ]𝟏/𝟐 = 𝟎. 𝟒𝟒 × 𝟓. 𝟒 × 𝟐. 𝟕[𝟏𝟗. 𝟔𝟐 × 𝟑. 𝟐𝟑]𝟏/𝟐 = 𝟓𝟏. 𝟎𝟕𝒎𝟑 /𝒔
Alternatively, using
𝑸 = 𝑪𝒅 𝑨𝑩 [𝟐𝒈𝑯𝟏 − 𝒀/𝟐]𝟏/𝟐 as a check,
𝑸 = 𝟎. 𝟔𝟐 × 𝟓. 𝟒 × 𝟐. 𝟕[𝟏𝟗. 𝟔𝟐 × (𝟑. 𝟐𝟑 − 𝟐. 𝟕/𝟐)]𝟏/𝟐 = 𝟓𝟒. 𝟗𝟎𝒎𝟑 /𝒔

Thus, both equations indicate that at the values of H1/Y required for submergence the capacity of the
inlet far exceeds the actual discharge, so the culvert is not operating under inlet control and must be in
outlet control. Note if a higher headwater level and inlet control was acceptable then the culvert could
be made smaller.
4.5 Flow measuring structures
4.5.1 Weirs
The primary purpose of a weir is to measure discharge.
Rectangular Sharp crested weirs

107
The simplest method of developing a numerical model which represents a weir is to use the Bernoulli
equation as the starting point applied along one streamline (due to uneven distribution of volume)
Total energy of streamline A-A at 1
𝑃1 𝑈12 𝑈12
𝐻1 = + 𝑍1 + = 𝑦1 +
𝜌𝑔 2𝑔 2𝑔
At section 2, the liquid passes over the weir and forms an over spilling jet whose lower nappe is exposed
to the atmosphere.
𝑃 𝑈2
Total energy at 2 𝐻2 = 𝜌𝑔2 + 𝑍2 + 2𝑔2 ; 𝑃2 = 0
Assumptions
• Velocities upstream are uniform and steady therefore pressure there varies according to the
hydrostatic equation
• Pressure throughout nappe is atmospheric therefore P2 = 0
• Effects of viscosity and surface tension are negligible
From the energy principle H1 = H2
1⁄
𝑈12 𝑈22 𝑈12 2
𝑦1 + = 𝑍2 + ; 𝑈2 = [2𝑔 (𝑦1 + − 𝑍2 )]
2𝑔 2𝑔 2𝑔
The ideal discharge through an elemental strip of width b and depth δZ is;
1⁄
𝑈12 2
δQideal = √2𝑔 (𝑦1 + − 𝑍2 ) (𝑏𝛿𝑍)
2𝑔
Assumptions
i. Elevation of water surface at section 2 is the same as that at1 (elevation of free surface is horizontal)
ii. Datum is raised to crest of weir
Thus integrating to obtain Q at weir
ℎ1

𝑄𝑖𝑑𝑒𝑎𝑙 = ∫ 𝑑𝑄𝑖𝑑𝑒𝑎𝑙 = ∫ 𝑈2 𝑏𝑑𝑍


0
ℎ1 1⁄
𝑈12 2
= ∫ √2𝑔 (ℎ1 + − 𝑍2 ) (𝑏𝑑𝑍2 )
2𝑔
0
Integrating with respect to Z2

108
3⁄ 3⁄
2 𝑈12 2 𝑈12 2
𝑄𝑖𝑑𝑒𝑎𝑙 = 𝑏 √2𝑔 [(ℎ1 + ) −( ) ]
3 2𝑔 2𝑔
𝑈12
For practical purposes h1 >>>> therefore equation 1.4 reduces to:
2𝑔
2 3⁄
𝑏√2𝑔ℎ1 2
𝑄𝑖𝑑𝑒𝑎𝑙 =
3
The streamlines approaching the weir converge downstream of the weir. This causes a contraction or
venacontracta which implies that actual discharge is less than Qideal and account must be taken of this
by the coefficient of discharge Cd. Hence;
2 3
𝑄𝑎𝑐𝑡𝑢𝑎𝑙 = 𝐶𝑑 √2𝑔𝑏ℎ ⁄2
3
Triangular / Vee – Notch weir (sharp crested)
Rectangular weirs have the disadvantage of loss of accuracy at low flows. The vee – notch weir
overcomes this problem since b varies with height and therefore offers greater sensitivity. However, the
vee – notch weir is inappropriate for large flows because its construction would be very uneconomical.
The theory and assumptions for the vee – notch weir is the same as for the rectangular sharp crested
weir.
1⁄
𝑈2 2
δQideal = √2𝑔 (𝑦1 + 2𝑔1 − 𝑍2 ) (𝑏𝛿𝑍)
b

b is not constant
δZ
b = f (θ, Z)
𝜃
b = 2 Z2 tan ( )
Z 2
θ ℎ 𝑈2
Integrating; 𝑄𝑖𝑑𝑒𝑎𝑙 = ∫0 1 √2𝑔 (ℎ1 + 1 −
2𝑔
1⁄
2 𝜃
𝑍2 ) 2𝑍2 𝑡𝑎𝑛 ( ) 𝑑𝑍
2
The approach velocity, U1 is almost always negligible for vee – notch weirs in view of the smaller
discharges for which they are designed. Therefore:
ℎ1
1⁄ 𝜃
𝑄𝑖𝑑𝑒𝑎𝑙 = ∫ √2𝑔(ℎ1 − 𝑍2 ) 2 2 𝑍2 𝑡𝑎𝑛 ( ) 𝑑𝑍
2
0
8 𝜃 5⁄
𝑄𝑖𝑑𝑒𝑎𝑙 = √2𝑔 𝑡𝑎𝑛 ( ) ℎ1 2
15 2
8 𝜃 5⁄
2
From which; 𝑄𝑎𝑐𝑡𝑢𝑎𝑙 = 𝐶 2𝑔
15 𝑑 √
𝑡𝑎𝑛 (2) ℎ1
Broad crested Weirs
The broad crested weir is an obstacle and the water upstream of the weir needs to gather just enough
specific energy to overcome the obstacle. Thus, given a sufficient weir height ∆Z the flow over the weir
will be critical.

109
Broad crested weirs are robust structures that are generally constructed from reinforced concrete and
which usually span the full width of the channel. By virtue of being a critical depth meter, the broad
crested weir has the advantage that it operates effectively with higher downstream water levels than a
sharp crested weir. Mostly rectangular broad crested weirs will be considered below, although there are
a variety of possible shapes: triangular, trapezoidal and round crested all being quite common.

When the length, L, of the crest is greater than about three times the upstream head, the weir is broad
enough for the flow to pass through critical depth somewhere near to its downstream edge.
Consequently, this makes the calculation of the discharge relatively straightforward. Applying the
continuity equation to the section on the weir crest where the flow is at critical depth gives: Q = ACVC.
Now assuming that the breadth of the weir (b) spans the full width (B) of the channel and that the cross-
sectional area of flow is rectangular, then:
𝐴𝐶 = 𝑏𝐷𝑐 𝑎𝑛𝑑 𝑉𝑐 = (𝑔𝐷𝑐)1/2
𝑄 = 𝑏𝐷𝑐(𝑔𝐷𝑐)1/2
𝑄 = √𝑔𝑏𝐷𝐶 3/2
Using the weir crest as the datum level, and assuming no loss of energy, the specific energy at an
upstream section (subscript 1, Fig. 9.26) equals that at the critical section:
𝐻1 + 𝑉12 /2𝑔 = 𝐷𝐶 + 𝑉𝐶2 /2𝑔 𝑤ℎ𝑒𝑟𝑒 𝐷𝐶 + 𝑉𝐶2 /2𝑔 = 𝐸𝐶
Therefore 𝐸𝐶 = 𝐻1 + 𝑉12 /2𝑔
2 2
𝐷𝐶 = 𝐸 𝑠𝑜 𝐷𝐶 = (𝐻1 + 𝑉12 /2𝑔)
3 𝐶 3
Substituting this expression into equation 9.18 gives.
2 3/2
𝑄 = √𝑔𝑏[𝐻1 + 𝑉12 /2𝑔]3/2 = (9.81)1/2 ( ) 𝑏(𝐻1 + 𝑉12 /2𝑔)3/2
3
𝑄 = 1.705𝑏(𝐻1 + 𝑉12 /2𝑔)3/2

110
The term 𝑉12 /2𝑔 in the above equation is the velocity head of the approaching flow. As with the
rectangular sharp crested weir, the problem arises that the velocity of approach, V1, cannot be calculated
until Q is known, and Q cannot be calculated until V1 is known. Therefore, a coefficient of discharge,
C, can be introduced into the equation to allow for the velocity of approach, non-parallel streamlines
over the crest, and energy losses. C varies between about 1.4 and 2.1 according to the shape of the weir
and the discharge, but frequently has a value of about 1.6. Thus:
𝟑/𝟐
𝑸 = 𝑪𝒃𝑯𝟏
The broad crested weir will cease to operate according to the above equations if a backwater from further
downstream causes the weir to submerge.
Minimum height of a broad crested weir

So how can we work out the optimum height for the weir? What height will give supercritical flow
without unduly raising the upstream water level? The answer is obtained by applying the energy equation
to two sections (Fig. 9.27), one some distance upstream of the weir (subscript 1) and the second on the
weir crest where critical depth occurs (subscript c). In this case the bottom of the channel is used as the
datum level. Assuming that the channel is horizontal over this relatively short distance, that both cross-
sectional areas of flow are rectangular, and that there is no loss of energy, then:
𝑉12 /2𝑔 + 𝐷1 = 𝑉𝑐2 /2𝑔 + 𝐷𝐶 + 𝑝
Where 𝑉1 = 𝑄/𝐴1 , 𝐷𝐶 = (𝑄 2 /2𝐵 2 )1/3 𝑎𝑛𝑑 𝑉 = (𝑔𝐷𝐶 )1/2 . This is usually sufficient to enable equation
(9.21) to be solved for p when Q and D1 are known (Example 9.6). Alternatively, the depth, D1, upstream
of the weir can be calculated if Q and p are known (Example 9.7). When calculating the ‘ideal’ height
of weir, it must be appreciated that it is only ideal for the design discharge. The weir cannot adjust its
height to suit the flow, so at low flows it may be too high, and at high flows it may be too low.
Consequently ‘V’ shaped concrete weirs are often used, or compound crump weirs that have crests set
at different levels.

111
Example:
Water flows along a rectangular channel at a depth of 1.3m when the discharge is 8.74m3 /s. The channel
width (B) is 5.5m, the same as the weir (b). Ignoring energy losses, what is the minimum height (p) of
a rectangular broad crested weir if it is to function with critical depth on the crest?
𝑉 = 𝑄/𝐴 = 8.74/(1.3 × 5.5) = 1.222𝑚/𝑠
𝑦𝑐 = (𝑄 2 /𝑔𝐵 2 )1/3 = (8.742 /9.81 × 5.52 )1/3 = 0.636𝑚
𝑉𝑐 = (𝑔𝑦𝑐 )1/2 = (9.81 × 0.636)1/2 = 2.498𝑚/𝑠
Substituting these values into equation (9.21) and the solving for p gives:
1.2222 /19.62 + 1.300 = 2.4982 /19.62 + 0.636 + 𝑝
0.076 + 1.300 = 0.318 + 0.636 + 𝑝
𝑝 = 0.42𝑚
The weir should have a height of 0.42m measured from the bed level.
Example:
Water flows over a broad crested weir 0.5m high that completely spans a rectangular channel 10.0m
wide (b = B). When the discharge is 19.0m3/s, estimate the depth of flow upstream of the weir. Assume
no loss of energy and that critical depth occurs on the weir crest.
𝑦𝑐 = (𝑄 2 /𝑔𝐵 2 )1/3 = (19.02 /9.81 × 10.02 )1/3 = 0.717𝑚
𝑉𝑐 = (𝑔𝑦𝑐 )1/2 = (9.81 × 0.717)1/2 = 2.652𝑚/𝑠
Substitution of these values into equation (9.21) and the fact that 𝑉1 = 𝑄/𝐵𝑦1 gives:
𝑉12 /2𝑔 + 𝑦1 = 2.6522 /19.62 + 0.717 + 0.500
𝑄 2 /2𝑔𝐵 2 𝑦1 + 𝑦1 = 0.358 + 0.717 + 0.500
[19.02 /(19.62 × 10.02 × 𝑦12 )] + 𝑦1 = 1.575
[0.184/𝑦12 ] + 𝑦1 = 1.575
𝑦1 has to be found by trial and error but is often possible to make a reasonably accurate first guess
because the upstream velocity head is usually small. So, to begin with guess 𝑦1 = 1.55𝑚 and evaluate
the LHS of the equation, then adjust 𝑦1 until LHS and the RHS agree.
Try 𝑦1 = LHS= RHS=
1.55 1.627 1.575
1.50 1.582 1.575
1.48 1.564 1.575
1.49 1.573 1.575
The water upstream of the weir is approximately 1.49m deep.

112
4.5.2 Venturi Flume
With small open channels a throated flume may prove a better alternative than a weir, and they have
been used successfully to measure relatively large flows. The throated flume is basically a width
constriction that, when seen in plan, has a shape similar to a Venturi meter. Flumes are usually designed
to achieve critical flow in the narrowest section (throat). Flumes are especially applicable where
depositions of solids must be avoided e.g. in sewage works and irrigation canals. Deposition at weirs
results in gradual change of the weir coefficients. Additionally, the use of weirs results in a relatively
large head loss.
Advantages of the flume include:
(a) The obstacle to the flow is relatively small so there is little afflux or backwater (that is increase
in the upstream water level), which is an asset where the channel has little freeboard, or has a
very small slope.
(b) ) It is easily constructed and very robust, since there is little to damage.
(c) Easy maintenance, since there is unlikely to be any siltation, and there is little to trap floating
debris. Consequently, flumes are often used in sewage treatment works.
(d) Like a Venturi meter, there is little loss of energy when water flows through a flume, much less
than with a weir

Head–discharge relationship
Under normal operating conditions, throated flumes are designed so that the flow passes through
critical depth in the throat, with a weak hydraulic jump forming in the diverging.

113
Thus, they are sometimes called ‘standing wave flumes’ as well as ‘Venturi flumes.’ Being a critical
depth meter, these flumes have the advantage that all of the equations applicable to critical flow in
a rectangular channel can be applied to derive the head–discharge equation. The procedure is the
same as for the broad crested weir, and the result identical:
𝑸 = 𝟏. 𝟕𝟎𝟓𝒃𝑪 (𝑫𝟏 + 𝑽𝟐𝟏 /𝟐𝒈)𝟑/𝟐
where bC is the width of the throat where critical flow occurs (which is not the same as the full
channel width, B) and D1 is the depth of water above the flat bed of the flume (instead of the height
of water above the weir crest, H1). As with the broad crested weir, a coefficient of discharge should
be introduced to allow for energy losses, and the fact that the velocity of approach, V1, is often
assumed to be negligible, so:
𝟑/𝟐
𝑸 = 𝑪𝒃𝑪 𝑫𝟏
where C has a value of about 1.65, slightly higher than the coefficient for a broad crested weir.
Design of throated flumes
The design of these flumes requires some degree of compromise between ensuring that the throat is
narrow enough to control the flow and prevent submergence, but not so narrow as to cause excessive
afflux. The equations and principles presented earlier can be used as the basis of a simple design
procedure, as in the following example.
Example: Flume design
A throated flume is to be built on a uniform man-made rectangular channel like that in Fig. 9.28.
The flow in the channel is maintained at about 0.3m3 /s with a normal depth of around 0.35m. The
freeboard of the channel is very limited, so the afflux should not exceed 0.2m otherwise overtopping
and scouring of the banks will result. Determine a suitable throat width for a flatbed flume. Assume
a modular limit of 0.75 and a coefficient C of 1.65.

4.6 Sluice gates and other control gates


Sluice gates like the one in Fig. 1.18 are used to control the flow in rivers and man-made open
channels. They are sometimes referred to as underflow gates, since the flow passes under the bottom
edge of the gate, as shown diagrammatically in Fig. 9.12 (a and b). Once calibrated, either by
114
measurements in the field or model tests, they can also be used to measure the discharge. There are
similarities between the discharge through an orifice and under a sluice gate, but also important
differences:

The analysis of flow under a sluice gate is based on the energy equation applied to cross-section 1
upstream of the gate and section 2 at the vena contracta, as in Fig. 9.12a. Taking the head (depth) at
section 1 to be H1, as for an orifice, and the downstream depth as H2, then with no loss of energy
head:
𝐻1 + 𝛼1 𝑉12 /2𝑔 = 𝐻2 + 𝛼2 𝑉22 /2𝑔
So
𝛼2 𝑉22 /2𝑔 = 𝐻1 + 𝛼1 𝑉12 /2𝑔 − 𝐻2
1/2
2𝑔 𝛼1 𝑉12
𝑉2 = [ (𝐻1 + − 𝐻2 )]
𝛼2 2𝑔
The theoretical discharge under the sluice gate is QT = aOV2 where aO is the area of the opening; aO =
bY where b is the width of the gate across the channel and Y the height of the opening. The actual
discharge can be obtained by introducing the coefficient of discharge CD = CC×CV. The value of the
coefficient of contraction (CC) depends upon the shape of the gate and its relative height from the bed,
but often has a value of around 0.6. The coefficient of velocity (CV) has a value just less than unity.
Thus, the actual discharge QA = CDaOV2 is given by:

115
𝟏/𝟐
𝟐𝒈 𝜶𝟏 𝑽𝟐𝟏
𝑸 𝑨 = 𝑪 𝑫 𝒂𝟎 [ ( 𝑯 𝟏 + − 𝑯𝟐 )]
𝜶𝟐 𝟐𝒈
In free flow it is often assumed that H2 = CC Y. If an overall sluice gate coefficient (C) is introduced
that incorporates CD, a2, the velocity of approach and the depth H2 then equation (9.1) can be
conveniently simplified to:
𝑸𝑨 = 𝑪𝑫 𝒂𝟎 √𝟐𝒈𝑯𝟏
When the depth 𝐻2 ≤ the critical depth (DC) the flow is free (i.e. the jet can discharge freely into
the downstream channel) and the value of C for a sharp-edged sluice gate is between 0.5 and 0.6, as
shown in Fig. 9.13. Under these conditions the value of C depends largely upon the relative height
of the opening H1/Y. For example, if H1/Y = 4 then the free flow line gives C = 0.54. If the normal
depth of flow (yn) in the channel some distance downstream of the sluice gate is relatively high
compared to the height of the opening (Y) then it will submerge the jet and affect the discharge, as
with a drowned orifice. One test for submergence is to assume that yn = y2, the depth after a hydraulic
jump. The sequent depth (y1) required to initiate the jump can be obtained from equation (8.36). If
the actual depth at the vena contracta H2 > y1 then a jump cannot form and submergence is likely.
Alternatively, it is possible to start with y1 = H2 and to use equation (8.37) to calculate the sequent
depth y2. The submerged condition will occur if yn > y2. This is illustrated in Example 9.2. The
submerged condition is more complex and requires two variables to obtain C from Fig. 9.13. These
are yn/Y (used instead of H1/Y) and the Froude number in the opening FO = VO/(gY) 1/2, the values
of which are printed within the diagram. For example, if yn/Y = 4

and FO = 1.5 then C = 0.41, considerably less than for the free flow situation (to compensate for not
using the differential head HD = H1 - H2 in equation (9.2)). Equation (9.2) can also be applied to the
other underflow radial gates in Fig. 9.12. For the radial gate in (c) and (d) all of the following affect
the value of C: the upstream head H1, the radius of the gate R, the height of the opening Y and the
height of the pivot P. In the free flow condition, the value of C is about 0.50–0.59 when P/R = 0.1;
0.58–0.63 when P/R = 0.5; 0.68–0.76 when P/R = 0.9 (see Lewin, 1995; Roberson et al., 1998). As

116
a radial gate is opened and closed the angle q at which the flow hits the bottom edge of the gate
changes, which affects the contraction of the jet. In the free flow condition, with q in degrees, a
frequently quoted empirical expression for CC is:
𝑪𝒄 = 𝟏 − 𝟎. 𝟕𝟓(𝜽/𝟗𝟎) + 𝟎. 𝟑𝟔(𝜽/𝟗𝟎)𝟐
An underflow vertical sluice gate discharges freely 56m3 /s into a rectangular channel 7.00m wide. The
gate, which is the same width as the channel, is set at a height of 1.40m above the bed and the depth at
the vena contracta is 0.85m (Fig. 9.14a). The energy loss in the converging flow at a sluice gate is quite
small, so take the head loss between sections 1 and 2 as 0.05𝑉2 2 /2𝑔. The normal depth in the
downstream channel yn = 2.60m. (a) Calculate the depth upstream of the gate (take the energy
coefficients a as 1.05). (b) Confirm that the gate is able to discharge freely. Does a hydraulic jump occur
downstream of the gate and, if so, where? (c) Use the momentum equation to calculate the force on the
gate (assume the momentum coefficient b = 1.00).

(a) Apply the energy equation between sections 1 and 2:


𝐻1 + 𝛼1 𝑉12 /2𝑔 = 𝐻2 + 𝛼2 𝑉22 /2𝑔 + 0.05𝑉22 /2𝑔
For the first iteration, assume V1 = 0. With H2 = 0.85m then 𝑉2 = 56.00/(0.85 × 7.00) =
9.41𝑚/𝑠.
𝐻1 + 0 = 0.85 + (1.05 × 9.412 /19,62) + 0.05 × 9.412 /2𝑔
𝐻1 = 0.85 + 4.74 + 0.23
𝐻1 = 5.82𝑚
For the second iteration 𝑉1 = 56.00/(5.82 × 7.00) = 1.37𝑚/s and 𝑎1 𝑉12 /2𝑔 = 0.10𝑚, so:
𝐻1 + 0.10 = 0.85 + 4.74 + 0.23
𝐻1 = 5.72𝑚
For the third iteration 𝑉1 = 56.00/(5.72 × 7.00) = 1.40𝑚/𝑠 and 𝑎1 𝑉12 /2𝑔 = 0.11𝑚, so 𝐻1 =
5.71𝑚
The fourth iteration gives 𝑉1 = 56.00/(5.71 × 7.00) = 1.40𝑚/𝑠, so 𝐻1 = 5.71𝑚. This answer can
be checked by substituting the values into the discharge equation (9.2) and solving for H1:
𝑸 = 𝑪𝒂𝟎 √𝟐𝒈𝑯𝟏
With H1=5.71m and Y=1.40m, then H1/Y=4.08 so for free flow fig. 9.13, gives C=0.54
 56.00 = 0.54 × (1.40 × 7.00)[19.62𝐻1 ]1/2

117
H1=5.71m
Note that the exact agreement between the two answers is due largely to luck: the answer
obtained from the energy equations depends upon the value assumed for the energy head
loss and particularly the value of 𝑎2 .
(b) From the critical flow equation (8.32) the critical depth is:
𝑦𝑐 = (𝑄 2 /𝑔𝐵 2 )1/3 = (56.002 /9.81 × 7.002 )1/3 = 1.87𝑚
This indicates that the flow at the vena contracta is supercritical (0.85m < 1.87m) so the jet is likely to
discharge freely. However, this can be confirmed using equation (8.36), with the depth after a hydraulic
jump has occurred y2 = yN = 2.60m. With 𝑉2 = 56.00/(2.60 × 7.00) = 3.08𝑚/𝑠 and 𝐹2 =
1
𝑉2 /(𝑔𝐷2 )2 = 3.08/(9.81 × 2.60)1/2 = 0.61 then the sequent depth to 2.60m, i.e. the initial depth
y1 required for the jump to form, is:
𝑦2 2.60
𝑦1 = (√1 + 8𝐹𝑟2 − 1) = (√1 + 8 × 0.622 − 1) = 1.29𝑚
2 2
The depth at the vena contracta is H2 = 0.85m so H2 < D1 which means the jump can form and the gate
is not submerged (if H2 > y1 then a jump cannot form, so the gate would probably be submerged).
(c) Take a control volume between sections 1 and 2 as in Fig. 9.14b, then considering the external
forces acting horizontally:
𝑝1 𝐴1 − 𝐹𝑅 − 𝑝2 𝐴2 = 𝜌𝑄(𝑉2 − 𝑉1 )
Assuming a hydrostatic pressure distribution at sections 1 and 2 then the average pressures are
𝑝1 = 𝜌𝑔𝐻1 /2 𝑎𝑛𝑑 𝑝2 = 𝜌𝑔𝐻2 /2, thus:
0.5𝜌𝑔𝐻1 𝐴1 − 𝐹𝑅 − 0.5𝜌𝑔𝐻2 𝐴2
Taking the values from part (a):
0.5 × 1000 × 9.81 × 5.71 × (5.71 × 7.00) − 𝐹𝑅 − 0.5 × 1000 × 9.81 × 0.85 × (0.85 × 7.00)
= 1000 × 56.00(9.41 − 1.40)
1119.46 × 103 − 𝐹𝑅 − 24.81 × 103 = 448.56 × 103
𝐹𝑅 = 646.09 × 103 𝑁

118

You might also like