You are on page 1of 143

DEFINITION OF A FLUID

In it‘s physical form matter may exist as solids, fluids (liquids and gases) and plasma.

The molecules of solids are so closely packed together that the attractive forces between the
molecules are so large that a solid tends to retain it‘s shape unless compelled by some external
forces to change it. Fluids are composed of molecules with relatively larger distances between
molecules and therefore the attractive forces between molecules are smaller than in solids. As a
result, fluids do not have a definite shape but rather assume the shape of the container in which
they are placed.

F shear stress ( ) F t2 shear stress ()


t0 t1 t3 t2 t3

3
2
 1

Solid, a) Fluid, b)

Fig. 1. Effect of external force on a) Solid and b) Fluid

Consider an external force F applied to the solid body Fig. 1a. The shear component  of this
force changes the shape of this force characterized by the angle . As long as the shear does not
change in magnitude the deformation of the body will be characterized by  independent of the
time the force is applied. Now if the force is applied to the fluid Fig. 1b resulting in a shear
component  , the deformation of the fluid will be characterized by 3 > 2 > 1 where 3 , 2 ,
1 are the deformations of fluid after a time t3 > t2 > t1 of application of shear stress. It must be
observed that the deformation of the fluid will continue indefinitely as long as the shear stress
exists.

A fluid may be defined in two perspectives:-

a) The form in which it occurs naturally :- a substance that is capable of flowing and has no
definite shape but rather assumes the shape of the container in which it is placed.

b) The deformation characteristics when acted upon by a shear stress :- a fluid is a substance that
offers no resistance to shear stress and will deform or continue to change shape as long as there
is applied shear stress no matter how small the stress.

There are plastic solids which flow under the proper circumstances and even metals may flow
under high pressures. On the other hand there are viscous fluids which do not flow readily and
one may easily confuse them with solid plastics. The distinction is that any fluid, no matter how
viscous will yield in time to the slightest shear stress. But a solid, no matter how plastic, requires
a certain limiting value of stress to be exerted before it will flow. Also when the shape of a solid

1
is altered (without exceeding the plastic limit) by external forces, the tangential stresses between
adjacent particles tend to restore the body to its original shape. With a fluid, these tangential
stresses depend on the velocity of deformation and vanish as the velocity approaches zero.
When motion ceases, the tangential stresses disappear and the fluid does not tend to regain its
original shape.

A fluid can be either a dropping liquid or a gas. A liquid is composed of relatively closed packed
molecules with strong cohesive forces. Liquids are relatively incompressible. As a result, a
given mass of fluid will occupy a definite volume of space if it is not subjected to extensive
external pressures. Gas molecules are widely spaced with relatively small cohesive forces.
Therefore if a gas is placed into a container and all external pressure removed, it will expand
until it fills the entire volume of the container. Gases are readily compressible. A gas is in
equilibrium only when it is completely enclosed. The volume (or density ) of a gas is greatly
affected by changes in pressure or temperature or both. It is therefore usually necessary to take
account of changes of pressure and temperature whenever dealing with gases.

FLUID MECHANICS

Fluid mechanics is the science of the mechanics of liquids and gases and is based on the same
fundamental principles that are employed in solid mechanics. It studies the behaviour of fluids at
rest and in motion. The study takes into account the various properties of the fluid and their
effects on the resulting flow patterns in addition to the forces within the fluid and forces
interacting between the fluid and its boundaries. The boundaries may be either solid surfaces or
other fluid surfaces. The study also includes the mathematical application of some fundamental
laws :- conservation of mass - energy, Newton‘s law of motion ( force - momentum equation ),
laws of thermodynamics, together with other concepts and equations to explain observed facts
and to predict as yet unobserved facts and to predict as yet unobserved fluid behaviour. The
study of fluid mechanics subdivides into fluid statistics, fluid kinematics and fluid dynamics.

Fluid statics : is the study of the behaviour of fluids at rest. Since for a fluid at rest there can be
no shearing forces all forces considered in fluid statics are normal forces to the planes on which
they act.

Fluid kinematics deals with the geometry (streamlines and velocities ) of motion without
consideration of the forces causing the motion. Kinematics is concerned with a description of
how fluid bodies move.

Fluid dynamics is concerned with the relations between velocities and accelerations and the
forces causing the motion.

― If the larger public may be permitted to have inadequate realizing sense of the part which fluid
mechanics plays not only in our daily lives, but throughout the entire domain of nature, the
engineer may not be permitted. For the flow of rivers and streams in their boundaries, the
circulation of blood in the arteries and veins, the flight of insects and birds and airplanes, the
movement of ships in waters or fish in the depths are all, in major degrees, varied expressions of
the laws of fluid mechanics.‖ - Durand W. F

2
The Engineer should therefore approach the study of fluid mechanics with the purpose:

a) to obtain an understanding of and feeling for the behaviour of fluids.


b) to solve problems of the type encountered in Engineering practice using the
knowledge and understanding of the behaviour of fluids.

If the engineer can predict a fluid‘s behaviour regardless of the constraints, he can design,
construct and exploit machines that use fluids (e.g. engines, ships, pumps, turbines etc. )

A fluid like oil or steam has stored energy. If we know how to predict the fluids behaviour, we
can transport that stored energy from one location to another or we can transform that stored
energy into useful work. A fluid energy can be converted into other forms of energy by
combustion, mixing, impact reaction etc. Blood is an example of fluid that transports energy in
the form of nutrients, chemicals and life- giving substances to our energy- burning cells.

SYSTEMS AND CONTROL VOLUMES

We did mention earlier that in the study of fluid mechanics, we make use of the basic laws in
physics namely
i) The conservation of matter (which is called the continuity equation)
ii) Newton‘s second law (momentum equation)
iii) Conservation of energy (1st law of thermodynamics)
iv) Second law of thermodynamics

In addition to these general laws, there are numerous subsidiary laws that apply to specific types
of media. (eg. equation of state for perfect gas, Newton‘s viscosity law for certain viscous fluids,
Hooke‘s law etc. )

In employing the basic and subsidiary laws, either one of the following models of application is
adopted ;
1) The activities of each and every given mass must be such as to satisfy the basic laws and the
pertinent subsidiary laws - SYSTEM
2) The activities of each and every volume in space must be such that the basic and the pertinent
subsidiary laws are satisfied
A system is a predetermined identifiable quantity of fluid. It could be a particle or a collection of
particles. A system may change shape, position and thermal conditions but must always contain
the same matter. ( close system.)

A system
Control volume

A control volume refers to a definite volume


Control designated
surface in space usually with fixed shape. The
boundary of this volume is known as the control surface. A control volume mode is useful in the

3
analysis of situations where flow occurs into and out of a space. In thermodynamics it
corresponds to open system.

FORCES (BODY AND SURFACE)

In the study of continua, we distinguish between two types of forces

1) There forces distributions which act on matter without the requirement of direct contact are
called body forces (e.g. gravity, magnetic, inertia, etc. Body forces are given on the basis of the
force per unit mass of the material acted on.
2) These forces on a body that arise from direct contact of this body with other surrounding
media are called surface forces eg. pressure force, frictional force, surface tension.

FLUID PROPERTIES

Property :- Is a characteristic of a substance which is invariant when the substance is in a


particular state. In each state the condition of the substance is unique and is described by its
properties. The properties of a fluid system uniquely determine the state of the system.

In thermodynamics a distinction between those properties of the substance whose measure


depends on the amount of the substance present - extensive and those properties whose measure
is independent of the amount of substance present - intensive properties. Example of extensive
properties :- weight, momentum, volume,energy. Clearly, changing the amount of mass directly
changes the measure of these properties.

For each extensive variable eg. Volume, Energy E on can introduce by distributive
measurements the corresponding intensive properties namely volume per unit mass v and energy
per unit mass e do not depend on the amount of matter present. eg. temperature, pressure,
viscosity, surface tension, mass density etc.

PHYSICAL PROPERTIES OF FLUIDS

Fluid properties are dependent upon the molecular characteristics of the fluid, including
intermolecular spacing, attraction and activity.

Each fluid property is important in a particular field of application. For example, applied
hydraulics is concerned with the physical action and the interaction of particular masses of fluid
in either the static or kinematic state. The actions and interactions are analysed by attributing to
the fluid certain physical properties each of which is defined to be the controlling property for a
particular type of action. Viscosity plays an important role in the problems of hydraulic friction.
Mass density is important in uniform flow. Compressibility is a factor in water hammer. Vapour
pressure is a factor in high velocity flow.

Mass density and Unit (specific) weight.

4
Mass density and unit weight are the two important parameters that tend to indicate heaviness of
a substance.
Mass density is the mass per unit volume usually denoted by the Greek letter ―rho‖

M
 kg/m3
V
At standard pressure (760 mmHg) and 4o C density of water  = 1000 kg/mm3
Specific volume : Is the reciprocal of the density ie. the volume occupied per unit mass of fluid.
1 V
Vs =  ( m3 / kg)
 M
Specific (unit ) weight  (gamma) - Is the weight per unit volume of the substance (is and
indication of how much a unit volume of a substance weighs.)

W Mg
   g ( kgm/s2)
V V
Specific Gravity : Is the ration of the weight of a substance to the weight of an equal volume of
water at standard conditions.
Viscosity : Is the property of a fluid to offer resistance to shear stress. Fluids offer resistance to a
shearing force. Viscosity is a property of a fluid that determines the amount of resistance.
Viscosities of liquids vary inversely with temperature, while viscosities of gases vary directly
with temperature.

Deformation of fluid resulting from application of a constant shear force.

As was previously defined, fluid is a substance which deforms continuously under shearing
(tangential) stress no matter how small the stress. The magnitude of the stress depends on the
rate of angular deformation. A shear force is the force component tangent to a surface, and this
force divided by the area of the surface is the average shear stress over the area. Shear stress at a
point is the limiting value of shear force to area as the area is reduced to the point.
The diagram shows a substance placed between two closely spaced parallel plates. The lower
plate is fixed, and a force F is applied to the upper plate which exerts a shear stress F/A on the
substance where A- the area of the upper plate. When a force F causes the upper plate to move
with a steady velocity no matter how small the magnitude of F, one can conclude that the
substance between the plates is a fluid.

The fluid in immediate contact with a solid boundary has the same velocity as the boundary. ie.
there is no slip at the boundary. Therefore the fluid just immediately on the lower plate has a
zero velocity while the fluid just immediately below the upper plate moves with the same
velocity as F ie. U. Therefore within some time dt, the fluid in area a b c d flows to the new
position a bl cl d each particle moving parallel to the plate. At any point at a distance y from the
lower plate, the velocity

y
U(y) = Uo *
Y

5
du Uo

dy Y
Experiments show that, other quantities being held constant F is directly proportional to the A
(area) and the velocity U and inversely proportional to the distance between the plates Y
AUo
F where  is the proportionality factor.
Y
The shear stress is defined as Γ (Tau)
F Uo du
Γ=  ; = 
A Y dy

The shear stress at any point in the fluid ,


du
= 
dy

NEWTONIAN AND NON- NEWTONIAN FLUIDS

Not all fluids show exactly the same relation between stress and the rate of deformation.

Newtonian fluids: are fluids for which shear stress is directly proportional to the rate angular
deformation or a fluid for which the viscosity  is a constant for a fixed temperature and
pressure. eg. Air, water, etc. Petroleum, kerosene, steam.
Non-Newtonian fluids : are fluids which have a variable proportionality (viscosity  ) between
stress and deformation rate. In such cases, the proportionality may depend on the length of time
of exposure to stress as well as the magnitude of the stress eg. Plastics, paint, blood, ink, etc.

COMPRESSIBLE AND INCOMPRESSIBLE FLUIDS.

Compressible fluids are fluids whose specific volume V (density) is a function of pressure. An
incompressible fluid is a fluid whose density is not changed by external forces acting on the
fluid.

Hydrodynamics is the study of the behaviour of incompressible fluids whereas gas dynamics is
the study of compressible fluid.

PERFECT GAS :
In this treatment, thermodynamics relations and compressible fluid flow cases have been limited
to perfect gas.

BULK MODULUS OF ELASTICITY

For most purposes a liquid may be considered as incompressible, but for situations involving
either sudden or great changes in pressure. Compressibility of a fluid is a measure of the change
in volume of the fluid when it is subjected to outside force. It is defined in terms of an average
bulk modulus of elasticity K.

6
F
A strain

P
F/A K=
V
V
V
dp

V/ V

strain V1 / V

where V is the decrease in volume conforming to the increase in pressure P, and V is the
original volume of fluid.

SURFACE TENSION
By the molecular theory liquid consists of molecules which under certain conditions do attract
each other. If we should consider the liquid to be made up of a column of spheres, attracting
each other, then it can be seen that apart from the extreme ones A and B, all the rest will be in
equilibrium. Two forces F acting on each particle balances one another. For the two extreme
particles A and B, each will be acted upon by one force F and so the column will be under
compaction force F. Now consider a vessel with a fluid. We can therefore
n=
say that the liquid in the container will be acted upon by the sum of  F force of the molecules
n=1
at the surface.

The molecular pressure exerted by the thin film of surface molecules (which acts as a piston
pressure) as experiments show are very high eg. for water it is approximately 11 105 kPa (or
11,000 atmospheres). It may be said that the low compressibility of liquids may be explained by
the fact that the external forces applied to the surface are relatively small in comparison to the
intermolecular forces (causing the surface tension). These forces F tend to pull the surface
molecules tightly to the lower layer and cause the surface to behave as though it were a
membrane. The magnitude of this force per unit length is defined as surface tension  (sigma).
For a small spherical droplet of radius r, the internal pressure P necessary to balance the tensile
force due to surface tension  is given by

7
p  r2 = 2  r  
p
2  r 2
p=  
r2 r

Assuming the surface of the liquid is made up of an infinite number of particles A, then we can
show that the molecular pressure

Pconcave<Pflat < Pconvex

Radius of influence

a11 b11

a1 b1

a2 ba2

Let‘s examine a molecule m near the surface of a liquid a, b, other1


molecules in the vicinitybof
2
2

the radius of influence of the molecule m will be attracting the a2 molecule m. Let‘s neglect the
attraction due to the air above the surface. The forces of attraction of molecule m by molecules
in region A will be balanced by the forces of attraction of molecules in region B. We can see
that the molecule m is being pulled down by those molecules in the region C bounded at the top
by line a2 b2 (in the case of flat surface
a1 b1 ) or curve a2l b2l (in case of concave surface a1l b1l . Since the area of C bounded by the line
a2 b2 is greater than C bounded by the curve a2l b2l then one can conclude that Pflat < Pconvex .

CAPILLARITY :

Is the rise or fall of a column of fluid (in a narrow tube called capillary tube) inserted in the fluid.

8
In the contact area between the fluid and container, we can have two cases ;
1) If the intermolecular cohesive forces between two molecules of the fluid is greater
than the adhesive forces between the molecules of the container and the molecule of
the fluid, case (a) - convex meniscus is obtained.
2) On the other hand if the adhesive force of molecule of the container and fluid is
greater than the cohesive force of the fluid molecules, case (b) - concae meniscus
is obtained.

Now assume that the surface of the liquid in vessel is being acted upon by a pressure Pflat, and the
surface in the capillary tube is acted upon by Pcomncave or Pconvex depending on the type of liquid.
In case (a) we have convex surface and therefore Pflat < Pconvex and therefore the level falls.
In case (b) we have concave surface and therefore Pflat > Pconcave. There is therefore a rise due to
Pflat - Pconcave.
The rise or fall in the capillary tube is given by

2Cos
h= Surface tension force Fs = 2rCos. Weight of column of

fluid being supported by surface tension
G = r2h
G = Fs => 2rcos= r2h
2Cos
h=
r

Kinematic Viscosity : (nu) is the ratio of the dynamic viscosity to the density of the
fluid.

 =  / Ns / m2 kgm-3 = m2 / s

9
HYDROSTATICS
HYDROSTATIC PRESSURE
Hydrostatic deals with fluid at rest. Hydrostatics studies the laws governing the behaviour of
fluid at equilibrium when it is subjected to external and internal forces and bodies at equilibrium
when they are immersed in the fluid.
Shear stress (viscosity) in a fluid at rest is always zero i.e. =0. It should also be noted that
tensile stresses cannot exist in fluid. Therefore in fluid at rest we shall be talking only of normal
stresses. I.e. =n
The forces acting on a given isolated fluid element may be subdivided into
1. body and
2. surfaces forces.
Body forces act directly on the particles of fluid filling a particular volume (e.g the force of
gravity). Surfaces forces act only on the surface of the isolated volume of fluid (such as the
pressure of a solid on the fluid which is flowing past it)
The basic concept of hydrostatics is the concept of hydrostatic pressure

HYDROSTATIC PRESSURE
Consider an arbitrary volume in a fluid at rest. Let us select any point M in this volume and draw
an arbitrary plane AB through the point M. The plane AB will divide the chosen volume into two
parts I and II. Part I will exert a force on part II through the surface AB. Let P be that part of the
force which acts on the area S. The force P acting on the surface S is called Hydrostatic force.

Note: the force P relative to part Ii is an external surface force, but to the whole volume it is an
internal force. There is a reaction of force P from part II to I.
If we divide the force (modulus) P by the area S we shall obtain
P
2.1--------------  Pcp ---- average hydrostatic pressure.
S
The average hydrostatic pressure Pcp is the hydrostatic force per unit area.

10
P
If the element of area S is made externally small and the ratio  Pcp is taken to the limit, we
S
obtain he pressure at a given point.
P
2.2-------------- P  lim it   ----pressure of a given point in a fluid or simply hydrostatic
S 0  S 
 
pressure.

PROPERTIES OF HYDROSTATIC PRESSURE

1). Hydrostatic pressure is a compression stress and always acts along the inside normal to the
element of area.
2). The hydrostatic pressure p at a given point in a fluid does not depend on the orientation of the
surface i.e. on the incline of the surface. It is the same in all direction.

SURFACE AND BODY FORCES

In the study of continua, we distinguish between two type pf forces.


1. Those force distributions, which act on matter without the requirement of direct contact,
are called body-forces (e.g. gravity, magnetic force, inertia, etc). body forces are given on
the basis of the force per unit mass of the material acted on.
2. Those forces on a body that arise from direct contact of this body with other surrounding
media are called surface forces. (e.g. pressure forces, frictional forces surface tension,
etc). They are given on the basis of the force per unit area of the material acted on.

DIFFERENTIAL EQUATION OF A FLUID AT REST (EQUILIBRIUM)

EULER‘S EQUATION

11
Let us consider the equilibrium of an elemental parallelepiped in a fluid.
Since it is in equilibrium, the projection of all forces on the x, y, z axis should be equal to zero
i.e. Fx=0, Fy=0, Fz=0. We shall consider only the x-axis Fx=0.

Projection of surface forces on the x-axis


Force on side ABCD
dFx=pdydx-----------------2.7
Where p is the average hydrostatic pressure on side ABCD
Force on side A1B1C1D1
dF1=p1dydz----------------2.8
Where p1 is the average the average hydrostatic pressure on side A1B1C1D1
Let find the value of p1
Since p=f (x, y, z) p is a function of x,y,z, then the transition from point A to A1, B to B1, C to C1
and D to D1 depends on only the coordinate x, there fore the change in pressure depends only on
the coordinate x. with this in mind,
p
p1=p+ dx
x
1 1 p
dF =p dydz=(p+ dx ) dydz---------2.9
x

Projection of body forces on the x-axis


The projection of body forces on the x-axis is the product of the mass of fluid and the projection
of acceleration on the x-axis. I.e.
dRx=dxdydz.-------------2.10

Where =density of fluid

--Projection of acceleration on the x-axis


dxdydz=dV—volume of elemental parallelepiped
Fx=0---sum of surface and body forces on the x-axis equals zero.
p
Fx=pdydz - (p+ dx ) dydz + dxdydz.=0----------------2.11
x

Expanding the bracket and dividing through by dydxdz


p
  X  0
x

1 p
  X  0 ------------------2.12
 x
By analogy Fx=0; Fz=0
Then we shall obtain
1 p
 X 0
 x

12
1 p
 Y  0 -------------2.13: Euler‘s Differential Equation for a Fluid in
 y
Equilibrium
1 p
 Z 0
 z

Now let us transform Euler‘s equation a little by multiplying the first by dx; the second by dy;
and the third by dz, and add all the three.
1 p
 dx  Xdx  0
 x
1 p
 dy  Ydy  0 ----------------------------------------2.14
 y
1 p
 dz  Zdz  0
 z
 p p p 
1

 dx  dy  dz   Xdx  Ydy  Zdz --------------2.15
 x y z 
Since hydrostatic pressure is a function of independent coordinates x, y, z, then the first three
functions on the left side of the above equation being the sum of three partial differential equals
the exact (total) differential. Therefore the above equation can be written as

1
 dp  Xdx  Ydy  Zdz ---------Basic differential equation of hydrostatic

Since the left hand side of equation is an exact (total) differential, then the right hand side shall
also be an exact differential of a certain function say U (x, y, z)
Therefore we can write

Xdx+Ydy+Zdz= U (x, y, z)------------------------------------2.17

We can write the exact differential dU(x, y, z) into partial differential i.e.
U U U
dU ( x, y, z )  dx  dy  dz -------------------2.18
x y z
therefore
U U U
Xdx  Ydy  Zdz  dx  dy  dz
x y z
1 p U U U
 X 0 X  ; Y ; Z -------------------------------2.19
 x x y z

Since U is a function of only coordinates (x, y, z) and its partial differential gives the
corresponding projection of body forces per unit mass (X, Y, Z), then the function U is a
Potential Function.

13
INFERENCE

Fluid can be in a state of equilibrium (rest) when and only when forces, which have potentials,
act it upon
From equation 2.16 and 2.17 we have

1
 dp  dU ------------------------------------2.20

Integrating 2.20, we have
P=U+C--------------------------------2.21
To find C let us consider a point in a fliud with p and U know.
Assuming at this point p=p0 when U=U0

Po=U0+C

Therefore C=p0 - U0-------------------------2.22

Substituting C into 2.21 we obtain

p-po-------------------------2.23

or
p=p0 + (U- U0)

This is the general equation of hydrostatics for the pressure at a point in a fluid.

Note: =constant

HYDROSTATIC PRESSURE AT A POINT IN A FLUID AT REST WHEN THE FORCE


OF GRAVITY IS THE ONLY BODY FORCE ACTING

Let us consider a closed container with a fluid at equilibrium. Let po be the external surface
pressure. Let us put the axis at the surface as shown in the diagram. Let the point M isolate a
fluid with a unit mass i.e. m=1unit.
Since force of gravity is the only body force acting, we shall have the following

X=0; Y=0; Z=-g-----------------2.24

Substituting 2.24 into the basic differential equation of hydrostatics 2.16 we shall obtain.
1 dp
dp   gdz Or   -------------------2.25
 dz
Integrating 2.25, we have
p=-gz + C
Or ---------------2.26

14
p=-z + C since g=

where C is the constant of integration.


To find C let us consider a point at the surface of fluid. At that point O, Z=0; p=po

po=C---------------------2.27

Putting C into 2.26 we have

p=-z + po------------------------2.28

Now let h be the depth of immersion of the point M. h=-z


Finally we have

p=po + h-------fundamental equation of hydrostatics-----2.29

In the equation 2.29


p---- is known as the absolute hydrostatic pressure at the point M
h--- is the body pressure i.e. pressure due to the body of column of fluid
above M.

Conclusion: the absolute pressure at a point is the sum of the external surface pressure and the
body pressure (pressure created by the column of fluid on point).
If the container is opened, then
po=pa

pa= atmospheric pressure or barometric pressure.

15
Equation 2.29 is then written as
p=pa + h---------------------2.30

p-pa=h-------Gauge or manometric pressure

Manometric Pressure: is the differential (excess) pressure above atmospheric pressure at a


point in a fluid.
In practice we often use the manometric pressure instead of the absolute pressure. So from now
we shall denote;
PA=absolute
p=excess or manometric pressure

 PA=pa + p

When p-pa=h is negative, we have vacuum pressure.

PA=po + p ------------2.31

In this formula
po---- external press
p----- excess pressure due to the column of fluid

PASCAL’S LAW: - HYDRAULIC PRESS

From 2.29 it can be seen that the pressure at any point in a homogeneous liquid at rest depends
(apart from external pressure) only on the height of the column of liquid above the point and,
therefore the surfaces of equal pressure are horizontal.

p= po + h--------------------2.29

It can also be seen that if the external pressure changes from po to po1=po + po the pressure at
all point in the fluid at rest also changes by the same value po. it is therefore evident that liquid
possesses the property of total transmissibility of the external pressure.
Pascal’s Law therefore states: pressure (external) which arises (or which is applied) at the
surface of a liquid at rest is transmitted throughout the liquid in all direction without any change.

16
HYDRAULIC PRESS: the distinctive characteristic of the hydraulic press is its ability to
produce great forces by expending fairly small original forces.
A hydraulic press is shown in the diagram. The material 3 to be pressed is on the table 4 and is
held against fixed traverse 2. The cylinder 1 of the press, the cylinder 7 of the pump and the
pipeline 6 are filled with a liquid.
The force F1 acts on the piston pump 8 of area A1 causes it to travel downwards and to exert
pressure on the liquid surface below. This pressure is
P=F1/A1
From Pascal‘s law, this pressure is transmitted to the piston 5. The result is a useful force F2
under whose action the material is pressed.
P=F2/A2
Where A2 is the area of piston 5
Therefore F1/A1 =F2/A2  F2/F1 =A2/A1=D2/d2
Expressing the areas of the piston in terms of their diameters and performing the required
operation, we have
F2/F1 =D2/d2 -------------2.32
From 2.32, if the diameter of the greater piston is 10times that of the smaller one, then the force
on the greater piston is 100times that on the smaller one.

PIEZOMETRIC (HEAD) HEIGHT

It can be shown that the pressure (absolute manometric or atmospheric) can be represented by a
height of a column of fluid. The diagram below shows a closed container partially filled with a
fluid with specific weight . The glass tubes To is closed to the atmosphere and T is opened to the
atmosphere and both are linked to the container. We assume that tube To is completely rarefied
(POT=0).

17
PA  Pa P
hex   [M]
 
where P-excess or gauge or differential pressure.

PA  Pa P
hex   --------2.33-----piezometric head (height)
 
INFERENCES

hA---absolute piezometric height: - is the height of that column of the fluid which with its weight
is able to create a pressure equal to the absolute pressure.
hex---piezometric height: - is the height of that column of fluid which by its weight is able to
create a pressure equal to the excess/gauge (manometric, differential) pressure.---gauge pressure.

POTENTIAL ENERGY OF A LIQUID AT REST

POTENTIAL HEAD
Liquid at rest or in motion possesses a certain amount of mechanical energy i.e. possesses the
ability to do a certain amount of work liquid at rest possesses only potential energy (energy by
virtue of its position relative to a certain level (datum)
Let us look at fig. 2.6. the ordinate Z (i.e distance from the datum to any point in the liquid
under consideration) is called ―head‖. At the point n in the liquid, as a result of the differential
(excess) pressure (PA-Pa), a column of liquid of hex relative to the plane MN and H relative to the
datum.
We can therefore say that liquid, located at the point n can do these useful works.

i) by falling from the plane MN onto the datum O-O

(P.E)Z = z x G ----------------------------------( )
ii) by rising to the height hex as a result of the differential pressure PA -Pa =p
(P.E.)p = hex x G ------------------------------------( )
Total work that can be done by the liquid of weight G located at n is:
P.E. = z x G + z x G = = (P.E.)z + (P.E.)p -------------------------------- ( )
P.E. is called the potential energy of the liquid of weight G located at the point n.
Ted at point n.
SPECIFIC POTENTIAL ENERGY; is defined as the potential energy per unit weight of the
fluid.
S.P.E. = P.E./G = {(z x G) + (hex x G )}/G = z + hex = H --------------------- ( )

18
Specific potential Energy is the sum of
i specific potential energy by virtue of position (z).
ii). Specific potential energy due to pressure hex = p/γ

POTENTIAL HEAD: In hydraulics (or fluid mechanics) ―head‖ is used to denote specific
potential energy; i.e. a measure of energy per unit weight of the liquid.
Therefore the potential head H can be written as
H = z + hex
Z – is called the geometric head
Hex is called the pressure (or piezometric) head.

H = z + hex = z + p/γ

Inference: In a fluid at rest, the value of the potential head is the same at all points within the
liquid.
Home work: show that the potential head is the same for all points in a liquid at rest; i.e H 1 = H2
= H3 etc.

.3

.1
.2

MEASUREMENT OF FLUID PRESSURE

There are generally two types of pressure measuring devices:


1. Tube gauges: - are those instruments that work on the principle that a particular pressures
can support a definite weight of a fluid and this weight is defined by definite column of
fluid.
2. Mechanical gauges: - work on the principle that the applied pressure will create a
deformation in either a spring or a diaphragm.

1.Tube Gauges

a.) Piezometric Tube: the simplest pressure measuring tube is the piezometer and it consists
of a narrow tube so chosen that the effect of surface tension is negligible. When connected to
the pipe whose pressure is to be measured, the liquid rises up to a height h, which is an
indicative of the pressure in the pipe p=h. the piezometer is somewhat limited in use
because even moderately high pressures require usually long tubes and also gas pressures
cannot be measures with it (since gases cannot form free surface). Piezometers are not

19
suitable of measuring negative pressures, in this case atmospheric air will enter the pipe
through the tube.

b.) Manometers: to overcome the above mentioned limitations of the piezometer, an


improved form of the piezometer consisting of a bent tube containing one or more fluids of
different specific gravities is used. Such a tube is called a manometer. There are different
types of manometers
i. Simple manometer
ii. Micro manometer
iii. Differential manometer
iv. Inverted differential manometer

i. Simple Manometer:

A simple manometer consists of a tube bent in U-shape, one end of which is attached to the
gauge point and the other is opened to the atmosphere. The fluid used in the bent tube is
called the manometric fluid (usually mercury) and the fluid whose pressure is to be measure
(and therefore exerts pressure on the manometric fluid is referred to as the working fluid.

20
By using the principle that the pressure on the horizontal and in the same continuous fluid is
the same, we shall state that:
For diagram A
P1=P2
P1=PA + h11
P2=Pa + h22
 PA + h11= Pa + h22
PA-Pa= h22- h11
For diagram B
P2=Pa=P1
P1=PB + h11 + h22
 PB + h11 + h22=Pa
PB-Pa= -h11 - h22 = vacuum gauge

ii. Micro manometers

Micro manometers are used for measuring how pressures, where accuracy is of much
importance. Micro manometer is a modified form of the simple manometer with the cross-
sectional area of one or both of the limbs made much larger (about 100times) than the other.
O-O & O1-O1 are the levels in the limbs when there is no pressure applied. When pressure is
applied there will be changes in the limbs as shown in the diagram. To find the pressure in A
we choose a horizontal on which the pressure in the half limb equals the pressure in the right
limb. E.g. horizontal 1-1----2-2
The pressure on level 1-1 is P1 and pressure on level 2-2 is P2.
P1=PA + w (h1+dh)
P2=m (h2+dh)
But P1=P2------on the same horizontal and in a continuous fluid.
PA + w (h1+dh) = m (h2+dh)
PA = mh2 + mdh - wh1 - wdh
21
By equation of volumes,
D2dh/4=d2h2/4  dh=(d/D)2h2
PA= mh2 + mh2 (d/D) 2 - wh1 - w (d/D) 2h2
= mh2[1+ (d/D) 2] - w[h1- (d/D) 2h2]
= mh2 - wh1 assuming (d/D) 2 to be very small.

Inclined tube Micro manometer


This type is more sensitive than the vertical tube type. Due to the inclination the distance
moved by the manometric fluid in the narrow tube will be comparatively more and thus give
a higher reading for a given pressure.

iii. Differential Manometer

It is a device used for measuring the difference in pressure between two points in a pipe, or in
two different pipes.
A differential manometer consists of a U-tube containing the manometric fluid. The two ends
of the tubes are connected to the points, whose differential pressure is to be measured.
The pressure at 1-1 equals the pressure at 2-2.
The pressure at 1 is given by
P1=PA + (x + h) w
The pressure at 2 is given by
P2=PB + xw + hm
But P1=P2
PA + (x + h) w=PB + xw + hm
PA – PB= xw + h m - xw - hw
PA – PB= h (m - w)
Converting this differential pressure to an equivalent height of the working fluid,
hw = (PA – PB)/w = h (m/w – 1)
hw = (PA – PB)/w = h (S.Gm/S.Gw – 1)
If the manometric fluid is mercury and the working fluid is water then hw can be
approximated as hw = 12.6h

iv. Inverted U-tube Differential Manometer

An inverted U-tube differential manometer is used for measuring difference of low pressures,
where accuracy is the prime consideration. It consists of an inverted U-tube containing a light
liquid. The ends are connected to the pipes whose difference in pressure is to be found. By
assuming a equality of pressures at horizontal 1-2, the pressure difference between A and B
could be found.

22
FORCES OF HYDROSTATIC PRESSURE ON PLANE SURFACES IMMERSED IN
FLUIDS

Let us consider in Fig. (below) an open container, filled with a fluid and an inclined plane OM.
On thye inclined plane OM is an arbitary plane figure AB with area A. It is known that the force
of hydrostatic pressure FA acting on the figure AB will be normal to the surface AB. Our task is
two folds:
i) to find the magnitude of the force
ii) to find the point (position) of action of this force.

O(x)

θ
h A
F hC
Z
m
ZD

ZC
B

M A
Z

D C m
X
B dA

Task 1; finding the magnitude of the force FA

Let us choose an arbitrary point m on the surface AB immersed in the fluid at a depth h, and at a
distance z from the axis OZ. At the point m, we choose an elemental area dA, surrounding the
point m. Since the area dA is elemental, we can assume that the pressure within it is the same
and equals to pm = pa + γh. Therefore the hydrostatic force on the area dA is given by:
dF  p m dA  ( p a  h)dA
but...h  z sin  --------------------( )
dF  ( p a  z sin  )dA

The total force acting on the surface A is obtained by integrating dF over the whole surface A
FA   ( p a  z sin  )dA   p a dA   sin   zdA...      (2.4)

FA  p a A   sin   zdA...                  (2.5)


But ..the.. exp ression.. zdA  ( St ) ox  z C A..          (2.6)

Where (St)ox is the static moment of area A about the axis Ox;
23
zC – is the coordinate of the centre of gravity of the plane surface AB from axis Oz
Now substituting equation (2.6) into equation (2.5), we shall obtain:

FA  p a A  z C A sin 
But ..z C sin  ..  hC
FA  ( p a  hC ) A  p a A  hC A..            (2.7)

Where paA is the force due to atmospheric pressure, which is transmitted through out the fluid
onto the planes surface AB and γhCA is the force due to pressure of the column of fluid on the
surface AB
Since in most cases, we shall be interested only in the gauge pressure, the total force on a plane
surface immersed in a fluid may be written finally as:
FAB = γhCA ------------------------(2.8)
Task 2. Finding the point of action (Centre of Pressure) of the resultant force of a plane
surface immersed in fluid

To find the centre of pressure, we are going to use the theory of moments which states that the
moment of the resultant force about a point (or axis) equals the sum of moments of all the forces
about the same point (or axis).
Let ZD be the centre of pressure and let us write the equation of moments about the axis Ox.
The moment dM of the elemental force dF about Ox equals

M (dF ) Ox  dF .z  (hsA) ----------------------(2.9)

The sum of moments of all the individual forces is given by:

 M (dF ) Ox   z sin  .dA    sin  .z 2 dA -----------(2.10)

The moment of the resultant force about the same axis Ox is given by:

FA.zD = γ.hC .A.zD = γzC.sinθ.A.zD ---------------------(2.11)

Equating equations (2.10) and (2.11) according to the theory of moments, we have

z
2
dA I ox
zD   ---------------------------------(2.12)
zC .A ( St ) ox

Where Iox = ⌠z2dA – 2nd moment of area or moment of inertia of AB about the
axis ox

and (St)ox = zc.A - 1st moment of area or the static moment of AB about the axis
Ox.

24
It is also known from the theory of moments that the moment of inertia of a body about a given
axis equals the moment of inertia about an axis parallel to the given axis and passing through the
centre of gravity (centroid) plus the product of the area and the square of the distance between
the axes; i.e.

Iox = IC + z2c.A ----------------------------------------(2.13)

Substituting equation (2.13) into equation (2.12) we shall obtain

I c  z c2 . A I
zD   z c  c -------------------------(2.14)
zc .A zc .A

Ic
hD  hc  sin 2 
hc .A

Ic I
The quantity  sin 2   e -- is called the eccentricity.
z c . A hc . A

The eccentricity is always positive and indicates that the centre of pressure is always below the
centre of gravity of the object. The value of e decreases with increasing depth of immersion of
the object. There are two situations for which the value of e equals to zero; ie. When the depth of
immersion ≈ ∞ and when the surface is horizontal i.e. sin θ=0

GRAPHICAL (PRESSURE DIAGRAMS) METHOD FOR FINDING HYDROSTATIC


FORCE ON PLANE SURFACES IMMERSED IN FLUIDS

Ox
O x

h
b

m m
H
C C

F D D

B A A

γH

25
We are supposed to find the hydrostatic force acting on the vertical gate OA with width b
immersed in a fluid.
Assume a particle on the gate OA and imagine the particle moving along the gate from the point
O to the point A. At any depth h, the particle will experience a pressure of γh. It can be seen
that the hydrostatic pressure on the particle will change from zero at point O to γH at point A.
The pressure change is linear and the gradient of the pressure line is γ.
Therefore if at the point A, we put a segment AB equals to γH and join point B to point O, we
shall obtain the pressure diagram on the gate OA The area of this pressure diagram OAB
multiplied by the width of the gate b gives the magnitude of the hydrostatic force acting on the
gate

1 2
F  .b  H  .b F ---------------------------------(2.15)
2
The force F is perpendicular to OA and passes through the centre of gravity of the pressure
diagram. Since the pressure diagram has a triangular distribution, the centre of gravity lies 1/3 H
from the base of the gate.
It is often advisable (easy) when drawing the pressure diagram to put H instead of γH. In this
case, the hydrostatic pressure is then given by

F = Ω.γ.b
Properties of the pressure diagram
1. Every ordinate on the pressure diagram gives the hydrostatic pressure at the point
2. The area under the pressure diagram gives the value of the hydrostatic force per unit
width of the gate.

O O

F
h1
F1

F
F1 h2
F2
B A F2
A
γh1
γh2

LOCK GATES

26
C
Hinge

B β H1
H
N θ
F1
D
H2
θ
P R θ F2

A Hinge
PLAN Elevation

Lock gates are hydraulic structures used in navigation for regulating water levels in channel for
the purposes of creating necessary levels for navigation.
AB and BC are two lock gates. Each gate is held in position by two hinges. In the closed
position, the gates meet at B exerting thrust on one another. Now let us consider the equilibrium
of one of the gate eg. gate AB.
Let H1 and H2 be the depths of water on the two sides of the gate and let F1 and F2 be the
hydrostatic forces exerted by the water on the two sides of the gate. The resultant hydrostatic
force on the gate is F1 – F2
Let N be the reaction at the common contact surface of the two gates. Let R be the resultant
reaction of the top and bottom hinges. The three forces, F, N, and R will all be in the same
horizontal plane (i.e coplanar). Since F, N and R are coplanar and they bring about the
equilibrium of the gate AB, then, the three forces must be concurrent at a point. i.e point D and
Angle DBA = angle DAB = θ -------------------------------(2.16)

Resolving forces along AB


N.cosθ = Rcosθ ---------------------------------------------(2.17)
N =R

Resolving forces normal toAB


F = N sin θ + R sin θ = 2Rsin θ ----------------------------(2.18)

Since F = F1 – F2 is known, R can be found.

Reaction at the top and bottom hinges


We know the resultant water pressure F acts normal to the gate and acts at the middle of the gate
AB. Thus one half of this force is transmitted to the hinges of the gate and the other half to the
reaction of the common contact. Let RT and RB be the reactions of the top and bottom hinges so
that RT + RB = R
Taking moments about the bottom hinge, we have

27
F.H  F H 
RT . sin  .H   1 1    2 . 2  --------------------------(2.19)
 2. 3   2 

Resolving forces in the horizontal direction,

F1 F2 F
RT . sin   RB . sin     --------------------------(2.20)
2 2 2

HYDROSTATIC PRESSURE FORCES ON CURVED SURFACES

It well be recalled that in the case of plane surfaces all the forces due to the pressure on its
different elements are parallel, thereby making the resultant easy to find. The forces of
hydrostatic pressure on different elements of a curved surface have different directions and
so it is impossible to predict before hand the direction of the resultant pressure force R.
Let us consider a curved surface ABC with length b. let Px and Pz be the horizontal and
vertical components of the force due to hydrostatic pressure acting on the curved surface. To
find these components lets erect the plane DE. The plane DE will isolate that volume of
liquid ABCED whose equilibrium we wish to investigate.

D A
A

Rx

Ph B
A
Rz
G
C
E A
A

RD

The volume ABCED is acted upon by the following:


1.the force Ph acting on the vertical side DE
2.the force RD-reaction of the base EC
RD=[area (C1CED)] b
3. the reaction R from the curved surface. Rx,Rz is the horizontal and vertical components
respectively.
4.force due to liquid‘s own weight G
G=[area (ABCED)] b

28
Now lets resolve all forces acting on the volume ABCED onto the x- and z- axis.
Rx=0; Ph – Rx = 0-------------2.67
Ph = Rx = Px---------2.68
Rz=0; G + Rz - RD = 0---------2.69
 Rz = RD-G
Pz = -Rz = G - RD
Pz= [area ABCED – area C1CED] b

Pz= - [area ABCC1] b


--------------2.70

1. The horizontal component Px of the force on a curved surface equals the force of hydrostatic
pressure on the plane vertical figure DE, which is a projection of the curved surface on the
vertical plane.
2. The vertical component Pz equals the weight of the imaginary free body of the fluid ABCC 1.
This imaginary free body of the fluid we shall called "pressure body".
Let represent [area ABCC'] b = Go
Pz = - Go------ pressure body.
P x = Rx = P h
F= Ph2  Go2 -------------2.71
tg = Go/Ph -------------2.72

+
+

2nd Case of Cylindrical Surface:


Considering this case we can proceed as before to find Px and Py. Reasoning as before, we can
show that the horizontal component Px is given by exactly as the previous case. For the vertical
component Py=Go.
As can be seen, in this case the "pressure body " is real (i.e. lies in the real zone not the
imaginary zone of fluid) fluid, not an imaginary fluid. With this in mind the pressure body is
positive.

29
x
+
z

Ph=Px

3rd Case of Cylindrical Surface:


In this case the cylindrical surface ABC is divided into two at the points N into the component
real and imaginary zone. In this case summing the positive and negative obtain the vertical
component of the hydrostatic force.

1.1.1.1.1.4 C1.1.1.1.1.3 A
' +
1.1.1.1.1.2 N
Fy2
Fy2
1.1.1.1.1.1
- B
Py2

A special Cylinder case

4th Case of Cylindrical Surface (A rectangular gate the projection of which is a straight line)
This is a special case of cylindrical surface and the method of finding the force of hydrostatic
pressure is the same as in the case of cylindrical surface.
C' A

+
B

Fy
C

INFERENCE: In constructing the section of the pressure body i.e. the expression of the of the
vertical component hydrostatic force Py, one can be guided by the following:
The cylindrical surface ABC is the surface whose body is to be found. First fix the extreme ends
A and C; then draw a vertical from these points to the water surface; finally note the contour of
the pressure body A'ABCC' with the principle (law).

30
The cross-section of the pressure body (positive or negative) is the area between the two
verticals, the cylindrical surface ABC and the surface of the fluid (or their continuation). If the
pressure body does not wet the cylindrical surface, then we have negative body pressure;
however if the pressure body wets the surface, then the pressure body is positive.

Py

1.1.1.1.1.5 A
'

BUOYANCY
FLOATING BODIES: - ARCHIMEDES PRINCIPLE
Archimedes Principle (about 250BC): the force, which a fluid exerts on a body immersed in it
equals the weight of the fluid displaced by the body or when a body is placed (submerged) in a
fluid, it experiences an upward (upthrust) force which is equal to the weight of fluid the body
displaces.

F dF1 h1 h2
ds
dp
B
A D C dA

G dF2

Fig 2.15

Consider a body AB with volume V completely submerged in a fluid as shown in the diagram. It
is desire to find the force with which the fluid acts on the body. As we have already noted the
force is the resultant of all forces due to pressure acting on the surface element of the body. We
may find it by referring to the principle of forces on a curved surface.

R  R x2  R y2  R z2
But Rx=0; Ry=0

31
To find the vertical component Rz let us break the body AB into small volumes with cross
sectional area dA. Considering one of such volumes we see that from the top the body is acted on
by the weight of a column of liquid equals h1dA; and from the bottom by the weight of column
equals h2dA.
The difference in the pressure force on the strip is:
dFb=(h2-h1)dA …………………….2.76

It is therefore evident that the body will experience an upward (buoyancy) force equals
The sum of all elemental buoyancy force on whole body AB is
v V V
F   (h
v 0
2  h1 )dA   dV  V
0
The buoyancy force acts at the centre of gravity of the displaced liquid AB. The point D is called
centre of buoyancy. Generally the centre of buoyancy D does not coincide with the centre of
gravity of the body G.

Therefore the basic equation of floating bodies is

Rz=0
Fb-G=0 or V-G=0 …………………..2.78

BUOYANCY: Is the tendency for fluid to exert a supporting force on a body immersed in it. As
can be seen from 2.78 there can be three cases

1. Fb < G--- Body sinks and fall to the bed of the fluid where the reaction of the bed will
support to bring the body to equilibrium
2. Fb > G--- Body floats partially submerged in fluid (FLOATING BODIES)
3. Fb= G--- Body floats totally submerged in the fluid (SUBMERGED BODIES)

STABILITY OF SUBMERGED BODIES

i.e. Situation when Fb=G

(a) (b) (c)


Fb Fb
A G Fb
Fb
D
D C C
C C D
C D D

G
G G G

Fb

Fig 2.16 32
a) Stable equilibrium
b) Unstable equilibrium
c) Indifferent i.e. neutral equilibrium

Referring to the diagram above, the line passing through C and D is called axis of floatation O-O
Assuming that under normal circumstances the body floats when the point A is at the top (i.e.
vertical on O-O). If in this equilibrium state a force is applied (eg. wind blows) to rotate the body
through a small angle the forces Fb and G will constitute a restoration moment. In the opposite
case (i.e. if the point D lies lower than C) the body will never return to its initial position but
would be overturned.
Therefore the criterion of stability for submerged bodies is the relative positions of D and C. For
submerged body to be stable, i) the weight of the body G must be equal to the buoyancy force Fb
and ii) the centre of buoyancy D must always be above the centre of gravity C of the body.
Submarines are submerged bodies, which use balancing tanks to make Fb equal to G and
trimming tanks to bring the centre of buoyancy above the centre of gravity. In order to bring the
submarine on the free surface, the weight G must be reduced by pumping out water, such that
Fb > G.
Fig.a is a submerged body in a stable equilibrium. A body is said to be in a stable equilibrium , if
a slight displacement generates forces which oppose the change of position and tend to bring the
body to its original position. Fig. b is an unstable equilibrium. Fig c is said to be in a neutral
equilibrium.

SITUATION WHEN Fb>G: - Floatation when body is partially submerged

When Fb>G the body will be surfacing out of the liquid until Fb=G
O

G M

C C Fb

D D
G D1

Fb

Fig. a

33
O
O
θ

C C
M
l
d D D
D1
C

O O Fb

Fig. b

A floating body is a partially submerged body in a liquid, which is subjected to a buoyant force
that is equal in magnitude to the weight of the liquid displaced by the body. The condition for
angular stability of a body floating in a liquid is a little more complicated than for submerged
bodies. This is because, when the body undergoes an angular displacement about a horizontal
axis, the shape of the immersed volume in general changes, so that the centre of buoyancy moves
relative to the body.

Some basic terms in floating bodies


O – O – axis of floatation
W-L: - water line –the line of intersection of the free surface of the fluid with the body.
C- centre of gravity of the body
D – centre of buoyancy of the body when it is upright
D1 – centre of buoyancy of the body when body is rotated through a small angle θ
M- Metacentre – is the point of intersection of the axis of floatation and the vertical through D 1.
MC – metacentric height-the distance between the metacentre and the centre of gravity.
MD – metacentric radius: - distance between the meatcentre and the centre of buoyancy when
object is upright.
l – height of floating body
d – draft of floating body

The figures shown above represent floating bodies. Fig a represents bodies in equilibrium. The
net force on the body is zero so it means the buoyancy force Fb equals in magnitude to the
weight of the body. There is no moment on the body so it means the weight acting vertically
downwards through the centre of gravity C must be in line with the buoyancy force acting
vertically upwards through the centre of buoyancy D. Fig (*) (a) shows the situation after the
body has undergone a small angular displacement (angle of heel θ). It is assumed that the

34
position of the centre of gravity C remains unchanged relative to the body. The centre of
buoyancy D, however, does not remain fixed relative to the body. During the movement, the
volume immersed on the right side increases while that on the left side decreases; so the centre of
buoyancy moves to the new position D1. The line of action of the buoyancy force will intersect
the axis of floatation at the point M. For small angles of inclination, θ the point M is practically
constant in position and it is known as the metacentre. For Fig. (*) (a), the point M is above the
point C and the couple acting on the body in its displaced position is a restoring couple that tends
to restore the body to its original position. On the other hand in Fig (*)(b), the point M is below
the point C and the couple thus formed is an overturning couple and the original equilibrium
would be unsafe. The distance MC is known as the metacentric height and for stability of the
body, it must be positive (i.e.M above C). The greater the magnitude of MC, the greater is the
stability of the body. The magnitude of MC serves as a measure of stability of a floating bodies.
Since it is important that all floating bodies do not capsize in water, it is essential that we are able
to determine its stability before it is put in water.

Experemental Determination of metacentric Height.

C C

D
D D1

The experiment consists of moving a weight P across the deck through a certain distance x and
observing the corresponding angle of heel or roll θ. The shifting of the weight P through a
distance x produces a moment Px which causes the vessel to tilt through an angle θ. This
moment Px is balanced by the righting moment G x CM θ.

35
Thus Px = G x MC tan θ
Px Px cot 
MC  
G tan  G

It must be noted that the vessel, before the weight was moved, was in an upright (vertical)
position and G is the total weight of the vessel (including the weight P)

Since the point M corresponds to the metacentre for small angles of heel only, the true
metacentric height is the limiting value of MC as θ→0.

The metacentric radius DM = I/V0

Where I – second moment of area of the plane of floatation about centroidal axis perpendicular
to plane of rotation; V0 – immersed volume.

Periodic Time of Oscillation.

The displacement of a stable vessel through an angle θ from its equilibrium position produces a
righting moment (or torque).
T = G x MC x θ

This torque, when the force bringing about the displacement is removed, will produce an angular
acceleration d2θ/dt2
If I is the mass moment of inertia of the vessel about its axis of rotation, then

d 2 T G.CM . CM . .g
   
dt 2 I (G )k 2 k2
g
Where k – radius of gyration from its axis of rotation. The negative sign indicates that the
acceleration is in the opposite direction to displacement. The above equation corresponds to a
simple harmonic motion with the period given by:

 
 Displacement      k2 
t  2    2    2  
 Accelerati on   CM . . g 2   CM . g 
 k 

From above, it can be inferred that although a large metacentric height ensures improved
stability it produces a short periodic time of oscillation, which results in discomfort and
excessive stress on the structure of the vessel.

The Hydrometer

36
The principle of buoyancy is used to determine the specific gravity of liquids using a
hydrometer. The hydrometer consists of a bulb weighted at the bottom to make it float upright in
liquid and a stem of smaller diameter and usually graduated.
Let the hydrometer read 1.0 when floating in distilled water of specific gravity 1. The
corresponding weight of water displaced will be Voγw; where Vo is the volume of distilled water
displaced. In another liquid of higher (or lower) density, the hydrometer will pop up (or down)
by an amount Δh. If the stem of the hydrometer is of cross-sectional area a, the reduction (or
increase) in volume of fluid displaced will be a.Δh
Since the weight of the hydrometer is equal to the weight of the volume of fluid displaced in
each case
G = γwVo = γf(Vo-a.Δh)

 f Vo   wV o
h  
Vo
S.G f  1
f a f

+Δh
-Δh

Vo
V=Vo+aΔh V=Vo-aΔh

GENERAL EQUATION FOR STEADY FLOW OF FLUID (FOR A FILAMENT).


(BERNOULL‘S ENERGY EQUATION)

The general equation of steady flow of fluid can be obtained by applying the first law of
thermodynamics to a fluid filament.
The first law of thermodynamics states that for a steady flow of fluid, the external work done on
any system plus the thermal energy transferred into or out of the system is equal to the changes
of energy of the system; i.e.

Work + Heat = Change in Energy -----------(1)

37
Let us apply the first law of thermodynamics to the fluid system defined by the fluid mass
contained at time t in the control volume between the section 1-1 and 2 -2.This fluid system
moves to a new position during time interval dt.

2
P2
machine
v2

A2
1 v1
ds2
QH z2

ds1
P1 z1
A1
O O

During this short time interval, the fluid moves a short distance ds1 at section 1-1 and dS2 at
section 2-2.For steady flow, the weight of fluid entering at section 1 must be equal to that leaving
section 2; therefore
γ1A1dS1 = γ2A2dS2 = γAdS

In moving this short distance, work done on the fluid system by the pressure forces called the
flow work is given by:

Flow work = P1 A1dS1 – P2A2dS2 ---------------(2)

In addition to flow work, there ia a machine between sections 1 and 2 which either impact or
extract energy to (from) the fluid called the shaft work. If hm is the energy per unit weight of
fluid impacted to the fluid, then the shaft work will be given by

Workshaft = γ1A1dS1hm -------------------- (3)

hm is positive if the machine is a pump because it gives energy to the fluid but negative if it is a
turbine because it extracts energy from the fluid.
The heat transferred from an external source into the fluid within the time interval is:

Heat = γ1A1dS1QH ----------------------(4)

Where QH is the energy put into the flow by the external heat source per unit weight of flowing
fluid.
38
The energy of the fluid system at time t+dt is E2 equals the energy of the system at time t plus
the energy that flowed out of the system within the time dt minus the energy that flowed into the
system within the same time dt

E2 = E1 + ΔEout -ΔEin

Hence the change in energy ΔE of the fluid system within the time dt

ΔE = E2 – E1 = ΔEout – ΔEin -------------(5)

The energy ΔEin which enters at section 1 during the time interval dt is

ΔEin = γ1A1dS1(Z1 + U1/2g +I1)

I= is the internal energy per unit weight of fluid

Eout = γ2A2dS2 (Z2 +U22/2g +I2)

ΔE = Δ E out – ΔEin = γ2A2dS2 (Z2 +U22/2g +I2)- γ1A1dS1(Z1 + U1/2g +I1)

Now, applying the 1st law of thermodynamics, i.e. Work + Heat = Change in Energy

P1 A1dS1 – P2A2dS2 + γ1A1dS1hm + γ1A1dS1QH =


= γ2A2dS2 (Z2 +U22/2g +I2)- γ1A1dS1(Z1 + U1/2g +I1)

Note that γ1A1dS1 = γ2A2dS2 = γAdS, so dividing through by γAdS we shall obtain

P1 P2 u 22 u2
  hm  QH  ( Z 2   I 2 )  Z1  1  I1
  2g 2g
(6) General equation of steady flow
 P u2   P u2 
or... Z 1  1  1  I 1   hm  QH   Z 2  2  2  I 2 
  2g    2g 

Energy Equation for Steady Flow of Incompressible fluids.


Daniel Bernoulli’s Energy Equation

 p u2 p u2 
 z1  1  1  hm  QH  z 2 2 2   I 2 I 1  --------------(6a)
  2g  2g 

Turbulence and eddies during fluid flow produce fluid friction and these forms of kinetic energy
are eventually transformed into thermal energy. If in the system described above, there is no heat
transfer, then the effect of friction is to produce an increase in temperature so that I2 should be
greater than I1. A change in the internal energy of a fluid is accompanied by a change in
temperature and is equal to the external heat added or taken from the system plus the heat
generated by fluid friction; i.e

39
Change in internal energy per unit mass, Δi = i2-i1 = c(T2-T1) or
Change in internal energy per unit weight, ΔI = Δi/g = c/g(T2-T1) ----------(7)

Where c = specific heat of the fluid.


If hl is the fluid friction energy loss per unit weight of fluid, then we can write that

I2-I1= hl+QH ------------------------------(8)

which implies that the change in internal energy of fluid is brought about by the heat due to
friction and the heat introduced or taken away from system.
Putting eq. (8) into eq.(6a) we shall obtain

 p u2 p u2 
 z1  1  1  hm  z 2  2  2   hl -------------------(9)
  2g  2g 

Assuming further that there is no machine between the sections 1 and 2 within the system, then
equation (9) becomes

 p u2 p u2 
 z1  1  1  z 2  2  2   hl ------------------------(10)
  2g  2g 
Equation (10) is Bernoulli‘s Energy equation for a real fluid written for a streamtube

 p1 u12 p 2 u 22 
 z1    z2    -------------------(11)
  2g  2 g 
Equation (11) represents Bernoulli‘s Energy equation for an ideal fluid written for a streamtube.

For a real fluid of a stream, the equation becomes

 p u 2 p u 2 
 z1  1  1  z 2  2  2 
  2g  2g 

Worked example:
Water flows at 10m3/s in a 150cm diameter pipe; the head loss in a 1000m length of this pipe
is2o m. Find the increase in water temperature assuming no heat enters or leave the pipe. C for
water is 4187 N.m/kg(K)

Solution:

hl + QH = I2 –Ii = c/g(T2T1)

QH = 0

40
hl = (c/g) x ΔT

ΔT = (g/c)ΔT = (9.81 x 20) x 20 /4187

ΔΔT = 0.047 oK

41
HEAD LOSSES

Equation of steady uniform flow. Incompressible steady and uniform flow in pipes

Let τo τ be the shear stress on the surface of the pipe due to fluid flow and let us also assume that
the value is constant along the flow.
The aim is to find the relation between linear head losses and the force of friction.
To do this, let us consider a section of a pipe of length l enclosed in the plane 1-1 and 2-2 and let
the axis of flow direction be S.

p1/γ
1 hl

p1

p2/γ 2

β
1 G
τo
l
z1 p2
z2 S
2
2
0 1
0

The

Since flow is uniform, then according to Newton‘s 2nd law, the sum of all forces
projected in the direction of flow must be equal zero.
Let us now find all the forces on the fluid enclosed within the section 1-1 and 2-2.
1) Fluid‘s own weight
G = Alγ --------------------------------------------(HL-1)
Projection of this force on the S-axis
Gs = Alγ.sinβ --------------------------------------(HL-2)
Where A – cross section of pipe, β – angle of inclination of pipe to the horizontal.
But lsinβ = z1 – z2 and therefore

42
Gs = Aγ(z1 – z2) ------------------------------------------------(HL -3)

2. Force of hydrodynamic pressure


p1A – p2A ----------------------------------------------------(HL-4)
3. Frictional force acting on the fluid from the pipe boundary
Ff = τo. χ .l -----------------------------------------------(HL-5)
Applying Newton‘s law, we obtain;
Aγ(z1 – z2) + p1A – p2A - τo. χ .l = 0 --------------------(HL -6)
Dividing through by Aγ and re-arranging, we obtain;
(z1 + p1/γ) – (z2 + p2/γ) = (τ0 .χ l)/γA --------------------(HL -7)
It must be seen that (z1 + p1/γ) – (z2 + p2/γ) = hl ; therefore
hl = (τ0 .χ l)/γA -------------------------------------------(HL-8)
From whence
τ0/γ = RS ------------------------------------------------- -(HL-9)
Though the formula obtained is for the shear at the boundary, it can be shown that the formula is
true for any distance within the fluid. Therefore the formula can be written as;
τ/γ = R’S ---------------------------------------------------(HL-10)

τo

τ
ro
r

Shear stress distribution in a pipe of laminar flow

43
Newton’s law of internal fluid friction
The force of internal fluid friction, which arises in fluids as a result ofone layer slipping over
another layer is:
i) directly proportional to the velocity gradient
ii) directly proportional to the area A of surface of contact
iii) dependent on the physical properties of the fluid Newton‘s law can be written as
Ff = μA(du/dr) ----------------------------------------------(HL-11)
Ff/A = τ = μdu/dr -----------------------------------------(HL- 12)
Where τ – shear stress in the fluid; μ – coefficient of dynamic viscosity and
du/dr –velocity gradient or du/dr = tgθ tangent of the velocity curve (epure)

VELOCITY DISTRIBUTION IN A STEADY LAMINAR FLOW


Let us consider a circular pipe of radius ro. Within the fluid, let us cut an imaginary column of
radius r. The shear stress acting on the surface of this imaginary cylindrical column can be
computed by:
1. The basic equation of steady uniform flow
τ = γRS = γ(r/2) S ----------------------(HL-10)
2. Newton‘s law of internal friction in fluids
τ = - μ(du/dr) ---------------------------(HL -12)

dr A

r umax
C

dr
B
dA=2πrdr

Equating the two equations since both are defining one and the same quantity, we have:
γ(r/2)S = - μ(du/dr) --------------------(HL-13)

44

du   Srdr
2
Integrating,

 du  2 S  rdr
S 2
u r C
4
U sin g .the.boundary.condition ..when.r  r0 ,.u  0
S 2
C r0
4
S 2
u
4
r0  r 2 ...              ( HL  14)

Equation (HL-14) gives the velocity distribution in a laminar flow in pipes. As can be seen, the
S 2
velocity distribution is parabolic, with the maximum value U max  r0 is at r = 0 i.e at
4
the centre of the pipe.
LINEAR HEAD LOSSES IN A STEADY LAMINAR FLOW
Let us find the elemental flow rate of fluid passing through the elemental cross section dA in the
form of a ring with radius r.
dQ = udA = u.2πr.dr -------------------------------------(HL-15)
Substituting u from (HL-14) into (HL -15), we shall obtain:

dQ  S r02  r 2 2rdr
4
r  r0

Q   dQ  S  r02  r 2 rdr
2 r 0
 4 
Q Sr0  SD 4      Hagen..Poiseuille ' s.law.of .la min ar. flow.    ( HL  16)
8 128

M
Let 128
Then...Q  MSD...                  ( HL  17)
Note that M depends only on the nature of fluid.
Now let us try to express Q in terms of the mean velocity.
Q = v.A

45
 4
Sr0
Q 8  1
v   Sr02  U max . ------------(HL-18)
A r02
8 2
Conclusion: In a steady laminar flow, the mean velocity is half the maximum real velocity, which
occur at the centre of the pipe.

From equation (HL-18), v  Sr02 and substituting S =hl/l
8
 hl 2  hl 2
v r0  D ..        ( HL  19)
8 l 32 l
 l
hl  32 v      ( HL  20)..   Linear .head .losses.in.steady.la min ar. flow
 D2
Equation (HL-20), the following conclusions can be made.
In a steady laminar flow, the linear head losses hl
1. depend on the properties of the fluid.(ie μ, γ)
2. directly proportional to the mean velocity, v
3. Does not depend on the pipe roughness
Transforming equation (HL-20) as
 l  v 2.v  l v 2 64 l v 2 l v2
hl  32  32 . l  64 . .  .  f . ------(HL-21)
 D 2v g D 2 2.v Dv d 2 g Re D 2 g D 2g
Equation (HL-21) is the Darcy-Weisbach formula for head losses in pipes
Note from above that for laminar flow, the frictional factor f (skin friction) is given by
f = 64/Re

LINEAR HEAD LOSSES IN A STEADY TURBULENT FLOW


The Darcy-Weisbach equation in (HL-21) above is applicable for computing linear head losses
in a steady turbulent flow. However, since the velocity distribution in a turbulent flow is not
parabolic, and since the shear stress τo cannot be expressed by Newton‘s law of internal friction,
the frictional factor is not equal to 64/Re. In turburent flow the dependence of the frictional
factor on the Reynolds number is determined by the roughness of the pipe walls. This
dependence of the frictional factor on the Reynolds number was studied by J. Nikuradze (1932)
in his experiments on artificially roughened pipes with the interior surface covered with sand
grains of different grades uniformly distributed.

46
Velocity distribution in a turbulent flow

The scheme of Nikuradze‘s experiment consists of a long pipe with diameter D, a tap K to
regulate the flow and two piezometers placed at relatively long distance apart (l). There is also a
velocity measuring device to register the velocity in the pipe.

P1
hl P2
K

Scheme of Nikuradze‘s experiment


With the help of the tap K, the velocity in the pipe can be changed and with every change in the
velocity, the piezometer readings change, indicating the dependence of head loss on velocity
within the segment of length l. The frictional factor can be computed by;
hl D
f  .2 g. 2
l v
The results of the experiments are presented in a graphical form, where the frictional factor f is
plotted as ordinate and Re as abscissa.

47
Scan the Moody diagram and place here

The result presents a series of curves obtained for pipes with various relative surface
roughnesses, characterized by the dimensionless quantity ε = k/D. Here k – is the absolute
roughness height, D – diameter of pipe.
Nikuradze‘s result make it possible to understand the question of head losses and make the
following inferences:
i) The frictional factor in the formula (HL- 21) generally depends the relative roughness
and the Reynolds‘ number.

48
ii) There are particular cases of flow, when f depends only either on the relative
roughness or only on the Reynolds‘ number.
iii) There are combinations of f and Re for which the proportionality hl ~ vm, where the m
take a definite value of m =1; m =1.75; m = 2.
iv) The physical difference between laminar flow and turbulent flow is indicated by the
change in the relationship of f to Re near the critical Reynolds‘ number of 2100.
v) The laminar flow regime is characterized by a single curve given by the equation f =
64/Re for all surface roughness. This shows that the head loss in laminar flow is
independent of surface roughness and that hl ~ v.
vi) In turbulent flow, a curve of f versus Re exists for many relative roughness
ε =k/D, and the horizontal aspect of the curve confirms that for rough pipes,
the roughness is more important than Re number in determining the
magnitude of the frictional factor.
vii) At high Re, the frictional factors of rough pipes become constant and
dependent wholly on the roughness of the pipe and independent of the
Reynolds‘ number. Here, hl ~v2
viii) the series of curves for the rough pipes diverges from the smooth pipe curve
as the Reynolds number increases. In other words, pipes that were smooth at
low values of Reynolds‘ numbers become rough at high values of Reynolds‘
numbers. This can be explained by the thickness of the viscous sublayer
decreasing as the Re number increases, thus exposing smaller roughness
protuberances to the turbulent region and causing the pipe to exhibit the
properties of rough pipe.
From Nikuradze‘s results, one can see that the whole flow region can be divided into three
zones;
1. The First Zone: (Re < 2300) is the field of laminar flow. Here curves for pipes with
different degree of roughness merge into a straight line, which is a graphical
representation of Hagen-Poiseuille equation (HL -16). Here, the Reynolds‘ number is
small,< 2300 and the head loss hl do not depend on the roughness and it is proportional
the velocity ie m=1.

49
2. The Second zone (2300,Re<40000, zone of unstable flow). It is the zone of transition
from laminar flow to turbulent flow.
3. The Third zone (The field of turbulent flow) It is the field right of the vertical B. This
zone can be divided into three sub-zones:
a) The zone of smooth pipes. It is the line 3 -4. In this zone, head losses are
proportional to the velocity raised to the power 1.75. The frictional factor is given
0.3164
by the Blasius equation as f  1
. As can be seen, in this sub-zone the
4
Re
frictional factor depends only on the Reynolds‘ number. The reason for this is
that for the Re in this zone, the thickness of the laminar (viscous) sub-layer is
sufficient to cover the wall roughness and make all the protuberances of
roughness to be immersed in the viscous layer, making the pipe walls to behave as
if they were smooth.
b) The pre-quadratic field or zone of rough pipes. It is the field between line 3-4
and line CD. Here the frictional factor and the head losses depend on both the
Reynolds‘ number Re and the relative roughness ε.
c) The zone of quadratic flow. This zone lies to the right of the line CD. Here head
losses are proportional to the square of the velocity, ie hl~v2 (m=2). In this zone,
the frictional factor f does not depend on the Reynolds‘ number and is a function
only of the relative roughness ε.

Linear head losses in a steady & uniform turbulence flow in the quadratic field.
The basic equation of steady uniform flow was obtained from (HL-) to be:
τ = γRS
In laminar flow, τ can be expressed as τ = μ(du/dy) but in turbulent flow, this is not applicable.
In order to express τ in the above equation for turbulent flow, the frictional factor is expressed by
the empirical factor based on experiment and is true for both laminar and turbulent flow and is
based on the assumption that the force resisting the flow (τoPL) per unit wetted area is
proportional to the square of the velocity head. i.e τo = K (v2/2g) where K = const.

 o f v2
 ---------------------(HL-22)
 4 2g

50
8g  o 8  o
v .  .
f  f 
o f
 v --------------(HL-23)
 8

Equation (HL-23) is the basic relation between the frictional stress τo and the friction factor. In
this fundamental equation relating wall shear to the friction factor, density and mean velocity of
flow; it is apparent that with f – dimensionless, then √(τo /ρ) must have the dimension velocity.
This quantity √(τo /ρ) is known as the friction velocity v* which is related to the friction factor
and the mean velocity by:
o f
v*   .v -------------------------------------(HL-24)
 8
This friction velocity embodies only the wall shear and the fluid density, therefore it is defined
by the equation only whatever the flow regime or whatever the boundary texture.

LOCAL (MINOR) HEAD LOSSES IN A STEADY TURBULENT FLOW

51
(Losses in pipes other than pipe friction)
Obstruction in pipelines such as partially opened valves, bends, taps, tees, diffusers, contractions
and expansion of cross-sections, entry into pipes, exits from pipes are called local hydraulic
resistance and they bring about energy losses called local head losses. They are so called because
these losses, though occur in certain length of the pipe, but since this length is so small compared
with the entire length of the pipe, are deemed to be concentrated at the point, where the
resistance is located.
All local losses are caused by increase in turbulence due to boundary layer separation.

b c
h

a d
L = (5 to 10)h
Obstruction
Whirlpool area, A
Causes of local head losses

When a turbulent stream flows over an object, there is a separation of the transit stream from the
walls of the stream. As a result, there arises areas, A, where fields of vortex of random fluid
motion appear. The main transit flow is separated from the field of vortex flow (whirlpool) by an
interface abcd. . The area A is characterized by reverse eddies flow, in which the velocity
distribution varies substantially with negative velocities at certain places. The end of the
whirlpool area is characterized by high velocity and pressure fluctuations. It is also characterized
by great deformation of the velocity diagram. After this section, a gradual attenuation of
velocity and pressure fluctuation takes place and there is a gradual leveling of the velocity
diagram until it takes the form of a steady uniform turbulent flow at some distance from the
obstruction.
Through the interface abcd and as a result of fluctuations of the transverse velocities, there is an
interchange of fluid between the whirlpool area and the main transition stream. As a result of the

52
interaction, the turbulent shearing stresses acting along the interface abcd become relatively big
and therefore causes relatively large head losses.

Local head losses due to sudden expansion of pipe

A
1 a 2

v2
d p1
p2
v1
D

A1

1 a 2
A A2
Sudden expansion of pipe

The figure above shows a fluid flowing from a pipe of diameter d into a pipe of greater diameter
D. From the first pipe, the fluid within the segment 1-1 to a-a (as a result of the presence within
the fluid frictional forces) expands and at section a-a fills completely the cross-section of the
pipe. Within this section, there is separation of fluid from the pipe walls and there is the
formation of whirlpool area, A. From section a-a there is a gradual leveling of the velocity and
pressure until at section 2-2 a normal steady uniform flow is established. Within sections 1-1
and 2-2 therefore occurs a local head loss due to the expansion of the pipe.
The magnitude of the head loss due to the expansion can be computed using the Bernoulli‘s
Energy equation and the momentum equation applied to the section of the fluid within 1-1 and 2-
2.
Applying Bernoulli‘s energy‘s equation to sections 1-1 and 2-2;

53
p1v12 p 2 v 22
z1   z 2    h je
 2g  2g
--------------(HL-25)
 p v2   p v2 
h je  H 1  H 2   z1  1  1    z 2  2  2 
  2g    2g 
To simplify the derivation, let us assume the pipe to be horizontal. In that case the above
equation becomes:
 v12 v22   Fp1 Fp 2 
h je         -------------------------------(HL-26)
 2 g 2 g     
To find the value of Fp1 – Fp2, we apply the momentum equation to the fluid within 1-1 and 2-2.
ρQv2 – ρQv1 = To(s) + Gs + Fp(s) + Rs ------------------------(HL -27)
To(s) – the projection in the direction of flow of frictional forces To. Since the length of the
expansion section is relatively very small, the value of To(s) is also very small and therefore can
be neglected. An assumption which is fairly true.
Gs – is the projection of weight of fluid in the direction of flow, which is zero because the pipe is
horizontal.
Fp(s) – sum of projections on the s-axis hydrodynamic force
Rs - = R where R is the pressure of the vertical wall 1-1 in the form of ring on the fluid.
Fp(s) + Rs = (Fp1 –Fp2) + R = (Fp1 + R) – Fp2 --------------(HL-28)
Fp1 + R ~ p1A2 a very credible assumption. -----------(HL-29)
Substituting (HL-29) and (HL-28) into (HL-27), we shall obtain:
ρQ(v2 – v1) = p1A2 – p2A2 -------------------------------(HL-30)
Which eventually gives us:
Fp1  Fp 2

Q
v2  v1    v2  v1 v2 --------------(HL-31)
 A2 g  g 
Substituting (HL-31) into equation (HL-26), we have
v12 v22 2  v2  v1  v 2  v22  2v22  2v1v2
hje    . v2  1 ..      ( HL  32
2g 2g 2  2g  2g
2
v12  v22  2v1v2  v1  v2 
hje              ( HL  33)
2g  2 g 

54
2
 v v 
hje   1 2               ( HL  34)
 2g 

(v1-v2) –lost velocity


Equation (HL-34) can be transformed by bringing v1 as:
2 2 2
 v v   v  v12  A  v2
hje   1 2   1  2   1  1  1 --------(HL-35)
 2g   v1  2g  A2  2 g

By representing (1 –A1/A2)2 = ξje then the head loss due to expansion can now be written as:
v12
hje   je ---------------------------------------(HL-36)
2g
Where ξje – resistance coefficient for sudden expansion of pipe.
Similarly, if v2 is brought outside the bracket, we shall obtain
v 22
hje   !
je where ξ!je = (A2/A1 – 1)
2g
Local head loss when a pipe discharges into a reservoir of big surface area.
In this case, we have a specific case in which A2 is by far bigger than A1. From equation (HL-
35), we observe that the resistance coefficient ξje =(1 –A1/A2)2~1 since A1/A2~0. In this case
equation (HL-36) becomes
v12
h je 
2g
SUDDEN CONTRACTION IN PIPES
The diagram below shows the most sudden contraction in pipes. Conditions of flow at the change
of diameter is as follows:
A fluid particle M moving along the wall ab must at the point b change its direction in the
opposite way. As a result of its inertia, the particle M will break away from the wall and follow
the path bef creating a ring-like whirlpool area A. In the region of the whirlpool A, we can
distinguish two sections of the main transition flow; a) contracting fluid before the vena
contracta C-C and b) the expanding flow after the vena-contracta C-C.
Experiments show that the local head loss in the section of contracting fluid before the
contracted section is relatively small due to the fact that velocity fluctuation in this region always
reduces. Apart from this, the length of the section is relatively small
55
(about 0.5D2). Therefore the local head loss is generally concentrated at the expanding section.

E.L
hj
H.G.L

2
3

W1
c
1
M
A
b W2
a f
D1 e
wc
D2
e
a f
b
M A

c
L

1 2 3

With the above explanation, one can agree that the head loss in sudden contraction could be
calculated by the same expression as that in the sudden expansion as
2
 v  v2 
h je   1  but instead of v1, we shall use the velocity in the contracted section vc
 2g 
Vc = Q/Ac where Ac is the cross-sectional area of the vena contracta, which can be found as: Ac
= ε.A2, where ε = Ac /A2 is the contraction coefficient.
2 2 2
 v v  v  v2  A  v2  1  v2
2

hjc   c 2    c  1 2   2  1 2    1 2 ------(HL-37)


 2g   v2  2 g  Ac  2g    2g

v 22
hjc   jc ----------------------------------------(HL-38)
2g
Where ξjc = (1/ε -1)2 – resistance coefficient of contraction.
Generally local head losses due to contraction may be computed by formula (HL-38)
v2
hjc   jc
2g
But ξjc must in every case be found experimentally or be given.

56
Other cases of local head losses
Abrupt changes in fluid direction such as bend or elbow, gates valves, tees etc. are local
resistances. As shown above, the local head losses hj are generally all concentrated in section
after the contracted section where we have the expanding flow. Therefore in every case, we can
find the local head loss using the general formula as:
v2
hj   j -----------------------------------(HL-39)
2g
However, in each case, the resistance coefficient ξj must be determined experimentally or could
be obtained from reference books.

FLOW IN PIPES
Fluid flow in closed and completely filled pipes is widely used for transporting various fluids
(water, oil, solutions, e.t.c) both in experimental research laboratories and in industry. Pipelines
may be classified as simple and complex, according to the hydraulic circuit. A simple pipeline
has no branches; the transported fluid moves directly from intake to the point of discharge.
Complex pipelines consist of a main pipe and a number of smaller ones branching off from it.
Pipeline could also be divided into those with constant flow and those with variable flow.
Transportation of fluids in pipes is generally forced under a certain given pressure, during which
time the fluid exerts a corresponding pressure from within the pipe that tends to disrupt the walls.
Pipes used under pressures below atmospheric (suction pipelines) must be rigid enough to
withstand flattering induced by the pressure difference in the pipes and the atmosphere.
The basic equations for hydraulic design of pipelines are:
1- the continuity equation; Q = v.A = const
2- Bernoulli‘s energy equation or its modifications
P1v12 P2 v22
z1    z2    hf
 2g  2g
3. Equations for determination of head losses (linear and local).

Complex head losses. Total resistance coefficient


Consider a pipe line of diameter D and length l with a number of local resistances as shown in
the diagram below. We shall assume that the distances between these

57
2
1 1
D,L
T

resistances are relatively large (> 20D – 30D) and as a result the effect of one resistance on the
other is totally absent.
The total head loss between sections 1-1 and 2-2 is given by:
hf = hl +∑hj ------------------------------------(HL-40)
a) The sum of local head losses is given by:
∑hj = hj el +hj v +hj e
Where hj el, hj v and hj e are local head losses in the elbow, valve and expansion respectively.
But the value of hj el = ξel (v2/2g); hj v = ξv (v2/2g); and hj e = ξe (v2/2g)
Therefore ∑hj =( ξel + ξv + ξe)v2/2g or in a more general form
∑hj = v2/2g∑ξj ----------------------------------(HL-41)
b) Linear head loss between sections 1-1 and 2-2 is:
hl = f (l/D).v2/2g and if we let f.(l/D) = ξl
then we have hl = ξl .v2/2g --------------------------------------(HL–42)
Substituting a) and b) into eq. HL-40 , we shall obtain
hf = ξl.(v2/2g) + ∑ξj.(v2/2g) = (ξl + ∑ξj).v2/2g
Representing (ξl + ∑ξj) = ξf --Total resistance coef
hf = ξf .(v2/2g) ----------------------( HL-43) --Total head loss.

The concept of “long” and “short” pipes


If a pipe is long enough so that the value ∑hj in comparison with the value hf could be neglected
or in order words, hf ≈ hl, then such a pipe is called ―long‖. In general, if ∑hj is less than 5% of
hf then pipe is referred to as ―long‖.
Short pipes

58
a). Flow (efflux) into a reservoir
Consider a steady flow of fluid from reservoir A into reservoir B through a pipe of diameter D
and length L. The difference in level of fluid in A and B is Z and we need to
Po
Q

▼=constant
1 1

E
VA=0 Z
p

αv*v/2g Q

Po
▼=constant
0 2
T 2
F
A
VB

v
B
D,L

find the flow rate through the pipe. To determine the flow rate, let us choose the cross-sections at
1-1 and 2-2 and apply Bernoulli‘s energy equation. For these sections, it is known that PA
=PB=pat and vA ≈ vB ≈ 0. Let the datum pass through the section 2-2.
Now applying the energy equation we have:
p1 v12 p2 v22
z1    z2    hf --------------------HL-44
 2g  2g
But z1 = Z; v1 ≈v2 =0; p1=p2 =pat and we assume α =1.0 Putting these values into eq. 44, we
shall obtain;
Z = hf -----------------------------HL-45
As can be seen, in a flow into a reservoir, the head difference Z is spent on head losses in the
pipe system from A to B. These head losses are made up of local and linear head losses in the
pipe system. Now let us express the head losses in terms of the velocity in the pipe
hf = ξf.(v2/2g)
1
Therefore the velocity in pipe v  2 gZ and hence the flow rate by the continuity
f

equation is given by:


D 2 1
Q  v. A  . . 2 gZ  C d . A. 2 gZ
4 f

59
Where Cd = 1/√ξf – discharge coefficient of the pipe system.

b) Flow (effluent into the atmosphere

1 1

Va≈0 P

h
2
A
F
v Pa
0 0

D,L 2

Applying Bernoulli‘s energy equation to sections 1-1 and 2-2


p1 v12 p2 v22
z1    z2    hf
 2g  2g
and noting that vA ≈ 0; vB ≈ 0; v2=v; p1=p2=pat; z1 = h; z2 = 0; α ≈ 1.0, we shall obtain:

h
v2
 hf 
v2
 f
v2

v2
1   f 
2g 2g 2g 2g
Therefore when flow is discharged into the atmosphere, the head difference is spent in
overcoming the head losses in the system and creating a velocity head at the exit into the
atmosphere.
1
From the above, v  2 gh
1  f

D 2 1
Q  v. A  . 2 gh .....or
And the rate of flow becomes: 4 1  f
Q  C D . A. 2 gh

The Siphon: a special case of flow in a simple short pipe.


The siphon is a curved pipe, so arranged that liquid flows up through it above the source and
then down. The characteristic feature of the siphon is that part of the pipe is higher than the
source of the fluid.

60
2
n
Pc Pd

n V

D,W,L
H’
Pa T
Q H’’
2
1 1
0 0
Pa
Z
A Q

3 3

It the pipe in someway is filled with the fluid and inserted into the vessel, the fluid will begin to
flow from A into B. To find out why the fluid flows, let us consider any section in the pipe e.g
section n-n, which is at a height h‘ from the vessel A and h‖ from vessel B.
The pressure in section n-n from the left side is given by:
pc = pa + (- γh‘)
The pressure in section n-n from the right side is given by:
pd = pa +(- γh‖)
As can be seen, pc is greater than pd since h‖>h‘ so the fluid flows from A to B.
To determine the flow rate, Q we apply Bernoulli‘s equation to sections 1-1 and 3-3.
p1 v12 p3 v32
z1    z3    hf
 2g  2g
Z1=0; v1≈0; z3=-z; v3 ≈0; p1=p3=pa; α≈1.0
Substituting into the above equation, we shall obtain:
v2
z  hf   f
2g
1
v 2 gz
f
D 2 1
Q  v. A  . . 2 gz
4 f

61
The most distinguishing feature of the siphon is the presence of vacuum throughout the pipe T.
The maximum value of this vacuum is at the highest point of the pipe; i.e at section n-n. To find
the maximum vacuum, we can apply the energy equation to the section 1-1 and n-n.
p1 v12 pn vn2
z1    zn    h fn
 2g  2g
Where z1=0; zn=h‘; p1=pa; (αv12/2g)≈0; (αvn2/2g) =(αv2/2g), where v is the velocity in the
pipe. Therefore
pa pn v2 p v2 v2
 h'    h fn  h '  n    fn
  2g  2g 2g

pa  pn
Or

 h  1   fn
'
  v2
2g
-----------------(HL-46)

Since (pa-pn)/γ = (hvac)max

(hvac)max = h’+ (1+ξfn)(v2/2g)

By using the equation (HL-46) we can find value of the vacuum at any section of the pipe. In this
case, h‘ must be the height of the section above the vessel A and ξfn should be computed to the
point in question. As can be seen from the equation, (hvac)max depends mainly on the height h
above the vessel A. As h gets bigger, (hvac)max also gets bigger. If however, (hvac)max becomes
greater than a certain value, there will be a break in flow and the siphon will cease to function.
To avoid a break in flow, the following conditions must be satisfied:

(hvac)max ≤ (hvac)permissible

(hvac)permissible for water = (6-7)m water column.

The theoretical value is 10.3m water column

The suction pipe of pumps

Suction pipe of pumps is that pipe which takes fluid from the reservoir or sump. This pipe, like
the siphon is also characterized by the presence of vacuum throughout the pipe. The maximum

62
value of the vacuum is at the pump i.e section 2-2. The vacuum pressure at any section can be
obtained by applying Bernoulli‘s equation to sections 1-1 and the point of interest.

2
V

Pa

a
1 1

The value of the pressure at the pump‘s centerline is:


(hvac) pump = a + (1+ξf).v2/2g
If the value of hvac becomes too big, there could arise a break in flow or cavitation will set in,
which can reduce the efficiency of the pump or damage the blades of the pumps by erosion.
Generally the permissible vacuum in a pump should be:
(hvac)permissible ≤6.5 m of water column.
Hence apermissible =(hvac)permissible - 1+ξf).v2/2g

Long Pipes
In the case of long pipes, local head losses are neglected due to their smallness in comparison to
the linear head losses. The main equation used in long pipes is the Darcy Weisbach or the
Chezy‘s equation or any other head loss equation.
fl v 2 fl Q 2 fl
hl  .  . 2  K .Q 2 ....where..K  or
D 2 g D A .2 g 2 g.D. A 2
Q2 hl Q2
Q  v. A  AC RI  K I   I   hl  .l
K2 l K2
Series and parallel connections of pipes

63
1. Series connection: In series connection, the same flow passing through all the pipes are the
same, but the head across each pipe is different.
A

Hl1 B
Hl2
Hl3

The head loss within the section A and B will be given by:

(hl)AB = hL1 + hL2 +hL3

Thus in series connection, the total head loss is the sum of the head losses of the individual pipes.

Parallel Connection

The diagram below shows a parallel connection of pipes with piezometers have been connected
at points A and B. The loss from A to B can be written as:

(hl)AB = HA – HB

Where HA is the head at A and HB is the head at B. On the other hand we can write the following
true statements:

hL1 = HA -HB

hL2 = HA -HB

hL3 = HA -HB

Therefore (hL) AB = hL1 = hL2 = hL3 = HA -HB --------------(HL-47)

As can be seen, the head losses across all pipes in a parallel connection are the same.

But the head loss can be written as hl = (Q2/K2).l. Therefore the above equation (HL-47) can be
written as:

Q12 Q22 Q32


hl AB  .l1  .l 2 
K 12 K 22 K 32

64
hl AB
Hence Q1  K1 -----------------------(i)
l1

hl AB
Q2  K 2 -----------------------(2) ---------------(HL-48)
l2

hl AB
Q3  K 3 ------------------------(iii)
l3

And also Q = Q1 + Q2 +Q3 -----------------------(iv)

Therefore if the flow Q is given and the characteristics (i.e D, l) of the pipes are given, then we
have a system of four equations (i, ii, iii, iv) with four unknowns
(Q1; Q2; Q3; hlAB). By solving the system of equations, all the unknowns can be determined.

THE PROBLEM OF THREE RESERVOIRS


In the network of water supply systems, for the purposes of balancing flows, reservoirs are
usually interconnected with the pipes meeting at a common junction and coming from reservoirs
of different surface levels.
The diagram below shows three reservoirs (I, II, III) connected together by pipes of D1,l1; D2, l2;
D3l3. Let us denote  1  2  3 the levels of the heads in I, II, and III respectively and we assume
that these levels are constant. From the levels given, it is obvious that flow direction in I is
known and that in II is also know; however it is not obvious the direction of flow in pipe 3. We
are to find the direction of flow in pipe 3 and the flow rates Q1, Q2, Q3

 1= const
 3 = const
0 I3
I I1 I2 2 =const
III

Q1 1 II
Q3

3
D3 , l 3 Q2

D1, l1 D2, l2 2

65
1. Direction of flow in pipe 3.
Let the piezometric surface at the common junction be  o as shown on the diagram. The
following arguments can be put forward.
i) If reservoir III is fed by reservoir I, then liquid will flow up in pipe 3 and this will
mean that  o >  3
ii) If reservoir III feeds reservoir II, then in pipe 3 liquid flows downwards and it
implies that  o <  3
iii) If there is no flow in pipe 3, then it means  o=  3
To find out which of the three assumptions is true, we use the 3rd assumption that there is no flow
in pipe three which will imply that all flow from reservoir I goes into reservoir II.
With this assumption, we find the flows in pipe 1 and pipe 2.

hl1 1   0 1   3
Q1  K1  K1  K1
l1 l1 l1

hl 2 0  2 3   2
Q2  K 2  K2  K2 ------------(HL-49)
l2 l2 l2
From the above equations, the values of Q1` and Q2 can be calculated.

If Q1 =Q2

Then it means Q3 is really equals to zero and  o=  3

If Q1 > Q2

then it means the difference in flow goes into reservoir III and therefore reservoir III is being fed
by reservoir I ( i.e.  o >  3)

If Q1 < Q2

Then the difference is made up by the flow from reservoir III into reservoir II
i.e  o <  3

2. Values of Q1, Q2, and Q3


Once the direction has been established, it is now easy to write out the following equations
assuming that reservoir III feeds reservoir II.
Q2
1   0  12 .l1 ------i)
K1
Q32
3  0  .l3 -------ii) ---------------(HL-50)
K 32
Q22
0  2  .l 2 -------iii)
K 22
Q1 + Q3 =Q2 ---------iv)

66
By solving the system of equations given above, all the unknowns can be found.

Equivalent length and equivalent pipes

At times for simplicity of solution and for convenience, losses in pipe fittings are expressed in
terms of equivalent length, which is the length of a straight pipes which will give an equal head
loss for the same discharge as that of the fitting. The equivalent length is given by:
v2 Leq v 2 d
KL  f. and therefore Leq  K L --------(HL-51)
2g d 2g f
Therefore in computation of head losses, the effective length of pipe is obtained by adding to the
actual length of pipe the equivalent lengths for different fittings in the piping system.

Equivalent Pipes
Two pipes of characteristics (L1, d1, f1) and (L2, d2, f2) are said to be equivalent if for the same
fL fL
discharge the head loss across both pipes are the same. i.e. 1 5 1  1 5 2
d1 d2
For pipes in series, since the head loss is the sum of the individual head losses, then for a set of
pipes (L1, d1, f1), (L2, d2, f2), (L3, d3, f3) connected in series, the equivalent pipe
(Le, de, f3) can be given by
f e Le fL fL f L
5
 1 5 1  1 5 2  3 5 3  ......... ------------(HL-52)
de d1 d2 d3
For a set of pipes in parallel between two points, the equivalent pipe is given by:
1 1 1 1
 d 
5 2
 d 
5 2  d 
5 2  d 
5 2
 e
   1
   2
   3
         ------(HL-53)
 e e
f L  1 1
f L  2 2
f L  3 3
f L

TRANSMISSION OF HYDRAULIC POWER THROUGH PIPES


Suppose we want to transmit power by means of water under pressure through a certain distance
by pipes, it is clear that the power that can be transmitted will depend on: i) the flow rate, Q and
ii) the total head (or pressure) available. Let the velocity in the pipe be V

hl

H1
H
Hav
V

Inlet Outlet
D, L, f
67
For the pipe shown above, the total head at the inlet to the pipe is H and the head loss due to
friction is hl. The total head available at the outlet of the pipe is given by:
fl v 2
H av  H  hl  H  . ---------------(HL-53)
D 2g
Efficiency of transmission is defined as the ratio of energy head at outlet head to the energy head
H  hl
at inlet and is given by:  transm  -----------(HL-54)
H
The power available at the outlet is given by; P=γQHav.=γQ.(H-hl) = γQ{H-(fl/D).v2/2g}
Expressing the flow rate Q = v.A = v.(πD2/4), the power available at the outlet is:
D 2  fl v 2  D 2  fl v 3 
P  QH av  QH  hl    .v  H  .     Hv  . 
4  D 2g  4  D 2g 
For the power to be maximum, (dP/dv) = 0

dP D 2  3 fl 2 
 . H  .v  0
dv 4  2 gD 
-----------------------------(HL-55)
fl v 2
 H  3. . 0
D 2g
H
 H  3hl  0...or...hl  ----------------------------------(HL-56)
3
The power transmitted is maximum when the head loss due to friction is one-third the total head
supplied. Hence the efficiency of power transmission from eq.HL-54 is equal to:

H  hl H  3 2
H
Transm    or in percentage = 66.67%.
H H 3

HYDRAULIC ANALYSIS OF PIPE NETWORKS


Pipe systems with which the practicing engineer has to deal with may not be as simple as those
we have already dealt with.
Cities‘ water distribution systems are ordinary design to adequately satisfy the water
requirements for a combination of domestic, commercial, industrial and fir e-fighting purposes.
The system should be capable of meeting the demands placed on it at all times and at satisfactory
pressures.
Distribution systems may be classified as grid system, branching systems or a combination of
these. The configuration of the system is dictated primarily by street patterns, topography, degree
and type of development of the area, and location of treatment and storage works.

The grid system is usually preferred to a branching system, since it can furnish a supply to any
point from at least two directions. The branching system does not permit this type of circulation,
since it has numerous terminals or dead ends
A city‘s supply network is often complex and to design an effective distribution system
throughout the city, it is necessary to calculate the pressure and discharge at all points in the
network. Irrespective of how complex a city‘s network might be, it can be broken down into a

68
convenient number of loops or rings. The theoretical requirements that must be satisfied in all
loops are:

Branch line

Flow Main supply line

Branching System Grid system

Combined grid and branching system

1. Continuity relationship must be satisfied; ie; the flow into a junction of the network must be
equal to the flow out of the junction. (in other words, the quantity of flow reaching any junction
must be zero).
2. The algebraic sum of the head loss between any two junctions is the same regardless of the
path in the series of pipe taken to get from one junction point to the other. The algebraic sum of
the head losses around any closed circuit must be zero.
3. The flow in each pipe must satisfy the appropriate head loss equation (e.g Darcy-Weisbach
equation).
The head loss against friction and other local losses in each pipe is expressed in terms of the
flowrate through the Darcy-Weisbach equation as;
fl v 2 fl 16Q 2 8 fl
hl  .  . 2 5  Q2 v=Q/A = 4Q/πd2
d 2g 2g  d g d
2 5

The above equation can be generalized in the form:

hl = KQn
where K and n are assumed constants and Q discharge through the pipe.
Pipe network problems are generally solved by trial and error method. The most practical and
widely used method of flow analysis is that of successive approximations developed by Hardy
Cross. The method consists of making a trial guess for the values of flow in every circuit of the
network which satisfies continuity equation. These trial guesses are improved by employing
corrections. The process being repeated until successive corrections become insignificant.
Two approaches are available; the head balancing and the quantity balancing.

69
HEAD BALANCING METHOD

Q QR

QL 1 2

3 4

Let QR and QL be the first guess and QR* and QL* be the correct values for a loop in a complex
pipe network. We can then write:
QR* = QR + ∆Q and
QL* = QL - ∆Q
Where ∆Q is the error in the assumed values. The head loss along either route from junction 1 to
junction 4 must be the same according to the theoretical requirement 2 above. Ie.
∑ hL 1→ 2 → 4 = ∑ hL 1 → 3 → 4

∑ K(QR*)n = ∑ K(QL*)n

∑K(QR + ∆Q)n = ∑K(QL - ∆Q)n

Expanding the above by the binomial theorem;

 K Q n
R  
 nQRn 1 Q   K QLn  nQLn 1 
Q  
 KQ   KQ  
n
R  KQ   KQ
n
L
n
R
n
L

n KQ   KQ 
n 1
 n 1
Q  QL
R
n  KQ . L
  KQ n 1
R
R
n 1
L .

Q R  QL

Q  
 hR   hL ......or....Q    h..( paying ..attention ..to..sign)
 h h  h
n ..(ignoring ..signs )
n  R   L 
 QR QL  Q

The sign convention for flow around the network loop is as follows: clockwise direction is
considered positive (+ve), whiles anticlockwise is negative (-ve).
Therefore ∆Q for any give loop can be computed and then added to the original values and the
process repeated until insignificant values of ∆Q are produced.

The procedure for the computation is as follows:

70
1. Assume any reasonable distribution of flow, which satisfies continuity requirements at
each junction and for the whole loop
2. Calculate in each pipe the head loss hL= KQn. The head loss is positive (head drop) if
the flow is in the fixed direction and negative (head rise) if the flow is in the opposite
direction.
3. Sum up the head losses algebraically, i.e. ∑hL = ∑KQn
4. Calculate the sum of hL/Q; ie; ∑n.(hL/Q)

5. Evaluate Q  
h
h
n
Q
6. Make the necessary adjustment QR* = QR + ∆Q
QL* = QL - ∆Q
7. Repeat the whole calculation with the new value QR* and QL* and calculate the new ∆Q‘
8. The calculation ends when ∆Q is very small, the magnitude which will be determined by
the accuracy being sought.
9. For common members among two loops, both (∆Q) corrections should be made one for
each loop by adding them algebraically.

Some examples. (scan some examples)

The computation can be quite tedious and laborious when the loops are more than two, however,
there exist computer programmes, which allows a network design to be input in response to a
number of prompts. The programme ‗HARDY C‘ and the EPANET are examples of such
programmes.

71
OPEN CHANNEL FLOW

Any fluid flow that is bounded by the sides and the bottom but free on the surface and subjected
to a constant (usually atmospheric) pressure is referred to as open channel
Unlike in pipe flow, where the pressure difference between two points within the pipe is the
driving force of flow, in open channel flow, gravity force is the main driving force of flow and
therefore elevation changes (drop) in the direction of flow is very essential.

As in highway drains
As in urban storm drains As in hydrotechnical pipes

As in river or stream flows


As in sanitary sewers
Examples of open channel flows

Since open channel flow is characterized by a constant pressure on the free surface, then the
pressure distribution with depth follows the hydrostatic law.
Depending upon the shape, open channels can be either prismatic or non-prismatic.
A prismatic channel is one in which the cross-sectional area along the direction of flow remains
constant. This implies that along the direction of flow, the area is only dependent on the depth of
flow. In non-prismatic channel, the cross-sectional area changes with distance along the direction
of flow. In this case the area is a function of depth and distance along the flow.
Depending upon the form, an open channel is either exponential or non-exponential. A channel is
exponential when its area of cross-section may be expressed in the form:

A = khm --------------------(OC-1)

Where A = area of cross-section, k = constant; d = depth of flow and m = exponent.


Rectangular, parabolic and triangular channels are exponential, whereas trapezoidal and circular
channels are non-exponential.
Classification of flow in open channel

72
Below is a classification attributed to Ven Te Chow taken from (Fluid Mechanics – A. K. Jain,
2003).

Open channel flow

Steady flow Unsteady flow

Uniform flow Non uniform or Uniform flow Non-uniform flow


varied flow or varied flow

Gradually Rapidly
Gradually varied flow Rapidly varied varied flow varied flow
flow

HYDRAULIC (GEOMETRIC) ELEMENTS OF CROSS-SECTION

Hydraulic elements of some common open channels

73
B B
B

h 1 h m m h

m
b b

B = b + 2mh B=b B = 2mh

A = (b+mh)h A = bxh A = mh2

P=b+2h(1+m2)1/2 P = b +2h P = 2h(1 +m2)1/2

UNIFORM FLOW IN OPEN CHANNELS

Consider a uniform flow in a prismatic channel shown below. The channel has a bed slope of So
Since the flow is uniform, it implies that the depth in section (1) and section (2) are the same and
therefore the slope of the water surface S must be equal to the slope of the bed. Also since flow
is uniform, then the velocity v1 must be equal to the velocity v2 and hence the velocity heads at
section (1) and section (2) must also be the same. This means the slope of the water surface is
equal to the slope of the energy line. Therefore by the condition of uniformity of flow in the
channel, we have:

So = S= SE ----------------------------(OC-2)

Now let us consider the flow of the mass of fluid within the sections 1 and 2 with cross –
sectional area of A. Since the flow is uniform, then according to Newton‘s second law the sum of
forces on the fluid mass in the direction of flow must be equal to zero.
Let us find all the forces acting on the fluid mass which include i) the weight of fluid, ii) the
frictional force acting on the wetted perimeter and iii) the pressure forces
However since the fluid is subjected to a constant atmospheric pressure and h1 =h2 then the
pressure forces resolve in the direction of flow to be zero.
Now, resolving all forces in the direction of flow, we shall obtain:
γALsin θ - τoPL = 0 ------------------------(OC-3)

where A – cross-sectional area; τo = average shear stress on the wetted perimeter;


P = wetted perimeter and L = length of section.
The average shear stress is then given by:

74
(2)
(1)

Slope SE Energy line

V12/2g
Slope S Liquid surface
V22/2g

h1
Slope So = sin θ h2

Z1 θ L Z2
Channel bottom
O O

Arbitrary datum O –O

τo = γ (A/P) L sin θ = γRS ----------------(OC -4)

At this stage, we make the assumption in respect to the flow resistance, according to which the
force resisting the flow (τoPL) per unit wetted area is proportional to the square of the velocity
head. i.e τo = K (v2/2g) where K = const.
The above assumption can be put in this form:

τo/γ = (f/4)x(v2/2g) ------------------------(OC-5)

where K = γ.f/4; f is the frictional resistance factor.


Substituting eq. (OC-5) into eq. (OC-4) and making velocity the subject, we shall obtain:

8g
v . RS  C RS -----------------(OC-6)
f
where C = (8g/f)1/2 referred to as the Chezy‘s coefficient

Q  Av  AC RS

If we represent K =AC(R)1/2 -------conveyance of channel

QK S
Then, Q2
S
K2

75
Equation (OC-6) represents the Chezy‘s formula; a formula developed by Antoine Chezy, a
French engineer in 1770 for uniform flow.
The Chezy‘s formula is simple and easy to use; however the main difficulty with it is the correct
determination of the constant C in the formula, which is generally obtained by empirical means.
Fortunately, a lot of work has been done on the constant C and there are recommendations for
determining the value of C. Below are a few of the available recommended formulae.

1) Ganguillet and Kutter’s formula


In 1819, two \Swiss engineers, based on extensive experimental studies on European rivers
proposed the following empirical formula to determine the Chezy‘s constant.

1 0.00155
23  
C n S -------------------(OC-7)
 0.00155  n
1   23  
 S  R

2. Manning’s Formula
Robert Manning, an Irish engineer in 1889 presented a formula for velocity of flow in open
channel from which the Chezy‘s constant could be defined as:

C = 1/nR1/6 ---------------------------(OC -8)

3. Bazin’s formula

The French Hydraulic engineer proposed the following formula for the Chezy‘s constant in 1897.

87
C -----------------------(OC -9)
k
1
R

where k = m/1.81. The values of m is given in a table below


n = Manning‘s roughness coefficient given for various surfaces in a table below.

Table X. Manning’s Roughness Coefficient

No. Description of channel surface Manning’s Roughness


coefficient, n
1 Well planed timber, glass or brass 0.009
2 Wood-shaved flumes, finished concrete 0.010- 0.013
3 Glazed tiles, vitrified sewers, concrete pipes 0.013-0.016
4 Bricks in cement mortar, ashlars 0.012- 0.20
5 Rubble masonry in cement 0.020 – 0.030
6 Straight unlined canals in earth 0.020
7 Unlined canals in gravel, unlined canals in earth 0.225
with some curves

76
8 Corrugated metal flumes, unlined canal winding 0.025
9 Canals with rough stony bed or weeds 0.030
10 Winding natural streams in good conditions 0.035
11 Rivers of irregular cross –section winding 0.04 to 0.10
Cast iron 0.013 – 0.017
Riverted steel 0.017 – 0.020

Table XX. Bazin’s Coefficient

No. Description of Channel Surface M


1 Very smooth cement or planned wood 0.11
2 Unplanned wood, concrete or bricks 0.21
3 Planks, ashlars 0.29
4 Rubble masonry 0.83
5 Earth channels in perfect condition 1.54
6 Earth channel in average condition 2.36
7 Earth channel in rough condition 3.17

HYDRAULIC MOST EFFICIENT CROSS-SECTION

Let us assume that we are to design a channel (i.e. to determine the dimensions) with the
following initial parameters given:
i) form of the cross-section or channel= trapezoidal
ii) the longitudinal slope S = So
iii) Side slope of channel m = mo
iv) The roughnest coefficient n = no and
v) Flow rate Q = Qo
The above problem has so many solutions; in fact an infinite number of solutions

77
h1 m1
β1=b1/h1
b1

m1=m2=m3= --=mo
h2 m2 β2=b2/h2
i1=i2=i3----=io
b2
n1=n2=n3=---no

h3 β3=b3/h3
Q1=Q2=Q3=-----Qo m3

b3

In the diagram above, only three of such solutions are given and it is clear that the first solution
is characterized by wide channel and shallow depth while the last is characterized by small base
width and deep depth.
We see that the first and third solutions are characterized by relatively high friction, defined by
the magnitude of the wetted perimeter and hence their velocities will consequently be small.
β1 ≠ β2 ≠ β3 and
P1 ≠ P2 ≠ P3
Where β = relative width
From the above discussion, it can be seen that, among the infinite number of solutions available,
there must be one which has the least wetted perimeter and cross-sectional area and hence the
maximum velocity of flow and least resistance to flow.

Hydraulic most efficient cross-section: is that cross-section which for the given parameters Qo,,
mo, no, io will characterize the maximum possible velocity of flow and hence the minimum wetted
perimeter and minimum cross-sectional area or a channel that gives the maximum discharge for
a given shape, slope area and roughness. Such a channel offers the least resistance to flow.

78
P, Q, V
P, A, V A = f(β)
Q=f(β)

P=f(β)

V=f(β) V=f(β)

P= f(β)

βb.s βb.s

Best Section: Q, m, i, n, = const Best Section: A, m, i, n, = const

To obtain the best section, the mathematical concept of minimum and maximum is used.
For a given discharge, the best section is obtained when the area and wetted perimeter are
minimum or the velocity is maximum.

For a trapezoidal channel, the area


A = (b + mh)h = h2β + h2m and
P = b +2h(1 + m2)1/2= hβ + 2h(1+m2)1/2
Therefore δA = (2hβ δh +2hm δ h) +h2 δβ and
δP = β δh + 2(1+m2)1/2 δh +h δβ

A h
 2h  2hm   h2  0
 
P

 
   2 1 m2 
h

h 0

By solving the two systems of equations, we shall obtain the best section for trapezoidal channel
as
b
 
 b.s     2 1  m 2  m --------------(OC-10)
 h  b. s

Uniform flow in close conduits

Examples of conduits designed on uniform flow concepts could be sewage pipes, drainage pipes,
hydrotechnical conduits etc. They come in different forms with relatively complex cross-
sectional areas, and wetted perimeter.

79
Closed conduits working under uniform flow conditions have the following specific
characteristics:
1) The maximum velocity of flow in the conduit occurs at a depth:

h = (0.80 – 0.85)D ----------------------(OC-11)

2) The maximum discharge Qmax occurs at a depth

h = (0.93 – 0.95)D ---------------------(OC-12)

It is quite difficult to understand why maximum velocity and discharge do not occur at the top
when conduit is full but one could explain the observation as follows:
At higher depths, (h.>0.8D), any small increase in depth brings about increases in cross-sectional
area and wetted perimeter of flow. However, the changes in the wetted perimeter far exceed the
increases in cross-sectional area. And since the wetted perimeter defines the frictional losses of
flow, the velocity consequently reduces and hence the flow rate.
Also, between the free surface and the soffit of the pipe is always a layer of a gaseous mixture of
gas and liquid, which is in motion, the motion getting faster as the depth in the conduit increases.
The presence of this gaseous mixture creates resistance at the free surface, the magnitude getting
bigger as depth of flow increases.
Equation (OC-11 & OC -12) are quite complex to prove for most sections because of the
complexity of the cross-sections, however, for circular pipes, they can be done analytically.

Conditions for maximum discharge and velocity in circular conduits under uniform flow
conditions
Consider a pipe of diameter D with changing depth of flow, h. In circular pipes, the cross-
sectional area and the wetted perimeter are a function of the angle β, where 2β is the angle
subtended at by the free surface at the centre
Now, using the Chezy‘s formula to define the discharge, we have

Q  AC RS  AC A PS  C A PS 3
---------------(1*)

For maximum discharge, δQ/δ β=0

1/ 2 3/ 2

Q   A 3 / 2 P 1 / 2    
3  A 1  A
A    P
2P 2P
-------------(2*)
Q A P
 3P A 0
  

From geometry,
P = D β and -----------------------------------(3*)

80
β β
D O
h

2 D 2
A
2
x
1 D D


 2 x sin x cos 180 o    
4 2 2 2 
D  D
2 2
A  sin  cos  -------------------(4*)
4 4
D2 D2
A  sin 2
4 8

Substituting the wetted perimeter and area from (3*) and (4*) into (2*) we shall ontain:

  D2 D2   D2   D 


3D   sin 2    sin 2  0
  4 8   4  
 D2 D2   D2 D2 
or.........3D   cos 2     sin 2  D  0
 4 8   4 8 
3 3 3 D 3 D 3
or........ D  D  cos 2 
3
 sin 2  0
4 4 4 8
or.........4  6 cos 2  sin 2  0

The solution of this equation is obtained by trial and error as:

β = 2.68 radians or β = 154o


Therefore the depth of flow for maximum discharge is

81
h = D/2 + (D/2)cos (180- β)
h = (D/2) (1-cosβ)

And substituting β = 154o,

h = (D/2) (1-cos 154o)


or h =(D/2) (1+0.90) = 0.95D β

Therefore the maximum discharge in a circular pipe occurs when the depth of flow is 0.95 D.
If the Manning‘s formula were used, the result would have been slightly different i.e
Hmax. Q = 0.94D
Generally the maximum flow occurs at h = (0.93 -0.95) D.

By similar analysis as above, the maximum velocity could be obtained as:

hmax. V = 0.81 D

Generally the maximum velocity occurs at h = (0.80 – 0.85)D


Although the above conclusion has been obtained for circular pipes, it hold true for all closed
conduits.

Problems encountered in the computation of channels in uniform flow conditions


As an example, we shall consider trapezoidal channel. Trapezoidal channel is characterized by
six parameters; b, h, m, (for cross-sectional parameters), n, S, Q or v. Some of these parameters
would be given and so the hydraulic computation of the channel involves finding the parameters
that are not known. Below are some examples of some of the problems that could be encountered
in practice.

a) Cross-sectional parameters to be found. Given: m, b, n, I, Q; and required to find h


This problem can be solved either analytically or graphically.
i) Analytical solution.
The discharge , Q = vA = AC(RS)1/2 = A(1/n)R2/3 S1/2

A = (b+mh)h; P = b +2h(1=m2)1/2
From these, the hydraulic radius is obtained as:

R = A/P = [(b+mh)h]/[b+2h(1+m2)1/2]
Therefore Q = (b+mh)h x (1/n) x {[(b+mh)h]/[b+2h(1+m2)1/2]}2/3 x S1/2

In the above equation of Q, the only unknown is h which can be solved by the method of
successive approximation.

ii) Graphical solution.


1. For the graphical we use the formulae Q = AC(RS)1/2 and K = ACR1/2
2. We determine the necessary conveyance to carry the flow Q

82
Knec = Q/(S1/2)
3. By giving different values of h we try to find the value of h that will give us the necessary
conveyance. This is better done in a tabular form

No. Quantitties Different values of given depths


h h1 h2 h3 h4 h5
2 A = (b +mh)h
3 P = b + 2h(1+m2)1/2
4 R= A/P
5 C = (1/n))xR1/6
6 K = AC(R)1/2

4. For each depth, h, we obtain the conveyance, K and therefore we can plot K =f(h)

h
K =f(h)

hnec

h
m

O
b Knec K
Fig. Graphical Solution

5. From the plotted graph and knowing the necessary conveyance, we determine the depth hnec

b) Given: b, h, m, S, n. Required to find Q


Solution: 1. Calculate A and P
2. Calculate R = A/P
3. Calculate C = (1/n)R1/6
4. Calculate v = C(RS)1/2
5. Calculate Q = vxA

c). Given: b, h, m, n, Q. Required to find S

From v = C(RS)1/2 then S= v2/(C2R.)

1. Calculate A and P
2. Calculate R
3. Calculate C

83
4. Calculate v = Q/A
5. Calculate S = v2/(C2R)

d) Given: m, n, Q, S, v. Required to find b, and h


From v = C(RS)1/2 Let CR1/2 = W then v = WS1/2
1. Calculate A = Q/v
2. Calculate W = v/S1/2
3. Write an expression for A= (b +mh)h = f(h, b)
4. Write an expression for W = CR1/2 = f(h,b)

Worked Examples
1. Water flows uniformly at a depth of 1.2 m in a rectangular canal 3m wide laid on a slope of 1
in 1000. What is the mean shear stress on the sides and bottom of the canal?.
Based on the Reynolds and Froude numbers, determine the type of flow in the canal. Take
Chezy’s C = 70 and ν = 10-6m2/s. (Taken from A. K. Jain)

Solution
The mean shear stress on sides and bottom of canal is given by

τ0= γRS

The hydraulic radius R = A/P = (b x h)/(b + 2h) = (2 x 1.2)/(3 + 2 x1.2) = 0.667m

The mean shear stress τ0 = 9810 x 0.667 x(1/1000) = 6.54 N/m2

Average velocity, v = C(RS)1/2 = 70x(0.667 x/1000) = 1.806 m/s

Reynolds number of flow Re = vR/ν = (1.806 x 0.667)/10-6 = 1.203 x104


Since Re is greater than 500, then flow is not laminar but turbulent since it is far greater than
2000.
The Froude Number is given by Fr = v2/gD; D = A/B = b xh /b = h where D = hydraulic mean
depth..
Fr = 1.806)2/(9.81 x1.2) = 0.527
Since Fr, 1, then flow in channel is sub-critical
Therefore flow is turbulent regime and sub-critical state.

2. A wide channel carrying silt free water has a depth of 2.0 m. The maximum tractive stress
permissible on the bed to prevent scour is 1.96 N/m2. What is the maximum slope that can be
given to the channel?

Solution
For a wide channel the hydraulic radius R ≈ the depth of flow.
The critical tractive stress τc = γRcS
from which S = τc/γRc = 1.96/(9810 x 2 = 1/104
Therefore the maximum slope that can be given to the channel is 1 in 10,000.

84
3. Show that the friction factor of the Darcy-Weisbach formula, and the Manning’s roughness
factor are inter-related.

Solution
We showed that the Chezy‘s coefficient C = (8g/f)1/2
Also the Manning‘s formula is given by C = (1/n) xR1/6
Therefore equating the two, we have
(8g/f)1/2 = (1/n) R1/6
Squaring both sides and making f the subject, we shall obtain
f = (8g/R1/3) x n2+
f = 8 x 9.81/R1/3) xn2 = 78.5n2/R1/3

4. The cross-section of an open channel is a trapezium with a bottom width of 4 m and side slope
of 1V to 2H. Calculate the discharge if the depth of water is 1.5 m and S = 1/1000 Use
a)Chezy’s formula, C.= 50, b) Bazin’s formula, m =2.30

Solution:
a) i) Use the continuity equation Q = vA
ii) Express the velocity by Chezy‘ formula v = C√RS; Q = AC√RS
iii) Determine the area of flow A = (b+mh)h = (4+2x1.5)1.5 = 10.5m2
iv) Determine the wetted perimeter P = (b+2h √(1+ m2) = (4+2x1.5√5)=10.71 m
v) Determine the hydraulic radius R = A/P = 10.5/10.71 = 0.98 m
vi) Determine Q = AC√ RS = 10.5 x 50√0.98 x (1/1000) = 13.02 m3/s

b) Step i) to v) is the same as in a) above.


vi) Determine C using Bazin formula C = (87)/[1+(k/√R)]
but k = m/1.87; therefore C = (1.87 x87)/[1.87+2.30/√0.98 = 38.13
vii) Determine Q = AC√RS = 10.5 x 38.13√0.98x1/1000 = 9.90 m3/s

5. A rectangular channel has a width of 2.5 m and a slope of 1:400. Find the depth of flow if the
dischargeis 10 cumecs. Use Chezy‘s formula, C = 50

i) Q = vA = AC√RS
ii) Determine A = bxh =
iii) Determine P = b + 2h
iv) Determine R = A/P = (bxh)/(b+2h)
v) Determine Q = 10 = AC√RS = (2.5 x h) x 50 x √(2.5 x h)/2.5+2h)x1/400
vi) Solving for h, we shall obtain h = 1.85 m

Example 6.Ariver used for navigation has the cross-section shown in the figure below. If the
channel cross-section if of earth (Manning‘s n = 0.025) and has a slope of 1 in 1000, what rate of
flow should it carry at the given depth?

85
J I

A H
3

B G
C F
7

D E
6 6
30 15 125 15 30
6

Solution:
The channel is a complex one consisting of the main channel and the flood plains so we are
going to divide it into three parts: ABCJ, JCDEFI and FGHI and determine the flow passing
through each.

a) Portion JCDEFI
i) The flow is determined by Q = vA = AC√RS = (A/n)R22/3S1/2
ii) Determine A =area CDEF +area JCFI = (b+mh)h +(b+2mh)h*
A = [125+(15/7)x7] 7 + [125 + 2(15/7)x7] x 3= 980+465 = 1445m2
iii) Determine the wetted perimeter P = b +2h√(1+m2) = 125+2x7√(1+4.5)= 158.1m
iv) Determine the hydraulic radius, R = A/P = 1445/158.1 = 9.15m
v) Determine the flow Q = (A/n)R22/3S1/2 =(1445/0.025) x 9.152/3 x (1/1000)1/2=
Q2= 8320m3/s

b) Portion ABCJ
i) Determine the area A = (BC+AJ)/2 x CJ = [30+(30+6)]/2 x 3 = 99m2
ii) Determine the wetted perimeter, P = AB + BC = (h*√1+m*2)+30 = 3√5+30=36.71m
iii) Determine the hydraulic radius, R = A/P = 99/36.71 = 2.7m
v) Determine the flow rate Q1 = (A/n)R22/3S1/2 = (99/0.025) x 2.72/3(1/1000)1/2 = 243 m3/s

c) Portion FGHI; Q3 = 243 since it is similar to portion ABCJ

Therefore total flow rate Q = 8320 + 2 x 243 = 8806 m3/s

Example 7.
A trapezoidal channel with side slope 1 :1 has to be designed to convey 10 m3/s at a velocity of 2
m/s, so that the amount of concrete lining for the bed and sides is the minimum.
Calculate i) the area of lining required for one metre length of the canal
ii) If the rugosity coefficient n = 0.015, calculate the bed slope of the canal for

86
uniform flow.

Solution.
i) For minimum lining, means we must use the best section. From eq. (OC-10), best section of a
trapezoidal channel is given by :
βbs = b/h = 2(√1 + m2) –m = 2(√2)= 2x1.42 -1= 1.84
b = 1.84h
ii) Determine the cross-sectional area, A = Q/v = 10/2 = 5 m2 = (b + mh)h =
= (1.84h + 1 x h)h=2.84h2 therefore, h = √(5/2.84) = 1.32
b = 1.84 x1.32 = 2.43m
iii) Determine the wetted perimeter, P = (b + 2h(√1 + m2)= (b+2h√ 2) = =(1.84h+2.84h)=4.68h =
4.6 x 1.32 = 6.07
iv) Determine the hydraulic radius , R = A/P = 2.8h2/4.68h = 0.598h = 0.82
v) Determine the flow rate Q = (A/n)R2/3S1/2 = (5/0.015)x0.822/3S1/2 = 10
By making S the subject, we find S = [(10)/(333.3 x 0.879)]2=(0.03413)2 =0.001165
S = 1.165 x 10-3 = 1 in 858

Example 8.

A hydraulically efficient trapezoidal channel has side slopes of 1 : 1. It is required to discharge


14 m3/s with a gradient (channel slope)of 1 in 1000. If unlined, the value of Chezy‘s C is 45. If
lined with concrete, the value is 70. If the cost per m3 of excavation is three times the cost per
m2 of lining, will the lined or the unlined channel be cheaper?

Solution:
To understand and appreciate the problem, one must know that the lined channel will be able to
pass the required flow with much smaller cross-sectional area due to the larger velocities it can
pass.
Let us find the areas of both unlined and lined channels to pass the required flow
b
 
For a trapezoidal section, A hydraulic efficient section is given by:  b.s   2 1  m 2  m =
h
2(1.42)-1= 1.84 Therefore b = 1.84h.
a) for the unlined channel; m = 1, S = 1/1000, Q = 14; C = 45
i) Determine the area A= (b+mh)h= (1.84h+h)h= 2.84h2
ii) Determine the wetted perimeter P = b+2h(1+m2)1/2 = 1.84h+2h x 1.42= 4.68h
iii) Determine the hydraulic radius R = A/P = (2.84h2)/4.68h= 0.607h
iv) Determine the flow rate Q = AC√RS = 2.84h2 x 45 (0.607h x 1/100)1/2 =14
v) By method of successive approximation, we obtain h = 1.82m
Hence b = 1.84 x1.82 =3.35m
Area of section = 2.84 h2 = 9.407m2
If E is the cost per m2 of lining of the channel, then 3E will be the cost of excavation per m3 .
Considering a unit length of channel, cost of excavation of unlined channel will be:
Costunlined= Area x1 x 3E = 9.407 x1 x 3E = 28.22E

b) i) We do same for the lined and since slope and flows are the same, everything will be the
same except the flow equation where instead of C = 45 we put in C = 70

87
ii) Q = AC√RS = 2.84h2 x 70 (0.607h x 1/100)1/2 =14
form whence h = 1.53; b= 1.84 x 1.53 = 2.815m
iii) Area of section A2 = 2.84 x 1.532 = 6.648m2
iv) Cost of excavation = Area x 1 x 3E = 6.648 x 1 x 3E = 19.944E
v) Cost of lining = Area x E = 6.648 x E = 6.648E
vi). Total cost of lined channel = cost of excavation + cost of lining =
= 19.944E + 6.648E= 26.692E

Therefore the lined channel is cheaper.

Example 9.
Water flows in a channel of the shape of an isosceles triangle of bed width a and side making an
angle of 45o with the bed. Determine the relation between depth of flow d and the bed width for
a maximum velocity and maximum discharge conditions. Use the Manning‘s formula and note
that d is less than 0.5a

Solution

D F

G H

a/2
d

45o 45o
B
E C
a

Manning‘s formula for velocity


2 1
1
v  R3S 2
n
i) Determine the area of flow A = ½(GH + BC) x d = ½{(a-2d) +a}x d = (a - d)d
ii) Determine the wetted perimeter, P = a + 2BG = a +2√2 d
iii) Determine R = A/P = R  
A a  d d
P a  2 2d

1  a  d d 
2/3

v  S 1/ 2
n  a  2 2d 

88
For maximum velocity, dv/dd = 0

dv
 
 
S  a  2 2d a  2d   a  d dx 2 2  2  a  d d 
x  
1 / 3

dd n 
 
a  2 2d
2
  3  a  2 2d 

which upon simplification, results in for maximum velocity


a2 – 2ad - 2√2d2 =0 from which
d/a = 0.3382

For maximum discharge, dQ/dd = 0


1  a  d d 
2/3

Q  A.v  a  d dx   S 1/ 2
n  a  2 2d 


S 

a  2 2d  2/3
x
5

a  d d 2 / 3 xa  2d   a  d d 5 / 3  2 a  2 2d 
1 / 3 
x2 2  
dQ 3 3 

a  2 
x
dd n  2d
4/3

 
 

which for maximum discharge simplifies to


5a2 – 1.5147 ad – 22.6274 d2 =0

from which d/a = 0.4378

Velocity Distribution in Open Channel


The point velocity in an open channel keeps changing across the cross-section, just like in pipes
because of friction near the boundaries. In an open channel, an additional factor which influences
the velocity profile is the presence of a free surface. The surface tension of the liquid-air
interface produces a resistance to flow and causes the maximum velocity not to occur at the

89
surface, where the shear stress is negligible but to occur at some distance below the free surface.
Another factor that can affect and modify the velocity profile is the presence of wind. A strong
wind blowing upstream or downstream can modify the velocity distribution considerably. In a
vertical, the minimum velocity occurs at the bottom, whiles the maximum value occurs at a
distance from the free surface which varies from 0.05 to 0.25 times the depth of flow. The higher
values are applicable to sections closer to the boundary

y
Z Y X

at Z

at Y
y

at X

v
b
Position of maximum velocities
Channel cross-section with distances from boundary

The sketches above show the pattern of velocity distribution in the horizontal direction of a
rectangular channel
The velocity distribution in channels depends upon the shape of channel section, roughness and
its alignment. However, in wide channels lateral boundaries have negligible effects on the
velocity distribution in the central portions. A wide channel, according to Franzini and
Finnemore is one with the width of channel being more than 10 times the depth. The velocity
distribution in wide channels according to (Vanoni, 1941) can be approximated by;
1  y
v( y )  V  gdS o 1  2.3 log 
k  d
where V = depth average velocity, k = von Karman constant (≈ 0.4), d = depth of flow, So =
slope of channel and y = the distance from the bottom of the channel

The average velocity can be estimated by

V = v(0.4d) or

V = {v(0.2d) + v(0.8d)}/2

90
Vanoni in general showed that the Plandtl-Karman universal logarithmic velocity distribution
law for pipes is also applicable to two-dimensional open channel flow. Using Nikuradze‘s
constants for pipe he obtained the following velocity distribution in open channel flow as:
9 yu *
v  5.57u * log 10 .......................( for ..smooth..surfaces )
.
30 y
v  5.57u * log 10 .......................( for ..rough..surfaces )
k
u *  gRS ........hear ..or.. friction ..velocity

where v = velocity at a distance y from channel bed, k = height of surface roughness


The theoretically average velocity analytically obtained by Keulegan,using Plandtl-Karman
universal velocity distribution law and Nikuradze and Bazin‘s data is given below as:

V Ru *
3.25  557 log 10 ,............... for ..smooth..channels ...and
u* 
V 10
 6.25  5.75 log 10 ................ for ..rough...channels
u* k
V= mean velocity of flow.

From measurements of velocities in natural streams and velocity distribution curves plotted from
such observations, one can make the following inferences:

1) The mean velocity in a vertical is at 0.55 to 0.65 of the flow depth. The velocity at 0.6
depth is usually within 5%of mean velocity.
2) 2The mean velocities at 0.2 depth and 0.8 depth usually gives the meanb velocity in the
vertical within 2% The maximum velocity in the vertical is at 0.80 to 0.95.

91
GRADUALLY VARIED, NON-UNIFORM STEADY FLOW IN CHANNELS

We shall be considering only turbulent flow i.e. flow in the quadratic region of resistance. If a
solid body (object) begins to fall through the atmosphere, at the beginning the motion of body
will be accelerated (ie the body will have acceleration).
As the velocity of fall increases, the resistance to motion by the atmosphere also increases and
sothe accelerated motion of the object gradually changes into a uniform motion; ie a uniform
motion such that the force of gravity of the body is balanced by the resistance of the air.
Similarly, if we should discharge water into a prismatic canal, then at the beginning the water
entering the canal will have a non-uniform motion. However as velocities change, so also
changes the resistance (frictional forces), until at a very far distance from the beginning the
frictional forces balance the body (gravity) forces of the fluid, and we shall have a uniform flow.
Every stream seeks to have a uniform flow where the work done by the body (gravity) forces
will be equal to the work of the frictional forces.
Non-uniform flows in channels arise when we in one way or the other disturb (disrupt) the
regime of uniform flow.

Examples of non-uniform flow: which are characterised by these general conditions:


i) hconst in the direction of flow
ii) const in the direction of flow

1. PRISMATIC (CONSTANT GEOMETRY) CHANNELS WITH (LONGITUDINAL)


SLOPE OF BED i 0

In this case the uniform flow will be disrupted in the following cases.

a) When a dam is constructed across the channel.

The water begins to flow over the dam. As can be seen as a result of constructing the dam, we
fix (artificially) in the channel the point A on the free surface of the stream and also the depth hf
which differs from the normal depth which characterises a uniform flow NN.

In this case, we have hf  huniform

Therefore in a certain segment of channel AB we shall have non-uniform flow.

b) When a drop structure is constructed across the channel.

92
At the crest of the structure, we shall have a depth hf, which is less than the normal depth, which
characterises uniform flow. In this case also we fix (artificially) a depth in the channel, which
differs from the normal depth.
hf  huniform

1.1.1.1.1.6 N

1.1.1.1.1.7 B

Q huniform N
Ib>0 A
hf
c) When a sluice gate is constructed across the channel
Drop structure
1.1.1.1.1

Q huniform flow B N
hf

Here too we have


hf  huniform
Inference:
Non-uniform flow in prismatic channels with bed slope ib 0 arises only when we in some way
artificially fix the depth hf which differs from the normal depth which characterises uniform
flow.

2. PRISMATIC CHANNEL WITH BED SLOPE ib =0 or ib < 0

Q
i=0 Q
i<0

By analysing Chezy‘s formula


 = CRi

93
If i = 0 then  = 0
Therefore we may conclude that if i=0 and more especially if i < 0 then uniform flow cannot
generally exist. In this case there can only be non-uniform flow.

3. NON-PRISMATIC CHANNELS
(EITHER EXPANDING OR CONTRACTING CHANNELS)

Q Q

Expanding Channel Contracting Channel

In this case there can only be a non-uniform flow. In conclusion, there is the need to know and
remember the following. Non-uniform flow occurs:

I) In the case of prismatic channels


a) When ib<0 and when we artificially fix a depth in the channel which differs from the
normal depth of uniform flow.
b) When bed slope ib=0 or ib<0

II) In the case of a non-prismatic channel.

Therefore uniform flow can only exist in a prismatic channel with positive bed slope (ib=0) on
condition that the channel is long enough and that we do not in any way (ie by putting across the
channel structures.) disturb the uniform regime.
In the study of non-uniform flow in channels, the basic problem that usually arises is the
determination of the free surface line (i.e. the line of intersection of a vertical longitudinal plane
with the free surface line), which is very important in hydraulic engineering and has so many
practical uses.

For example:
1 2

h1 hconst h2 h1 h2
1.1.1.1.2
1 2

94
a) If we draw the free surface curve for a canal, we can find the depth h at any cross section.
Knowing the depth of water, we can then solve the problem of how deep the cut or fill
should be.

b) By drawing the free surface curve for a river on which a dam is to be built, we can know the
area to be unadated (flooded) and take the necessary precautions to reduce the damage of
flooding.

BASIC DIFFERENTIAL EQUATIONS OF GRADUALLY VARIED NON-UNIFORM


FLOW
E Energy Line
1.1.1.1.3

dhw
P Piezometric line E

h dH
O H h1 P
h2
L
z 
z1 dL z2 S

O O

We shall at first consider the general case of a non-prismatic channel. To remind you once again

1. The shape (form) and width of cross section of a prismatic channel is the same all along the
direction of flow. Therefore for such channels the cross sectional area A is a function of only
the depth ie A = f(h)…………………………(1)
2. The shape (form) of cross section of a non prismatic channel changes in the direction of flow
and so the cross sectional area A is a function of both depth and distance L from the origin ie
A = f (h,l)……………………….(2)
Let us consider a longitudinal section of a stream h - depth of flow; z – height of bed above
an arbitrary chosen datum. H = h + z
Let us further consider cross section 1-1 at a distance S from the reference section W-W and
2-2 situated at an infinitesimal distance dS from 1-1.
Now let us write the Bernoulli's equation for the sections 1-1 and 2-2.
p v  2 2
2
p
3. z1  h1   1  1  z1  h2   2   hw  const ……………….(3)
 2g  2g
P1 = P2 = Pa. Lets represent z1 +h = H
After differentiation we shall obtain

95
  2 
4. dH  d    dhw  0 …………………………………………………...(4)
 2g 

  2
dhw  dH  d  
 2g 
 2
Now let us represent  hv
2g
dhw = -(dh + dhv)
or

5. dH = dhw + dhv ……………………………………………..(5)

Eqn 5 is the basic differential equation of non-uniform flow.


Here dH is the change in free surface level within the distance dL (dH) may be positive or
negative. From equation 5 we can see that the fall (or increase) in the specific potential energy of
stream equals the increase (or decrease) in specific kinetic energy plus the losses due to friction.
Let us divide equation 5 by dL.
dh dhw dhv
   …………………………………………………..…(6)
dL dL dL
dh
  J - free surface slope
dL

Since in free surface flow the piezometric line coincides with the free surface of the fluid dh/dS
is also the piezometric slope.
dhw
 i f - Hydraulic gradient (slope of the energy line)
dL
Therefore equation (6) becomes
d   2 
J  if    ……………………………………………………(7)
dL  2 g 

Further on we make the following assumptions:

The head loss due to friction hw in a non uniform flow may be expressed by the same
relationship as in a uniform flow.
With this assumption we can now use Chezy's formula to find the value of if.
 = CRif
2 Q2 Q2
if  2  2 2  2
C R A C R K
Where , C, R, K are the flow and channel characteristics at cross section 1-1.

96
2 d  2 
J    ………………………………………..…A
C2R dS  2 g 

Equation A is the first form of the differential equation of a non-uniform flow.

SECOND FORM OF THE DIFFERENTIAL EQUATION OF A NON UNIFORM


GRADUALLY VARIED FLOW
Let us transform equation (6). Let us begin with the left hand side.
Since H = z + h, then dH = dz + dh
dH dz dh
  
dL dL dL
dH dz
But J ;   ib - bed slope
dL dL
dH
 J  ib  ……………………………………………………………….(8)
dL
Let us transform the right hand side of equation (6)
dhv d   2  d  Q 2  Q 2 d  1 
        …………………………(9)
dS dS  2 g  dS  2 gA 2  2 g dS  A 2 
Since in the general case
A = f(h,s)

Then the total differential of the function A will be given by


A A
dA  dS  dh
S h
 1  2 2  A A 
d  2    3 dA   3  dS  dh 
A  A A  S h 

Therefore
d  1  2  A A dh 
  3   
dS  A 2  A  S h dS 

A
But equals the width of the channel at the surface of water.
h

Equation 9 can therefore be written as


d   2  Q 2  A dh 
    3   B 
dS  2 g  gA  S dS 
hw 2 Q2
The value  if  2  2 2
dS C R A C R

97
Therefore equation 6 becomes after substituting all values
dh dQ 2  A dh  Q2
ib     B   ………………………………………….(10)
dS gA 3  S dS  A 2 C 2 R
dH
Solving equation 10 relative to we shall obtain
dS
Q 2 A Q2
ib  
dh gA 3 S A 2 C 2 R
 ………………………………………………….B
dS Q 2 B
1
gA 3

Which is the second form of the differential equation for a non-uniform flow.

I. For a prismatic channel,


A
A = f(h) and so the partial differential  0 and the basic equation B takes the form
S

Q2
i 
dh b A 2 C 2 R
 ……………………………………………B'
dS Q 2 B
1
gA 3

Which is the 2nd form of differential equation for a non-uniform, gradually varied flow for
prismatic channels.

II. If channel has a zero slope. (i.e. ib =0)

Q2
dh 2 2
 A 2C R ……………………………………………….B''
dS Q B
1
gA 3

III. For a channel with reverse slope (ib 0)


Let's adopt the symbol ib = -ib' (where i' 0)

Q2
i 
dh A2C 2 R
 ………………………………………………B'''
dS Q 2 B
1
gA 3

98
IV. For a channel with reverse slope (ib 0)
Let's adopt the symbol ib = -ib' (where i' 0)


l 
ib  i f
1   2
L 
ib  i f

FREE SURFACE NON-UNIFORM FLOW

4 AUXILIARY CONCEPTS

1) SPECIFIC ENERGY OF A CROSS SECTION

Energy line
P Piezometric line
 M free surface
h M
z
Obed
Obed z' Obed
Obed
O O O O
Datum

The total energy (Ht) per unit weight of a flowing fluid is given by the sum of the potential head
above an arbitrary datum, the pressure head and the kinetic head. For a particle at the point M,
the total energy Ht is given by
P  2
HT  z  
 2g
By definition, specific energy  of a cross section is the total energy when the bed level is chosen
as the datum i.e. when the datum is chosen to coincide with the bed of the channel.
The specific energy for the particle at the point M is given by
P  2
z 
 2g
………………………………....(10)
P
But z h

99
 2
 h
2g
………………………………….….(11)
2
 h
2g
 Q2  1 q2
 h  
 h
 bh  2 g 2 gh 2
q2
h3   h 2  0
2g
…………………………………..(+)
Equation (+) has three roots of which two are positive and real h
and the other is negative and unreal. For a constant value of
specific energy  is constant, there are two and only two alternate
depths for a given discharge. The larger of the two values
corresponds to conditions of deep slow flow (tranquil, or
straining flow). The smaller value is that for shallow fast flow hcr
(shooting flow). As the discharge per unit width increases for
tranquil flow, the depth decreases while for shooting flow the
depth increases. qmax

There is a depth at which the two depths coincide, when the two discharges for a given specific
Q
energy is a maximum and the energy required for a given discharge is a minimum.  
A
Q 2
  h 
2gA 2
………………………………………….(11a)
Q q
For a rectangular channel A = bh  V  
bh h
Q
q – unit discharge – discharge per unit length of channel q 
b
q 2
  h 
2gh 2
………………………………………….(11b)
Let us analyse the equation 11 or 11a.

In analysing it we shall assume the flow is given and that the given flow through a given cross
section may pass at different depths depending on the bed slope, roughness etc. For different
depths we shall obtain different specific energies. We can therefore say that  = f(h)

From the equations 11, 11a, 11b we can easily see that:
a) As h 0;   better seen in equation 11b, since the second term on the right hand
side tends to infinity.

100
b) As h 0;  

If a continuous function tends to plus infinity at the extreme of the dependant variables then it
means in between extreme values there must be a minimum value of the function.

The graph shows the function .


The shaded portion gives the changes in  M
the velocity head v2
2g I
In the case of a uniform flow h
(where h=constant in the direction v2
of flow) the total energy Ht hcr 2g
decreases downstream (due to h O
frictional losses) 450 II
but the specific energy E remains N
constant since the datum is at the
min 
bed level.

A plot of specific energy of a cross-section

2) CRITICAL DEPTH hcr


By definition critical depth is the depth corresponding to the minimum value of the specific
energy of a section. The velocity corresponding to this depth is called the critical velocity. To
find the value hcr we find the differential of the function and equate it to zero.

 f h   0 ……………………………..(12)
dE
dh h A3 =f(h)
B
dE d  Q 2  Q 2 dA
Then   h  1  0 …....(12a)
dh dh  gA 2  gA 3 dh hcr
dA
Since  B we have
dh
Q 2
1 B0
gA 3
A 3 Q 2 A3
or  …………………………………(12b)
Q2 B
B gA 2
g
1.1.1.1.1.7.1 Find
ing the critical depth
The solution of the equation 12b gives you the value of the critical depth. Equation 12b can
easily be solved by plotting a graph of A3 = f(h)
B
From 12b we may write
A Q 2
hcr   ………………………………(12c)
B gA 2

101
12b is the general expression for the critical depth in any open channel. For a special case of a
rectangular channel (A=Bh)

Q 2
hcr  ………………………………….(12d)
gB 2 hcr2
Q 2
hcr3 
gB 2

Q 2 q 2
hcr  3  3
where q=Q/B
gB 2 g

Eqn. (12d) can be rewritten as

Q 2  cr2 v2=1hk
hcr   2g 2
g ( Bhcr ) 2 g E E
P P
 cr2 1
hcr
2g 2 h=hk
Therefore in rectangular channels, when h =hcr
the velocity head equals ½ the depth in the channel.

The specific energy for a rectangular section is

z 2 q 2
Eh h
2g 2 gh 2 h

At critical conditions
hkp
q 2
h 33
E  hcr  2
 hcr  h  hcr
cr
kp2
2 ghcr 2h 2 cr

Thus the critical depth is two-thirds of the specific energy. E


2
hcr  E
3
Note: Critical depth does not depend on the slope of the bed.
N
3) NORMAL DEPTH ho:
K
N
hcr ho
B K
A 102
Q
The diagram shows a non-uniform
flow with a free surface curve AB.
The flow is considered given Q.
For a given discharge, normal
depth is that depth which will be
established in the channel if the
flow were to be uniform.
Q   o co Ro i from here ho can be obtained where o,
co, Ro, are parameters of normal depth.

4) CRITICAL SLOPE
Let us assume that we are given the flow Q, the cross section of a channel and the roughness.
Then for a given bed slope, we shall have one normal depth ho which can be obtained from
Chezy's formula. Fig A below is a channel with adjustable bed slope. By giving the bed different
slopes we can find the corresponding normal depth ho against bed slope using Chezy's formula.
ho

ho = f(ib)

ho'' ho' hcp = f2(ib)


i' A
i'' hcr ho=hcr

icp ib
fig A fig B
The critical depth, unlike the normal depth which (for a given cross-sectional dimensions)
depends on the bed slope, depends only on the discharge. The critical depth does not depend on
the bed slope. As can be seen from the right diagram. There exist a slope of channel bed for
which the normal depth ho will be equal to the critical depth hcr ie ho=hcr.
That slope for which ho=hcr is called the CRITICAL SLOPE icr.

DEFINITION
CRITICAL SLOPE
It is the slope, which must be given a cylindrical prismatic channel so that the normal depth ho
equals the critical depth (ho=hcr)
It is evident that;
a. If icr  ib , then hcr  ho ib - is called mild slope
b. If icr  ib, then hcr  ho ib - is called steep slope
c. If icr = ib, then hcr = ho ib - is called critical slope
Let us find an expression for critical slope
Q = AC(Ri)
Qcr = AcrCcr(Rcricr)

103
Earlier on hcr was obtained
Q 2
3
Acr
 ………………………………………..(*)
Bcr g
By putting Ccr into (*) we shall obtain
Acr
3



 Acr C cr Rcr icr 
2

Bcr g
Acr C cr Rcr icr
2

Or 
Bcr g
gAcr
Hence icr 
C cr2 Bcr Rcr
g cr
icr 
C cr2 Bcr

CRITICAL, SUPERCRITICAL AND SUBCRITICAL FLOW


CRITICAL FLOW:
When channel depth equals the critical depth ie h = hcr, the flow is called critical and its
corresponding velocity called the critical velocity. Fr=1

SUPERCRITICAL (TORRENTIAL) FLOW


When h hcr, the velocity is greater than the critical velocity and the flow is said to be
supercritical.
Fr1

SUBCRITICAL FLOW (TRANQUIL) FLOW


When h hcr, the velocity is lower than the critical velocity and the flow is said to be subcritical.
Fr1
Subcritical flow lies in the region I of the specific energy curve i.e. E = f(h) where we can see
that subcritical flow is characterised by the condition. dE  0
dh
Supercritical flow lies in the region II of the specific energy curve i.e. E = f(h) and is
characterised by the condition.
dE  0
dh
Note:
If flow in a channel is subcritical, then changes in the flow pattern in the lower parts (reaches) of
the channel affect the flow pattern in the upper parts (reaches) of the channel; however changes
in the upper reaches do not affect the flow pattern in the lower reaches.
Therefore in subcritical flow, the depth in a given cross-section is determined by the conditions
of the lower sector of the channel.

104
If flow is supercritical we have an opposite situation. i.e. the depth in a given cross-section is
determined by the conditions in the upper reaches of the channel.
Let us look at a situation in which we have two flows; a subcritical and supercritical flow. Let us
examine how the transition is going to take place. The equation (2nd differential) of a non-
Q2
1 2
dh K
uniform flow is given by  …………………………..(#)
ds Q 2 B
1 
g 3
It was shown above that when h = hcr
 3 Q 2

B g
Q 2 B
  Fr  1.0
g 3
If this is the case then the denominator of the 2nd differential equation tends to zero and therefore
with the condition h hcr the value dh 
ds
This means that the curve h = f(s) when h = kcr have a vertical tangent at the point where h = hcr
i.e.
W - W. It must be noted that the above inference was derived on the basis of formal analysis of
the equation (2nd differential). In practice at the point h = hcr we obtain a very (sharp) abrupt in
the flow pattern.

Experiments and practice show the following.


1. Changes from supercritical flow to C
subcritical flow, as a rule, are
accomplished only through hydraulic Cr
jumps.
Hydraulic jump is the abrupt increase
in stream depth from a value h less
than hcr to a value h greater than hcr. hydraulic jump

2. Changes from subcritical flow


to supercritical flow is
accomplished through a drop
which must pass through the
critical depth.

INVESTIGATION OF FLOW PROFILE IN A STEADY GRADUALLY VARIED


(NON-UNIFORM) FLOW – CYLINDRICAL, PRISMATIC CHANNEL

105
FLOW PROFILE:
It is the outline of the water surface in a longitudinal section of a channel.
Let us write the 2nd form of differential equation for a non-uniform gradually varied flow.
Q2
i 
dh b A 2 C 2 R

ds Q 2
1 3 B
gA
Let us transform this equation so that it becomes convenient for analysis.
Let us first consider the numerator of the right-hand side.
It is known that Q=koi where ko= conveyance of channel (moduli of discharge) for the normal
depth ho.
Therefore Q2 = Ko2ib
K = ACR
K2 = A2C2R
2
K o ib  K o2 
Therefore the numerator becomes ib   i 
b 1  
K2  K 2 
Now let us consider the denominator
Q 2 Q 2 B V 2 B
1 B 1  1
gA 3 gA 2 A gA
For prismatic channel A=Bh
V 2 V 2
1 but  Fr - Froude Number
gh gh
The denominator becomes 1-Fr
 K 
2

i1  o2 
 K 
dh 
 
ds 1  Fr
Now let us consider a longitudinal section of a channel.

N Zone A
C Zone A
N
C ho Zone B
N Zone B C
hcr C
Zone C N
ho Zone C
i  icr
i > icr

106
Three possible zones can be defined
Zone A – the zone of space above the critical and normal depths.
Zone B - the zone of space between the normal depth and critical depth
Zone C – the lowest zone of space above the channel bed and below both the critical and normal
depth lines.
As three types of slopes exist (mild M, steep S and critical C) and there are three zones for C
slope, eight different curves can be drawn.
Now let us consider mild curves i.e. when ib icr
M1A Curve-backwater flow
N N

MB curve-drawdown curve

C M2
C
M3
MC curve
flow through sluice gate

i < icr

First let consider zone A. Here the depth is greater than the normal depth. Therefore the modulus
2
of discharge KKo. The ratio K o 2 1 . Therefore the numerator is positive. Since h  hcr then it
Ko
means Fr1 and therefore the denominator is also positive.

dh  
 i  0 This means the depth must increase in the downstream direction and the flow
ds  
profile is of the type MA or backwater curve.
Now let us examine the conditions at the limits. Downstream, the depth gets larger and in the
limit tends to infinity (i.e h ). In this case K  and therefore the numerator tends to 1.
As h 
The value of Froude Fr 0 and the denominator also tends to 1.
dh
Therefore  i and dh = ids …………………….(*)
ds
The equality (*) shows that the water surface tends to the horizontal, as an increase in depth dh
equals the drop in the channel bed ids. Therefore the curve Ma downstream tends to be
asymptotic to the horizontal line.
Upstream, the depth decreases and remains in zone A and tends to ho. But in this case K Ko
2
and the ratio K o 2  1 and therefore the numerator tends to zero. The denominator remains
Ko
dh
positive since h is still greater than hcr. Therefore i
0  0 . If dh decreases and tends to
ds   ds
zero in the limit, then it means that upstream the MA curve is asymptotic to the normal depth.

107
Zone B:
Here the depth h ho, but h hcr and therefore K  Ko and Fr 1.
2 dh  
Therefore K o 1 and the numerator becomes negative. The denominator is positive i 0
Ko
2
ds  
This implies that the depth decreases downstream and the profile thus obtained is the M2 type
(drawdown curve).

The Limit Conditions


In the zone B the depth h is greater than hcr but downstream tends to hcr. Then KKcrKo u
Ko/K1.
As the depth tends to hcr, Fr 1
dh  
 i  
ds 0
i.e. the tangent to the water surface tends to the vertical. Upstream the depth increases and h
ho, then K Ko and Fr becomes less than unity.

dh
i
0  0
ds  
Therefore the MB curve upstream is asymptotic to the normal.

Zone C:

Here h hcr, therefore K Kcr Ko and Fr 1


Therefore
dh 
 i 0
ds 
Therefore the depth increases downstream and the profile is the MC type
In the downstream limit h hcr and therefore
dh
i

 
ds  0
Therefore the water surface sharply rises and tends to be tangent to the vertical.
Upstream h can tend to zero, which does not have any physical sense (i.e. if h=0 Q=0)

STEEP CURVES (i  icr)


SA curve (ponding curve)

C
zone A

C
N zone B

N
zone C

108
Zone A
THE SA CURVE
dh  
 i 0
ds  
The depth increases downstream. The condition at the extremities.
Downstream h  K   Ko2 0
K2
Fr 0 dh i or dh = ids
ds
The water surface tends to the horizontal.
Upstream h decreases and tends to hcr  ho  Ko2  0
K2
As h hcr Fr 1
dh
 i  
 
ds 0
Therefore the water surface falls and tends to be tangent to the vertical.
1.1.1.1.3.1.1.1
1.1.1.1.3.1.1.2 Zone B
1.1.1.1.3.2 S2 THE SB CURVE
dh
i

0
ds 
The depth decreases downstream
The conditions at the limits.
2
Ko
Downstream h decreases and tends to ho. K Ko. At the extreme, 1 . Fr > 1
K2
dh 0
  0
ds  1
Downstream the curve is asymptotic to the normal depth. Upstream h increases and tends to h cr>
ho.
2
Ko
 2 1 . At the limit h = hcr  Fr 1
K
dh
i
   
ds  0
Upstream the water surface rises and tends to the vertical.

109
1.1.1.1.3.2.1.1 Zone C
S3 THE SC CURVE
dh 
 i 0
ds 
The depth increases downstream.

The conditions at the limits.


2
Ko
Downstream h decreases and tends to ho : K < Ko. K Ko. At the extreme,  1 ; Fr > 1
K2


dh

 0  0
ds 
Downstream, the curve is asymptotic to the normal depth. Upstream h decreases and tends to
zero which is meaningless since h=0 means Q=0.

1.1.1.1.3.3 S1 CRITICAL, C-CURVES


1.1.1.1.3.4 S3

C-N Zone A

i = icr N-C
Zone C

If i = icr, the lines N-N and C-C coincides, so zone B does not exist and we are left with zones A
and C.
1.1.1.1.3.4.1.1.1 Zone A
dh  
 i 0 The depth increases downstream
ds  
Extreme conditions
2
K dh
Downstream h increases and h  K   o2  0 Fr 0  i
K ds
Therefore curve tends to the horizontal.
Upstream h decreases and h hcr (ho)
K
2
dh  0
K Ko  o2  0 Fe (1)    not defined.
K ds  0

Zone C
dh 
 i 0
ds 
Downstream h increases.

Extreme conditions

110
Downstream h increases and h hcr (ho), K Ko from (-)
dh
i
 0  not defined
ds  0
If the channel is wide enough and the Chezy‘s coefficient do not change with depth (C=const)
then these curves are straight horizontal lines.

CHANNEL WITH HORIZONTAL BED (i=0)


If i=0 then ho = , therefore zone A does
not exist since line N-N is located at Zone B
infinity from the bed. Therefore we shall
be left with zones B and C. Zone C

CHANNEL WITH ADVERSE SLOPE (i0)


Just like the (i=0) no normal depth can
occur [see ho = f(I)] so the normal depth Zone B
disappears and we have zones B and C.
Zone C

INTEGRATING THE GRADUALLY VARIED FLOW DIFFERENTIAL EQUATION


(We shall consider i0 only). By B. A. Bachmetot’s method.

We must note that for non prismatic channels solution of the differential equation B does not
A
exist since in general the function A = f(L) is unknown and therefore the differential is also
L
not known. Therefore, solution of the differential equation is only applicable to prismatic
channel.
Q2 k o2
i 2 2 1 2
dh A C R i k
We have the equation  ………………..(*)
ds Q 2
Q 2 B
1 B 1
gA 3 gA 3
K = wCR C = 1/nRy
In the equation above the parameters K, B, A are all functions of depth h and the remaining
parameters are constants. Therefore the right side of the above equation can be expressed as a

111
certain function of depth ie F(L). This function is quite complex and more especially when we
are to look for the integral for the above equation. To make the solution of the problem easier B.
A. Bachmetot proposed (for the purpose of integrating (*) a more simpler relationship between K
 k    h  
2 X

and h. This relationship is of the form     


 k   h 
Where h‘‘ and h‘ are two arbitrary depths taken for a given cross section K‘‘ and K‘ modulus of
discharge corresponding to these depths.
X – HYDRAULIC INDEX OF A CHANNEL
X – is a constant for a given channel and does not depend on the value of the depths h‘ and h‘‘.
2 log K   2 log K 
X
log h   log
K o2
1
dh K2
i
ds Q 2 B
1
gA 3
The numerator may be written as
2 X
 Ko   ho 
1   1  
 K   h
Let us transform the denominator
Q 2 B iK o2 B iK 02 B C 2  C 2 i B K 02
1 1 1   1   2 2
gA 3 gA 2 A gA 2 A C 2  g  A C A /  
2 2 2
But A/X = R ; A C R = K
C 2 i B  K 0  C 2 i B  h0 
2 X

1    1   
g   K  g  h
Therefore
X
h 
1  0 
dh  h
i
ds C i B  h0  X
2
1   
g  h
C 2 i B
Let   j
g 
Dividing both numerator and denominator by (ho/h)x we shall obtain
X
 h 
i   1
dh  h0 
 X
…………………… ( )
dL  h 
   1
 h0
Let us now represent h/ho by a new variable 

112
h dh
  d   dh  h0 d
h0 h0
h0 d  X 1
i X
dL  j
idL  X  j  X 11 j 1 j
 X d  d  1d  X d
h0  1  1X
 1
Integrating
S  S1  2 d
i 2   2  1   1  j 
h0 1 1 X
d
Introducing javerage and representing       C ………we shall finally obtain;
1 X
S  S1 
i 2   2  1  1  j ave        ……………()
h0
In this equation,
S2-S1 = l - length of segment
2 and 1 corresponds to h2/h0 and h1/h0
C 2 i B
j av   for the given sector
g 
(2) and (1) corresponds to 2 = h2/h0 and 1 = h1/h0
dz
The value of the function  ( )    C is obtained from a table for different values of the
1 x
hydraulic index of channel, X.
N

C
h1
h0
S1 h2

L=S2 – S1
S2

Equation () makes it possible to solve 2 main problems.


Problem 1.
Finding the distance l=s2-s1 between two cross sectional areas of a given channel, when the
depths h1 and h2 are known. It is supposed that all parameters of channel Q, ho, i, n and channel
form are also known. In this case

113
j1  j 2
j ave 
2
2 log( K 2 / K 1 )
X
log h2 / h1
With X calculated we find the values of the function(2) and (1) from a table.

Problem 2.
Finding the depth h1 in a given cross section at a known distance from another cross section
whose depth h2 is known. All other parameters of the stream and channel form are known. The
solution of this problem is more complex since the depth h1 would be found from the condition
h1 = h01.
Therefore the value 1 from equation ( ) can be found by method of approximation (by giving
arbitrary values h1, h1‘, h1‘‘, …. then finding l1, l1',…. ) or by constructing a graph of l = f (h1).

Note :
Remark:
Since the first problem is less difficult solving, then in practical calculations (for example in the
construction of flow profiles) it is necessary to use the first.
For example if we are requested to draw the flow profile of a river, which has been dammed,
knowing h2, we give arbitrary values of h and find the lengths l1.

Before demonstrating the use of the different curves in predicting the channel profiles that occur
in various circumstances, it is necessary to examine a phenomenon that is of frequent occurrence
in open channel, namely the formation of hydraulic jumps.

114
THE HYDRAULIC JUMP
Hydraulic jump: is an abrupt (rapid) increase in depth of a stream from a value less than the
critical depth to a value greater than the critical depth.

GEOMETRIC ELEMENTS OF THE HYDRAULIC JUMP


Height of jump aj = h'' –h'
Length of jump lj = horizontal projection of the roller.
The depths h' and h'' measured just before and after the jump are called conjugate depths.
Note that whenever there is an increase in free surface depth to pass through the critical depth
hydraulic jump occurs.
Let us examine the flow in the region of the jump. In the stream between the sections 1-1 and 2-2
we observe a separation plane ABC; below this plane, we have a stream (transitional stream)
which rapidly expands from a depth h' to a depth h'' above ABC we have surface roller,
backward flow. In this zone, strong turbulence is developed and this results in large energy
losses.
Just after the jump we have the post-jump section. In section 2-2 just after the jump the velocity
profile (average) is shown, the velocity at C is zero and the velocities just near the bed are
relatively high. The jump facilitates the rapid increase in the fluctuation of the actual velocities
and pressure. As a result, the post-jump region is characterised by a very strong turbulence.
The presence of large velocities at the bottom and high pulsation of actual velocity give rise to a
high erosion ability of the stream. Between section 2-2 and 3-3 the velocity profile gradually
changes until it attains the normal profile corresponding to a uniform flow.
The length lpj is given by lp.j = (1030) h''
Let us examine hydraulic jump from the energy viewpoint.
Let us consider a hydraulic jump caused as a liquid flows through a sluice gate as shown in the
diagram. The flow profile is also shown in the diagram.

Sluice gate II
Fr <1

c c'
C C
e d d'
a b b' Fr >1
a a' 115
Let us consider also the curve of specific energy of this channel discharging the given flow. The
points a‘b‘c‘d‘ on the specific energy curve corressponds to points a,b,c,d on the flow profile. As
we move in the direction of flow from point a to point d on the flow profile, we move on specific
energy curve E=f (h) following the route a‘b‘c‘d‘. At the point b‘ (corresponding to point b)
there is a jump (transition) from the lower part I of the specific energy curve E=f (h) to part II
and the point falls to point C‘. Lastly the stream obtains its minimum energy E at the point d‘,
and therefore here critical depth is established.
If however we were to suppose that in nature jumps do not exist, then the curve ab at a certain
point e would have come to the critical depth C-C and the stream having minimum energy could
not have moved on further since there would not be energy to overcome frictional losses.

BASIC EQUATION OF HYDRAULIC JUMP


Let us consider a longitudinal cross section of a jump as shown on the diagram. Our task is
to find the analytical relationship between the conjugate depths h’ and h’’.
1 2

C y2 C
C.G
B h''
y1 C.G G P2
h' C.P(centre of pressure)
Q P1 C.P
A T0 D S

1 V 2

We are going to use the momentum equation and we are going to apply it to the section of fluid
ABCD.
Fdt = MV  MV/dt = F
0Q(v2 – v1) = T0 + Gs + Rs + Ps ……………………..(1)
Where
v1, v2 – average velocities in sections AB & CD
Tos – projection of frictional force on section ABCD in the s-dxn.
Tos – will be neglected since it is small compared to other forces
(1-2 is considered very small); Tos = 0
Gs – projection of weight of liquid on the s-axis; Gs – 0
Rs – projection of reaction from the bed on the liquid ABCD; Rs = 0
Ps – P1 – P2 = A1(y11) – A2(y22)
 Q Q
 0 Q    A1 y1  A2 y 2 ..........................(2)
  A2 A1 
 0Q 2  0Q 2
 A2 y 2   A1 y1 ................................(3)
gA2 gA1
Eqn. (3) is the general equation of hydraulic jum

116
FUNCTION OF THE HYDRAULIC JUMP (HYDRAULIC JUMP FUNCTION)

Q = const h
E (h)
E2

y (h) E
h

E1 y(h') = y(h'')

For a given cross section and discharge, and since A=f (h) and y = f (h) and the rest of the
parameters being constant we can write the following.
 oQ 2
 yA   h  ……………………..(4)
gA
where h – depth in the section
A- Area of section
y – position of centre of gravity of section
The function (h) is called hydraulic jump function.
We may therefore write the general equation of hydraulic jump as
 (h‘) =  (h‘‘) ………………………..(5)
where
 (h‘) – hydraulic jump function when h = h‘
 (h‘‘) – hydraulic jump function when h = h‘‘

From equation (5) it is evident that for conjugate depths the hydraulic jump function has the
same value or hydraulic jump functions for conjugate depths are equal to each other.

The hydraulic jump function has the following properties


i) The hydraulic jump function has a minimum and the minimum coincides with the
minimum of the specific energy curve E(h).
ii) When h 0;  (h) 
iii) When h  ;  (h) 

Using the hydraulic jump function, we can find the conjugate depth h' when given h'' and vice
versa. Approximate empirical formula for conjugate depth.

Rachmanov's formula

117
1.2 h
  0.2 
  hcr
1 h 
     
0.167  0.834  hcr

HYDRAULIC JUMP EQUATION FOR A RECTANGULAR CHANNEL


For a rectangular channel
A = bh ; y = h/2 ; Q/b = q;  Q = qb
Where q – discharge per unit length of channel.
The hydraulic jump function becomes
 oQ 2  o q 2b 2 h
 h    yA   bh
gA gbh 2
  oq 2 h2 
 h   b   ……………………………………….….(6)
 gh 2 
 h 
 h   - Unit function of hydraulic jump
b
 q2 h2 
 h   b o   ………………………………..…….….(7)
 gh 2 
Instead of equation (5) we may write y (h') = y (h'') ……………(8)
q 2
But h 
3
cr
g
3
hcr h 2
 h  
h 2
Substituting into eqn (8)
hcr3 h  2 hcr3 h  2
  
h 2 h  2
3  1 1  h   h  2
2
hcr    
 h  h   2
h  2  h  2
2hcr3 
1 1

h  h 
Or hhh  h  2hcr3
Solving it relative to h' or h'' we shall obtain

118
h   
3
 hcr 
h  1  8   1
2 
  h   

h   
3
 hcr 
h   1  8   1 ………………………(9)
2
  h  

 hcr 
3
q 2

Let us consider 8  h 3
But cr 
 h   g

Therefore,

h 
2

h 

1  8Fr  1

h 
h
2
1  8Fr  1  …………………………..(10)

ENERGY LOSS IN A HYDRAULIC JUMP


Energy loss in a hydraulic jump for a rectangular channel can be obtained using the Bernoulli's
equation and hydraulic jump equation.
Thus applying Bernouilli's equation to sections 1-1 and 2-2 we shall obtain

v12 v2
E  E1  E 2  (h '  )  (h ''  2 ) --------------(11)
2g 2g

By finding hcr3 from equation (9) and putting into equation (11) we shall obtain

E 
h  h'
''

3

4h ' h ''

Using the hydraulic jump function, we can find the conjugate depth h' when given h'' and vice
versa.
Approximate empirical formula for conjugate depth by A .N Rachmanov:

1.2
'   0.2;
 ''
1
 '' 
0.167  0.834 '
where ξ‘, ξ‘‘ are the relative depths; ξ‘ = h‘/hcr, ξ‘‘=h‘‘/hcr

119
LENGTH OF JUMP
Length of jump are found by empirical formuli
1. Pavloski's formula (1937)
Lj = 2.5(1.9h''-h')
2. Safranecs' formula
Lj = 4.5h''
3. Baulmelov's formula
Lj = 5(h'' – h')
4. Aivazian's formula
L = k(h'' – h')3
4h'h''
Where
k = 8.10 + Fr
Fr

TYPES OF HYDRAULIC JUMPS


Hydraulic jumps appear in various shapes. The value of the Froude number of the approach flow
decides the nature of the jump. The U.S. Bureau of Reclamation has classified S patterns of the
hydraulic jump.

FLOW THROUGH ORIFICE AND NOZZLES

An orifice may be defined as an opening provided in the side or bottom of a tank for the
purposes of discharging the liquid contained in the tank. It should be noted that the opening will
be considered as an orifice only when the liquid surface in the tank is above the upper edge of
the opening. The characteristic feature of an orifice is the relative smallness of the thickness of
wall to the diameter of the opening. Orifice may be classified in three ways based o the size,
shape and sharpness.
Based on size, orifice may be classified as small and large orifices. I a small orifice, the size of
the opening is so small compared with the head of liquid over it; the velocity at the level of the
centre of the orifice may be taken as the mean velocity through the orifice. In large orifice, this
assumption is not correct.
Based on their shape, orifice may be classified into sharp-edged and bell-mouth.

120
FLOW FROM SMALL ORIFICE INTO THE ATMOSPHERE

Po
1 1

v≈0
2
H
C A0

O vc O
Ao

let us consider a flow fro a small orifice made in a vertical thin wall. of a vessel. Let p0 be the
pressure on the surface of fluid; Ao – area of orifice; Ac – area of contracted section of jet; H –
depth of centre of gravity of orifice from the surface of fluid.
The fluid at the outlet of the orifice sharply contracts. This contraction is caused by the inertia of
the fluid particles moving to the outlet in a curved path.

Let us denote ε = Ac/Ao - coefficient of contraction

Now, let us try to find the mean velocity vc in the contracted section and the flow rate Q from the
vessel. To do this, we write the Bernoulli‘s equation for the sections 1–1 and 2-2
The datum passes through the centre of gravity of the orifice.

p1 v1 P2 v2
z1    z2    h f        (1)
 2g  2g
p1 P0 v1 P2 pa v22 v22 v2
z1  H ,...,  .......  0.....z2  0;...  ;...   c
  2g   2g 2g 2g

Now let us express the head loss hf in terms of mean velocity at the contracted section

vc2
hf   f
2g
where δf = coefficient of resistance accounting for head losses between sections 1-1 and 2-2.
Now, substituting all the values into equation (1) we shall obtain

121
Pa pa vc2 v2
H     f c ....            (2)
  2g 2g
P P 
H   a  a   1   f  c \
v2
   2g

H con  1   f  c ....                  (3)


2
v
2g
1
whence....vc  . 2 gH con ..  cv 2 gH con    (4)
1  f
where cv – velocity coefficient , and Hcon – converted head
Now let us consider partial cases of formula (4)
1) when Po = Pa ie when the vessel is opened

VC = CV (2gH)1/2 ---------------------------------------------(5)

2) For an ideal fluid, hf = δfvc2/2g = 0 hence

vc = (2gH)1/2 ------------------------------------(6)

Formula (6) is the Torricelli theorem defining the velocity of efflux of a jet. The velocity of
the issuing jet is proportional to the square root of the head producing the flow.

Knowing the velocity at the contracted section, we can now find the flow rate Q when po =
pa.

Q = Acvc = Ac.cv.(2gH)1/2 = (Ao./Ao)Ac.cv(2gH)1/2 = Aoεcv(2gH)1/2

Q = cdAo(2gH)1/2

Where cd = εcv – coefficient of discharge; Ac/Ao = ε = contraction coefficient

Cv = 0.96 - 0.98 (0.97) ε = 0.67 - 1.0 , cd = 0.60 - 1.0, (0.62)

II) FLOW TRAJECTORY

Flow trajectory: - is the path described by the axis of a free falling fluid when issued from an
orifice.

122
H
vc
O x

To find the trajectory, we put an imaginary particle at the point O. with an initial velocity vc.
Now using the equation of motion of a particle , we can write:

X = vct; z = (gt2)/2 --------------(10)

Eliminating t from equation (10), we obtain the equation of trajectory

Z = (gx2)/(2vc2) where vc = cv (2gH)1/2 ------(11)

III) Flow from orifice into a reservoir

123
1 1

3 3
2

vc

Writing Bernoulli‘s equation for sections 1-1 and 3-3 and expressing the head loss in terms
of mean velocity at the contracted section

Hf = z = δ vc2/2g = (δ1-2 + δ2-3) vc2 = (δ1-2 +1)vc2/2g

We get the final equation of flow from orifice into a reservoir.

Q cd Ao(2gz)1/2 where z = head difference of the two reservoirs

FLOW THROUGH ORIFICE WITH CHANGING HEAD

124
QI

1 1 A

dH

dV
2 2

Qo

Let us consider a container filled with fluid. The cross-sectional area A in general is a function of
H ie A = f(H)

Qo – is the outlet discharge Qo = cdAo(2gH)1/2

QI – is the inlet discharge

1. If QI > Qo , then the container will be filling up until such time that QI = Qo

2. If QI < Qo then the level of fluid in container will be falling until such time when QI =Qo

Let us look at a situation when QI < Qo and find the time for the level to fall from 1-1 to 2-2.

Within the short time interval dt, the volume of fluid flowing out of the container

Vo  Qo dt  cd Ao 2 gH dt

Within this time the volume of water flowing into container

VI = QIdt

Change in volume of fluid in container can be expressed in two ways;

i) dV = VI – Vo = QIdt – cdAo(2gH)1/2dt

ii) dV = AdH

125
Equating i) and ii)

QIdt - cdAo(2gH)1/2dt = AdH

A
dt  dH
QI  cd Ao 2 gH
H2
A
Integratin g....t   Q dH
H1  cd Ao 2 gH

I )Particular case when QI = const; A = const


H2 1 H
A A
t  dH   dH
H 1 QI  cd Ao 2 gH H 2 cd Ao 2 gH  QI

II) Time to empty a reservoir, QI = 0; and A = const

 
H1 H1
A A dH A 2A
t  
H1
dH   2H 1 / 2  H1  H 2
H2
H 2 cd Ao 2 gH cd Ao 2 g H2 H cd Ao 2 g cd Ao 2 g

Time for total discharge of fluid from container when H2 = 0

2A 2 AH1 2 AH1
T H1  
cd Ao 2 g cd Ao 2 gH1 Q1

Flow through Nozzles

In practice the need frequently arises for increasing the coefficient of discharge or maintaining
the shape of the jet (as in hydraulic giants, fire holes, lances etc). To achieve such purposes
different nozzles are used.

A nozzle: - a very short completely filled pipe, in which the head loss (linear) is negligible
during calculations. The main types of nozzles are represented below:

126
A B C D E

A – external cylindrical nozzle (Venturi nozzle); E – Bell-mouth nozzle

B- internal cylindrical nozzle ( Borda nozzle)


C – Conical converging nozzle; D - Conical Diverging nozzle

Equations for v and Q

a) Flow through nozzle into the atmosphere

v  cv 2 gH    12)
1
cv     coefficien t.of .velocity    13)
1   noz
Q  cd ( noz A 2 gH              14)
cd   .cv    disch arg e.  coefficien t
 noz  1.0,...cv.noz  0.82;..    cd  0.82
v and Q are the velocity and discharge at the end of the nozzle.

Flow through nozzle into a reservoir

v  cv 2 gH            
Q  cd A 2 gH ...      where..cd  0.82

127
Comparing flow from orifice and flow through nozzle (Venturi nozzle)

In the case of flow into the atmosphere through nozzle

Qnoz = 0.82A(2gH)1/2; vnoz = 0.82(2gH)1/2

In the case of flow into the atmosphere from orifice;

Qori = 0.62A(2gH)1/2 ; vorif = 0.97(2gH)1/2

If heads and cross-sectional area of nozzle and orifice are the same, then we have the following

Qnoz 0.82
  1.34
Qorif 0.62
vnoz 0.82
  0.85
vorif 0.97

As can be seen, if we connect a nozzle to an orifice, we achieve the following:

i) flow velocity into the atmosphere is reduced by about 15 %

ii) flow rate (discharge) is increased by 34%

128
FLOW OVER WEIRS AND NOTCHES

A weir in general is any regular obstruction over which flow occurs. Weirs, especially the sharp-
crested ones are commonly used for measuring large and small open flows in the field or
laboratory.

Head race

H Ho
vo
z
δ
P

Tail race

LB
B

The zone of fluid before the wall of the weir is called the head race while the zone of water after
the weir is called the tail race

The distance LB from the wall of the weir to the point where a fall in the level of stream is
noticeable, known as the drawdown section is given by:

LB = (3- 5)H

H - measured at the section B-B is called the geometric head which is the rise of fluid above the
crest of the weir

B – width of weir going into the paper

δ – thickness of the wall of the weir

z – geometric drop in the level of fluid at the weir. It is the difference between the head race and
the tail race.

vo – approach velocity

Ho – The total head which is given by the sum of the geometric head and the velocity head.

129
zo – total drop in head over the weir which is the sum of the geometric drop plus the velocity
head.

Classification of weirs

1. By form of the opening

a) Rectangular

b) Triangular

c) Trapezoidal

d) Circular

e) Parabolic

f) With inclined crest

2. By the thickness of the cross-section of weir wall

a) Sharp-crested weir δ ≤ (0.1 – 0.5)H: In sharp-crested weirs, the nappe springs free as
it leaves the upstream face

b) Broad – crested weir 2H ≤ δ < 8H: In broad-crested weir, the falling nappe is
supported on the crest and does not allow the nappe to fall free.

c) Weir with practical profile (Orgee weir)

3. By width of obstacle in relation to width of stream

a) Suppressed weir: - When the length of the weir is the same as the width of the channel
(stream) It is so called because in this case, the effect of sides or ends of the weir on contraction
of the nappe is eliminated or suppressed. The width of the nappe is the same as the width of the
stream.

b) Weirs with end contractions: - In this case, the length of the weir is less than the width
of the channel. Therefore there is a lateral contraction of the nappe so that the length of the nappe
is less than the length of the weir crest.

4. By the level of downstream water level relative to the crest of the weir.

a) Un-submerged weir: - When the water level on the downstream side of the weir is
below the crest of the weir, it is said to be un-submerged.

b) Submerged weir: When the water level on the downstream side of the weir rises above
the level of the crest, the weir is said to be submerged.

130
The flow of fluid over weirs is extremely complex phenomenon and virtually impossible of exact
analytical solution. As result, all weir formula are derived by making lots of simplification of the
problem which leads to an approximate solution which is somehow rectified by the use of
experimental coefficients derived from experimental set ups.

To derive a simplified relationship for a weir, the following assumptions are made:

i) velocity distribution upstream from the weir is uniform

ii) all fluid particles move horizontally as they travel pass the weir crest

iii) the pressure in the nappe is zero; ie atmospheric

iv) the influence of viscosity, surface tension, turbulence are all neglected

v12/2g 2
1 1 h

H H
dh

2
Weir crest
v1
P

B
Fig. Simplified flow over weir with the above assumptions

After simplification, the sharp-crested weir is considered as a large orifice of rectangular shape
placed in the channel such that the head on its upper edge is zero. As a result, the upper edge is
eliminated leaving only the lower edge or the crest. By considering the weir as a large orifice and
by considering an elementary strip of area Bdh below the free surface, the velocity of flow is
given by:

u  cv 2gh..

The flow rate through the strip

dQ  Bdhcv 2 gh

131
The theoretical flow passing over the weir is obtained by integrating the above equation over the
geometric head.
H
2
Q   Bcv 2 ghdh  Bcv 2 g H 3 / 2
0
3

To obtained the actual flow rate over the weir due to the above assumptions, we introduce an
experimental coefficient of discharge

2
Q CD B 2 g H 3 / 2
3

If the approach velocity vo is appreciable (>1.0), then instead of geometric head one has to use
the total head Ho= H + vo2/2g

As can be seen, the formula for flow over weir seems simple and easy to use however, the
difficulty is with the correct determination of the discharge coefficient. In the laboratory, this
could be determined quite easily and accurately, however, it becomes extremely difficult to
determine this accurately in the field. A few empirical formulae exist for the determination of the
discharge coefficient for a rectangular weir.

CD = 0.40 + 0.05 H/P ------------Chugaev R. R. (2/3 in front of flow rate absorbed)

CD = 0.405 + 0.003/H -----------Basen (2/3 in front of formula absorbed)

CD = 0.397 + 0.00015/H3/2 Fteley and Stearn (2/3 in front of formula absorbed)

CD = 605 + 0.08H/P + 1/1000H ------Rehbock

CD = 0.611 + 0.075H/P -----------taken from A. K. Jain.

Contracted Weirs

For weir with end contractions, Francis observed that the effect of the contraction is to reduce the
length of nappe by 0.1H. Thus if there are n end contractions, then the length of the nappe will
be reduced by 0.1nH and will be equal to (B – 0.1nH) and the discharge equation will then be
given as:

2
Q CD ( B  0.1nH ) 2 g H 3 / 2
3

Flow over a Triangular Weir (Notch)

132
A triangular weir (notch) is particularly useful where the discharge is to vary over a large range
and the same accuracy is desired for both small and large discharges because of the following
advantages:

i) The coefficient of discharge for a triangular weir is practically independent of the head. This is
because, for all heads the ratio of the head to the wetted length of crest is constant. But in a
rectangular weir, the ratio of head to the wetted length of crest is not constant.

ii) When the discharge is small, a triangular notch provides a greater head than the rectangular
notch. Hence head measurement can be done more accurately over the triangular notch than
over the rectangular notch.

iii) When the flow rate is small, there is the tendency of a clinging nappe in a rectangular notch.
However, for the same flow rate over a triangular notch, the head will be greater and clinging
nappe will be avoided.

b h

dh
H

The equation for the discharge over the vee-notch can be obtained just like we did in the
rectangular weir by considering an elementary strip of width b and height dh. Considering the
notch as a large orifice, the velocity of flow is given by:

u  cv 2gh..

The flow rate through the strip

dQ  bdhcv 2 gh

But b = 2(H – h)tan θ/2

133

Putting b into the discharge equation, dQ  2( H  h) tan cv 2 ghdh
2

 
 
H
8
Q   2 tan cv 2 g Hh1 / 2  h3 / 2 dh  cv 2 g tan H 5 / 2
0
2 15 2

The above equation represents the theoretical discharge without consideration of the contraction
of the flow area. The actual discharge is obtained by applying the coefficient of discharge CD. to
obtain:


5
8
Q  CD 2 g tan .H 2
15 2

By representing k= (8/15)CD(2g)1/2 and taking the average value of CD =0.60

k = 8/15 x 0.60 x 4.43 = 1.42 and hence

Q = 1.42 x tan(θ/2) H5/2

For a right angled Vee-notch, θ = 90o and hence

Q = 1.42 H5/2

Empirical formulae 1) By Thompson

Q = 1.4 H5/2

2. By King

Q = 1.343 H2.47

By Grave When 22o ≤ θ ≤ 118o Q = 1.331 (tan θ/2)0.996 H2.47

The coefficient of discharge is made up mainly of the contraction coefficient and the velocity
coefficient. The contraction coefficient of a notch depends on the wetted perimeter. For a
rectangular notch, the wetted perimeter does not vary with head because the base, which forms
part of the wetted perimeter, remains constant for all heads. Consequently, the coefficient of
contraction and hence the coefficient of discharge will not be constant for all heads. In a
triangular notch, there is no base and therefore the contraction is due to the sides only. The
wetted perimeter depends on the length of the sides which in turn depend on the head.
Consequently, the coefficient of discharge is fairly constant in a triangular notch.

Trapezoidal Weir

A trapezoidal weir has an opening of a trapezoidal shape and may be considered as a


combination of a rectangular and a triangular weir.

134
H
θ/2
θ/2

Trapezoidal Weir

The trapezoidal notch is equivalent to two notches; one rectangular with length b and the other a
triangular with an apex angle of θ. The total discharge over the trapezoidal notch would be equal
to the sum of discharges over the equivalent rectangular and triangular notches; thus
Qtrap = Q1 + Q2 =
 2 
5
2 8 2 8
CDb 2 g H  CD 2 g tan H  CD* 2 g H 3 / 2  b  H tan 
3/ 2

3 15 2 3 15 2

Cippoletti Weir

The principle behind the Cippoletti weir is based on the fact that in a rectangular weir, due to end
contractions, the effective length of the weir is reduced and consequently the discharge. On the
other hand, by the addition of side slopes to a rectangular weir, gives an increase in discharge.
Therefore if the side slopes of a trapezoidal weir are so adjusted such that the reduction in
discharge due to end contractions is just equal to the increase in discharge due to the addition of
side slopes, the net effect in discharge is zero and therefore the standard weir formula without
end contraction can be used.
The Cippoletti Weir is a trapezoidal weir with side slopes 1 horizontal to 4 vertical such that the
discharge in effect equal to that of a rectangular weir without end contraction.

3
2
For Cippoletti weir, Q  CDb 2 g H 2
3
Cipolletti gave the equation for the discharge through such a weir as:
Q = 1.86 b H3/2

Submerged Weir

135
A submerged weir (notch) is one in which the level of fluid in the downstream side of the weir is
above the crest of the weir.

1 H
A
H1 B
2 H2

Flow over submerged weir

The flow over such a weir can be obtained by dividing the flow into two portions:
i) Flow over the upper part of the section (above line AB) of depth (H1 – H2) may be
considered as a free from a weir into the atmosphere and
ii) Flow through the lower part of depth (below line AB) may be considered as a
discharge through a submerged orifice.

The discharge over the upper portion behaving as a free weir is given by

CDb 2 g H1  H 2   CDb 2 g H 3 / 2


2 2
Q1 
3/ 2

3 3

and the lower portion acting as an orifice is given by:

Q2  CDbH 2 2 g H1  H 2   CDbH 2 2 gH

Therefore the total flow is given by:

2
Q CDb 2 g H 3 / 2  CDbH 2 2 gH
3
Experimental works by Herschel, Villemonte and Mavis on the discharge characteristics of a
number of sharp-crested submerged weirs on rectangular, triangular. Parabolic and proportional
weirs showed the results for all types could be represented by the equation:

0.385
  H n 
Q  Q1 1   2  
  H1  

136
where Q = discharge for submerged condition
Q1 = free discharge given by Q = K H1n
n = exponent in the free discharge equation
n = 1.44 for contracted weir
n = 1.5 for suppressed weir
n = 2.50 for 90o notch.

Broad-Crested Weir
Broad-crested weirs (long-based weirs) are weirs that have crest lengths that are sufficiently
longer (δ > 0.67H) to prevent the nappe from springing free. These weirs are usually constructed
of concrete, have rounded edges, and are capable of handling much larger discharges than sharp-
crested weirs.
Some advantages of broad-crested weirs.
i) It is cheaply and easily fabricated (can be done in-situ or off-site)
ii) It is easy to install
iii) It possesses a wide modular range;
iv) It produces minimum afflux(increases in upstream depth due to weir installation)
v) It requires a minimum of maintenance.

Flow over a broad-crested weir must pass through the critical depth somewhere near the
downstream corner. Broad-crested weirs are characterized by the presence of two drops in the
water level over the weir.

1 2
ZB(o) ZB

Ho H
Zc
v h

By taking Bernoulli‘s equation in sections 1 and 2 , the velocity of flow on the weir is obtained
as
v  cv 2 gZ B0  cv 2 g H o  h

137
For a rectangular weir, the discharge is given by:
Q = Av = b x h x v = bhcv 2 g H o  h ----------(*)
q  hcv 2 g H o  h  f h ---------------------(**)
As can be seen from the above equation, in order to determine the discharge, there is the need to
first find h, the depth of flow on the weir. There are many methods of finding h and Q but the
Belanch‘s method based on the principle of maximum discharge is much more appealing.
As can be seen from the above equation, the discharge is only a function of h and the value of h
varies between 0 ≤ h ≤ Ho however the discharge goes to zero as h = 0 and as h = Ho
Since the function (**) is a continuous function and the function goes to zero at the two limits,
then it follows that there must be a turning point (maximum or minimum) within the limits. This
is the basis of Belanch‘s principle which states that: for a given head Ho on a broad-crested weir,
the depth that will be established on the weir is such as to result in the maximum discharge over
the weir.
With the above postulate, the required depth is obtained by differentiating q with respect to h and
putting everything to zero as:

dq 
d c v h 2 g H o  h  
 0
dh dh

Ignoring the constants and differentiating

 
1
d h Ho  h h
Ho  h  2 0
dh Ho  h

which gives h = 2/3Ho


Substituting this value of h into (*) we

2  2  4
Q  b. H o cv 2 g  H o  H o   bcv 2 g H o3 / 2  0.385cvb 2 g H o3 / 2
3  3  27
Q  CDb 2 g H o3 / 2
q  CD 2 g H o3 / 2

where CD = 0.385 Cv , CD ranges between 0.32 to 0.36.

Ventilation of Weirs
When flow takes place over a suppressed weir with wing walls on the downstream side, the free
access of air below the nappe is prevented. The air entrapped between the nappe and the weir
face is gradually evacuated by the dynamic action of the flowing fluid. This eventually leads to
subatmospheric pressure in the underside of the nappe bringing about increases in the discharge.

138
Investigations by Bazin have shown that if the flow is sufficient to prevent the air having access
to the underside of the nappe, it may assume one of the forms shown below

Sub-atmospheric
pressure

Atmospheric
Pressure

Fully aerated nappe Drowned nappe Clinging nappe

Fully aerated (free, ventilated) nappe: In this type, the weir discharges free as the pressure
below the nappe is maintained at atmospheric by proper ventilation.

Drowned (depressed) nappe: When the underside of the nappe is poorly aerated, the air enclosed
between the nappe and the downstream face of the weir gradually mixes up with the flowing
fluid and is carried away. This causes the pressure under the napper to fall below atmospheric.
The pressure difference on the two sides of the nappe-(the upper side exposed to atmospheric
while the bottom side has sub-atmospheric pressure) –causes the nappe to deflect towards the
weir. The discharge over such a weir is greater than that of the ventilated weir.

Clinging nappe: This type occurs when there is no ventilation and all the air under the napper
has been carried away by the dynamic action of the flowing fluid. The nappe is forced to adhere
to the downstream face of the weir by the atmospheric pressure on the upper side. In this type,
the discharge is 20 to 30 per cent more than in free nappe.

Knowledge of these types of nappes is necessary especially when the weir is used for flow
(discharge) measurement. All the discharge formulae have been derived for free nappe and
therefore some adjustment in the measured flow must be made when other forms of nappe occurs
during discharge measurement. For discharge measurement it is always advisable to ensure that
the nappe is fully aerated.

139
FLUMES
Although weirs are the cheapest and simplest structures for flow measurement in open channels,
however the relatively high losses caused by weirs and the tendency of sedimentation of
suspended particles due to reduced velocities in the vicinity of the weir, could in certain cases
pose important limitations Flumes provide a convenient alternative to weirs for flow
measurement in open channels where high head losses and sedimentation are of concern. Such
cases include flow measurement in wastewater treatment plants and irrigation channels with
suspended particles

Flumes are devices in which the flow is locally accelerated by a streamlined lateral contraction in
the channel sides. A flume has: a convergent section, in which the flow accelerates, a throat section
and a divergent section, in which the flow returns to normal
Flumes are of two types: non-modular or the venturi flume and the modular or the standing
wave flume.

In the non-modular flume, the velocity at the throat is maintained below the critical value so that
no standing wave is produced. However, in the modular type, the flume is designed such that the
velocity of flow at the throat is greater than the critical velocity thereby resulting in a standing
wave within the flume.

Non-Modular Flume (Venturi Flume)


When the width of a channel is reduced while the bed remains flat, the discharge per unit width
increases. If the losses are neglected, the specific energy remains constant. If the conditions are
made such that the free surface does not pass through the critical depth, the arrangement forms a
Venturi flume.
Referring to the diagram above, for continuity of flow,

h1 x b1 x v1 = h2 x b2 x v2 --continuity equation for section 1-1 and 2-2.

Applying Bernoulli‘s equation to sections 1-1 and 2-2

h1 + v12/2g = h2+v22/2g

Substituting v1 for continuity equation,

h1-h2 = v22/2g[1-(h2b2/h1b2)2]

2 g h1  h2 
v2 
  b h 2 
1   2 2  
  h1b  
 
2 g h1  h2 
Q  v2 A2  b2 h2 ..    Theoretica l..disch arg e
  b h 2 
1   2  
  b h1  
 

140
1
2

b1 b2

PLAN
1 2
v12/2g
v22/2g
h1
h2
Critical depth

ELEVATION

Introducing the discharge coefficient, CD = (Actual discharge)/(Theoretical discharge)

2 g h1  h2 
2 g h1  h2 
A1 A2
Q  CDb2 h2  CD
  b h 2  A12  A22
1   2 2  
  b1h1  

CD = 0.95 to o.99

Therefore to measure the flowrate, water depth are measured at two locations, one at section 1-1
and the other at section 2-2.

Modular Flume or Standing Wave Flume (eg. Parshall Flume)


As already mentioned, modular flumes are designed to make the flow pass through the critical
depth in some section of the throat.
By considering a section upstream with depth h1 and another section at the throat with depth h2,
and applying Bernoulli‘s equation,

h1 + v12/2g = h2 + v22 = H

v2 = (2g(H-h2)1/2

Therefore the discharge Q = b2 h2v2 = b2h2[2g(H-h2)] 1/2

141
  
Q  b2 2 g Hh22  h3  b2 2 g Hh2  h3 
1/ 2

For the maximum discharge, the quantity (Hh2-h3)1/2 should be maximum


d Hh22  h3 
0
dh
Therefore 2 Hh2  3h22  0
2
h2  H
3
Thus for the flow to be maximum, the depth at the throat section should be two-thirds the total
energy.
 3h 
For this condition, the velocity v2  2 g H  h2   2 g  2  h2   gh2
 2 

This implies that for maximum discharge, the depth at the throat must be equal to the critical
depth and the velocity critical velocity.
Substituting h2 = 2/3H in the discharger expression, we shall obtain

1/ 2
2  2  2
Q  b2 2g H  H  H   2 gb2 H 3 / 2  1.705b2 H 3 / 2  1.705b2 (h1  v12 / 2 g )3 / 2
3  3  3 3

Although flat bed modular flume is easier to construct, it is sometimes necessary to raise the
invert in the throat to attain critical conditions. For the flat or raised floor, the throat length
should ideally be sufficient to ensure that the curvature of water surface is small to make the
water surface parallel to the invert. Such ideal flumes become relatively long and expensive. A
more compact, short and less expensive groups of flumes have been constructed. However, the
surface water profile varies rapidly and as such the theoretical analysis for ordinary flumes is not
wholly applicable. Therefore empirical relationship are usually developed for such flumes (e.g;
Parshall Flume), which make each such flume standard with its unique calibration curve.

The Parshall Flume


The Parshall flume is a widely-used discharge measurement structure. Figure 76 shows its
general form. The characteristics of Parshall flumes are:
i) Small head losses, ii) Free passage of sediments, iii) Reliable measurements even when
partially submerged and iv) Low sensitivity to velocity of approach

The Parshall flume consists of a converging section with a level floor, a throat section with a
downward sloping floor and a diverging section with an upward sloping floor. Flume sizes are
known by their throat width and each size has its own characteristics, which is unique (see Table
31).

142
The flow through the Parshall flume can occur either under free flow or under submerged flow
conditions. Under free flow the rate of discharge is solely dependent on the throat
width and the measured water depth, ha. The water depth is measured at a fixed point in
the converging section.
The upstream water depth-discharge relationship, according to empirical calibrations, has the
following general form:

Q = K x (ha)u

Where: Q - Discharge (m3/s); ha Water depth in converging section (m); K -A fraction, which
is a function of the throat width; u Variable, lying between 1.522 and 1.60

Table 32 gives the values for K and u for each flume size.

When the ratio of gauge reading hb to ha exceeds 60% for flumes up to 9 inches, 70% for flumes
between 9 inches and 8 feet and 80% for larger flume sizes, the discharge is reduced due to
submergence. The upper limit of submergence is 95%, after which the flume ceases to be an
effective measuring device because the head difference between ha and hb becomes too small,
such that a slight inaccuracy in either head reading results in a large discharge measurement error

143

You might also like