You are on page 1of 86

COLLEGE OF ENGINEERING

& TECHNOLOGY (CoETEC)

EMG 2205: FLUID MECHANICS I

Gı̃tahi

2023/2024 Academic Year


TABLE OF CONTENTS

PREAMBLE v

1.0 PROPERTIES OF FLUIDS AND FLUID STATICS 1


1.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.1 Fluid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.2 Types of fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1.3 Fluid mechanics . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Properties of fluids . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.1 Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.2 Mass density . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.3 Relative density . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.4 Surface tension . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.5 Capillarity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.6 Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2.7 Vapour pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3 Variation of pressure with depth . . . . . . . . . . . . . . . . . . . . . . 12
1.4 Pressure measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.4.1 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.4.2 The barometer and atmospheric pressure . . . . . . . . . . . . . 16
1.4.3 Measurement of gauge pressure . . . . . . . . . . . . . . . . . . 17
1.4.4 Modifications of the simple U-tube manometer . . . . . . . . . . 20
1.4.5 Measurement of differential pressure (pressure difference) . . . . 21
1.5 Hydrostatic thrust on submerged surfaces . . . . . . . . . . . . . . . . 27
1.5.1 Pressure force (thrust) . . . . . . . . . . . . . . . . . . . . . . . 27
1.5.2 Centre of pressure for the plane surface . . . . . . . . . . . . . . 28
1.6 Fluids in relative equilibrium and under constant acceleration . . . . . 31
1.6.1 Fluids subjected to horizontal acceleration . . . . . . . . . . . . 31
1.6.2 Fluids subjected to vertical acceleration . . . . . . . . . . . . . 33
1.7 Buoyancy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
1.7.1 Concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
1.7.2 The hydrometer . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
1.8 Stability of immersed and floating bodies . . . . . . . . . . . . . . . . . 38
1.8.1 Immersed body . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
1.8.2 Floating bodies . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

i
1.8.3 Determination of metacentric height . . . . . . . . . . . . . . . 41

2.0 FLUID DYNAMICS 43


2.1 Lagrangian and Eulerian viewpoints . . . . . . . . . . . . . . . . . . . . 43
2.1.1 Lagrangian approach . . . . . . . . . . . . . . . . . . . . . . . . 43
2.1.2 Eulerian approach . . . . . . . . . . . . . . . . . . . . . . . . . 44
2.1.3 Pathline, streamline and streamtube . . . . . . . . . . . . . . . 45
2.1.4 Real and ideal fluids . . . . . . . . . . . . . . . . . . . . . . . . 46
2.1.5 Steady and unsteady flow . . . . . . . . . . . . . . . . . . . . . 47
2.1.6 Uniform and non-uniform flow . . . . . . . . . . . . . . . . . . . 47
2.1.7 Laminar flow and turbulent flow: Reynolds experiment . . . . . 47
2.2 Continuity of flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
2.2.1 Continuity equation for a general case . . . . . . . . . . . . . . 48
2.2.2 Continuity equation for steady, one-dimensional flow . . . . . . 50
2.3 Navier-Stokes equations . . . . . . . . . . . . . . . . . . . . . . . . . . 55
2.3.1 Newton’s law of viscosity in Cartesian co-ordinates (x,y,z) . . . 55
2.3.2 The momentum equation . . . . . . . . . . . . . . . . . . . . . . 57
2.4 Euler’s and Bernoulli’s equations . . . . . . . . . . . . . . . . . . . . . 61
2.4.1 Euler’s equation of motion along a streamline . . . . . . . . . . 61
2.4.2 Bernoulli’s equation . . . . . . . . . . . . . . . . . . . . . . . . . 63
2.4.3 Energy approach to Bernoulli’s equation . . . . . . . . . . . . . 65
2.4.4 Application of Bernoulli’s equation . . . . . . . . . . . . . . . . 66
2.4.5 Notches and weirs . . . . . . . . . . . . . . . . . . . . . . . . . . 73

ii
LIST OF FIGURES

Figure 1.1 Pascal’s principle . . . . . . . . . . . . . . . . . . . . . . . . . 3


Figure 1.2 Surface tension . . . . . . . . . . . . . . . . . . . . . . . . . . 4
Figure 1.3 Capillarity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
Figure 1.4 Viscosity: Newton’s law . . . . . . . . . . . . . . . . . . . . . 7
Figure 1.5 Shear stress vs shear rate curves for various substances . . . . 9
Figure 1.6 Cylindrical element at arbitrary orientation . . . . . . . . . . . 13
Figure 1.7 Liquid with a free surface . . . . . . . . . . . . . . . . . . . . . 14
Figure 1.8 Pressure in glycerine . . . . . . . . . . . . . . . . . . . . . . . 15
Figure 1.9 Mercury barometer . . . . . . . . . . . . . . . . . . . . . . . . 17
Figure 1.10 Aneroid barometer . . . . . . . . . . . . . . . . . . . . . . . . 18
Figure 1.11 Piezometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Figure 1.12 Simple U-tube manometer . . . . . . . . . . . . . . . . . . . . 19
Figure 1.13 Single-column manometer . . . . . . . . . . . . . . . . . . . . 20
Figure 1.14 Inclined leg manometer . . . . . . . . . . . . . . . . . . . . . . 21
Figure 1.15 Differential U-tube manometer . . . . . . . . . . . . . . . . . . 22
Figure 1.16 Inverted differential U-tube manometer . . . . . . . . . . . . . 23
Figure 1.17 Bourdon gauge . . . . . . . . . . . . . . . . . . . . . . . . . . 24
Figure 1.18 Simple U-tube manometer . . . . . . . . . . . . . . . . . . . . 25
Figure 1.19 Differential U-tube manometer . . . . . . . . . . . . . . . . . . 26
Figure 1.20 Inverted differential manometer . . . . . . . . . . . . . . . . . 26
Figure 1.21 U-tube manometer . . . . . . . . . . . . . . . . . . . . . . . . 27
Figure 1.22 Hydrostatic thrust on a submerged plane . . . . . . . . . . . . 27
Figure 1.23 Submerged circular surface . . . . . . . . . . . . . . . . . . . . 30
Figure 1.24 Fluid mass subjected to horizontal acceleration . . . . . . . . . 31
Figure 1.25 Forces on fluid element A . . . . . . . . . . . . . . . . . . . . . 32
Figure 1.26 Fluid mass subjected to vertical acceleration . . . . . . . . . . 33
Figure 1.27 Hydrometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
Figure 1.28 Iceberg in sea water . . . . . . . . . . . . . . . . . . . . . . . . 37
Figure 1.29 Floating body . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
Figure 1.30 Hydrometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
Figure 1.31 Stability of a ship . . . . . . . . . . . . . . . . . . . . . . . . . 40
Figure 1.32 Metacentre and metacentric height . . . . . . . . . . . . . . . 40
Figure 1.33 Determining the metacentric height . . . . . . . . . . . . . . . 41

Figure 2.1 Streamline . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

iii
Figure 2.2 Streamtube . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
Figure 2.3 Velocity profiles for one dimensional flow . . . . . . . . . . . . 47
Figure 2.4 Reynolds apparatus . . . . . . . . . . . . . . . . . . . . . . . . 48
Figure 2.5 Cartesian control volume for the continuity equation . . . . . . 49
Figure 2.6 Continuous flow through a streamtube . . . . . . . . . . . . . 51
Figure 2.7 A branched pipe . . . . . . . . . . . . . . . . . . . . . . . . . . 52
Figure 2.8 Pipe network . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
Figure 2.9 Viscous stresses . . . . . . . . . . . . . . . . . . . . . . . . . . 57
Figure 2.10 Cartesian control volume for the momentum equation . . . . . 58
Figure 2.11 Deriving Euler’s equation . . . . . . . . . . . . . . . . . . . . . 62
Figure 2.12 Converging pipe . . . . . . . . . . . . . . . . . . . . . . . . . . 64
Figure 2.13 The venturimeter . . . . . . . . . . . . . . . . . . . . . . . . . 66
Figure 2.14 Venturimeter for Example 10 . . . . . . . . . . . . . . . . . . . 69
Figure 2.15 The flow-nozzle . . . . . . . . . . . . . . . . . . . . . . . . . . 70
Figure 2.16 Pipe orifice meter . . . . . . . . . . . . . . . . . . . . . . . . . 71
Figure 2.17 Stagnation point . . . . . . . . . . . . . . . . . . . . . . . . . 72
Figure 2.18 Pitot tube . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Figure 2.19 Discharge from a rectangular notch . . . . . . . . . . . . . . . 74
Figure 2.20 Discharge from a vee notch . . . . . . . . . . . . . . . . . . . . 75
Figure 2.21 Flow over a weir . . . . . . . . . . . . . . . . . . . . . . . . . . 76

iv
PREAMBLE

Purpose
The aim of this course is to enable the students to;

1. understand the nature of fluids and their behaviour as distinct from that of solids.

2. understand fluid statics as applicable in manometry and forces in submerged


surfaces.

3. apply Bernoulli’s equation in measurements of fluid flow.

Learning outcomes
At the end of this course, the students should be able to:

1. Describe what a fluid is and distinguish between liquids and gases.

2. Analyze the behaviour of a liquid at rest and in motion and apply the knowledge
in manometry and calculation of forces in submerged surfaces.

3. Identify and use flow measurement devices to measure common fluid parameters
like pressure, velocity and discharge, in closed conduits and open channels.

Course Description
Properties of fluids: Nature, density, viscosity, vapour pressure, surface tension and
capillarity.
Fluid statics: Pressure distribution, Pascal’s law, pressure gauges and manometers.
Forces on submerged surfaces. Fluids in relative equilibrium and under constant ac-
celeration.
Fluid dynamics: Conservation equations; mass conservation, steady flow energy
equation, Navier-Stokes, Euler and Bernoulli’s equations. Flow measurement in closed
conduits and open channels; venturimeter, orifice meters, flow nozzle, rotameter, rect-
angular weir and triangular weir. Pitot tubes.

Prerequisite(s)
SPH 2173 PHYSICS FOR ENGINEERS I, SPH 2174 PHYSICS FOR ENGINEERS II

v
Prescribed Textbook(s)
1. Douglas, J. F., Gasiorek, J. M. & Swaffield, J. A. (2001) Fluid Mechanics, Pren-
tice Hall, 2nd Ed.

2. Munson, B. R., Young, D. F. & Okiishi, T. H. (1998) Fundamentals of Fluid


Mechanics, John Wiley & Sons, 3rd Ed.

References
1. Robertson, J. A. & Crowe, C. T. (1997) Engineering Fluid Mechanics, John Wiley
& Sons, 9th Ed.

2. Bansal, R. K. (1992) A Textbook of Fluid Mechanics and Hydraulic Machines,


Laxmi Publications, 4th Ed.

3. Journal of Fluids Engineering.

Teaching Methodology
2-hour lecture and 1-hour tutorial every week. At least two laboratory sessions in
the semester.

Assessment
70% written exam, 30% continuous assessment

Important Requirements
1. A scientific calculator

2. A workbook for tutorials and other exercises

vi
Semester Activities
WEEK DATES ACTIVITY
1 Jan 4 - 5 Registration
2 Jan 8 - 12 Lecture: Course description; FLUID STATICS:
Introduction; Definitions & properties of fluids.
3 Jan 15 - 19 Lecture: Viscosity - Newtonian & Non-Newtonian fluids;
Pressure measurements - atmospheric pressure;
Gauge pressure - piezometers & U-tube manometers.
4 Jan 22 - 26 Lecture: Modifications of U-tube manometers;
Hydrostatic thrust on submerged surfaces. Assignment 1.
5 Jan 29 - Feb 2 Lecture: Fluids in relative equilibrium; Buoyancy;
Stability of immersed & floating bodies.
6 Feb 5 - 9 CAT 1
7 Feb 12 - 16 Lecture: FLUID DYNAMICS:
Lagrangian & Eulerian viewpoints; Types of flows -
steady/unsteady, uniform/non-uniform, laminar/turbulent;
Continuity of flow.
8 Feb 19 - 23 Lecture: Navier-Stokes equations.
9 Feb 26 - Mar 1 CAT 2
10 Mar 4 - 8 Lecture: Euler’s & Bernoulli’s equations;
Application of Bernoulli’s equation - Venturimeter.
Assignment 2 & Beginning of Practicals
11 Mar 11 - 15 Lecture: More applications of Bernoulli’s equation -
flow nozzle, orifice meter, Pitot tube; Notches & weirs.
12 Mar 18 - 22 CAT 3
13 Mar 25 - 29 Tutorials
14 Apr 1 - 5 Revision
15 Apr 8 - 12 EXAMS
16 Apr 15 - 19 EXAMS

Notable Dates
Continuous Assessment Tests (CATS)
• CAT 1 - Week 6: Feb 8, Time; 2.00 pm - 4.00 pm

• CAT 2 - Week 9: Feb 29, Time; 2.00 pm - 4.00 pm

vii
• CAT 3 - Week 12: Mar 21, Time; 2.00 pm - 4.00 pm

Laboratory Sessions
Week 7 - Week 12

End-of-Semester Examinations
April 8 - April 19

viii
1.0 PROPERTIES OF FLUIDS AND
FLUID STATICS
1.1 Definitions
1.1.1 Fluid
A fluid is a material substance that can flow. A fluid cannot sustain a shearing stress
when at rest. Indeed, it is the inability of fluids at rest to resist shearing stresses that
gives them their characteristic ability to flow, i.e., change shape [1, 2, 3].

1.1.2 Types of fluids


Fluids are conventionally classified as either liquids or gases.

1. A liquid is a substance that under suitable conditions of temperature, will in time


deform to take up the shape of any container into which it is placed. A liquid
is relatively incompressible, and if all pressure, except its own vapour pressure
is removed, the cohesion between the molecules holds them together so that the
liquid does not expand indefinitely. Therefore, a liquid may form a free surface.

2. A gas is a substance that will deform and expand to occupy the entire volume
of any container in which it is placed without developing a free surface. Unlike
liquids, gases are highly compressible.

1.1.3 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 the mechanics of solids. Primarily,
fluid mechanics deals with the action of forces on fluids. Fluid mechanics may be
divided into three branches:

1. Fluid statics: This is the study of the mechanics of fluids at rest.

2. Kinematics of fluids: This deals with the velocities and streamlines of fluid flow
without considering the forces and energy involved.

3. Fluid dynamics: This is concerned with the relations between velocities and
accelerations of fluid flow and the forces exerted by or upon fluids in motion.

Other terms encountered under fluid mechanics include:

1
• Hydrodynamics - this deals primarily with the flow of fluids for which there is
virtually no density changes, e.g., the flow of liquid or gas at low speeds.
Hydraulics (study of liquid flows in pipes or open channels) falls within this
category.

• Gas dynamics - this deals with fluids that undergo significant density changes,
e.g., high speed flows of gas through a nozzle.

• Aerodynamics - this deals with the flow of air past aircraft or rockets, whether
low-speed (incompressible) or high-speed (compressible).

1.2 Properties of fluids


1.2.1 Pressure
The way a surface force acts on a fluid differs from the way it acts on a solid. As stated
earlier, a fluid at rest cannot resist tangential (shearing) stresses. This is because the
fluid layers would simply slide over one another when subjected to such forces/stresses.
It is therefore convenient to describe the forces acting on a fluid by specifying the pres-
sure p, which is defined as the magnitude of the normal force per unit surface area.
Note that in a solid, because of the possibility of tangential stresses between adjacent
particles, the stresses at any given point may be different in different directions. How-
ever, in a fluid at rest, no tangential stresses can exist, and the only forces between
adjacent surfaces are pressure forces normal to the surfaces [2, 4, 5].
STATEMENT: The pressure at any given point in a fluid at rest is the same
in every direction.
PROOF:
Figure 1.1 shows a very small wedge-shaped element of fluid at rest whose thickness
perpendicular to the plane of the paper is constant and equal to δy. Let p be the
average pressure in a direction in the plane of the paper as shown. Let px and pz be
the average pressures in the horizontal and vertical directions respectively. The specific
weight of the fluids is γ = ρg. The forces acting on the element of fluid in the plane
shown can be resolved in the x and z directions; and since the fluid is at rest (at a state
of equilibrium), the algebraic sum of the force components in any direction is equal to
zero.

2
Figure 1.1: Pascal’s principle

In the x -direction:

pδlδycosα − px δyδz = 0, (1.1)


δlcosα = δz, (1.2)
∴ pδyδz − px δyδz = 0, (1.3)
p = px (1.4)

In the z -direction:
1
pz δxδy − pδlδysinα − γδxδyδz = 0, (1.5)
2
but the third term (weight of the fluid element) is small enough compared to the
pressure forces and may thus be neglected. Therefore,

pz δxδy − pδlδysinα = 0, (1.6)

but again from the triangle, δlsinα = δx, such that

pz δxδy − pδxδy = 0, (1.7)


pz = p. (1.8)

Therefore, p = px = pz . This result is independent of α and shows that pressure at


any point in a fluid at rest is the same in all directions. This is Pascal’s Principle.

3
1.2.2 Mass density
This, also simply ecalled density and denoted by ρ, is the mass per unit volume of a
fluid. Its units of measure are kg/m3 , e.g., the mass density of water at 40 C is 1000
kg/m3 .

1.2.3 Relative density


It is also called specific gravity and is the ratio of the density of a substance to that of
water at 40 C. It is a numeric without units.

1.2.4 Surface tension


Liquids have cohesion and adhesion, both of which are forms of molecular attraction.
Cohesion enables a liquid to resist tensile stress, while adhesion enables it to adhere
to another body. The attraction between molecules is greatest near a free surface or
an interface, and leads to formation of an imaginary film which acts like a stretched
membrane capable of resisting tension at the free surface or at the interface between two
immiscible liquids or at the interface between a liquid and a gas. The liquid property
that creates this capability is known as surface tension, σ [2, 3, 4].
The value of surface tension depends on:

1. The nature of the liquid

2. The nature of the substance with which it is in contact at the surface

3. The temperature and pressure

Figure 1.2: Surface tension

Surface tension increases the pressure inside a small droplet of liquid, which is measured
with reference to the atmospheric pressure. To determine a relation between the inside

4
pressure, and the surface tension, consider a hemispherical portion of this droplet,
i.e., one half of the droplet. The surface tension force and the pressure force are in
equilibrium and so must be equal.
NOTE: The surface tension force acts along the circumference while the pressure force
d
acts over the cross-sectional area. Therefore, since the radius r = ,
2

σ(2πr) = p(πr2 ) or p = . (1.9)
r
1.2.5 Capillarity
This is due to both cohesion and adhesion. When cohesion has less effect than adhesion,
the liquid will wet a solid surface with which it is in contact and rise at the point of
contact. If cohesion predominates, the liquid surface will be depressed at the point of
contact. For example, capillarity makes water rise much higher than mercury in a glass
tube.
From Fig. 1.3, equating the lifting force created by surface tension to the gravity force,

Figure 1.3: Capillarity

2πrσcosθ = πr2 hρg, (1.10)


2σcostθ
h= , (1.11)
rρg

5
where σ = surface tension (units of measure: force per unit length, e.g., N/m); θ =
wetting angle; ρ = mass density of liquid; r = radius of tube; h = capillary rise.
Note that for tube diameters larger than 12 mm, capillary effects are virtually negligi-
ble. In a clean glass tube, θ = 00 for water and 1400 for mercury.
Tutorial Problems 1

1. A small drop of water is in contact with air and has a diameter of 0.05 mm. If
the pressure within the droplet is 565 Pa greater than the atmosphere, what is
the value of the surface tension? [7.06 x 10−3 N/m]

2. A spherical soap bubble has a radius R, a film thickness t, and a surface tension
σ. Derive a formula for the pressure within the bubble relative to the outside
atmospheric pressure. What is the value of this pressure for a bubble of 4 mm
radius? Assume σ is the same as for pure water at room temperature, i.e., 0.073
N/m. [73 N/m2 ]

3. Given that at 21.10 C, water has a surface tension of 0.0725 N/m, estimate the
height to which water at 21.10 C will rise in a capillary tube of diameter 3 mm.
[9.85 mm]

1.2.6 Viscosity
The viscosity of a fluid is its internal friction or its tendency to resist flow. It is a
measure of the fluid’s resistance to shear or angular deformation. As the temperature
increases, the viscosity of liquids decreases, while the viscosity of gases increases.
Consider a 3D rectangular element of fluid as shown in Fig. 1.4. The shearing force F
acts tangentially on the area on the top of the element. This area is given by A = δzδx.
The shear stress can be calculated as
Tangential Force, F
shear stress = (1.12)
Area, A
Shear stress is a vector that possesses both magnitude and direction. The symbol for
shear stress is τ . The deformation which this shear stress causes is measured by the
size of the angle φ and is known as shear strain. In a solid, shear strain φ is constant
for a fixed shear stress γ. In a fluid, φ increases for as long as τ is applied (the fluid
flows).
It is found experimentally that in a fluid, the rate of shear strain is directly pro-
portional to the shear stress, i.e.

shear stress ∝ shear strain rate (1.13)

6
Figure 1.4: Viscosity: Newton’s law

If the particle at a point E (which is at a height y above point D in the figure) moves
under the shear stress to point E’ and it takes time t to get there, it will have moved
a distance x. For small deformations, the shear strain is expressed thus,
x
φ= , (1.14)
y
while the shear strain rate is expressed,
φ
rate of shear strain =
t
x
=
ty
u
= , (1.15)
y
x
where = u, is the velocity of the particle at E. Using the experimental result that
t
shear stress is proportional to the rate of shear strain, then, from Eq. (1.13),
 
u
τ = constant × . (1.16)
y
u
The term is the change in velocity with y, or the velocity gradient, and may be
y
du
written in the differential form . This velocity gradient is also termed as the shear
dy
rate, denoted by the symbol γ̇ and expressed in the units of s−1 . The constant of
proportionality is known as the viscosity or dynamic viscosity, µ, of the fluid, giving
du
τ =µ or, (1.17)
dy
τ = µγ̇. (1.18)

7
Equation (1.17), or (1.18) is known as Newton’s Law of Viscosity. Fluids obeying
this law, i.e., those that have a constant µ, are called Newtonian fluids, e.g., water,
while fluids in which µ is not constant are non-Newtonian, e.g., milk. A fluid is said
to be ideal if it is frictionless or inviscid (µ = 0) - but this is only an idealization, it
does not occur in practice.
In many problems involving viscosity, there frequently appears the value of viscosity
divided by density. This is defined as the kinematic viscosity, ν
µ
ν= . (1.19)
ρ

EXAMPLE 1
The viscosity of water at 200 C is given as 0.01008 poise. Compute

1. the dynamic viscosity in Pa.s

2. the value of the kinematic viscosity in m2 /s, if the relative density at 200 C is
0.998

SOLUTION 1

1. Dynamic viscosity (µ);

µ = 0.01008 poise × 0.1 = 1.008 × 10−3 Ns/m2 .

2. Kinematic viscosity (ν);


µ
ν=
ρ
1.008 × 10−3
= = 1.01 × 10−6 m2 /s.
0.998 × 1000

1.2.6.1 Newtonian fluids


In Newtonian fluids, the shear stress is directly proportional to the shear rate and the
plot begins at the origin. A Newtonian fluid can be defined by a single value of viscosity
at a specified temperature.
Examples
Water, mineral and vegetable oils, pure sucrose solutions, most carbonated beverages
[6, 7].

8
Figure 1.5: Shear stress vs shear rate curves for various substances

1.2.6.2 Non-Newtonian fluids


These are materials which cannot be defined by a single value of viscosity at a specified
temperature. The viscosity must be stated together with a corresponding temperature
and shear rate.

1. Time-independent non-Newtonian behaviour


These non-Newtonian fluids could be shear thinning, shear thickening or plastic.
Their viscosities, though varying with magnitude of shear rate, do not vary with
respect to the duration of application of shear [6].

(a) Shear thinning flow behaviour


The curve begins at the origin of the shear stress - shear rate plot but an
increasing shear stress gives a more than proportionate increase in the shear
rate. Shear thinning fluids are also regarded as pseudoplastic fluids.
Examples
Blood, clay, yoghurt, fruit juice concentrates, salad dressings, tomato ketchup,
banana puree, milk.
(b) Shear thickening flow behaviour
The curve begins at the origin of the τ − γ̇ plot but an increasing shear stress
gives a less than proportionate increase in shear rate. Shear thickening fluids
are also referred to as dilatants. This type of flow behaviour is generally

9
found among suspensions of very high concentrations.
Examples
Concentrated starch suspension, wet sand.
(c) Plastic flow behaviour
A fluid which exhibits a yield stress is called a plastic fluid. For this type of
flow behaviour, a significant force must be applied before the material starts
to flow. Below the yield stress, the fluid does not flow (strain rate = 0).
Once the yield stress is exceeded, the fluid can flow either like a Newtonian
fluid - hence a Bingham plastic fluid; or it could flow like a shear thinning
fluid - hence a viscoplastic fluid. Plastic flow is common in foods.
Examples
High pectin pineapple juice concentrate, tomato paste, certain ketchups.
Other examples include handcream and toothpaste.
NOTE: A simple way of checking a fluid’s possible plastic properties is to
turn its container upside down. If the fluid will not flow by itself, it probably
has a significant yield value. If it flows by itself but very slowly, it probably
has no yield value but a high viscosity.

2. Time-dependent non-Newtonian behaviour


Non-Newtonian fluids may be time-dependent in which case the viscosity is a
function not only of the magnitude of the shear rate but also of the duration of
shear application.

(a) Thixotropic flow behaviour


A thixotropic fluid can be described as a time-dependent shear thinning
system where the viscosity decreases with increasing shear rate as well as
the duration of application of shear. This type of flow behaviour is exhibited
by all gel-forming systems.
Examples
Yoghurt, mayonnaise, ice-cream, soft cheese, tomato ketchup. Most paints
also fall in this category.
(b) Rheopectic flow behaviour
A rheopeptic fluid can be described as a time-dependent shear thickening
system, i.e., where the viscosity increases with increasing shear rate and
also with the duration of application of the shear.
Examples
Whipping cream, some fruit juice concentrates.

10
1.2.6.3 Flow behaviour models
For the adaptation of viscosity measurement data to process design calculations, some
kind of mathematical description of the flow behaviour is required. Several models are
available e.g., Ostwald, Herschel-Bulkley, Steiger-Ory, Bingham, Ellis & Eyring. These
models relate the shear stress of a fluid to the shear rate, thus enabling the apparent
viscosity to be calculated as a ratio between the shear stress and the shear rate [6].
The most general model is the Herschel-Bulkley or the generalized power equation.
It is applicable to a great number of non-Newtonian fluids over a wide range of shear
rates. Additionally, the power equation lends itself readily to mathematical treatment
e.g., in pressure drop and heat transfer calculations.

(τ − τ0 ) = K γ̇ n , (1.20)

where,
τ = Shear stress (units: Pa or N/m2 )
τ0 = Yield stress (units: Pa or N/m2 )
K = Consistency (units: Pa sn )
γ̇ = Shear rate (units: s−1 )
n = Flow behaviour index (dimensionless)
du
• For Newtonian fluid, τ0 = 0, K = µ and n = 1 such that τ = µγ̇ = µ .
dy
• For shear thinning and shear thickening fluids, τ0 = 0 and τ = K γ̇ n ; where n <
1 and n > 1 respectively.

• For plastic fluids, (τ −τ0 ) = K γ̇ n ; where n < 1 for viscoplastic behaviour and n =
1 for Bingham plastic behaviour.

For time-dependent fluids, the equations are much more complex.


1.2.6.4 Measurement of viscosity
Viscometers are used to measure viscosity and the main types of viscometers are ro-
tational and capillary [6, 7].

• Rotational viscometers could be concentric cylinders (also called coaxial or Cou-


ette type), cone-plate, plate-plate or spindle type.

• Capillary viscometers may be atmospheric or pressurized type.

Rotational viscometers are easier to use and more flexible than capillary viscometers.
On the other hand, capillary viscometers are more accurate at low viscosities and at
high shear rates.

11
1.2.7 Vapour pressure
All liquids tend to evaporate (vapourize), which they do by projecting molecules into
the space above their surfaces. If this space is confined, the partial pressure exerted
by the molecules increases until the rate at which molecules re-enter the liquid is equal
to the rate at which they leave. For this equilibrium condition, the vapour pressure
is known as the saturation pressure [8]. This saturation pressure may also be referred
to as the boiling pressure because, should the equilibrium be disturbed such that the
vapour pressure exceeds the pressure on the liquid even slightly, rapid evaporation
occurs in an attempt to re-establish the equilibrium. Vapour pressure increases with
temperature.

1.3 Variation of pressure with depth


If a fluid is in equilibrium, every portion of the fluid is in equilibrium. Consider the
cylindrical element in Fig. 1.6 inclined at an angle θ to the vertical, length δs, cross-
sectional area A, in a static fluid of mass density ρ. The pressure at the end with
height z1 is p1 and at the end with height z2 = z1 + δz, is p1 + δp.
The forces acting on the element are:

• p1 A; acting at right angles to the end of the face at z1

• (p1 + δp)A; acting at right angles to the end of the face at z1 + δz

• mg = the weight of the element acting vertically down through the centroid of
the element, G (= mass density x volume x gravity = ρAδsg)

For equilibrium of the element, the resultant of forces in any direction is zero. Resolving
the forces in the direction along the central axis gives

p1 A − (p1 + δp)A − ρgAδscosθ = 0 (1.21)


δp = −ρgδscosθ (1.22)
δp
= −ρgcosθ, (1.23)
δs
or in the differential form
dp
= −ρgcosθ. (1.24)
ds
If θ = 900 , i.e. cylindrical element were horizontal, then
 
dp dp dp
= = = 0, (1.25)
ds θ=900 dx dy

12
confirming that pressure change on any horizontal plane is zero, i.e., points that lie at
the same horizontal level in a continuous fluid at equilibrium have equal pressures.
If θ = 00 , i.e. cylindrical element were vertical, then
 
dp dp
= = −ρg, (1.26)
ds θ=00 dz
confirming the result that pressure increases with depth (or decreases with height).
This is the hydrostatic law of pressure [2, 4].

Figure 1.6: Cylindrical element at arbitrary orientation

If a liquid has a free surface, this is the natural level from which to measure dis-
tances, and if the surface is open to the atmosphere, pressure at this surface is atmo-
spheric, p0 . Applying Eq. (1.26) to the situation in Fig. 1.7,
p0 − p1
= −ρg, (1.27)
z2 − z1
∴ p1 = p0 + ρgh, (1.28)

13
Figure 1.7: Liquid with a free surface

where p1 is the absolute pressure at point 1. The height h = z2 − z1 for the static fluid
(fluid at rest), is the head of the fluid which produces the gauge pressure ρgh at point
1.
EXAMPLES 2

1. Convert a pressure head of 4.6 m of water to metres of oil of specific gravity 0.75.

2. Find the pressure at the bottom of a tank containing glycerine under pressure as
shown in Fig. 1.8. Take specific weight of glycerine as 12.34 kN/m3

SOLUTIONS 2

1. Conversion,

ρw ghw = ρo gho
ρw hw 1000 × 4.6
∴ ho = =
ρo 0.75 × 1000
= 6.13 m.

2. Pressure at the bottom of the tank,

p = po + ρgh
= 50 × 103 + 12.34 × 103 × 2 = 74, 680 Pa(or 74.68 kP a).

14
Figure 1.8: Pressure in glycerine

Tutorial Problems 2

1. Calculate the dynamic viscosity of an oil, which is used for lubrication between
a square plate of size 0.8 m x 0.8 m, and an inclined plane with an angle of
inclination of 300 . The weight of the plate is 300 N and it slides down the
inclined plane with a uniform velocity of 0.3 m/s. The thickness of the oil film is
1.5 mm. [1.17 Ns/m2 ]

2. Determine the intensity of shear of an oil having a viscosity of 0.1 Ns/m2 , which
is used for lubricating the clearance between a shaft of diameter 10 cm and its
journal bearing, given that the clearance is 1.5 mm and the shaft rotates at 150
rpm. [52.33 N/m2 ]

3. What will be

(a) the gauge pressure


(b) the absolute pressure

of water at a depth of 12 m below the free surface?


Take the density of water to be 1000 kg/m3 and the atmospheric pressure to be
101 kN/m2 [117.72 kN/m2 ; 218.72 kN/m2 ]

4. What depth of oil of specific gravity 0.8 will produce a gauge pressure of 120
kN/m2 ? What would be the corresponding depth of water? [15.3 m; 12.2 m]

15
5. What would be the pressure in kN/m2 if the equivalent head is measured as 400
mm of

(a) mercury of specific gravity 13.6? [53.4 kN/m2 ]


(b) water? [3.92 kN/m2 ]
(c) oil of specific weight 7.9 kN/m3 ? [3.16 kN/m2 ]
(d) a liquid of density 520 kg/m3 ? [2.04 kN/m2 ]

1.4 Pressure measurements


1.4.1 Definitions
1. Atmospheric pressure
The earth is surrounded by an atmosphere, many miles high. The pressure due
to this atmosphere at the surface of the earth depends upon the head of air above
the surface.
Atmospheric pressure at sea-level is about 101.325 kN/m2 , equivalent to a head
of 10.35 m of water or 760 mm of mercury approximately, and decreases with
altitude.

2. Vacuum
A perfect vacuum is a completely empty space, in which the pressure is absolute
zero.

3. Gauge pressure
This is the intensity of pressure measured above or below atmospheric pressure.

4. Absolute pressure
This is the intensity of pressure measured above absolute zero (absolute zero of
pressure occurs in a perfect vacuum).

• Absolute pressure = atmospheric pressure + gauge pressure

1.4.2 The barometer and atmospheric pressure


1. Mercury barometer
In its simplest form, it consists of a glass tube about 1-m long that is closed at
one end, and which is completely filled with mercury and inverted in a bowl of
mercury. A vacuum, known as Torricellian vacuum, forms at the top of the tube.
The Torricellian vacuum is not a perfect vacuum because of the vapour pressure
of the vapour given off the surface of the mercury.

16
Figure 1.9: Mercury barometer

Pressure at A equals that at B (atmospheric) because the two points are at the
same level, and

pA = pV + ρgh. (1.29)

Under normal circumstances of temperature, pV is too small compared to pA such


that it can be neglected. Therefore,

pA = ρgh, (1.30)

where ρ = density of mercury and h is the rise in the tube. At sea-level, h =


760 mm of mercury. A mercury barometer is thus used to measure atmospheric
pressure. Mercury is employed because its density is sufficiently high for a fairly
short column to be obtained.

2. Aneroid barometer
In this barometer, a metal bellows containing a near-perfect vacuum is expanded
or contracted according to variations in the pressure of the surrounding atmo-
sphere. This movement is transmitted by a suitable mechanical linkage to a
pointer moving over a calibrated scale (students to make a sketch) [1, 4].

1.4.3 Measurement of gauge pressure


1. Piezometer
This is a simple device for measuring moderate pressures of liquids. It consists
of a tube in which the liquid can freely rise. The height of the liquid in the tube
will give the value of pressure head directly [1, 4, 8].

17
Figure 1.10: Aneroid barometer

Figure 1.11: Piezometer

As the top of the tube is open to the atmosphere, the pressure measured is gauge
pressure. Clearly, the device is only suitable for moderate pressures otherwise
the liquid will rise too high in the piezometer and probably overflow.

2. Simple U-tube manometer


The principle employed here is the same as that of a piezometer only that now, the
difficulties that would be associated with an excessively long tube are overcome
by use of a U-tube containing a manometric fluid (usually mercury). While
a piezometer cannot measure pressure in gases, a manometer can be used to
measure the pressure of either liquids or gases. The manometric liquid Q is
immiscible with the fluid R. The mass density of Q is ρman , that of R is ρ.
If B is the level of the interface in the left-hand limb and C is a point at the same
level in the right-hand limb, pressure pB at B = pressure pC at C.

• For the left-hand limb:

18
Figure 1.12: Simple U-tube manometer

pB = pressure pA at A + pressure ρgh1 , due to depth h1 of fluid R, i.e.,

pB = pA + ρgh1 (1.31)

• For the right-hand limb:


pC = pressure pD + pressure ρman gh2 due to the depth h2 of liquid Q; but
pD = atmospheric pressure = 0 gauge pressure. Therefore,

pC = ρman gh2 . (1.32)

Equating pB to pC , i.e., Eqs. (1.31) and (1.32), gives,

pA + ρgh1 = ρman gh2 ,


pA = ρman gh2 − ρgh1 . (1.33)

EXAMPLE 3
A U-tube manometer similar to that shown in Fig. 1.12 is used to measure the gauge
pressure of a fluid R of density ρ = 800 kg/m3 . If the density of the manometric liquid
Q is 13.6 x 103 kg/m3 , what will be the gauge pressure at A if;

1. h1 = 0.5 m and D is 0.9 m above BC ?

2. h1 = 0.1 m and D is 0.2 m below BC ?

19
SOLUTION 3
With reference to Eq. (1.33),

1. pA = 13.6 x 103 x 9.81 x 0.9 - 800 x 9.81 x 0.5 = 116.15 kPa.

2. Students to do. [Ans: -27.47 kPa].

1.4.4 Modifications of the simple U-tube manometer


1. U-tube manometer with one leg enlarged (or single-column manometer)
Industrially, the simple U-tube manometer has the disadvantage that the move-
ment of the liquid in both limbs must be read. But by making the diameter of

Figure 1.13: Single-column manometer

one leg very large compared with the other (Fig. 1.13), it is possible to make the
movement in the large leg very small such that it is only necessary to read the
movement of the liquid in the narrow leg.
Let;
z = rise of liquid 2 in the right limb
h = height of centre of pipe above X-X
pA = pressure at A, which is to be measured
At datum Y-Y:

pA + ρ1 (h + ∆h)g = ρ2 g(z + ∆h),


∴ pA = ρ2 gz − ρ1 (h + ∆h)g + ρ2 g∆h
= ρ2 gz − ρ1 gh + ∆h(ρ2 − ρ1 )g. (1.34)

20
NOTE: X-X is the level of liquid 2 in both limbs when the pressure difference
is zero, i.e., before the manometer is connected to the pipe carrying liquid 1.
When the manometer is connected to the pipe and a difference in pressure arises
between the two limbs, liquid 2 goes down a height ∆h on the left side and rises
a height z on the right side. Therefore, the volume of liquid 2 in the left limb
π
goes down by an amount D2 ∆h. This volume is also equal to the total volume
4
π
rise in the right limb, i.e., d2 z. Equating the two gives,
4

D2 ∆h = d2 z,
2
d2

d
∆h = 2 z = z.
D D

If D is large compared to d, ∆h is negligible, therefore we can approximate from


Eq. (1.34) that

pA = ρ2 gz − ρ1 gh, (1.35)

meaning that, as long as the relative height between the centre of the pipe and
the level of the manometric liquid is known prior to connecting up (height h),
then once the manometer is connected, only height z needs to be measured.

2. U-tube manometer with inclined leg (students to do)

Figure 1.14: Inclined leg manometer

1.4.5 Measurement of differential pressure (pressure differ-


ence)
1. Differential U-tube manometer
The pressure difference between two points in a fluid can be measured either by

(a) measuring the pressure at each point separately and subracting or

21
Figure 1.15: Differential U-tube manometer

(b) using a differential pressure gauge or manometer which directly measures


the difference in pressure.

The pressure at the same level CD in the two limbs must be the same, since the
fluid at the bottom of the U-tube is at rest.

• For the left limb:

pC = pA + ρga. (1.36)

• For the right:

pD = pB + ρg(b − h) + ρman gh. (1.37)

Since pC = pD , we can equate (1.36) and (1.37) thus,

pA + ρga = pB + ρg(b − h) + ρman gh, (1.38)

and the pressure difference is given as

pA − pB = ρg(b − h) + ρman gh − ρga


= ρg(b − a) + (ρman − ρ)gh (1.39)

EXAMPLE 4
A differential U-tube manometer as in Fig. 1.15 is used to measure the pressure
difference between two points A and B in a pipeline conveying water of density ρ
= 103 kg/m3 . The density of the manometric liquid Q is 13.6 x 103 kg/m3 , and
point B is 0.3 higher than point A. Calculate the pressure difference when h =

22
0.7 m.
SOLUTION 4
With reference to Eq. (1.39),
pA − pB = 103 x 9.81 x 0.3 + (13.6 x 103 - 103 ) x 9.81 x 0.7 = 89.467 kPa.

2. Inverted differential U-tube manometer


This is used to measure pressure difference in liquids. The top of the U-tube
is filled with a fluid, which is less dense than that in which pressure is to be
determined.
Since the fluid at the top is at rest, pressures at level X-X will be the same in

Figure 1.16: Inverted differential U-tube manometer

both limbs.

• For the lift limb:

pXX = pA − ρga − ρman gh. (1.40)

• For the right limb:

pXX = pB − ρg(b + h). (1.41)

Equating (1.40) and (1.41),

pA − ρga − ρman gh. = pB − ρg(b + h),


pB − pA = ρg(b − a) + (ρ − ρman )gh. (1.42)

If A and B are at the same level,

pB − pA = (ρ − ρman )gh. (1.43)

23
Further, if the top of the tube is filled with air, ρman is negligible compared to ρ
and so,

pB − pA = ρgh, (1.44)

when A and B are at the same level.


EXAMPLE 5
An inverted U-tube manometer of the type shown in Fig. 1.16 is used to mea-
sure the pressure difference between two points A and B in an inclined pipeline
through which water is flowing. The difference of level, h = 0.3 m, a = 0.25 m
and b = 0.15 m.
Calculate the pressure difference, pA − pB if the top of the manometer is filled
with: (a) air (b) oil of relative density 0.8
SOLUTION 5

(a) In Eq. (1.42), ρman ' 0 since it is air. Therefore,

pB − pA = ρg(b − a) + ρgh
= 1000 × 9.81 × (0.15 − 0.25) + 1000 × 9.81 × 0.3
= 1, 962 N/m2

(b) Students to do [Ans: -392.4 N/m2 ]

3. The Bourdon gauge


Where high precision is not required, a pressure difference may be indicated by
the deformation of an elastic solid.
A bourdon gauge is based on this and is compact, reasonably robust and simple

Figure 1.17: Bourdon gauge

24
to use. A curved tube of elliptical cross-section is closed at one end; this end
is free to move (Fig. 1.17). The other end, through which the fluid enters,
is rigidly fixed to the frame as shown. The movement of the free end of the
tube is transmitted by a suitable mechanical linkage to a pointer moving over a
scale. Zero reading is obtained when the pressure inside the tube equals the local
atmospheric pressure.

Tutorial Problems 3

1. In Fig. 1.18, fluid R is water and fluid Q is mercury. If the specific weight is
13.6 times that of water and the atmospheric pressure is 101.3 kN/m2 , what is
the absolute pressure at A when h1 = 15 cm and h2 = 30 cm? [59.8 kN/m2 ]

Figure 1.18: Simple U-tube manometer

2. The U-tube differential manometer in Fig. 1.19 measures the pressure difference
between two points A and B in a liquid of density ρ1 . The U-tube contains
mercury of density ρ2 . Calculate the difference of pressure if a = 1.5 m, b = 0.75
m and h = 0.5 m given that the liquid at A and B is water and ρ2 = 13.6 ρ1 .
[54.4 kN/m2 ]

3. The liquid in a piezometer stands 1.5 m above a point A in a pipeline. What is


the pressure at A in N/m2 if the liquid is:

(a) water? [14.7 kPa]


(b) oil of specific gravity 0.85 [12.5 kPa]
(c) mercury of specific gravity 13.6? [200 kPa]
(d) brine of specific gravity 1.24? [18.2 kPa]

25
Figure 1.19: Differential U-tube manometer

4. In Fig. 1.20, fluid A is water and fluid B is oil of specific gravity 0.9. If h = 69
cm and z = 23 cm, what is the difference of pressure in kN/m2 between M and
N ? [1.58 kPa]

Figure 1.20: Inverted differential manometer

ASSIGNMENT 1

1. What is the gauge pressure of the air in the tank shown in Fig. 1.21 if: l1 is
40cm, l2 is 100 cm and l3 is 80 cm? Take the specific gravity of mercury as 13.6,
that of the oil as 0.8 while the density of air can be neglected.

2. A shaft of diameter 100 mm is rotating inside a journal bearing of diameter 102


mm at a speed of 360 rpm. The space between the shaft and the bearing is filled
with a lubricating oil of viscosity 5 poise. If the length of the bearing is 200 mm,
find the power absorbed in the lubricating oil.

26
Figure 1.21: U-tube manometer

1.5 Hydrostatic thrust on submerged surfaces


1.5.1 Pressure force (thrust)
Figure 1.22 shows a plane surface of arbitrary shape, wholly submerged in a liquid in
equilibrium. Let the liquid be of density ρ. The plane makes an angle θ with the free

Figure 1.22: Hydrostatic thrust on a submerged plane

27
surface (horizontal). Every element of the area of the plane, e.g. δA, is subjected to a
force due to the pressure of the liquid. At any element of area δA, at a depth h below
the free surface, the gauge pressure is p = ρgh and the corresponding force is

δF = pδA = ρghδA = ρgysinθδA. (1.45)

Total force on the upper side of the plane is therefore


Z Z
F = δF = ρgysinθδA, (1.46)
A Z
= ρgsinθ yδA, (1.47)
A
R
but A
yδA is the first moment of area of the plane surface about the x -axis and
may be represented by Aȳ (A is the total area of the surface in consideration).
Therefore,

F = ρgsinθAȳ = ρg h̄A, (1.48)

but ρg h̄ is the pressure at the centroid of the plane surface and so, whatever the slope
of the plane, the total force (or thrust) exerted on it by the static fluid is given by the
product of area of the surface and the pressure at the centroid of the surface.

1.5.2 Centre of pressure for the plane surface


Besides knowing the magnitude of the total force F, we need to know the position of its
line of action. The total force is perpendicular to the plane and the point at which its
line of action meets the plane is known as the centre of pressure (or centre of thrust).
We shall determine this point by use of the principle of moments about the axis O - x,
in three steps:

1. From Eq. (1.45), force on an element of area δA is, δF = ρgysinθδA. Moment


of this force about O − x, δM , is

δM = δF (y)
= ρgy 2 sinθδA (1.49)

2. Let the centre of pressure be at (x’,y’ ) as shown in Fig. 1.22. The moment of
the resultant force F about any axis must be the same as the sum of the moments
of the individual forces (of each element) about the same axis. The moment of
the resultant force F about O − x, δM , is given by

M = F y0 (1.50)

28
3. Equating the moment due to the resultant force to the sum of the moments due
to elemental forces gives,
Z
0
Fy = ρgy 2 sinθ δA, (1.51)
A

but from Eq. (1.48), F = ρg h̄A = ρgsinθAȳ. Therefore,


Z Z
0
(ρgsinθAȳ)y = ρgy sinθδA = ρgsinθ y 2 δA,
2
(1.52)
A R R 2 A

0 ρgsinθ A y 2 δA y δA
y = = A , (1.53)
ρgsinθAȳ Aȳ
R
but A y 2 δA is the second moment of area of the immersed surface about
O − x, and is equal to I0 . Therefore
I0
y0 = . (1.54)
Aȳ

From the parallel axes theorem, I0 = IG + Aȳ 2 , where IG is the second moment
of area of the immersed surface about an axis through its centroid G, parallel to
the axis O − x, while ȳ is the perpendicular distance between the two axes.
Equation (1.54) then becomes

IG + Aȳ 2 IG
y0 = = ȳ + . (1.55)
Aȳ Aȳ
From this, it can be seen that the centre of pressure will always be below the
centroid G except when the surface is horizontal (i.e., for θ = 00 in Fig. 1.22).

In terms of vertical depth:


h0 h̄
Since y 0 = and ȳ = , then Eq. (1.55) becomes
sinθ sinθ
h0 h̄ IG sinθ
= + . (1.56)
sinθ sinθ Ah̄
Therefore,

IG sin2 θ
h0 = h̄ + . (1.57)
Ah̄
This equation is especially useful when θ = 900 , i.e., when the plane is submerged
IG
vertically, whereby h0 = h̄ + .
Ah̄
EXAMPLE 6
A circular plate 3.0 m in diameter is immersed in water is such a way that its greatest
and least depths from the free surface are 4 m and 1.5 m respectively. Determine the

29
total pressure force on the upper face of the plate and also the position of the centre
πd4
of pressure measured vertically from the free surface. For a circle, IG =
64
SOLUTION 6
d = 3.0 m; h1 = 1.5 m; h2 = 4 m; ρ = 1000 kg/m3 ; G = centroid; C = centre of
pressure; F = total thrust

Figure 1.23: Submerged circular surface

F = ρgAh̄,
π π
A = d2 = × 3.02 = 7.069 m2 ,
4 4
h̄ = h1 + r sin θ,
h2 − h1 4 − 1.5
sin θ = = = 0.833,
d 3.0
∴ h̄ = 1.5 + 1.5 × 0.833
= 2.75 m,
F = 1000 × 9.81 × 7.069 × 2.75
= 190.67 kN.

30
IG sin2 θ
h0 = h̄ + ,
Ah̄
πd4 π × 3.04
IG = = = 3.976 m4 ,
64 64
3.976 × 0.8332
h0 = 2.75 +
7.069 × 2.75
= 2.89 m.

1.6 Fluids in relative equilibrium and under con-


stant acceleration
1.6.1 Fluids subjected to horizontal acceleration
If a container containing a liquid and initially at rest is made to move with a constant
acceleration a in a horizontal direction, the liquid particles will initially move relative
to each other. After some time, there will be no more relative motion between the liquid
particles themselves and between the particles and the boundaries of the container, i.e.,
the liquid will come to rest in a new position relative to the container. Since the liquid
will be static relative to the container, the laws of hydrostatics can be applied. The
initial horizontal free surface of Fig. 1.24(a) will become a downward sloping inclined
plane as in Fig. 1.24(b).

Figure 1.24: Fluid mass subjected to horizontal acceleration

1. Equation of the slope: To determine the equation for the slope, we consider
the equilibrium of a fluid element A lying on the free surface as in Fig. 1.24(b).
The forces acting on the fluid element A are:

(a) Pressure force P exerted by the surrounding fluid. This force acts normal
to the free surface.
(b) Weight of the fluid element, i.e., W = m × g acting vertically downward.
(c) Accelerating force, i.e., F = m × a acting in the horizontal direction.

31
Figure 1.25: Forces on fluid element A

Resolving the forces horizontally,

P sinθ + m × a = 0,
∴ P sinθ = −ma. (1.58)

Resolving the forces vertically,

P cosθ − mg = 0,
∴ P cosθ = mg. (1.59)

Dividing Eq. (1.58) by (1.59), we get,


a
tan θ = − . (1.60)
g
The negative sign shows that the free surface is sloping downward in the direction
of acceleration.

2. Maximum and minimum pressure intensities at bottom of tank: The


pressure intensity will be a maximum where the depth of the liquid is a maximum,
i.e., if the density of the liquid is ρ,

pmax = ρgh1 . (1.61)

The minimum pressure intensity on the bottom of the tank will occur where the
depth is a minimum, i.e.,

pmin = ρgh2 . (1.62)

32
3. Forces on the ends of the tank: The total force F1 acting on the left end of
the tank (Fig. 1.24(c)) is given by,

F1 = ρg h̄1 A1 , (1.63)

h1
where, A1 = h1 × width of tank, while h̄1 = .
2
The total force F2 acting on the right side of the tank (Fig. 1.24(c)) is given by

F2 = ρg h̄2 A2 , (1.64)

h2
where A2 = h2 × width of tank, while h̄2 = .
2

1.6.2 Fluids subjected to vertical acceleration


Consider a tank containing a liquid, initially at rest, that is then moved vertically
upward at a constant acceleration a as in Fig. 1.26.
To obtain the expression for the pressure at any point in the liquid mass, consider a

Figure 1.26: Fluid mass subjected to vertical acceleration

vertical prism of liquid of cross-section area dA and height h.


The forces acting on the prism are:

1. Pressure force p0 dA acting vertically downward on the top face.

2. Pressure force pdA acting vertically upward on the bottom face.

3. Weight of prism W = ρ × dA × h × g acting vertically downward through the


prism’s centre of gravity.

From Newton’s second law of motion, the net force in a given direction must equal the

33
mass multiplied by the acceleration in the same direction. Therefore,

pdA − p0 dA − ρgdAh = (ρdAh)a,


p − p0 = ρgh + ρha
 
a
= ρgh 1 + . (1.65)
g

p is the absolute pressure while p0 is the atmospheric pressure, hence, p − p0 is the


gauge pressure.
Using a similar argument, it may be shown that if the tank were accelerating vertically
downward, then the gauge pressure at any point would be given by,
 
a
p − p0 = ρgh 1 − . (1.66)
g

Note:

• If the tank is moving downward with a constant acceleration equal to g (i.e.,


a = g), then p − p0 = 0, i.e., the pressure at any point in the liquid would be
equal to the surrounding atmospheric pressure.

• If the tank is stationary (i.e., a = 0), then the gauge pressure at any point in the
liquid is simply equal to ρgh, where h is the depth from the free surface.

Tutorial Problems 4

1. Determine the total pressure force and depth of centre of pressure on a vertical
plane rectangular plate 1 m wide and 3 m depth when its upper edge

(a) coincides with the free water surface [44.145 kN; 2 m]


(b) is 2 m below the free water surface [103 kN; 3.714 m]
bd3
For a rectangle of width b and depth d, IG = .
12
2. A circular opening 3 m in diameter, in a vertical side of a tank, is closed by a
disc of 3 m diameter, which can rotate about a horizontal diameter. If the head
of water above the horizontal diameter is 6 m, calculate

(a) the force on the disc [416.05 kN]


(b) the depth of the centre of pressure, measured from the horizontal diameter
[9.37 cm]

34
3. A rectangular tank is moving horizontally in the direction of its length with a
constant acceleration of 2.4 m/s2 . The length, width and depth of the tank are
6 m, 2.5 m and 2 m respectively. If the depth of water in the tank is 1 m and
the tank is open at the top, calculate,

(a) the angle of the water surface to the horizontal, [13.750 ]


(b) the maximum and minimum pressure intensities at the bottom, [pmax =
17010 N/m2 , pm in = 2610 N/m2 ]
(c) the total force due to water acting on each end of the tank. [F1 = 36,870.35
N, F2 = 867.65 N]

4. A tank containing water upto a level of 500 mm is moving vertically upward at


a constant acceleration of 2.45 m/s2 . Find the force exerted by the water on the
side of the tank if the width of the tank is 2 m. [3,065 N]

1.7 Buoyancy
1.7.1 Concept
A fluid will exert a resultant upward force on any body which is wholly or partly
immersed in it. This force is known as the buoyancy, and always acts vertically upwards.
A body, whether partly or wholly submerged in a fluid, will experience an upthrust
(buoyant force, FB ) equal to the weight of fluid it displaces. This is Archimedes prin-
ciple. The buoyancy acts through the centroid of the displaced volume of fluid, i.e.,
it acts through the centre of gravity of the displaced volume of fluid. This centroid is
referred to as the centre of buoyancy. The buoyant force is expressed,

FB = ρgV, (1.67)

where ρ is the density of fluid, while V is the volume of fluid displaced.

1.7.2 The hydrometer


Precise measurements of the specific weight of liquids is done by utilizing the principle
of buoyancy and the Archimedes’ principle.
The device used for this is the hydrometer. It is essentially a glass bulb that is
weighted on one end to make it float in a vertical position, and has a graduated stem
of constant diameter extending from the glass bulb. The level at which the hydrometer
floats depends on the density of liquid and it is this level that is used to indicate
the density (or specific gravity) of the liquid. Commercial hydrometers are usually
calibrated for ordinary room temperature, which is taken to be 200 C. Because of the

35
Figure 1.27: Hydrometer

variation in the depth to which the instrument sinks in heavy and light liquids, one
type is made for use in liquids more dense than water and another for use in liquids
less dense than water. A typical hydrometer is illustrated in Fig. 1.27.
EXAMPLES 7

1. What percentage of the total volume of iceberg of density 912 kg/m3 will extend
above the surface of sea water of density 1025 kg/m3 ?

2. Find the density of a solid body that floats on the interface of mercury of specific
gravity 13.6, and water. 40% of the body’s volume is submerged in the mercury
while 60% is in water.

SOLUTIONS 7

1. Let: FB be the buoyancy on the iceberg, Vi be the total volume of the iceberg,
Vs be the volume of sea water displaced by the iceberg, Vo be the volume of the
iceberg outside the water, W be the total weight of the iceberg, ρi be the density
of the iceberg and ρs be the density of the sea water.

FB = W,
ρs Vs g = ρi Vi g,
∴ ρs Vs = ρi Vi ,
Vs ρi 912
= = = 0.89,
Vi ρs 1025
Vo Vs
=1− = 0.11, i.e., 11%.
Vi Vi

36
Figure 1.28: Iceberg in sea water

2. Let:
Density of water, ρw = 1000 kg/m3
Density of mercury, ρm = 13.6 x 1000 kg/m3 = 13,600 kg/m3
Volume of body in water = Vw
Volume of body in mercury = Vm
Total volume of body, V = Vw + Vm
Buoyant force on body due to the water = FBw , acting through the centroid of
Vw
Buoyancy on body due to the mercury = FBm , acting through the centroid of
Vm
Density of body = ρ, while its weight = W
The total buoyant force on the body balances its weight, i.e.,

W = FBw + FBm ,
ρV g = ρw Vw g + ρm Vm g,
ρV = ρw Vw + ρm Vm .

Therefore,
Vw Vm
ρ = ρw + ρm
V V
= 1000(0.6) + 13600(0.4)
= 6, 040 kg/m3 .

37
Figure 1.29: Floating body

Tutorial Problems 5

1. A wooden block of width 2 m, depth 1.5 m and length 4 m floats upright in water.
Find the volume of water displaced and position of centre of buoyancy measured
from the top of the block. Specific gravity of the wooden block is 0.7. [8.4 m3 ;
0.975 m]

2. A piece of wood of specific gravity 0.651 is 80 mm square and 1.5 m deep. What
weight of lead of specific weight 110 kN/m3 must be fastened centrally at the
bottom end of the wood so that it will float upright with 0.3 m out of water?
[15.4 N]

3. Consider the hydrometer shown in Fig. 1.30.

(a) If the hydrometer sinks z = 5.3 cm in water at room temperature, the bulb
displaces 1.0 cm3 , and the stem area is 0.1 cm2 , find the weight of the
hydrometer. [1.5 x 10−2 N]
(b) If the hydrometer is placed in a brine solution and sinks such that z = 3.3
cm, what is the specific gravity of the brine? [1.149]

1.8 Stability of immersed and floating bodies


1.8.1 Immersed body
Here, we invariably refer to a body whose whole volume is inside the fluid of interest.
The stability of such a body depends on the relative positions of its centre of gravity
and the centre of buoyancy (the centroid of the volume of fluid it displaces).

38
Figure 1.30: Hydrometer

• If the centre of gravity is below the centre of buoyancy, any tipping of the body
produces a righting moment and hence the body is stable.

• If the centre of gravity is above the centre of buoyancy, any tipping produces an
overturning moment, hence the body is unstable.

• If the centre of gravity and centre of buoyancy coincide, the body is neutrally
stable, i.e., it has neither a righting nor an overturning tendency.

1.8.2 Floating bodies


For floating bodies, the relative positions of the centre of gravity and centre of buoyancy
may vary depending on the shape of the body and the position in which it floats.
Consider the cross-section of a ship floating in water as in Fig. 1.31. Since the centre
of gravity G is above the centre of buoyancy C, it gives the impression of instability
hence a tendency to flip over. However, after the ship has taken a small angle of heel
in the clockwise direction, G is in the same position but C has moved outward to C1 ,
resulting in a righting moment, while the waterline shifts from O - O to O1 − O1 (Fig.
1.32).
Let us focus on this change further with reference to Fig. 1.32. The change of
position C is because part of the original buoyant volume (wedge ABA1 ), is transferred
to the right to a new buoyant volume (wedge DBD1 ). Clearly then, since C is at the
centroid of the original displaced volume, then it must also shift towards the right. The
ship is hence stable.

39
Figure 1.31: Stability of a ship

Figure 1.32: Metacentre and metacentric height

The point of intersection of the lines of action of the buoyant force before and after
the heel is called the metacentre, M, and the distance GM is called the metacentric
height. It can be shown that the metacentric height is determined thus,
I00
GM = − CG, (1.68)
V
where I00 is the second moment of area of the body defined at the waterline, V is
the total displaced volume of fluid, while CG is the distance between the centre of
buoyancy (C) and the centre of gravity (G). Equation (1.68) is used to determine the
stability of floating bodies.

• If GM is positive (M above G), the floating body is stable.

• If GM is negative (M below G), the floating body is unstable.

40
There are in fact two metacentric heights for a ship: One for rolling and the other for
pitching. The former will always be less than the latter and unless otherwise stated,
the metacentric height given will be for rolling.

1.8.3 Determination of metacentric height


Let w1 be a known weight placed centrally over a floating vessel as shown in Fig.
1.33(a). The weight w1 is moved laterally across the vessel a distance x so that the
vessel tilts through an angle of heel θ as shown in Fig. 1.33(b). The centre of gravity
of the combined weight of vessel and load w1 moves from G to G1 while the centre of
buoyancy moves from C to C1 . Under equilibrium, the moment caused by the moving
load w1 through the distance x equals the moment caused by the shift of centre of
gravity from G to G1 . Therefore, if the combined weight of vessel and load w1 is W,
then, taking moments about M gives w1 cosθ · x = W sinθ · GM ,

w1 x = W (GM tanθ),
w1 x
∴ GM = . (1.69)
W tanθ
The angle θ can be measured using a plumbline and a protractor attached to the vessel.

Figure 1.33: Determining the metacentric height

Tutorial Problems 6
1. A block of wood 30 cm square in cross section and 60 cm long weighs 318 N. If
the block is placed in water with its length horizontal and the short sides vertical,
investigate its stability about its longitudinal axis. For a rectangle of length l and
lb3
breadth b, IOO = . [GM = -1.833 cm, unstable]
12
41
2. A solid cylinder of 15 cm diameter and 60 cm long, consists of two parts made
of different materials. The bottom part is 1.2 cm long and of specific gravity 5.0,
while the top part is made of material having specific gravity 0.6. Investigate
whether it can float vertically in water. [GM = -5.26 cm, unstable]

3. A body has a cylindrical upper portion of 4 m diameter and 2 m deep. The lower
portion is a curved one, which displaces a volume of 0.9 m3 of water. The centre
of buoyancy of the curved portion is at a distance of 2.1 m below the top of the
cylinder. The centre of gravity of the whole body is 1.5 m below the top of the
cylinder. The total displacement of water is 4.5 tonnes. Find the metacentric
height of the body. [2.387 m]

42
2.0 FLUID DYNAMICS
Fluid dynamics deals with fluids in motion, i.e., fluid flow, which is made up of a
large number of individual fluid particles moving in the general direction of flow. The
prediction of the conditions encountered by, and as a result of fluids in motion presents
a range of problems that must be resolved by reference to the fundamental laws of
physics, coupled with fluid properties.
There is no shear force in a fluid at rest but when in motion, shear forces can
be set up due to viscosity and turbulence which oppose motion, producing frictional
effects. Many problems can be solved at least partially, by assuming an ideal frictionless
(inviscid) fluid but where viscosity plays a major part, this assumption is untenable
and simplification is often obtained by assuming the flow does not change with time
(steady flow is assumed).
In this chapter, the fundamental equations used in the analysis of fluid motion will
be developed. Typical assumptions that are normally invoked to simplify analysis, e.g.,
inviscid flow, will also be discussed.

2.1 Lagrangian and Eulerian viewpoints


Equations for fluids in motion can be expressed in either of two ways: Lagrangian or
Eulerian.

2.1.1 Lagrangian approach


This approach considers an individual fluid particle for all time. For that reason, the
position of a specific particle is tracked as time passes. The velocity of the particle is
obtained by differentiating the particle’s position vector with respect to time. Assuming
a Cartesian co-ordinate system, the position vector may be expressed,

r(t) = xi + yj + zk, (2.1)

where i, j and k are the unit vectors in the x, y and z directions respectively. The
velocity of the fluid particle is then obtained through differentiation,
dx dy dz
v(t) = i+ j+ k
dt dt dt
= ui + vj + wk. (2.2)

To describe the entire flow field, the motion of all fluid particles must be considered
simultaneously. This is an enormous task particularly because the relative positions of
the particles continuously change with time. For that reason, the Eulerian approach is

43
generally preferred, unless one is specifically concerned with the motion of individual
particles, e.g., in rarefied gas dynamics [3].

2.1.2 Eulerian approach


In this approach, we observe the motion of particles passing a specific point in the flow
field, i.e., the focus is on a certain point in space and the motion of all fluid particles
that pass that point as time passes is considered. The fluid particle velocity is then a
function of space and time. This can be expressed in Cartesian co-ordinates thus,

u = f1 (x, y, z, t), (2.3)


v = f2 (x, y, z, t), (2.4)
w = f3 (x, y, z, t), (2.5)

where u is the x-component of velocity, etc. We can further work out the acceleration
of such a particle by differentiating the expression for velocity with respect to time and
space. For example, the acceleration of a fluid particle in the x-direction is given by
du
ax = . (2.6)
dt
We can then use the chain rule for differentiation,
∂u dx ∂u dy ∂u dz ∂u
ax = + + + , (2.7)
∂x dt ∂y dt ∂z dt ∂t
where,
dx
= u, (2.8)
dt
dy
= v, (2.9)
dt
dz
= w. (2.10)
dt
Therefore,
∂u ∂u ∂u ∂u
ax = u +v +w + . (2.11)
∂x ∂y ∂z ∂t
|{z}
| {z }
convective accelerations local acceleration

Notes:

• Convective accelerations are associated with velocity changes that occur because
of changes in position in the flow field. They occur when the flow is non-uniform.

• Local accelerations are the ones due to velocity changes with respect to time at a
given point. For that reason, local acceleration results when the flow is unsteady.

44
Unsteady and non-uniform flows will be described in subsequent sections.
Equation (2.11) may be written in contracted form as,

Du
ax = , (2.12)
Dt
D ∂ ∂ ∂ ∂ D
where = +u +v + w . The operator is called the substantive or
Dt ∂t ∂x ∂y ∂z Dt
material derivative.
Equations (2.11) and (2.12) can be written for the y- and z-components of accel-
eration as well (students to do).

2.1.3 Pathline, streamline and streamtube


A fluid consists of a large number of individual particles moving in the general direction
of flow but usually not parallel to each other. The velocity of any particle is a vector
quantity having magnitude and direction which vary from moment to moment [5, 8].

Figure 2.1: Streamline

The path followed by a particle is called a pathline. At any given instant of time,
the positions of successive particles can be joined up by a curve, which is tangential
to the direction of motion (and velocity) of the particle at that instant [2, 4]. This
curve is called a streamline (Fig. 2.1). Since the velocity of a particle at any point
on a streamline is tangential to it, there can be no flow across a streamline. If a series
of streamlines are drawn through every point on the perimeter of a small area of the
stream cross-section, they will form a streamtube (Fig. 2.2).
Since there is no flow across a streamline, the fluid inside the streamtube cannot
escape through its walls and behaves as if it were contained in an imaginary pipe. This
is a useful concept in dealing with the flow of a large body of fluid, since it allows
elements of the fluid to be isolated for analysis.

45
Figure 2.2: Streamtube

2.1.4 Real and ideal fluids


When a real fluid flows past a boundary, the fluid immediately in contact with the
boundary will have the same velocity as the boundary. When the boundary is sta-
tionary, the fluid in contact with the boundary will also be stationary. The velocity of
successive layers of the fluid will increase as we move away from the boundary until a
point will be reached beyond which the velocity will approximate to the free stream
velocity, and the drag exerted by the boundary will have no effect [5, 8].
The part of the flow adjoining the boundary in which change of velocity occurs
is known as the boundary layer. In this region, shear stresses are developed between
the layers of fluid moving with different velocities. The shear forces are developed as
a result of viscosity and the interchange of momentum due to turbulence. Outside the
boundary layer, in a real fluid, the effect of the shear stresses due to the boundary can
be ignored and the fluid can be treated as if it were an ideal fluid. But even in cases
where the effects of viscosity and turbulence cannot be neglected, it is often convenient
to carry out the mathematical analysis assuming an ideal fluid and correction factors
can be employed to bring the results into agreement with the behaviour of a real fluid.
In general, all fluid flow occurs in three dimensions, so that the velocity, pressure
and other flow parameters vary with reference to three orthogonal axes. In some
problems however, the major changes occur in two directions or even in only one
direction.
If the flow parameters, for example, pressure and velocity, describing the flow at
any given instant, vary only along the direction of flow, the flow is described as one

46
dimensional.

2.1.5 Steady and unsteady flow


Steady flow is one in which the flow parameters such as velocity, pressure and cross-
section of the stream do not change with time. They are the same at all instants of
time.
Flow in which the flow parameters change with time is called unsteady flow.

Figure 2.3: Velocity profiles for one dimensional flow

2.1.6 Uniform and non-uniform flow


Flow is described as uniform at a particular instant if the flow parameters, e.g., velocity
and cross-section of the stream remain constant at every point along the flow over a
specified region.
If the flow parameters change from point to point, the flow is said to be non-uniform.

2.1.7 Laminar flow and turbulent flow: Reynolds experiment


Osborne Reynolds performed an experiment in 1883 in which water was discharged
from a tank through a glass tube. The rate of flow could be controlled by a valve at
the outlet, and a fine filament of dye injected at the entrance to the tube. The way in
which the water was flowing along the glass tube could thus be studied by observing
the behaviour of the stream of dye.
2.1.7.1 Laminar flow
At low velocities of water flow, it was found that the dye filament remained intact
throughout the length of the glass tube, showing that particles of water moved smoothly
in parallel lines. This type of flow is known as laminar, viscous or streamline flow. The
particles of the fluid move in an orderly manner and retain the same relative positions
in successive cross-sections.

47
Figure 2.4: Reynolds apparatus

2.1.7.2 Turbulent flow


As the velocity of water in the tube was increased by opening the outlet further, the
dye filament began to waver and break up so that the colour was diffused through the
water. Further increase in the velocity of the water caused the dye to mix more or
less completely with the water, showing that the particles of fluid no longer moved in
an orderly manner but occupied different relative positions in successive cross-sections.
This type of flow is known as turbulent flow.
The type of fluid flow is distinguished by means of the Reynolds number, Re. In
pipe flow, Re is defined,
ρvD
Re = , (2.13)
µ
where, v = mean flow velocity, D = diameter of pipe, ρ = fluid density and µ =
fluid viscosity. If Re > 3000 in pipe flow, the flow is generally turbulent and if Re <
2000, the flow is laminar. The flow in between the two values is transitional. Re is
dimensionless and in terms of forces, it is a ratio of the inertia force to the viscous
force in a flow.

2.2 Continuity of flow


Matter is neither created nor destroyed. This is the principle of conservation of mass
and is the basis for the derivation of the continuity equation for fluid flow.

2.2.1 Continuity equation for a general case


Based on mass conservation, the mass rate of fluid flow out of a region of space (such
as a control volume) minus the mass rate of fluid flow into the region, is equal to the
rate of change of mass within the region. The continuity equation can be derived for

48
Cartesian co-ordinates as follows:
Consider a fixed Cartesian control volume with sides ∆x, ∆y and ∆z (Fig. 2.5). The
total rate of change of mass with time for the control volume comprises:

• The rate of change of the total mass inside the control volume.

• The difference between the rates at which mass leaves and enters the control
volume (i.e., net mass flux).

Thus,
d
(mass) + net outward mass flux = 0. (2.14)
dt
If the volume of the control volume is V and a typical face has area A,

mass of fluid in control volume = density x volume = ρV, (2.15)


mass flux = ρuN A. (2.16)

Equation (2.14) may then be written using Eqs. (2.15) and (2.16) as,

Figure 2.5: Cartesian control volume for the continuity equation

d X
(ρV ) + ρuN A = 0. (2.17)
dt faces

49
Note that uN is the velocity component normal to the face in consideration, A. Based
on Eq.2.17 and with reference to Fig. 2.5, we may write,
d
(ρV ) + (ρuA)e − (ρuA)w + (ρvA)n − (ρvA)s + (ρwA)t − (ρwA)b = 0.
|dt {z } | {z }
net mass flux OUT of the control volume through the faces
rate of change of mass

(2.18)

Substituting the dimensions of the control volume,


d
(ρ∆x∆y∆z) + [(ρu)e − (ρu)w ]∆y∆z
dt
+ [(ρv)n − (ρv)s ]∆x∆z
+ [(ρw)t − (ρw)b ]∆x∆y = 0. (2.19)

Dividing through by the volume ∆x∆y∆z,

∂ρ (ρu)e − (ρu)w
+
∂t ∆x
(ρv)n − (ρv)s
+
∆y
(ρw)t − (ρw)b
+ = 0. (2.20)
∆z
Proceeding to the limit as ∆x, ∆y and ∆z → 0,

∂ρ ∂(ρu) ∂(ρv) ∂(ρw)


+ + + = 0. (2.21)
∂t ∂x ∂y ∂z

2.2.2 Continuity equation for steady, one-dimensional flow


The principle of conservation of mass can be applied to steady flow in a streamtube
having a cross sectional area, δAi , that is small enough for the velocity to be considered
constant across any given cross section at any point i along the streamtube. Consider
the streamtube in Fig. 2.6. Since there can be no flow across a stream tube, mass
entering per unit time at section 1 = mass leaving per unit time at section 2.
At section 1:

• δA1 is the cross section area of the stream tube.

• u1 is the velocity of the fluid.

• ρ1 is the fluid density.

δA2 , u2 and ρ2 are the corresponding values at section 2. Therefore,


mass entering per unit time at 1 = ρ1 δA1 u1 ,

50
Figure 2.6: Continuous flow through a streamtube

mass leaving per unit time at 2 = ρ2 δA2 u2 .


Thus, for steady flow,

ρ1 ∂A1 u1 = ρ2 ∂A2 u2 = Constant. (2.22)

This equation is the equation of continuity for flow of a fluid through a streamtube,
where u1 and u2 are the velocities measured at right angles to the cross-sectional areas
δA1 and δA2 respectively.
For flow of a real fluid through a pipe, the velocity varies from wall to wall (refer
to Fig. 2.3) but using the mean velocity ū, the continuity equation for steady flow
becomes

ρ1 A1 ū1 = ρ2 A2 ū2 = ṁ, (2.23)

(A1 and A2 are the cross-sectional areas of flow at section 1 and 2 respectively while
ṁ is the mass flow rate of the fluid with units of kg/s). If the fluid is incompressible,
ρ1 = ρ2 , and the continuity equation reduces to

A1 ū1 = A2 ū2 = Q, (2.24)

(Q is the volume flow rate with units m3 /s).


Note that Eq. (2.23) can also be recovered from the general expression of Eq. (2.18) in
the following manner:

51
• The flow is steady, therefore
d
(ρV ) = 0. (2.25)
dt

• The flow is one-dimensional, therefore, we consider fluxes in only one direction,


say x-direction. This gives,

(ρuA)e − (ρuA)w = 0. (2.26)


∴ (ρuA)e = (ρuA)w = constant. (2.27)

Consider a branched pipe:


For steady conditions:

Figure 2.7: A branched pipe

ṁ1 = ṁ2 + ṁ3 ,


∴ ρ1 A1 ū1 = ρ2 A2 ū2 + ρ3 A3 ū3 .

For an incompressible fluid, ρ1 = ρ2 = ρ3 such that A1 ū1 = A2 ū2 + A3 ū3 , meaning,


Q1 = Q2 + Q3 .
EXAMPLE 8
Water flows through a pipe AB of diameter d1 = 50 mm, which is in series with a pipe
BC of diameter d2 = 75 mm in which the mean velocity ū2 = 2 m/s. At C, the pipe
forks and one branch CD is of diameter d3 and mean velocity ū3 = 1.5 m/s. The other
branch CE is of diameter d4 = 30 mm and conditions are such that the discharge Q2
Q3
from BC divides so that Q4 = . Calculate the values of Q1 , ū1 , Q3 , d3 , Q4 and ū4 .
2
SOLUTION 8

52
Figure 2.8: Pipe network

Given:
d1 = 0.05 m d2 = 0.075 m d3 =? d4 = 0.03 m
ū1 =? ū2 = 2 m/s ū3 = 1.5 m/s ū4 =?
Q3
Q1 =? Q2 = Q1 Q3 =? Q4 =
2
• ū1 and Q1 :

A1 ū1 = A2 ū2 ,
∴ d21 ū1 = d22 ū2 ,
 2
d2
ū1 = ū2
d1
 2
0.075
= × 2 = 4.5 m/s,
0.05
Q1 = A1 ū1
π
= d21 ū1
4
π
= × 0.052 × 4.5 = 8.836 × 10−3 m3 /s
4

53
• d3 and Q3 :
Q3 3Q3
Q2 = Q1 (pipes in series ) = Q3 + Q4 = Q3 + = ,
2 2
2Q2
∴ Q3 =
3
2 × 8.836 × 10−3
= = 5.89 × 10−3 m3 /s,
3
π
Q3 = A3 ū3 = d23 ū3 ,
s 4

4Q3
∴ d3 =
πū3
s
4 × 5.89 × 10−3

= = 0.0707 m.
π × 1.5

• ū4 and Q4 :
Q3
Q4 =
2
5.89 × 10−3
= = 2.945 × 10−3 m3 /s,
2
π
Q4 = A4 ū4 = d24 ū4 ,
4
4Q4
ū4 =
πd24
4 × 2.945 × 10−3
= = 4.166 m/s.
π × 0.032

Tutorial Problems 7

1. A 40 cm diameter pipe containing water branches into two pipes of diameter 30


cm and 20 cm respectively. If the mean velocity in the 40 cm-diameter pipe is 3
m/s, find the discharge in the pipe. [0.377 m3 /s]
Also determine the velocity in the 20 cm-diameter pipe, if the mean velocity in
the 30 cm-diameter pipe is 2 m/s. [7.5 m/s]

2. When 1800 litres/min flow through a 0.3 m-diameter pipe which later reduces to
0.15 m-diameter, calculate the average velocities in the two pipes.

3. A fluid of constant density flows at the rate of 15 litres/s along a pipe AB of


100 mm diameter. This pipe branches at B into three pipes BC, BD and BE.
BC and BD are each of 25 mm diameter while BE is of 50 mm diameter. The
flowrates are such that, that through BC is three times that through BE and
the velocity through BD is 4 m/s.

54
Find the flow rates in the three branches BC, BD and BE and the velocities in
the pipes AB, BC and BE.
[9.777 x 10−3 m3 /s; 1.063 x 10−3 m3 /s; 3.26 x 10−3 m3 /s; 1.91 m/s; 19.92 m/s;
1.66 m/s]

2.3 Navier-Stokes equations


The Navier-Stokes equations are a set of partial differential equations which describe
the motion of a fluid. These equations establish that changes in momentum (accelera-
tion) of the particles of a fluid are a function of the changes in pressure and dissipative
viscous forces (friction) acting inside the fluid, i.e., they focus on the rates of change or
fluxes of these quantities. In mathematical terms these rates correspond to their deriva-
tives. There are cases where body forces, e.g., gravitational force, become significant.
Special cases of Navier-Stokes equations include the Poiseuille’s law and Bernoulli’s
equation. In a nutshell then, the Navier-Stokes equations are a dynamical statement
of the balance of forces acting at any given region of a fluid.
The Navier-Stokes equations are very useful in the study of a wide variety of fluid
flow phenomena of academic and economic interest, including: Modelling of weather,
study of ocean currents (climate), study of water flow in pipes, study of motion of
stars inside a galaxy, study of flow around a wing of an aircraft, design of aircraft and
cars, the study of blood flow, the design of power stations, the analysis of the effects
of pollution, and many more.
The equations apply to both laminar and turbulent flow. For complex applications,
the analysis becomes very complex and analytical solutions cannot be realized. This
necessitates the use of numeric computer models, in the context of Computational Fluid
Dynamics (CFD).

2.3.1 Newton’s law of viscosity in Cartesian co-ordinates (x,y,z)


The generalized Newton’s law of viscosity can be written as
    
∂vj ∂vi 2 ∂vx ∂vy ∂vz
τij = µ + + µ−κ + + δij , (2.28)
∂xi ∂xj 3 ∂x ∂y ∂z

which may also be written in vector-tensor notation as


2
τ = µ(∇v + (∇v)T ) + ( µ − κ)(∇ · v)δδ , (2.29)
3
in which δ is the unit tensor with components δij (note, δij = 1 for i=j and δij =
∂vj
0 for i 6= j), ∇v is the velocity gradient tensor with components , (∇v)T is the
∂xi

55
∂vi
transpose of the velocity gradient tensor with components and ∇·v is the divergence
∂xj
of the velocity vector [9].
Equation (2.29) is written here for Cartesian co-ordinates in three dimenstions
   
∂u 2
τxx = µ 2 + ( µ − κ)(∇ · v) , (2.30)
∂x 3
   
∂v 2
τyy = µ 2 + ( µ − κ)(∇ · v) , (2.31)
∂y 3
   
∂w 2
τzz = µ 2 + ( µ − κ)(∇ · v) , (2.32)
∂z 3
 
∂v ∂u
τxy = τyx = µ + , (2.33)
∂x ∂y
 
∂w ∂v
τyz = τzy = µ + , (2.34)
∂y ∂z
 
∂u ∂w
τzx = τxz = µ + . (2.35)
∂z ∂x
• If the fluid is a gas, it is often assumed to act as an ideal monoatomic gas for
which κ is identically zero [9].

• If the fluid is a liquid, it is often assumed incompressible for which the divergence
of the velocity vector is zero [9], i.e.,

∇ · v = 0, or (2.36)
∂vx ∂vy ∂vz
+ + = 0. (2.37)
∂x ∂y ∂z
Equations (2.30) - (2.35) are then written for an incompressible fluid as,
∂u
τxx = 2µ , (2.38)
∂x
∂v
τyy = 2µ , (2.39)
∂y
∂w
τzz = 2µ , (2.40)
∂z  
∂v ∂u
τxy = τyx = µ + , (2.41)
∂x ∂y
 
∂w ∂v
τyz = τzy = µ + , (2.42)
∂y ∂z
 
∂u ∂w
τzx = τxz = µ + , (2.43)
∂z ∂x
where Eqs. (2.38) - (2.40) represent normal stresses, while Eqs. (2.41) - (2.43) represent
the shear stresses. Consider a 2D case given in Fig. 2.9, which corresponds to the top

56
face ABCD of Fig. 2.10. In this 2D case, the dimensions are ∆x in the x-direction
and ∆y in the y-direction. It is assumed that the third dimension (normal to the page)
is unity.
2D CASE - IMPORTANT

Figure 2.9: Viscous stresses

For our 2D case of Fig. 2.9, the total viscous force in the x direction is a
function of:

1. The normal stresses pointing in the x-direction and acting on the x-faces,
i.e, τxx . The x-faces are the faces normal to the x-direction. Therefore, the net
force in the x-direction because of the normal stresses would then be expressed
as τxx,BC ABC − τxx,AD AAD = (τxx,BC − τxx,AD )∆y.

2. The shear stresses pointing in the x-direction and acting on the y-faces, i.e.,
τxy . The y-faces are the faces normal to the y-axis. Therefore, the net force in
the x-direction because of these shear stresses would be expressed as τxy,AB AAB −
τxy,DC ADC = (τxy,AB − τxy,DC )∆x.

2.3.2 The momentum equation


The momentum equation for a fluid is based on Newton’s second law, i.e., rate of
change of momentum = force. The momentum equation can be expressed in terms of
the shear stress, τ thus [9],

Dv
ρ = −∇p + ∇ · τ + ρg. (2.44)
Dt

57
Figure 2.10: Cartesian control volume for the momentum equation

This equation can be derived for Cartesian co-ordinates as follows:


Consider a fixed Cartesian control volume with sides ∆x, ∆y and ∆z (Fig. 2.10).
The total rate of change of momentum with time for the control volume comprises:

• The rate of change of the total momentum inside the control volume.

• The difference between the rates at which momentum leaves and enters the
control volume (i.e., net flux).

Thus,
d
(momentum) + net outward momentum flux = force. (2.45)
dt
The force on the right-hand side of Eq. (2.45) is made up of surface forces and body
forces.

• Surface forces act on the control volume faces and are proportional to the area,
e.g., pressure (p) and viscous stress (τ ).

• Body forces are a function of the volume, e.g., gravitational force (ρg) is the force
per unit volume.

If the volume of the control volume is V and a typical face has area A,

momentum in control volume = mass x velocity = ρV u, (2.46)


momentum flux = mass flux x velocity = (ρuN A)u. (2.47)

58
Equation (2.45) may then be written using Eqs. 2.46 and (2.47) as,
d X
(ρV u) + (ρuN A)u = F . (2.48)
dt faces
Note:
• u and F are vectors with 3 components each, i.e., u, v and w for velocity and Fx ,
Fy and Fz for the force, in the x-, y- and z-directions respectively.

• uN is the velocity component normal to the face in consideration, A.


To simplify our analysis, we shall consider only the x-component of the rate of change
of momentum, and thus the net force in the x-direction in Eq. (2.48).
3D CASE - IMPORTANT
For our 3D case of Fig. 2.10, the total viscous force in the x direction is a
function of:
1. The normal stresses pointing in the x-direction and acting on the x-faces,
i.e., τxx . The x-faces are the faces normal to the x-direction. Therefore, the net
force in the x-direction because of the normal stresses would then be expressed
as (τxx )e Ae − (τxx )w Aw .

2. The shear stresses pointing in the x-direction and acting on the y-faces, i.e.,
τxy . The y-faces are the faces normal to the y-axis. Therefore, the net force in
the x-direction because of these shear stresses would be expressed as (τxy )n An −
(τxy )s As .

3. The shear stresses pointing in the x-direction and acting on the z-faces, i.e.,
τxz . The z-faces are the faces normal to the z-axis. Therefore, the net force in
the x-direction because of these shear stresses would be expressed as (τxz )t At −
(τxz )b Ab .
Based on these relations and with reference to Eq. (2.48), we may then write for the
x-component,
d
(ρV u) +
|dt {z }
rate of change of momentum

(ρuA)e ue − (ρuA)w uw + (ρvA)n un − (ρvA)s us + (ρwA)t ut − (ρwA)b ub


| {z }
net momentum flux OUT of the control volume through the faces

= (pw Aw − pe Ae ) + ((τxx )e Ae − (τxx )w Aw ) + ((τxy )n An − (τxy )s As ) + ((τxz )t At − (τxz )b Ab )


| {z } | {z }
net pressure force net viscous force

+ (ρV g)x . (2.49)


| {z }
body force

59
Substituting the dimensions of the control volume,
d
(ρ∆x∆y∆z u) + [(ρu)e ue − (ρu)w uw ]∆y∆z
dt
+ [(ρv)n un − (ρv)s us ]∆x∆z
+ [(ρw)t ut − (ρw)b ub ]∆x∆y
= (pw − pe )∆y∆z
+ ((τxx )e − (τxx )w )∆y∆z
+ ((τxy )n − (τxy )s )∆x∆z
+ ((τxz )t − (τxz )b )∆x∆y
+ (ρg)x ∆x∆y∆z. (2.50)

Dividing through by the volume ∆x∆y∆z, and changing the order of pe and pw for
convenience,

∂(ρu) (ρuu)e − (ρuu)w


+
∂t ∆x
(ρvu)n − (ρvu)s
+
∆y
(ρwu)t − (ρwu)b
+
∆z
(pe − pw ) (τxx )e − (τxx )w (τxy )n − (τxy )s (τxz )t − (τxz )b
=− + + +
∆x ∆x ∆y ∆z
+ ρgx . (2.51)

Proceeding to the limit as ∆x, ∆y and ∆z → 0,

∂(ρu) ∂(ρuu) ∂ρvu ∂(ρwu) ∂p ∂τxx ∂τxy ∂τxz


+ + + =− + + + . (2.52)
∂t ∂x ∂y ∂z ∂x ∂x ∂y ∂z

We can then proceed to express the viscous stresses in Eq. (2.52) using constitutive
relations, i.e., Eqs. (2.38), (2.41) and (2.43) thus,
 
∂(ρu) ∂(ρuu) ∂(ρvu) ∂(ρwu) ∂p ∂ ∂u
+ + + =− + 2µ
∂t ∂x ∂y ∂z ∂x ∂x ∂x
  
∂ ∂v ∂u
+ µ +
∂y ∂x ∂y
  
∂ ∂u ∂w
+ µ +
∂z ∂z ∂x
+ ρgx , (2.53)

60
which can be written (students to show),
 2
∂ u ∂ 2u ∂ 2u
  
∂u ∂u ∂u ∂u ∂p
ρ +u +v +w =− +µ + +
∂t ∂x ∂y ∂z ∂x ∂x2 ∂y 2 ∂z 2
 
∂ ∂u ∂v ∂w
+µ + + + ρgx , (2.54)
∂x ∂x ∂y ∂z
or,
 
Du ∂p 2 ∂ ∂u ∂v ∂w
ρ =− + µ∇ u + µ + + + ρgx , (2.55)
Dt ∂x ∂x ∂x ∂y ∂z
D
where is the substantive derivative, while ∇2 is the Laplacian operator, which
Dt
∂2 ∂2 ∂2
stands for + + .
∂x2 ∂y 2 ∂z 2
For incompressible flow, the divergence is zero, i.e.,
∂u ∂v ∂w
+ + = 0, (2.56)
∂x ∂y ∂z
and the x-momentum equation reduces to,
Du ∂p
ρ =− + µ∇2 u + ρgx . (2.57)
Dt ∂x
The y- and z-momentum equations can be similarly derived by inspection (students
to do). The complete set of the three-dimensional equations is what forms the Navier-
Stokes equations, named after L. M. Navier (1785 - 1836) and G. G. Stokes (1819 -
1903), who are credited with their development [3].

2.4 Euler’s and Bernoulli’s equations


Bernoulli’s equation, which is one of the most useful equations in fluid mechanics, can
be obtained in several ways including:

1. Deriving Euler’s equation then integrating it.

2. Applying the principle of conservation of energy.

3. Obtaining Euler’s equation from Navier-Stokes equations, then integrating it


(Assignment 2).

2.4.1 Euler’s equation of motion along a streamline


If we consider the rate of change of momentum from point to point along a streamline,
and the forces acting due to the effects of the surrounding pressures and changes of
elevation, it is possible to derive a relationship between velocity, pressure, elevation
and density along a streamline.

61
Figure 2.11: Deriving Euler’s equation

Figure 2.11 shows a short cylindrical section of a streamline having a length δs


and a cross-sectional area, A that is small enough for the velocity to be considered
constant over the cross-section; and let us assume flow conditions to be steady. The
forces acting on the fluid in the direction of flow are:

1. Force due to pressure p acting in the direction of motion = pA.

2. Force due to pressure p + δp opposing motion = (p + δp)A.

3. Force due to the weight of the element, producing a component opposing motion
= mgcosθ.

Let θ be the angle between the direction along which the fluid flows, and the line of
action of the weight of the element.
Resultant force in the direction of motion, FR is given as,

FR = pA − (p + δp)A − mgcosθ,

but m = density x volume = ρAδs,

FR = pA − (p + δp)A − ρgAδs cosθ


= pA − pA − δpA − ρgAδs cosθ
= −δpA − ρgAδs cosθ,

62
δz
but cosθ = , therefore,
δs
FR = −Aδp − ρgAδz. (2.58)

But by Newton’s second law, the resultant force on the fluid element must equal the
product of mass and acceleration in the direction of flow, i.e.,

FR = m as = ρAδs as . (2.59)
du
The acceleration as is given by, as = where u is a function of both position and
dt
time. Therefore, applying partial differentiation with respect to position and time,
 
∂u ds ∂u
as = + , (2.60)
∂s dt ∂t
ds
= u (velocity component in s-direction in this case) and since the flow is steady,
dt
∂u du
then = 0. Therefore, as = u (notice now the full derivative since the velocity is
∂t ds
varying only with position).
 
du
FR = (ρAδs) u . (2.61)
ds
 
du
Equating (2.58) and (2.61), (ρAδs) u = −Aδp − ρgAδz; dividing through by Aδs,
ds
and taking the limit δs → 0,
du ∂p ρg∂z
ρu =− − or,
ds ∂s ∂s

ρas = − (p + ρgz), (2.62)
∂s
du
where as = u , is the fluid acceleration in the s-direction (direction of flow).
ds
Going back to the Navier-Stokes equations, we may then express the fluid acceleration
in the x-direction for steady flow from Eq.(2.11) as,
∂u ∂u ∂u
ax = u +v +w .
∂x ∂y ∂z
Equation (2.62) is known as Euler’s equation.

2.4.2 Bernoulli’s equation


Integrating Eq. (2.62) for an incompressible fluid (constant ρ), along the streamline
with respect to s gives [5],
p u2
+ + gz = constant, dividing through by g,
ρ 2
p u2
+ + z = constant. (2.63)
ρg 2g

63
Equation (2.63) is the Bernoulli’s equation for an incompressible frictionless fluid.
The terms represent energy per unit weight. In terms of energy per unit volume,
Eq. (2.63) is multiplied through by ρg such that,
1
p + ρu2 + ρgz = constant. (2.64)
2
EXAMPLE 9
The diameter of a pipe carrying water changes gradually from 150 mm at a point A, 6
m above the datum, to 75 mm at B, 3 m above the datum. The pressure at A is 103
kN/m2 while the velocity is 3.6 m/s. Neglecting any losses, determine the pressure at
B.
SOLUTION 9
Given:

Figure 2.12: Converging pipe

dA = 0.15 m dB = 0.075 m
zA = 6 m zB = 3 m
pA = 103 kN/m2 pB =?
uA = 3.6 m/s uB =?
From the continuity equation,

AA uA = AB uB ,
∴ d2A uA = d2B uB
 2
dA
uB = uA
dB
 2
0.15
= × 3.6 = 14.4 m/s.
0.075

64
From Bernoulli’s equation,
u2A pA u2 pB
+ + zA = B + + zB ,
2g ρg 2g ρg
ρ
∴ pB = pA + (u2A − u2B ) + ρg(zA − zB )
2
1000
= 103 × 103 + (3.62 − 14.42 ) + 1000 × 9.81 × (6 − 3) = 35.23 kPa.
2

2.4.3 Energy approach to Bernoulli’s equation


The one-dimensional energy equation for steady-in-the-mean flow may be written
2 2
     
p p Cout − Cin
ṁ uout − uin + − + + g(zout − zin ) = Q̇net − Ẇnet ,
ρ out ρ in 2
(2.65)

where C is the fluid flow velocity. This equation is valid for both compressible and
incompressible flows. The heat transfer rate, Q̇ is positive if heat is transferred into
the flow system and negative if the heat is transferred out of the flow system. The rate
of doing work, Ẇ (otherwise called power ) is positive if the flow system does work on
the surroundings, and negative if work is done on the flow system. Normally, the rate
of doing work associated with the fluid flow is called shaft power, Ẇshaf t , because a
rotating shaft is involved, e.g., in a pump. In that case, Ẇshaf t = Torque × ω.
If the flow is incompressible, ρout = ρin = ρ, therefore,
2 2
 
pout pin Cout − Cin
ṁ uout − uin + − + + g(zout − zin ) = Q̇net − Ẇnet , (2.66)
ρ ρ 2
and in terms of energy per unit mass, we divide through by ṁ,
2 2
 
pout pin Cout − Cin
uout − uin + − + + g(zout − zin ) = qnet − wnet . (2.67)
ρ ρ 2
Flows obeying Eq. 2.67 include flows through pumps, blowers, fans, turbines etc.
If the flow is steady throughout, one-dimensional and only one fluid is involved, the
shaft work is zero (note that fluid machines involve locally unsteady flow), therefore,
2 2
 
pout pin Cout − Cin
uout − uin + − + + g(zout − zin ) = qnet . (2.68)
ρ ρ 2
Re-organizing,
2 2
pout Cout pin Cin
+ + gzout = + + gzin − (uout − uin − qnet ). (2.69)
ρ 2 ρ 2
If the flow is frictionless, (uout − uin − qnet ) = 0 and Eq. (2.69) reduces to,
p C2
+ + gz = constant = useful or available energy. (2.70)
ρ 2

65
We may then consider the term (uout −uin −qnet ) as the loss of useful or available energy
that occurs in an incompressible fluid flow because of friction. With this knowledge,
we can derive the extended Bernoulli’s equation, otherwise called mechanical
energy equation, from Eq. (2.67) thus,
2 2
pout Cout pin Cin
+ + gzout = + + gzin − wshaft,net − loss. (2.71)
ρ 2 ρ 2

2.4.4 Application of Bernoulli’s equation


2.4.4.1 The Venturimeter
Pressure differences between any two points in a tapering pipe through which a fluid
is flowing depends on the difference of levels (z2 − z1 ), the mean velocities u2 and u1
and therefore, on the volume rate of flow, Q through the pipe [5, 8].

Figure 2.13: The venturimeter

Pressure difference can then be used to determine the volume rate of flow for any
particular configuration. This is the effect the venturimeter uses for the measurement
of flow in pipelines. The venturimeter consists of a short converging tube leading to
a short cylindrical portion, called throat (of a smaller diameter than the main pipe),
which is followed by a long diverging section in which the diameter increases again to
that of the main pipe.

66
The pressure difference from which the volume rate of flow can be determined is
measured between the entry section 1 and the throat section 2 by means of a U-tube
manometer. The axis of the venturimeter may be horizontal, or inclined at any angle.
For an incompressible flow, the volume rate of flow (or discharge) is determined from
continuity as

Q = A × u, (2.72)

where A is the cross-section area of the flow while u is taken to be the mean velocity
of flow. The cross-section area of flow at any point would be determined from the
dimensions of the venturimeter as provided by the manufacturer/vendor, and therefore
the main task would be to determine the mean velocity of flow.
Assuming that there are no energy losses and applying Bernoulli’s equation to section
1 and 2;
p1 u2 p2 u2
z1 + + 1 = z2 + + 2
ρg 2g ρg 2g
2 2
u2 − u1 p1 − p2
= + (z1 − z2 ). (2.73)
2g ρg
For continuity of flow,

A1 u1 = A2 u2 ,
A1
u2 = u1 ,
A2
 2
2 A1
∴ u2 = u21 ,
A2
substituting this in Eq. (2.73),
 2  
2 A1 2 p1 − p2
u1 − u1 = 2g + z1 − z2 ,
A2 ρg
 2   
2 A1 p1 − p2
u1 − 1 = 2g + z1 − z2 ,
A2 ρg
 2 2
  
2 A1 − A2 p1 − p2
u1 = 2g + z1 − z2 ,
A22 ρg
A22
 
2 p1 − p 2
u1 = 2 2g + z1 − z2 ,
A1 − A22 ρg
s  
A2 p1 − p 2
u1 = p 2 2g + z1 − z2 . (2.74)
(A1 − A22 ) ρg
Volume rate of flow, Q = A1 u1 = A2 u2 . Therefore,
s  
A1 A2 p1 − p2
Q= p 2 2g + z1 − z2 .
(A1 − A22 ) ρg

67
For a horizontal pipe, z1 − z2 = 0, thus,
s  
A1 A2 p1 − p2
Q= p 2 2g .
(A1 − A22 ) ρg

p1 − p2
To simplify, we define H = where H is the pressure difference expressed as
ρg
a head of the liquid flowing in the pipe. It is NOT the difference in the levels of the
manometric fluid.
A1 A2 p
Q= p 2 2gH. (2.75)
(A1 − A22 )

Equation (2.75) then gives the theoretical discharge through a horizontal venturimeter.
To get the actual discharge, Qact , we multiply Q by a coefficient of discharge, Cd,
which is found experimentally.
Although Cd varies somewhat with the rate of flow, the viscosity of the fluid and
the surface roughness, a value in the range of 0.95 - 0.99 is usual with fluids of low
viscosity.
A1 A2 p
Qact = Cd p 2 2gH. (2.76)
(A1 − A22 )

H can be expressed in terms of the difference in the levels of the manometric fluid as
 
ρman
H= − 1 h.
ρ

EXAMPLE 10
A venturimeter having a throat diameter d2 of 100 mm is fitted into a horizontal
pipeline which has a diameter d1 = 250 mm through which oil of specific gravity 0.9 is
flowing. The pressure difference between the entry and the throat tappings is measured
by a differential U-tube manometer, containing mercury of specific gravity 13.6.
If the difference of level indicated by the mercury in the U-tube is 0.63 m, calculate
the theoretical flow rate through the meter. Give your answer in litres/s.
SOLUTION 10
Given:
d1 = 250 mm, d2 = 100 mm, h = 0.63 m, ρ = 0.9 x 1000 = 900 kg/m3 and ρman =

68
Figure 2.14: Venturimeter for Example 10

13.6 x 1000 = 13,600 kg/m3 , then


A1 A2 p
Q= p 2gH,
(A21 − A22 )
 
ρman
H= −1 h
ρ
 
13600
= − 1 × 0.63 = 8.89 m,
900
π
A1 = d21
4
π
= × 0.252 = 0.049 m2 ,
4
π 2
A2 = d2
4
π
= × 0.12 = 7.854 × 10−3 m2 ,
4
0.049 × 7.854 × 10−3 √
∴Q= p 2 × 9.81 × 8.89
(0.0492 − (7.854 × 10−3 )2 )
= 0.105 m3 /s ≡ 105 l/s.

Tutorial Problem 8
A pipe of diameter 100 mm has to carry a maximum flow rate of water of 0.0341 m3 /s
which is to be measured by means of a horizontally mounted venturimeter. If the
maximum difference in the levels of the manometric fluid in the U-tube manometer
at maximum flow is 600 mm, determine the throat diameter of the venturimeter, for
maximum flow. Take Cd = 0.96 for the meter and the specific gravity of mercury to
be 13.6. [59.1 mm]
ASSIGNMENT 2
1. (a) List the various assumptions used in the derivation of Bernoulli’s equation.
(b) Apply the above assumptions to the x-component of the Navier-Stokes equa-
tions so that it simplifies to Euler’s equation.

69
2. A horizontally-mounted venturimeter is used to measure the discharge of a fluid
of density ρ, flowing in a horizontal pipe. If the manometric fluid in the U-tube
manometer has density ρman , and the difference in the levels of the manometric
fluid is h, show, with the aid of a well-labelled sketch, that
 
ρman
H= − 1 h,
ρ

where H is the difference in pressure head of the flowing fluid between the inlet
and the throat of the venturimeter.

2.4.4.2 The Flow-nozzle


The nozzle-meter or flow-nozzle, illustrated in Fig. 2.15, is essentially a venturimeter
with the divergent part omitted, and the basic equations are the same as those for the
venturimeter.

Figure 2.15: The flow-nozzle

Much of the kinetic energy of the jet from the flow-nozzle is dissipated in eddies
downstream and so, the overall loss of useful energy is considerably larger than for a
venturimeter. This disadvantage however, is offset by the lower cost of the nozzle.
2.4.4.3 The Orifice meter
An orifice plate inserted in a pipeline produces the same effect as a venturimeter. The
orifice plate has an opening in it that is smaller than the internal diameter of the
pipeline.

70
Figure 2.16: Pipe orifice meter

This arrangement is cheap compared to the cost of a venturimeter, but there are
substantial energy losses because of the abrupt expansion after the plate. Just like
in a venturimeter, the theoretical discharge, Q, can be calculated from Eq. (2.75),
though the actual discharge may be as little as two-thirds of the theoretical discharge.
A coefficient of discharge must therefore be introduced in the same way as for the
venturimeter.
Tutorial Problems 9

1. An orifice meter consists of a 100 mm diameter orifice in a 250 mm diameter


pipe, and has a coefficient of discharge of 0.65. The pipe conveys oil of specific
gravity 0.9 and the pressure difference between the two sides of the orifice plate
is measured by a mercury manometer, the leads to the manometer being filled
with oil. If the difference in mercury levels in the gauge is 760 mm, calculate the
rate of flow of oil in the pipeline. [0.075 m3 /s]

2. With the aid of a sketch, explain what a rotameter is.

2.4.4.4 Pitot tube


A point in a fluid stream where velocity is reduced to zero is called a stagnation point,
and a non-rotating obstacle in the path of a flowing stream produces a stagnation point
on its upstream surface, e.g., point B on Fig. 2.17.
On either side of the central streamline AB, the flow is deflected round the object
but along AB, the velocity decreases until it is zero at point B. From Bernoulli’s

71
Figure 2.17: Stagnation point

equation,

pA u2 pB u2
+ zA + A = + zB + B . (2.77)
ρg 2g ρg 2g
For a horizontal application, zA = zB and at point B, uB = 0. Therefore,

pA u2A pB
+ = , (2.78)
ρg 2g ρg
or,

ρu2A
pB = pA + . (2.79)
2
Pressure pB is called stagnation pressure of the central streamline AB and is made up
ρu2
of the static pressure pA and the dynamic pressure A . The stagnation pressure
2
represents the conversion of all kinetic energy along a streamline into a pressure rise.
Revisiting Bernoulli’s equation, we may write,

ρu2
p + + ρgz = total pressure = constant. (2.80)
|{z} 2 |{z}
static pressure hydrostatic pressure
|{z}
dynamic pressure

In other words, Bernoulli’s equation is a statement that the total pressure remains
constant along a streamline.

Figure 2.18: Pitot tube

The Pitot tube (named after H. de Pitot (1695 - 1771)), applies the principle of
stagnation.

72
It is a right-angled glass tube, large enough for capillary effects to be neglected, that
is inserted into the flow as shown in Fig. 2.18. The Pitot tube measures the stagnation
pressure. A piezometer tube is also connected to measure the static pressure. The
velocity of the flow is determined from Eq. (2.79), by measuring the difference between
the static pressure pA and the stagnation pressure pB , i.e.,
ρu2A
∆p = pB − pA = ,
2
∆p pB pA
∴ = −
ρg ρg ρg
= htotal − hstatic
= h (velocity head)
u2A
= ,
2g
p
∴ uA = 2gh. (2.81)

This velocity is the theoretical velocity and the actual velocity can then be determined
by incorporating the co-efficient of the Pitot tube, Cv , therefore,
p
uA,act = Cv 2gh. (2.82)

Piezometer tubes as shown in Fig. 2.18 are suitably replaced by connections to a dif-
ferential manometer. Frequently, the tubes recording the static and dynamic pressures
are conveniently combined into one instrument known as a Pitot-static tube (students
to sketch and describe a Pitot-static tube).
Tutorial Problems 10
1. A submarine submerged in sea-water of density 1026 kg/m3 , travels at 16 km/hr.
Calculate the pressure at the front stagnation point situated 15 m below the
surface. [161.1 kPa]

2. A submarine moves horizontally in the sea and has its axis 20 cm below the
surface of water. A Pitot-static tube placed in front of the submarine and along
its axis, is connected to two limbs of a U-tube manometer containing mercury.
The difference of mercury level is found to be 20 cm. Find the speed of the
submarine, if the specific gravities of mercury and sea-water are 13.6 and 1.026
respectively. [24.958 km/hr]

2.4.5 Notches and weirs


2.4.5.1 Notches
A notch is an opening in the side of a measuring tank or reservoir extending above
the free surface. It is, in effect, a large orifice which has no upper edge, so that it has

73
a variable area depending upon the level of the free surface. Two common types of
notches are rectangular and vee notches [5, 8].

• Discharge from a rectangular notch


Consider a horizontal strip of width B and height dh, at a depth h below the free
surface.

Figure 2.19: Discharge from a rectangular notch

– Area of the shaded strip, dA = Bdh.



– Velocity through the strip, dv = 2gh.
– Discharge through the strip, Q = area x velocity = dA dv.

p
∴ dQ = B 2gh dh
p 1
= B 2g h 2 dh,

integrating from h = 0 at the free surface, to h = H at the bottom of the notch


gives the total theoretical discharge;
p Z H
1
Q = B 2g h 2 dh
0
2 p 3
= B 2g H 2 . (2.83)
3

• Discharge from a Vee-notch


Consider a horizontal strip of width b and height dh, at a depth h from the free
θ
surface, in a vee-notch of included angle θ. From Fig. 2.20, b = 2(H − h)tan .
2
θ
– Area of the strip, dA = bdh = 2(H − h)tan dh.
2

– Velocity through the strip, dv = 2gh.

74
Figure 2.20: Discharge from a vee notch

– Discharge through the strip, dQ = dA dv.

θp
∴ dQ = 2(H − h)tan 2gh dh
2
p θ 1
= 2 2g tan (H − h)h 2 dh,
2
the total theoretical discharge, Q is then determined by integration thus,
θ H
Z
p 1 3
Q = 2 2g tan (Hh 2 − h 2 )dh
2 0
 H
p θ 2 3 2 5
= 2 2g tan Hh 2 − h 2
2 3 5 0
 H
p θ 2 5 2 5
= 2 2g tan H2 − H2
2 3 5 0
8p θ 5
= 2g tan H 2 . (2.84)
15 2
As is the case of orifices, the actual discharge through a notch can be found by
multiplying the theoretical discharge by a co-efficient of discharge to allow for
energy losses.

2.4.5.2 Weirs
A weir is a notch on a large scale. It is a barrier or dam placed in the channel so that
water flows over the top.
A weir is normally made of concrete or masonry, and has long been used as a stan-
dard device for the measurement of water flow in an open channel. The sheet of liquid
escaping over the weir is called a nappe or vein.

75
Figure 2.21: Flow over a weir

Some types of weirs

• Sharp-crested weir : This is the term more often used for a large rectangular
notch.

• Suppressed rectangular weir : This type of weir is as wide as the channel, and the
width of the nappe is the same as the length of the crest.

Tutorial Problems 11

1. In an experiment on a 900 vee notch, the flow is collected in a 0.9 m diameter


vertical cylindrical tank. It is found that the depth of water increases by 0.685
m in 16.8 s when the head over the notch is 0.2 m. Determine the coefficient of
discharge of the notch. [0.615]

2. Water flows through a triangular right-angled weir first and then over a rectangu-
lar weir of 1 m width. The discharge co-efficients of the triangular and rectangular
weirs are 0.6 and 0.7 respectively. If the depth of water over the triangular weir
is 360 mm, find the depth of water over the rectangular weir. [141.5 mm]

76
REFERENCES
[1] Douglas, J. F, Solving Problems in Fluid Mechanics volume 1. Longman Singapore,
1986.

[2] Francis, J. R. D., Fluid Mechanics for Engineering Students. Edward Arnold, 1975.

[3] Robertson, J. A. and Crowe, C. T., Engineering Fluid Mechanics. John Wiley &
Sons, Inc., 6th ed., 1997.

[4] Gupta, V. and Gupta, S. K., Fluid Mechanics and its Applications. Wiley Eastern
Limited, 1984.

[5] Bansal, R. K., Fluid Mechanics and Hydraulic Machines. Laxmi, 1989.

[6] Bolmstedt, U., “Viscosity and rheology; theoretical and practical considerations in
liquid food processing,” New Food, vol. 3, Issue 2, 2000.

[7] Fung, D. Y. C. and Matthews, R. F., Instrumental Methods for Quality Assurance
in Foods. ASQC Quality Press, 1991.

[8] Massey, B. S., Mechanics of Fluids. Van Nostrand Reinhold (UK), 1983.

[9] Bird, Byron R., Warren, E., and Lightfoot, Edwin N., Transport Phenomena. John
Wiley & Sons, Inc., 2nd ed., 2002.

77

You might also like