MSC - Maths - 4sem - 4.3 Fluid Mechanics (Math 4.3) PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 124

KARNATAKA STATE OPEN UNIVERSITY

Mukthagangotri, Mysore – 570 006

M.Sc. MATHEMATICS
(FOURTH SEMESTER)

COURSE CODE: MATH 4.3

FLUID MECHANICS
Math 4.3

Fluid Mechanics

K.V. Prasad
Professor and Chairman
Department of Mathematics,
VSK University, Bellary,

i
Course Design and Editorial Committee
Prof. M.G. Krishnan Dr. S.N. Vikram Raj Urs
Vice-Chancellor & Chairperson Dean(Academic) & Convenor
Karnataka State Open University Karnataka State Open University
Mukthagangothri, Mysore - 570 006 Mukthagangothri, Mysore - 570 006

COURSE WRITER BLOCK - 1, 2 3 & 4

K. V. Prasad
Department of Mathematics,
Vijayanagara Sri Krishnadevaraya University,
Bellary 583 104

COURSE CO-ORDINATOR

Mr. Chandru Hegde


Assistant Professor and Chairman
Department of Studies in Mathematics
Karnataka State Open University
Mukthagangothri, Mysore - 570 006

PUBLISHER

The Registrar
Karnataka State Open University
Mukthagangothri, Mysore - 570 006

Developed by Academic Section, KSOU, Mysore


Karnataka State Open University, 2013

All rights reserved. No part of this work may be reproduced in any form, by mimeograph or any other means,
without permission in writing from the Karnataka State Open University.

Further information on the Karnataka State Open University Programmes may obtained from the University’s
office at Mukthagangothri, Mysore - 6

Printed and Published on behalf of Karnataka State Open University. Mysore – 6 by Registrar
(Administration)

ii
BLOCK INTRODUCTION

Fluid mechanics deals with fluid flow, which is the natural science of fluids (liquids
and gases) in motion. Fluid dynamics has a wide range of applications, including calculating
forces and moments on aircraft, determining the mass flow rate of petroleum through
pipelines, predicting weather patterns, understanding nebulae in interstellar space and
reportedly modelling fission weapon detonation. Some of its principles are even used in
traffic engineering, where traffic is treated as a continuous fluid. This course is designed to
give an overview of fluid dynamics from a mathematical viewpoint, and to introduce students
to areas of active research in fluid dynamics.

In Block 1, the basic concepts required in the development of the theory of viscous
flow have been discussed and the constitutive equation for an isotropic Newtonian fluid is
derived in Cartesian tensors.

The Block 2 deals with the fundamental equations, which govern the motion of a
viscous compressible fluid. The derivation of the equations has been made simple and
concise with the help of Cartesian tensors. The governing equations, for compressible and
incompressible fluid motion, in Cartesian, cylindrical and spherical polar coordinates are
given in tabular form for ready reference.

In Block 3 the topics like Dynamical similarity and Dimensional analysis have been
discussed and the physical importance of the non-dimensional parameters and coefficients,
which play an important role in the study of the flow of viscous fluids, is given.

Since the Navier-Stoke’s (N-S) equations of motion are non- linear in character, there
is no known general method to solve these equations. Only in few special cases exact
solutions can be obtained by making certain assumptions about the state of the fluid and a
simple configuration for the flow pattern. The exact solutions of N-S equations discussed in
Block 4, which are valid for all ranges of viscosity but are few in number. However, there
appear, at present, no such exact solution of the flow past bodies of finite size and, therefore,
in order to study such flows the N-S equations have been simplified to mathematically
tractable forms for two extreme cases.

iii
TABLE OF CONTENTS

Block 1 Viscous Fluid Dynamics

Basic concepts: Fluid, Continuum hypothesis, Viscosity, General 1-11


Unit-1
motion of a fluid element.

Rate of Strain quadric, stress at a point, Symmetry of stress matrix, 12-20


Unit-2
Stress quadric, stress in a fluid at rest, stress in a fluid in motion.

Relation between stress and strain components (Stokes law of 21-24


Unit-3
friction)

Unit-4 Thermal conductivity, generalized law of heat conduction. 25-28

Block 2 Fundamental Equations of the Flow of Viscous Fluids.

Introduction, Equation of state, Equation of continuity 29-33


Unit-5
conservation of mass

Unit-6 Equation of motion: conservation of momentum 34-36

Unit-7 Equation of energy: conservation of energy 37-42

Unit-8 Vorticity and circulation in a viscous incompressible fluid motion 43-47

Block 3 Dynamical Similarity, Inspection and Dimensional Analysis

Unit-9 Introduction: dynamical similarity (Reynolds Law) 48-52

Inspection analysis, dimensional analysis, Buckingham pai- 53-57


Unit-10
theorem

Unit-11 Physical importance of non-dimensional parameters 58-64

Important non-dimensional co efficient in the dynamics of viscous 65-68


Unit-12
fluids

Block 4 Exact Solutions of the Navier Stokes Equations

Introduction: Flow between parallel plates: velocity and 69-80


Unit-13
Temperature distribution

Unit-14 Flow in a circular pipe 81-86

Unit-15 Flow between two concentric rotating cylinders 87-99

Unit-16 Flow due to a rotating disc 100-113

Appendix 114-119

iv
Block – 1
Viscous Fluid Dynamics

1
Unit – 1
Basic concepts: Fluid, Continuum hypothesis, Viscosity,
General motion of a fluid element

Structure:
1.0. Objectives
1.1. Introduction
1.2. Basic concepts, fluid, Continuum hypothesis
1.3. Viscosity
1.4. General motion of a fluid element
1.5. Exercises
1.6. Keywords
1.7. References

1.1. Objectives

The basic concepts required in the development of the theory of viscous flow have
been discussed. Mainly in this unit we introduce Continuum hypothesis, Viscosity and
General motion of a fluid element.

1.2. Basic Concepts

All materials exhibit deformation under the action of forces. If the deformation in
the material increases continually without limit under the motion of shearing forces,
however small, the material is called a ‘fluid’. This continuous deformation under the
action of forces is manifested in the tendency of fluids to flow. Fluids are usually
classified as liquids or gases. A liquid has interamolecular forces which hold it together
so that it possesses volume but no definite shape. When it is poured in to a container will
fill the container up to the volume of the liquid regardless of the shape of the container.
Liquids have but slight compressibility. For most purposes it is, however, sufficient to
regard liquids as ‘incompressible fluids’. A gas, on the other hand, consists of molecules
in motion which collides with each other tending to disperse it so that a gas has no set
volume or shape. The interamolecular forces are extremely small in gases. A gas will fill
any container in to which it is placed and is therefore known as a (highly) ‘compressible
fluid’.

It is proper to remark here that for speeds which are not comparable with that of
sound i.e., if the Mach number (the ratio of the velocity of flow to the velocity of sound)
is small compared with unity the effect of compressibility on atmospheric air can be
neglected and it may be considered to be a liquid, and in this sense it is called

2
incompressible air. Taking the velocity of sound about 330m/sec a flow velocity of
99m/sec may be accepted as the upper limit when a gaseous flow can be considered
incompressible.

1.3. Continuum hypothesis

In its most fundamental form, at the microscopic level, the description of the
motion of a fluid involves a study of the behaviour of all the discrete molecules which
make up the fluid. However, when one is dealing with problems in which some
characteristic length in the flow is very large compared with molecular distances, it is
convenient to think of a lump of fluid sufficiently small from macroscopic point of view
but large enough at the microscopic level so as to contain a large number of molecules
(for instance, at normal temperature and pressure a volume of 1012 cc of a gas contains
about 27  107 molecules) and to work with the average statistical properties of such
large number of molecules. In such a case the detailed molecular structure is washed out
completely and is replaced by a continuous model of matter having appropriate
continuum properties so defined as to ensure that on the macroscopic scale the behaviour
of the model resembles with the behaviour of the real fluid. When the characteristic
length in the flow is not large compared with molecular distances, the continuum model
is invalid and the flow must be analyzed on the molecular scale. The smallest lump of
fluid material having sufficiently large number of molecules to allow statistically of a
continuum interpretation is here called a ‘fluid particle’. The material in this book will
deal primarily with fluids obeying continuum hypothesis.

1.4. Viscosity

Viscosity of the fluid is that characteristics of real fluids which exhibit a certain
resistance to alterations of form. Viscosity is also known as internal friction. All known
fluids (gases or liquids) possess the property of viscosity in varying degrees. The nature
of viscosity can be best illustrated with the aid of the following experiment of Newton:

Fig 1.1. Fluid motion between a stationary plate and a moving parallel plate.
3
Consider the motion of a fluid between two very long parallel plates at a distance, say, ‘d’
apart, one of which is at rest and the other moving with a constant velocity U parallel to
itself, as shown in Fig. 1.1. Because of viscosity, the fluid will also be in motion.
Experiment teaches us that the fluid adheres to both walls (no slip condition), so that its
velocity at the lower plate, which is stationary, is zero and at the upper plate, which is
moving, is equal to the velocity of the plate U. Experiments have shown that for a large
class of fluids the tangential stress  acting on either of the plates is proportional to the
relative velocity between the plates and inversely proportional to the distance d. Thus we
have
U
  , (1.1)
d
Where  is a constant of proportionality, it is independent of U and d and depends only
on the nature of the fluid. This constant is a measure of viscosity of the fluid and is called
the ‘coefficient of shear viscosity’ or simply the ‘coefficient of viscosity’.

For ordinary fluids, since there is no slip on the walls and the fluid is displaced in such a
manner that the various layers of fluid slide uniformly over one another, the viscosity u of
a layer of the fluid at a distance y from the lower plate is then

y
u U (1.2)
d

It may be seen from (1.2) that U d in (1.1) may be replaced by the velocity gradient
du dy , hence

du
  (Newton’s law of viscosity) (1.3)
dy

Equation (1.3) may be regarded as the definition of viscosity. Thus the coefficient of
viscosity of a fluid may be defined as the tangential force required per unit area to
maintain a unit velocity gradient, i.e., to maintain unit relative velocity between two
layers unit distance apart.

The dimension of the coefficient of viscosity  can be found as follows:


shearing stress force/area
 =  M L1T 1
velocity gradient velocity/length
As we shall see later, the effect of viscosity on the motion of a fluid is determined by the
ratio of  to the density  rather than by  alone. This ratio is known as the
‘kinematic viscosity’ and is usually denoted by , thus

4
 M L1T 1
   L2T 1 .
 ML 2

Some typical values of  and  are given below in c.g.s. units for gases and liquids at

150 C and under atmospheric pressure.

Gases Liquids

   

Air 0.00018 0.15 Water 0.0114 0.0114

Nitrogen 0.00017 0.15 Mercury 0.016 0.0012

Oxygen 0.0002 0.15 Paraffin oil 0.2 0.25

Hydrogen 0.00009 1.5 Glycerine 13 10

Helium 0.0002 0.12 Castor oil 15 15

Carbon dioxide 0.00014 0.077 Pitch 1010 1010

For liquids the viscosity  is nearly independent of pressure and decreases


rapidly with increasing temperature. In the case of gases, to a first approximation, the
viscosity can be taken to be independent of pressure but it increases with temperature. It
is small for ‘thin’ fluids, such as water or air, but large in the case of very viscous liquids,
such as oil or glycerine. It is appropriate to remark here that the example considered in
Fig 1.1 constitutes a particularly simple case of fluid motion. The generalization to the
case of three- dimensional flow is contained in ‘Stokes law of friction’, the theory of
which will now be developed.

1.5. General motion of a fluid element

Fig 1.2. Relative motion between two neighboring points of a fluid element.

5
In general, the motion of a fluid element consists of three parts (i) a translation,
(ii) a rotation, and (iii) a deformation. We shall show this by considering the relative
motion between two neighboring points of a fluid element. Let P and Q denote any two
  
such points, at a time t, and let r and r  dr be their position vectors referred to a fixed
 
point O (fig. 1.2). If V (r , t ) is the velocity at the point P at a time t, the velocity at Q at
the same instant is expressed to a first order by

         
V ( r  dr , t )  V (r , t ) d V  V (r , t ) (dr , grad )V . (1.4)
 
It is clear from above that in the velocity of the point Q, there is a part viz., V (r , t )
which is the same as that of P. This part is the same for all points of the fluid element
and, therefore, corresponds to a velocity of translation of the fluid element as a whole.

   
The component d V or (dr grad ) V is the relative velocity between P and Q , can be
shown to be made up of a rotation and a deformation. We will carry out the proof in
Cartesian tensors.
  
Let us denote the Cartesian components of r , dr and V by xi , dxi and vi (i  1, 2,3)

  
respectively. The j th component of (dr grad ) V is then given by

v j  1  v j vi  1  v j vi 
dxi  dxi          (1.5)
xi  2  xi x j  2  xi x j  

1  v j vi 
Let eij     (1.6)
2  xi x j 

1  v j vi 
ij     (1.7)
2  xi x j 
Clearly eij  e ji and ij  ji (1.8)

Then (1.5) may be written as


v j
dxi  eij dxi  ij dxi (1.9)
xi
v j 
We note the tensor , the tensor gradient of V , is a second order tensor. Its symmetric
xi

and anti-symmetric parts are eij and ij respectively. Let us first consider the anti-

symmetric tensor ij . Explicitly the tensor is given by the array

6
 0 12 13 
 
ij   12 0 23  (1.10)
  23 0 
 13
Since there are only three independent components of ij we may associate ij with a

vector  k  such that ij   ijk k , (1.11)

where  ijk is the permutation tensor and is defined as

 0, if any two of i, j, k are the same



 ijk  1, if i, j, k is an even permutation of 1,2,3 (1.12)
 1, if i, j, k is an odd permutation of 1,2,3

In order to express k in terms of ij we notice that if we form  ijk ij , then for any fixed

k only the terms with i  k , j  k will appear. Thus


 ij 3 ij  123 12   213 21
 12  21  23

1
It follows that k   ijk ij (1.13)
2
Now, the second term in the right hand side of (1.9), in view of (1.11), may be written as
ij dxi   ijk k dxi (1.14)

The right hand side of equation (1.14) is the jth component of the vector product
 
  dr . This product, as is well known, is the velocity at the point Q due to a rigid body
rotation of the fluid element as a whole about an instantaneous axis through the point P

with an angular velocity  .
 
From equations (1.7) and (1.13) the angular velocity  , in terms of the velocity V , can
be written as
 1 
  curlV (1.15)
2

The vector curlV is usually called as the ‘vorticity vector’ or simply vorticity.

We now turn to the other part eij dxi of the velocity of Q, which we will show is the

velocity due to the deformation of the fluid element. Explicitly we have

v1 v2 v3 1  v v 


e11  , e22  , e33  , e12   2  1   e21 ,
x1 x2 x3 2  x1 x2 
(1.16)
1  v v  1  v v 
e23   3  2   e32 , e31   1  3   e13 ,
2  x2 x3  2  x3 x1 
7
Each of these terms can be given a kinematic interpretation, and we now proceed to
obtain it.
v1 ( x1   x1 , x2 , x3 , t )  v1 ( x1 , x2 , x3 , t )
e11  lim
 x1 0  x1
rate of increment in length of  x1
 lim
 x1 0 original length  x1
Thus e11 denotes the relative rate of elongation in the x1 -direction suffered by he

fluid element. We call it the normal rate of strain in the x1 -direction. Similarly the terms

e22 and e33 may be interpreted to mean normal rate of strains in the x2 -and x3 - directions

respectively.

It is now easy to see the distortion imparted to a fluid element by the simultaneous
action of all the three terms e11 , e22 and e33 . The fluid element expands in all the three
directions with the result that the change in the length of its three sides produces a change
in volume at a relative rate

 v1  v  v 
 dx1  dx1 dt  dx2  2 dx2 dt   dx3  3 dx3 dt   dx1 dx2 dx3
x1 x2 x3
   
dx1 dx2 dx3 dt (1.17)
v1 v2 v3 
    divV
x1 x2 x3

to a first order in derivatives. During this distortion only the size of the element changes
while the shape remains similar, since all right angles at the vertices continue to be that
way. This relative rate of increase in the volume, i.e.,  r is called the dilation of the fluid
element and is zero when the fluid is incompressible, as must be the case.

Consider next the term e12 given by

v2 v1
2e12  
x1 x2

v2
For the kinematic interpretation of the derivative , we have
x1
v2 v ( x   x1 , x2 , x3 , t )  v2 ( x1 , x2 , x3 , t )
 lim 2 1
x1  x1 0  x1
rate of displacement between the end points of  x1 in the x2 - direction
 lim
 x1 0  x1
We see that due to this relative rate of displacement the line element  x1 , which was

originally in the x1 -direction, now turns through a small angle, say 1 .

8
Fig 1.3. Relative displacement between the end points of x1 in x2 direction
In fact, see fig.1.3,
v2 v2 1
 t  tan 1  1 or  (1.18)
x1 x1 t
Similarly, we find that
v1  2
 (1.19)
x2 t
 
   1   2 
It follows that 2e12  2  (1.20)
t

Fig 1.4. Relative displacement between the end points of x2 in x1 direction

Fig 1.5. Distortion of a fluid element.

9
Thus a rectangular element of fluid centred at P now distorts in to a parallelogram as
indicated in fig.1.5. Hence 2e12 represents the rate of the decrease of the angle between

the lines which were originally parallel to x1 and x2 axes respectively. We call e12 the

rate of shearing strain in the ( x1 , x2 ) -plane. Similarly we may interpret e23 and e31 as the

rate of shearing strain in the ( x2 , x3 ) -plane and in ( x3 , x1 ) -plane respectively. This


distortion is volume preserving and effects only the shape of the element.

 e11 e12 e13 


 
Thus, eij   e21 e22 e23  (1.21)
e e33 
 31 e32

is a symmetric second order tensor, the diagonal elements of which represent rates of
normal strains while the non-diagonal terms represent rates of shearing strains. The tensor
eij is therefore known as the rate of strain tensor. We see that the rate of normal and

shearing strains give rise to deformation (change of shape and size) of the fluid element.
Thus the component velocity eij dxi represents the velocity due to the deformation of the

fluid element.

We have thus shown that the most general motion of a fluid element consists of a
translation, a rotation and a deformation.

It may be remarked here that the mathematical properties of the tensor eij are
analogous to those of a symmetric second order tensor, and therefore the scalar invariants
of it are

a) e11  e22  e33  2,

b) e11e22  e22 e33  e33e11  e122  e23


2
 e31
2
,

 
c) det eij

1.6. Summary:
In this chapter the authors explains the basic concepts required in the development
of the theory of viscous flow, predominantly definition of viscosity, Newton’s law of
viscosity and general motion of fluid particles.

1.7. Exercises:

1) Define the terms fluid, compressible and incompressible fluids.


2) Explain the concept of continuum hypothesis.
3) Write short note on viscosity and hence explain coefficient of viscosity.
4) Explain in detail the general motion of fluid element.
10
5) Define viscosity of the fluid.
6) Define (i) viscous fluid (ii) shearing stress (iii) laminar flow (iv) turbulent
flow (v) steady and unsteady flow (v) uniform and non-uniform flow.

1.8. Keywords:
Fluid, stress and strain tensor, viscosity,

1.9. References:
1) Bhatnagar. P. L, ‘Elements of Dimensional Analysis’, National Pub House, New Delhi
2) Binder R. C., ‘Advanced Fluid Mechanics’, Prentice-Hall of India, New Delhi (1964).
3) Bird. R. B. et al: ‘Transport Phenomena’, John Wiley, New York (1960).
4) Chorlton. F. ‘Text Book of Fluid Dynamics, ELBS (1970).
5) Bansal. J. L: ‘Viscous Fluid Dynamics’, Oxford and IBH publishing Co. Pvt. Ltd., New
Delhi (1977).

11
Unit – 2
Rate of Strain quadric, Stress at a point, Symmetry of stress
matrix, Stress quadric, Stress in a fluid at rest

Structure:
2.0. Objectives
2.1. Rate of strain quadric
2.2. Stress at a point
2.3. Symmetry of stress matrix
2.4. Stress quadric
2.5. Stress in a fluid at rest
2.6. Stress in a fluid in motion
2.7. Exercises
2.8. Keywords
2.9. References

2.0. Objective
At the end of reading this unit students are able to understand
a) Rate of Strain quadric
b) Stress at a point
c) Symmetry of stress matrix
d) Stress quadric, Stress in a fluid at rest
e) Stress in a fluid in motion
2.1. Rate of strain quadric

The coordinates of Q relative to P are ( x1 ,  x2 ,  x3 ) which, for the sake of convenience,


we propose to write as ( X1 , X 2 , X 3 ) then


1
2
 e11 X 12  e22 X 22  e33 X 32  2e23 X 2 X 3  2e31 X 3 X 1  2e12 X 1 X 2 
(1.22)
 constant

is the equation of a quadric, with its centre at P, called the rate of strain quadric.
Properties of 

(i) In Cartesian tensor notation  can be written as


1
  eij X i X j (1.23)
2
Differentiating  with respect to X j , we get

12
 1 1 X i
 eij X i  eij X k
x j 2 2 xk
1 1
 eij X i  eij X k  ik (1.24)
2 2
1 1
 eij X i  eij X i  eij X i  eij  xi
2 2
Where  ik is the Kronecker delta and is defined as

 1, i  k
 ik   (1.25)
0, i  k

Since eij  xi is the jth component of the deformation velocity,  can be regarded as the

potential giving the deformation velocity.

(ii) Let r be the length of the radius of the rate of strain quadric drawn from P in the
direction  whose direction cosines are li , then

X i  rli (1.26)

1
Therefore,   eij X i X j =constant,
2
becomes r 2 li l j eij  constant

or r 2 e  constant ,

or e  constant (1.27)


r2
Since li l j eij  e (1.28)

Form (1.27) we conclude that the rate of normal strain at any point P in any direction  is
inversely proportional to the square of the length of the radius vector, of the rate of strain
quadric at P, drawn in the direction of  .

(iii) If the rate of strain quadric be referred to a set of axes through P parallel to a set
of new axes ( X '1 , X '2 , X '3 ) , then the coefficients in the transformed equation of the

quadric will be e11' ' ,....., e2'3' ,.... when the new axes are taken along the principal axes of

the quadric, then e2'3' , e31' ' , and e12' ' vanish and the rate of strain quadric take the form

e11' ' X12  e2'2' X 22  e33' ' X 32  constant.


' '
(1.29)

It follows that the rate of shearing strains along these axes are zero. Hence at any point of
a fluid in motion there is a set of three mutually perpendicular straight lines such that, if
these lines move with the fluid, after a small time  t ' the angle between them continues

13
to be right angle to the first order in  t ' . These lines are called the principal axes of the

rate of strain.

2.2. Stress at a point

Fig 1.6. Specification of stress at a point.

Let us consider a point P in the fluid and take an infinetisimal area  s surrounding the

point P. We denote by n the unit vector in the direction of the normal to the plane area
 s and consider the forces excreted across  s by the portion of the fluid which lies on

the side of n . The fluid may be either in motion or at rest. These surface forces are not,
in general, distributed uniformly across  s . We represent these forces as equivalent to a
 
force  F acting at the point P and a couple of moment  M about some axis through P.
If we gradually shrink the area of the plane surface to the point P (within the limitations
 
of the continuum hypothesis), both  F and  M will tend to zero. However, for a

vanishingly small area the equivalent surface force  F may be assumed to be
proportional to the surface area. Then the ratio of force to the area tend to a definite limit
as the area shrinks to a point, and it follows that the ratio of the couple (with vanishingly
small moment arm) to the area must ultimately vanish.

This means that the action of the internal forces in the immediate neighborhood of a point

P across the surface, normal to n , can be specified by the limit of the ratio of forces to

 

area  lim  F  s  , i.e., by force per unit area at that point. This is called the stress at the
  s 0 

point P across a plane n . It is a vector and its direction, in general, different from that of

14

n . For a fluid at rest, it is normal to the surface and is in the nature of a pressure but
when the fluids are in motion, there are also shearing stresses in addition to normal stress.

The definition of a stress vector involves, naturally, two directions, viz., (i) the direction
of the normal to the area, and (ii) the direction of the stress vector itself.

The above definition of a stress vector enables us to specify the state of internal forces at

a point P across only a certain plane n . In order to have a complete specification of the
internal forces at P we should know the stress vectors at P across all the planes, infinite in
number that can be drawn through P. However, by considering the equilibrium of an
infinitesimal tetrahedron, it can be easily verified that if the stress vectors across three
independent planes passing through a point are given, the stress vector across any other
plane passing through that point is determined.

It thus follows that the state of stress at any point, is completely specified by giving a set
of three independent stress vectors. Equivalently we can give the components of these
stress vectors. When the normal is in the direction of x2 -axis the components of the

stress parallel to the three coordinate axes are denoted by 11 , 12 , and 13 respectively,

when the normal is in the direction of the x2 -axis the components are denoted by

 21 ,  22 , and  23 and when it is in the direction of the x3 -axis then by  31 ,  32 , and  33


respectively. Hence the stress at a point is completely defined by its nine components,
which may be represented, in short, by  ij or explicitly by the matrix

  11  12  13 
 ij    21  22  23  (1.30)
 
 31  32  33 

where the first subscript indicates the axis to which the plane face is perpendicular and
the second indicates the direction to which the stress is parallel or we may say that  ij is

the stress acting, on an area perpendicular to the xi -axis, taken in the direction parallel to

x j -axis. The diagonal elements of the matrix (1.30) are known as normal stresses and the

non-diagonal elements as the shearing stresses. We shall now show that  ij is a

symmetric second order Cartesian tensor.

15
2.3. Symmetry of stress matrix  ij

Fig 1.7. Derivative of the symmetry of stress matrix.

In general, the motion of a fluid element can be separated in to an instantaneous


translation and an instantaneous rotation. Only the later needs to be considered for our
purpose.

Denoting the instantaneous angular acceleration of the element by   1 ,  2 ,  3  , we


. . ..

 
can write for the rotation about x3 axis

  12  x1   x1   21  x2   x2
3 I 3    12    x2 x3    21    x1 x3 (1.31)
 x1 2  2  x2 2  2

where  I 3 is the moment of inertia of the element about x3 axis. The moment of inertia is
proportional to the fifth power of the linear dimensions of the elementary parallelepiped.
Hence, dividing throughout by  x1  x2  x3 and taking the limit when the fluid element
reduces to the point P, we get
1 1
0   12   21
2 2 (1.32)
or ,  12   21
In a similar way, we find that
 23   32 and  31  13 (1.33)

In other words  ij   ji , (1.34)

which shows that  ij is symmetric.

16
Tensor character of stress matrix  ij

Fig 1.8. Derivative of the tensor character of stress matrix

Let us consider an element of fluid in the form of a tetrahedron (0, ABC) having the edges
OA,OB,OC parallel to the axes of reference. Let the direction cosines of the normal to the
plane ABC be (ln1 , ln 2 , ln3 ) . If S be the area of the face ABC, then the area of the faces

OBC, OCA, and OAB will be Sln1 , Sln 2 and Sln3 respectively.

Since the surface force are proportional to the surface area, whilst the body forces are
proportional to the volume, the surface forces must balance exactly in the limit as the
tetrahedron reduces to a point,
Revolving along OX 1 ,we get S n1  Sln1 11  Sln 2 21  Sln3  31

Therefore  n1  ln1 11  ln 2 22  ln3  31 or, in general,

 nj  lni  ij (1.35)

Let us now consider the transformation of the coordinate axes from X i to X j . Let  i ' j '

denotes the stress, on the plane perpendicular to the X i ' -axis, in the direction of X j '

axis. Evidently this is equal to the algebraic sum of the resolved parts  f of the stress
acting on this plane in the direction of X1 , X 2 , X 3 axes, i.e.,

 i ' j '   i ' j ' li ' j (1.36)

Taking the direction of n parallel to X i ' -axis, from (1.35), we have

 i ' j '  li ' i  ij (1.37)

therefore , from (1.36)and (1.37), we conclude


 i , j ,  li ,i l j , j  ij (1.38)

Hence  ij is a symmetric second order Cartesian tensor.

17
As in the case of rate of strain tensor the three scalar invariants of the symmetric stress
tensor  i , are,

i) 11   22   33

ii) 11 22   22 33   3311   212   223   231

iii)  
det  ij

2.4. Stress quadric

The quadric    ij X i X j  constant (1.39)

is called the stress quadric at the point P  xi  at which  ij are the stress components and

X i are the coordinates of Q, a neighboring point of P, relative to P.

Properties of 
i) Let r be the length of the radius vector, of the quadric drawn from P in the direction 
whose direction cosines are li . Then

xi  rli (1.40)

Therefore  ij r 2 li l j  constant or r 2  = constant

or    constant r 2 (1.41)

which shows that the normal stress on a plane, through P, perpendicular to  direction is
inversely proportional to the square of the radius vector of the stress quadric at P drawn
in the direction of  .
ii) We may define the principal axes of stress, in exactly the same way, as we
defined the principal axes of rate of strain and these axes will have analogous properties.
The planes perpendicular to the principal axes of stress are called principal planes of
stress. Recorded to the principal axes, the stress quadric will take the form
1'1' X 21'   2 ' 2 ' X 22 '   3'3' X 23'  cons tan t (1.42)
meaning thereby that in the principal planes there are no tangential or shearing stresses.
The stress across each of these principal planes is purely normal, and these three normal
stresses are called the principal stresses.

2.5. Stress in a fluid at rest

When the fluid is at rest, it is a fact of experience that the tangential stresses do not exist.
This means that the stress vector at any point of the fluid, which is at rest, is wholly
normal to any plane surface passing through that point. In such a case the stress tensor
takes on a particularly simple form

18
  11 0 0 
 ij   0  22 0 

(1.43)
 0 0  33 

For such a state of stress, considering the equilibrium of an infinitesimal tetrahedron, we
may see that the magnitude of  11 ,  22 ,  33 at a point is the same for all elemental planes
passing through that point. In such a case all that is required to specify the stress at a
point is simply a single number, which we may denote by  , for the time being. If  is
positive,  ij represents a tensile stress which is contrary to the experience, since in the

interior of the fluid no tensile stresses occur. This means that the nature of  ij is of

compression and it will be appropriate to replace  by a negative number, say –P. Then
the stress in a fluid at rest is represented by
p 0 0 
 ij   0  p 0 

or  ij   p ij (1.44)
 0 0  p 

for all orientations of the coordinate axes. The scalar p is called the hydrostatic pressure
at the point considered.

2.6. Stress in a fluid in motion

Both the tangential and normal stresses occur when the fluid is in motion. In such a case
the complete stress tensor can be expressed as the sum of two parts- one which represents
compressive stresses equal in all directions i.e., it represents pressure and this pressure is
also denoted by p, which is similar to but not identical with the hydrostatic pressure. The
other part represents viscous or frictional stresses. Thus the state of stress in a moving
fluid can be expressed as
 ij   p ij   ij (1.45)

It will be seen that the viscous stresses are assumed to be proportional to the rates of
strain occurring at the point considered (c.f.I.12). The proportionality constants, known as
the viscosity coefficients, depend on the nature of the fluid. Thus the viscous stresses
occur only when the fluid is in non-uniform motion, i.e., when there are velocity
gradients. When the motion is uniform, i.e., when the rate of strain disappear the viscous
stresses also disappear and leaving the stress tensor as that of uniform pressure in all
directions. Similarly, if the fluid is at rest, i.e., the velocity of the fluid is zero
everywhere, the viscous stresses become zero, and in that case the uniform pressure is
known as the hydrostatic pressure. It is because of this behaviour the stress tensor  ij is

known as the viscous stress tensor or deviotoric stress tensor.

19
2.7. Exercises
1) Explain the term “rate of strain quadric”.
2) Describe Tensor character of stress matrix.
3) Give brief explanation about the fluid at rest and in motion.

2.8. Keywords

Rate of strain, Stress, Stress quadric.

2.9. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House,


New Delhi (1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi
(1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2,
Oxford University Press (1970).
6. Schlichting. H., Boundary Layer Theory, McGraw-Hill (1968).
7. Streeter .V. L. (Ed)., Hand Book of Fluid Dynamics, McGraw2-Hill (1961).
8. Yuan. S.W: Foundational of Fluid Mechanics, Prentice. Hall of India Pvt.
Ltd., New Delhi (1969).
9. Bansal. J. L., Viscous Fluid Dynamics, Oxford and IBH publishing Co. Pvt.
Ltd., New Delhi (1977).
10. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition,
(2011).

20
UNIT – 3
Relation Between Stress and Strain Components
(Stokes law of friction)

Structure:
3.0. Objectives
3.1. Relation Between Stress and Rate of Strain Components (Stoke’s Law of Friction)
3.2. Interpretation of the constants  and 
3.3. Stoke’s hypothesis
3.4. Exercises
3.5. Keywords
3.6. References

3.0. Objective
At the end of reading this unit students are able to understand
a) Stoke’s law of friction
b) Stoke’s Hypothesis
3.1. Relation Between Stress and Rate of Strain Components (Stoke’s Law of
Friction)
We shall not able to deduce the equations of fluid motion in a mathematically
solvable form unless we know some relationship between the stress and rate of strain
components. Such a relation can be either derived theoretically or on the basis of
experimental results. No experimental results could indicate, so far, any perfect
relationship between these components. Hence, there is only one alternative left with us,
i.e., to assume, arbitrarily, some theoretical relation between these two components and
study the implications of the assumptions. If the theoretical results agree favourably with
the experimental results of some fluids then one could assert that the assumed
relationship is plausible for those fluids.

Stokes made the following assumptions in order to find one such relationship connecting
the stress and rate of strain components:
i) The stress components are linear functions of the rate of strain components
(Newtonian fluid),
ii) The relations between stress components and rate of strain components are invariant
to orientation of the coordinate axes i.e., they remain unchanged by a rotation of the
system of coordinates or by an interchange of axes (Isotropic fluid)
iii) When all the velocity gradients are zero the stress components must reduce to
hydrostatic pressure.

21
In general, for arbitrary choice of coordinate axes, there are six independent stress
components and six rate of strain components and each of the stress component is a linear
function of the rate of strain component or vice versa. To arrive at the final form of stress
rate of strain relation, it is convenient to start with the principal axes of the two quadrics
at a point, because of isotropy the principal axes of the stress quadric coincide with those
of rate of strain quadric at every point in the continuum.

It is clear that referred to the principal axes  ij  0 and eij  0 , when i  j . In view of

the first and third assumptions the non-zero components of the stress tensor  11 ,  22 ,  33

are related to the non-zero components of the rate of strain tensor e11 , e22 , e33 in the
following manner:
 11   p  a11e11  a12 e22  a13 e33
 22   p  a21e11  a22 e22  a23 e33 (1.46)
 33   p  a31e11  a32 e22  a33 e33
where the aij are constants to be determined.

However, we observe that the assumption of isotropy implies that any permutation of the
e ' s must effect the same permutation of  ' s .
Now, permute that e11 , e22 , e33 to e22 , e33 , e11 (rotation of axes) and rearrange to obtain

 22   p  a13 e11  a11e22  a12 e33


 33   p  a23 e11  a21e22  a22 e33 (1.47)
 11   p  a33 e11  a31e22  a32 e33
The left hand of these equations has been obtained by making the same permutation on
the  ' s .
Also, by interchanging the axes 1 and 2, the first relation of the set (1.46) becomes
 22   p  a11e22  a12 e11  a13e33 (1.48)
Comparing the two sets of equations (1.46) and (1.47) and the equation (1.48), we find
a12  a21  a31  a13  a23  a32   (say )
(1.49)
and a11  a22  a33    2 ( say )
where  and  are, for the moment, numbers whose physical meaning have to be
obtained. The equations (1.46), in view of (1.49), become
 11   p  2  e11    e11  e22  e33 
 22   p  2  e22    e11  e22  e33 
 33   p  2  e33    e11  e22  e33 

Now these three equations plus the six equations implicit in the fact that  ij  eij  0 ,

when i  j , may all be combined in to a single tensor equation

22
 ij   pij  2  eij   ekk ij (1.50)

Since this is a tensor equation, it holds true regardless of the axes chosen and hence will
be taken as the required relationship between the stress components and rate of strain
components for arbitrary choice of the coordinate axes. It remains now to determine the
physical meaning of the constants  and  .

3.2. Interpretation of the constants  and 


Form equation (1.45) and (1.50), we conclude that the relationship between the
components of the viscous stress tensor and those of rate of strain tensor is given by
 ij  2  eij   ekk  ij (1.51)
Consider a state of simple shearing motion, as in fig.1.1 described by the velocity field

V :  v1  x2  , 0, 0  (1.52)
For this all the eij are zero except
1 dv1
e12  e21  (1.53)
2 dx2
Thus, form equation (1.51)
dv1
 12   21   (1.54)
dx2
And all the other viscous stresses are zero. Equation (1.54) is the same as equation (1.3)
and, therefore,  is the coefficient of viscosity.
The interpretation of the second constant  requires further discussion, but we note that
in an incompressible fluid it does not play any part, because in such a case from equation
(1.17),   ekk  0 and therefore the term containing  in equation (1.50) disappears

altogether. Thus  is important only in the case of compressible fluid.

3.3. Stoke’s hypothesis


Equation (1.51) can be written as
2  2 
 ij    ekk  ij  2  eij       ekk  ij (1.55)
3 3  
Clearly
 ii   3  2  eii (1.56)

 ii / 3  2 
or       k ( say) (1.57)
eii  3 
The bulk viscosity is defined as the ratio of the mean normal viscous stress to the rate of
volumetric strain in a state of pure dilation. Hence K is the bulk viscosity.
Stokes assumed that K=0, i.e.,
2
  (1.58)
3
23
giving  in terms of  . It can be easily seen from equation (1.50) that the Stoke’s
hypothesis is equivalent to the assumption that the pressure is the mean of the normal
stresses ( ii / 3   p).
The kinetic theory of gases has given some support in favour of taking K=0 for ideal
monatomic gases. However, for fluids in general, there is no justification for taking K=0.
Inspite of this drawback we will continue with the relation (1.58), since the equations of
motion which result from the substitution of (1.58), as we shall will see later, subjected to
a large number of problems whose results have been verified experimentally.
Thus from equation (1.55), in view of (1.58), the constitute equation for an isotropic
2
Newtonian fluid is  ij  2  eij   ekk  ij (1.59)
3
This is the required relationship between viscous stress tensor and rate of strain tensor
and is known as Stoke’s law of friction. In fact it is a generalization of Newton’s law of
viscosity, which is in one dimension, to the three dimensional space.

3.4. Exercises

1) Derive relation between stress and Rate of strain component.


2) Explain the concept of Stokes hypothesis.

3.5. Keywords
Stokes Law, Stokes Hypothesis.

3.6. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House,


New Delhi (1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi
(1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2,
Oxford University Press (1970).
6. Bansal. J. L., Viscous Fluid Dynamics, Oxford and IBH publishing Co. Pvt.
Ltd., New Delhi (1977).

24
UNIT-4
Thermal Conductivity,
Generalised law of Heat Conduction

Structure:
4.0. Objectives
4.1. Thermal Conductivity
4.2. Generalized law of heat conduction
4.3. Summery
4.4. Exercises
4.5. Keywords
4.6. References

4.0. Objective
At the end of reading this unit students are able to understand
a) Stoke’s law of friction
b) Stoke’s Hypothesis
4.1. Thermal conductivity
Temperature differences in a fluid, in the course of time, are reduced by heat flowing
from regions of higher temperature to those of lower temperature. Knowledge of the laws
governing the process of heat transfer is of great importance throughout the field of fluid
dynamics. There are three basic modes of heat transfer viz., conduction, convection and
radiation. Heat radiation is negligible unless the temperature is very high and we shall not
discuss it in this book. The transfer of heat by convection depends on the velocity field
and will be discussed in Chapter 2 in connection with the equation of energy. Only the
law governing the heat conduction will be taken up in this section.

Fig 1.9. Heat conduction between parallel layers of a fluid

Two parallel layers of fluid, at a distance d apart, are kept at different temperatures T1

and T2 (one of the layers may be a solid surface). Fourier noticed that a flow of heat is set
up through the layer such that the quantity of heat q transferred through unit area in unit

25
time is directly proportional to the difference of temperature between the layers and
inversely proportional to the distance d. Thus he found
T1  T2
qk (1.60)
d
where k is the constant of proportionality and is known as the coefficient of thermal
conductivity. If the distance d between the two layers of fluid is infinitesimal the above
law can be written in the differential form as
dT
q  k ,(Fourier’s law of heat conduction), (1.61)
dy
where the negative sign has been taken because the heat flows in the direction of
decreasing temperature. The dimensions of the coefficient of thermal conductivity can be
determined as follows:

Heat flux Btu / ft 2 hr Btu


k  0 
Temperature gradient F / ft ft  hr  0 F

Some typical values of k are given below for gases and liquids at 32 0 F and under
atmospheric pressure.

Gases Liquids

Air 0.0140 Water 0.319

Hydrogen 0.1003 Mercury 4.74

Carbon dioxide 0.0083 Glycerine 0.140

It may be pointed out here that the thermal conductivity k is not necessarily a constant
but, in fact, depends on temperature and pressure both. The example considered in fig.1.9
constitutes a particularly simple case of heat conduction. The generalization to the case of
three dimensional flow will now be taken.

4.2. Generalized law of heat conduction

In an isotropic medium in which the temperature varies in all three directions we may
write an equation like equation (1.61) for each of the coordinate directions;
T T T
q1  k , q2  k , q3  k (1.62)
x1 x2 x3
These three relations are the components of single vector equation, in Cartesian tensor
notation

26
T
qi  k (1.63)
xi

which is the three dimensional form of Fourier’s law. It states that the heat flux vector q
is proportional to the temperature gradient T and is oppositely directed.

It may be pointed out that there is a striking similarity between the equation (1.3),
Newton’s law of viscosity, and equations (1.61), Fourier’s law of heat conduction for one
dimensional case. But in more complicated situation in which the temperature and
velocity vary in all the three directions, however , we find that equation (1.63) for the
heat flux is simpler than equation (1.59) for the momentum flux. This difference in form
is mainly due to the fact that temperature is a scalar quantity while velocity is a vector
quantity. Therefore the gradient of velocity will be a tensor. We can thus anticipate that
the problems of velocity distribution will not be mathematically analogous to the
temperature distribution except in certain simple flows.

As we shall see later, the effect of conductivity on the temperature field is determined by
the ratio of k to the product of density  and specific heat c p rather than by k alone.

This ratio is known as the thermal diffusivity and is usually denoted by a . Thus
k
a (Thermal diffusivity) (1.64)
 cp

4.3. Summary:
In this Block the authors explains the basic concepts required in the development of the
theory of viscous flow and the constitutive equation for an isotropic Newtonian fluid is
derived in Cartesian form. This chapter also includes the basic concepts of thermal
conductivity and the generalized law of heat conduction.

4.4. Exercises

1. Obtain the relationship between the stress and strain for a viscous fluid in usual
notation and interpret the physical constants
2. If the velocity distribution is q  Ax2 y iˆ  By 2 zt ˆj  Czt 2 kˆ A, B, C, are constants,
find the acceleration and velocity components.
3. Determine the acceleration at the point (2,1,3) at t =0.5 sec if
u  yz  t , v  xz  t , w  xy.
4. The velocity components of a flow in cylindrical polar coordinates are
 r z cos , rz sin  , z t  determine
2 2
the components of acceleration of a fluid
particle.
5. Write short note on thermal conductivity and generalized law of heat conduction.

27
4.5. Keywords

Thermal Conductivity, Heat Conduction.

4.6. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New Delhi
(1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi (1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2, Oxford
University Press (1970).
6. Schlichting. H., Boundary Layer Theory, McGraw-Hill (1968).
7. Streeter .V. L. (Ed)., Hand Book of Fluid Dynamics, McGraw2-Hill (1961).
8. Yuan. S.W: Foundational of Fluid Mechanics, Prentice. Hall of India Pvt. Ltd., New
Delhi (1969).
9. Bansal. J. L., Viscous Fluid Dynamics, Oxford and IBH publishing Co. Pvt. Ltd.,
New Delhi (1977).
10. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).

28
Block – 2
Fundamental Equations of
the Flow of Viscous Fluids

29
UNIT – 5
Equation of state, Equation of continuity,
Conservation of Mass.

Structure:
5.0. Objectives
5.1. Introduction
5.2. Equation of state
5.3. Equation of continuity – Conservation of mass
5.4. Exercises
5.5. Keywords
5.6. References

5.0. Objectives:

The Block 2 deals with the fundamental equations, which govern the motion of a viscous
compressible fluid. At the end of reading this unit students are able to understand
a) Equation of state
b) Equation of continuity – Conservation of mass

5.1. Introduction
The fundamental equations of the flow of viscous compressible fluids are
(i) Equation of state, (one)
(ii) Equation of continuity, (one)
(iii) Equation of motion, (three) and
(iv) Equation of energy. (one)
These equations are mathematical expressions of basic physical concepts. These are six
in number and therefore determine the six unknowns of the fluid motion viz., the three
components of velocity (v1, v2, v3 ) , the temperature T, the pressure p and the density  ,
which are functions of both space coordinates and time. The deviation of these equations
will be the subject matter of this chapter.

5.2. Equation of state

Variables that depend only upon the state of a system are called variables of state. The
variables of state are the pressure p , the density  and the temperature T . It is an
experimental fact that a relationship between these three thermodynamic variables exists
and can be written as
F ( p,  , T )  0 (2.1)
which is commonly called the ‘Equation of state’. For gases at reasonably high
temperature or low pressure the above functional relationship can always be written as
p
 1  B(T )   C (T )  2  ...... (2.2)
 RT
30
where B(T ) , C (T ) and so on are the functions of temperature only and R is the gas
constant.
As a first approximation equation (2.2) may be written as
p   RT . (2.3)
This is known as the equation of state of a perfect gas. This equation has also been
derived from the kinetic theory of gases when the interamolecular action among
molecules of a gas is neglected, the distance between molecules is assumed to be much
greater than the diameter of the molecules and the volume of the molecules is considered
negligible in comparison of the total apparent volume of the gas. In this book the
equation of state of a viscous compressible fluid will be taken as the equation of state of a
perfect gas. When the fluid is considered as incompressible the equation of state is
simply.
  const. (2.4)
5.3. Equation of continuity – Conservation of mass

This equation expresses that the rate of generation of mass within a given volume is
entirely due to the net inflow of mass through the surface enclosing the given volume
(assuming that there are no internal sources). It amounts to the basic physical law that the
matter is conserved; it is neither being created nor destroyed.

Fig 2.1. A closed surface S on closing an arbitrary fixed volume V in the region of a moving fluid.

Let us consider a closed surface S , enclosing a fixed (arbitrarily chosen) volume V in


the region occupied by the moving fluid. If n j is the unit vector in the direction of the

outward drawn normal to the element dS of the surface S and v j be the velocity of the


fluid at that point, then the inward normal velocity is v j n j . 

31
Thus the mass of the fluid entering, per unit time, through the element dS is equal to

  v j n j  dS. From the above it follows that the mass of the fluid entering, per unit time,

in the controlled surface S is    v j n j dS . (2.5)


S

Also the mass of the fluid within S is 


V
dV . (2.6)

Therefore, the rate at which the enclosed mass increases is simply


 
t V  t
 dV , or dV , (2.7)
V

the differentiation and integration being interchangeable because a fixed volume is


considered.
Now, by the conservation of mass; the expression in (2.5) should be the same as in (2.7).
Hence

 t
V
dV     v j n j dS
S
(2.8)

Applying Gauss’ theorem, we get


 
 t dV   
x j
  v j  dV ,
V V

 
  

or   t 
x j
  v j  dV  0. (2.9)
V  

Since, V is an arbitrary chosen volume, we deduce that
 

t x j
  v j   0, (2.10)

which is the required equation of continuity in the Cartesian tensor notations. In vector

notation it can be written as  div( V )  0. (2.11)
t
In case of steady compressible flow the equation of continuity reduces to

x j
  v j   0, (2.12a)

or div( V )  0. (2.12b)
If the fluid is considered as incompressible it further reduces to
v j
 0, (2.13a)
x j

or div(V )  0. (2.13b)

32
5.4. Exercises

1. Derive the mass conservation equation in the form  div   v   0,
t
2. With usual notation derive equation of state of a perfect gas.

For more Problems see end of the Block, page number 46

5.5. Keywords

State equation, Conservation of mass.

5.6. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New Delhi
(1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi (1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2, Oxford
University Press (1970).
6. Schlichting. H., Boundary Layer Theory, McGraw-Hill (1968).
7. Streeter .V. L. (Ed)., Hand Book of Fluid Dynamics, McGraw2-Hill (1961).
8. Yuan. S.W: Foundational of Fluid Mechanics, Prentice. Hall of India Pvt. Ltd., New
Delhi (1969).
9. Bansal. J. L., Viscous Fluid Dynamics, Oxford and IBH publishing Co. Pvt. Ltd.,
New Delhi (1977).
10. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).

33
UNIT – 6
Equation of Motion: Conservation of Momentum.

Structure:
6.0. Objectives
6.1. Equations of motion (Navier-Stokes’ equations) – Conservation of momentum
6.2. Exercises
6.3. Keywords
6.4. References

6.0. Objectives:

At the end of reading this unit students are able to understand Navier-Stokes’ equations
for Conservation of momentum.

6.1. Equations of motion (Navier-Stokes’ equations) – Conservation of


momentum

The equations of motion are derived from Newton`s second law of motion which states
that “ Rate of change of linear momentum = Total surface”.
Let us consider a closed surfaces S , as earlier (see Fig 2.1) enclosing a volume V in the
region occupied by the moving fluid. The rate at which momentum entering the element

dS is vi    dSv j n j  . Therefore, the rate at which the momentum enters the controlled

surface S is

  vi    v j n j dS  . (2.14)
S

Also the rate at which the momentum increases in the enclosed volume V is

t V
vi dV . (2.15)

Hence, from (2.14) and (2.15), the rate of change of linear momentum is given by

vi dV   vi   v j n j  dS .
t V
(2.16)
S

In fluid motion it is necessary to consider the following two classes of forces, (i) forces
acting throughout the mass of the body of fluid, such as gravitational forces, known as
body forces, and (ii) forces acting on the boundary, the fluid stresses, and are known as
surface stresses. If f i denotes the body forces per unit mass and Pi the forces on the
boundary per unit area, the equation of motion can be written as

34

t V
vi dV   v   v n  dS.
S
i j j

rate at which the momentum rate of outflow of momentum


increases in the enclosed volume through the controlled surface

   fi dV   PdS.
i
V S

body forces acting on the surface forces acting on the (2.17)


enclosed volume controlled surface
where the stress vector Pi is given by

Pi   ij n j , (2.18)

and  ij   p  ij   ij . (1.45)

Substituting (2.18) and (1.45) in equation (2.17) and changing the surface integrals into
volume integrals by Gauss’ theorem and noting that V is an arbitrary chosen volume, we
get the equation of motion as
  p 
  vi     vi v j    fi   ij (2.19)
t x j xi x j

or , using the equation of continuity (2.10),


 vi vi  p vij
  vj    fi   . (2.20)
 t x j  xi x j

It should be kept in mind that equation (2.20) is valid for any continuous fluid medium.
In order to use these equations to determine velocity distribution, however, we must
insert expressions for the viscous stresses in terms of velocity gradients and fluid
properties. For isotropic Newtonian fluid these expressions are given by the constitutive
equation
2
 ij  2 eij   ekk  ij (1.59)
3

1  vi v j 
where eij     (1.6)
2  x j xi 

Substituting (1.59), with (1.6), in equation (2.20), we finally get


 vi vi  p    v v j 2 v 
  vj    fi      i    ij k   (2.21)
 t x j  xi x j   jx xi 3 xk  
These are known as Navier-Stokes equations for the motion of a viscous compressible
fluid and are three in number. In case of incompressible fluid flow the equation of
continuity is

35
vk
 0, (2.13a)
xk
and if  is also regarded as constant, the equation (2.21) can be further simplified to

 vi vi  p  2vi


  vj    fi   , (2.22)
 t x j  xi x j x j

keeping in view that

  v j    v j 
     0 (2.23)
x j  xi  xi  x j 
Equation (2.22) in vector notation can be written as

DV
   F  p   2V . (2.24)
Dt
where
D 
  (V .) (2.25)
Dt t
and is known as the ‘material derivative’.

6.2. Exercises

1. Derive the N-S equation for a viscous incompressible fluid in the form;
 q 
   qq   p   g   2 q explain the meaning of each term.
 t 

For more Problems see end of the Block, page number 46

6.3. Keywords
Conservation of momentum.

6.4. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New Delhi
(1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi (1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2, Oxford
University Press (1970).
6. Streeter .V. L. (Ed)., Hand Book of Fluid Dynamics, McGraw2-Hill (1961).
7. Yuan. S.W: Foundational of Fluid Mechanics, Prentice. Hall of India Pvt. Ltd., New
Delhi (1969).
8. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).

36
UNIT – 7
Equation of Energy: Conservation of Energy.

Structure:
7.0. Objectives
7.1. Equations of energy – Conservation of energy
7.2. Summary of fundamental equations
7.3. Exercises
7.4. Keywords
7.5. References

7.0. Objectives:

At the end of reading this unit students are able to understand Equations for Conservation
of energy.

7.1. Equation of energy-Conservation of energy

To obtain the energy equation we have to apply the law of conservation of energy which
requires that, the difference in the rate of supply of energy to a controlled surface S
enclosing a volume V in the region occupied by a moving fluid and the rate at which the
energy goes out through S must be equal to the net rate of increase of energy in the
enclosed volume V. Thus the following equation in Cartesian tensors is obtained:
Q
 t dV   vi  ij n j  dS   Et  v j n j dS
v S S

rate at which heat rate at which heat is rate of energy loss


is produced by produced by the work by heat convection
external agencies of the surface stresses


t V
  q j n j dS   Et dV
S

rate of energy loss rate of increaseof (2.26)


by heat conduction energy in the volume
Where
Et  Total energy of thesystem per unit mass
1
 vi v j  K  I , (2.27)
2
( K .E ) ( P.E ) ( I .E )

and the heat flux vector q j is given by the generalized heat conduction law

T
q j  k (1.63)
xi

37
Substituting (1.63) in (2.26) and changing the surface integrals by Gauss’ theorem and
noting that V is an arbitrary chosen volume, we get the equation of energy as

Q   T  
 vi ij   Et v j    k   E t    0 (2.28)
t xi x j x j  xi  t
The following relations will help us to simplify the equation (2.28).

(i) Combining the third and fifth terms of equation (2.28), we get
   

Et v j    E t    Et     v j     E s  v j Et    DE t , (2.29)
x j t  t x j   t x j  Dt

where the equation of continuity (2.10) has been taken into consideration and
D  
  vj , (2.30)
Dt t x j

is the material derivative.


The right hand side of (2.29), with the help of (2.27), can be written as

DE t  Dvi DI K 
    vi   vj , (2.31)
Dt  Dt Dt x j 
 
Since
K
 0 (Potential energy is independent of time; it depends only on space
t
coordinates) (2.32)
(ii) We have
 ij   p ij   ij (1.45)

The equation of motion (2.20), in view of (1.45), is




Dvi
 f i   ij  (2.20)
Dt x j

K
But fi   (since K is the P.E.) (2.33)
xi


 ij     Dvi  K  (2.34)
x j  Dt xi 
Now the second term of equation (2.28), in view of (1.45) and (2.34), can be written as


vi ij   vi  Dvi  K    p ij   ij  vi
x j  Dt xi  x j

 Dv K  v
 vi  i    p i   , (2.35)
 Dt xi  xi

38
vi
where    ij
x j

 1  v
 2  eij  ekk  ij  i
 3  x j


 v v j  2 v k   vi
   i     , (2.36)
  3 x
 x j xi 
ij
  k 
 x j

is the heat generated due to frictional forces and is usually known as ‘dissipation
function’.
Thus the equation of energy (2.28), with the help of equations (2.29), (2.31) and (2.35),
can be simplified to

Q Dv K v Dv DI K   T 
 vi i  vi  p i    vi i    vi  k   0 (2.37)
t Dt xi xi Dt Dt xi x j  xi 

Q v DI   T 
or,  p i    k     0 . (2.38)
t x i Dt x j  x i 
Further, making use of the equation of continuity (2.10),
vi p Dp D1
p   p   (2.39)
xi  Dt Dt   
Therefore, the energy equation (2.38) is written as
 DI D  1   Q   T 
 p     k   (2.40)
 Dt Dt     t x j  xi 
For most fluid problems, it is convenient to have the energy equation in terms of fluid
temperature and heat capacity rather than the internal energy. For a perfect gas, we have
the following relations:
p  RT , (2.3)

c p  c0  R , (2.41)

I  c0T   c p  R  T  cPT 
p
(2.42)

where c0 and c p are specific heats at constant pressure and at constant volume

respectively.
Hence for a perfect gas the terms on the left hand side of equation (2.40) may be written
as follows:
 DI D  1  D D  p D  1 
 p       c pT      p  
 Dt Dt      Dt Dt    Dt    

39

D
c pT   Dp , (2.43) and
Dt Dt
the energy equation (2.40) may , then , be written as
Dp Q   T 
 c pT  
D
  k   . (2.44)
Dt Dt t xi  xi 

vi
In case of an in compressible fluid,  0 ; I  c0T , the equation of energy (2.38), with
xi
constant viscosity and heat conductivity, becomes

DT Q   T 
cv  k     , (2.45)
Dt t xi  xi 
 v v j  v
where     i   i

(2.46)
 x j xi  x j
7.2. Summary of the fundamental equations

For a viscous compressible fluid, we have obtained the following equations in Cartesian
tensors:
Equation of state: - p  RT , (2.3)
 
Equation of conductivity:-  v j   0, (2.10)
t x j

 v v  p  
  vi v j 2 vk 
Equation of motion :-   i  v j i   f i   
     
ij 
 t x j  xi x j
   x j xi 3 xk
 

(2.21)
and are three in number.
Equation of energy :-
 T 

D
c pT   Dp  Q    k    (2.44)
Dt Dt t xi  xi 
D  
where   vj (2.30)
Dt t x j


 vi v j  2 vk   vi
and      
 3 x
  (2.36)
 xi  x j
ij

 x j  k 

As has already been pointed out in Chapter 1 that the coefficient of viscosity  and
thermal conductivity k are not constants but depend on temperature and this dependence
is much significant in the case of flow compressible fluids. Hence we have eight
unknowns, viz., vi  i  1, 2,3 , p,  , T ,  and k instead of six and therefore two more

equations are required. These two equations, in general, are


   T  and k  k T  (2.47)

40
The variation of viscosity  with absolute temperature T, in the case of air, may be
expressed quite accurately by Sutherland’s formula,
3
  T  2 T0  S1
  , (2.48)
0  T0  T  S1

where  0 denotes the viscosity at a reference temperature T0 and S1 is a constant, for air

S1  110  K .
In view of the complicated nature of the formula (2.48) it is assumed , in theoretical
calculations, that

 T 
   with 0.5    1 (2.49)
0  T0 
It has been found that at high temperature the relation (2.48) can be well approximated by
(2.49) when  the values of lie between 0.5 and 0.75 and at low temperature   1
appears to be adequate.
The dependence of thermal conductivity on temperature is of similar nature as that of
viscosity. In fact if we form a non-dimensional quantity
cp
Pr  , (2.50)
k
Then Pr , which is called the Prandtl number, is found to be constant for air even at large
temperature difference. The specific heat c p is assumed to be constant with a satisfactory

degree of approximation. Although the value of Pr is 0.72 for air, in theoretical


calculations it is generally taken as 1 since it leads to useful analytical simplifications.
Having given the summary of the equations which govern the motion of a viscous
compressible fluid we now turn to the case of incompressible flow, where it may be
assumed that the fluid properties such as density  , viscosity  and conductivity k are
nearly constant. An immediate consequence of assuming the fluid properties to be
constant is that the number of unknowns is reduced, and the equation of state becomes
redundant. We have only five unknowns, viz., vi i  1,2,3, p and T , which are determined
from the following five equations:
Viscous incompressible flow
v j
Equation of continuity :- 0 (2.13a)
x j

 v v  p  2 vi
Equation of motion :-   i  v j i   f i   (2.22)
 t x j  xi x j x j

and are three in number.

41
DT Q   T 
Equation of energy: - cv  k     , (2.45)
Dt t xi  xi 
D  
where   vj (2.30)
Dt t x j

 v v  v
    i   i .
j
and (2.46)
  
x
 j x i  x j

One of the essential difference between the compressible flow and incompressible flow is
that, in compressible flow the equations of motion and energy are coupled whereas in an
incompressible flow, with constant fluid properties, they are uncoupled and therefore in
this case the equations of continuity and equations of motions are first solved for the three
velocity components vi and the pressure p and thereafter the equation of energy is solved
for the temperature field.
The solution of the above equations become fully determined physically when the
boundary and initial conditions are specified. Although the initial conditions will be
discussed in specific flow problems the boundary conditions are those required by
geometrical considerations, together with the no-slip condition which states that on a wall
the tangential component of relative velocity must be zero. Finally, some conditions must
be imposed on the temperature on the boundary and will be taken up in specific flow
problems.

7.3. Exercises

1. Derive the equation of energy for an incompressible viscous fluid in the form
DT
 cv  K 2T  Q   .
Dt

For more Problems see end of the Block, page number 46

7.4. Keywords
Conservation of Energy.

7.5. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New Delhi
(1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi (1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2, Oxford
University Press (1970).
6. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).

42
UNIT – 8
Vorticity and Circulation in a Viscous Incompressible
Fluid Motion

Structure:
8.0. Objectives
8.1. Vorticity and circulation in a viscous incompressible fluid motion
8.2. Summary
8.3. Exercises
8.4. Keywords
8.5. References

8.0. Objectives:

At the end of reading this unit students are able to understand Vorticity and circulation in
a viscous incompressible fluid motion. Predominantly (a) Vorticity transport equation and
(b) Circulation

8.1. Vorticity and circulation in a viscous incompressible fluid motion

(a) Vorticity transport equation


The Navier-stokes equations for a viscous incompressible fluid motion may be
interpreted as the Vorticity transport equations if we assume that the external forces are
conservative, i.e. they can be derived from a force potential V f such that

F  V f (2.51)

Using Lagrange’s vector identity

   1 2   
 V   V    V  V  (2.52)
  2 

 
where     V (2.53)

is the velocity vector, the Navier-stokes equations (2.24) may be written as



V   p 1   
 V       V 2  V f   v 2
V (2.54)
t  2 

Taking the curl of both sides of equation (2.54) and keeping in view that the curl of a
gradient is identically zero, we obtain

43

   
    V     v 2  . (2.55)
t  

Using the vector identity

     
  V    . V  V .   V .   .V

   
 . V  V .  , (2.56)

.  0 and .  0 , the equation (2.55) is finally written as


Since V

D
Dt
 
 . V  v 2  (2.57)

This equation is referred to as vorticity transport equation. The first term on the right

hand side of (2.57) represents the rate at which  varies for a given particle when the
vortex lines move with the fluid, the strengths of the vortices remaining constant. The
second term represents the rate of dissipation of vorticity through friction.

In case of two-dimensional motion, if

V  ux, y i  vx, y  j, (2.58)

then it can be easily verified that

.V  0 (2.59)

and, therefore, the equation (2.57) reduces to

D
 v 2  (2.60)
Dt

This equation is of the same form as the equation of heat conduction in a liquid (c.f.eqn.
2.45), without heat addition and neglecting the frictional heat. Thus we may conclude that
the vorticity diffuses through a viscous liquid in much the same manner as heat does, and
therefore can assert that the vorticity cannot originate within the interior of a viscous fluid
but must be diffused from the boundaries into the fluid.

44
(b) Circulation
We now introduce an important concept of fluid dynamics, the circulation, which is
defined as the line integral of the velocity along a closed curve. Thus

   V.dr (2.61)
C

We would now like to know the time rate of change of circulation if the closed curve,
drawn in a viscous incompressible fluid, moves with the fluid. We have

D  DV

 
Dt C 
.dr  V .
D 

dr 

  (2.62)
 Dt Dt 

But
D
Dt
 
dr is the difference of the velocities at the end of the line vector dr , so that

(c.f.eqn.1.4)
D
Dt
  
dr  dr. V  (2.63)

Also, if the external forces are conservative, the equation of motion (2.24) can be written

DV p 
as   V f   v curl  (2.64)
Dt p 

where the help of the vector identity

   
Curl       V   .V   2 V

  2 V , since  V  0, (2.65)

has been taken

Substituting (2.63) and (2.64) in equation (2.62), we find

D p 
Dt 
1
      Vr  .V 2  dr  v  curl  dr
2 
 
C C

p 
1
   d   Vr  .V 2  dr  v  curl  dr
 2 
 
C C

 v   curl  
C
dr (2.66)

45
Thus, the rate of change of circulation in a closed curve, drawn in a viscous
incompressible fluid, moving with the fluid depends on the kinematic viscosity and on
the space rate of change of the vorticity components at the contour.

If v  0 i.e., if the fluid is taken as in viscid, we get the well-known Kelvin’s circulation
theorem, viz., the circulation round any closed curve moving with the fluid does not
change with time, provided the fluid is in viscid , the field of force is conservative and
pressure is a single valued function of density only.

8.2. Summary

In this chapter the author explains the fundamental equations which govern the motion of
a viscous compressible fluid. The derivations of the equations have been made simple
and concise with the help of Cartesian tensors. The governing equations for compressible
and incompressible fluids motion in Cartesian, polar and spherical coordinates are given
in tabular form for ready reference.

8.3. Exercise

1. With usual notation prove the following for the vorticity transport equation
DT
 cv  K 2T  Q   and hence derive the expression for circulation in a viscous
Dt
incompressible fluid motion.

More Problems covering entire block

1. The particle of a fluid move symmetrically in space with regard to a fixed centre
    2
prove that the equation of continuity is
t
u 
p r 2 r
 
r u  0 where u is the

velocity at the distance r.


2. A mass of a fluid moves in such a way that each particle describes a circle in one
    w
plane about a fixed axis show that the equation of continuity is  0
t 
where w is the angular velocity of a particle whose azimuthal angle is θ at time t.
3. Derive Eulers equation of motion in cylindrical coordinates.
4. Prove that, if the motion of an ideal fluid, for which density is a function of pressure
p only, is steady and external forces are conservative then there exist a family of
surfaces which contains the stream line and vertex lines.
 a2   a2 
5. The velocity components ur  r ,   V 1  2  cos  , u  r ,   u 1  2  sin  ,
 r   r 
satisfy the equation of motion for a steady two dimensional inviscid incompressible
flow. Find the pressure associated with this velocity field u and v are constants.
6. A steady inviscid incompressible fluid flow has a velocity field
u  f x, v   fy, w  0 where f is a constant . Derive an expression for pressure field
p(x,y,z) if the pressure p(0,0,0)=p0 and F= -g i z.

46
7. Show that if the velocity field
  x  , v  x, y   2Bxy x  w  x, y   0 satisfies
2 2
u  B x2  y 2 2
 y2 2
 y2 the
equation of motion for inviscid incompressible flow then determine the pressure
associated with this velocity field, B is being constant.
8. Show that for an incompressible steady flow with constant viscosity the velocity
U h2  p  y  y 
components u  y   y     1   , v  w  0 satisfy the equation of
h 2  x  h  h 
p
motion when the body forces are neglected, , h,U are cons tan ts and p  p  x  .
x
9. Define circulation. Derive the time rate of change of circulation of a closed curve
drawn in a viscous incompressible fluid , moving with the fluid.
10. Write the Navier Stokes equations in Cartesian co ordinates. Simplify the equations
when (i) fluid is incompressible and dynamic viscosity is constant. (ii) the fluid is
incompressible and viscous effects are neglisible.
11. Derive the equation of energy with constant viscosity and heat conductivity of the
fluid.
12. Derive the equation of continuity in usual form.

8.4. Keywords:

Transport Equation, Vorticity

8.5. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New Delhi
(1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi (1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2, Oxford
University Press (1970).
6. Schlichting. H., Boundary Layer Theory, McGraw-Hill (1968).
7. Streeter .V. L. (Ed)., Hand Book of Fluid Dynamics, McGraw2-Hill (1961).
8. Yuan. S.W: Foundational of Fluid Mechanics, Prentice. Hall of India Pvt. Ltd., New
Delhi (1969).
9. Bansal. J. L., Viscous Fluid Dynamics, Oxford and IBH publishing Co. Pvt. Ltd.,
New Delhi (1977).
10. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).

47
Block – 3
Dynamical Similarity,
Inspection and
Dimensional Analysis

48
UNIT – 9
Dynamical Similarity and Inspection Analysis

Structure:
9.0. Objectives
9.1. Introduction
9.2. Dynamical Similarity (Reynolds’ Law)
9.3. Inspection Analysis
9.4. Exercises
9.5. Keywords
9.6. References

9.0. Objectives:

At the end of reading this unit students are able to understand Reynolds law.

9.1. Introduction

Although we have developed the fundamental equations governing the flow of a viscous
compressible fluid, there are no known general methods to solve these equations. The
main reason of the absence of such a general method is the non-linear character of the
governing equations. Only in few particular cases and that too under restricted conditions
exact solutions of these equations, for all ranges of viscosity, exist and will be taken up in
the next chapter. However, attempts have been made to simplify these equations for two
extreme cases of viscosity ,very large and very small, and we have well established
theories for these cases which are known as ‘Theory of slow motion’ and ‘Theory of
boundary layers’, but the cases of moderate viscosities cannot be interpolated from these
two theories. Even in these two extreme cases we find great mathematical difficulties
and therefore most of the research on the behavior of viscous fluids has been carried out
by experiments. In experiments, generally a prototype (geometrically similar but reduced
in size) of the actual body is taken and the flow around it is investigated in the wind
tunnel in order to reduce the cost of the full scale test and to have better control over the
conditions. This always raises the question of ‘dynamic similarity’ of fluid motions
because, naturally, we would like to know that how far the results obtained on the
prototype can be considered the same as on the full scale body.

9.2. Dynamical similarity (Reynolds’ law)

Two fluid motions are said to be ‘dynamically similar’ if with geometrically similar
boundaries the flow patterns are geometrically similar. We now discuss the conditions
under which the fluid motions are dynamically similar. In other words we have to find
49
out those parameters which characterise a flow problem. There are two methods for
finding out these parameters (i) inspection analysis, and (ii) dimensional analysis. In the
inspection analysis, we reduce the fundamental equations to a non-dimensional form and
obtain the non-dimensional parameters from the resulting equations. In dimensional
analysis we form non-dimensional parameters from the physical quantities occurring in a
problem, even when the knowledge of the governing equations is missing. We will now
discuss these with particular reference to the flow of a viscous compressible fluid.

9.3. Inspection analysis

The fundamental equations governing the flow of a viscous compressible fluid (c.f.g2.6)
are
p  RT (2.3)
 
 ( v j )  0 (2.10)
t x j

 vi v  p  
  vi v j 2 v k 
  vj i    fi   
      ij 

(2.21)
 t x j xi x j
    x j xi 3 xk
 

Dp   T 

D
Dt
 c pT    k  ,
Dt xi  xi 
(2.44)

D  
where  vj , (2.30)
Dt t x j


 vi v j 2 vk  v
and         ij   i , (2.36)
  x j
 x j xi 3 xk
 
and the heat produced by the external agencies is assumed to be zero.
We are now going to put the above equations in a non-dimensional form by setting
xi * vi * tU 0 *  * 
xi*  , vi  ,t  ,  ,  ,
L0 U0 L0 0 0

p f T k cp R 0T0
p*  , f i *  i , T *  , k *  , c *p  and R *  ,
p0 f0 T0 k0 c p0 p0

where the quantities with subscript ‘o’ are certain reference values associated with the
flow.
Thus the governing equations of a viscous compressible fluid in the non-dimensional
form are
p *   * R *T * (3.1)

 * 
t *
x j

 *  * v*j  0.  (3.2)

50
 v*i * vi
*  f 0 L0 * * p0 p*
*   v    f 
 t * j
x*j 2 i
0U 02 xi*
  U0

0   *  v* v*j 2 v* *  


    *  *   
* ij 
i k
(3.3)
0U 0 L x*j x x 3 x 
  j i k  

  * T *  0
*
D * *
c 
T 
p0
U 02 Dp*
+
k0
 k  +
U 02
*,
 0U 02 c p 0 T0 Dt * c*  0U 0 L xi*  xi*   0U 0 L0 c p 0 T0
p
Dt *

(3.4)
D  
where  *  v*j * , (3.5)
Dt *
t x j

 v * v *j 2 v *  v *j
* 
* 
   * * *
i k
  ij  * , (3.6)
 x j xi 3 x k  x
*
 j

and  ij*   ij
It can be easily recognized that the solutions of the above equations depend on the
following five non-dimensional groups:
0 f 0 L0 p0 U 02 k0
, , , , and ,
 0U 0 L0 U0 2
 0U 02 c p T0
0
c p0  0U 0 L0

which can be expressed in terms of the following, commonly used , dimensionless


numbers :
0U 0 L0
(i) Re  (Reynolds’ number)
0

U 02 U2
(ii) Fr   o (Froude number, when the body force is the gravitational
f 0 L0 g 0 L0
force , i.e., f 0  g 0 )

0
(iii) Ma  (Mach number) where c 0 is the velocity of sound at the reference
c0

 p0
values and is given by co2 
0
c p0
(iv)   (Ratio of specific heats)
c

c p0  0
(v) Pr  (Prandtl number)
k0
Sometimes in place of Mach number or ratio of specific heats another number known as
U 02
Eckert number is used, which is defined as: Ec  (Eckert number)
c p T0
0

51
Its relationship with Mach number and ratio of specific heats can be seen from the
following calculations: Equation of state can be written as
 1
p0
0

 RT0  c p0  c T0  c T
 p 00

 p0 U 02 c 2 U 02
Therefore, c02     1 c p T 0 and Ec   0     1 Ma 2 .
0 0 2
c p T0 c p0 T0 c0
0

Thus the five dimensionless groups, obtained above, in terms of the commonly used
dimensionless numbers are expressed as
0 1
 ,
 0U 0 L0 Re

f 0 L0
2

1
if f 0  g 0 ,
U0 Fr

p0 1
 ,
0U 0  Ma 2
2

U 02
    1 Ma 2  Ec ,
c p T0
0

k0 1
and  ,
c p 0  0U 0 L0 Pr Re

Hence, for the complete dynamical similarity of the flows of viscous compressible fluids
past geometrically similar bodies, when the body force is the gravitational force only, we
must have the same Reynolds number, same Froude number, same ratio of specific heats,
same Mach number and same Prandtl number. The Eckert number will be the same if the
ratio of the specific heats and the Mach number are the same.

9.4. Exercises

For Problems see end of the Block, page number 67

9.5. Keywords
Dynamical Similarity, Reynolds’ Law

9.6. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New Delhi
(1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi (1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).

52
UNIT – 10
Dimensional analysis and Buckingham Pai-Theorem

Structure:
10.0. Objectives
10.1. Dimensional Analysis
10.2. Buckingham  -theorem
10.3. Applications of Buckingham  -theorem
10.4. Exercises
10.5. Keywords
10.6. References

10.0. Objectives:

At the end of reading this unit students are able to understand Analysis of dimensions,
Buckingham  -theorem and its applications.

10.1. Dimensional Analysis

In g 3.3 we reduced the governing equations of a viscous compressible fluid to a non-


dimensional form and obtained the dimensionless parameters. An alternative method,
with which the non-dimensional parameters may be formed from the physical quantities
occurring in a flow problem is known as ‘dimensional analysis’. They are derived on the
basis of the dimensions in which each of the quantities involved in a phenomenon is
expressed, and hence , must not depend on the units chosen for the calculations.
In dynamics of viscous compressible fluids there are four fundamental units, viz., length,
mass, time and temperature in which the dimensions of all the physical quantities
occurring in such a flow problem can be expressed .We shall denote the dimensions of
these fundamental units by L , M  , t  and   respectively .We give below the
dimensions of some of the quantities which commonly occur in fluid dynamics.

Quantity Dimension Quantity Dimension


Length L Acceleration Lt 
2

Mass M  Density ML 


3

Time t  Force MLt  2

Temperature   Stress ML t 


1 2

Velocity Lt  1
Viscosity ML t 
1 1

Kinematic viscosity  L2 t 1  Thermal Conductivity MLt  


3 1

Specific heat L t
2 2
 1 

53
We shall now give an elementary proof of the Buckingham  -theorem which is the basic
theorem in dimensional analysis and shall see its application to the flow of viscous
compressible fluids.

10.2. Buckingham  -theorem

The  -theorem makes use of the following assumptions:


(i) It is always possible to select m independent fundamental units in a physical
phenomenon. (In dynamics of viscous compressible fluids m  4 ,i.e., mass, length, time
and temperature).
(ii) There exist n quantities, say Q1 , Q2 ,......., Qn involved in a physical phenomenon
whose dimensional formulae may be expressed in terms of m fundamental units.
(iii) There exist a functional relationship between the n dimensional quantities
Q1 , Q2 ,......., Qn , say  Q1 , Q2 ,......., Qn   0 ,and this equation is independent of the types

of units chosen and is dimensionally homogeneous.


The  -theorem may now be stated as follows;
An equation in physical variables, which is dimensionally homogeneous, can be reduced
to a relationship among a complete set * of dimensionless products. Or
In other words if Q1 , Q2 ,......., Qn be n physical quantities involved in a physical
phenomenon and if there are m independent fundamental units in this system then a
relation  Q1 , Q2 ,......., Qn   0 is equivalent to the relation f  1 ,  2 ,.......,  n1   0 ,
where  1 ,  2 ,.......,  n1 are the dimensionless quantities formed by the Qn ' s and r is the

rank of the dimensional matrix of the given physical quantities. The proof of the  -
theorem is based on the following theorem of matrix algebra on the solution of linear
algebraic equations:
If we have m homogeneous equations in n unknowns, then the number of independent
solutions is n-r, where r is the rank of the matrix of coefficients and any other solution
can be expressed as a linear combination of these linearly independent solutions. Further,
there will be only r independent equations in the set of equations. The proof of this
theorem can be found in any book on matrix algebra, we shall only explain the
application of it to the proof of  -theorem.
Let Q1 , Q2 ,......., Qn be n given physical quantities and let their dimensions be expressed in

terms of m fundamental units u1 , u 2 ,......., u m in the following manner:

Q1   u1a
11
, u 2a21 ,......., u mam1 
Q2   u1a 12
, u 2a22 ,......., u mam 2
………… ………. ….

54
Qn   u1a 1n

, u 2a2 n ,......., u mamn ,

So that a ij is the exponent of ui in the dimension of Q j . The dimensional matrix (matrix

of dimensions) of the given physical quantities is written as


Q1 : Q2 : Qn :

u1 :  a11 a12 ... a1n 


u2 :  a21 a22

... a2 n 

...  ... ... ... ... 
 
um :  am1 am 2 ... amn 

and this m n matrix is usually denoted by A.

Now let us form a product  of powers of Q1 , Q2 ,......., Qn , say   Q1x1 Q2x2 ......., Qnxn .

    u1a  u  u  
x1 x2 a a xn
Then 11
u2a21 .......umam1 a12
u2a22 .......umam 2 a1n
u2 2 n .......ummn  .

1 1

In order that the product  is dimensionless, we must have


a11 x1  a12 x2  .....  a1n xn  0 ,

a21x1  a22 x2  .....  a2n xn  0 ,

…. ….. …. ………..,
am1 x1  am 2 x2  .....  amn xn  0 ,

Which is a set of m homogeneous equations in n unknowns and in matrix form can be


written as
AX  0 . The number of linearly independent solutions of this equation are, evidently, n-r
where r is the rank of the dimensional matrix A. Thus corresponding to each independent
solution of X we will have a dimensionless product  and therefore the number of
dimensionless products in a complete set will be n-r. The mathematical reasoning given
above leads to the Buckingham  theorem stated earlier.

10.3. Application of  - Theorem to viscous compressible fluid motion *

In the dynamics of viscous compressible fluid the physical quantities involved are
L,U ,  ,  , k , g , p, c p , T and the fundamental units in which the dimensions of all these

quantities can be expressed are length, mass, time and temperature.


(i) The dimensional matrix in the present problem is
L U   k g p cp T

L: 1 1  3 1 1 1 1 2 0
 
M: 0 0 1 1 1 0 1 0 0
t : 0 1 0 1  3  2  2  2 0
 
 :  0 0 0 0  1 0 0  1 1  55
(ii) The rank of the matrix, which can easily be calculated, is 4. Hence the number
of independent dimensionless products will be 5.
(iii) Let us take L,U ,  and k as base quantities.
(iv) Let

 1  Lx U x  x k ,
1 2 3
x4

 2  Lx U x  x k
x8
5 6 7
g,

 3  Lx U x  x k
x12
9 10 11
p,

 4  Lx U x  x k x c p ,
13 14 15 16

 5  Lx U x  x k
x20
17 18 19
T,

(v) Now,  1   L 1 Lt 1  x


  L
x2 3
M  LMt
x3 3
 1 
x4
L
1
Mt 1 

 Lx1  x2 3 x3  x4 1 M x3  x4 1 t  x3 3 x4 1   x4 
If  1 is dimensionless, then we must have
x1  x2  3x3  x4  1  0,

x3  x4  1  0,

 x2  3x4  1  0,
 x4  0,
therefore, x1  1, x2  1 , x3  1 , and x4  0 .


Hence,  1  L1 U 1  1  
UL

Lg p LUc p
In a similar manner, we find that  2  , 3  , 4  and
U2 U 2 k
kT
5  ,
LU 8 
From these dimensionless products we can easily construct the five dimensionless
numbers, listed in g3.3 as shown below:
1 1  1 3 1
Re  , Fr  , pr   1 4 ,  and Ma 2  .
1 2   4 5  3
Thus in the dynamics of viscous compressible fluids there are only five independent
dimensionless groups. This is in complete agreement with results obtained in inspection
analysis.

56
10.4. Exercises

For Problems see end of the Block, page number 67

10.5. Keywords
Dimensional Analysis and  - Theorem

10.6. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New Delhi
(1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi (1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2, Oxford
University Press (1970).
6. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).

57
UNIT – 11
Physical Importance of Non-Dimensional Parameters

Structure:
11.0. Objectives
11.1. Reynolds Number
11.2. Froude Number
11.3. Mach Number
11.4. The ratio of specific heats
11.5. Prandtl Number
11.6. Eckert Number
11.7. Peclet Number
11.8. Grashoff Number
11.9. Brinkman Number
11.10. Exercises
11.11. Keywords
11.12. References

11.0. Objectives:

At the end of reading this unit students are able to understand Physical Importance of
non-dimensional parameters like Reynolds’ Number, Froude number, Mach number, The
ratio of specific heats, Eckert number, Peclet number, Grashoff number and Brinkman
number

11.1. Reynolds Number

The dimensionless quantity Re defined as


UL UL
Re   ,
 v
Where U , L,  and  are some characteristic values of the velocity, length, density and
viscosity of the fluid respectively, is known as the Reynolds number in honour of the
British scientist Osborne Reynolds, who in 1883 demonstrated the importance of Re in
the dynamics of viscous fluids. It can be easily seen from the equation of motion that the
u 
inertia forces  terms like   are of the order of U
 2
and the viscous forces
 x  L

 2u   U . Therefore,
 terms like  2  are of the order of
 x  L2

U 2
inertia forces L  UL  Re.

viscous forces  U v
L2

58
for this reason Reynolds number is sometimes spoken of as the ratio of inertial to viscous
forces. It is in fact a parameter for viscosity, for if Re is small the viscous forces will be
predominant and the effect of viscosity will be felt in the whole flow field. On the other
hand if Re is large the internal forces will be predominant and in such a case the effect of
viscosity can be considered to be confined in a thin layer, known as boundary layer,
adjacent to a solid boundary, However if Re is very large the flow ceases to be laminar
and becomes turbulent * . The Reynolds number at which the transition, from laminar to
turbulent, occurs is known as critical Reynolds number. For example, Reynolds in 1883
found out that for flow in a circular pipe
 ud 
Re crit     2300,
 v  crit.

where u is the mean flow velocity and d is the diameter of the pipe. In this book we shall
consider only laminar flow so that Re will always be less than Re crit .

11.2. Froude Number

The ratio of the inertial force  U 2


L  to the gravity force g  is given by the non-

dimensional parameter U 2 gL which is usually known as Froude number.


Thus we have

inertia force

 

U 2 L U 2
 Fr.
gravity force g gL
It is important only when there is a free surface, e.g., in an open channel flow problem. In
such cases too the force due to gravity may be neglected in comparison to the inertial
Fr vU viscous forces
forces if    1..
Re gL2 gravity forces
Moreover, in ordinary problems of fluid dynamics, the body is generally taken to be
completely submerged in the fluid and in such a case the effect of gravitational force may
be completely eliminated by a proper choice of the fluid pressure. Consider the flow due
to a body moving in an incompressible viscous fluid. Let, initially the fluid with body

immersed in it is at rest under the action of the body force F (may be gravitational
force) per unit mass. By letting the body move, the fluid is set in motion and the body
forces continue to act on the fluid. Let p n be the pressure when the fluid is at rest and p
be the pressure when it is in motion. For hydrostatic equilibrium, we have

F  p n, and for dynamic equilibrium

DV
   F  p   2 V .
Dt
59
From these two equations, we get

  p  p n    2 V ,
DV

Dt
 p   2 V , Where p  p  p n.

It thus follows that for the problem under consideration the body force, in the equation of
motion, may be eliminated if the pressure in the resulting equation is regarded as the
difference of actual pressure from the hydrostatic.

11.3. Mach Number

The ratio of the flow velocity to the velocity of sound,


material velocity U U
i.e.    Ma,
sound velocity c p 
is designated as the Mach number named after the German scientist Ernst Mach. It is a
measure of the compressibility of the fluid as can be seen by the following example:
Consider a steady flow of a non-viscous, non-heat conducting, compressible fluid in the
absence of external forces. The governing equations for such a flow are

 
. V  0, (3.7)

 V V  p. (3.8)

c V T  V p.
p (3.9)


Denoting the derivatives along a stream line by , and let U be the velocity along it , the
s
last two equations can be written as
U p
U  (3.10)
s s
T p
and c p  (3.11)
s s
Also from the gas equation (2.3), on differentiation,
dp d dT
  (3.12)
p  T,
1 p 1 p 1 T
Or   , (3.13)
p s p s T s
In view of the equation (3.11) and the gas equation p  RT together with the relation

R  c p  c0 , we find from equation (3.13) that

p  p p
 . (3.14)
s  s

60
Therefore, from equations (3.10) and (3.14) we conclude
d U 2 dU dU
   Ma 2 . (3.15)
 p  U U
Hence for very small Mach number the variation of density due to the variation of
velocity of the flow field is negligibly small and can be neglected and the fluid may be
considered as incompressible. In the case of air, with the velocity if sound 330m./sec, a
flow velocity of 99m/sec may be taken as the outside limit for the gaseous flow to be
considered as incompressible. However, for large Mach numbers the effect of
compressibility must be taken into account. According to the magnitude of the Mach
number the flows are, generally, classified as follows:
Mach number Type of flow
Ma  1 Subsonic
Ma  1 Transonic
Ma  1 Sonic
Ma  1  6 Supersonic
Ma  6 Hypersonic

11.4. The ratio of specific heats

The ratio of specific heat at constant pressure , c p to that at constant volume , c 0 is

usually designated as  , therefore   c p cv .

It is a measure of the relative complexity of the gas molecules. A simple expression for
c p in terms of n, the number of degrees of freedom of the approximate molecular model,
is obtained in statistical mechanics in the form
n2
cp  R, where R is the gas constant
2
n
Also, c v  c p  R  R,
2
cp n  2
therefore,   .
cv n
For monoatomic gases such as He, A, etc., the molecule is generally considered as a
smooth sphere or a mass point, has only three degrees of freedom, and therefore the value
of  is 5 3.

61
Fig 3.1. Rigid dumball having translation and rotational degrees of freedom.

For a diatomic gas such as oxygen, nitrogen or air the simplest model is a rigid dumbbell,
there are five degrees of freedom, three translation and two rotations about the two axes
perpendicular to the dumbbell axis. Hence the value of  for a diatomic gas is 7 5 , i.e.,
1.4, at normal temperature. However, at high temperature it further decreases, because the
atoms in a diatomic molecule are not rigidly bound but can vibrate and this vibration,
classically, would add two degrees of freedom and therefore the value of  is to be
expected as 9 7.

Fig 3.2. Diatomic molecules having translation, rotational and vibrational degrees of freedom.

Ordinarily we will assume that a diatomic gas has only five degrees of freedom and the
value of  will be taken as 1.4.
11.5. Prandtl Number

The ratio of the kinematic viscosity to the thermal diffusivity of the fluid, i.e.,
kinematic vis cos ity v   cp
    Pr ,
thermal diffusivity a k c p k

62
is designated as the prandtl number named after the German scientist Ludwig prandtl. It
is a measure of the relative importance of heat conduction and viscosity of the fluid. The
prandtl number, like the viscosity and thermal conductivity, is a material property and it
thus varies from fluid to fluid. For air pr  0.7 (approx.) and for water at 60F ,
Pr  7.0 (approx.) . For liquid metals the prandtl number is very small, e.g., for mercury
Pr  0.044 , but for highly viscous fluids it may be very large, e.g., for glycerin
Pr  7250 .
Note: Beside these non-dimensional quantities of the complete set the following other
non-dimensional quantities also occur in the study of the dynamics of viscous fluids:

11.6. Eckert number

The dimensionless quantity “Ec” defined as Ec  U 2 c p T , Where U, c p and T are some

reference values of the velocity, specific heat at constant pressure and the temperature, is
known as the Eckert number named after the German scientist E.R.G Eckert. In
compressible fluids it determines the relative rise in temperature of the fluid due to
adiabatic compression. It can also be retained in incompressible flow, if the frictional heat
is to be considered, but the interpretation with reference to adiabatic compression will no
longer be true. As already pointed out, in high speed flow, for gases the Eckert number
becomes equivalent to Mach number (c.f. g3.3),
Ec    1 Ma 2 .

11.7. Peclet number

In heat transfer theory the peclet number is defined as


UL UL v
Pe'    Re.Pr. It is the product of Reynolds number and Prandtl number.
a v a
11.8. Grashoff number

The dimensionless quantity Gr, which characterizes the free convection * is known as
the Grashoff number and is defined as
gL3 Tw  T 
Gr  , where g is the acceleration due to gravity and Tw and T are two
v 2T
representative temperatures.
11.9. Brinkman number
U 2
The dimensionless quantity “Br” defined as Br  , where  ,U , k , Tb and T0
k Tb  T0 
are some reference values of the viscosity, velocity conductivity, and two different
temperatures, is known as the Brinkman number named after professor H.C.Brinkman,

63
who solved the problem of flow in a circular tube with viscous heat effects . It is a
measure of the extent to which viscous heating is important relative to the heat flow
resulting from the impressed temperature difference ( Tb  T0 ).

11.10. Exercises

For Problems see end of the Block, page number 67

11.11. Keywords
Reynolds’ Number, Froude number, Mach number, The ratio of specific heats, Eckert
number, Peclet number, Grashoff number and Brinkman number

11.12. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New Delhi
(1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi (1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2, Oxford
University Press (1970).
6. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).

64
UNIT – 12
Important Non-Dimensional Coefficient in the Dynamics of
Viscous Fluids

Structure:
12.0. Objectives
12.1. Important non-dimensional coefficients in the dynamics of viscous fluids
12.2. Dynamical Similarity (Reynolds’ Law)
12.3. Summary
12.4. Exercises
12.5. Keywords
12.6. References

12.0. Objectives:

At the end of reading this unit students are able to understand important non-dimensional
coefficients in the dynamics of viscous fluids.

12.1. Important non-dimensional coefficients in the dynamics of viscous fluids

In order to complete the studies of non- dimensional quantities, which occur in the
dynamics of viscous fluids, let us mention some important dimensionless coefficients
which are usually calculated in the analysis and their values are compared with the
experimental results.
(i) Lift and drag coefficients

If F is the force on an obstacle , placed in an otherwise undisturbed stream, due to the

system of stresses over its surface, then the component of F in the direction of the
undisturbed stream is called the drag force and denoted by D, and the component at right
angles to this called the lift ant is denoted by L. The expressions
D L
CD  , CL  ,
U S 2
2
U 2 S 2

where S represents a typical area associated with the obstacle , are called the drag
coefficient and lift coefficient respectively.
(ii) Local skin-friction coefficient
The dimensionless shearing stress on the surface of a body, due to a fluid motion, is
w
known as ‘local skin-friction coefficient’ and is defined as C f  , where  w is the
U 2 2

local shearing stress on the surface of the body.

65
(ii) Dimensionless coefficient of heat transfer (Nusselt Number)
In the dynamics of viscous fluids one is not much interested to know all the details of the
velocity and temperature fields but would certainly like to know quantity of heat
exchanged between the body the fluid. This quantity of heat transfer can be calculated
with help of a coefficient of heat transfer  x  , which is defined by Newton’s law of
cooling as follows:

If qx  is the quantity of heat exchanged between the wall and the fluid, per unit area
per unit time, at a point x, then

qx   xTw  T  , (Newton’s law of cooling)

where ( Tw T ) is the difference between the temperature of the wall and that of the
fluid.
Since at the boundary the heat exchanged between the fluid and the body is only due to
conduction, according to Fourier’s law, we have

 T 
q  x   k  
 n n 0
where n is the direction of the normal to the surface of the body. From these two laws we
can define dimensionless coefficient of heat transfer which is generally known as the
Nusselt number as follows:
 x L L  T 
Nu  
 Tw  T   n n 0
, where L is some characteristic length in the problem.
k

(iii)Temperature recovery factor


The temperature which a surface assumes under the influence of internal friction is called
the recovery temperature or adiabatic wall temperature. The difference between the
recovery temperature T, and the temperature of the steam T is often made dimensionless
by dividing it by the quantity U 2 2c p and the resulting expression is known as the
temperature recovery factor or simply the recovery factor and is usually denoted by ‘r’.
Tr  T
Thus r  ,
U 2 2c p

It is generally found to be a function of Prandtl number and is a measure of the energy


distribution in the flow field. It is particularly important in a high speed flow in which the
frictional heat plays an important role.

66
12.2. Summary:

Here we discuss the topics of Dynamical similarity and Dimensional analysis have been
discussed and the physical importance of the non-dimensional parameters and
coefficients, which play an important role in the study of viscous fluids.

12.3. Exercises
1. Explain the physical meaning of Reynolds number and Frude number.
2. State and prove Buckingham theorem for dimensional analysis.
3. Explain Reynolds principle of similarity.
4. Explain the application of Buckingham theorem to viscous compressible fluid
motion.
5. Explain the physical importance of non-dimensional parameters.
6. Explain non-dimensional coefficients in the dynamics of viscous fluids.
7. Check the dynamical similarity of the following equations (i)
p q2 u v
  z  cons tan t. (ii)  0.
 2g x y
8. An oil of specific gravity 0.85 is flowing through a pipe of 5cm diameter at the
rate of 3 liters/sec. Find the type of flow, if the viscosity for the oil is 3.8poise.
9. A one fourth scale of a helicopter is to be tested in a wind-tunnel such that the
kinematic similarity is obtained. The helicopter is designed to fly in level flight 40
meter/sec in a region where v=1.4*10-5ite shore note on inspection analysis.
10. Define and give the physical importance of the following non-dimensional
parameters (i) local skin friction coefficient (ii) Prandtl number.
11. Define and give the physical importance of the following non-dimensional
parameters (i) Eckert number (ii) Reynolds number.
12. Show that in the dynamics of compressible fluids there are only five independent
dimensionless groups.
13. The pressure difference in a pipe of diameter D and length L due to turbulent flow
depends on the velocity, density, and roughness. Using Buckingham’s theorem
obtain an expression for pressure difference.
14. A cylindrical tank of diameter D discharges water through a short pipe of
diameter d into the atmosphere. The water surface is at a distance L above the
level of the pipe. The mass flow rate m is known to depends on the lengths D, d
and L, the acceleration due to gravity g density of water ρ. Determine the
dimensionless numbers of the physical problem just described.
15. Write short note on dimensional analysis.
67
12.4. Keywords
Non-dimensional Coefficients, Dynamics of Viscous Fluids

12.5. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New


Delhi (1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi
(1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2, Oxford
University Press (1970).
6. Schlichting. H., Boundary Layer Theory, McGraw-Hill (1968).
7. Streeter .V. L. (Ed)., Hand Book of Fluid Dynamics, McGraw2-Hill (1961).
8. Yuan. S.W: Foundational of Fluid Mechanics, Prentice. Hall of India Pvt. Ltd.,
New Delhi (1969).
9. Bansal. J. L., Viscous Fluid Dynamics, Oxford and IBH publishing Co. Pvt. Ltd.,
New Delhi (1977).
10. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).

68
Block – 4
Exact Solutions of the
Navier Stoke’s Equations

69
UNIT – 13
Introduction – Flow between Parallel Plates

Structure:
13.0. Objectives
13.1. Introduction
13.2. Flow between parallel plates (velocity distribution)
13.3. Flow between parallel plates (temperature distribution)
13.4. Exercises
13.5. Keywords
13.6. References

13.0. Objectives:

Since the Navier-Stoke’s (N-S) equations of motion are non- linear in character, there is
no known general method to solve these equations. Only in few special cases exact
solutions can be obtained by making certain assumptions about the state of the fluid and a
simple configuration for the flow pattern. The exact solutions of N-S equations discussed
in this chapter are valid for all ranges of viscosity but are few in number. However, there
appear, at present, no such exact solution of the flow past bodies of finite size and,
therefore, in order to study such flows the N-S equations have been simplified to
mathematically tractable forms for two extreme cases.

13.1. Introduction

Since the Navier-Stoke’s equations are non-linear in character there is no known general
method to solve these equations. Only in few special cases exact solutions can be
obtained, where we have to make certain assumptions about the state of the fluid and a
simple configuration for the flow pattern is to be considered. We shall divide some of
these known cases into the following five categories:
(A) Steady incompressible flow with constant fluid properties,
(B) Steady incompressible flow with temperature dependent viscosity,
(C) Unsteady incompressible flow with constant fluid properties,
(D) Steady compressible flow, and
(E) Steady incompressible flow with fluid suction/injection on the boundaries.
(A) Steady incompressible flow with constant fluid properties
If at various points of the flow field all quantities (such as velocity, density, pressure)
associated with the flow field remain unchanged with time, the motion is said to be
steady; otherwise it is called unsteady. Thus in steady motion time drops out of the
independent variables and the various field quantities simply become functions of the

70
space coordinates. If the density of a fluid is regarded as constant throughout the flow
field, it is said to be incompressible fluid. Such an assumption is possible for liquids and
also for gases at low speeds  Ma  1  . As pointed out in §1.3 the viscosity of a fluid, in

general, depends on the temperature and this dependence on temperature is more


significant in case of gases and lubricating oils. However, for most of the incompressible
fluids the viscosity may be regarded as constant. Moreover, this assumption has an
additional advantage that in such a case the velocity field is independent of the
temperature field so that the equation of continuity and the equations of motion can be
solved first to get the velocity and pressure distribution and then the results can be
employed to evaluate the temperature field.

13.2. Flow between parallel plates (velocity distribution)

A very simple solution of the equations  B1 to  B4  , c.f. Table 2.2, can be obtained for

the flow between two parallel plates which are kept at a finite distance apart. We shall
assume that the x -axis is along the direction of the flow, the y -axis being at right angles
to it and the width of the plates, parallel to the z -direction, be large compared to the
distance between the plates. The motion is two-dimensional and therefore all the
variables will be independent of z -coordinate. Hence,

( )  0, u  u ( x, y ), v  0, w  0, and p  p( x, y ). (4.1)
z
Further, the equation of continuity and the equations of motion reduce to
u
0 , (4.2)
x
p  2u
0  , (4.3)
x y 2
p
0 , (4.4)
y
where the equation of continuity (4.2) is used in obtaining the equation (4.3). From
equations (4.2) and (4.4), we conclude that u will be a function of y only and p will be
a function of x only, therefore, the equation (4.3) can be written as
d 2u 1 dp
 . (4.5)
dy 2
 dx
In which the left-hand side is a function of y only and the right-hand side is a function of
x only and hence for its validity both should be constant.
On integrating equation (4.5), we find

71
1 dp 2
u y  Ay  B (4.6)
2 dx
where A and B are arbitrary constant to be determined by the boundary conditions of
the following three different types of flows:
(i) Plane Couette flow

Fig 4.1. Velocity distribution in plane Couette flow

The so called plane Couette flow or simple shear flow is the flow between two parallel
plates one of which is at rest and the other moving with a uniform velocity U in its own
plane and the pressure gradient is taken to be zero. If the x -axis is taken along the
stationary plate and the distance between the plates be denoted by h , then the boundary
conditions are
y0:u 0

y  h : u U (4.7)
Hence the velocity field, in the present case, from equation (4.6) in non – dimensional
form is given by
u y
 . (4.8)
U h
The velocity distribution is linear as shown in fig 4.1.

(ii) Plane Poiseuille flow

In plane Poiseuille flow the pressure gradient is non-zero, but the plates are kept
stationary. Let the distance between the plates be 2b and the axis of x , for the sake of
convenience, be taken in the middle of the channel. In such a case the boundary
conditions are

72
Fig 4.2. Velocity distribution in plane Poiseuille flow
y  b : u  0. (4.9)
Therefore, the equation (4.6) becomes
1 dp 2
u (b  y 2 ) (4.10)
2 dx
The velocity distribution is parabolic, as shown in fig. 4.2, and the maximum velocity
occurs in the middle of the channel which is
b 2 dp
um   . (4.11)
2 dx
Hence, the velocity distribution in a plane Poiseuille flow, in non-dimensional form, is
given by
u y2
 1 2 . (4.12)
um b
(iii) Generalized plane coquette flow
It is a plane Couette flow with non-zero pressure gradient. Therefore, the boundary
conditions are the same as that in Case (i), viz.,

Fig 4.2. Velocity distribution in generalized plane Couette flow

73
y 0:u 0,
y  h : u U . (4.7)
Hence, equation (4.6) becomes
y h2 dp y y
u U (1  ) . (4.13)
h 2 dx h h
Lets us introduce a dimensionless pressure P , defined as
h 2  dp 
P   (4.14)
2U  dx 
Thus the velocity field in a generalized plane Couette flow, is non – dimensional form, is
given by
u y y y
  P 1   (4.15)
U h h h
It is clear that the velocity field will depend on the nature of the non-dimensional pressure
gradient P and therefore three different cases are possible.
Case I: P0
When the pressure is decreasing in the direction of the flow or in other words for a
favorable pressure gradient the velocity distribution is positive over the whole width of
the channel, as can be seen in fig. 4.3.
Case II: P0
When the pressure is increasing in the direction of the flow or in other words for an
adverse pressure gradient, we find that a back flow begins to occur near the stationary
plate as P < -1. It is due to the influence of the adverse pressure gradient which surpasses
the dragging action of the faster layers on fluid particles in that region.
Case III: P0
In the case of zero pressure gradients i.e., when the pressure is constant throughout flow
field the equation (4.15) reduces to equation (4.8) and in this case the velocity
distribution is linear.
Characteristic parameters
(a)Volume rate of flow
The volume of flow Q per unit width, per unit time, at any normal section of the channel
is given by
h
Q   u dy . (4.16)
0

Substituting the value of u from equation (4.15) in (4.16), we find


Uh PUh
Q  . (4.17)
2 6
74
If P  3 , the volume rate of flow becomes zero. The means that there is no net flow
across any section perpendicular to the direction of motion. In such a case the dragging
action of the faster layers will be balanced exactly by the influence of the adverse
pressure gradient.
(b) Coefficient of skin-friction
The shearing stress on the stationary plate is given by
 dw  U
w      (1  P) . (4.18)
 dy  y 0 h

Therefore, the coefficient of skin-friction in the present case is


w 2
cf   (1  P) , (4.19)
U / 2
2
Re
hpU
Where, Re  (Reynolds number).

Clearly, c f will be positive if P is positive, it will be negative if P  1 (case of back

flow) and zero if P  1. For plane Coquette flow (i.e., P  0 ) c f will have a fixed

value 2 / Re .

13.3. Flow between parallel plates (temperature distribution)

Having determined the velocity distribution, in the above three types of flows between
two parallel plates, we are in a position to calculate the temperature distribution from
energy equation (B5), c.f. Table 2.2.
Substituting equation (4.1) into equation (B5) and taking the help of the continuity
equation (4.2), the energy equation, for the steady flow and without heat addition,
becomes
2
T   2T  2T   u 
 c pu k 2  2     , (4.20)
x  x y   y 
where k and  are taken to be constant. If the plates are kept at constant temperature,
T
then  0, and hence the equation (4.20) reduces to
x
2
 2T  u 
k 2     . (4.21)
y  y 
We will now calculate the temperature distribution for the three cases discussed in §4.2.

75
(i) Plane Couette flow
In this case the velocity distribution is given by equation (4.8) and therefore the equation
(4.21) becomes
 2T U2
k    (4.22)
y 2 h2
If the plates are kept at different temperatures, the boundary conditions are
y  0 : T  T0 ,
T1  T0 . (4.23)
y  h : T  T1 ,
The solution of equation (4.22), with the boundary condition (4.23), is
T  T0 y 1 y y
  Ec.Pr 1   , (4.24)
T1  T0 h 2 h h

U2
where, Ec  (Eckert number),
c p (T1  T0 )

c p
and Pr  (Prandtl number).
k

Fig 4.4. Temperature distribution in plane Couette flow plates are kept at unequal temperatures.

( T0 = Temperature of the stationary plates, T1 = Temperature of the moving plate)

It can be seen from equation (4.24) that the temperature distribution in the present case
depends on the product of Eckert and Prandtl numbers and for different values of this
product the dimensionless temperature distribution is plotted against the dimensionless
distance between the plates in fig. 4.4. Now in order to see the heat transfer at the upper
plate, let us calculate the dimensionless coefficient of heat transfer, viz., Nusselt number
defined as

76
h  T 
Nu  
T1  T0   y  y h
(4.25)

T
Substituting the value of from equation from (4.24) in (4.25) ,we find
y
Ec.Pr
Nu   1. (4.26)
2
Thus the Nusselt number will be positive if Ec. Pr > 2 and in this case the heat will
transferred from fluid to the upper plate. If Ec. Pr < 2, the Nusselt number will be
negative i.e., the reversal in the heat transfer will take place and the heat will be
transferred from upper plate to the fluid. If Ec. Pr = 2, there will be no transfer of heat
between the fluid and upper plate. Hence, we may say that
if Ec. Pr > 2, the upper plate will be heated,
and if Ec. Pr < 2, the upper plate will be cooled.
In the similar way we may calculate the Nusselt number for the heat transfer at the lower
plate and may find that at the stationery plate the heat is always transferred from fluid to
plate irrespective of the range of Ec. Pr.
If both the plates are kept at the same constant T0 (i.e.,if T1 = T0 ) the temperature
distribution from equation (4.24) becomes
U 2 y  y
T  T0  1   . (4.27)
2k h  h 

Fig 4.5. Temperature distribution in plane Couette flow plates are kept at equal temperatures.

( Tm = Temperature in the middle of the channel)

It is parabolic, as shown in fig. (4.5), and the maximum temperature exist in the middle of
the channel, which is given by
U 2
Tm  T0  . (4.28)
8k
The Nusselt number in the present case, defined as

77
h  T 
Nu  
T0  Tm   y  y 0
,

for the lower plate has a constant value 4.


Another simple solution of the equation (4.22) can be obtained if we assume that at one
of the plate, say the stationery plate, no heat transfer takes place (adiabatic wall). The
boundary conditions, therefore are ,
T
y0:  0,
y (4.29)
y  h : T  T1.
With these the temperature distribution is given by

U 2  y2 
T  T1   1   (4.30)
2k  h 2 

Fig 4.6. Temperature distribution in plane Couette flow, the stationary plates is insulated.

( T1 = Temperature of the moving plate, Tr = recovery temperature)

The temperature distribution is shown in fig 4.6. The temperature which an insulated
surface assumes under the influence of internal friction is known as recovery

temperature. The difference between the recovery temperature Tr   T  y 0  and the


 
temperature of the upper plate, from equation (4.30) is given by
U 2
Tr  T1  . (4.31)
2k
Thus the recovery factor in a plane Couette flow is
Tr  T1
r  Pr (Prandtl Number). (4.32)
U 2 / 2c p
78
(ii) Plane Poiseuille flow
In this case, the velocity distribution is given by equation (4.12) and therefore the
equation (4.21) becomes
d 2T 4  um 2 2
k   y . (4.33)
dy 2 b4
Let both the plates are kept at the same constant temperature T0 . Therefore the boundary

conditions are y   b : T  T0 . (4.34)

The solution of equation (4.33) under the boundary conditions (4.34) is

 um 2  
4
 y
T  T0  1    . (4.35)
3k   b  

Fig 4.7. Temperature distribution in plane Poiseuillo. Plates are kept at equal temperatures.

( Tm = Temperature in the middle of the channel)

The temperature distribution is shown in fig. (4.7). The maximum temperature exists in
the middle of the channel and is given by
 um 2
Tm  T0  . (4.36)
3k
An extension of this to the temperature dependent viscosity will be taken up in section
 B .
(iii) generalized plane Couette flow
In this case the velocity distribution is given by equation (4.15) and therefore equation
(4.21) becomes

U 2  y
2
d 2T 2 y
k 2   2 (1  P)  4 P    4 P(1  P) 
2
(4.37)
dy h  h h 

79
Let, both the plates be kept at the same constant temperature T0 . Therefore, the boundary
conditions are
y  0 : T  T0 ,
(4.38)
y  h : T  T0 .
The solution of equation (4.37), with the boundary conditions ( 4.38), is

U 2  2 y   y 2    y 3  y
2
T  T0  3(1  P) 1    4 P(1  P) 1      2 P 1     . (4.39)
6k   h   h     h   h

The temperature gradient at the lower plate is given by


 T  U 2
    2  (1  P)2  . (4.40)
 y  y 0 6 k

This shows the heat will always be transferred from the fluid to the lower plate
irrespective to the sign of P.

13.4. Exercises

For Problems see end of the Block, page number 112

13.5. Keywords
Velocity, temperature, skin friction, uniform tubes,

13.6. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New Delhi
(1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi (1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2, Oxford
University Press (1970).
6. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).
7. Bansal. J. L., Viscous Fluid Dynamics, Oxford and IBH publishing Co. Pvt. Ltd.,
New Delhi (1977).
8. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).

80
UNIT – 14
Introduction – Flow between Parallel Plates

Structure:
14.0. Objectives
14.1. Flow in the circular pipe (Hagen-Poiseuille flow)
14.2. Exercises
14.3. Keywords
14.4. References

14.0. Objectives:

At the end of reading this unit students are able to understand velocity and temperature
disributions of flow in the circular pipe.

14.1. Flow in the Circular Pipe (Hagen-Poiseuille flow)

(a) Velocity distribution


The fully developed steady laminar flow through a long straight circular pipe, without
body forces, is another example of the exact solution of the Navier-Stokes equations of
motion. Such a motion is maintained by the presence of a pressure gradient along the
axis of the pipe. Let the axis of the pipe be taken as z -axis along which the flow takes
place and r denotes the radial distance measured outward from the z -axis. Due to axial
symmetry
d
( )0 (4.41)
d
or in other words, all variable will be independent of  . Also the only non-zero
component of velocity is vz . Hence the equation of continuity and the equation of
motion, c.f. equation (D1) to (D4), Table 2.4, cylindrical coordinates reduce to
vz
0 , (4.42)
z
p
0 , (4.43)
r
  2vz 1 vz  p
and    . (4.44)
 r r r  r
2

81
Fig 4.8. Laminar flow in a circular pipe (Hagen poiseeuille flow)

From equations (4.42) and (4.43) it is clear that vz is a function of r only and p is a
function of z only. Therefore the equation (4.44) can be written as
  2vz 1 vz  p
   . (4.45)
  r 2
r r  z
Since p is not a function of r and vz is not a function of z , equation (4.45) can only be

dp
valid when is a constant.
dz
Integrating equation (4.45), we get
1 dp 2
vz  r  A ln r  B , (4.46)
4 dz
where A and B are arbitrary constants to be determined by the following conditions:
dvz
r 0:  0 (due tosymmetry),
dr (4.47)
r  R : vz  0 (no-slip condition),
where R is the radius of the pipe.
Therefore, the equation (4.46), for the velocity distribution, becomes,

R 2 dp   r  
2

vz   1     . (4.48)
4 dz   R  

which is of the form of a paraboloid of revolution as shown in fig.4.8. The pressure


dp p  p1
gradient is clearly equal to 2 , where L is the distance between the two
dz L
sections of the pipe where the pressures p1 and p2 are measured.
The maximum velocity occurs on the axis of the pipe and is given by
R 2 dp p1  p2 R 2
(v z ) m    . (4.49)
4 dz L 4
82
Hence the velocity field in a non-dimensional form in Hagen-Poiseuille flow is given by
2
vz r
 1   , (4.50)
 vz  m R

The average velocity over a cross section can be obtained as


R

v z 2 r dr
R 2 dp 1
vz  0
  (vz ) m , (4.51)
 R2 8 dz 2
and the volume rate of flow Q becomes
 R 4  dp   R 4 p1  p2
Q   R 2 vz     . (4.52)
8  dz  8 L
This relation was first obtained by Hagen and then independently, by Poiseuille. This
relation can determine the viscosity of a fluid if other quantities are measured.
Coefficient of skin-friction
The shearing stress on the wall of the pipe is given by, c.f. (D9), Table 2.4,
 dv  R dp 4
( rz ) w     z     vz . (4.53)
 dr r  R 2 dz R
Therefore, the coefficient of skin friction in the present case is
( rz ) w
cf   16 / Re , (4.54)
 (v z ) 2 / 2

where, Re  2 R vz /  (Reynolds number).

The laminar flow described above occurs in practice only when the Reynolds number Re
is less than 2300, which is the critical Reynolds number in the present case as has been
found by experiments.
The value of the coefficient of skin-friction given by the equation (4.54), for various
values of Re in the laminar range, has been verified experimentally by G. Hagen. Since
the Hagen-Poiseuille flow is an exact solution of the Navier-Stokes’ equations, seeking
the excellent agreement between the theoretical results and experimental data for the
coefficient of skin-friction, it can be said that the Navier-Stokes’ equations of motion are
true at least for parallel flows.
(b) Temperature distribution
Having determined the velocity distribution, we are in a position to calculate the
temperature distribution from the energy equation (D5), c.f. Table 2.4.
Substituting equation (4.41) in the equation (D5), keeping in view that the only non-zero
component of the velocity is vz , the energy equation for the steady flow and without heat
addition becomes

83
  2T 1 T  2T   vz 
2
T
 c p vz k 2       . (4.55)
z  r r r z 2   r 
where k and  are taken to be constant. We shall now study two cases (i) when the wall
of the pipe is kept at a constant temperature, and (ii) when the wall of the pipe is kept at a
uniform temperature gradient.
(i) Wall at constant temperature
T
If the wall of the pipe is kept at a constant temperature, then = 0 and hence the
z
equation (4.55) reduces to

  2T 1 T   vz  4  ( vz ) m 2 r 2
2

k 2          . (4.56)
 r r r   
4
 r R

The boundary conditions are


r  0 : T  finite,
(4.57)
r  R : T  T0 .

Fig 4.9. Temperature distribution in Hagen poiseuille flow(Wall at constant temperature T0 )

The solution of equation (4.56) with the boundary conditions (4.57) is

 (vz )m2  r4 
T  Tc   1  4 
. (4.58)
4k  R 
The maximum temperature exists on the axis of the pipe and is given by
 (vz )2m
Tm  Tc  (4.59)
4k
Hence the dimensionless temperature distribution in Hagen Poiseuille flow is given by
T  T0 r4
 1 4 . (4.60)
Tm  T0 R

84
The mean temperature over a cross-section can be obtained as

 T 2 r dr  (v z ) 2 m
Tmean  0
  T0 (4.60)
 R2 6k
 (v z ) 2 m
or Tmean  T0  . (4.61)
6k
The Nusselt number, in the present case, defined as
2 R  T 
Nu     (4.62)
Tmean  T0  r r  R
=12.
(ii) Wall at uniform temperature gradient
If the wall of the pipe is kept at a constant temperature gradient, say, A, then, keeping in
view that the conditions will be similar at each section of the pipe, we may seek the
solution of equation (4.55) in the form
T  Az  g (r ) (4.63)
Substituting (4.63) and (4.50) in equation (4.55), and neglected its last term i.e., the heat
due to dissipation, we find
  r 2   d 2 g 1 dg 
 c p (vz )m 1     A  k  2   (4.64)
  R    dr r dr 

The solution, which is free from singularities, and satisfies the condition that at r = R : g
= 0, is

 c p (vz )m AR 2  3 1 r  1r
2 4

g (r )           (4.65)
k 10 4  R  16  R  
The maximum temperature exists on the axis of the pipe and is given by
3
Tw  Az   c p (vz )m AR 2 . (4.66)
16
Let us now calculate the mean temperatures Tmean and TMean viz., the unweighted mean and
mean weighted with respect to the velocity-i.e. the temperature which is measured in
fluid which is mixed after passing through the pipe, respectively as follows:
R

 T 2 r dr 1  c p (vz )m AR
2

Tmean  0
 Az  , (4.67)
 R2 12 k
R

 Tv z 2 r dr
11  c p (vz ) m AR
2

and TMean  0
R
 Az  . (4.68)
96 k
v
0
z 2 r dr

85
The Nusselt number, based on the unweighted mean temperature Tmean , is given by
2R  T 
Nu    6,
Tmean  Tw   r r  R
(4.69)

where Tw  Az .
But, when Nusselt number is based on the weighted mean temperature TMean ,we have
2R  T  48
Nu      . (4.70)
TMean  Tw   r r  R 11

14.2. Exercises

For Problems see end of the Block, page number 112

14.3. Keywords
Flow in Circular pipe, Velocity Distribution, and Temperature Distribution

14.4. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New Delhi
(1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi (1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2, Oxford
University Press (1970).
6. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).
7. Bansal. J. L., Viscous Fluid Dynamics, Oxford and IBH publishing Co. Pvt. Ltd.,
New Delhi (1977).
8. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).

86
UNIT – 15
Flow between two Concentric Rotating Cylinders

Structure:
15.0. Objectives
15.1. Flow between two concentric rotating cylinders (Couette flow)
15.2. Flow in convergent and divergent channels (Jeffery Hamel flow)
15.3. Stagnation point flows
15.4. Exercises
15.5. Keywords
15.6. References

15.0. Objectives:

At the end of reading this unit students are able to understand velocity and temperature
disributions of flow between two concentric Rotating Cylenders.

15.1. Flow between two concentric rotating cylinders (Couette flow)

(a) Velocity distribution


Let us consider the flow between two concentric rotating cylinders. Let r1 , 1 and r2 ,

2 be the radius and angular velocity of the inner and of the outer cylinders
respectively.
Let the z -axis is along the common axis of the cylinders, and r denotes the radial
direction measured outward from the z -axis. The nature of the problem suggests that the
only non-zero component of velocity is v and there is no pressure gradient in the θ-
direction.

Fig 4.10. Flow between two concentric rotating cylinders (Couette flow)

87
Hence the equation of continuity and the equations of motion, c.f. equations (D1) to (D4),
Table 2.4, in cylindrical polar coordinates reduce to
 v
 0, i.e., v  v (r ) , (4.71)

v 2 dp
  , (4.72)
r dr
d 2 v d  v 
  0
dr 2 dr  r
and (4.73)

Integrating equation (4.73), we get
v  Ar  B / r , (4.74)
where the arbitrary constants A and B are to be determined by the following boundary
conditions:
r  r1 : v  r1 1 ,
(4.75)
r  r2 : v  r2 2 .

Therefore the equation (4.74), for the velocity distribution, becomes

1  r12 r2 2 
v  
 22
( r 2
  r 2
) r  (   ) , (4.76)
(r2 2  r12 ) 
11 2 1
r 

The radial pressure distribution may, now, be calculated from equation (4.72).
If the inner cylinder is at rest, then 1  0 and

2 r2 2  r12 
v   r   (4.77)
(r2 2  r12 )  r 

The only viscous stress in the fluid is  r (c.f. Table 2.4 (D9) and is given by

d  v  22 r1 r2
2 2
 r   r 
dr  r  (r 2  r 2 )r 2 , (4.78)
 2 1

when the inner cylinder is taken at rest.

Fig 4.11 Laminar flow in the space between two coaxial cylinder. Inner cylinder stationary and the outer
rotating with an uniform angular velocity

88
The torque M required to turn the outer cylinder may now be easily calculated as the
product of the force and the arm of the couple.
r12 r2 2
M  L( r )r r2 .2r2  L z , (4.79)
(r2 2  r12 )
where L is the length of the cylinder.
The system considered here is frequently used to determine the fluid viscosities from
observations of torque and angular velocities. It was Couette, who first used this model
to determine the coefficient of viscosity and that is why this flow is, generally, known as
Couette flow.
(b) Temperature distribution
Having determined the velocity distribution, the temperature distribution can now be
obtained from equation (D5) of table 2.4, the equation of energy in cylindrical polar
coordinates. In the present problem it reduces to
2
1 d  dT   d  v 
0k r  r   (4.80)
r dr  dr   dr  r 
The boundary conditions are
r  r1 : T  T1 ,
(4.81)
r  r2 : T  T2 .

Substituting the value of v / r from equation (4.76) in the equation (4.80) and performing
the indicated differentiation, it becomes

1 d  dT  4 2  1  r14 r2 4
r  . (4.82)
r dr  dr  k (r2 2  r12 )2 r 4
The solution of equation (4.82), with the boundary conditions (4.81), is given by

T  T1 
r 2  r12 r2 2 
 N 2 2 2  (1  N )
l n(r / r1 )
, (4.83)
T2  T1 (r2  r1 )r l n(r2 / r1 )

 (2  1 )2 r12 r2 2
where, N  (Non dimensional parameter). (4.84)
k (r2 2  r12 )(T2  T1 )
This solution is of some interest in connection with the heat effects in viscometry.

15.2. Flow in convergent and divergent channels (Jeffery Hamel flow)

Let us consider the steady flow of a viscous incompressible fluid, in the absence of
external forces, between two non-parallel plane walls. We use the cylindrical polar
coordinates  r ,  , z  , in which the line of intersection of the planes is taken as the z -axis

and r is the distance from this line. The walls are in the planes,     .
89
Fig 4.12. Flow in convergent and divergent channel (Jeffery Hemel flow)

If the motion is purely radial, then the only non-zero component of velocity is vr and
hence the equation of continuity and momentum (c.f. Table 2.4) reduce to

(rvr )  0 , (4.85)
r

vr p   2v 1 vr 1  2vr vr 


 vr      2r    . (4.86)
r r  r r r r 2  2 r 2 

1 p 2 v
and, 0    2 r . (4.87)
r  r 
The boundary conditions are,
    , vr  0 . (4.88)

The equation of continuity (4.85) suggests


vr  f ( ) / r , (4.89)

where f ( ) is an arbitrary function of  to be determined.


Substituting equation (4.89) in equations (4.86) and (4.87), we get
f2 p f ''
p   3 , (4.90)
r 3
r r
p f'
0  2 2 , (4.91)
 r
and the corresponding boundary conditions are
    : f  0. (4.92)
Eliminating p between (4.90) and (4.91), we have
2 ff   v(4 f   f )  0 , (4.93)
where a prime denotes differentiation with respect to  .
Integrating equation (4.93) once, we find

90
f 2  v(4 f  f )  K , (4.94)
where K is a constant of integration.
Multiplying (4.94) by 2 f  and integrating once again, we get
2
f 2  (h  3Kf  6vf 2  f 3 ), (4.95)
3v
where h is a second constant of integration.
The problem is to solve the equation (4.95) with the help of the boundary conditions
given in (4.92), which are two in number. But the solution of the equation (4.95) requires
three boundary conditions and, therefore, an additional condition is to be prescribed.
When the flow is purely divergent or purely convergent the function f will be
symmetrical about  = 0; and in such a case the value of f at  = 0 may be prescribed.
Equation (4.95) can be written as
2
f 2  ( f1  f )( f 2  f )( f3  f ), (4.96)
3v
where the constant f 1 , f 2 and f 3 are connected by the relations

f1  f 2  f3  6v , (4.97)

f1 f 2  f 2 f3  f3 f1  3K , (4.98)

and, f1 f 2 f3  h . (4.99)
Integrating equation (4.96), we get
0 f0
3v df
 

d  
2  ( f
0  f )( f 2  f )( f 3  f )
1/2
(4.100)
1

where f 0 is the value of f at  = 0.


The solution of equation (4.100) may be expressed in terms of the elliptic function as has
been done by Millsaps and Pohlhausen, who have discussed the different possible cases
of this type of flow. Here we shall restrict ourselves to the flows with large Reynolds
number.

Case I – flow in a divergent channel


For the flow in a divergent channel the radial velocity will be positive and, therefore,
from equation (4.89) we conclude that in this case f  0 . Since in the middle of the
channel f   0 , we find from equation (4.96) that f must be equal to
f1 , f 2 or f3 . Let f  f1  f0 .

91
Fig 4.13. Flow in a divergent channel
The equation (4.100), in view of equations (4.97) and (4.99), may be written as
f
3v 1 df
  . (4.101)
 1 1
2 0 ( f  f ) f 2  (6v  f ) f  h / f 1/2
1  
Since, on the walls f vanishes, it follows from equation (4.95) that

f '2  2h / 3v . (4.102)
Hence h is positive. Therefore, α has its greatest permissible value for a given value of f1
when h=0, i.e. either f 2 or f3 is zero.

Let f  f1 cos2  , Re  f1 / v  r (vr )max / v (Reynolds number) (4.103)


Substituting (4.103) in the equation (4.101) and neglecting the term containing 6v, when
Re is large, we get
 /2
3v d

f1   1 2 
1/2
,
1  sin  
0

 2 
 /2
d
 Re  3   3.211
1/2
or 1/2
(4.104)
 1 2 
1  sin  
0

 2 
Thus  Re1/2 has an upper limit when Re is large and  is small. Hence if the angle 
and the Reynolds number Re are specified, then the velocity profiles may be calculated
from equation (4.96) taking either f2 or f3 equal to zero.

92
Case II – flow in a convergent channel

Fig 4.14. Flow in a convergent channel


For the flow in a convergent channel f must be negative and, therefore, from equation
(4.99) and (4.102) it follows that one of the root should be positive and the other two
must be negative. Let f1 be positive and f2 , f3 be negative. Further, one notes that
f 2  f  0 (it is only in the middle of the channel, where f  f 2 ) ,and f3  f 2 .

Let, Re   f 2 / v , f / f 2  w,  f1 / f 2  w1 and f3 / f 2  w3 , (4.105)


so that the w’s are positive, and w3 lies between 0 and 1.
Substituting equation (4.105) in (4.97), we find
1  w1  w3  6 / Re. (4.106)
Therefore, from equation (4.96)
1
3 dw

2 Re  (1  w)(w 3  w)( w1  w)
1/2

(4.107)
w 
Hence,
1
3 dw
Re    (1  w)(w
1/2
. (4.108)
 w)( w1  w)
1/2
2 0 3

Now, there is no restriction on the upper limit of Re1 2  . In order, however, that Re1 2 
may be large w3 must be nearly equal to 1, and therefore, if 6 Re is neglected, from

equation (4.106) we conclude that w1 is nearly equal to 2. With these values of w1 and

w3 , we find from equations (4.107) and (4.108)


w
3 dw
  
2 Re  (1  w)(2  w)
0
1/2
. (4.109)

Putting, 2  w  3tanh 2  , and integrating (4.109), we get


f vr
 3tanh 2  Re/ 2       1.146   2 ,
1/2
w  (4.110)
f 2  vr max  

satisfying the boundary conditions (4.92).

93
Since tanhx is close to unity when x is close to 2.5, it follows from equation (4.110) that
for large Re the velocity vr will be equal to (vr)max everywhere except in a thin layer near
each wall of thickness proportional to Re-1/2.
Further, from equation (4.91), on integration, we find
p /   2vf / r 2  F (r ). (4.111)
p
Since from equation (4.90) we conclude that r 2 should be a function of  only, F(r)
r
should have the following form:
F (r )   K / 2r 2 . (4.112)

Hence, p /   2vf / r 2  K / 2r 2 . (4.113)


Now, when w1  2 and w3  1 , we find from (4.105) and (4.98) that

K  f 22  r 2 (vr )2 max .

Therefore, p /    f 22 / 2r 2  2vf / r 2 ,

or p /   (vr )2 max / 2  2vf / r 2  0 (vr )2 max / Re. (4.114)

i.e., for large Re the pressure at the walls is equal to the pressure of the main flow.
The results obtained above concerning the velocity and pressure near the walls, for large
Reynolds number, are in exact agreement with the basic assumptions of the theory of
boundary layers, which will be taken up in Chapter 6.

Fig 4.15. Velocity distribution in a convergent and a divergent channel after K.Millsaps and K.Pohlhausen
We shall now discuss, briefly, the character of the velocity distribution, as calculated by
Millsaps and Pohlhausen, which differ markedly for convergent and divergent channels
and for different values of the Reynolds number, A total angle of 100 has been chosen for
the divergence or convergence of the fixed walls. In a convergent channel the velocity

94
distribution for a sufficiently large Reynolds number ( Re  5000 ) remains almost
constant everywhere except near the walls, where it decreases rapidly to the value zero,
exhibiting a boundary layer phenomenon. In a divergent channel the velocity is positive
over the whole cross-section if the Reynolds number is small. However, for large
Reynolds number ( Re  5000 ) it becomes negative near the walls and a back flow
occurs.
It may also be pointed out that Millsaps and Pohihausen have given a numerical solution
for the temperature distribution for the present problem and report that in convergent
channels, for large Reynolds number, there are definite thermal boundary layers near the
walls.

15.3. Stagnation point flows

(a) stagnation in two-dimensional flow (Hiemenz flow)


An exact solution of the flow of a viscous incompressible fluid in the neighbourhood of a
stagnation point in a plane may be obtained by considering the flow at a large distance
from the stagnation point to be the potential flow.

Fig 4.16. Flow in the neighborhood of a stagnation point in a plane (Hiemenz flow)
The velocity and pressure in a potential flow in the vicinity of the stagnation point
x  0, y  0 in a plane is given by
U  ax , V  ay , (4.115)

and p /   U 2  V 2  / 2  p0 /  ,


or p0  p 
2
 
a2 x2  y 2 , (4.116)

where a is constant and p0 is the pressure at the stagnation point.


When viscosity is included, we take the following form of the velocity and pressure
distribution:
u  xf '( y), v   f ( y) ; (4.117)
95

and p0  p 
2
 
a2 x2  F ( y) , (4.118)

where the prime denotes differentiation with respect to y. It can easily be seen that the
equation of continuity ( B1 ), Table 2.2. is identically satisfied and the equations of motion
( B 2 )and ( B3 ), in the absence of external forces, reduce to
f 2  ff   a 2  vf  , (4.119)
1 2
and ff   a F   vf  . (4.120)
2
The boundary conditions for f and F are obtained from the consideration that at y = 0, u =
0 = v and at x = y = 0, p = p0 and at a large distance from the wall, i.e. as y  , u  U
. Hence,
y  0 : f  0, f '  0 and F  0,
y   : f '  a. (4.121)
In order to solve the equation (4.119) for f, let us make the following transformation:

a
 y , f ( y )  av  ( ). (4.122)
v
Substituting (4.122) in equation (4.119), we get
 '''  ''  '2  1  0 , (4.123)
and the corresponding boundary conditions for  are
  0 :   0,  '  0,
   :  '  1, (4.124)
where a prime indicates differentiation with respect to  .
Equation (4.123) was first solved, numerically, by Hiemenz and the solution was later
improved by L.Howarth .Some values of  ,   and   for different values of  are
tabulated here from the paper of Howarth.The dimensionless velocity in the x -direction
is given by
u 1
 f ( y)   ( ) (4.125)
U a
Having determined the value of  ( ) or in other words value of f ( y), the unknown
function F ( y) will be obtained by integrating equation (4.120)

Hence, a 2 F  f 2  2vf ' ,


v 2
or F  (  2 ) , (4.126)
a

96
Table 4.1

a u
y     
v U
0 0 0 1.2326
0.2 0.0233 0.2266 1.0345
0.6 0.1867 0.5663 0.6752
1.0 0.4592 0.7779 0.3980
1.4 0.7967 0.8968 0.2110
1.8 1.1689 0.9568 0.1000
2.0 1.3620 0.9732 0.0658
2.4 1.7553 0.9905 0.0260
2.8 2.1530 0.9970 0.0090
3.0 2.3526 0.9981 0.0051

where the constant of integration vanishes with the help of the boundary conditions
(4.121). Substituting the value of F in equation (4.118) we will get the required pressure
distribution.

Fig 4.17. velocity distribution in plane stagnation point flow

The dimensionless velocity, fig. 4.17, begins to increase linearly at   0 and tends to its
maximum value unity only asymptotically. However, 99% of the maximum velocity is
reached at about  = 2.4, and the corresponding value of y, which we will call as  from
equation (4.158) is

v
  2.4 (4.127)
a
If v is small than  will be small and it can be said that the viscous effect are confined in

a very thin layer near the wall and the thickness of this layer is proportional to v.

97
Secondly, the pressure gradient in the y  direction, as can be seen from equations

(4.118) and (4.126), is also proportional to v and is small if v is small. It will be seen
in Chapter 6 that these two results confirm the basic assumptions of the Prandtl`s
boundary layer theory. An analogous solution for the flow near a stagnation point in an
axially symmetrical case, which we now propose to discuss, was obtained by Homann
and was later improved by Frocessling.
(b) Rotationally symmetrical flow with stagnation point (Homann flow)
In this case, we shall use the cylindrical polar coordinates  r , , z  and put

 
( )  0, and ( )  0 .
 t

Fig 4.18. Rotationally symmetrical flow with stagnation point (Homann flow)

The velocity and pressure distribution in the potential flow in the vicinity of the
stagnation point at r  z  0 are given by
Vr  br , Vz  2bz, (4.128)


and p0  p 
2

b2 r 2  4 z 2 .  (4.129)

The corresponding viscous solutions may be taken as


vr  rf   z  , vz  2 f  z  , (4.130)


and p0  p  b 2  r 2  F ( z )  . (4.131)
2
Substituting equations (4.130) and (4.131) in the equations of motion in Table 2.4, in the
absence of body forces, we get
f 2  2 ff   b2  vf  , (4.132)

and 2 ff   b2 F   vf  , (4.133)
98
where a prime denotes differentiation with respect to z .
The boundary conditions for f and F are obtained from the consideration that at
z  0, vr  vz  0 and r  z  0, p  p0 , also at large distance from the surface i.e as

z  , vr  Vr . Hence
z  0 : f  f   0, F  0,
(4.134)
z : fb
Using the transformation

b
 z, f ( z )  bv ( ), (4.135)
v
Equation (4.132) becomes
 ''' 2 '' 1   '2  0 (4.136)
and the corresponding boundary conditions are
  0 :      0,
(4.137)
:    1.
The solution of equation (4.136) was worked out by Homann and later by other workers.
It is also a particular case of Hartree`s, solutions, which will be taken up in §6.5.

15.4. Exercises

For Problems see end of the Block, page number 112

15.5. Keywords
Couette flow, Jeffery Hamel flow, Hiemenz flow, Homann flow.

15.6. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New Delhi
(1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi (1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2, Oxford
University Press (1970).
6. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).
7. Bansal. J. L., Viscous Fluid Dynamics, Oxford and IBH publishing Co. Pvt. Ltd.,
New Delhi (1977).
8. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).

99
UNIT – 16
Flow due to Rotating Disc

Structure:
16.0. Objectives
16.1. Flow due to a rotating disc (Ka’rma’n flow)
16.2. Variable viscosity plane Couette flow
16.3. Variable viscosity plane Poiseuille flow
16.4. Summery
16.5. Exercises
16.6. Keywords
16.7. References

16.0. Objectives:

At the end of reading this unit students are able to understand steady incompressible flow
with variable viscosity, variable viscosity in plane coutte flow, variable viscosity in place
poiseuille flow.

16.1. Flow due to a rotating disc (Ka’rma’n flow)

Let us consider the flow due to a disc which rotates with an angular velocity  about an
axis perpendicular to it plane, in a fluid otherwise at rest. In order to avoid the edge
effect the disc is considered to be of infinite radius. Due to the action of the centrifugal
forces the fluid near the disc will be thrown outward and to compensate this a flow in the
axial direction towards the disc will follow.

Fig 4.19. Flow due to a rotating disc (Ka’rma’n flow)

100
Thus in this case all the three components of velocity in the cylindrical polar coordinates
will exit and the boundary conditions for such a fluid motion are
z  0 : vr  0, v  r, vz  0,
(4.138)
z  : vr  0, v  0.
We now look for a solution of the Navier-Stokes equations in cylindrical polar
coordinates, in the absence of body forces, for a steady motion which is independent of 
. Then, guided partly by the boundary conditions (4.138) and partly by the dimensional
consideration, we consider the following possibilities for the velocity components:
v   rg( z ),
vr   rf ( z ) or (v )1 2 f ( z ) , (4.139)
vz   rh( z ) or (v )1 2 h( z ).

It may be noted that r and   each have the dimension of velocity. By examining
12

the equation of continuity and the equations of motion it is found that a solution of this
nature is possible only if we take the following form for the velocity and pressure
distribution
vr  rf ( z ), v  rg( z ), vz  (v)1 2 h( z ) ,

and p   v p( z ) . (4.140)
Substituting (4.140) in equations (D1) to (D4), Table 2.4, for the steady motion and under
the absence of external forces, we find
2 f  (v /  )1 2 h  0,
f 2  g 2  (v /  )1 2 f  h  (v / w) f ,
(4.141)
2 fg  (v /  )1 2 g h  (v /  )1 2 g ,
hh   p  (v /  )1 2 h,
where a prime denotes differentiation with respect to z .
In order to remove the constant coefficients (v /  )1 2 and (v /  ) we make the following
transformation:
  ( / v)1 2 z ,
f ( z)  F ( ), g ( z)  G( ), h( z)  H ( ) and p( z)  P( ) . (4.142)
In view of (4.142), the equations (4.141) reduce to
2 F  H   0,
F 2  G 2  F H  F ,
(4.143)
2 FG  G H  G ,
HH    P   H ,
where a prime, now, denotes differentiation with respect to  .
The boundary condition (4.138) are now written as

101
  0 : F  0, G  1, H  0,
(4.144)
   : F  0, G  0.
The solution of the first three equations of (4.143) can be obtained with the help of the
boundary conditions in (4.144) and then the last of equations (4.143) will give P in terms
of H . Von Ka’rma’n was the first to obtain a solution of this system of equations by an
approximate method which was later improved by W.G. Cochran and other workers. We
now propose to discuss the method of Cochran in which a solution is obtained by a series
expansion for F,G, and H near   0 and an asymptotic expansion for large values of 
and then the two expansions are joined at a suitable value of  .
The expansion for small 

1 1 1
F  a1   2  b1 3  b12 4  ....,
2 3 12
1 1
G  1  b1  a1 3   a1b1  1 4  ...., (4.145)
3 12
1 1
H  a1 2   3  b1 4  ....,
3 6
which satisfy the differential equation (4.143) and the boundary conditions at   0 . Also
a1 and b1 are arbitrary constants to be determined.

The expansion for large value of 


For large values of  , first of all we note that by virtue of the conditions at infinity the
middle two equations of (4,143) may be approximated by
F H   F , (4.146)

and GH   G. (4.147)

Thus , on integration, we find that F  and G are both proportional to exp ( H  ) . For

consistency H  must be negative and let is value be c .Then for large 

F    ec , G    ec and H    c. (4.148)

Hence, it is reasonable to look for the expansion of F , G and H , for large values of  , in

powers of e c and such an expansion satisfying the differential equations and the
boundary conditions at infinity are:

F  A1e  c

A12  B12 2c A1 A1  B1 3c
e 
2 2

e
 ....,

2c 2 4c 4

G  B1e  c


B1 A12  B12 e 3c
 ....,
4
12c

H  c 
2 A1  c A12  B12 2c A1 A1  B1 3c
e  e 
2 2

e  ....,
  (4.149)
c 2c3 6c5

102
where A1 , B1 and c and are arbitrary constants to be determined.

We now choose A1 , B1 , c, a1 and b1 in such a way so that F , G, H , F  and G are


continuous where the expansions (4.145) and (4.149) are joined. By considering
sufficient number of terms in each expansion any desired accuracy can be achieved, and
the numerical results as obtained by Cochran are
A1  0.934, B1  1.208, c  0.886, a1  0.510 and b1  0.616 .
He also gave tables and graphs of F , G, H , F  and G . The graphs of F , G and H are
shown in fig. 4.20.
It can be seen from the figure that although F , G and H tend to their limiting values only
asymptotically but for all practical purposes these values are reached at about   5 and
the corresponding value of z , which we will call as  , from equation (4.142), is

  5v   .
1/2
(4.150)

Fig 4.20. velocity distribution near a rotating disc in a fluid otherwise at rest.

If  v   is small then  will be small and once again we will have a boundary layer type

of flow.
Strictly speaking the results obtained here are applicable to an infinite disc only, however,
they may be used to calculate an approximate value of the frictional moment on a finite
disc of radius a , provided a is large in comparison of  .
The circumferential component of the shearing stress on the wall, c.f. (D9) Table 2.4, is
given by
 v 
 z      ,
 z  z 0 (4.151)
  r  v  G(0).
12

103
Hence the frictional moment on one side of the disc is
a
M  
0
r  z 2 r dr

1
 
1/ 2
  a 4 v 3 G (0), (4.152)
2
 
1/ 2
or 2M  0.616  a 4  3 ,

where the value of G(0) is b1 and is equal to -0.616.


The dimensionless moment coefficient is given by
2M 1.232
CM   , (4.153)
1/ 2    a   a 2a
2
Re1/2

where, Re  a 2 w / v (Reynolds number).


For Reynolds number less than 105 about the value of the moment coefficient CM given
by equation (4.153) is in close agreement with the experimental results but for higher
Reynolds number the flow becomes turbulent and the present theory based on the laminar
flow ceases to apply.
It may be pointed out here that Millsaps and Pohlhausen have given a numerical solution
for the associate heat transfer problem.

(B) Steady incompressible flow with variable viscosity

For lubricating oils, which are of great use in engineering problems, the temperature rise
due to friction, event at moderate velocity, is so large that the dependence of viscosity on
temperature becomes important and it can no longer be regarded as constant. In the
present case the velocity field is not independent of the temperature field and hence the
problem is more difficult in comparison to the flow with constant fluid properties even
when a simple configuration for the flow pattern is considered.
Here we shall consider only the case of a two dimensional steady incompressible flow
with variable viscosity. The equations of continuity, momentum and energy are
u v
  0, (4.154)
x y

 u u  p   u     u v  
 u  v      2         , (4.155)
 x y  x x  x  y   y x  

 v v  p   v     u v  
 u  v      2         , (4.156)
 x y  y y  y  x   y x  

 T T    2T  2T 
cp  u v   k  2  2   , (4.157)
 x y   x y 
104
 u 2  v 2   u v 
2

  2            ,
 x   y    x y 
where, (4.158)

The body forces are absent and there is no heat addition


  
 c. f .Table 2.1, put  = const, w  0 and ( )  0 .
 z 

16.2. Variable viscosity plane Couette flow

An exact solution of the equation (4.154) to (4.158) can be obtained for plane Couette
flow, i.e. the flow between two parallels plates one of which is at rest and the other
moving with a uniform velocity U in its own plane and there is no pressure gradient in
the direction of the flow, also the plates are kept at the same constant temperature T0
(say).

Fig 4.21. Plane to Couette flow when the viscosity of the fluid depends on temperature

We assume that the x -axis is placed along the stationary plate and y -axis is being taken
at right angle to it. Since v  0 and there is no pressure gradient in the direction of x
along which the flow takes place, the set of equations (4.154) to (4.158) reduce to
d  du 
 0, (4.159)
dy  dy 
2
d 2T  du 
and k 2      0. (4.160)
dy  dy 
The boundary conditions are
y  0 : u  0, T  T0 ,

y  h : u  U , T  T0 , (4.161)

where h is the distance between the plates,

105
Let us introduce the following non-dimensional quantities:
u*  u / U ,   y / h, T *  T  Tw  / T0 ,  *   / 0 ,

Pr  0c p / k (Prandt number), and Ec  U 2 / c pT0 (Eckert number),

where 0 is the viscosity of the fluid at temperature T0 .


The equation (4.159) and (4.160) take the form

d  * du* 
   0, (4.162)
d  d 
2
d 2T * *  du 
*
and  Ec.Pr.    0, (4.163)
d 2  d 
and the corresponding boundary conditions are
  0 : u*  0, T *  0,
  1: u*  1, T *  0. (4.164)
Integrating equation (4.162) we get
du*
 *
 const   w* , (4.165)
d

 du* 
where,  w*    *  , (4.166)
 d  0

is the dimensionless shearing stress at the stationary plate, to be determined:


Now, equation (4.163), with the help of equation (4.165), can be written as
d 2T *  w*2
 Ec.Pr  0. (4.167)
d 2 *
In order to solve the equation (4.167) we require an empirical relation between viscosity
and temperature. In the present analysis this relation, as suggested by Prandtl on the basis
of experimental results, is taken to be

1  *  eT
*
(4.168) where,
  b / T0 , b has the dimension of temperature and depend on the nature of the fluid.
Substituting (4.168) in equation (4.167), we get
d 2T *
 Ec.Pr  w*2e  T  0.
*
(4.169)
d 2

An exact solution of the equation (4.169) was first given by Nahme and was later studied
by Jain and Bansal in details.
1
Thus, T *  lnC1  2ln cosh (C2  C3 ) , (4.170)

106
2
where, C1  C2 . (4.171)
 Ec.Pr  w*2 2
Now, substituting equations (4.170) and (4.168) in (4.165) and on integration, we find
 w* C1
u*  tanh(C2  C3 )  C4 , (4.172)
C2

where C2 , C3 , C4 and  w* are the unknown constants to be determined.


Applying the boundary conditions (4.164) in equations (4.170) and (4.172) the value of
these unknown constants are obtained as follows:

 
 w*  tanh 1 uh* / 1  uh*2 / uh* 1  uh*2

C2  2 w* uh* 1  uh*2 ,

C3   w* uh* 1  uh*2 ,

C4  1/ 2 , (4.173)

where, uh*   Ec.Pr / 8 (4.174)

Also the value of C1 , from equation (4.171), in terms of uh* can be written as

C1  1  uh*2 . (4.175)

It is important to note that for the case of constant viscosity   0 , therefore, uh*  0 and
then

(4.176)
Hence the expressions in (4.172) and in (4.170), after a little simplification, reduce to
u*   , (4.177)
Ec.Pr
and T*   (1   ), (4.178)
2
which are same as obtained in §4.2 (c.f. equations (4.8) and (4.27)).

107
Fig 4.22. Velocity distribution in variable viscosity plane Couette flow.

The velocity profiles and the shearing stress are shown in figs. 4.22 and 4.23, for different
values of uh* respectively. In the present case the velocity ceases to be linear and the

shearing stress goes on decreasing as uh* increases, in comparison to the case of constant
viscosity flow.

Fig 4.23. Dimensionless shearing stress on the stationary plate in variable viscosity plane Couette flow

16.3. Variable viscosity plane Poiseuille flow

In plane Poiseuille flow the pressure gradient is non-zero, but the plates are kept
stationary. Let the distance between the plates be 2b and the axis of x be in the middle
of the channel.

108
Fig 4.24. Plane Poiseuille flow when the viscosity of the fluid depends on temperature.

Let us also suppose that both the plates are kept at the same constant temperature T0
(say). In such a case the equations (4.154) to (4.158) reduce to
d  du  dp
  , (4.179)
dy  dy  dx
2
d 2T  du 
and k 2      0. (4.180)
dy  dy 
The boundary conditions are
y   b : u  0, T  T0 (4.181)
The solution of the above set of equations, which we now propose to discuss, was
obtained by Hausenblas .
Let us introduce the following non-dimensional quantities.
u*  u / um , x*  x / b,   y / b,  *   / 0 ,

T *  T  T0  / T0 , Pr  0c p / k (Prandt number),

and Ec  um2 / c pT0 (Eckert number) .

where um is the velocity in the middle of the channel, when viscosity is regarded as
constant, i.e.
b2 dp
um   , (c. f .eqn.4.11)
20 dx

0 is the viscosity of the fluid at temperature T0 .


The equations (4.179) and (4.180), now, take the form

d  * du* 
   2, (4.182)
d  d 

109
2
d 2T * *  du 
*
 Ec.Pr.    0, (4.183)
d 2  d 
and the corresponding boundary conditions are
  1: u*  0, T *  0. (4.184)
Integrating equation (4.182), we get
du*
*  2 , (4.185)
d
where the constant of integration vanishes from symmetry considerations.
Substituting equation (4.185) in (4.183), we get
d 2T * 4 Ec.Pr 2
   0. (4.186)
d 2 *
In order to solve the equation (4.222) the viscosity temperature relation, as taken by
Hausenblas, is 1  *  1   T *   * (say), (4.187)
which is in fact a first order approximation of the Prandtl`s relation (4.168). Hausenblas
has also worked out the solution by retaining upto quadratic terms in the expansion of
(4.168) and report that the analysis becomes more complicated but the accuracy is not
much improved.
Substituting (4.187) in (4.186), we get
d 2 *
 4 N 2 *  0, (4.188)
d 2

where, N   .Ec.Pr. (4.189)

The boundary conditions on  * are


  1:  *  1. (4.190)
The series solution of the equation (4.188), satisfying the conditions (4.190), is

 *  A sv  4 , (4.191)
r 0

2
1 (1 4)!  N 
where sv  i  , (4.192)
v !  1 4  v !  2 
1
 
and A    sv  . (4.193)
 v 0 
Now, equation (4,185), with the help of (4.187) and (4.191) becomes

du*
 2 *  2 A  sv 4 (4.194)
d v 0

The solution of equation (4.194), subject to the boundary conditions (4.184), is

110

u*  2 A
sv
 
1   4  2 , (4.195)
v  0  4v  2 

For the case of constant viscosity   0 , therefore, N  0, and it can be easily seen that

0, if v  0,
sv  
1, if v  0,
and A  1. (4.196)
Therefore, the expression in equation (4.195), for the velocity distribution, reduce to
u*  1  2  , (4.197)

which is the same as obtained in §4.2, equation (4.12).


However, for the temperature distribution the following limit is to be evaluated

(4.198)
and is the same as obtained in §4.2, equation (4.35).
In fig.4.25 the velocity profiles for different values of N are plotted. The curve
corresponding to N  0 is for constant viscosity flow and is a parabola. It is clear from
the graph that the velocity at any section increases as N increases and is always greater
than the velocity in constant viscosity flow.
It may be pointed out that Hausenblas, in the same article, has also worked out the
variable viscosity flow in a circular tube.

*
Fig 4.25. Velocity distribution in variable viscosity plane poiseuille flow ( N  Ec, Pr) .

(C) Unsteady incompressible flow with constant fluid properties

In the previous two sections we studied few examples of the exact solutions of the
Navier-Stokes equations for steady incompressible flow. We will now take up the study
111
of some simple unsteady motions. We shall concentrate ourselves only on those
configurations of fluid motion in which, besides time as one of the independent variable,
the velocity field is a function of only one space coordinate.

16.4. Summary:

Navier- Stokes equations of motion are non-linear in character, there is no known general
method to solve these equations. Only in few special cases exact solutions can be
obtained by making certain assumptions about the state of the fluid and a simple
configuration for the flow pattern. Some of these exact solutions have been discussed in
this chapter.

16.5. Exercise:

1. Obtain the velocity distribution for a generalized plane Coutte flow in the form
y y y
u   p 1   .
h h h

2.
 2  2
Prove that      

  ,  2 where  is a stream function for a two
 t    x, y 
dimensional motion of a viscous fluid.
3. Prove that in the steady motion of a viscous fluid in two dimensions
4  dX dy  dY dx where (X,Y) is the impressed force per unit area.
4. Two dimensional potential flow of an inviscid and incompressible fluid near the
stagnation point at the origin at a fixed point taken as y=0 is given u = bx show
that the corresponding problem for a viscous liquid has the solution

u  bx , v  b      b  y where        2  1  0 .

5. Viscous incompressible fluid is steady two dimensional radial motion between
two non-parallel walls r and  q are polar coordinates, r being the distance from
the line of intersection of the walls which are    show that the velocity is
given by qr  f   r where f 2  
2
3

h  2 kf  6 f 2  f 3 where h and k are
constants.
6. Prove that for a plane coutte flow solution for heat conduction equation can be
T  T0 y Ec Pr y  y
obtained in the form   1   .
T1  T0 h 2 h h
7. Discuss the flow of an incompressible viscous fluid between two rotating
concentric cylinders.
8. Determine the average velocity distribution shear stress and drag coefficient for
laminar flow in a plane poiseuille flow.
9. Find the velocity distribution in the steady flow of a viscous incompressible fluid
along an infinitely long circular pipe due to an applied pressure gradient. Also
determine the expressions for volumetric rate of flow and coefficient of friction at
the wall of the pipe.
10. Define poiseuille flow. For such a flow obtain the velocity distribution , average
velocity mass f flow rate and skin friction.
11. Derive the expression for Coutte flow between two concentric rotating cylinders.
112
12. Write short on Jeffery-Hamel flow.
13. Explain stagnation point in two dimensional flow.
14. Obtain the expression for two dimensional steady incompressible flow with
variable viscosity.
15. Write short note on Karman flow.
16. Explain in detail for two dimensional steady incompressible poiseuille flow with
variable viscosity.
17. In a plane poiseuille flow, water at 20oC flows between two large plates at a
distance 1.5 mm apart. If the average velocity is 0.15 m/sec. Evaluate (i) the
maximum velocity (ii) the pressure drop (iii) the wall shearing stress (iv) the
frictional coefficient.
18. In a plane poiseuille flow, water at 70oC flows between two large plates at a
distance 1/16 inch apart. If the average velocity is ½ feet/sec. . Evaluate (i) the
maximum velocity (ii) the pressure drop (iii) the wall shearing stress (iv) the
frictional coefficient.
19. In a plane poiseuille flow, water at 70oC flows between two parallel plates one of
which is at rest and other is moving with a velocity U. If the volumetric flow Q
per unit width is zero, find the relation between U and the pressure gradient in
terms of the dynamic viscosity and the spacing of the plates h. calculate the
pressure gradient fo that a distance 1.5 mm apart. If the average U=1/2 feet/sec
and h=1/2 inch..
20. Incompressible liquid is flowing steadily through a circular pipe prove that the
mean pressure is constant over the cross section and that the rate of flow is
 a 4  p1  p2  8l where p1 and p2 r1 and r2 are pressure over sections at
distance l apart.
21. Determine the maximum value of the velocity profile in the annular space
between two coaxial cylinders.
22. Determine the maximum value of the velocity profile in the annular space
between two coaxial cylinders. If a=50mm, b=75mm and volumetric flow of
water Q=0.006 m3/sec, calculate (i) pressure drop (ii) the maximum value of u
(iii) the shearing stress at the wall of both cylinders. Assume that =1.01g/ms.

16.6. Keywords
Flow in Circular pipe, Velocity Distribution, and Temperature Distribution

16.7. References

1. Bhatnagar. P. L., Elements of Dimensional Analysis, National Pub House, New Delhi
(1970).
2. Binder R. C., Advanced Fluid Mechanics, Prentice-Hall of India, New Delhi (1964).
3. Bird .R. B. et. al., Transport Phenomena, John Wiley, New York (1960).
4. Chorlton. F., Text Book of Fluid Dynamics, ELBS (1970).
5. Goldstein. S. (Ed), Modern Development in Fluid Dynamics, Vol.1 and 2, Oxford
University Press (1970).
6. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).
7. Bansal. J. L., Viscous Fluid Dynamics, Oxford and IBH publishing Co. Pvt. Ltd.,
New Delhi (1977).
8. Dr. M. D. Raisinghania, Fluid Dynamics, S. Chand Publishing 10th Edition, (2011).

113
APPENDIX

Equation of motion in orthogonal curvilinear coordinates


1. Orthogonal curvilinear coordinates
In analyzing many problems of fluid dynamics it is often advantageous to use curvilinear
coordinates instead of the ordinary Cartesian coordinates. We shall now see how
curvilinear coordinates are introduced.
Let the Cartesian coordinates  x, y, z  of any point P in space be expressed as functions

of  ,  ,  

x  x  ,  ,   y  y  ,  ,   z  z  ,  ,   (1)

but with x, y, z completely defined and continuous in the first derivative the above
equations may be solved for  ,  ,  in terms of x, y, z, i.e.

    x, y, z      x, y, z      x, y, z  (2)

Thus to each point P  x, y, z  we can assign a unique set of new coordinates  ,  ,  

called the curvilinear coordinates of P . In this sense the equations (1) or (2) may be
interpreted as defining a transformation of coordinates.
The surface   x, y, z   const.,   x, y, z   const., and   x, y, z   const., are called

coordinates surfaces and each pair of these surfaces intersect in curves called coordinate
curves or lines. The surfaces   const., and   const., intersect in a curve along which
the coordinate ' ' alone varies and therefore known as  curve or line. Similarly we
have  -line and  -line. If the coordinate surfaces intersect at right angles, the
curvilinear coordinates system is called orthogonal. We shall restrict ourselves to
orthogonal coordinates only as required in the text.
At the point P we draw tangents to each of the coordinate lines. These tangents, which
are mutually perpendicular in orthogonal system, are taken as the coordinate axes at the
point P. The axes are chosen positive in the direction in which  ,  ,  increase from
  
the point . Along the coordinate axes we take unit vectors e1 , e2 , e3 respectively. It should
be noted that the directions of these unit vectors vary, in general from point to point and
in this sense they are not constant vectors.
The line element ds in Cartesian coordinates is given by

 ds    dx    dy    dz 
2 2 2 2
(3)

but from (1) we have

114
x x x
dx  d  d  d , (4)
  
and similar two results.
Substituting the expression for dx, dy, and dz by and from (4), in (3) we find that the
line element in orthogonal curvilinear coordinates is given by.

 ds   h12  d   h22  d    h32  d   ,


2 2 2 2

(5)

 x   y   z 
2 2 2

where h   2
     ,
        
1

2 2 2
 x   y   z 
h22        ,
        
2 2 2
 x   y   z 
h32          ,
         (6)
h1 , h2 , h3 are known as scale factors.
The gradient, divergence and curl in orthogonal curvilinear are given by the values of
and can be obtained by cyclic permutation.
1  1  1  
 , , , (7)
 h1  h2  h3  

if V   v1 , v2 , v3  (8)

1     
then divV    h2 h3v1    h3h1v2    h1h2v3 
h1h2 h3      (9)
also   curl V  1 , 2 , 3 
(10)
1   
where 1    h3v3    h2v2   ,
h2 h3     (11)
The values of 2 and 3 can be obtained by cyclic permutations.
In short
h1eˆ1 h2 eˆ2 h3eˆ3
1   
1 , 2 , 3  
h1h2 h3   
h1v1 h2 v2 h3v3
(12)

The components of V   are still given by


v23  v32 , v31  v13 , v12  v21
(13)
The rate of strain components are (see Love`s Mathematical theory of elasticity)

115
1 v1 v2 h1 v3 h1
e    ,
h1  h1h2  h3 h1 
1 v2 v h2 v h2
e   3  1 ,
h2  h3 h2  h2 h1 
1 v3 v h3 v h3
e   1  2 ,
h3  h3 h1  h3h2 
h3   v3  h2   v2 
e     ,
h2   h2  h3   h2 
h1   v1  h3   v3 
e     ,
h3   h1  h1   h3 
h2   v2  h1   v1 
e     ,
h1   h2  h2   h1 
(14)
The relationship between stress components and the rate of strain components for an
isotropic Newtonian incompressible fluid is still given by
    p  2 e ,        e ,
    p  2 e ,        e ,
    p  2 e ,        e .
(15)
where   , for example, is the component of the stress at  ,  ,   in the direction of 

increasing across the surface   const. The dissipation function  , in orthogonal


curvilinear coordinates, in the case of an incompressible fluid is given by


   2  e
2
 e
2
 e2   e
2
 e
2
 e
2
.  (16)
The above formulae will now be used to write the equations governing the motion of a
viscous incompressible fluid in cylindrical and spherical polar coordinates.
2. Governing equations
We have seen in Chapter 2 that the equations governing the motion of an incompressible
fluid and the equation of energy are
Equation of continuity : div V  0.
DV
Equation of motion :    F  p   2V
Dt
D 
where   V . .
Dt t
 
using the vector identities

V ...V    12 V .V   V  curl V , (17)

and  
2V  grad div V  curl curl V   (18)
the equation (2.24) may be written as

116
 V 
1 
     V .V   V  curl V
 t 2 
   
   F  p   grad div V  curl curl V


(19)
equation (19) is convenient for transformation to any system of orthogonal curvilinear
coordinates, as it does not involve differential coefficients of unit vectors.
Equation of energy:
DT Q
cp   k2T  ,
Dt t
The dissipation function  , in orthogonal curvilinear coordinates, is given by (16) and
the Laplacian of T is

1     h2 h3 T    h3h1 T    h1h2 T  

T        (20)
   h1     h2     h3   
h1h2 h3  

3. Cylindrical polar coordinates


With cylindrical polar coordinates r , , z such that x  r cos , y  r sin  ,
if r , , z are taken as  ,  ,  respectively , then from (6) h1  1, h2  r and h3  1.

 1   
hence    r , r  , z  .
(21)
also, if V   vr , v , vz  ,

1  1 v v
div V   rvr     z
r r r  z
vr vr 1 v vz
   
r r r  z (22)

The components of the vorticity vector curl V in view of (11) are


1   1 vz v
1    v    rv    ,
r   z  r  z
  v v
2    vr    v   r  z ,
 z r  z r
1    v v 1 vr
3    rv    vr        .
r  r   r r r  (23)

 
Accordingly, the components of the vector curl curl V , i.e. curl  may be deduced from

1    v v 1 vr    vr v 


     r r z  ,
r    r r r   z  z r 
   1 vz v    v v 1 vr 
(23) as         , (24)
 z  r  z  r  r r r  
1    vr v    1 vz 1 v 
 r r z    .
r  r  z r    r  r z 

117
 
The components of grad div V can easily be obtained from (21) and (22) as

  vr vr 1 v vz 


    ,
r  r r r  z 
1   vr vr 1 v  vz
    , (25)
r   r r r   z
  vr vr 1 v vz 
    .
z  r r r  z 
subtracting (24) from (25) component wise, it may be found, after some simplification,

  
that the components of grad div V  curl curl V , 
 v 2 v   2 v 2 vr 
are  2vr  r2  2   ,   v  2  2  ,  vz
2
(26)
 r r    r r  
2 1  1 2 2
where   2
2
  . (27)
r r r r 2  2 z 2
1 
The components of   V .V   V  curl V , in view of (13), (21) and (23) may be
2 
calculated as
 1 2 1 2 1 2  r v 1 vr   vr vz 
 vr  v  vz   v      vz   ,
r  2 2 2   r r r    z r 
1  1 2 1 2 1 2  1 vz v   v v 1 vr 
 vr  v  vz   vz     vr    , (28)
r   2 2 2   r  z   r r r  
 1 2 1 2 1 2  vr vz   1 vz v 
 vr  v  vz   vr     v   .
z  2 2 2   z r   r  z 
which , on simplification become
vr v vr v v 2
vr   vz r   ,
r r r z r
v v v v v v
vr      vz   r  , (29)
r r r z r
vz v vz vz
vr   vz .
r r  z
Now , the equation of continuity follows by equating to zero the expression given in (22)
and the equations of motion, substituting (26) and (29) in (19) component wise, become
 Dvr v2  p  v 2 v 
     f r      2 vr  r2  2   ,
 Dt r  r  r r  
 Dv v v  1 p  v 2 v 
    r     f      2 v  2  2 r , (30)
 Dt r  r   r r  
Dv p
 z   f z    2vz ,
Dt z
D  v  
where  vr   vz , (31)
Dt r r  z
118
and  2 is defined by the equation (27).
The rate of strain components, given in (14) become
vr 1 vz v
err  , e z   ,
r r  z
1 v vr v v
e   , ezr  r  z , (32)
r  r z r
v   v  1 vr
ezz  z , er  r     .
z r  r  r 
The dissipation function  , in cylindrical polar coordinates, may now be obtained from
(16) and (32) as
  v 2  1 v vr   vz      v
2 2
 1 vr 
2

c    2  r
 

     r   r  
  r   r  r   z    r  r  
(33)
 1 vz v   vr vz  
2 2

    
 r  z   z r  

4. Spherical polar coordinates


With spherical polar coordinates r ,  ,  such that
x  r sin  cos , y  r sin  sin , z  r cos 
If r ,  ,  are taken as  ,  ,  respectively, then from (6)
h1  1, h2  r and h3  r sin 
The remaining analysis will be parallel to the one adopted in cylindrical polar
coordinates.

119

You might also like