Thanks to visit codestin.com
Credit goes to www.scribd.com

0% found this document useful (0 votes)
1K views593 pages

Fluid Mechanics-An Introduction

The document is the fourth edition of 'Fluid Mechanics: An Introduction' by Ethirajan Rathakrishnan, published in 2022. It covers fundamental concepts, properties, and laws of fluid mechanics, along with detailed discussions on fluid flow, dimensional analysis, boundary layers, vortex theory, and flow through pipes. The book serves as a comprehensive resource for students and professionals in the field of aerospace engineering and fluid dynamics.

Uploaded by

roal732001
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
1K views593 pages

Fluid Mechanics-An Introduction

The document is the fourth edition of 'Fluid Mechanics: An Introduction' by Ethirajan Rathakrishnan, published in 2022. It covers fundamental concepts, properties, and laws of fluid mechanics, along with detailed discussions on fluid flow, dimensional analysis, boundary layers, vortex theory, and flow through pipes. The book serves as a comprehensive resource for students and professionals in the field of aerospace engineering and fluid dynamics.

Uploaded by

roal732001
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 593

Fourth Edition

FLUID
MECHANICS
An Introduction
Dye Tank

Water

Glass tube

Jet Dye streak

Water flow

Ethirajan Rathakrishnan
FLUID MECHANICS

ISTUDY
ISTUDY
FLUID MECHANICS
An Introduction
FOURTH EDITION

Ethirajan Rathakrishnan
Professor of Aerospace Engineering
Indian Institute of Technology Kanpur

Delhi-110092
2022

ISTUDY
FLUID MECHANICS: An Introduction, Fourth Edition
Ethirajan Rathakrishnan

© 2022 by PHI Learning Private Limited, Delhi. Previous editions © 2012, 2007, 1993.

All rights reserved. No part of this book may be reproduced in any form, by mimeograph or
any other means, without permission in writing from the publisher.

ISBN-978-93-89347-91-3 (Print Book)


ISBN-978-93-89347-92-0 (e-Book)

The export rights of this book are vested solely with the publisher.

Published by Asoke K. Ghosh, PHI Learning Private Limited, Rimjhim House, 111, Patparganj
Industrial Estate, Delhi-110092 and Printed by Syndicate Binders, A-20, Hosiery Complex,
Noida, Phase-II Extension, Noida-201305 (N.C.R. Delhi).

ISTUDY
To
my parents
Thammanur Shunmugam Ethirajan
and
Aandaal Ethirajan

ISTUDY
ISTUDY
Contents
Preface.......................................................................................... xiii
Preface to the Third Edition............................................................xv
Preface to the Second Edition........................................................xvii
Preface to the First Edition........................................................... xix

1.SOME BASIC FACTS ABOUT FLUID MECHANICS............. 1–29


1.1 General Description.................................................................. 1
1.2 Fluids and the Continuum .................................................... 3
1.3 Dimensions and Units.............................................................. 5
1.4 Law of Dimensional Homogeneity........................................... 5
1.5 The Perfect Gas: Equation of State..................................... 6
1.6 Regimes of Fluid Mechanics.................................................... 7
1.7 Fluid Statics ........................................................................... 9
1.7.1 Scalar, Vector, and Tensor Quantities......................... 9
1.7.2 Body and Surface Forces........................................... 10
1.7.3 Forces in Stationary Fluids........................................ 10
1.7.4 Basic Equation of Fluid Statics................................. 12
1.7.5 Hydrostatic Forces on Submerged Surfaces............... 16
1.7.6 Pressure Variation with Elevation for a Static
Compressible Fluid..................................................... 23
1.8 Summary......................................................................... 25
1.9 Problems.......................................................................... 25

2. FUNDAMENTALS OF FLUID MECHANICS....................... 30–172


2.1 Introduction............................................................................ 30
2.2 Properties of Fluids............................................................... 30
2.2.1 Pressure....................................................................... 31
2.2.2 Temperature................................................................ 32
2.2.3 Density........................................................................ 32
2.2.4 Viscosity...................................................................... 33
2.2.5 Absolute Coefficient of Viscosity............................... 33
2.2.6 Kinematic Viscosity Coefficient.................................. 36
2.2.7 Thermal Conductivity of Air..................................... 36
2.2.8 Compressibility............................................................ 37
2.3 Thermodynamic Properties.................................................... 37
2.3.1 Specific Heat............................................................... 38
2.3.2 The Ratio of Specific Heats....................................... 38
vii

ISTUDY
viii CONTENTS

2.4 Surface Tension...................................................................... 39


2.5 Analysis of Fluid Flow.......................................................... 41
2.5.1 Lagrangian and Eulerian Specifications..................... 41
2.5.2 Relation Between Local and Material
Rates of Change......................................................... 42
2.5.3 Graphical Description of Fluid Motion...................... 45
2.6 Basic and Subsidiary Laws for Continuous Media............... 46
2.6.1 Systems and Control Volumes................................... 47
2.6.2 Integral and Differential Analysis.............................. 47
2.6.3 State Equation............................................................ 47
2.7 Kinematics of Fluid Flow...................................................... 48
2.7.1 Rotational and Irrotational Motion............................ 58
2.8 Stream Function.................................................................... 62
2.8.1 Relationship Between Stream Function and
Velocity Potential....................................................... 66
2.9 Potential Flow........................................................................ 69
2.9.1 Bernoulli Equation...................................................... 71
2.9.2 Two-Dimensional Source and Sink............................. 74
2.9.3 Simple Vortex............................................................. 75
2.9.4 Source–Sink Pair......................................................... 78
2.9.5 Doublet....................................................................... 78
2.10 Combination of Simple Flows................................................ 81
2.10.1 Flow Past a Half-Body.............................................. 81
2.11 Flow Past a Circular Cylinder.............................................. 88
2.11.1 Flow without Circulation............................................ 88
2.11.2 Flow with Circulation................................................ 90
2.12 Viscous Flows......................................................................... 94
2.12.1 Drag of Bodies............................................................ 97
2.12.2 Turbulence.................................................................102
2.12.3 Flow Through Pipes..................................................108
2.13 Gas Dynamics.......................................................................116
2.13.1 Perfect Gas................................................................117
2.13.2 Speed of Sound..........................................................117
2.13.3 Flow with Area Change............................................118
2.13.4 Normal Shock Relations............................................124
2.13.5 Oblique Shock Relations............................................126
2.13.6 Flow with Friction....................................................129
2.13.7 Flow with Simple T0-Change....................................133
2.14 Summary........................................................................135
2.15 Problems.........................................................................143

3. DIMENSIONAL ANALYSIS AND SIMILARITY................173–194


3.1 Introduction...........................................................................173
3.2 Dimensionless Groups...........................................................173
3.3 Dimensional Homogeneity Principle......................................174

ISTUDY
CONTENTS ix

3.4 Buckingham’s p-Theorem......................................................177


3.5 Dimensionless Group in Fluid Dynamics.............................177
3.6 Calculation of the Dimensionless Groups.............................178
3.7 Similarity...............................................................................180
3.8 Relationship Between Dimensional Analysis
and Similarity.......................................................................181
3.9 Similarity Requirements from the Equations of Flow.........185
3.10 Scale Factors .......................................................................186
3.11 Summary........................................................................191
3.12 Problems.........................................................................192

4. BOUNDARY LAYER............................................................195–242
4.1 Introduction...........................................................................195
4.2 Boundary Layer Development..............................................196
4.2.1 Velocity Profile............................................................197
4.3 Boundary Layer Thickness...................................................198
4.3.1 Displacement Thickness.............................................198
4.3.2 Momentum Thickness................................................200
4.3.3 Kinetic Energy Thickness..........................................201
4.3.4 Non-Dimensional Velocity Profile..............................201
4.3.5 Types of Boundary Layer.........................................202
4.4 Boundary Layer Flow...........................................................204
4.5 Boundary Layer Solutions....................................................206
4.6 Momentum-Integral Estimates..............................................207
4.6.1 Conservation of Linear Momentum...........................207
4.6.2 Karman’s Analysis of the Flat Plate
Boundary Layer.........................................................208
4.7 Boundary Layer Equations...................................................209
4.8 Flat Plate Boundary Layer..................................................214
4.8.1 Laminar Flow............................................................215
4.8.2 Boundary Layer Thickness........................................216
4.9 Turbulent Boundary Layer for Incompressible Flow
Along a Flat Plate...............................................................224
4.10 Flows with Pressure Gradient..............................................227
4.11 Laminar Integral Theory......................................................228
4.12 Summary........................................................................235
4.13 Problems.........................................................................239

5. VORTEX THEORY...............................................................243–318
5.1 Introduction...........................................................................243
5.2 Vorticity Equation in Rectangular Coordinates...................244
5.2.1 Vorticity Equation in Polar Coordinates...................246
5.3 Circulation................................................................................. 248
5.4 Line (Point) Vortex................................................................... 252
5.5 Laws of Vortex Motion.............................................................. 254

ISTUDY
x CONTENTS

5.6 Helmholtz’s Theorems............................................................... 255


5.7 Vortex Theorems....................................................................... 256
5.7.1 Stokes’ Theorem............................................................ 261
5.8 Calculation of uR, the Velocity due to Rotational Flow............ 265
5.9 Biot–Savart Law........................................................................ 269
5.9.1 A Linear Vortex of Finite Length.................................. 271
5.9.2 Semi-infinite Vortex ..................................................... 272
5.9.3 Infinite Vortex............................................................... 273
5.9.4 Helmholtz’s Second Vortex Theorem............................ 278
5.9.5 Helmholtz’s Third Vortex Theorem.............................. 281
5.9.6 Helmholtz’s Fourth Vortex Theorem............................. 282
5.10 Vortex Motion........................................................................... 282
5.11 Forced Vortex........................................................................... 285
5.12 Free Vortex............................................................................... 287
5.12.1 Free Spiral Vortex......................................................... 288
5.13 Compound Vortex..................................................................... 292
5.14 Physical Meaning of Circulation................................................ 293
5.15 Rectilinear Vortices................................................................... 298
5.15.1 Circular Vortex.............................................................. 298
5.16 Velocity Distribution................................................................. 300
5.17 Size of a Circular Vortex........................................................... 302
5.18 Point Rectilinear Vortex........................................................... 302
5.19 Vortex Pair................................................................................ 304
5.20 Image of a Vortex in a Plane.................................................... 305
5.21 Vortex between Parallel Plates................................................. 305
5.22 Force on a Vortex .................................................................... 307
5.23 Mutual Action of Two Vortices................................................. 308
5.24 Energy due to a Pair of Vortices............................................... 308
5.25 Line Vortex............................................................................... 310
5.26 Summary...................................................................................311
5.27 Problems....................................................................................317

6. FLOW THROUGH PIPES....................................................319–388


6.1 Introduction...........................................................................319
6.2 Flow in Circular Pipes.........................................................319
6.3 Laminar to Turbulent Transition.........................................320
6.4 Fully Developed Laminar Flow in a Pipe...........................325
6.4.1 Shear Stress Distribution...........................................328
6.4.2 Volume Flow Rate.....................................................328
6.4.3 Average Velocity........................................................329
6.4.4 Point of Maximum Velocity......................................329
6.5 Shear Stress Distribution in Fully Developed
Pipe Flow .......................................................................... 332
6.6 Head Loss Due to Friction...................................................333
6.7 Poiseuille’s Law.....................................................................335

ISTUDY
CONTENTS xi

6.8 Friction Factor Variation.....................................................336


6.9 Velocity Distribution in Turbulent Flow.............................343
6.10 Velocity Distribution in Smooth Pipes................................345
6.10.1 Friction Factor for Smooth Pipes.............................349
6.10.2 Velocity Distribution and Friction Factor for
Rough Pipes...............................................................355
6.10.3 Universal Features of the Velocity Distribution
in Turbulent Flow ................................................. 359
6.11 Energy Considerations...........................................................360
6.11.1 Kinetic Energy Coefficient.........................................362
6.11.2 Head Loss..................................................................362
6.12 Minor Losses.........................................................................365
6.12.1 Inlets and Exits.........................................................365
6.12.2 Contractions and Enlargements.................................368
6.12.3 Loss in Bends............................................................372
6.12.4 Valves and Fittings...................................................372
6.13 Noncircular Ducts.................................................................377
6.14 Pipe Flow Solution...............................................................380
6.14.1 Single-Path Systems...................................................380
6.15 Multiple-Path Systems..........................................................381
6.16 Summary...............................................................................381
6.17 Problems................................................................................386

7. FLOW WITH FREE SURFACE...........................................389–486


7.1 Introduction...........................................................................389
7.2 Steady-Flow Energy Equation for Open Channels..............391
7.2.1 Energy Gradient .................................................... 393
7.3 Steady Uniform Flow............................................................394
7.4 Boundary Layer in Open Channels......................................398
7.4.1 Partially Filled Flow in Closed Conduits.................402
7.5 Waves and Surges in Open Channels..................................405
7.6 Specific Energy and Alternative Depth of Flow..................411
7.6.1 Dimensionless Forms of Specific-Energy Curves.......420
7.7 The Hydraulic Jump.............................................................424
7.7.1 Force on an Obstacle in a Stream...........................432
7.8 Critical Flow.........................................................................435
7.9 The Broad-Crested Weir.......................................................436
7.9.1 Drowned Weir and Free Outfall...............................443
7.10 Rapid Flow Approaching a Weir or Other Obstruction.....444
7.10.1 Flow Through Venturi..............................................447
7.11 Gradually Varied Flow.........................................................451
7.12 Governing Equations of Gradually Varied Flow..................452
7.13 Classification of Surface Profile............................................459
7.13.1 The Basic Equations of Motion................................461
7.14 Gravity Waves......................................................................465

ISTUDY
xii CONTENTS

7.15 Capillary Waves....................................................................466


7.16 Motion of Individual Particles..............................................466
7.16.1 Wave Energy..............................................................467
7.17 Energy Transmission Rate....................................................469
7.18 Group Velocity......................................................................470
7.19 Waves Moving into Shallower Liquid...................................472
7.19.1 Standing Waves..........................................................473
7.20 Summary...............................................................................475
7.21 Problems................................................................................482

8. HYDRAULIC MACHINERY—PUMPS AND TURBINES.....487–563


8.1 Introduction...........................................................................487
8.2 Pumps ...................................................................................487
8.3 Head Developed by a Pump.................................................490
8.3.1 Effect of Blade Angle on Pump Head......................494
8.4 Efficiency of Pump...............................................................499
8.5 Similarity Laws for Pumps...................................................505
8.6 Performance Characteristics of Pumps at
Constant Speed.....................................................................507
8.7 Performance Characteristics at Different Speeds
and Sizes...............................................................................509
8.8 Operating Point of a Pump.................................................510
8.9 Specific Speed of Pumps......................................................512
8.10 Peripheral-Velocity Factor.....................................................515
8.11 Cavitation in Pumps.............................................................516
8.12 Selection of Pumps...............................................................520
8.13 Pumps Operating in Series and in Parallel.........................521
8.14 Hydraulic Turbines................................................................521
8.15 Impulse Turbine....................................................................522
8.16 Impulse Turbine Operation...................................................525
8.17 Head and Efficiency of Impulse Turbine..............................532
8.18 Nozzles for Impulse Turbines...............................................539
8.19 Reaction Turbine...................................................................539
8.20 Performance of Reaction Turbines........................................541
8.21 Draft Tubes and Effective Head on Reaction Turbines......548
8.22 Efficiency of Turbines...........................................................549
8.23 Similarity Law for Reaction Turbine....................................551
8.24 Peripheral-Velocity Factor and Specific Speed of
Turbines.................................................................................553
8.25 Cavitation in Turbines..........................................................554
8.26 Pump Turbines......................................................................555
8.27 Summary...............................................................................555
8.28 Problems................................................................................560

Bibliography.................................................................... 565

Index........................................................................567–571

ISTUDY
Preface
The third edition of this book, developed to serve as text for a course
fluid mechanics at the introductory level for undergraduate course and
for an advanced level course at graduate level was well received all over
the world, because of its completeness and proper balance of theoretical
and application aspects of this science.
Over the years, the feedback received from the faculty and students
made the author to realize the need for adding following material to
serve as text for students of all branches of engineering. Considering
the feedback from faculty and students the following material is added
in this edition.
• Three new chapters; Pipe Flows, Flow with Free Surface and
Hydraulics Machinery have been added to this edition.
• Large number of solved examples are included in all the chapters to
enable the user to gain an insight in to the theory and application
aspects of the concepts introduced.
I would like to thank the faculty and students all over the world for
adopting this book for their courses. I thank my doctoral and masters
students, for checking the material added in this edition and the Solution
Manual.
For instructors, a companion Solutions Manual that contains solutions
to all the end-of-chapter problems is available from the publisher.

Ethirajan Rathakrishnan

xiii

ISTUDY
ISTUDY
Preface to the Third Edition
My sincere thanks to the students and instructors who adopted this book
for their courses. In this edition, the subject matter has been given a
fine tuning, clarifying the vital aspects of the processes associated with
potential and viscous flows. This exercise is made to make the book
effective for both theory and application. Few new examples are added.
Some new problems along with answers are added at the end of
Chapter 4. A new chapter on Vortex Theory is added, beginning from
the definition of vortex and covering all the fundamental and application
aspects of the vortices, which play a dominant role in dictating the
performance of almost all engineering devices.
For instructors only a companion Solutions Manual, that contains
typed solutions to all the end-of-chapter problems, is available from
PHI Learning. I am grateful for the financial support extended by the
Continuing Education Centre of the Indian Institute of Technology
Kanpur, for the preparation of the manuscript.
My sincere thanks to my undergraduate and graduate students at
Indian Institute of Technology Kanpur, who are directly and indirectly
responsible for the development of this book.

Ethirajan Rathakrishnan

xv

ISTUDY
ISTUDY
Preface to the Second Edition
This book was originally written as an attempt to provide an overall
view of fluid mechanics in a concise form. To make this book simple
and easy to understand, many rigid proofs of mathematical formulae
are omitted, and simplifications made to others. Also, detailed tables
of experimental data have been avoided as far as possible; the students
are encouraged to acquaint themselves with these at a later stage by
consulting standard reference books. On the other hand, in this book,
considerable attention has been paid to explain the limitations of any
derived equations. Chapter 2 has been completely revised to include the
vital aspects of potential flow, vortex motion and pipe flow. A new chapter
(Chapter 4) on boundary layer theory has been added to this edition.
Throughout the book, considerable emphasis is placed on the physical
phenomena of fluid flows, and their limitations of applicability are
stressed. A large number of solved numerical examples are presented to
demonstrate the application of basic principles. Problems with answers
are provided at the end of each chapter to provide the students with an
exercise to check and augment their understanding of the fundamental
principles of the subject. A list of selected references is given to serve as
a guide for those students who wish to study in more detail the various
branches of fluid mechanics.
In this revised augmented edition, special attention has been given to
the second chapter. Direct definitions and descriptions of the concepts
introduced are expected to provide a valuable insight into the subject
in an easy but effective manner.
For instructors only, a companion Solutions Manual is available from
Prentice-Hall of India that contains typed solutions to all of the end-
of-chapter problems. The financial support extended by the continuing
Education Center of Indian Institute of Technology Kanpur for the
preparation of the Solutions Manual is gratefully acknowledged.
I deeply appreciate the many comments and suggestions that I received
from the users of the first edition of this book. My sincere thanks go
to my doctoral students Professor V.N. Sukumar, Shibu Clement, P.
Lovaraju and B.R. Vinoth and masters students Amit Kumar, Mohan
Murali and Jayaprakash for their help during the preparation of this edition.
The editorial and production staff at Prentice-Hall of India have
been a great help and I sincerely thank them.
Ethirajan Rathakrishnan
xvii

ISTUDY
ISTUDY
Preface to the First Edition
Fluid Mechanics is a basic science that deals with the motion of fluids
such as gases and liquids and has long been considered as an essential part
of engineering education all over the world. This concise and condensed
book is intended for use by students and practising engineers to have
an overall view of the subject in a short span of time.
The entire spectrum of the subject is briefly covered in this book,
with the necessary explanations on every aspect. This approach is meant
to arouse the interest in the subject in the minds of the readers.
Fluid mechanics is often perceived as a difficult subject. However, in
my opinion, it is a simple subject, and an observant mind approaching
it with proper perspective should have no difficulty in understanding
it. For, the basic laws involved in any fluid flow analysis are just four:
conservation of mass, conservation of momentum, conservation of energy,
and the second law of thermodynamics.
The material covered in this book is so designed that any beginner
can follow it easily. The order of coverage followed is such as to enable
the reader to get a complete picture of the subject after having gone
through the material covered in the text.
Diagrams are used wherever necessary to elucidate the concepts, which
cannot be effectively explained otherwise. The examples given should be
of interest in understanding the concepts covered.
I would like to express my sincere thanks to my Ph.D. students
K. Srinivasan, Ignatius John and Himanshu for their help and suggestions
while preparing the manuscript. Further, I wish to thank Prof. S.
Elangovan, Department of Mechanical Engineering, Gurunanak Dev
Engineering College, Bidar, Karnataka, for his valuable suggestions.
The financial support given by the Continuing Education Centre of
the Indian Institute of Technology Kanpur, for the preparation of the
manuscript is gratefully acknowledged.

Ethirajan Rathakrishnan

xix

ISTUDY
ISTUDY
Chapter 1

Some Basic Facts About


Fluid Mechanics

1.1 GENERAL DESCRIPTION


Fluid Mechanics is a science of fluid flow where the temperature changes
encountered are less than five per cent. This corresponds to speeds up to
650 kmph or Mach numbers up to 0.5, under standard sea level conditions.
Thus, for the study of fluid mechanics, among the four main governing
equations—continuity, momentum, energy equations and increase-of-entropy
principle—the energy equation assumes no significance. On the other hand,
when the temperature change associated with a flow is more than five per
cent, the flow science is termed gas dynamics, for which the energy equation
is of primary importance. Thus, fluid mechanics essentially deals with sub-
sonic flows up to Mach 0.5 and which is incompressible up to Mach 0.3 and
compressible from Mach 0.3 to 0.5.
Fluid Mechanics, the subject dealing with the investigation of the
motion and equilibrium of fluids, is one of the oldest branches of physics and
the foundation for the understanding of many other aspects of the applied
sciences and engineering. It is a subject of enormous interest in numerous
fields like biology, biomedicine, geophysics, meteorology, physical chemistry,
plasma physics, and almost all branches of engineering.
Nearly two hundred years ago, man thought of laying down scientific rules
governing the motion of fluids, water and air, mainly to use the rules to under-
stand these elements so that he could not only protect himself from their fury
during natural calamities like cyclone and floods but also utilize their pow-
ers to develop fields like civil engineering and naval architecture. In spite of
this common origin, two distinct schools of thought gradually developed. On
the one hand, through the concept of an ideal fluid, mathematical physicists
developed the theoretical science, known as classical hydrodynamics. On the
other hand, realizing that idealized theories were of no practical application

ISTUDY
2 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

without empirical correction factors, engineers developed from experimental


studies the applied science known as hydraulics, for the specific fields of irri-
gation, water supply, river flow control, hydraulic power, and so on. Further,
the development of aeronautical, chemical and mechanical engineering during
the last few decades, and the exploration of space from 1960s have increased
the interest in the study of fluid mechanics. Thus, it now ranks as one of the
most important basic subjects in engineering science.
The science of Fluid Mechanics has been extended into fields like regimes
of hypervelocity flight and flow of electrically conducting fluids. This has
introduced new fields of interest such as hypersonic flow and magneto-fluid
dynamics. In this connection, it has become essential to combine the
knowledge of thermodynamics, heat transfer, mass transfer, electromagnetic
theory and fluid mechanics for the complete understanding of the physical
phenomenon involved in any flow process.
Fluid dynamics is one of the rapidly growing basic sciences whose principles
find application even in daily life. For instance, the flight of a bird in air and
the motion of fish in water are governed by the fluid dynamic rules. The
design of various types of aircraft and ships are based on the principles of
fluid dynamics. Even natural phenomena like tornadoes and hurricanes can
also be explained by the science of fluid dynamics. In fact, the science of
fluid dynamics dealing with such natural phenomena has been developed to
such an extent that they can be predicted well in advance. Since the earth is
surrounded by an environment of air and water to a very large extent, almost
everything that happens on earth and its atmosphere, in some way or the
other, is associated with the science of fluid dynamics.
This science is referred to as the Mechanics of Fluids, an allied subject
of the mechanics of solids and engineering materials, and is built on the same
fundamental laws of motion. Therefore, unlike empirical hydraulics it is based
on the physical principles, and has close correlation with experimental studies
which both complement and substantiate the fundamental analysis, unlike the
classical hydrodynamics which is based purely on mathematical treatment.
For understanding the fluid flows, it is essential to know the properties of
fluids. Before discussing the fluid properties, it will be useful if the differences
between solids and fluids are understood. From the basic studies on physics,
it is known that solids, liquids, and gases are the three states of matter. In
general, the liquids and gases are called fluids. It can be shown that this
division into solid and fluid states constitutes a natural grouping of matter
from the standpoint of internal stresses and strains in elastic media, that is, the
stress in a linear elastic solid is proportional to strain, while the stress in a fluid
is proportional to its time rate of strain. In fact, among fluids themselves, only
some fluids exhibit the above stated stress–strain relation, and they are called
newtonian fluids. The fluids which do not obey this stress–strain relation are
termed non-newtonian fluids. The above mentioned behaviour of solids and
fluids may also be expressed in a simpler way as follows:

ISTUDY
1.2. FLUIDS AND THE CONTINUUM 3

• When a force is applied to a solid, deformation will be produced in the


solid. If the force per unit area, namely the stress, is less than the yield
stress, i.e. within the proportional limit of the material, the deformation
disappears when the applied force is removed. If the applied stress is
more than the yield stress of the material, it will acquire a permanent
setting or may even break.
• If a shearing force is applied to a fluid, it will deform continuously as
long as the force is acting on it, regardless of the magnitude of the force.

This difference in behaviour between solids and liquids can be explained


by their molecular properties. The existence of very strong intermolecular
attractive forces in solids lends them rigidity. These forces are comparatively
weaker in liquids and extremely small in gases. This characteristic enables the
liquid molecules to move freely within a liquid mass while still maintaining
a close proximity to one another, whereas the gas molecules have freedom to
the extent of completely filling any space allotted to them.
The study of all the aspects of fluid behaviour can be divided into:

1. Statics: This study deals with fluid elements, at rest with respect to
one another and therefore is free of shearing stresses. The static pres-
sure distributions in a fluid and on bodies immersed in a fluid can be
determined from a static analysis.
2. Kinematics: This study deals with the translation, the rotation and
the rate of deformation motion of a fluid element and with the analy-
sis of flow patterns. However, the velocity and acceleration of the fluid
elements cannot be obtained from kinematic study alone, since the inter-
action of fluid elements with one another makes the fluid a distributed
medium.
3. Dynamic analysis: This study deals with the determination of the
effects of the fluid and its surroundings on the motion of the fluid. This
involves the consideration of forces acting on the fluid elements in motion
with respect to one another. Since there is relative motion between
fluid elements, shearing forces must be taken into consideration in the
dynamic analysis.

1.2 FLUIDS AND THE CONTINUUM


Fluid flows may be modelled either on a macroscopic level or on a micro-
scopic level. The macroscopic model regards the fluid as a continuum and
the description is in terms of the variations of the macroscopic velocity, den-
sity, pressure and temperature with distance and time. On the other hand,
the microscopic or molecular model recognizes the particulate structure of a
fluid as a myriad of discrete molecules and ideally provides information on
the position and velocity of every molecule at all times.

ISTUDY
4 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

The description of a fluid motion essentially involves a study of the


behaviour of all the discrete molecules which constitute the fluid. In liquids,
the strong intermolecular cohesive forces make the fluid behave as a continuous
mass of substance and, therefore, these forces need to be analyzed by the
molecular theory. Under normal conditions of pressures and temperatures,
even gases have a large number of molecules in unit volume (e.g. under normal
conditions, for most gases the molecular density is 2.7 × 1025 molecules per
m3 ) and, therefore, they also can be treated as a continuous mass of substance
by considering the average effects of all the molecules within the gas. Such a
fluid model is called continuum.
The continuum approach must be used only where it may yield reasonably
correct results. For instance, this approach breaks down when the mean free
path, the average distance travelled by the molecules between two successive
collisions, of the molecules is of the same order of magnitude as the smallest
significant length in the problem being investigated. Under such circumstances,
the detection of meaningful and gross manifestation of molecules is not possible.
The action of each molecule or a group of molecules is then of significance and
must be treated accordingly.
To understand this, it is essential to investigate the action of a gas on
an elemental area inside a closed container. Even if the quantity of gas is
assumed to be small, innumerable collisions of molecules on the surface result
in the gross, non-time-dependent manifestations of force. That is, the gas
acts like a continuous substance. On the other hand, if only a tiny amount of
gas is kept in the container so that the mean free path is of the same order of
magnitude as the sides of the area element, an erratic activity is experienced,
as individual or groups of molecules bombard the surface. This cannot be
treated as a constant force, but one must deal with an erratic force variation,
as shown graphically in Figure 1.1.

FIGURE 1.1 Force variation with time.

A continuous distribution of mass cannot exhibit this kind of variation.


Thus, it is seen that in the first case the continuum approach would be appli-
cable but in the second case the continuum approach would be questionable.
From Figure 1.1, it is clear that when the mean free path is large in compari-
son with some characteristic length, the gas cannot be considered continuous

ISTUDY
1.3. DIMENSIONS AND UNITS 5

and hence must be analyzed on the molecular scale. The mean free path,
the statistical average distance which molecules travel between collisions, of
atmospheric air is between 50 nm and 70 nm. The other factor which influ-
ences the molecular activities of a gas is the elapsed time between collisions.
The elapsed time must be sufficiently small so that the random statistical
nature of the molecular activity is preserved.
This book deals only with continuous fluids. Further, it will be assumed
that the elastic properties are the same at all points in the fluid and are
identical in all directions from any specified point. These stipulations make
the fluid both homogeneous and isotropic.

1.3 DIMENSIONS AND UNITS


In fluid dynamics, mostly the gross, measurable molecular manifestations
such as pressure and density as well as other equally important, measurable
abstract entities, e.g. length and time, will be dealt with. These manifesta-
tions which are characteristic of the behaviour of a particular fluid, and not of
the manner of flow, may be called fluid properties. Density and viscosity are
examples of fluid properties. In order to adequately discuss these properties, a
consistent set of standard units must be defined. Table 1.1 gives the common
systems of units and their symbols.

TABLE 1.1 Common systems of units


Quantity Unit SI CGS FPS MKS
Mass kilogram kg g lb kg
Length metre m cm ft m
Time second s s s s
Force newton N = kg·m/s2 dyne pdl kgf
◦ ◦ ◦
Temperature kelvin K C F C

Throughout this book we shall use the SI system of units. However, the
other systems of units as mentioned in Table 1.1 are equally applicable to all
the equations.

1.4 LAW OF DIMENSIONAL


HOMOGENEITY
This law states: “An analytically derived equation representing a physical
phenomenon must be valid for all systems of units”.√ Thus, the equation for
the frequency of a simple pendulum, f = (1/2π) g/l, is properly stated for
any system of units. This explains why all natural phenomena proceed com-
pletely in accordance with man-made units, and hence fundamental equations
representing such events should have validity for any system of units. Thus the
fundamental equations of physics are dimensionally homogeneous, and conse-
quently all relations derived from these equations must also be dimensionally

ISTUDY
6 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

homogeneous. For this to occur under all systems of units, it is necessary that
each grouping in an equation must have the same dimensional representation.
Examine the following dimensional representation of an equation:
L = T2 + T
where L denotes length and T the time. Changing the units of length from
feet to metres will change the value of the left-hand side while not affecting
the right-hand side, thus making the equation invalid in the new system of
units. Dimensionally homogeneous equations only will be considered in this
book.

1.5 THE PERFECT GAS: EQUATION


OF STATE
Gases are basically divided into two broad categories: perfect gases and real
gases.
Perfect gas is a gas in which the intermolecular forces are negligible.
Real gas is a gas where the intermolecular forces are important and must
be accounted for.
In this book, we are concerned only with the fluids which can be regarded
as perfect gases. For perfect gases, the kinetic theory of gases indicates that
there exists a simple relation between pressure, specific volume, and absolute
temperature. For a perfect gas at equilibrium, this relation has the following
form:

pv = RT (1.1)
Equation (1.1) is called the ideal gas equation of state or simply the ideal
gas relation, and a gas which obeys this relation is called an ideal gas. In
this equation p is the absolute pressure, T is the absolute temperature, and
v is the specific volume. The gas constant R is different for each gas and is
determined from
Ru
R= [kJ/kg · K or kPa · m3 /kg · K]
M
where Ru is the universal gas constant and M is the molar mass (also called
the molecular weight).
The constant Ru is same for all substances and its value is



 8.314 [kJ/kmol · K]



 8.314 [kPa · m3 /kmol · K]

0.08314 [bar · m3 /kmol · K]
Ru =

 1.986 [Btu/lbmol · R]



 10.73 [psia · ft3 /lbmol · R]

1545.00 [ft · lbf/lbmol · R]

ISTUDY
1.6. REGIMES OF FLUID MECHANICS 7

The molar mass M can be simply defined as the mass of one mole of
a substance in grams, or the mass of one kmol in kilograms. It is essential
to realize that an ideal gas is an imaginary substance that obeys the relation
pv = RT . It has been experimentally observed that the ideal gas relation given
above closely approximates the p–v–T behaviour of real gases at low densities.
At low pressures and high temperatures, the density of a gas decreases, and
the gas behaves as an ideal gas under these conditions.
In the range of practical interest, many familiar gases such as air, nitrogen,
oxygen, hydrogen, helium, argon, neon, krypton and even heavier gases such
as carbon dioxide can be treated as ideal gases with negligible error (often less
than 1%). However, dense gases such as water vapour in steam power plants
and refrigerant vapour in refrigerators should not be treated as ideal gases.
Essentially, the perfect gases are those which have constant specific heats
and obey the perfect gas law
p
= pv = RT
ρ

This law relates the various gas properties at a particular state; it is known
as the equation of state and as property relation. Perfect gases are sometimes
called the ideal gases. One should not confuse a perfect (ideal) gas with an
ideal fluid.
An ideal fluid is usually defined as a fluid in which there is no friction;
it is inviscid (its viscosity is zero). Thus the internal forces at any section
within it are always normal to the section, even during motion. Therefore,
these forces are purely pressure forces. Although such a fluid does not exist in
reality, many fluids approximate frictionless flow at sufficient distances from
solid boundaries, and so we can often conveniently analyze their behaviour by
assuming them to be an ideal fluid.
In a real fluid, either liquid or gas, tangential or shearing forces always
develop whenever there is motion relative to a body, thus creating fluid fric-
tion, because these forces oppose the motion of one particle (molecule) past
another. These friction forces give rise to a fluid property called viscosity.

1.6 REGIMES OF FLUID MECHANICS


Based on the fluid properties which characterize the physical situation, the
flows are classified into various types as follows:

Ideal fluid flows


It is only an imaginary situation where the fluid is assumed to be inviscid
or non-viscous and incompressible. Therefore, there is no tangential force
between the adjacent fluid layers. An extensive mathematical theory is avail-
able for the ideal fluid. Although the theory of ideal fluids fails to account
for viscous and compressibility effects in actual fluid flow processes, it gives

ISTUDY
8 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

reasonably reliable results in the calculation of lift, induced drag and wave
motion for gas flow at low velocity and for water. This branch of fluid
dynamics is called classical hydrodynamics.

Viscous incompressible flows


The theory of viscous incompressible fluids assumes fluid density to be con-
stant. It finds widespread application in the flow of liquids and the flow of air
at low velocity. The phenomena involving viscous forces, flow separation and
eddy flows are studied with the help of this theory.

Gas dynamics
The theory of gas dynamics deals with the dynamics and thermodynamics of
the flow of a compressible fluid. Based on the dimensionless velocity, namely
Mach number M , defined as the ratio of flow velocity and the local speed of
sound, gas dynamics can be further divided into the fields of study commonly
referred to as subsonic (M < 1), transonic (M ≈ 1), supersonic (1 < M < 5),
and hypersonic (M > 5) gas dynamics.

Rarefied gas dynamics


The concept of continuum fails when the mean free paths of fluid molecules
are comparable with some characteristic geometrical length in the flow field.
A dimensionless parameter, the Knudsen number Kn, defined as the ratio
of mean free path to a characteristic length, aptly describes the degree of
departure from continuum flow. Based on the Knudsen number, the flow
regimes are grouped as follows:
Continuum (Kn < 0.01): All equations of viscous compressible flow are
applicable in this regime. The no-slip boundary condition is valid.
Slip flow (0.01 < Kn < 0.1): Here again the continuum fluid dynamic
analysis is applicable provided the slip boundary conditions are employed.
That is, the no-slip boundary condition of continuum flows, dictating zero
velocity at the surface of an object kept in the flow, is not valid. The fluid
molecules move (slip) with a finite velocity, called the slip velocity, at the
boundary.
Transition flow (0.10 < Kn < 5): In this regime of flow, the fluid
cannot be treated as continuum. At the same time, it cannot be treated
as a free molecular flow since such a flow demands the intermolecular force of
attraction to be negligible. Hence, it is a flow regime between continuum and
free molecular. The kinetic theory of gases must be employed to adequately
describe this flow.
Free molecular flow (Kn > 5): In this regime of flow the fluid molecules
are so widely dispersed that the intermolecular forces can be neglected.
All these regimes of rarefied gas dynamics or super aerodynamics are
encountered at high altitudes, where the molecular density is very low. This
branch of fluid flow is also called the low-density flow.

ISTUDY
1.7. FLUID STATICS 9

Magnetofluidmechanics
The subject of magnetofluidmechanics is an extension of fluid mechanics
with thermodynamics, mechanics, materials and the electrical sciences. This
branch was initiated by astrophysicists. The other names which are used to
refer to this discipline are magnetohydrodynamics, magnetogasdynamics and
hydromagnetics.
Magnetofluidmechanics is the study of the motion of an electrically charged
conducting fluid in the presence of a magnetic field. The motion of the elec-
trically conducting fluid in the magnetic field will induce electric currents in
the fluid, thereby modifying the field. The flow field will also be modified by
the mechanical forces produced by it. The interaction between the field and
the motion makes magnetofluiddynamic analysis difficult.
A gas at normal and moderately high temperatures is a nonconductor. But
at very high temperatures of the order of 10,000 K and above, thermal excita-
tion sets in. This leads to dissociation and ionization. Ionized gas is called a
plasma, which is an electrically conducting medium. Electrically conducting
fluids are encountered in engineering problems like re-entry of missiles and
spacecraft, plasma jet, controlled fusion research and magnetohydrodynamic
generator.

Flow of multicomponent mixtures


This field is simply an extension of basic fluid mechanics. The analysis of
flow of the homogeneous fluid consisting of a single species, termed basic fluid
mechanics, is extended to study the flow of chemically reacting component
mixtures, made of more than one species. All the three transports, namely
the momentum transport, the energy transport, and the mass transport, are
considered in this study, unlike the basic fluid mechanics where only the trans-
ports of momentum and energy are considered.

Non-newtonian fluid flow


Fluids for which the stress is not proportional to time rate of strain are called
non-newtonianf luids. Such fluids show a nonlinear dependence of shearing
stress on velocity gradient. Examples of non-newtonian fluids are: honey,
printers’ ink, paste, and tar.

1.7 FLUID STATICS


1.7.1 Scalar, Vector, and Tensor Quantities
Before entering the field of fluid mechanics, it will be useful to classify certain
types of quantities which are essential for the study of the subject. Some of
the very useful quantities necessary for the study of fluid mechanics are scalar,
vector, and tensor quantities.

ISTUDY
10 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

Scalar quantities require only a magnitude to be specified for a complete


description. For example, temperature is a scalar quantity.
Vector quantities require, in addition to magnitude, a complete direc-
tional specification for their description. Usually three values associated with
orthogonal directions are used to specify a vector. These quantities are called
scalar components of a vector. For example, velocity is a vector quantity.
T ensors require the specification of nine or more scalar components for a
complete description. For example, stress, strain, and mass moment of inertia
are tensor quantities.
A field is a continuous distribution of a scalar, vector, or a tensor quantity
described by continuous functions of space coordinates and time. For example,
the scalar quantity temperature at all points in a field may be expressed
mathematically as T (x, y, z, t). In a similar fashion, a vector field like velocity
may also be described mathematically as V (x, y, z, t). However, normally
three scalar fields are employed to designate a vector field. Thus, the velocity
field may be expressed as

Vx = f (x, y, z, t)
Vy = g(x, y, z, t) (1.2)
Vz = h(x, y, z, t)

where Vx , Vy , Vz are the velocity components along x, y, and z directions,


respectively. Likewise, the tensor fields may be designated mathematically by
employing nine or more scalar fields.

1.7.2 Body and Surface Forces


The forces we come across in continuum fluid mechanics may broadly be
divided into body forces and surface forces. All external forces acting on
any material, which are developed without physical contact, are called body
forces. Gravitational force, the effect of the earth on a mass manifesting itself
as a force distribution throughout the material, directed towards the earth’s
centre, is a body force. Body forces are usually expressed per unit mass of
the material acted on. All forces exerted on a boundary by its surroundings
through direct contact are termed surface forces, e.g. pressure.

1.7.3 Forces in Stationary Fluids


Consider an infinitesimal prismatic element in a stationary fluid, as shown in
Figure 1.2. The triangular prismatic element considered has dimensions of δx,
unity and δz along x, y, and z directions, respectively. The force of gravity
is assumed to be the only body force acting on the fluid element. There are
no shear forces acting on the element, since a fluid element cannot withstand
a shear stress, a stationary fluid must necessarily be completely free of shear
stress.

ISTUDY
1.7. FLUID STATICS 11

FIGURE 1.2 A stationary fluid element.

The pressures acting at the faces are shown as px , pz and pn . The pressure
may be defined as “the force per unit area which acts normal to the surface of
any object which is immersed in a fluid”. For equilibrium, the net force acting
on the fluid element along the x and z directions must be zero. Therefore,
δz
px (δz · 1) − pn sin θ = 0 (1.3)
sin θ
δx 1
pz (δx · 1) − pn cos θ − (δxδz · 1)ρg = 0 (1.4)
cos θ 2
Now, letting the size of the element to shrink to zero, we see from Eqs. (1.3)
and (1.4) that

px = pn = pz ≡ p (1.5)
That is, the pressure in a stationary fluid is equal in all directions.

Pressure force on a fluid element


Consider a small fluid element of length δx, δy and δz along x, y and z direc-
tions, respectively, in a stationary fluid, as shown in Figure 1.3. The corner
of the fluid element close to the origin is taken as the position (x, y, z).

FIGURE 1.3 Pressure force on a fluid element.

ISTUDY
12 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

The pressure force acting on the fluid element along the x-direction is
given by
( )
∂p ∂p
dFp,x = p(δyδz) − p + δx (δyδz) = − (δxδyδz) (1.6)
∂x ∂x
Similarly, the pressure force acting on the fluid element along the y and z
directions can be expressed as

∂p
dFp,y = − (δxδyδz) (1.7)
∂y
and
∂p
dFp,z = − (δxδyδz) (1.8)
∂z
respectively. Combining the pressure force components given by Eqs. (1.6)–
(1.8), the net pressure force acting on the fluid element can be written as
( )
∂p ∂p ∂p
dFp = − i+ j+ k (δxδyδz) (1.9)
∂x ∂y ∂z
where i, j, and k are the unit vectors along x, y, and z directions, respectively.
Then the net force per unit volume is
( )
dFp ∂p ∂p ∂p
=f =− i+ j+ k (1.10)
dx dy dz ∂x ∂y ∂z
If cylindrical coordinates rather than Cartesian coordinates were used, f in
Eq. (1.10) would have taken a form different from the one given above. How-
ever, all such formulations have identically the same physical meaning which
is independent of the coordinate system used for evaluation purposes. Hence,
Eq. (1.10) can also be written as

f = −grad p (1.11)
or
f = −∇p (1.11a)
where the operator ∇ is called the gradient operator and has a form dependent
on the coordinate system used. For Cartesian coordinates,

∂ ∂ ∂
grad ≡ i +j +k (1.12)
∂x ∂y ∂z

1.7.4 Basic Equation of Fluid Statics


Consider the element of sides δx, δy, and δz of a stationary fluid, as shown
in Figure 1.4. It is at rest under the action of pressure forces and the gravity
force. The pressure force acting on it is obtained from Eq. (1.9) as

ISTUDY
1.7. FLUID STATICS 13

FIGURE 1.4 Fluid element at equilibrium.

( )
∂p ∂p ∂p
dFp = − i+ j+ k (δxδyδz) = −(∇p)δV (1.13)
∂x ∂y ∂z
where δV = δxδyδz is the volume of the fluid element. The gravity force
acting on the element is
dFg = ρgδV (1.14)
For equilibrium, from Eqs. (1.13) and (1.14), we have

−∇p + ρg = 0

or
∇p = ρg (1.15)
Equation (1.15) is the basic equation of fluid statics.
If g is taken as acting in the negative z-direction, i.e. g = – gk, the three
components of Eq. (1.15) will then be
∂p
=0 (1.16)
∂x
∂p
=0 (1.16a)
∂y

∂p
= − ρg (1.16b)
∂z
From Eqs. (1.16), (1.16a) and (1.16b), it is seen that pressure in a stationary
fluid can vary only in the z-direction, which has been selected as the direction
of gravity. In other words, the pressure in a stationary fluid varies only in the
vertical direction, and is constant in any horizontal plane. At this stage, it is
important to note that in the preceding formulations it is assumed that the
free surface of a liquid at rest (or the interface between a liquid and a gas or
between two immiscible liquids) is at right angles to the direction of gravity.

ISTUDY
14 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

Hydrostatic pressure distribution


Using the ordinary derivative and realizing that the pressure varies only in
the z-direction and is not a function of x and y, Eq. (1.16b) may be expressed
as
dp
= − ρg (1.17)
dz
For incompressible fluids, i.e. for fluids with constant density ρ, Eq. (1.17)
can be integrated between any desired limits to evaluate the pressure distri-
bution in the static fluid under consideration. Choosing the subscript “0” to
represent conditions at the free surface, integrate Eq. (1.17) from any position
z, where the pressure is p, to position z0 , where the pressure is atmospheric
and denoted as patm . Thus, we have
∫ z0 ∫ patm
−ρg dz = dp
z p

Since ρ and g are constants, the above equation readily gets integrated to
yield
patm − p = −ρg(z0 − z)
or
p = patm + ρg(z0 − z) (1.18)
From Eq. (1.18) it is seen that, in stationary fluids, the pressure increases
linearly with depth (negative z). This linear pressure distribution is called
the hydrostatic pressure distribution.
Usually the term (p – patm ), i.e. the pressure above the atmospheric
pressure, is known as the gauge pressure, and is denoted by pg . So,

pg = ρg(z0 − z)

where (z0 − z) in the above equation is the depth h below the free surface.
Therefore,
pg = ρgh (1.19)
In all engineering flow problems the p to be measured by pressure gauges are
above or below that of atmosphere. Therefore, in engineering work the gauge
pressure pg can be negative, with a maximum possible negative value equal
to –patm .
The hydrostatic pressure distribution, given by Eq. (1.18), holds for mov-
ing fluids as well, provided there is no acceleration in the direction normal to
the flow. This finds a very good application in manometry.

EXAMPLE 1.1 A water-in-glass manometer is connected to two vessels


A and B at pressures 100 kPa and 10 kPa, as shown in Figure 1.5. Calculate
the water column height h.

ISTUDY
1.7. FLUID STATICS 15

FIGURE 1.5 Manometer under differential pressure.

Solution For pressure balance,

pA = 10 kPa + ρgh
∴ ρgh = pA − 10 kPa
= (100 − 10) kPa (∵ pA = 100 kPa)
= 90 × 103 Pa
90 × 103
∴ h= = 9.174 m
103 × 9.81

EXAMPLE 1.2 In Figure 1.6, if the pressure pA = 10 kPa, find pB . Take


ρHg = 13.6 × 103 kg/m3 .

FIGURE 1.6 A mercury manometer reading pressures


(all dimensions in mm).

ISTUDY
16 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

Solution For equilibrium of mercury, as shown in Figure 1.6

pM = pN + ρHg ghHg
pM = pA = 10 kPa
∴ pN = (10 × 103 − 13.6 × 103 × 9.81 × 30 × 10−3 ) Pa
= 5.998 kPa
i.e. pB = pN = 5.998 kPa

1.7.5 Hydrostatic Forces on Submerged Surfaces


Consider a two-dimensional curved plate submerged in a stationary incom-
pressible fluid, as shown in Figure 1.7. Examine the upper surface. The
pressure force acting on the elemental area dA is –p dA (i.e. inwards). By
Eq. (1.18), the pressure p is

p = patm + ρgh

In terms of gauge pressure,


pg = ρgh

FIGURE 1.7 Forces on a submerged body.

Therefore, the force acting on the elemental area dA is

dF = −ρgh dA (1.20)

The total force F acting on the upper surface of the plate, is obtained by
integrating dF over the area as
∫∫
F = −ρg h dA (1.21)
Area

ISTUDY
1.7. FLUID STATICS 17

Equation (1.21) gives the resultant force acting at the upper surface of the
plate due to the gauge pressure.
The vertical component of the resultant force is given by
∫∫ ∫∫
Fz = −ρg h(k · dA) = −ρg h dAz (1.22)
Area Area

where k is the unit vector along the vertical direction and dAz is the projected
area in the vertical direction. But hdAz is the volume dV of the fluid prism
that stands vertically on the area element dA. Therefore,

Fz = −ρgV (1.23)

The negative sign in Eq. (1.23) shows that Fz acts downwards. For thin
plates, the vertical force on the bottom surface is also the same as Fz given by
Eq. (1.23), except that it acts upwards. Similarly, the horizontal component
of the force is given by
∫∫ ∫∫
Fx = −ρg h(i · dA) = ρg h dAx (1.24)
Area Area

where i is the unit vector along the x-direction and dAx is the area projected
on the x-direction. Thus, the horizontal and vertical forces on curved surfaces
are simply estimated with the projected area of the surfaces in the respective
directions.

Buoyancy
The buoyant force on a body is defined as the vertical force due to the fluid or
fluids in contact with the body. A body in floatation is in contact only with
fluids, and the surface force from the fluids is in equilibrium with the force of
gravity on the body.
Consider a three-dimensional body completely submerged in a fluid of
density ρ, as shown in Figure 1.8. The downward force, due to fluid acting
on the upper surface of the body, by Eq. (1.23), is ρgV1 , where V1 is the
volume of fluid prism that stands on the upper surface and extends to the
fluid level. Similarly, the upward force due to fluid acting on the lower surface
is ρgV2 , where V2 is the sum of the volume V1 and the volume of the body.
The net upward force, due to fluid, acting on the body, termed buoyancy, is
given by

FB = ρg(V2 − V1 ) (1.25)
where (V2 − V1 ) is the volume of the body Vs and, therefore, the buoyant
force becomes

FB = ρgVs (1.26)

ISTUDY
18 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

FIGURE 1.8 A body experiencing buoyancy.

Equation (1.26) is the mathematical form of the law of buoyancy due to


Archimedes which states that “a body immersed in a fluid experiences a buoy-
ant force equal to the weight of the fluid displaced by it”. From Eq. (1.26) it
is also seen that, when ρ is constant, the buoyancy force does not depend on
the depth of submergence.

EXAMPLE 1.3 Compute the magnitude, the direction, and the point of
action of the resultant pressure force, due to water, acting on the parabolic
dam section of width 10 m, as shown in Figure 1.9.

FIGURE 1.9 Parabolic profile of a dam.

Solution The equation of the parabolic profile of the dam is x2 = 10z. The
horizontal component Fx of force by Eq. (1.24), is
∫∫
Fx = ρg h dAx
∫∫
= ρg (90 − z)dAx

ISTUDY
1.7. FLUID STATICS 19

∫ 90∫ 10
= ρg (90 − z) dzdy
0 0
[ ]90
z2
= ρg 90z − [y]10
0
2 0
( )
90 × 90
= 10 × 9.81 90 × 90 −
3
(10)
2

= 397.305 MN

Similarly, the vertical component Fz of the force, using Eq. (1.22), is de-
termined to be Fz = –176.58 MN acting downwards. The resultant force is,
therefore, given by

F = Fx2 + Fz2 = 434.8 MN
The resultant force is inclined to the vertical at an angle θ, which is given by
( )
−1 Fx
θ = tan = 66◦ 2′ 15′′
Fz

The point of action of the resultant force is obtained as follows:


x-location:
∫ ( ∫ 30∫ )
10
x2
xf Fz = x dFz = ρg 90 − x dx dy
0 0 10
[ ]30
x2 x4
= 10 × 9.81 90 −
3
[y]10
0
2 40 0

= 0.1986525 × 1010 N·m

0.1986525 × 1010
∴ xf = = 11.25 m from the z-axis
0.017658 × 1010
z-location:
∫ ∫ 90∫ 10
zf Fx = z dFx = ρg (90 − z)z dz dy
0 0

= 1.191915 × 1010 N·m

1.191915 × 1010
∴ zf = = 30 m above the x-axis
0.0397305 × 1010
y-location:
yf = 5 m from the x-z plane because of symmetry.

ISTUDY
20 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

EXAMPLE 1.4 Calculate the force exerted by a water jet of 10 mm


diameter and velocity 10 m/s which strikes a flat plate at an angle of 45◦
to the jet axis.

Solution From Figure 1.10 it is seen that the momentum of the given jet
has to be balanced by the force F acting on the flat plate to keep the plate
in equilibrium. For equilibrium,

FIGURE 1.10 Water jet striking a plate.

Fx = momentum of the jet in the x-direction


= (ρAV1 )V1
[ ( )2 ]
π 10
= 10 ×
3
× 10 × 10
4 1000
π
= × 10 = 7.85 N
4
Also,

Fx = F cos 45◦

Therefore,

F = 11.10 N

EXAMPLE 1.5 The cylindrical gate of mass 1000 kg, shown in Figure 1.11,
is 2.5 m long and is hinged at O. What is the torque necessary to hold the
gate in position?

Solution The projected area of the gate in the x-direction is Ax =


(2 × 2.5) m2 = 5 m2 . The projected area of the gate in the z-direction is
Az = (2 × 2.5) m2 = 5 m2 . The fluid pressure acting on the gate in the
x-direction is
px = ρg(1) N/m2

ISTUDY
1.7. FLUID STATICS 21

FIGURE 1.11 A hinged gate.

Therefore, the force Fx = px × Ax = 5ρg N. The moment due to Fx about O


is
10ρg
Mx,O = 5ρg(2/3) = N·m (clockwise)
3
The force due to the fluid on the gate in the z-direction is Fz = ρgV, where V
is the volume of the fluid displaced by the presence of the gate which is equal
to one-fourth of the volume of gate. Fz is the buoyant force and thus given
by
π × 22 × 2.5
Fz = ρg = 2.5πρg N
4
The moment due to Fz about O is
4 20
Mz,O = (2.5πρg) (2) = ρg N·m (clockwise)
3π 3
Therefore, the torque necessary to hold the gate in position is equal to

(Mx,O + Mz,O )counterclockwise = 98.1 kN.m (counterclockwise)

EXAMPLE 1.6 Find the free-fall velocity of a sphere of diameter 200 mm


and weight 5 N, falling through air at sea level state.
Solution Given, d = 0.2 m, W = 5 N. For air at sea level state,
ρ = 1.225 kg/m3 .
In free-fall, the weight of the sphere is balanced by the buoyant force FB and
the drag force FD . Thus

Fz = W − FB − FD = 0

where z is the normal direction.


(a) The buoyant and drag forces are given by

FB = ρ V g
( 3)
πd
=ρ× ×g
6

ISTUDY
22 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS
( )
π × 0.23
= 1.225 × × 9.81
6
= 0.0503 N
1
FD = ρV 2 SCD
2
where
S = πr2

= π × 0.12

= 0.0314 m2
Therefore,
( )
1
5 − 0.0503 − × 1.225 × 0.0314 × V 2 CD =0
2

5 − 0.0503 − 0.0192 V 2 CD = 0

V 2 CD = 257.8
This has to be solved by trial and error, by assuming a value for CD . Using
the velocity obtianed, the drag coefficient for the Reynolds number for this
velocity has to be checked with the Moody’s chart.
Let us assume the drag coefficient to be 0.2. Thus

257.8
V =
0.2
= 35.9 m/s
For this velocity,
ρV d
Re =
µ
For sea level air,
288.153/2
µ = 1.46 × 10−6 ×
288.15 + 111
= 1.789 × 10−5 kg/(m s)
Therefore,
1.225 × 35.9 × 0.2
Re =
1.789 × 10−5
= 4.92 × 105
For this Reynoldy’s number, from Moody’s chart, CD ≈ 0.2. Therefore, the
velocity is
V ≈ 36 m/s

ISTUDY
1.7. FLUID STATICS 23

1.7.6 Pressure Variation with Elevation for a Static


Compressible Fluid
The hydrostatic pressure distribution Eq. (1.18) relates the pressure, the
specific weight ρg, and the elevation. For compressible fluids, ρ is a variable
and cannot be treated as a constant. For the discussion of pressure variation
with elevation, let us assume a perfect gas medium like air. Air or most of
its components behave like a perfect gas for relatively large ranges of pressure
and temperature. In general, the density of air is a function of pressure and
temperature. For a perfect gas, the equation of state relating the density to
pressure and temperature is

p = ρRT (1.27)

where R is a constant which is a characteristic of the gas. The variation of


pressure in the z-direction, given by Eq. (1.17), is

dp = −ρg dz

Eliminating ρ in the preceding equation, by using Eq. (1.27), we get

dp g
=− dz (1.28)
p RT

The pressure variation with elevation can be determined by solving Eq. (1.28)
provided the temperature and elevation relation is known. Let us solve
Eq. (1.28) for the specific field, namely, atmosphere. Here the elevation is
altitude, Therefore, to solve Eq. (1.28), the temperature and altitude relation
must be known.
To illustrate the application of Eq. (1.28) to get the pressure variation
with elevation, we consider below the troposphere and the stratosphere of the
earth’s atmosphere.

The troposphere
The layer of air above the earth’s surface up to 11 km altitude is called the
troposphere. The temperature–altitude variation in the troposphere has been
found to be linear and may be expressed as

T = T0 − λz (1.29)

where T0 = 288 K is the standard sea-level temperature and λ is known as


the lapse rate. In the international standard atmosphere, the lapse rate is
assigned a value of 6.5 K/km.
Substituting for dz from Eq. (1.29) into Eq. (1.28), we get

dp g dT
= (1.30)
p Rλ T

ISTUDY
24 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

Integrating Eq. (1.30), we get


g
ln p = ln T + const. (1.31)

The integration constant in this equation can be evaluated as follows: At sea
level, p = p0 and T = T0 . Using these values in Eq. (1.31), we get
g
const. = ln p0 − ln T0

Therefore, Eq. (1.31) becomes
( )
p g T g λz
ln = ln = ln 1 − (1.32)
p0 Rλ T0 Rλ T0

Equation (1.32) may also be written as


( )g/Rλ
p λz
= 1− (1.32a)
p0 T0
Combining Eqs. (1.27) and (1.32a), the density–altitude relation for a
perfect gas can be written as
( )(g/Rλ)−1
ρ λz
= 1− (1.33)
ρ0 T0

From Eqs. (1.32a) and (1.33), the relation between pressure and density
is obtained as ( )1−(Rλ/g)
ρ p
= (1.34)
ρ0 p0

The stratosphere
The layer of air between 11 km and 20 km above the earth’s surface is
known as the stratosphere. The temperature in the stratosphere is approxi-
mately constant at –56◦ C. Now, let the temperature in the stratosphere be T1 .
Equation (1.28) for this isothermal zone can be written as

dp g
=− dz (1.35)
p RT1

Integration of Eq. (1.35) results in


p
= exp[−(g/RT1 )(z − z1 )] (1.36)
p1

where p1 is the pressure at altitude z1 . From Eq. (1.27), we have


p
ρ=
RT1

ISTUDY
1.8. SUMMARY 25

Therefore, the density ratio can be expressed as


ρ p
= = exp[−(g/RT1 )(z − z1 )] (1.37)
ρ1 p1
In the above discussion, the acceleration due to gravity g was assumed to be a
constant. Actually, it varies inversely with the square of the distance from the
centre of the earth. However, the error due to this assumption is insignificant.

1.8 SUMMARY
The forces we come across in continuum fluid mechanics may broadly be
divided into body forces and surfaces forces. All external forces acting on
any material, which are developed without physical contact, are called body
forces. Gravitational force, the effect of the earth on a mass manifesting itself
as a force distribution throughout the material, directed towards the earth’s
centre, is a body force. Body forces are usually expressed per unit mass of
the material acted on. All forces exerted on a boundary by its surroundings
through direct contact are termed surface forces, e.g. pressure.
For a fluid at rest the pressure on a surface must act in the direction
perpendicular to that surface. In other words, the pressure acting at a point
in a fluid at rest is the same in all directions. This is known as Pascal’s
principle.
The pressure in a stationary fluid varies only in the vertical direction, and
is constant in any horizontal plane.

p = patm + ρ g (z0 − z)

From this equation it is seen that, in stationary fluids, the pressure increases
linearly with depth (negative z). This linear pressure distribution is called
the hydrostatic pressure distribution. Usually, the term (p − patm ), i.e. the
pressure above atmospheric pressure, is known as the gauge pressure, and is
denoted by pg . So,
pg = ρg(z0 − z)
where (z0 − z) in the above equation is the depth h below the free surface.

1.9 PROBLEMS
1.1 What is the pressure difference between the points A and B in the tanks
shown in Figure P1.1?

ISTUDY
26 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

FIGURE P1.1

[Ans. pA − pB = ρHg gh1 − (h2 + h3 ) ρH2 O g]


1.2 Determine the pressure at a depth of 10 km below the surface of a sea.
Take the average specific gravity of water to be 1.3.
[Ans. 1.2763 × 108 Pa(abs)]
1.3 Determine the pressure at the point A in the tank shown in Figure P1.3.

FIGURE P1.3

[Ans. 119.5716 kPa]


1.4 Consider a U–tube and a funnel arrangement shown in Figure P1.4.
Mercury is poured into the funnel to trap the air in the tube. The tube has
10 mm inside diameter and 1 m total length. Assuming that the trapped
air is compressed isothermally, determine h at which the funnel will begin to
overflow.

ISTUDY
1.9. PROBLEMS 27

FIGURE P1.4

[Ans. 0.3942 m]
1.5 Find the total force on door AB, shown in Figure P1.5, and the
moment of this force about the bottom of the door. The width of the door is
2 m.

FIGURE P1.5

[Ans. 2601612 N, 2601612 N·m]

1.6 A rectangular plate ABC, shown in Figure P1.6, can rotate about the
hinge B. Determine the length l of BC for which there is no torque about B
due to water weight and pressure. The plate weight is 500 N/m. The width
of the plate is 2 m.
[Ans. 4.86 m]

ISTUDY
28 CHAPTER 1. SOME BASIC FACTS ABOUT FLUID MECHANICS

FIGURE P1.6

1.7 A nitrogen tank contains 8 kg of nitrogen at 10 MPa(absolute) and 30◦ C.


If the tank is spherical in shape, determine its volume and diameter.
[Ans. 0.072 m3 , 0.516 m]
1.8 A swimmer can move at a maximum steady speed of Vr = 2 m/s in
still water. To do so, the swimmer must produce enough power to overcome
water resistance. The drag force acting on the swimmer is estimated to be
FD = k V 2 , where, k = 25 N·s2 /m2 , and V is the speed of the swimmer
relative to the water. Evaluate the power produced by the swimmer in still
water. If the swimmer can maintain the same power output in a river current
that moves at 3.5 km/h, estimate the maximum speeds that the swimmer can
reach (a) swimming upstream and (b) swimming downstream.
[Ans. 200 W; (a) 1.03 m/s; (b)2.97 m/s]
1.9 If a pitot-static tube connected to a water manometer has to measure an
air speed of 0.6 m/s, within ± 1 per cent accuracy, under standard sea level
conditions, what should be the sensitivity of the manometer?
[Ans. 0.022 mm of water]
1.10 The lower-half of a 5 m high circular cylinder container is filled with
water and the upper-half with an oil of specific gravity 0.8. Determine the
pressure difference between the top and bottom of the cylinder.
[Ans. 44.145 kPa]
1.11 A gas is contained in a vertical, frictionless piston–cylinder device. The
piston has a cross-sectional area of 30 cm2 and a mass of 3 kg. A compressed
spring above the piston exerts a force of 50 N on the piston. Calculate the
pressure inside the cylinder. Take atmospheric pressure to be 100 kPa.
[Ans. 126.476 kPa]
1.12 A right-circular conical tank of height 2 m and base diameter 1 m is
filled with 0.5 m3 of water. Determine the height of the free surface of water
from the tank base.
[Ans. 1.552 m]

ISTUDY
1.9. PROBLEMS 29

1.13 A pressure gauge connected to a tank reads 500 kPa at a place where
the atmospheric pressure is 100 kPa. Determine the tank pressure in absolute
units.
[Ans. 600 kPa]
1.14 A vacuum gauge connected to a tank reads 50 kPa at a place where the
barometric reading is 760 mm of mercury. Determine the absolute pressure
in the tank. Take ρHg = 13.6 kg/m3 .
[Ans. 51.396 kPa]
1.15 A 10 m diameter spherical balloon is filled with helium gas of density
0.3 kg/m3 . If the balloon has to carry a load of 200 kg, determine the accel-
eration of the balloon when it is first released. Assume the air density to be
1.225 kg/m3 , and neglect the weight of the ropes and the cage.
[Ans. 7.81 m/s2 ]
1.16 Determine the maximum amount of load that the balloon described in
Problem 1.15 can carry.
[Ans. 641.4 N]
1.17 Determine the pressure that can act on a diver at 50 m depth in a
sea. Assume the specific gravity of sea water to be 1.03 and the atmospheric
pressure to be 101 kPa.
[Ans. 606.215 kPa]
1.18 A perfect gas at constant temperature is at rest in equilibrium in a
uniform gravitational field. Find the pressure as a function of height z, given
that, at z = 0, p = p0 .
[Ans. p0 exp (− gz/RT )]
1.19 A mercury manometer connected to a wall pressure tap of a duct with
air flow shows a pressure of 25 mm suction. If the atmospheric pressure is
101 kPa, determine the absolute static pressure at the duct wall.
[Ans. 97.665 kPa]
1.20 A cylinder contains a fluid at a gauge pressure of 350 kPa. Express
the pressure in terms of (a) water, (b) mercury. What would be the absolute
pressure in the cylinder if the atmospheric pressure is 101.3 kPa?
[Ans. (a) 35.68 m; (b) 2.62 m; 451.3 kPa]
1.21 A spherical balloon of diameter 1.5 m and weight 8.50 kN is anchored to
the sea floor with a cable. The balloon is completely immersed in the water.
If the specific weight of sea water is 10.1 kN/m3 , calculate the tension in the
cable. [Ans. 9.35 kN]
1.22 A parajumper deploys his parachute of diameter 8.5 m, at 2500 m
altitude. If the parachute and the person weigh 825 N, (a) determine the
terminal velocity of the parachute at 2500 m and (b) determine the time to
fall from 2500 m to 1000 m, assuming that the terminal velocity is constant
through every 500 m height.
[Ans. (a) 5.03 m/s, (b) 283.4 s]

ISTUDY
Chapter 2

Fundamentals of Fluid
Mechanics

2.1 INTRODUCTION
Gases and liquids are generally termed fluids. Though the physical properties
of gases and liquids are different, they are grouped under the same heading
since both can be made to flow unlike a solid. Under dynamic conditions, the
nature of the governing equations is the same for both gases and liquids. Hence
it is possible to treat them under the same heading, namely fluid dynamics or
fluid mechanics. However, certain substances known as visco-elastic materials
behave like liquid as well as like solid depending on the rate of application of
the force. Pitch and silicone putty are typical examples of the visco-elastic
material. If a force is applied suddenly the visco-elastic material will behave
like a solid, but with a gradually applied pressure the material will flow like
a liquid. The properties of such materials are not considered in this book.
Similarly, non-newtonian fluids, low-density flows, and two-phase flows such
as gas–liquid mixtures are also not considered in this book.

2.2 PROPERTIES OF FLUIDS


Fluid may be defined as a substance which will continue to change shape as
long as there is a shear stress present, however small it may be. That is,
the basic feature of a fluid is that it can flow, and this is the essence of any
definition of it. Examine the effect of shear stress on a solid element and
a fluid element as shown in Figure 2.1. It is seen from this figure that the
change in shape of the solid element is characterized by an angle ∆α when
it is subjected to a shear stress. On the other hand, for the fluid element
there is no such fixed ∆α even for an infinitesimal shear stress. A continuous
deformation persists as long as the shearing stress is applied. The rate of

30

ISTUDY
2.2. PROPERTIES OF FLUIDS 31

deformation, however, is finite and is determined by the applied shear force


and the fluid properties.

FIGURE 2.1 Solid and fluid elements under shear stress.

2.2.1 Pressure
Pressure may be defined as the force per unit area which acts normal to the
surface of any object which is immersed in a fluid. For a fluid at rest, at any
point the pressure is the same in all directions. The pressure in a stationary
fluid varies only in the vertical direction and is constant in any horizontal
plane. That is, in stationary fluids the pressure increases linearly with depth.
This linear pressure distribution is called the hydrostatic pressure distribu-
tion. The hydrostatic pressure distribution is valid for moving fluids as well,
provided there is no acceleration in the vertical direction. This distribution
finds a very good application in manometry.
When a fluid is in motion the actual pressure exerted by the fluid is known
as the static pressure. If there is an infinitely thin pressure transducer which
can be placed in a flow field without disturbing the flow and it can be made
to travel with the same speed as that of the flow, it will then record the exact
static pressure of the flow. From this stringent requirement of the probe for
static pressure measurement, it can be inferred that exact measurement of
static pressure is impossible. However there are certain phenomena, like the
static pressure which at the edge of a boundary is impressed through the layer,
which are made use of for the proper measurement of static pressure. The
total pressure is that pressure which a fluid flow will experience if it is brought
to rest isentropically. It is also called the impact pressure. The total and static
pressures are used for computing the flow velocity.
Since pressure is the intensity of force, it has the dimensions
Force M LT −2
= = M L−1 T −2
Area L2
and is expressed in the units of newton per square metre (N/m2 ) or simply
pascal (Pa). At standard sea level conditions, the atmospheric pressure is
101325 Pa, which corresponds to 760 mm of mercury column height.

ISTUDY
32 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.2.2 Temperature
In any form of matter the molecules are in motion relative to each other. In
gases the molecular motion is a random movement of appreciable amplitude
ranging from about 76 × 10−9 m under normal conditions to some tens of
millimetres at very low pressures. The distance of free movement of a molecule
of a gas is the distance it can travel before colliding with another molecule or
the walls of the container. The mean value of this distance for all molecules in
a gas is called the molecular mean free path length. By virtue of this motion
the molecules possess kinetic energy, and this energy is sensed as temperature
of the solid, liquid or gas. In the case of a gas in motion it is called the
static temperature. Temperature has the units kelvin (K) or degree Celsius
(◦ C), in SI units. For all calculations in this book, the temperature will be
expressed in kelvin, i.e. from absolute zero. At standard sea level conditions
the atmospheric temperature is 288.15 K.
Although it is well known that temperature is a measure of the intensity of
‘hotness’ or ‘coldness’, it is not easy to give an exact definition to it. Based on
psychological sensations, we express the level of temperature qualitatively
with words like hot, red-hot, warm, cold, and freezing cold. However, we
cannot assign numerical values to temperature based on our sensations alone.
Detailed information about temperature and its measurements can be seen in
E. Rathakrishnan, Fundamentals of Engineering Thermodynamics, 2nd ed.,
Prentice-Hall of India, New Delhi, 2005.

2.2.3 Density
The density is a measure of the amount of material contained in a given
volume. In other words, the material contained in a unit volume is a measure
of the density ρ of the substance. It is expressed as mass per unit volume, say
kg/m3 . Mass is defined as weight divided by acceleration due to gravity. At
standard atmospheric temperature and pressure (288.15 K and 101325 Pa)
the density of dry air is 1.225 kg/m3 .
The density of a material is a measure of the amount of matter contained
in a given volume. In a fluid, density may vary from point to point. Consider
the fluid contained within a small spherical region of volume δV centred at
some point in the fluid, and let the mass of the fluid within this spherical
region be δm. Then the density of the fluid at the point on which the sphere
is centred can be defined by

δm
ρ = lim (2.1)
δV→∞ δV

There are practical difficulties in applying the above definition of density to


real fluids composed of discrete molecules, since under the limiting condition
the sphere may or may not contain any molecule. If it contains a molecule the
value obtained for the density will be fictitiously high. If it does not contain
a molecule the resultant value of density will be zero. This difficulty can be

ISTUDY
2.2. PROPERTIES OF FLUIDS 33

avoided, over the range of temperatures and pressures normally encountered


in practice, by the following two ways.

1. The molecular nature of a gas may be ignored and the gas be treated
as continuum, i.e. the gas does not consist of discrete particles.

2. The decrease in size of the imaginary sphere may be assumed to be


carried to a limiting size. This limiting size of the sphere is such that
although it is small compared to the dimensions of any physical object
present in a flow field, e.g. an aircraft, it is large compared to the
fluid molecules and, therefore, contains a reasonably large number of
molecules.

2.2.4 Viscosity
The property which characterizes the resistance that a fluid offers to shear
force applied is termed viscosity. This resistance, unlike for solids, does not
depend upon the deformation itself but on the rate of deformation. Viscosity
is often regarded as the stickiness of a fluid and its tendency is to resist sliding
between layers. There is very little resistance to the movement of a knife-blade
edge through air, but to produce the same motion through a thick oil needs
much more effort. This is because the viscosity of oil is higher than that of
the air.

2.2.5 Absolute Coefficient of Viscosity


The absolute coefficient of viscosity is a direct measure of the viscosity of a
fluid. Consider a fluid sheared between two parallel plates placed at a distance
h apart, as shown in Figure 2.2. The space between the plates is filled with

FIGURE 2.2 A fluid sheared between two plates.

a fluid. The bottom plate is fixed and the other is moved in its own plane
at a speed u. The fluid in contact with the lower plate will be at rest, while
that in contact with the upper plate will be moving with speed u, because
of no-slip condition. In the absence of any other influence, the speed of the

ISTUDY
34 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

fluid between the plates will vary linearly, as shown in Figure 2.2. As a direct
result of viscosity, a force F has to be applied to each plate to maintain the
motion, since the fluid will tend to retard the motion of the moving plate and
will also tend to drag the fixed plate in the direction of the moving plate. If
the area of each plate in contact with the fluid is A, then the shear stress
acting on the plate is F/A. The rate of slide of the upper plate over the lower
is u/h.
The above quantities are connected by Maxwell’s equation, which serves
to define the absolute coefficient of viscosity µ. The equation is

F (u)
=µ (2.2)
A h
Hence, [ ] [ ] [ ]
ML−1 T−2 = [µ] LT−1 L−1 = [µ] T−1
i.e. [ ]
[µ] = ML−1 T−1
and the unit of µ is therefore kg/m·s. At 0◦ C the absolute coefficient of
viscosity of dry air is 1.71 ×10−5 kg/m·s. The absolute coefficient viscosity µ
is also called the dynamic viscosity coefficient.
Equation (2.2) with µ remaining constant does not apply to all fluids. For
a class of fluids, which includes blood, some oils, some paints and the so-called
thixotropic fluids, µ is not constant but is a function of du/dh. The derivative
du/dh is a measure of the rate at which the fluid is shearing. Usually, µ is
expressed as N·s/m2 or gram/cm·s. One gram/cm·s is known as the poise.
Newton’s law of viscosity states that, the stresses which oppose the shearing
of a fluid are proportional to the rate of shear strain, i.e. the shear stress τ is
given by

∂u
τ =µ (2.3)
∂y
where µ is the absolute coefficient of viscosity and ∂u/∂y is the velocity gra-
dient. The viscosity µ is a property of the fluid. Fluids which obey the above
law of viscosity are called newtonian fluids. Some fluids such as silicone oil,
viscoelastic fluids, sugar syrup, tar, etc. do not obey the Newton’s law of
viscosity given by Eq. (2.3) and they are called non-newtonian fluids.
We know that, in incompressible flow, it is possible to separate the cal-
culation of (hydrodynamic) velocity boundary layer from that of the thermal
boundary layer. But in compressible flow this is not possible, since the veloc-
ity and thermal layers interact intimately and therefore, they must be consid-
ered simultaneously. This is because for high-speed flows (compressible flows)
heating due to friction as well as temperature changes due to compressibility
must be taken into account. Further, it is essential to include the effects of
viscosity variation with temperature. Usually, large variations in temperature
are encountered in high-speed flows.

ISTUDY
2.2. PROPERTIES OF FLUIDS 35

The relation µ(T ) must be found by experiment. The voluminous data


available in literature leads to the conclusion that the fundamental relation-
ship is a complex one and that no single correlation function can be found to
apply to all gases. Alternatively, the dependence of viscosity on temperature
can be calculated with the aid of the method of statistical mechanics but, as
yet, no completely satisfactory theory has been evolved. Also, these calcu-
lations lead to a complex expression for the function µ(T ). Therefore, only
the semi-empirical relations appear to be the means to calculate the viscosity
associated with compressible boundary layers. It is important to realize that
even though semi-empirical relations are not extremely precise, they are rea-
sonably simple relations. For air, it is possible to use an interpolation formula
based on D.M. Sutherland’s theory of viscosity. This will lead to an expression
for µ in the following form
( )3/2
µ T T0 + S
=
µ0 T0 T +S
where µ0 denotes the viscosity at the reference temperature T0 , and S is a
constant which for air assumes the value S = 110 K.
For air the Sutherland’s relation can also be expressed [Heiser, W.H. and
Pratt, D.T. Hypersonic Air Breathing Propulsion, AIAA Education Series,
1994] as
( )
−6 T 3/2
µ = 1.46 × 10 [kg/(m s)] (2.4)
T + 111

where T is in kelvin. This equation is valid for the static pressure range
of 0.01 to 100 atm, which is ordinarily encountered in atmospheric flight.
The temperature range in which this equation is valid is up to 3000 K. The
reasons that the absolute viscosity is a function only of temperature under
these conditions, are that the air behaves as a perfect gas, in the sense that
intermolecular forces are negligible, and that viscosity itself is a momentum
transport phenomenon caused by the random molecular motion associated
with thermal energy or temperature.

EXAMPLE 2.1 Determine the absolute viscosity of air at temperatures


0◦ C, 5◦ C, and 10◦ C.

Solution By Sutherland’s relation, we have


( )
−6 T 3/2
µ = 1.46 × 10 [kg/(m s)]
T + 111
Substituting the above temperatures in the Sutherland’s relation, we get

µ0 = 1.71 × 10−5 kg/(m s)

µ5 = 1.74 × 10−5 kg/(m s)

ISTUDY
36 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

µ10 = 1.76 × 10−5 kg/(m s)

The following program can calculate the viscosity of air at the desired
temperatures.

PROGRAM
-----------------------------------------------
c Estimation of viscosity
real mu
do it = 0,2000,10
t=float(it)
t = t + 273.15
mu = 1.46E-6 *( t**(1.5)/(t + 111.0))
print *, it, mu
enddo
stop
end
--------------------------------------------

2.2.6 Kinematic Viscosity Coefficient


The kinematic viscosity coefficient is a convenient form of expressing the vis-
cosity of a fluid. It is formed by combining the density ρ and the absolute
coefficient of viscosity µ according to the equation
µ
ν=
ρ

The kinematic viscosity coefficient ν is expressed in m2 /s or cm2 /s and is


known as stoke.
The kinematic viscosity coefficient is a measure of the relative magnitudes
of viscosity and inertia of the fluid. Both dynamic viscosity coefficient µ and
the kinematic viscosity coefficient ν are functions of temperature. For liquids,
µ decreases with increase in temperature, whereas for gases, µ increases with
increase in temperature. This is one of the fundamental differences between
the behaviour of gases and liquids. The viscosity is practically unaffected by
the pressure.

2.2.7 Thermal Conductivity of Air


At high speeds, heat transfer from vehicles becomes significant. For example,
re-entry vehicles encounter an extreme situation where ablative shields are
necessary to ensure protection of the vehicle during its passage through the
atmosphere. The heat transfer from a vehicle depends on the thermal con-
ductivity k of air. Therefore, a method to evaluate k is also essential. For this
case, a relation similar to Sutherland’s law for viscosity is found to be useful

ISTUDY
2.3. THERMODYNAMIC PROPERTIES 37

and it is ( )
−3 T 3/2
k = 1.99 × 10 [J/(m K s)]
T + 112
where T is the temperature in kelvin. The pressure and temperature ranges
in which this equation is applicable are from 0.01 to 100 atm and from 0 to
2000 K. For the same reason as given for the viscosity relation, the thermal
conductivity also depends only on the temperature.

2.2.8 Compressibility
The change in volume of a fluid associated with change in pressure is called
compressibility. When a fluid is subjected to pressure it gets compressed and
its volume changes. The bulk modulus of elasticity is a measure of how easily
the fluid may be compressed, and is defined as the ratio of pressure change
to volumetric strain associated with it. The bulk modulus of elasticity K is
given by
K=
pressure increment
volume strain
=−
Vdp
dV
V (2.5)

It may also be expressed as

−∆p dp
K = lim = (2.6)
△v→0 ∆v/v dρ/ρ
where v is the specific volume. Since dρ/ρ represents the relative change
in density brought about by the pressure change dp, it is apparent that the
bulk modulus of elasticity is the inverse of the compressibility of the substance
at a given temperature. For instance, K for water and air are approximately
2 GN/m2 and 100 kN/m2 , respectively. This implies that air is about 20,000
times more compressible than water. It can be shown that K = a2 ρ, where
a is the speed of sound. The compressibility plays a dominant role at high
speeds. The Mach number M (defined as the ratio of local flow velocity to
local speed of sound) is a convenient non-dimensional parameter used in the
study of compressible flows. Based on M the flow is divided into the following
regimes. When M < 1 the flow is called subsonic, when M ≈ 1 the flow is
termed transonic flow, M from 1.2 to 5 is called the supersonic regime, and
M > 5 is referred to as hypersonic regime. When the flow Mach number is less
than 0.3 the compressibility effects are negligibly small and hence the flow is
called incompressible. For incompressible flows, density change with velocity
is neglected and the density is treated as invariant.

2.3 THERMODYNAMIC PROPERTIES


We know from thermodynamics that heat is energy in transition. Therefore,
heat has the same dimensions as energy, and is measured in units of joule (J).

ISTUDY
38 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.3.1 Specific Heat


The inherent thermal properties of a flowing gas become important when the
energies are considered. The specific heat is one such quantity. The specific
heat is defined as the amount of heat required to raise the temperature of a
unit mass of a medium by one degree. The value of the specific heat depends
on the type of process involved in raising the temperature of the unit mass.
Usually, the constant volume process and the constant pressure process are
used for evaluating specific heat. The specific heats at constant volume and
constant pressure processes, respectively, are designated by cv and cp . The
specific heat at constant volume is defined as
( )
∂u
cv ≡ (2.7)
∂T v
where u is internal energy per unit mass of the fluid, which is a measure of the
potential and more particularly the kinetic energy of the molecules comprising
the gas. The specific heat cv is a measure of the energy-carrying capacity of
the gas molecules. For dry air at normal temperature, cv = 717.5 J/kg·K.
The specific heat at constant pressure is defined as
( )
∂h
cp ≡ (2.8)
∂T p
where h = u + pv, the sum of internal energy and flow energy is known as the
enthalpy or total heat content per unit mass of fluid and cp is a measure of the
ability of the gas to do external work in addition to possessing internal energy.
Therefore, cp is always greater than cv . For dry air at normal temperature,
cp = 1004.5 J/kg·K.

2.3.2 The Ratio of Specific Heats


The ratio of specific heats given by
cp
γ= (2.9)
cv
is an important parameter in the study of compressible flows. This is a mea-
sure of the relative internal complexity of the gas molecules. It has been
determined from the kinetic theory of gases that the ratio of specific heats
can be related to the number of degrees of freedom n of the gas molecules by
the relation
n+2
γ=
n
At normal temperatures, there are six degrees of freedom, three translational
and three rotational, for diatomic gas molecules. For example, for nitrogen
which is a diatomic gas, n = 5 since one of the rotational degrees of freedom
is small in comparison with the other two. Therefore,
γ = 7/5 = 1.4

ISTUDY
2.4. SURFACE TENSION 39

A monatomic gas like helium has 3 translational degrees of freedom only and
therefore,
γ = 5/3 = 1.67
This value of 1.67 is the upper limit of the value which γ can take. In general,
γ varies from 1 to 1.67, i.e.

1 ≤ γ ≤ 1.67

The specific heats of a compressible gas are related to the gas constant R. For
a perfect gas, this relation is

R = cp − cv (2.10)

The enthalpy is h = u + pv. For a perfect gas, pv = RT . Thus, h = u + RT .


Differentiating with respect to T , we get
∂h ∂u
= +R
∂T ∂T
But by definition,
∂h ∂u
cp = , cv =
∂T ∂T
Therefore, we have

cp = cv + R

or

R = cp − cv

2.4 SURFACE TENSION


Liquids behave as if their free surfaces were perfectly flexible membranes hav-
ing a constant tension σ per unit width; this tension is called the surface
tension. It is important to note that this is neither a force nor a stress but a
force per unit length. The value of surface tension depends on
• the nature of the fluid,
• the nature of the substance with which it is in contact at the surface,
and
• the temperature and pressure.
Consider a plane material membrane possessing the property of constant ten-
sion σ per unit width. Let the membrane be a straight edge of length l. The
force required to hold the edge stationary is

p = σl

ISTUDY
40 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

Now, suppose that the edge is pulled so that it is displaced normal to itself by
a distance x in the plane of the membrane. The work done F , in stretching
the membrane is given by
F = σlx = σA (2.11)
where A is the area added to the membrane. We see that σ is the free energy
of the membrane per unit area. The important point to be noted here is that,
if the energy of a surface is proportional to its area, it will then behave exactly
as if it were a membrane with a constant tension per unit width and this is
totally independent of the mechanism by which the energy is stored. Thus,
the existence of a surface tension at the boundary between two substances is a
manifestation of the fact that the stored energy contains a term proportional
to the area of the surface. This energy is attributable to molecular attractions.
An associated effect of surface tension is the capillary deflection of liquids
in small tubes. Examine the level of water and mercury in capillaries as shown
in Figure 2.3.

FIGURE 2.3 Capillary effect of water and mercury.

When a glass tube is inserted into a beaker of water, the water will rise in
the tube and display a concave meniscus. The deviation of water level h in
the tube from that in the beaker can be shown to be
σ
h∝ cos θ (2.12)
d
where θ is the angle between the tangent to the water surface and the glass
surface. In other words, a liquid like water or alcohol which wets the glass
surface makes an acute angle with the solid wall and the level of the free surface
inside the tube will be higher than that outside. This is termed capillary
action. However, when wetting does not occur, as in the case of mercury in
glass, the angle of contact is obtuse and the level of free surface inside the
tube is depressed, as shown in Figure 2.3.
Another important effect of surface tension is that a long cylinder of liquid,
at rest or in motion, with a free surface is unstable and breaks up into parts,
which then assumes an approximately spherical shape. This is the mechanism
of the breakup of liquid jets into drops.

ISTUDY
2.5. ANALYSIS OF FLUID FLOW 41

2.5 ANALYSIS OF FLUID FLOW


2.5.1 Lagrangian and Eulerian Specifications
Basically two treatments are followed for fluid flow analysis. They are
Lagrangian and Eulerian descriptions. Lagrangian method describes the mo-
tion of each particle of the flow field in a separate and discrete manner. For
example, the velocity of the nth particle of an aggregate of particles moving
in space can be specified by the scalar equations

(Vx )n = fn (t)

(Vy )n = gn (t) (2.13)

(Vz )n = hn (t)

where Vx , Vy , and Vz are the velocity components in the x, y, and z directions,


respectively. They are independent of the space coordinates and functions of
time only. Usually, the particles are denoted by the space point they occupy
at some initial time t0 . Thus, T (x0 , t) refers to the temperature at time t of
a particle which was at location x0 at time t0 .
This approach of identifying the material points and following them along
is also termed the particle or material description. This approach is usu-
ally preferred in the description of low density flow fields (also called rarefied
flows), moving solids, like in describing the motion of a projectile and so on.
However, in a deformable system like a continuum fluid, since there are infinite
number of fluid elements whose motion has to be described, the Lagrangian
approach becomes unmanageable. Instead, we can employ spatial coordinates
to help us in identifying the particles in a flow. The velocity of all particles
in a flow, therefore, can be expressed in the following manner.

Vx = f (x, y, z, t)

Vy = g(x, y, z, t) (2.14)

Vz = h(x, y, z, t)

This is called the Eulerian or field approach. If the properties and flow
characteristics at each position in space remain invariant with time, the flow
is called steady flow. A time dependent flow is referred to as unsteady flow.
The steady flow velocity field would then be given as

Vx = f (x, y, z)

Vy = g(x, y, z) (2.15)

Vz = h(x, y, z)

ISTUDY
42 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.5.2 Relation Between Local and Material


Rates of Change
The rate of change of a property measured by a probe at a fixed location
is referred to as local rate of change and the rate of change of a property
experienced by a material particle is termed the material or the substantive
rate of change.
The local rate of change of a property η is denoted by ∂η(x, t)/∂t, where it
is understood that x is held constant. The material rate of change of property
η shall be denoted by Dη/Dt. If η is the velocity V, then DV /Dt is the rate
of change of velocity for a fluid particle and thus, is the acceleration that the
fluid particle experiences. On the other hand, ∂V /∂t is just a local rate of
change of velocity recorded by a stationary probe. In other words, DV /Dt is
the particle or material acceleration and ∂V /∂t is the local acceleration.
For a fluid flowing with a uniform velocity V∞ , it is possible to write the
relation between the local and material rates of change of property η as
∂η Dη ∂η
= − V∞ (2.16)
∂t Dt ∂x
Thus the local rate of change of η is due to the following two effects.
1. Due to change of property of each particle with time.
2. Due to combined effect of the spatial gradient of that property and the
motion of the fluid.
When a spatial gradient exists, the fluid motion brings different particles
with different values of η to the probe, thereby modifying the rate of change
observed. This latter effect is termed convection effect. Therefore, V∞ (∂η/∂x)
is referred to as the convective rate of change of η. Even though Eq. (2.16)
has been obtained with uniform velocity V∞ , note that, in the limit δt → 0,
it is only the local velocity V which enters into the analysis and therefore, we
have
∂η Dη ∂η
= −V (2.17)
∂t Dt ∂x
Equation (2.17) can be generalized for a three-dimensional space as
∂ D
= − (V · ∇) (2.18)
∂t Dt
where ∇ is the gradient operator (= i ∂/∂x + j ∂/∂y + k ∂/∂z) and (V · ∇) is
a scalar product (= Vx ∂/∂x + Vy ∂/∂y + Vz ∂/∂z). Equation (2.18) is usually
written as
D ∂
= +V · ∇ (2.18a)
Dt ∂t
When η is the velocity of a fluid particle, DV /Dt gives the acceleration of the
fluid particle and the resultant equation is
DV ∂V
= + (V · ∇)V (2.19)
Dt ∂t

ISTUDY
2.5. ANALYSIS OF FLUID FLOW 43

Equation (2.19) is known as the Euler’s acceleration formula. The total


rate of change D/Dt is generally called the material derivative. This is also
referred to as the substantial derivative, or the particle derivative, to empha-
size the fact that the derivative is taken following a fluid particle. As seen
above, it is made of two parts; ∂V /∂t is the local rate of change of V at
a given point, and is zero for steady flows. The second part (V · ∇)V =
Vx (∂Vx /∂x) + Vy (∂Vx /∂y) + Vz (∂Vx /∂z) is called the advective derivative,
because it is the change in V as a result of advection of the particle from one
location to another where the value of V is different.

EXAMPLE 2.2 A bus climbs up a hill by a straight road, making an angle


of 5◦ with the horizontal, during the night when the atmospheric temperature
decreases at a rate of 0.5 K per hour. The fall in atmospheric temperature
with increase in altitude is 6 K per km height. Determine the time rate of
change in temperature experienced by a passenger in the bus, travelling at a
speed of 30 kmph.

FIGURE 2.4 An inclined road.

Solution Given,
∂T
= −0.5 K/h
∂t
∂T
= −6 K/km
∂h
where T is the temperature, h is the altitude, and t is the time. Using
Eq. (2.18a), we can write

DT ∂T ∂T ∂T ∂T
= Vx + Vy + Vz +
Dt ∂x ∂y ∂z ∂t
For the present problem, the above equation reduces to
( )
DT ∂T ∂T ∂T ∂T
= + Vz ∵ =
Dt ∂t ∂h ∂z ∂h

Also, Vz = V sin 5◦ = 30 sin 5◦ km/h. Therefore,

DT
= (30 sin 5◦ × (−6)) − 0.5 = −16.19 K/h
Dt

That is, the passenger will feel a fall in temperature at a rate of 16.19 K per
hour.

ISTUDY
44 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

EXAMPLE 2.3 For the velocity field V = z i+x j+y k, obtain the material
acceleration vector at x = 1, y = 4, z = 1. Also, obtain the components of
acceleration parallel and normal to V at the same position.
DV ∂V ∂V ∂V ∂V
Solution = + Vx + Vy + Vz
Dt ∂t ∂x ∂y ∂z

= 0 + zj + xk + yi

= 4i + j + k at (1, 4, 1)

Acceleration parallel to V is
V
at = a .
V
But,

a.V = (4i + j + k).(z i + x j + y k)


= 4z + x + y
=4+1+4=9

and

V = 42 + 12 + 12

= 18

Thus, the tangential acceleration becomes


9
at = √
18

= 2.12 m/s2

The acceleration normal to V is


√ √
an = a2 − a2t = 18 − 4.494 = 3.68 m/s2

EXAMPLE 2.4 A velocity field is given by u = −3 x, v = 2 y, w = z. Is this


flow steady? Is it two- or three-dimensional? At (x, y, z) = (1, 1, 1), compute
(a) the velocity, (b) the local acceleration, and (c) the convective acceleration.

Solution The velocity field is given as

V = −3 x i + 2 y j + z k

The flow is steady since V is independent of time t. The flow is three-


dimensional since all the three components of velocity are finite.

ISTUDY
2.5. ANALYSIS OF FLUID FLOW 45

(a) At (1,1,1), V = −3 i + 2 j + k

∂V
(b) Local acceleration = = 0
∂t
du dv dw
(c) Convective acceleration = i+ j+ k
dt dt dt

du ∂u ∂u ∂u
=u +v +w
dt ∂x ∂y ∂z
= −3 x (−3) + 2 y (0) + z (0) = 9 x
dv ∂v ∂v ∂v
=u +v +w
dt ∂x ∂y ∂z
= −3 x(0) + 2 y(2) + z(0) = 4 y
dw ∂w ∂w ∂w
=u +v +w
dt ∂x ∂y ∂z
= −3 x (0) + 2 y(0) + z (1) = z

Thus, the convective acceleration at (1, 1, 1) is

acceleration = 9(1) i + 4(1) j + 1 k = 9 i + 4 j + k

2.5.3 Graphical Description of Fluid Motion


There are three important concepts for visualizing or describing the flow fields.
They are:
1. Pathline
2. Streakline
3. Streamline

Pathline
Pathline may be defined as a line in the flow field describing the trajectory
of a given fluid particle. From the Lagrangian viewpoint, for a closed system
with a fixed identifiable quantity of mass, the independent variables are the
initial position (with which each particle is identified) and the time. Hence,
the locus of the same particle over a time period from t0 to tn is called the
pathline.

Streakline
Streakline may be defined as the instantaneous loci of all the fluid elements
that have passed the point of injection at some earlier time. Consider a
continuous tracer injection at a fixed point Q in space. The connection of

ISTUDY
46 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

all elements passing through the point Q over a period of time is called the
streakline.

Streamlines
Streamlines are imaginary lines, in a fluid flow, drawn in such a manner that
the flow velocity is always tangential to them. Flows are usually depicted
graphically with the aid of streamlines. Streamlines proceeding through the
periphery of an infinitesimal area at some instant of time t will form a tube
called streamtube, which is useful in the study of fluid flow.
From the Eulerian viewpoint, for an open system with constant control
volume, all flow properties are functions of a fixed point in space and time, if
the process is transient. The flow direction of various particles at time ti forms
a streamline. The pathline, the streamline and the streakline are different in
general, but coincide in a steady flow.

Timelines
In modern fluid flow analysis, yet another graphical representation, namely
the timeline is used. When a pulse input is periodically imposed on a line
of tracer source placed normal to a flow, a change in the flow profile can be
observed. The tracer image is generally termed timeline. Timelines are often
generated in the flow field to aid the understanding of flow behaviour such as
velocity and velocity gradient.
From the above mentioned graphical descriptions, it can be inferred that:

• There can be no flow through the lateral surface of the streamtube.


• An infinite number of adjacent streamtubes arranged to form a finite
cross-section is often called a bundle of streamtubes.
• Streamline is an Eulerian (or field) concept.
• Pathline is a Lagrangian (or particle) concept.
• For steady flows, streamlines and streaklines are identical.

2.6 BASIC AND SUBSIDIARY LAWS FOR


CONTINUOUS MEDIA
In the range of engineering interest, four basic laws must be satisfied for any
continuous medium. They are:

1. Conservation of matter (continuity equation)


2. Newton’s second law (momentum equation)
3. Conservation of energy (the first law of thermodynamics)
4. Increase-of-entropy principle (the second law of thermodynamics).

ISTUDY
2.6. BASIC AND SUBSIDIARY LAWS FOR CONTINUOUS MEDIA 47

In addition to these primary laws, there are numerous subsidiary laws, some-
times called the constitutive relations that apply to specific types of media or
flow processes (e.g. equation of state for a perfect gas, Newton’s law of vis-
cosity for certain viscous fluids, the isentropic process relation, the adiabatic
process relation, etc.).

2.6.1 Systems and Control Volumes


In employing the basic and subsidiary laws, any one of the following modes
of application may be adopted.
• The activities of each and every given element of mass must be such
that it satisfies the basic laws and the pertinent subsidiary laws.
• The activities of each and every elemental volume in space must be such
that the basic laws and the pertinent subsidiary laws are satisfied.
In the first case the laws are applied to an identified quantity of matter
called the control mass system. A control mass system may change its shape,
position, and thermal condition with time or space or both, but must always
entail the same matter.
For the second case a definite volume called control volume is designated
in space, and the boundary of this volume is known as control surface. The
amount and identity of the matter in the control volume may change with
time, but both the shape and size of the control volume are fixed, i.e. the
control volume may change its position in time or space or both, but both its
shape and size are always preserved.

2.6.2 Integral and Differential Analysis


The analysis where large control volumes are used to obtain aggregate forces
or transfer rates is termed integral analysis. When the analysis is applied
to individual points in the flow field, the resulting equations are differential
equations and the method is termed differential analysis.
The science of fluid mechanics is based on the conservation laws of mass,
momentum, and energy. These laws can be stated in the integral form,
applicable to an extended region. In the integral form, the expressions of
the laws depend on whether they are related to a volume fixed in space, or
to a material volume, which consists of the same fluid particles and whose
bounding surface moves with the fluid.

2.6.3 State Equation


For air at normal temperature and pressure, the density ρ, the pressure p and
the temperature T are connected by the relation p = ρRT , where R is a
constant called the gas constant. This is known as the state equation for a
perfect gas. At high pressures and low temperatures the above state equa-
tion breaks down. At normal pressures and temperatures the mean distance

ISTUDY
48 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

between molecules and the potential energy arising from their attraction can
be neglected. The gas behaves like a perfect gas or an ideal gas in such a
situation. At this stage, it is essential to understand the difference between
the ideal and perfect gases. An ideal gas is frictionless and incompressible.
The perfect gas has viscosity and can therefore develop shear stresses, and it
is compressible according to perfect gas state equation.
Real gases below the critical pressure and above the critical temperature
tend to obey the perfect-gas law. The perfect-gas law encompasses both
Charles’ law and Boyles’ law. Charles’ law states that, for constant pressure
the volume of a given mass of gas varies directly with its absolute temperature.
Boyles’ law (isothermal law) states that, for constant temperature the density
varies directly with the absolute pressure.

2.7 KINEMATICS OF FLUID FLOW


Kinematics is the branch of mechanics that deals with quantities involving
space and time only. It treats variables such as displacement, velocity, accel-
eration, deformation, and rotation of fluid elements without referring to the
forces responsible for such a motion. In other words, kinematics essentially
describes the appearance of a motion. Another branch of mechanics which
deals with the forces due to motion is termed dynamics.
To simplify the discussions let us consider the flow to be incompressible,
i.e. the density is treated as invariant. The basic governing equations for
an incompressible flow are the continuity and momentum equations. The
continuity equation is based on the conservation of matter. For steady incom-
pressible flow the continuity equation in differential form is

∂Vx ∂Vy ∂Vz


+ + =0 (2.20)
∂x ∂y ∂z

Equation (2.20) may also be expressed as ∇.V = 0, where V is the velocity


vector. This is the differential form of the law of conservation of mass. The
same law can also be deduced from the general form of mass conservation
equation in an integral form for a fixed region. Consider a volume V fixed in
space as shown in Figure 2.5. The rate of increase of mass inside the volume
is the volume integral ∫ ∫
d ∂ρ
ρ dV = dV
dt V V ∂t

The time derivative is taken inside the integral on the right-hand side
because the volume is invariant. The rate of mass flow out of the volume is
the surface integral ∫
ρV · dA
A

ISTUDY
2.7. KINEMATICS OF FLUID FLOW 49

FIGURE 2.5 Mass conservation of a volume fixed in space.

By the law of conservation of mass, we have that “the rate of increase of


mass within a fixed volume must be equal to the rate of inflow through the
boundaries”. Therefore,
∫ ∫
∂ρ
dV = − ρV · dA (2.21)
V ∂t A

This is the integral form of the mass conservation law for a volume fixed in
space.
The differential form can be obtained by transforming the surface integral
on the right-hand side of Eq. (2.21) to a volume integral using the divergence
theorem, which gives
∫ ∫
ρV · dA = ∇ · (ρ V ) dV
A V

Equation (2.21) then becomes


∫ [ ]
∂ρ
+ ∇ · (ρV ) dV = 0
V ∂t

This relation holds for any volume. This implies that the integrand vanishes
at every point. If the integrand does not vanish at every point, we can choose
a small volume around that point and obtain a nonzero integral. Therefore,

∂ρ
+ ∇ · (ρV ) = 0 (2.22)
∂t

This is the general form of the continuity equation which expresses the differ-
ential form of the conservation of mass principle.
The continuity equation can be written in several other forms. Rewriting
the divergence term in Eq. (2.22) as
∂ ∂Vi ∂ρ
(ρVi ) = ρ + Vi
∂xi ∂xi ∂xi

ISTUDY
50 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

the continuity equation becomes


1 Dρ
+∇·V =0
ρ Dt
For incompressible flows, ρ is invariant, thus the continuity equation reduces
to
∇·V =0
This is same as Eq. (2.20).
The momentum equation which is based on Newton’s second law represents
the balance between various forces acting on a fluid element. The forces
encountered are:

1. Force due to rate of change of momentum, generally referred to as inertia


force
2. Body force such as buoyancy force, magnetic force, and electrostatic
force
3. Pressure force
4. Viscous forces (causing shear stress).

For a fluid element under equilibrium, by Newton’s second law, we have the
momentum equation as

Inertia force + Body force + Pressure force + Viscous force = 0

For a gaseous medium, the body forces are usually negligibly small compared
to other forces and hence can be neglected. For steady incompressible flows
the momentum equation can be written as
( 2 )
∂Vx ∂Vx ∂Vx 1 ∂p ∂ Vx ∂ 2 Vx ∂ 2 Vx
Vx + Vy + Vz =− +ν + +
∂x ∂y ∂z ρ ∂x ∂x2 ∂y 2 ∂z 2
( 2 )
∂Vy ∂Vy ∂Vy 1 ∂p ∂ Vy ∂ 2 Vy ∂ 2 Vy
Vx +Vy +Vz =− +ν + + (2.23)
∂x ∂y ∂z ρ ∂y ∂x2 ∂y 2 ∂z 2
( 2 )
∂Vz ∂Vz ∂Vz 1 ∂p ∂ Vz ∂ 2 Vz ∂ 2 Vz
Vx + Vy + Vz =− +ν + +
∂x ∂y ∂z ρ ∂z ∂x2 ∂y 2 ∂z 2
Equations (2.23) are the x, y, z components of the momentum equation,
respectively. These equations are generally known as Navier–Stokes equations.
They are nonlinear partial differential equations and there exists no known
analytical method to solve them. This poses a major problem in fluid flow
analysis. However, the problem is tackled by making some simplifications to
the equation depending on the type of flow to which it is to be applied. For
certain flows the equation can be reduced to an ordinary differential equation
of a simple linear type. For some other type of flows it can be reduced to
a nonlinear ordinary differential equation. For the above types of Navier–
Stokes equation governing special category of flows such as potential flow,

ISTUDY
2.7. KINEMATICS OF FLUID FLOW 51

fully developed flow in a pipe and channel, and boundary layer flows, it is
possible to obtain solutions.
It is essential to understand the physics of the flow before reducing the
Navier–Stokes equations to any useful form, by making suitable approxima-
tions with respect to the flow. For example, let us examine the flow over an
aircraft wing as shown in Figure 2.6.

FIGURE 2.6 Flow past an aircraft wing.

This kind of problem is commonly encountered in fluid mechanics. Air flow


over the wing creates a pressure at the bottom which is larger than that
at the top surface. Hence, there is a net resultant force component normal
to the freestream flow direction called lift, L. The velocity varies along the
wing chord as well as in the direction normal to its surface. The former
variation is due to the shape of the aerofoil and the later is due to the no-slip
condition at the wall. In the direction normal to the wing surface the velocity
gradients are very large near the wall and the flow reaches asymptotically
a constant velocity in a short distance. This thin region adjacent to the
wall where the flow velocity increases from zero to the freestream value is
known as the boundary layer. Inside the boundary layer the viscous forces
are predominant. Further, it so happens that the static pressure outside the
boundary layer, acting in the direction normal to the surface, is transmitted
through the boundary layer without any appreciable change. In other words,
the pressure gradient across the boundary layer is zero. Neglecting the inter-
layer friction between the streamlines, in the region outside the boundary
layer, it is possible to treat the flow outside the boundary layer as inviscid.
Inviscid flow is also called potential flow, and for this case the Navier–Stokes
equation can be made linear. It is possible to obtain the pressures in the field
outside the boundary layer and take this pressure to be invariant across the
boundary layer, i.e. the pressure in the freestream is “impressed through”
the boundary layer. For low-viscous fluids, such as air, we can assume with
a high degree of accuracy that the flow is frictionless over the entire flow
field except for a thin region near the solid surfaces. In the vicinity of a
solid surface, owing to high velocity gradients, the frictional effects become
significant. Such regions near the solid boundaries where the viscous effects
are predominant are termed boundary layers. A comprehensive picture of the
boundary layer can be seen in Chapter 4.

ISTUDY
52 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

In general, for streamlined bodies these boundary layers are extremely


thin. There may be laminar and turbulent flow within the boundary layer, and
its thickness and profile may change along the direction of the flow. Consider
the flow over a flat plate shown in Figure 2.7, wherein different zones of
boundary layer over the flat plate are shown. The laminar sublayer is that
zone near the boundary where the turbulence is suppressed to such a degree
that laminar effects predominate. This layer right next to the wall where
viscous effects dominate is also called the viscous sublayer. The various regions
shown in Figure 2.7 are not sharp demarcations of different zones. There is
actually a smooth variation from a region where a certain effect predominates
to another region where some other effect is predominant.

FIGURE 2.7 Flow over a flat plate.

Although the boundary layer is thin it plays a vital role in fluid dynamics.
The drag on ships and missiles, the efficiency of compressors and turbines used
for jet engines, the effectiveness of ram and turbojets and numerous other
engineering devices, i.e. the performance of each of these devices depends on
the behaviour of the boundary layer and its effect on the main flow. The
following are some of the important parameters associated with boundary
layers.

Boundary layer thickness


The boundary layer thickness δ may be defined as the distance from the wall
out to where the fluid velocity is within 1 per cent of the local main stream
velocity. It may also be defined as the distance δ normal to the surface, in
which the flow velocity rises from zero to some specified value (e.g. 99%) of its
local main stream flow velocity. The boundary layer thickness may be shown
schematically as in Figure 2.8.

Displacement thickness
The displacement thickness δ ∗ may be defined as the distance by which the
boundary would have to be displaced if the entire flow were imagined to be
frictionless and the same mass flow maintained at any section.

ISTUDY
2.7. KINEMATICS OF FLUID FLOW 53

FIGURE 2.8 Illustration of boundary layer thickness.

Consider a unit width in the flow across an infinite flat plate at zero angle
of attack, and let the x-component of velocity be Vx and the y-component Vy .
The volume flow rate ∆q̇ through this boundary layer segment of unit width
is given by ∫ ∞
∆q̇ = (Vm − Vx ) dy
0
where Vm is the main stream frictionless velocity component and Vx is the
actual velocity component. To maintain the same volume flow rate q̇ for the
frictionless case as in the actual case, the boundary must be shifted out by a
distance δ ∗ so as to cut off the amount ∆q̇ of flow. Thus,
∫ ∞

(Vm )(δ ) = ∆q̇ = (Vm − Vx )dy
0

or ∫∞
δ∗ = 0
(1 − Vx /Vm ) dy (2.24)

The displacement thickness is illustrated in Figure 2.9. The main idea of this
postulation is to permit the use of a displaced body in place of the actual body
such that the frictionless mass flow around the displaced body is the same
as the actual mass flow around the real body. The displacement thickness
concept is made use of in the design of wind tunnels, air intakes for jet engines,
etc.
There are other (thickness) measures pertaining to the thickness of bound-
ary layer, such as momentum thickness θ, and energy thickness δe . They are
defined mathematically as follows:
∫ ∞( )
Vx ρVx
θ= 1− dy (2.25)
0 Vm ρm Vm
∫ ∞( )
Vx2 ρVx
δe = 1− 2 dy (2.26)
0 Vm ρm Vm

where Vm and ρm are the velocity and density at the edge of the boundary

ISTUDY
54 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

FIGURE 2.9 Illustration of displacement thickness.

layer and Vx and ρ are the velocity and density at any y. In addition to
boundary layer thickness, the displacement thickness, the momentum thick-
ness and the energy thickness, we can also define the transition point and the
separation point with the help of boundary layer.

Transition point
Transition point may be defined as the end of the region at which the flow in
the boundary layer on the surface ceases to be laminar and becomes turbulent.

Separation point
Separation point is the position at which the boundary layer leaves the surface
of a solid body. If the separation takes place while the boundary layer is still
laminar, the phenomenon is termed laminar separation. If it takes place for
a turbulent boundary layer it is called turbulent separation.
The boundary layer theory makes use of Navier–Stokes equation
(Eq. (2.23)) with the viscous terms in it, but in a simplified form. On the
basis of many assumptions (such as the boundary layer thickness is small com-
pared to the body length, and similarity between velocity profiles in a laminar
flow), the Navier–Stokes equation can be reduced to a nonlinear ordinary dif-
ferential equation, for which special solutions exist. Some such problems, for
which the Navier–Stokes equations can be reduced to boundary layer equa-
tions and closed form solutions can be obtained, are flow past a flat plate or
Blassius problem, Hagen–Poiseuille flow through pipes, Couette flow and flow
between rotating cylinders.

EXAMPLE 2.5 Fluid flows down a wide inclined plate under the action of
gravity. The angle of inclination of the plate with horizontal is θ. The depth
of liquid normal to the plate is h. Determine the velocity profile considering

ISTUDY
2.7. KINEMATICS OF FLUID FLOW 55

the flow to be steady, two-dimensional, laminar, and incompressible. Assume


that the velocity is parallel to the plate.

Solution The flow field given is as shown in Figure 2.10.

FIGURE 2.10 Flow over an inclined plate.

The governing momentum equations for the flow described in the present
problem, as obtained from Eqs. (2.23), are:

∂Vx ∂Vx 1 ∂p µ 2
Vx + Vy =X− + (∇ Vx )
∂x ∂y ρ ∂x ρ

∂Vy ∂Vy 1 ∂p µ 2
Vx + Vy =Y − + (∇ Vy )
∂x ∂y ρ ∂y ρ
where X and Y are the body forces acting along the x and y directions,
respectively. The flow is parallel to the plane and is independent of x.
Therefore,
Vx = f (y), Vy = 0
With these, the above momentum equations reduce to

µ d2 Vx
0=X+
ρ dy 2
1 dp
0=Y −
ρ dy

The body force components are X = g sin θ and Y = g cos θ. Therefore,

dp
= ρg cos θ
dy
For small values of θ, dp/dy = ρg. This is the hydrostatic variation of pressure
in the y-direction. The integration of x-momentum equation yields

ρg sin θ y 2
Vx = − + C1 y + C2
µ 2

ISTUDY
56 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

where C1 and C2 are the integration constants to be evaluated as follows: By


no-slip condition, at y = 0, Vx = 0. With this condition, C2 = 0. At y = h,
τyx = 0.
∂Vx
τyx = µ
∂y
Therefore, dVx /dy = 0 at y = h. This gives
ρgh sin θ
C1 =
µ
Hence the velocity profile is
ρg sin θ
Vx = (2hy − y 2 )

EXAMPLE 2.6 A paper coating process consists of pulling a continuous


sheet of paper upwards through the surface of a pool of coating liquid at a
speed V = 0.1 m/s. A film of coating liquid of constant thickness h adheres
to the paper, as shown in Figure 2.11. The liquid speed u(h) at the outside
edge of the liquid layer (y = h) is zero. Calculate the film thickness h, if the
liquid kinematic viscosity ν = 1 × 10−6 m2 /s.

FIGURE 2.11 Paper with liquid film over it.

Solution Taking the x-axis vertically upwards and g downwards, the Navier–
Stokes equation in the vertical direction is
∂2u
0 = −g + ν
∂y 2
or
d2 u g
2
=
dy ν

ISTUDY
2.7. KINEMATICS OF FLUID FLOW 57

Therefore,
gy 2
u= + c1 y + c2

At y = h, du/dy = 0, therefore,

gh
0= + c1
ν
At y = 0, u = V , therefore,

V = 0 + 0 + c2
Thus, ( )
g y2
u= − hy + V
ν 2
If u = 0 at y = h, then

−gh2
0= +V

Therefore,

2νV
h=
g

2 × 1 × 10−6 × 0.1
=
9.81

= 1.428 × 10−4 m

EXAMPLE 2.7 A newtonian fluid having a specific gravity of 0.92 and a


kinematic viscosity of 4 × 10−4 m2 /s flows past a fixed surface. The velocity
profile near the surface is shown in Figure 2.12. Determine the magnitude and
direction of the shearing stress developed on the plate. Express your answer
in terms of U and δ, with U and δ expressed in units of metre per second and
metre, respectively.

FIGURE 2.12 Velocity profile.

ISTUDY
58 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

Solution Given,

ρ = 0.92 × 103 kg/m3


µ
ν = = 4 × 10−4 m2 /s
ρ
Shear stress is
∂u
τw = µ at y = 0
∂y
Thus,
πU
τw = µ
2 δ
πU
= νρ
2 δ
π U
= × 0.92 × 103 × 4 × 10−4 ×
2 δ

U
= 0.578 N/m2 (acting to right on the plate)
δ

2.7.1 Rotational and Irrotational Motion


In a fluid flow if the possible distortion of the fluid element by severe viscous
traction is ignored (this assumption is valid in much of aerodynamic theory),
there remain only three possible ways in which a particle can move.

1. In pure translation—the fluid element is free to move anywhere in space


but continues to keep its axes parallel to reference axes fixed in space.
Flow field outside the boundary layer over an aerofoil is a typical exam-
ple of pure translation.
2. In pure rotation—the fluid element rotates about its own axis which
remains fixed in space.
3. In the general motion—into which (1) and (2) above are compounded.
Such a motion is found in the wake of a bluff body.

When a fluid element is subjected to a shearing force, a velocity gradient


is produced perpendicular to the direction of shear, i.e. a relative motion
occurs between two layers. To achieve this relative motion the fluid elements
have to undergo rotation. A typical example of this type of motion is the
motion between two roller chains rubbing each other but moving at different
velocities. It is convenient to use an abstract quantity called circulation Γ,
defined as the line integral of velocity vector between any two points, to define
rotation of the fluid element. By definition, the circulation is given as
H
Γ ≡ C V · dl (2.27)

ISTUDY
2.7. KINEMATICS OF FLUID FLOW 59

where dl is an elemental length, and the loop through the integral sign signifies
that the contour is closed.
By Stokes’ theorem, we have
I ∫
V · dl = (curl V ) · dA
C A

which states that, “the line integral of V around a closed curve C is equal to
the ‘flux’ of curl V through an arbitrary surface A bounded by C”. Note that
the word ‘flux’ is generally used to mean the integral of a vector field normal
to a surface. Thus, Eq. (2.27) becomes

Γ = (curl V ) · dA
A

But curl V is known as the vorticity. Thus,



Γ= A
ζ · dA

Thus, the circulation around a closed curve is equal to the surface integral
of the vorticity, which can be termed flux of vorticity. In other words, the
vorticity at a point equals the circulation per unit area.
Circulation per unit area is known as the vorticity ζ, i.e.

ζ = Γ/A (2.28)

In vector form, ζ becomes

ζ = ∇ × V = curl V (2.29)

For a two-dimensional flow in xy-plane, ζ becomes


∂Vy ∂Vx
ζz = − (2.30)
∂x ∂y
where ζz is the vorticity about the z-direction which is normal to the flow
field. Similarly, the vorticity about the x and y directions can be respectively
written as
∂Vz ∂Vy
ζx = −
∂y ∂z
and
∂Vx ∂Vz
ζy = −
∂z ∂x
If the vorticity components are zero, the flow is known as irrotational flow.
Inviscid flows are essentially irrotational flows. An irrotational flow is that
in which each element of the moving fluid suffers no net rotation from one
instant to the next, with respect to a given frame of reference. The classic
example of irrotational motion (although not a fluid) is that of carriages on a

ISTUDY
60 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

giant wheel used for amusement rides. Each carriage describes a circular path
as the wheel revolves, but does not rotate with respect to the earth. However,
in an irrotational flow, a fluid element may deform, causing the axes of the
elements rotate equally towards or away from each other, so that there is no
net rotation. As long as the algebraic average rotation (or angular velocity)
is zero, the motion is irrotational.

EXAMPLE 2.8 Check whether the two-dimensional flow field given by

V = 2x3 yi − 3x2 y 2 j

is irrotational.

Solution We know that the vorticity is finite for rotational flows and zero
for irrotational flows.
The vorticity [i.e. circulation per unit area, by Eq. (2.30)] is

∂Vy ∂Vx
ζz = −
∂x ∂y

For the given flow field, Vx = 2x3 y and Vy = 3x2 y 2 . Therefore,


∂Vy ∂Vx
= −6xy 2 , = 2x3
∂x ∂y
Substituting these in the vorticity equation, we get

ζz = −6xy 2 − 2x3 ̸= 0

The vorticity is finite and, hence, the flow is rotational.

EXAMPLE 2.9 A large tank is fixed to a cart as shown in Figure 2.13.


Water flows from the tank through a nozzle of diameter 30 mm at a speed of
8 m/s. The water level in the tank is maintained constant by adding water
through a vertical pipe. Determine the tension in the wire holding the cart
stationary.

FIGURE 2.13 Tank on a moving cart.

ISTUDY
2.7. KINEMATICS OF FLUID FLOW 61

Solution Consider the control volume at the outer edge of the tank–cart
system, as shown in Figure 2.14.

FIGURE 2.14 Control volume for the tank–cart system.

The momentum of the water flow is the only force acting on the control
volume. Thus,

Fx = (ρAV ) V = ρAV 2
( )
π × 0.032
= 103 82
4

= 45.24 N

Therefore, the wire tension T = 45.24 N

EXAMPLE 2.10 A large tank is mounted on rollers, as shown in


Figure 2.15, and a water jet discharges water into the tank at the rate of
10 kg/s with a velocity of 25 m/s. Calculate the force Fx necessary to
(a) hold the tank stationary, and (b) to allow it to move to the left with
a steady velocity of 3 m/s.

FIGURE 2.15 Tank on rollers.

ISTUDY
62 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

Solution Since there is accumulation of mass, and therefore momentum, in


the tank with time, it is an unsteady problem. Selecting the control surface
as shown by dashed lines, the x-momentum equation can be written as
[ ]
d(ṁVx )cv
Fx = (ṁVx )e − (ṁVx )i +
dt

where the subscripts e, i, and cv refer to exit, inlet, and control volume,
respectively. But (ṁVx )e = 0, therefore,

(ṁVx )i = 10 (−25 × cos 30◦ )


= − 216.5 (kg m)/s2

(a) The tank is at rest and (Vx )cv = 0. Therefore, there is no accumulation
of momentum, even though there is accumulation of mass. For this case,

Fx = [0 − (−216.5) + 0]

= 216.5 N

(b) The tank is moving to the left, and there is a rate of increase in momentum
in the control volume given by

d(ṁVx )cv
= 10 (−3) = −30 (kg m)/s2
dt
Thus,

Fx = [0 − (−216.5) + (−30)]

= 246.5 N

2.8 STREAM FUNCTION


Streamlines are imaginary lines in the flow field such that the velocity at
all points on these lines is always tangential to the lines. Flows are usually
depicted graphically with the aid of streamlines. Streamlines proceeding thro-
ugh the periphery of an infinitesimal area at some time ‘t’ form a tube called
streamtube, which is useful for the study of fluid flow phenomena. From the
definition of streamlines, it can be inferred that:

• Flow cannot cross a streamline, and the mass flow between two stream-
lines is confined.

• Based on the streamline concept, a mathematical function ψ called


stream function can be defined. The velocity components of a flow field
can be obtained by differentiating the stream function.

ISTUDY
2.8. STREAM FUNCTION 63

In terms of stream function ψ the velocity components of a two-dimensional


incompressible flow are given as

∂ψ ∂ψ
Vx = , Vy = − (2.31)
∂y ∂x

If the flow is compressible the velocity components become

1 ∂ψ 1 ∂ψ
Vx = , Vy = − (2.32)
ρ ∂y ρ ∂x

It is important to note that the stream function is defined only for two-
dimensional flows and the definition does not exist for three-dimensional flows.
Even though some books define ψ for axisymmetric flows, they again prove to
be equivalent to two-dimensional flows. We must realize that the definition
of ψ does not exist for three-dimensional flows. This is because the definition
of streamline demands a single tangent at any point on a streamline, which is
possible only in two-dimensional flows.
It is sometimes convenient to work with polar coordinates, especially in
respect of problems involving circular boundaries. The expressions for velocity
components in terms of the stream function ψ given by Eq. (2.31) can be stated
in terms of the polar coordinates for such problems. Consider the continuity
equation for a two-dimensional incompressible flow in Cartesian coordinates,
∂Vx ∂Vy
+ =0
∂x ∂y
Now consider the following relation between the Cartesian and polar
coordinates,
x = r cos θ θ = tan−1 (y/x)

y = r sin θ r = x2 + y 2
Also, we know that ψ = f (x, y), and x and y themselves are functions of
r and θ.
In cylindrical polar coordinates the continuity equation for a two-dimensional
incompressible flow can be written as

1 ∂(rVr ) 1 ∂Vθ
+ =0
r ∂r r ∂θ
and the velocity components Vr and Vθ can be related to the stream function
ψ(r, θ), as
1 ∂ψ ∂ψ
Vr = , Vθ = −
r ∂θ ∂r

Substitution of these expressions for the velocity components into the conti-
nuity equation satisfies the continuity equation.

ISTUDY
64 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

EXAMPLE 2.11 An incompressible flow is squeezed outwards between


two large circular disks by uniform downward motion with velocity V0 of the
upper disk, as shown in Figure 2.16. The bottom disk is fixed. Assuming
one-dimensional radial outflow, derive an expression for V (r) in terms of h.

FIGURE 2.16 Flow between disks.

Solution The flow is incompressible, therefore, ρ = constant. For the


moving boundary the velocity V0 is uniform.
By continuity equation, we have
∫∫∫
∂ π
ρ D2 h = −ρV (r) πDh
∂t 4
where D is the disk diameter and is a constant. Hence, this equation simplifies
to
∂h D
= −V (r) h
∂t 4
where h is a function of time. Thus,
D
V0 = −V (r)h
4
Since,
∂h
= V0
∂t
we have
V0 D
V (r) =
4h

EXAMPLE 2.12 For steady compressible flow of a gas through a nozzle,


the quasi one-dimensional velocity distribution is given by u = U0 (1 + x/L),
where U0 and L are constants. Using continuity, find an expression for density
distribution ρ(x), if ρ = ρ0 at x = 0. At what position of x will the density
drop to 25 per cent below ρ0 ?

Solution By continuity equation, we have


∂ ∂ ∂
(ρ u) + (ρ v) + (ρ w) = 0
∂x ∂y ∂z

ISTUDY
2.8. STREAM FUNCTION 65

In this equation, the second and third terms become zero due to one-dimensional
nature of the flow. Therefore,

ρ u = constant

i.e.
ρ U0 (1 + x/L) = constant (i)
At x = 0, ρ = ρ0 . Using this condition in the above equation, we get

constant = ρ0 U0
Substituting this into Eq. (i), we get
ρ0
ρ(x) =
1 + x/L

Let x1 be the location at which ρ = 0.75 ρ0 . From the above equation for ρ,
we have
ρ0
0.75 ρ0 =
1 + x1 /L
This gives,
L
x1 =
3

EXAMPLE 2.13 A wooden beam of specific gravity 0.6 is 10 cm by


10 cm in cross-section and 5 m long. It is hinged at point A, as shown in
Figure 2.17. Find the angle θ at which the beam will float in the water. Take
the density of water to be 1000 kg/m3 .

FIGURE 2.17 A hinged beam in water.

Solution The forces acting on the wooden beam are its weight W and the
buoyant force FB .

ISTUDY
66 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

Volume of the wooden beam is

V = 0.1 × 0.1 × 5 = 0.05 m3

Weight of the wooden beam is

W = 0.6 × 103 × V × g
= 0.6 × 103 × 0.05 × 9.81
= 294.3 N

The buoyant force is given by

FB = ρgV
= (1000)(9.81) (0.1 × 0.1 × l)
= 98.1 l N

Taking moments about A, we have

ΣMA = 0 = (98.1 × l)(5 − l/2)cos θ = (294.3)(2.5 cos θ)

Solving for l, we get


l = 1.838 m
and
1
sin θ =
5 − 1.838
This gives
θ = 18.4◦

2.8.1 Relationship Between Stream Function and


Velocity Potential
For irrotational flows (the fluid elements in the flow field are free of angular
motion), there exists a function ϕ called the velocity potential or potential
function. For two-dimensional flows, ϕ must be a function of x, y, and t. The
velocity components are given by
∂ϕ ∂ϕ
Vx = , Vy = (2.33)
∂x ∂y
From Eqs. (2.31) and (2.33), we can write

∂ψ ∂ϕ ∂ψ ∂ϕ
= , =− (2.34)
∂y ∂x ∂x ∂y

These relations between stream function and potential function given by


Eq. (2.34) are the famous Cauchy–Riemann equations of complex-variable
theory. It can be shown that the lines of constant ϕ or potential lines form

ISTUDY
2.8. STREAM FUNCTION 67

a family of curves which intersect the streamlines in such a manner as to


have the tangents of the respective curves always at right angles at the point
of intersection. Hence the two sets of curves given by ψ = constant and
ϕ = constant form an orthogonal grid system or flow net.
Unlike the stream function, the potential function exists for three-
dimensional flows as well. This is because there is no condition (like the
local flow velocity must be tangential to the potential lines) imposed on the
definition of ϕ. The only requirement for the existence of ϕ is that the flow
must be potential.

EXAMPLE 2.14 A two-dimensional flow field is specified by V = 3yi+3xj.


State whether the flow is steady, irrotational, and check whether the given
field is feasible. Find the stream function and determine the volume flow rate
passing between streamlines through the points (1, 3) and (3, 3).

Solution Given, Vx = 3y and Vy = 3x. This field is steady since ∂V /∂t =


0. From Eq. (2.30) the vorticity is given by

∂Vy ∂Vx
ζz = − =3−3=0
∂x ∂y
Therefore, the field is irrotational.
For this field to be a feasible one, it must satisfy the continuity equation.
For steady, incompressible, two-dimensional flow the continuity equation in
the differential form is given by Eq. (2.20) as

∂Vx ∂Vy
+ =0
∂x ∂y
The present field satisfies the above continuity equation and hence is feasible.
Let ψ(x, y) be the stream function of the given field. In differential form,

∂ψ ∂ψ
dψ = dx + dy = −Vy dx + Vx dy
∂x ∂y
= −3xdx + 3ydy

Integrating this equation, we get the relation


3 2 3 2
ψ= y − x + constant
2 2
The discharge between streamlines through (1, 3) and (3, 3) is

q = qψ3,3 − qψ1,3 = ψ3,3 − ψ1,3

where q is the volume flow rate. Therefore,


3 2 3
q= (3 − 32 ) − (32 − 12 )
2 2

ISTUDY
68 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

i.e.
q = –12 units flowing from right to left.

EXAMPLE 2.15 Show that the velocity vector V is everywhere tangential


to the streamlines in the x–y plane.

Solution Let ψ(x, y) = constant, be the stream function representing the


streamlines in the x–y plane.
The stream function in general can be expressed by

∂ψ ∂ψ
dψ = dx + dy
∂x ∂y

Along the lines with ψ(x, y) = constant, dψ = 0. Therefore,

∂ψ ∂ψ
dx + dy = 0
∂x ∂y

By Eq. (2.31), Vx = ∂ψ/∂y and Vy = −∂ψ/∂x. Substituting these, the


above equation becomes

Vx dy − Vy dx = 0
or
dy Vy
=
dx Vx

This implies that V is tangent to lines of ψ(x, y) = constant.

EXAMPLE 2.16 A velocity field is given by u = V cos θ, v = V sin θ, and


w = 0, where V and θ are constants. Find an expression for the streamlines
of this flow.

Solution The equation of the streamlines is given by

dx dy dz
= =
u v w
dx dy dz
= =
V cos θ V sin θ 0

dy V sin θ
=
dx V cos θ

= tan θ

ISTUDY
2.9. POTENTIAL FLOW 69

Therefore,

y = x tan θ + constant

These lines are the lines which are inclined at an angle θ to the x-axis.

2.9 POTENTIAL FLOW


Potential flow is based on the concept that the flow field can be represented
by a potential function ϕ such that

∇2 ϕ = 0 (2.35)

This linear partial differential equation is popularly known as Laplace equa-


tion. The derivatives of ϕ give velocities, as given in Eq. (2.33), for a two-
dimensional flow. Unlike the stream function ψ, the potential function can
exist only if the flow is irrotational, that is, when the viscous effects are absent.
All inviscid flows must satisfy the irrotationality condition, namely

∇×V =0 (2.36)

For two-dimensional potential flows, by Eq. (2.30), we have the vorticity ζz


as

∂Vy ∂Vx
ζz = − =0
∂x ∂y
Equation (2.36) can be rewritten, using Eq. (2.33), as

∂2ϕ ∂2ϕ
− =0
∂x∂y ∂x∂y

This shows that the flow is irrotational. For two-dimensional flows, the con-
tinuity equation given by Eq. (2.22) simplifies to

∂Vx ∂Vy
+ =0
∂x ∂y
Using Eq. (2.33) this equation can be expressed as

∂2ϕ ∂2ϕ
+ 2 =0
∂x2 ∂y
i.e.
∇2 ϕ = 0 (2.37)

For flows with finite vorticity the potential function ϕ does not exist and
the linear equation ∇2 ϕ = 0 cannot be obtained.

ISTUDY
70 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

For potential flows, the Navier–Stokes equations (2.23) reduce to the form
∂Vx ∂Vx ∂Vx 1 ∂p
Vx + Vy + Vz =−
∂x ∂y ∂z ρ ∂x
∂Vy ∂Vy ∂Vy 1 ∂p
Vx + Vy + Vz =− (2.38)
∂x ∂y ∂z ρ ∂y
∂Vz ∂Vz ∂Vz 1 ∂p
Vx + Vy + Vz =−
∂x ∂y ∂z ρ ∂z
Equations (2.38) are known as Euler’s equations.
The Laplace equation (2.37) represents potential flows which are imprac-
tical. Therefore, it is a natural to have a doubt that “what is the use of such
an impractical equation?”. The answer to this question is the following. Even
though it describes an impractical flow field, it plays a dominant role in fluid
flow analysis. The basic solutions to this equation can be suitably combined
to represent any shape of engineering interest, along with the flow field around
it. The Laplace equation has many basic and simple solutions. Some of the
popular solutions are the source, sink, their combination, parallel flows, and
flows induced by vortex with vortex core neglected. These solutions could
be superimposed. With a suitable combination of these simple mathematical
models, many complicated flows encountered in practice can be simulated.
The stream function of these flows can be derived easily. The combination of
sources and sinks produces closed flows and by properly distributing sources
and sinks of different suitable strength, the flow over any desired body can be
obtained. That is, potential flow over an arbitrary body such as an airplane
can be simulated by adding sources and sinks, provided the right combina-
tion is known. In other words, the Laplace equation can serve to replace an
engineering shape by mathematical functions so that they can be analyzed
to get the flow properties over the shape of interest. The stream function
obtained from the addition of many stream functions is sufficient to solve the
potential flow, since the velocity can be obtained by differentiating the stream
function. Once the velocity is known the pressure can be determined from the
Euler’s equation (2.38). Thus, the Laplace equation forms the basis for fluid
dynamic computation.
A function satisfying the Laplace equation is sometimes called the harmonic
function. The Laplace equation is encountered not only in potential flows, but
also in heat conduction, elasticity, magnetism, and electricity. Therefore,
the solution in one field of study can be found from a known analogous
solution in another field. The Laplace equation is of elliptic type. It is
known that the solution of elliptic equations are smooth and do not have
discontinuities, except for certain singular points on the boundary of the
region, unlike hyperbolic equations such as wave equation which can have
discontinuous ‘wavefronts’ in the middle of a region.
The boundary conditions usually encountered in irrotational flows are:
Condition on solid surface: The velocity normal to a solid surface must
be equal to the velocity of the boundary normal to itself. This ensures that

ISTUDY
2.9. POTENTIAL FLOW 71

the fluid does not penetrate a solid boundary. For a stationary surface, the
condition is
∂ϕ ∂ψ
=0 or =0
∂n ∂s
where s is direction along the surface and n is normal to the surface.
Condition at infinity: For a body immersed in a uniform stream flowing
in the x-direction with speed U , the condition is
∂ϕ ∂ψ
=U or =U
∂x ∂y
Solving the Laplace equation with the boundary conditions of the above type
is not easy. Historically, the potential flow theory was developed by find-
ing a function that satisfies the Laplace equation and then determining what
boundary conditions are satisfied by that function. As the Laplace equa-
tion is linear, superposition of a known harmonic function gives another har-
monic function satisfying a new set of boundary conditions. After obtaining a
solution of the Laplace equation, the velocity components are determined by
taking the derivatives of ϕ or ψ. Finally, the pressure distribution is obtained
using the Bernoulli equation

p + 12 ρV 2 + ρgz = constant (2.39)

between any two points in the flow field. Thus, a solution of the non-
linear equation of motion(Euler equation) is obtained in irrotational flows in a
simpler manner.

2.9.1 Bernoulli Equation


The Bernoulli equation is a statement of conservation of mechanical energy
between any two points along a streamline. The constant in the equation,
known as the Bernoulli constant, in general varies from one streamline to
another but remains constant along a streamline in a steady, frictionless,
incompressible flow. These four assumptions are needed and must be kept in
mind while applying this equation. Each term in Eq. (2.39) has the dimensions
(L/T )2 or the units metre-newton per kilogram, i.e.

m·N (m · kg/s2 )(m) m2


= = 2
kg kg s
Therefore, the Bernoulli equation is interpreted as energy per unit mass.
When it is divided by g,
p V2
+ + z = constant
ρg 2g
it can be interpreted as energy per unit weight, metre-newton per newton.
This form is convenient for dealing with liquid problems with a free surface.

ISTUDY
72 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

Each term in the Bernoulli equation may be interpreted as a form of avail-


able energy. This equation is also referred to as a conservation of mechanical
energy equation. It is essential here to realize that this energy equation was
derived from the conservation of momentum equation. Thus, the Bernoulli
equation is essentially a momentum equation and not an energy equation.

Cautions for the use of Bernoulli equation


The Bernoulli equation is a useful equation because it relates pressure changes
to velocity and elevation changes along a streamline. But it gives correct
results only when applied to flows where all the four restrictions, namely
(a) steady flow, (b) incompressible flow, (c) frictionless flow, and (d) flow
along a streamline, are reasonable.
Flow through a nozzle can be modelled well by the Bernoulli equation.
Because the pressure gradient in a nozzle is favourable, there is no separation
and the boundary layers on the walls remain thin. Friction has negligible
effect on the flow velocity profile, therefore, one-dimensional flow is a good
model. The velocity at any section can be calculated from the corresponding
flow area.
Flow through a diffuser or a suddenly expanded duct should not be
modelled using the Bernoulli equation. Adverse pressure gradient prevailing
in such flows can cause rapid growth of boundary layers, severely distorted
velocity profiles, and possible flow separation. One-dimensional model is a
poor model for such flows. Because of area blockage due to boundary layer
growth, the pressure rise in an actual diffuser is always less than that predicted
by inviscid one-dimensional flow model.
Bernoulli equation would be a reasonable model for passages like siphon
tube with well-rounded entrance, gentle bends, and short length. Flow sepa-
ration that can occur at inlets with sharp corners and in abrupt bends causes
the flow to depart from that predicted by the one-dimensional model and the
Bernoulli equation. Friction effect would not be negligible if the tube were
long.
Bernoulli equation is a good flow model for open-channel flow without
hydraulic jump, since it is analogous to that in a nozzle. The hydraulic jump
causes an adverse pressure gradient. Flow through a hydraulic jump is mixed
violently, making it impossible to identify streamlines. Thus, the Bernoulli
equation cannot be used to model flow through a hydraulic jump.
Bernoulli equation is not applicable to flow through a machine such as
a propeller, pump, or windmill. The equation was derived by integrating
along a streamline or a streamtube. It is impossible to have steady flow or to
identify streamlines during flow through a machine. However, the Bernoulli
equation may be applied for the flow field ahead of and behind a machine if
the restrictions for using the equation are satisfied.
The compressibility must be considered for flow of gases when the Mach
number is more than 0.3. For flow Mach numbers less than 0.3, the density
changes caused by dynamic compression due to the flow motion are less than

ISTUDY
2.9. POTENTIAL FLOW 73

5 per cent (for air at sea-level condition) and hence can be neglected. But
when the density changes are more than 5 per cent, as in air flow with Mach
number greater than 0.3, such changes must be accounted for. For such flows,
the Bernoulli equation in its general form, namely
dp
+ V dV + g dz = 0
ρ
can be integrated if the relation between p and ρ is known. For example, for
an isentropic flow, the process equation is given by
p
= constant
ργ
where γ is the ratio of specific heats. For isentropic flow of a perfect gas with
negligible change in potential energy (which is the case usually), the Bernoulli
equation can be integrated to result in (E. Rathakrishnan, Gas Dynamics
Prentice-Hall of India, New Delhi, 1995)

γ p V2 γ p0
+ =
γ−1ρ 2 γ − 1 ρ0

where p0 and ρ0 are the stagnation pressure and density, respectively. This is
the well known compressible Bernoulli equation.
The continuity, irrotationality, and Laplace equation can be expressed in
terms of
∂ϕ 1 ∂ψ
Vr = =
∂r r ∂θ
and
1 ∂ϕ ∂ψ
Vθ = =−
r ∂θ ∂r
as follows:

Continuity equation:
1 ∂ 1 ∂Vθ
(rVr ) + =0 (2.40)
r ∂r r ∂θ
Irrotationality equation:
1 ∂ 1 ∂Vr
(rVθ ) − =0 (2.41)
r ∂r r ∂θ
Laplace equation:
( )
1 ∂ ∂ϕ 1 ∂2ϕ
∇2 ϕ = r + 2 2 =0 (2.42)
r ∂r ∂r r ∂θ

( )
1 ∂ ∂ψ 1 ∂2ψ
∇ ψ=
2
r + 2 2 =0 (2.43)
r ∂r ∂r r ∂θ

ISTUDY
74 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.9.2 Two-Dimensional Source and Sink


A type of flow in which the fluid emanates from the origin and spreads radially
outwards to infinity, as shown in Figure 2.18, is called a source. The volume
flow rate q̇ crossing a circular cylindrical surface of radius r and unit depth is
given by

FIGURE 2.18 A two-dimensional source.

q̇ = 2πrVr (2.44)
where Vr is the radial component of velocity. For a source, the radial lines
are the streamlines. Therefore, the potential lines must be concentric circles
represented by
ϕ = A ln(r)
where A is a constant. The velocity component, Vr = ∂ϕ/∂r = A/r. Substi-
tuting this into Eq. (2.44), we get
2πrA/r = q̇
A = q̇/2π
Thus, the velocity potential for a two-dimensional source of strength q̇
becomes

ϕ= ln r (2.45)

In a similar manner as above, the stream function for a source of strength q̇
can be obtained as

ψ= θ (2.46)

where θ stands for the location of the streamline in the θ-direction. Similarly,
for a sink, which is a type of flow in which the fluid at infinity flows radially
towards the origin, we can show that the potential and stream functions are
given by

ϕ=− ln r

and

ψ=− θ

ISTUDY
2.9. POTENTIAL FLOW 75

where q̇ is the strength of the sink. Note that the volume flow rate is termed
the strength of source and sink. Also, for both source and sink the origin is a
singular point.

2.9.3 Simple Vortex


Vortex motion is a fluid flow in which the streamlines are concentric circles.
The vortex motions encountered in practice may in general be classified as
(a) free vortex or potential vortex and (b) forced vortex or flywheel vortex. A
vortex is designated as free or potential if the fluid elements in the field do not
rotate about their own axes. That is, in a free vortex the fluid elements move
along circular streamlines without rotating about their own axes. Thus, the
vorticity content of a free vortex is zero. In contrast, in addition to tracing
circular paths, the fluid elements spin about their own axes in a forced vortex.
Thus, the vorticity content of a forced vortex is finite. A simple vortex or free
vortex is a fluid flow in which the streamlines are concentric circles. A free
vortex may be established by selecting the stream function ψ of the source to
be the potential function ϕ of the vortex. Thus, for a simple vortex


ϕ= θ (2.47)

It can be easily shown from Eq. (2.47) that the stream function for a simple
vortex is

ψ = − ln r (2.48)

It follows from Eqs. (2.47) and (2.48) that the velocity components of the
simple vortex, shown in Figure 2.19, are

Vθ = and Vr = 0 (2.49)
2πr

FIGURE 2.19 Simple vortex flow.

It is seen that the tangential velocity is inversely proportional to the radial


distance. This makes the velocity to become infinity at the origin. Thus,
a singularity appears at the origin, where the tangential velocity approaches
infinity, as may be seen from Eq. (2.49). Also, around the axis the velocity

ISTUDY
76 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

assumes large magnitudes. The flow in a simple or free vortex resembles a


common whirlpool found while paddling a boat or while emptying water from
a bathtub. An approximate profile of a whirlpool is as shown in Figure 2.20.
The shaded portion around the axis of the whirlpool rotates like a solid body
and due to this the flow around the axis is rotational. Thus, the flow in
the free vortex, such as a whirlpool, is potential everywhere, except a small
portion around the axis. This rotational portion around the axis is called the
core.

FIGURE 2.20 A whirlpool flow field.

For the whirlpool shown in Figure 2.20, the circulation along any path about
the origin is given by
I ∫ 2π
Γ= V · dl = Vθ rdθ (2.50)
0


By Eq. (2.49), Vθ = , therefore, the circulation becomes
2πr

∫ 2π

Γ= rdθ = q̇
0 2πr

Since there are no other singularities for the whirlpool shown in Figure 2.20,
this must be the circulation for all paths about the origin. Consequently, q̇
in the case of vortex is the measure of circulation about the origin and is also
referred to as the strength of the vortex.

EXAMPLE 2.17 Consider a flow field with pure tangential (circular)


streamlines; Vr = 0 and Vθ = f (r). Evaluate the vorticity and circulation
for (a) a rigid-body rotation and (b) a free vortex. Show that it is possible to
choose f (r) so that the flow is irrotational, i.e. to produce a free vortex.

Solution For motion in the rθ-plane, the only components of rotation and
vorticity are about the z-direction.

ISTUDY
2.9. POTENTIAL FLOW 77

Vorticity is given by

ζ = 2ω = ▽ × V

1 ∂ (rVθ ) 1 ∂Vr
ζz = 2 ωz = −
r ∂r r ∂θ
But Vr = 0. Thus,
1 ∂ (rVθ )
ζz = 2 ωz =
r ∂r
(a) For rigid-body motion, Vθ = ωr.

Therefore, the angular velocity is given by


( )
1 1 ∂ (r Vθ ) 1 1 ∂ r2 ω
ωz = = = ω
2r ∂r 2 r ∂r

ζz = 2 ω

The circulation is given by


I ∫
Γ= V ds = 2ωz dA
C A

Since ωz = ω = constant, the circulation about any closed contour is given


by Γ = 2 ω A , where A is the area enclosed by the contour. Thus, for rigid-
body motion (a forced vortex), the rotation and vorticity are constants and
the circulation depends on the area enclosed by the contour.

(b) For irrotational flow,

1 ∂ (rVθ )
=0
r ∂r
Integrating, we get
rVθ = constant
or
c
Vθ = f (r) =
r
For this flow, the origin is a singular point where Vθ → ∞.

The circulation for any contour, enclosing the origin, is


I ∫ 2π
c
Γ= V ds = r dθ = 2πc
C 0 r

The circulation around any contour not enclosing the singular point at the
origin is zero.

ISTUDY
78 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.9.4 Source–Sink Pair


This is a combination of a source and a sink of equal strength situated at
a distance apart. The stream function due to this combination is obtained
simply by adding the stream functions of source and sink. When the distance
between the source and sink is made negligibly small, in the limiting case, the
combination results in a doublet.

2.9.5 Doublet
A doublet or a dipole is a potential flow field due to a source and a sink of equal
strength, brought together in such a way that the product of their strength
and the distance between them remains constant. Consider a point P in the
field of a doublet formed by a source and a sink of strength q̇ and − q̇, kept
at a distance ds, as shown in Figure 2.21, with sink at the origin.

FIGURE 2.21 Source and sink.

By Rankine’s theorem, the velocity potential of the doublet, ϕD , can be ex-


pressed as the sum of the velocity potentials of the source and sink. Thus, we
have
q̇ q̇
ϕD = ln (r + dr) − ln r
2π 2π
( )
q̇ r + dr
= ln
2π r
( )
q̇ dr
= ln 1 +
2π r
( )
dr
Expanding ln 1 + , we get
r
( ) ( )2
dr dr 1 dr
ln 1 + = − + ...
r r 2 r

ISTUDY
2.9. POTENTIAL FLOW 79

dr
But ≪ 1, therefore, neglecting the second and higher order terms, we get
r
q̇ dr
ϕD =
2π r
By the definition of doublet, ds → 0. Therefore,

dr = ds cos θ

Hence,

ϕD = ds cos θ
2πr
Also, for a doublet, by definition, q̇ ds = constant. Let this constant, known
as the strength of the doublet, be denoted by m, then

m = q̇ds

and
m
ϕD = cos θ (2.51)
2πr
In Cartesian coordinates the velocity potential for the doublet becomes
( )
m x
ϕD =
2π x + y2
2

From the above equations for ϕD , the expression for the stream function ψD
can be obtained as
m
ψD = − sin θ
2πr
( )
m y
=−
2π x2 + y 2
In the above discussion the source and the sink were placed on the x-axis.
Such a doublet will look like the field shown in Figure 2.22.
If the source and the sink are placed on the y-axis, the resulting expressions
for ϕD and ψD will become
( )
m m y
ϕD(yy) = sin θ =
2πr 2π x2 + y 2
( )
m m x
ψD(yy) = − cos θ = −
2πr 2π x2 + y 2
The flow field will look as shown in Figure 2.23.
The expression for ψD(yy)
( )
m x
ψD(yy) = −
2π x2 + y 2

ISTUDY
80 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

FIGURE 2.22 Doublet with the source and the sink on the y-axis.

FIGURE 2.23 Doublet streamlines.

can be arranged in the form


cx
ψD(yy) = −
x2 + y 2
where m/2π = c, is a constant.
The equation for ψD(yy) can be expressed as
c
x2 + y 2 + =0
2ψD(yy)
i.e. ( )2 ( )2
c 2 c
x+ +y =
2ψD(yy) 2ψD(yy)

Thus the streamlines represented by ψD(yy) = constant are circles with their
centres lying on the x-axis and are tangent to the y-axis at the origin
(Figure 2.23). The direction of flow at the origin is along the negative y-axis,
pointing outwards from the source of the limiting source–sink pair, which is
called the axis of the doublet.
The potential and stream functions for the concentrated source, sink, vor-
tex, and doublet are all singular at the origin. It will be shown in the following

ISTUDY
2.10. COMBINATION OF SIMPLE FLOWS 81

section that several interesting flow patterns can be obtained by superposing


a uniform flow on these concentrated singularities.

EXAMPLE 2.18 Water approaches the intake of a pump with the velocity
varying inversely as the square of the radial distance from the intake. At a
radial distance of 1.5 m, the velocity is found to be 0.68 m/s. What is the
acceleration of flow at a radial distance of 1 m from the intake? [Hint: The
streamlines are radial.]

Solution Let the velocity be V = c/r2 , where c is a constant and r is the


radial distance from the intake.
Since the streamlines are radial, the normal acceleration is zero. Therefore,
by Eq. (2.19), the acceleration a is
∂V ∂V
a= +V
∂t ∂s
since the flow velocity depends only on r. Assuming the flow to be steady,
∂V /∂t = 0. Hence, a = V (∂V /∂s). For the present problem, the distance
s = r. Therefore,
∂V ∂V 2c
= =− 3
∂s ∂r r
2
2c
a=− 5
r
At r = 1.5 m, V = 0.68 m/s. Hence

c = 0.68 × 1.52 = 1.53 m3 /s

Therefore, at r = 1 m, the acceleration becomes

a = −2 × 1.532 = −4.68 m/s2

The negative sign implies that the acceleration is towards the intake.

2.10 COMBINATION OF SIMPLE FLOWS


We saw in Section 2.9 that the flow past the practical shapes of interest can
be represented or simulated with a suitable combination of source, sink, free
vortex and uniform flow. In this section, let us discuss some such flow fields.

2.10.1 Flow Past a Half-Body


An interesting pattern of flow past a half-body, shown in Figure 2.24, can
be obtained by combining a source and a uniform flow parallel to the x-axis.
By definition, a given streamline is associated with one particular value of
the stream function and, therefore, if we join the points of intersection of
the radial streamlines (source) with the rectilinear streamlines (uniform flow)

ISTUDY
82 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

where the sum of the stream functions is a given constant value, the resulting
line will be a streamline of the combined flow pattern. If this procedure is
repeated for a number of values of the combined stream function, the result
will be a picture of the combined flow pattern.
The stream function for the combined flow due to a source of strength q̇
at the origin, immersed in a uniform flow of velocity V∞ , parallel to the x-axis
is

ψ = V∞ r sin θ + θ

The plot of streamlines will be as shown in Figure 2.24.

FIGURE 2.24 Irrotational flow past a two-dimensional half-body.

The boundary streamline is given by


ψ=
2
It is seen that S is the stagnation point where the uniform flow velocity cancels
the velocity of the flow from the source. If the polar coordinates of S is (a, π),
then the cancellation of velocity requires

1 ∂ψ q̇
Vr = = V∞ cos θ + =0
r ∂θ 2π r
i.e.

V∞ − =0
2π a
Thus,

a=
2πV∞
Therefore, the value of the stream function at the stagnation point is

q̇ θ q̇ q̇
ψS = V∞ r sin θ + = V∞ a sin π + π=
2π 2π 2

ISTUDY
2.10. COMBINATION OF SIMPLE FLOWS 83

The equation of the streamline passing through the stagnation point is


obtained by setting ψ = ψS = q̇/2, resulting in

q̇ θ q̇
V∞ r sin θ + = (2.52)
2π 2

A plot of this streamline is shown in Figure 2.24. It is a semi-infinite body


with a smooth nose, generally called a half-body. The stagnation streamline
divides the field into a region external to the body and a region internal to it.
The internal flow consists entirely of fluid emanating from the source, and the
external region contains the originally uniform flow. The half-body resembles
several shapes of theoretical interest, such as the front part of a bridge pier
or an aerofoil; the upper-half of the flow resembles the flow over a cliff or a
side contraction in a wide channel.
The half-width of the body is given as

q̇ (π − θ)
h = r sin θ =
2 πV∞
As θ → 0, the half-width tends to a maximum of hmax = q̇/(2 V∞ ), i.e. the
mass flux from the source is contained entirely within the half-body, and
q̇ = (2 hmax ) V∞ at a large downstream distance where u = V∞ .
The pressure distribution can be found from the Bernoulli’s equation
1 1
p+ ρ u2 = p ∞ + ρ V ∞
2
2 2
where p and u are the local pressure and velocity of the flow, respectively.
The pressure can be expressed through the non-dimensional pressure dif-
ference called the pressure coefficient, defined as
p − p∞
Cp = 1 2
2 ρ∞ V∞

where p and p∞ are the local and freestream static pressures, respectively, ρ∞
is the freestream density and V∞ is the freestream velocity.
A plot of Cp on the surface of the half-body is shown in Figure 2.25. It is
seen that there is positive pressure or compression near the nose of the body
and the pressure becomes negative (or suction) beyond it. The net pressure
force acting on the body can easily be shown to be zero by integrating p on
the surface. The half-body obtained by a linear combination of the individual
stream functions of source and uniform flow as per the Rankine’s theorem,
which states that the resulting stream function of n potential flows can be
obtained by combining the stream functions of the individual flows, is also re-
ferred to as Rankine’s half-body.

ISTUDY
84 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

FIGURE 2.25 Pressure distribution for potential flow over a half-body.

EXAMPLE 2.19 A two-dimensional source of strength 4.0 m2 /s is placed


in a uniform flow of velocity 1 m/s parallel to the x-axis. Determine the flow
velocity and its direction at r = 0.8 m and θ = 140◦ .

Solution The given flow is an ideal flow around a half-body. The stream
function for the flow around a half-body is given by

ψ = V∞ r sin θ + θ

Given that V∞ = 1 m/s and q̇ = 4 m2 /s, we have
4
ψ = r sin θ + θ

The tangential and radial components of velocity, respectively, are
∂ψ
Vθ = −
∂r
= − sin θ = − sin 140◦

= − 0.643 m/s

and
1 ∂ψ
Vr =
r ∂θ
( )
1 4
= r cos θ +
r 2π
( )
1 ◦ 2
= 0.8 cos 140 +
0.8 π
= 0.0297 m/s

ISTUDY
2.10. COMBINATION OF SIMPLE FLOWS 85

The resultant velocity at r = 0.8 and θ = 140◦ is



V = Vr2 + Vθ2

= 0.02972 + (−0.643)2 = 0.644 m/s

The velocity field is as shown in Figure 2.26.

FIGURE 2.26 Velocity field.

If θ is the angle that the velocity makes with the horizontal, as shown in the
figure, then
θ = 140◦ − α
Also,
Vθ 0.643
tan α = = = 21.65
Vr 0.0297
Therefore, α = 87.36◦ . Thus,

θ = 140 − 87.36 = 52.64◦

EXAMPLE 2.20 A two-dimensional flow field is made up of a source at


the origin and a flow given by ϕ = r2 cos 2θ. Locate any stagnation points in
the upper-half of the coordinate plane (0 ≤ θ ≤ π).

Solution The flow field’s potential function is


m
ϕ= ln r + r2 cos 2θ

m
where ln r is the velocity potential for the source, with strength m in m2 /s.

ISTUDY
86 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

The velocity components are given by


∂ϕ m
Vr = = + 2r cos 2θ
∂r 2πr

1 ∂ϕ
Vθ = = −2r sin θ
r ∂θ
At the stagnation point, Vr = 0 and Vθ = 0. Thus, we have
m
cos 2θ = − (i)
4πr2
and
sin 2θ = 0
or
2θ = 0 or π
or
π
θ = 0 or
2
Thus, θS = π/2 at the stagnation point. Substitution of this into Eq. (i) gives

( m )1/2
rS =

EXAMPLE 2.21 A certain body has the shape of Rankine half-body


with a thickness of 0.5 m. If this body is placed in an air stream of velocity
30 m/s, what source strength is required to simulate flow around the body?

Solution The half-body can be represented as a combination of a source


and a uniform flow, as shown in Figure 2.27.

FIGURE 2.27 Half-body.

ISTUDY
2.10. COMBINATION OF SIMPLE FLOWS 87

The resulting stream function is

ψ = ψuniform flow + ψsource


m
θ
= U r sin θ + (i)

where m is the source strength. At the stagnation point S,
1 ∂ψ
Vr = =0
r ∂θ
i.e.
m
U cos θ + =0
2πr
But at the stagnation point S, θ = π. Thus,
m
U=
2πr
If x = − b is the stagnation point, then at r = b
m
U=
2πb
or
m
b=
2πU
The value of the stream function at the stagnation point can be obtained by
evaluating ψ at r = b and θ = π, which yields
m
ψS = = πbU (ii)
2
Thus from Eq. (i), we get
m
πbU = U r sin θ + θ

θ
= U r sin θ + πbU
π
Solving, we get
b (π − θ)
r= (iii)
sin θ
The width of the half-body asymptotically approaches 2(πb). Equation
(iii) can be written as
y = b(π − θ)
so that as θ → 0 or θ → 2π, the half-width approaches ± bπ. Thus, for U =
30 m/s and y = 0.5/2 m, by Eqs. (ii) and (iii), we have
mU
π bU = U yπ = Uy
2πU

ISTUDY
88 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

This gives

m = 2U y
= 30 × 2 × (0.5/2)

= 15 m2 /s

EXAMPLE 2.22 Check whether the flow represented by the stream


function

ψ = V∞ r sin θ + θ

is irrotational, where q̇ is the volume flow rate, which is a constant.

Solution The radial and tangential components of the given flow are
1 ∂ψ
Vr =
r ∂θ

= V∞ cos θ +
2πr
∂ψ
Vθ = −
∂r

= − V∞ sin θ

The irrotationality condition given by Eq. (2.41) is


1 ∂ 1 ∂Vr
(r Vθ ) − =0
r ∂r r ∂θ
Thus,
( )
1 ∂ 1 ∂ q̇
(−r Vθ sin θ) − V∞ cos θ + =0
r ∂r r ∂θ 2πr
or
1 1
− Vθ sin θ + Vθ sin θ = 0
r r
The irrotational condition is satisfied and hence the flow is irrotational.

2.11 FLOW PAST A CIRCULAR CYLINDER


2.11.1 Flow without Circulation
A flow pattern equivalent to that of an irrotational flow over a circular cylinder
can be obtained by combining a uniform stream and a doublet with its axis
directed against the stream, as shown in Figure 2.28.

ISTUDY
2.11. FLOW PAST A CIRCULAR CYLINDER 89

FIGURE 2.28 Irrotational flow past a circular cylinder without


circulation.

The points S are the stagnation points. The combined stream function
becomes

ψ = ψdoublet + ψuniform flow


m
=− sin θ + V∞ r sin θ
( 2πr
m )
= V∞ r − sin θ
2πr
The potential function for the flow is
( m )
ϕ = V∞ r + cos θ (2.53)
2πr
It is seen that ψ = 0 for all values of θ, showing √ that the streamline
ψ
√ = 0 represents a circular cylinder of radius r = m/(2πV ∞ ). Let r = a =
m/(2πV∞ ). For a given velocity of the uniform flow and a given strength
of the doublet, the radius a is constant. Thus, the stream function and the
potential function of the flow past a cylinder can also be expressed as
( )
a2
ψ = V∞ r − sin θ (2.54)
r
( )
a2
ϕ = V∞ r + cos θ (2.55)
r
From the flow pattern shown in Figure 2.28, it is evident that the flow inside
the circle has no influence on the flow outside the circle. The normal and

ISTUDY
90 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

tangential components of velocity around the cylinder are


( )
∂ϕ a2
Vr = = V∞ 1 − 2 cos θ
∂r r

( )
1 ∂ϕ a2
Vθ = = −V∞ 1 + 2 sin θ
r ∂θ r
The flow speed around the cylinder is given by

|V |r=a = |Vθ |r=a = 2 V∞ sin θ

where what is meant by |Vθ | is the positive value of sin θ. This shows that
there are stagnation points on the surface at (a, 0) and (a, π). The flow velocity
reaches a maximum of 2 V∞ at the top and bottom of the cylinder.
The non-dimensional pressure distribution over the surface of the cylinder
is given by
p − p∞ V2
Cp = 1 = 1 − 2 = 1 − 4 sin2 θ (2.56)
2 ρ V∞
V∞

The pressure distribution at the surface of the cylinder is shown by the


continuous line in Figure 2.29. The symmetry of the general result of irro-
tational flow supports the theory that a steadily moving body experiences no
drag. This result is at variance with observations and is known as d’Alembert’s
paradox. The discrepancy is because of
• the existence of tangential stress or ‘skin friction’ and
• the drag due to the separation of the flow from sides and the resulting
formation of eddies for the blunt bodies.
The surface pressure in the wake is lower than that predicted by irrotational
or potential flow theory (Figure 2.29), resulting in a pressure drag.

2.11.2 Flow with Circulation


We observed that there is no net force acting on a circular cylinder in a
steady irrotational flow without circulation. It can be shown that a lateral
force identical to a lift force on an aerofoil, results when circulation is intro-
duced around the cylinder. When a clockwise line vortex of circulation Γ is
superposed around the cylinder in an irrotational flow, the stream function
becomes ( )
a2 Γ
ψ = V∞ r − sin θ + ln r
r 2π
The tangential velocity component at any point in the flow is
( )
∂ψ a2 Γ
Vθ = − = −V∞ 1 + 2 − (2.57)
∂r r 2πr

ISTUDY
2.11. FLOW PAST A CIRCULAR CYLINDER 91

FIGURE 2.29 Comparison of irrotational and observed pressure over a


circular cylinder. The observed distribution changes with the
Reynolds number Re, a typical behaviour at high Re, is
indicated by the dashed line.

At the surface of the cylinder, the tangential velocity becomes


Γ
Vθ |r=a = −2 V∞ sin θ −
2πa
At the stagnation point, Vθ = 0. Thus,
Γ
sin θ = − (2.58)
4 π a V∞
For Γ < 4πaV∞ , two values of θ satisfy this equation. This implies that there
are two stagnation points on the surface, as shown in Figure 2.30.
When Γ = 4 πaV∞ , the stagnation points merge on the negative y-axis, as
shown in Figure 2.30(c). For Γ > 4πaV∞ , the stagnation points merge and
stay outside the cylinder, as shown in Figure 2.30(d). The stagnation points
move away from the cylinder surface, since sin θ cannot be greater than 1.
The radial distance of the stagnation points in this case is found from
( )
a2 Γ
Vθ |θ=− π/2 = V∞ 1 + 2 − =0
r 2πr
This gives [ √ ]
1 2
r= Γ ± Γ2 − (4 πaV∞ )
4πV∞
The one root of which is r > a, the second root corresponds to a stagnation
point inside the cylinder.

ISTUDY
92 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

FIGURE 2.30 Flow past a circular cylinder without and with circulation.

To determine the magnitude of the transverse force acting on the cylinder,


it is essential to find the pressure distribution around the cylinder. Since the
flow is irrotational, Bernoulli’s equation can be applied between a point in the
freestream flow and a point on the surface of the cylinder. Pressure is found
from the Bernoulli’s equation as
ρV 2 ρ V∞2
p+ = p∞ +
2 2
Using Eq. (2.57), the surface pressure is found to be
[ ( )2 ]
1 Γ
pr=a = p∞ + ρ V∞ 2
− − 2V∞ sin θ − (2.59)
2 2π a

The symmetry of flow about the y-axis implies that the pressure force on the
cylinder has no component along the x-axis. The pressure force along the
y-axis, called the ‘lift’ force in aerodynamics, is (Figure 2.31)
∫ 2π
L=− pr=a a sin θ dθ
0

ISTUDY
2.11. FLOW PAST A CIRCULAR CYLINDER 93

FIGURE 2.31 Calculation of pressure force on a circular cylinder.

Substituting Eq. (2.59) in the above equation and integrating, we obtain the
lift as
L = ρ V∞ Γ (2.60)
where we have used
∫ 2π ∫ 2π
sin θ dθ = sin3 θ dθ = 0
0 0

It can be shown that Eq. (2.60) is valid for irrotational flows around any
two-dimensional shape, not for just circular cylinders. The result that the
lift force is proportional to circulation is of fundamental importance in aero-
dynamics. Wilhelm Kutta(1902) the German mathematician and Nikolai
Zhukhovsky(1906) the Russian aerodynamicist, have proved the relation of
Eq. (2.60) independently and it is called the Kutta–Zhukhovsky lift theorem
(the name Zhukhovsky is transliterated as Joukowsky in older Western texts).
The circulation developed by certain two-dimensional shapes, such as aerofoil,
when placed in a stream can be explained with the help of vortex theory. It
can be shown that fluid viscosity is responsible for the development of circula-
tion. The magnitude of circulation, however, is independent of viscosity, and
depends on the flow speed V∞ and the shape and ‘attitude’ of the body.
For a circular cylinder, the only way to develop circulation is by rotating
it in a flow stream. Although the viscous effects are important in this case,
the observed pattern for large values of cylinder rotation displays a striking
similarity to the ideal flow pattern for Γ > 4πaV∞ . For lower values of cylinder
rotation, the retarded flow in the boundary layer is not able to overcome
the adverse pressure gradient behind the cylinder. This leads to separation
and therefore the real flow pattern is rather unlike the irrotational pattern.
However, even in the presence of separation, the observed speeds are higher
on the upper surface of the cylinder, implying the existence of a lift force.
A second reason for a rotating cylinder generating lift is the asymmetry
caused due to delay of separation on the upper surface of the cylinder. The
asymmetry results in the generation of lift force. The contribution of this

ISTUDY
94 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

mechanism is small for two-dimensional objects such as circular cylinder,


but it is the only mechanism for side force experienced by spinning three-
dimensional objects such as soccer, tennis and golf balls. The lateral force
experienced by rotating bodies is called the Magnus effect. It is interesting
to note that the horizontal component of the force on the cylinder due to the
pressure which in general is called drag, and for the cylinder the drag given
by
∫ 2π
D= pr=a a cos θ dθ
0

is equal to zero. It is important to realize that this result is obtained on the


assumption that the flow is inviscid. In real flows the cylinder will experience
a finite drag force acting on it due to viscous friction and flow separation.

2.12 VISCOUS FLOWS


In the previous sections of this chapter, we have seen many interesting con-
cepts of fluid flow. With this background, let us study some of the impor-
tant aspects of fluid flow from practical or application point of view. We are
familiar with the fact that viscosity produces shear force which tends to retard
the fluid motion. It works against the inertia force. The ratio of these two
forces governs (dictates) many properties of the flow and the ratio expressed
in the form of a non-dimensional parameter known as the famous Reynolds
number, Re, is given by
ρV L
ReL =
µ
where V is the flow velocity, ρ is the flow density, µ is the dynamic viscos-
ity of the fluid, and L is a characteristic dimension. The Reynolds number
plays a dominant role in fluid flow analysis. This is one of the fundamental
dimensionless parameters which must be matched in similarity considerations
in most of the fluid flow analysis. At high Reynolds numbers, the inertia force
is predominant compared to viscous forces. At low Reynolds numbers the
viscous effects predominate everywhere. On the other hand, at high Re the
viscous effect is confined to a small region just adjacent to the surface of the
object present in the flow, and it is termed boundary layer. Since the length
and the velocity scales are chosen according to a particular flow, when com-
paring the flow properties at two different Reynolds numbers, only the flows
with geometric similarity should be considered. In other words, the flow over
a circular cylinder should be compared only with the flow of another circular
cylinder whose dimensions can be different, but not the shape. Flow in pipes
with different velocities and diameters, flow over aerofoils of the same kind,
and the boundary layer flows are also some geometrically similar flows. From
the above mentioned similarity consideration, we can infer that the geometric
similarity is a prerequisite for dynamic similarity. That is, the dynamically
similar flows must be geometrically similar, but the converse need not be true.

ISTUDY
2.12. VISCOUS FLOWS 95

Only similar flows can be compared, that is, when comparing the effect of vis-
cosity, the changes in flow pattern due to body shape should not interfere
with the problem.
For calculating the Reynolds number, different velocity and length scales
are used. Some of the length scales we often encounter in fluid flow studies
are given below.
Circular cylinder:
Red = ρ∞ V∞ d/µ∞ d is the cylinder diameter
Aerofoil:
Rec = ρ∞ V∞ c/µ∞ c is the aerofoil chord
Pipe flow (fully developed):
Red = ρV d/µ V is the average velocity
d is the pipe diameter
Channel flow (two-dimensional
and fully developed):
Reh = ρV h/µ V is the average velocity
h is the height of the channel
Flow over a grid:
ReM = ρV M/µ V is the velocity upstream or
downstream of the grid
M is the mesh size
Boundary layer:
Reδ = ρV δ/µ V is the outer velocity
δ is the boundary layer thickness

Reθ = ρV θ/µ θ is the momentum thickness


Rex = ρV x/µ x is the distance from the leading edge
In the above descriptions of the Reynolds numbers the quantities with the
subscript ∞ are at freestream and the quantities without the subscript are
the local properties. The Reynolds number is basically a similarity parame-
ter. It is used to determine the laminar and turbulent nature of flow. Below
certain values of the Reynolds number the flow cannot be turbulent. Under
such a condition, any disturbance introduced into the flow will be dissipated
by viscosity. It is known as the lower critical Reynolds number. Some of the
well-known critical Reynolds numbers are listed below.

Pipe flow : Red = 2300 : based on mean velocity and diameter

Channel flow : Reh = 1000 (two-dimensional): based on height


and mean velocity

Boundary layer flow : Reθ = 350: based on freestream velocity and


momentum thickness

ISTUDY
96 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

Circular cylinder : Rew = 200 (turbulent wake): based on wake


width and wake defect

Flat plate : Rex = 5 ×105 : based on length from the leading


edge

Circular cylinder : Red = 1.66 ×105 : based on cylinder diameter

The lower critical Reynolds number is that Reynolds number below which the
entire flow is laminar. The Reynolds number above which the entire flow is
turbulent is termed the upper critical Reynolds number. The critical Reynolds
number is that at which the flow field is a mixture of laminar and turbulent
flows. It is essential to realize that the transition is not taking place at a fixed
Reynolds number. Flow changes over from laminar to turbulent over a range
of Reynolds numbers—from lower critical to upper critical Reynolds number.
Note: It is important to note that when Re is low due to large µ, the flow is
termed stratified flow. Flow of tar, honey, etc. are stratified flows. When Re
is low because of low density, the flow is termed rarefied flow. For example,
the flows in space and at very high altitudes are rarefied flows.

EXAMPLE 2.23 A 25 mm diameter water pipe is 30 m long and delivers


water at 2 litres per minute at 30◦ C. Determine whether the flow in the pipe
is laminar or turbulent.

Solution Given that the volume flow rate

q̇ = 2 lit/min
q̇ = 2 × 0.001 m3 /min
= 0.002 m3 /min

The average velocity at the pipe exit is

V = q̇/A = 0.0679 m/s

For water at 30◦ C, ν = 0.802 × 10−6 m2 /s. Therefore, the Reynolds number
based on pipe diameter is

Vd
Red = = 2116.6
ν
The Reynolds number at the pipe exit is less than 2300 (the transition Reynolds
number); hence the flow is laminar.

EXAMPLE 2.24 A journal bearing supports a circular shaft of diameter


D = 10 cm turning at 3600 rpm. The bearing length L = 10 cm. The gap
between the shaft and the bearing, h = 0.1 mm, is filled with a lubricant

ISTUDY
2.12. VISCOUS FLOWS 97

whose viscosity is µ = 6.7 × 10−5 Pa·s. Calculate the torque T applied to the
shaft to overcome the friction in the bearing and the power P consumed in
the bearing by friction.

Solution The flow of lubricating oil in the bearing gap is very nearly a
plane Couette flow because the gap width h is very much less than the radius
of curvature D/2 of the shaft surface. The stress τω (acting) exerted on the
surface of the shaft is
VP µω D
τω = µ =
h h 2
where ω is the angular velocity of the shaft. The torque T is the product of
the shear stress τw times the radius D/2 times the surface area πDL of the
bearing, i.e.
D
T = τω (πDL)
2
πµωLD3
=
4h
π(6.7 × 10−5 ) (2π × 60) (0.1) (0.1)
3
=
4 × (1 × 10−4 )

= 1.984 × 10−2 N·m

The power P consumed is the product of the torque T times the angular
speed ω, i.e.
P = T ω = (1.984 × 10−2 ) (2 π × 60)

= 7.479 W

2.12.1 Drag of Bodies


When a body moves in a fluid it experiences forces and moments due to the
relative motion of the fluid flow which is taking place around it. If the body
has arbitrary shape and orientation, the flow will exert forces and moments
about all the three coordinate axes, as shown in Figure 2.32. The force on
the body along the flow direction is called drag.
The drag is essentially a force opposing the motion of the body. Viscosity
is responsible for a part of the drag force, and the body shape generally
determines the overall drag. In the design of transport vehicles, the shapes
experiencing minimum drag are considered to keep the power consumption
to a minimum. Low drag shapes are called streamlined bodies and high drag
shapes are termed bluff bodies.
Drag arises due to (a) the difference in pressure between the front and
back regions and (b) the friction between the body surface and the fluid. The

ISTUDY
98 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

FIGURE 2.32 Forces acting on an arbitrary body.

drag force (a) is known as pressure drag and the drag force (b) is known as
skin friction drag or shear drag.

Pressure drag
The pressure drag arises due to the separation of boundary layer, whenever
it encounters an adverse pressure gradient. The phenomenon of separation
and how it causes the pressure drag can be explained better by considering
flow around a body like a circular cylinder. If there is no viscosity and hence
no boundary layer, the flow would have gone around the cylinder like a true
potential flow, as shown in Figure 2.33. For this flow the pressure distribu-
tion will be the same on the front and back sides. The net force along the
freestream direction is then zero. That is, there is no drag acting on the
cylinder.

FIGURE 2.33 Potential flow past a circular cylinder.

But in real flow, because of viscosity, a boundary layer is formed on the


surface of the cylinder. The flow experiences a favourable pressure gradient
from forward stagnation point S1 to the topmost point A on the cylinder at
θ = 90◦ , as shown in Figure 2.33. Therefore, the flow accelerates from θ = 0◦
to 90◦ . However, beyond θ = 90◦ the flow is subjected to an adverse pressure
gradient and hence decelerates. Under this condition the pressure is acting
against the fluid flow. In a boundary layer the velocity near the surface is
small and hence the force due to its momentum is unable to counteract the
pressure force. The boundary layer flow gets retarded and the velocity near

ISTUDY
2.12. VISCOUS FLOWS 99

the wall region reduces to zero and then the flow is pushed back in the opposite
direction, as sketched in Figure 2.34. This phenomenon is called separation.

FIGURE 2.34 Separation process.

Across the separated region the pressure is nearly constant and lower than
what it would have been if the flow did not separate. The pressure had not
completely recovered as in the case of potential flow. Thus, on account of the
incomplete recovery of pressure due to separation, a net drag force opposing
the body motion is generated. We can easily see that the pressure drag will
be less if the separation had taken place latter, that is, the area over which
the pressure unrecovered is less. To minimize pressure drag the separation
point should be as far as possible from the leading edge. This is true for any
shape. Streamlined bodies are designed on this basis and the adverse pressure
gradient is kept as small as possible by keeping the curvature very small.
The separation of boundary layer depends not only on the strength of the
adverse pressure gradient but also on the nature of the boundary layer, namely
laminar or turbulent. Laminar flow has the tendency to separate earlier than
a turbulent flow. This is because the laminar velocity profiles in a boundary
layer has less momentum near the wall. This is conspicuous in the case of flow
over a circular cylinder. Laminar boundary layer separates nearly at θ = 90◦ ,
whereas for a highly turbulent boundary layer the separation is delayed and
the attached flow continues up to as much as θ = 150◦ on the cylinder. The
reduction of pressure drag when the boundary layer changes from laminar
to turbulent is of the order of 5 times for bluff bodies. The flow behind a
separated region is called the wake. Thus, the wake may be defined as the
separated region behind a body where the pressure loss is severe. For low
drag, the wake width should be small.

ISTUDY
100 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

Although separation is shown to take place at well-defined locations on the


body, in the illustration in Figure 2.34, it actually takes place over a zone on
the surface which cannot be identified easily. Therefore, the theoretical esti-
mation of separation especially for a turbulent boundary layer is difficult and
hence pressure drag cannot be easily calculated. Some approximate methods
exist but they can serve only as guidelines for the estimation of pressure drag.
It is important to note that, for all irrotational flows, a velocity potential
(or potential) function ϕ exists whether the flow is compressible or incom-
pressible.
On the other hand, the stream function exists only when compressibility
is negligible, irrespectively of whether the motion is irrotational or rotational.
This is because the existance of stream function is merely a consequence of
the assertion of the continuity of incompressible flow.

Skin friction drag


The friction between the surface of the body and the fluid causes viscous shear
stress and this force is known as skin friction drag. Wall shear stress τ at the
surface of the body is given by Newtons’ law of viscosity (friction) as

∂Vx
τ =µ (2.61)
∂y

where µ is the dynamic viscosity coefficient and ∂Vx /∂y is the velocity gradient
at the body surface y = 0. If the boundary layer velocity profile is known,
the shear stress can be calculated.
For streamlined bodies, the separated zone being small, the major portion
of the drag is from skin friction. Turbulent boundary layer results in more
skin friction than a laminar one. Examine the variation of friction coefficient
Cf versus Reynolds number, for a flat plate kept at zero angle of attack in a
uniform stream, plotted in Figure 2.35.

FIGURE 2.35 Cf variation with Reynolds number.

ISTUDY
2.12. VISCOUS FLOWS 101

The characteristic length for Reynolds number is the plate length from
its leading edge. It can be seen from Figure 2.35 that the Cf is more for a
turbulent flow than that for a laminar flow. The friction coefficient is defined
as
total frictional force
Cf = 1 2
(2.62)
2 ρV∞ S

where V∞ and ρ are the freestream velocity and density, respectively, and S
is the wetted surface area of the flat plate.
For bluff bodies the pressure drag is many times larger than the skin
friction drag, and for streamlined bodies the condition is the reverse. In the
case of streamlined bodies like aerofoil the designer aims at keeping the skin
friction drag as low as possible. Maintaining the laminar boundary layer
conditions all along the surface is the most suitable arrangement. Though
such aerofoils have been designed, they have many limitations. Even a small
surface roughness and disturbance make the flow turbulent which defeats the
purpose. In addition, there is a tendency for the flow to separate even at
small angles of attack which severely restricts the use of such aerofoils. Thus,
it can justifiably be stated that the streamlined bodies are those for which
the skin friction drag constitutes the major portion of the total drag, and the
bluff bodies are those which have pressure drag as the major constituent of
the total drag.

Comparison of drag of various bodies


In low-speed flow past geometrically similar bodies with identical orienta-
tion and relative roughness, the drag coefficient should be a function of the
Reynolds number only, i.e.
CD = f (Re)
The Reynolds number is based upon the freestream velocity V and a charac-
teristic length L of the body. The drag coefficient CD could be based upon
L2 , but it is customary to use a characteristic area of the body instead of L2 .
Thus, the drag coefficient becomes

Drag
CD = 1 2
(2.63)
2 ρV S

The factor 1/2 is our traditional tribute to Euler and Bernoulli. The area S
is usually one of the following three types.

1. Frontal area of the body as seen from the stream. This is suitable for
thick stubby bodies, such as spheres, cylinders, cars, missiles, projectiles,
and torpedos.
2. Planform area of the body as seen from above. This is suitable for wide
flat bodies such as wings and hydrofoils.
3. Wetted area. This is appropriate for surface ships and barges.

ISTUDY
102 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

While using drag or other fluid–force data, it is important to note what


length and area are being used to scale the measured coefficients.
The drag coefficients of various bodies are given in Table 2.1. The influ-
ence of geometry on drag can be clearly seen from this table.

TABLE 2.1 Drag coefficients of various bodies at Re ≥ 104


[Fluid Dynamic Drag, Hoerner, 1965]

The sharp-edged bodies, which tend to cause flow separation regardless


of the nature of boundary layer, are insensitive to Reynolds number. Bodies
such as elliptic cylinders, being smoothly rounded, have the laminar turbulent
transition effect and are therefore quite sensitive to the nature of the boundary
layer.
For three-dimensional bodies, the sharp edges always cause flow separation
and high drag which is insensitive to Reynolds number.
Rounded bodies like the ellipsoid have drag which depends upon the point
of separation, so that both Reynolds number and the nature of boundary layer
are important. Increase in body length will generally decrease the pressure
drag by making the body relatively more slender, but sooner or later the skin
friction drag will catch up.

2.12.2 Turbulence
Turbulent flow is described as a flow with irregular fluctuations. In nature,
most of the flows are turbulent. Turbulent flows have properties which are

ISTUDY
2.12. VISCOUS FLOWS 103

appreciably different from those of laminar flows. Likewise, we have to ex-


plain all the properties of turbulent flow to completely describe it. Incor-
porating all the important properties the turbulence may be described as a
three-dimensional, random phenomenon, exhibiting multiplicity of scales, pos-
sessing vorticity and showing very high dissipation. Turbulence is described
as a three-dimensional phenomenon. It means that even in a one-dimensional
flow field the turbulent fluctuations are always three-dimensional. In other
words, the mean flow may be one- or two- or three-dimensional, but the tur-
bulence is always three-dimensional. From the above discussions, it is evident
that turbulence can only be described and cannot be defined.
A complete theoretical approach to turbulent flow paralleling that of lam-
inar flow is impossible because of the complexity and apparently random
nature of the velocity fluctuations in turbulent flow. Nevertheless, semi-
theoretical analysis aided by limited experimental data can be carried out
for turbulent flows, with instruments which have the capacity to detect high
frequency fluctuations. For flows at very low-speed, say around 20 m/s, the
frequencies encountered will be from 2 to 500 Hz. Hot-wire anemometer is
well suited for measurements in such flows. A typical hot-wire velocity trace
of a turbulent flow is shown in Figure 2.36.

FIGURE 2.36 Hot-wire trace of a turbulent flow.

Turbulent fluctuations are random in amplitude, phase and frequency.

ISTUDY
104 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

If an instrument, such as a pitot-static tube, which has a low frequency


response of the order of 30 seconds, is used for the measurement of velocity,
the manometer will read only a steady value ignoring the fluctuations. This
means that the turbulent flow consists of a steady-velocity component which is
independent of time, over which the fluctuations are superimposed, as shown
in Figure 2.36(b). That is,

U (t) = U + u′ (t) (2.64)

where U (t) is the instantaneous velocity, U is the time averaged velocity


and u′ (t) is the turbulent fluctuation around the mean velocity. Since U
is independent of time, the time average of u′ (t) should be equal to zero.
That is, ∫
1 t ′
u (t) = 0 ; u′ = 0
t 0
provided that the time t is sufficiently large. In most of the laboratory flows,
averaging over a few seconds is sufficient if the main flow is kept steady.
In the beginning of this section, we saw that the turbulence is always
three-dimensional in nature even if the main flow is one-dimensional. For ex-
ample, in a fully developed pipe or channel flow as far as the mean velocity
is concerned only the x-component of velocity U alone exists, whereas all the
three components of turbulent fluctuation u′ , v ′ and w′ are always present.
The intensity of the turbulent velocity fluctuations is expressed in the form
of its root mean square value. That is, the velocity fluctuations are instan-
taneously squared, then averaged over certain period and finally the square
root is taken. The root mean square (rms) value is useful in estimating the
kinetic energy of fluctuations. The turbulence level for any given flow with a
mean velocity U is expressed as the turbulence number n, which is defined as

u′2 + v ′2 + w′2
n = 100 (2.65)
3U
In the laboratory, turbulence can be generated in many ways. A wire-mesh
placed across an air stream produces turbulence. This turbulence is known
as grid turbulence. If the incoming air stream as well as the mesh sizes
are uniform, then the turbulent fluctuations behind the grid are isotropic in
nature, i.e. u′ , v ′ and w′ are equal in magnitude. In addition to this, the
mean velocity is the same across any cross-section perpendicular to the flow
direction. That is, no shear stress exists. As the flow moves downstream
the fluctuations die-down due to viscous effects. Turbulence is produced in
jets and wakes as well. The mean velocity in these flows varies and they are
known as free shear flows. Fluctuations exist up to some distance and then
slowly decay. Another type of turbulent flow often encountered in practice
is the turbulent boundary layer. It is a shear flow with zero velocity at the
wall. These flows maintain the turbulence level even at large distances, unlike
grid or free shear flows. In wall shear flows or boundary layer type flows,
turbulence is produced periodically to counteract the decay.

ISTUDY
2.12. VISCOUS FLOWS 105

A turbulent flow may be visualized as a flow made up of eddies of various


sizes. Large eddies are first formed taking energy from the mean flow. They
then breakup into smaller ones in a sequential manner till they become very
small. At this stage the kinetic energy gets dissipated into heat due to vis-
cosity. Mathematically it is difficult to define an eddy in a precise manner. It
represents, in a way, the frequencies involved in the fluctuations. Large eddy
means low frequency fluctuations and the smallest eddy means the highest fre-
quency fluctuations encountered in the flow. The kinetic energy distribution
at various frequencies can be represented by an energy spectrum as shown in
Figure 2.36(c).
The problem of turbulence is yet to be solved completely. Different kinds
of approaches are employed to solve these problems. The well-known method
is to write the Navier–Stokes equations for the fluctuating quantities and then
average them over a period of time. In the Navier–Stokes equations, substi-
tute the following.
Vx = V x + u′ , p = p + p′ (2.66a)
Vy = V y + v ′ , V z = V z + w′ (2.66b)

where Vx , u′ , Vy , v ′ , Vz and w are the mean and fluctuational components of
velocity along x-, y- and z-directions, respectively. The bar denotes the mean
value, that is, the time averaged quantity.
Let us now consider the x-momentum equation for a two-dimensional flow.
( 2 )
∂Vx ∂Vx 1 ∂p ∂ Vx ∂ 2 Vx
Vx + Vy =− +ν + (2.67)
∂x ∂y ρ ∂x ∂x2 ∂y 2
In Eq. (2.67), ν is the kinematic viscosity, given by
ν = µ/ρ
Substituting Eq. (2.66) into Eq. (2.67), we get

∂(Vx + u′ ) ∂(Vx + u′ ) 1 ∂(p + p′ ) ∂ 2 (Vx + u′ )


(Vx + u′ ) + (Vy + v ′ ) =− +ν
∂x ∂y ρ ∂x ∂x2

∂ 2 (Vx + u′ )

∂y 2
Expanding this, we obtain
∂Vx ∂u′ ∂Vx ∂u′ ∂Vx ∂u′ ∂Vx ∂u′
Vx + Vx + u′ + u′ + Vy + Vy + v′ + v′
∂x ∂x ∂x ∂x ∂y ∂y ∂y ∂y
1 ∂p 1 ∂p′ ∂ 2 Vx ∂ 2 u′ ∂ 2 Vx ∂ 2 u′
=− − +ν + ν + ν + ν (2.68)
ρ ∂x ρ ∂x ∂x2 ∂x2 ∂y 2 ∂y 2
In this equation, time averaging of the individual fluctuations is zero. Their
products or square quantities are not zero. Taking the time average of
Eq. (2.68), we get
∂V x ∂u′ ∂V x ∂u′ 1 ∂p ∂2Vx ∂2Vx
Vx + u′ +Vy + v′ =− +ν 2
+ν (2.69)
∂x ∂x ∂y ∂y ρ ∂x ∂x ∂y 2
Equation (2.69) is slightly different from the laminar Navier–Stokes equation.

ISTUDY
106 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

The continuity equation for the two-dimensional flow under consideration


is
∂(V x + u′ ) ∂(V y + v ′ )
+ =0
∂x ∂y
This can be expanded to result in

∂V x ∂V y ∂u′ ∂v ′
+ =0 and + =0 (2.70)
∂x ∂y ∂x ∂y

The terms involving turbulent fluctuational velocities u′ and v ′ on the left-


hand side of Eq. (2.69) can be written as

∂u′ ∂u′ ∂ ( ′2 ) ∂u′ ∂u′


u′ + v′ = u − u′ + v′
∂x ∂y ∂x ∂x ∂y

Using Eq. (2.70) the above equation can be expressed as

∂u′ ∂u′ ∂ ( ′2 ) ∂ ( ′ ′)
u′ + v′ = u + uv (2.71)
∂x ∂y ∂x ∂y

Combination of Eqs. (2.69) and (2.71) results in


( ) ( )
∂V x ∂V x ∂p ∂ ∂V x ∂∂V x
ρV x + ρV y =− + µ − ρu′2 + µ − ρu′ v ′
∂x ∂y ∂x ∂x ∂x ∂y∂y
(2.72)
The terms −ρu′2 and −ρu′ v ′ in Eq. (2.72) are due to turbulence. They are
popularly known as Reynolds or turbulent stresses. For a three-dimensional
flow the turbulent stress terms are ρu′2 , ρv ′2 , ρw′2 , ρu′ v ′ , ρv ′ w′ and ρw′ u′ . The
solutions of Eq. (2.72) are rather cumbersome. Assumptions like eddy viscos-
ity, mixing length are made to find a solution to this equation. The Reynolds
stresses are additional stresses acting in a mean turbulent flow. In fact, these
additional stresses on the mean
( field of a turbulent
) flow are much larger than
the viscous contribution µ ∂Vx /∂x + ∂Vx /∂y , except when very close to a
solid surface where the fluctuations are small and the mean flow gradients are
large.
The tensor −ρui uj is called the Reynolds stress tensor and has the nine
Cartesian components

−ρu2 −ρuv −ρuw


−ρuv −ρv 2 −ρvw
−ρuw −ρvw −ρw2

This is a symmetric tensor with the normal stresses as the diagonal compo-
nents and the shear stresses as the off-diagonal components. For isotropic
turbulence, i.e. when the turbulent fluctuations do not have any directional
preference, the off-diagonal components vanish and u2 = v 2 = w2 .

ISTUDY
2.12. VISCOUS FLOWS 107

At this stage, it is important to have proper clarity about the laminar and
turbulent flows. The laminar flow may be described as an orderly pattern
where the fluid layers are assumed to slide over one another, i.e. in laminar
flow the fluid moves in layers, or laminas, one layer sliding over an adjacent
layer with only a molecular interchange of momentum. Any tendency towards
instability and turbulence is damped out by the viscous shear forces that
resist relative motion of adjacent fluid layers. In other words, laminar flow
is an orderly flow in which the fluid elements move in an orderly manner
such that the transverse exchange of momentum is insignificant. On the
other hand, the turbulent flow is a three-dimensional random phenomenon,
exhibiting multiplicity of scales, possessing vorticity, and showing very high
dissipation. Turbulent flow is basically an irregular flow. Turbulent flow has
an erratic motion of fluid particles, with a violent transverse exchange of
momentum.
A laminar flow, though has irregular molecular motion, is macroscopically
a well-ordered flow. But in the case of turbulent flow there is the effect of
a small but macroscopic fluctuating velocity superimposed on a well-ordered
mean flow. A graph of velocity versus time at a given position in pipe flow
would appear as shown in Figure 2.37(a) for steady laminar flow, and as shown
in Figure 2.37(b) for turbulent flow. In Figure 2.37(b) for turbulent flow, an
average velocity denoted by V has been indicated. Because this average is
constant with time, the flow has been designated as steady turbulent flow. An
unsteady turbulent flow may prevail when the average velocity field changes
with time as shown in Figure 2.37(c).

FIGURE 2.37 Variation of flow velocity with time.

ISTUDY
108 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.12.3 Flow Through Pipes

Fluid flow through pipes with circular and non-circular cross-sections is one of
the commonly encountered problems in many practical systems. Flow through
pipes is driven mostly by pressure or gravity or both.
Consider the flow in a long duct as shown in Figure 2.38. This flow is
constrained by the boundary walls. At the entrance, the flow (assumed to be
inviscid) converges and enters the tube.

FIGURE 2.38 Flow development in a long duct.

Because of the viscous friction between the fluid and the pipe wall, the
viscous boundary layer grows downstream of the entrance. The boundary
layer growth makes the effective area of the pipe to decrease progressively
downstream, thereby making the flow along the pipe to accelerate. This
process continues up to the point where the boundary layer from the wall
grows and meets at the pipe centreline, i.e. fills the pipe.
The zone upstream of the boundary layer merging point is called the en-
trance or flow development length, and the zone downstream of the merging
point is termed the fully developed region. In the fully developed region,
the velocity profile remains unchanged. Dimensional analysis shows that the
Reynolds number is the only parameter influencing the entrance length. In

ISTUDY
2.12. VISCOUS FLOWS 109

the functional form, the entrance length can be expressed as

Le = f (ρ, V, d, µ)
( )
Le ρV d
= f1 = f1 (Re)
d µ

where ρ, V and µ are the flow density, velocity and viscosity, respectively, and
d is the pipe diameter.
For laminar flow, the accepted correlation is

Le
≈ 0.06 (Re)d (2.73a)
d

At the critical Reynolds number (Re)c = 2300, for pipe flow, Le = 138 d,
which is the maximum development length possible.
For turbulent flow the boundary layer grows faster, and Le is given by the
approximate relation
Le 1/6
≈ 4.4 [(Re)d ] (2.73b)
d

Now examine the flow through an inclined pipe, shown in Figure 2.39, con-
sidering the control volume between sections 1 and 2.

FIGURE 2.39 Fully developed flow in an inclined pipe.

ISTUDY
110 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

Treating the flow to be incompressible, by volume conservation, we have

Q̇1 = Q̇2 = constant

Q̇1 Q̇2
V1 = = V2 =
A1 A2

where Q̇1 and Q̇2 are the volume flow rates and A1 , A2 , V1 and V2 are the
local area and velocity at states 1 and 2. The velocities V1 and V2 are equal
since the flow is fully developed and also A1 = A2 .
By incompressible Bernoulli equation, we have
p1 1 p2 1
+ V12 + gz1 = + V22 + gz2 + hf (2.74)
ρ 2 ρ 2

Since V1 = V2 , the head loss due to friction can be expressed as


( ) ( )
p1 p2 ∆p
hf = z1 + − z2 + = ∆z + (2.75)
ρg ρg ρg

that is, the head loss in the pipe, due to friction, is equal to the sum of the
change in gravity head and pressure head.
By momentum balance, we have

∆p πR2 + ρg(πR2 )∆L sin θ − τw (2πR)∆L = ṁ(V1 − V2 ) = 0 (2.76)

Dividing throughout by (πR2 )ρg, we get

∆p 2τw ∆L
+ ∆L sinθ =
ρg ρg R

But ∆L sin θ = ∆z. Thus, ∆p 2τw ∆L


+ ∆z =
ρg ρg R
Using Eq. (2.75), we obtain

∆p 2τw ∆L
+ ∆z = hf = (2.77)
ρg ρg R

In the functional form, the wall shear τw may be expressed as

τw = F (ρ, V, µ, d, ϵ) (2.78)

where µ is viscosity of the fluid, d is the pipe diameter, and ϵ is the wall
roughness height. By dimensional analysis, Eq. (2.78) may be expressed as

8τw ( ϵ)
= f = F Red , (2.79)
ρV 2 d

where f is called the Darcy friction factor, which is a dimensionless parameter.

ISTUDY
2.12. VISCOUS FLOWS 111

Combining Eqs. (2.77) and (2.79), we obtain the pipe head loss as

LV2
hf = f (2.80)
d 2g

This is called the Darcy–Weisbach equation, valid for flow through ducts of
any cross-section. Further, in the derivation of the above relation, there was no
mention about whether the flow was laminar or turbulent and hence Eq. (2.80)
is valid for both laminar and turbulent flow. The value of friction factor f for
any given pipe (i.e. for any surface roughness ϵ and d) at a given Reynolds
number can be read from the Moody chart (which is a plot of f as a function
of (Re)d and ϵ/d).

EXAMPLE 2.25 Gasoline at 20◦ C is pumped at 0.2 m3 /s through a


15 km long cast-iron pipe of diameter 0.18 m. Compute the power required
in kilowatts if the pumps are 80 per cent efficient. [for cast iron ϵ = 0.26 mm;
for gasoline ρ = 680 kg/m3 , µ = 2.92 × 10−4 kg/(m s)].

Solution Flow velocity through the pipe is given by


V =
A
0.2
= 2
(π/4)(0.18)
= 7.86 m/s

The Reynolds number of the flow is

680 (7.86)(0.18)
Re =
2.92 × 10−4
= 3.29 × 106

From Moody chart, for Re = 3.29 ×106 and ϵ/d = 0.00026/0.18 = 0.00144,
we have the friction factor f = 0.0216.
The head loss through the pipe is given by

L V2
hf = f
D 2g
( )( )
15000 7.862
= 0.0216
0.18 2 × 9.81
= 5668 m

ISTUDY
112 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

The power required becomes

ρ g Q̇ hf
Power =
ηpump

(680)(9.81)(0.2)(5668)
=
0.8

= 9453 kW

EXAMPLE 2.26 Crude oil flows through a level section of a refinery


pipeline, shown in Figure 2.40, at a rate of 250,000 m3 /day. The inside diam-
eter of the pipe is 1200 mm, its roughness is equal to that of the galvanized
iron. The maximum allowable pressure is 8.3 MPa; the maximum pressure
required to keep the dissolved gas in solution in the crude oil is 340 kPa. The
2
specific gravity of crude oil is 0.93 and µ = 0.017 N · s/m at the pumping

temperature 60 C. For these conditions, determine the maximum possible
spacing between the pumping stations. If the pump efficiency is 85 per cent,
determine the power that must be supplied at each pumping station. Take
ϵ/D = 0.00012, where ϵ is the average surface roughness of the pipe.

FIGURE 2.40 Oil flow through a pipe.

Solution By Bernoulli equation, we have

p2 V2 p1 V2
+ 2 + gz2 = + 1 + gz1 + hf
ρ 2 ρ 2

Assume that, V12 = V22 , µ = constant, z1 = z2 (since the pipe is horizontal)


Also,
2
L V
hf = f
d 2

where V is the average velocity in the pipe and f is the friction factor.
Thus,
2
L V
△p = p2 − p1 = f ρ
d 2

ISTUDY
2.12. VISCOUS FLOWS 113

or

2d △p
L=
f ρV 2

Q̇ 250000 4
V = = ×
A 24 × 3600 π(1.2)2

= 2.56 m/s

ρV d (0.93 × 103 )(2.56)(1.2)


(Re)d = =
µ 1.7 × 10−2

= 1.68 × 105

From the Moody chart, for the above (Re)d , and ϵ/d = 0.00012, we have
f = 0.017. Therefore,

( )
2 1 1
L= (1.2)(8.3 × 10 − 340 × 10 )
6 3
0.017 0.93 × 1000 2.562

= 185 × 103 m

= 185 km

Now, applying the first law of thermodynamics to CV2, across the pumps
between the sections 3 and 2, we get (with e as the specific energy and u as
the specific internal energy)

∫ ∫ ( )
∂ V2 p
Q̇ − Ẇ = eρ dV + u+ + gz + ρV dA
∂t CV CS 2 ρ
( ) ( )
p2 p3
Ẇin = − W = u2 + ṁ + u1 + (−ṁ) − Q̇
ρ ρ
( )
p2 − p3
Ẇin = ṁ + ṁ (u2 − u1 ) − Q̇
ρ
( )
p2 − p3
= ṁ + losses
ρ

Also,

Losses = (1 − η) Ẇin

ISTUDY
114 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

Thus,
1 p2 − p3
Ẇin = ṁ
η ρ

1 V A ∆p
= (p2 − p3 ) V A =
η η
[ ]
1 π 2 ( ) 1
= (2.56) (1.2) 7.96 × 106
0.85 4 1000

= 27100 kW

EXAMPLE 2.27 A steady push on the piston shown in Figure 2.41 causes
a flow rate Q̇ = 0.4 cm3 /s through the needle. The fluid has ρ = 900 kg/m3
and µ = 0.002 kg/m·s. Determine the force F required to maintain the flow.

FIGURE 2.41 Flow caused by piston.

Solution Neglecting the head loss in the cylinder, we have the flow velocity
through the needle as


Vneedle =
A
0.4
= 2 = 8.15 m/s
(π/4)(0.025)

By energy equation, we have

p1 V2 p2 V2
+ 1 + z1 = + 2 + z2 + hf
ρg 2g ρg 2g
where the subscripts 1 and 2 refer to cylinder and needle, respectively.
ρV d2
Re =
µ
900(8.15)(0.00025)
= = 917
0.002

ISTUDY
2.12. VISCOUS FLOWS 115

Thus the flow is laminar. The head loss is given by

32 µ L V
hf =
ρ g d22

32 (0.002)(0.03)(8.15)
= 2 = 28.36 m
900 (9.81)(0.00025)

p1 − p2 V2
= hf + 2
ρg 2g

8.152
= 28.36 + = 31.75 m
2 × 9.81
∴ p1 − p2 = (31.75)(900)(9.81) = 280320.75 Pa

Thus, the force required becomes

F = (p1 − p2 )Ap
(π )
= 280320.75 × 0.012 = 22 N
4
EXAMPLE 2.28 A fluid flows through two horizontal pipes of equal length
which are connected together to form a passage of length 2L. The flow is
laminar and fully developed. The pressure drop in the first pipe is 1.44 times
greater than the pressure drop in the second pipe. If the diameter of the
first pipe is D, determine the diameter of the second pipe. Assume the same
friction factor f for both the pipes.

Solution The pressure drop ∆p can be expressed as

V2 L
∆p = f (i)
2g D

where V is the flow velocity, L and D are the length and diameter of the pipe,
g is the gravitational acceleration and f is the friction factor.
Given that,
∆p1 = 2.44 ∆p2 (ii)
where the subscripts 1 and 2 refer to pipes 1 and 2, respectively.
From Eqs. (i) and (ii), we have

V12 V2
= 2.44 2 (iii)
D1 D2

By volume flow conservation, we have

A1 V1 = A2 V2

ISTUDY
116 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

This gives
( )2
V1 D2
=
V2 D1
Therefore, Eq. (iii) becomes
( )5
D2
= 2.44
D1

Thus,
D2 = 1.195 D
since D1 = D.

2.13 GAS DYNAMICS


In the preceding sections of this chapter, the discussions concerned the science
of fluid mechanics, defined as the fluid flow in which the temperature changes
associated with the flow are insignificant. Thus, fluid mechanics is essentially
a science of isenthalpic flows and thus, the main equations governing a fluid
dynamic stream are only the continuity and momentum equations and the sec-
ond law of thermodynamics. The energy equation is passive as far as the fluid
dynamic streams are concerned. At standard sea level conditions, considering
less than five per cent change in temperature as insignificant, it can be seen
that a flow with Mach number less than 0.5 can be termed fluid mechanic
stream. For analyzing a fluid mechanic flow, the continuity and momentum
equations are sufficient. The energy equation need not be considered. Thus,
a fluid mechanics stream may be compressible or incompressible. For an in-
compressible flow both temperature and density changes are insignificant. For
a compressible fluid dynamic stream, the temperature change is insignificant
but the density change is finite.
But in many engineering applications, such as design of aeroplanes, missiles
and launch vehicles, the associated flow Mach number is more than 0.5 and
hence both the temperature and density changes associated with the flow
become significant. The study of such flows where the changes in both density
and temperature associated with pressure change become appreciable is called
gas dynamics. In other words, gas dynamics is the science of fluid flow where
both density and temperature changes become significant. The essence of gas
dynamics is that the entire flow field is dominated by Mach waves, expansion
waves and shock waves, when the flow speed is supersonic. It is through
these waves that the change in flow properties from one state to another takes
place. In the theory of gas dynamics, the change of state in flow properties
is achieved by three means: (a) with area change, treating the fluid to be
inviscid and the passage to be frictionless, (b) with friction, treating the heat
transfer between the surrounding and system to be negligible, and (c) with
heat transfer, assuming the fluid to be inviscid. These three types of flows are

ISTUDY
2.13. GAS DYNAMICS 117

called isentropic flow, frictional or Fanno type flow, and Rayleigh type flow,
respectively.
All problems in gas dynamics can be classified under the three flow
processes described above, of course with the assumptions mentioned. Al-
though it is impossible to have a flow process which is purely isentropic or
Fanno type or Rayleigh type, in practice, it is justified in assuming so, since
the results obtained with these treatments prove to be accurate enough for
most of practical problems in gas dynamics. Even though it is possible to
solve problems with mathematical equations and working formulae associated
with these processes, it is found to be extremely useful and time saving if the
working formulae are available in the form of tables with Mach number which
is the dominant parameter in compressible-flow analysis.

2.13.1 Perfect Gas


In principle, it is possible to do gas dynamic calculations with the general
equation of state relations, for fluids. But in practice most elementary treat-
ments are confined to perfect gases with constant specific heats. For most
problems in gas dynamics, the assumption of the perfect gas law is sufficiently
in accord with the properties of the actual gases, hence it is acceptable.
For perfect gases, the pressure–density–temperature relation or the state
equation, is given by
p = ρRT (2.81)
where R is the gas constant and T is absolute temperature. Further, for
perfect gases the specific heats are constants and independent of temperature.
For perfect gases,
R = cp − cv

cp
γ= (2.82)
cv
where cp and cv are the specific heats at constant pressure and constant
volume, respectively, and γ is the isentropic index. For all real gases cp , cv
and γ vary with temperature but only moderately. For example the cp of air
increases about 30 per cent as the temperature increases from 0 to 3000◦ C.
Since we rarely deal with such large temperature changes, it is reasonable to
assume specific heats to be constants in our studies.

2.13.2 Speed of Sound


In the beginning of this section, it was stated that gas dynamics deals with
flows in which both compressibility and temperature changes are important.
The term compressibility implies variation in density. In many cases, the
variation in density is mainly due to pressure change. The rate of change
of density with respect to pressure is closely connected with the velocity of
propagation of small pressure disturbances, i.e. with the speed of sound ‘a’.

ISTUDY
118 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

The speed of sound may be expressed as


( )
2 ∂p
a = (2.83)
∂ρ s

In Eq. (2.83), the ratio dp/dρ is written as partial derivative at constant en-
tropy because the variations in pressure and temperature are negligibly small,
and consequently, the process is nearly reversible. Moreover, the rapidity with
which the process takes place, together with the negligibly small magnitude
of the total temperature variation, makes the process nearly adiabatic. In
the limit, for waves with infinitesimally small thickness, the process may be
considered both reversible and adiabatic, and therefore, isentropic.
It can be shown that for an isentropic process of a perfect gas the velocity
of sound can be expressed as

a= γRT (2.84)

where T is absolute static temperature.

Mach number
Mach number M is a dimensionless velocity, expressed as the ratio between
the magnitudes of local flow velocity and local speed of sound, i.e.

local flow velocity V


M≡ = (2.85)
local speed of sound a

Mach number plays a dominant role in the field of gas dynamics.

2.13.3 Flow with Area Change


If the flow is assumed to be isentropic for a channel flow, all states along the
channel or streamtube lie on a line of constant entropy and have the same
stagnation temperature. The state of zero velocity is called the isentropic
stagnation state, and the state with M = 1 is called the sonic or critical state.

Isentropic relations
The relations between pressure, temperature, and density for an isentropic
process of a perfect gas are
( )γ ( )(γ − 1)/γ
p ρ T p
= and = (2.86)
p0 ρ0 T0 p0

Also, the pressure–temperature–density relation of a perfect gas is


p p0
= =R (2.87)
ρT ρ0 T0

ISTUDY
2.13. GAS DYNAMICS 119

The temperature, pressure, and density ratios as functions of Mach number


are
( )
T0 γ−1 2
= 1+ M (2.88)
T 2

( )γ/(γ − 1)
p0 γ−1 2
= 1+ M (2.89)
p 2

( )1/(γ − 1)
ρ0 γ−1 2
= 1+ M (2.90)
ρ 2

where T0 , p0 and ρ0 are the temperature, pressure and density, respectively


at the stagnation state. The particular value of temperature, pressure, and
density ratios at the critical state (i.e. at the minimum area) are found by
setting M = 1 in Eqs. (2.88)–(2.90). For γ = 1.4, the following are the
resulting formulae.

T∗ a∗
2
2
= 2 = = 0.8333 (2.91)
T0 a0 γ+1

( )γ/(γ − 1)
p∗ 2
= = 0.5283 (2.92)
p0 γ+1

( )1/(γ − 1)
ρ∗ 2
= = 0.6339 (2.93)
ρ0 γ+1

where, T ∗ , p∗ and ρ∗ are the temperature, pressure and density, respectively,


at the critical state.
The critical pressure ratio p∗ /p0 is of the same order of magnitude for all
gases. It varies almost linearly with γ, from 0.6065 for γ = 1 to 0.4867 for
γ = 1.67.
The dimensionless velocity M ∗ is one of the most useful parameters in gas
dynamics. Generally, it is defined as

V V
M∗ ≡ ∗
≡ ∗ (2.94)
a V

where a∗ is the critical speed of sound. This dimensionless velocity can also
be expressed in terms of Mach number M as

(γ + 1)M 2
M∗ =
2
(2.95)
(γ − 1)M 2 + 2

ISTUDY
120 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

Area–Mach number relation


For isentropic flow of a perfect gas through a duct, the Area–Mach number
relation may be expressed, assuming one-dimensional flow, as

( )2 [ ( )](γ + 1)/(γ − 1)
A 1 2 γ−1 2
= 2 1+ M (2.96)
A∗ M γ+1 2

where A∗ is called the sonic or critical throat area. From Eq. (2.96) we get
the striking result M = f (A/A∗ ), i.e. the Mach number at any location in
the duct is a function of the ratio of the local area of the duct to the sonic
throat area. The local area A of the duct must be larger than or at least
equal to A∗ ; the case in which A < A∗ it is physically impossible for an
isentropic flow to occur. Further, from Eq. (2.96), for any given A/A∗ > 1,
two values of M are obtained, a subsonic value and a supersonic value. Once
the variation of Mach number through the nozzle is known, the variation of
static temperature, pressure, and density can be determined from isentropic
relations, i.e. Eqs. (2.88)–(2.90). The pressure, the temperature, and the
density decrease continuously throughout the nozzle. Also, the exit pressure,
the density, and the temperature ratios depend only on the exit area ratio
Ae /A∗ . That is, if the nozzle is part of a supersonic wind tunnel, then the
test-section conditions are determined by Ae /A∗ (geometry of the nozzle) and
the stagnation pressure p0 and the temperature T0 (properties of the gas in
the reservoir).
A variety of flow fields can be generated in the convergent-divergent nozzle
by independently governing the back pressure downstream of the nozzle exit.
Consider the flow through a Laval nozzle as shown in Figure 2.42.

FIGURE 2.42 Flow in a convergent-divergent nozzle.

ISTUDY
2.13. GAS DYNAMICS 121

When pe = p0 , there will be no flow through the nozzle. Let pe be reduced


to pe1 slightly below p0 . This small favourable pressure gradient will cause a
flow through the nozzle at low subsonic speeds. The local Mach number will
increase continuously through the convergent portion of the nozzle, reaching
a maximum at the throat. In other words, the static pressure will decrease
continuously in the convergent portion of the nozzle, reaching a minimum at
the throat, as shown by curve a in the figure. Assume that pe is reduced
further to pe2 . Then the pressure gradient will be stronger, the flow accel-
eration will be faster and the variation of Mach number and static pressure
through the duct will be larger, as shown by curve b. Similarly, if pe is re-
duced continuously, at some value of pe , the flow will reach sonic velocity at
the throat, as shown by curve c. For this case, Ath = A∗ . Now, the sonic
flow at the throat will expand further in the divergent portion of the nozzle
as supersonic flow if pe /pth < 1, and will decelerate as subsonic flow as shown
by curve c, for pe3 /pth > 1.
For the cases discussed above, for a given p0 , the mass flow through
the duct increases as pe decreases. This mass flow can be calculated with
ṁ = ρth Ath Vth . When pe is equal to pe3 , where the sonic flow is attained
at the throat, ṁ = ρ∗ A∗ a∗ ; also, the Mach number at the throat is unity.
Hence, the flow properties at the throat, and indeed throughout the subsonic
section of the duct, become ‘frozen’ when pe < pe3 , i.e. the subsonic flow in
the convergent portion of the nozzle becomes ‘unaffected’ and the mass flow
remains constant for pe < pe3 . This condition for sonic flow at the throat is
called the choked flow. For further reduction of pe below pe3 , after the flow
becomes ‘choked’, the mass flow remains constant.
At this stage it is important to realize that the choked mass flow rate ṁ ∗
is the maximum only for a given p0 and T0 . However, when the stagnation
pressure and temperature are altered, ṁ∗ will have a different maxima cor-
responding to every set of p0 and T0 . The choked mass flow for gases with
γ = 1.4 can be expressed as

0.6847 ∗
ṁ∗ = √ A p0
R T0

From the foregoing discussions, it is clear that in the convergent portion of


the duct, the flow remains unchanged for back pressures below pe3 . But in
the divergent portion of the duct the flow expands as a supersonic flow for
pe < pe3 . However, pe should be adequately reduced to a specified low value
pec , for an isentropic expansion of flow in the divergent portion of the nozzle,
resulting in a shock-free supersonic flow; the variation of pressure for such an
isentropic expansion is shown by curve d in Figure 2.42.
For values of exit pressures between pec and pe3 , a normal shock wave
exists inside the divergent portion of the nozzle. The flow behind the normal
shock is subsonic; hence the static pressure increases to pe4 at the exit. The
normal shock moves downstream with decrease in pe below pe4 and will stand
precisely at the exit when pe = pe5 , where pe5 is the static pressure behind

ISTUDY
122 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

a normal shock at the design Mach number of the nozzle; this is shown in
Figure 2.43(a). When pe is further reduced such that pec < pb < pe5 , the
flow inside the nozzle is fully supersonic and isentropic where pb , the pressure
of the ambient atmosphere to which the flow is discharged, is called the back
pressure. Further increase in the flow pressure, resulting in equilibrium with
pb , takes place across an oblique shock attached to the nozzle exit outside
the duct, as shown in Figure 2.43(b). For further reduction in back pressure
below pec , equilibration of the flow takes place across expansion waves outside
the duct, as illustrated in Figure 2.43(c).

FIGURE 2.43 Flow with shock and expansion waves at the exit of a
convergent-divergent nozzle.

When the flow situation is as shown in Figure 2.43(b), the nozzle is said
to be overexpanded, since the pressure at the exit has expanded below the
back pressure, pe < pb . Conversely, when the situation is as shown in
Figure 2.43(c), the nozzle is said to be underexpanded, since the exit pressure
is higher than the back pressure, i.e. pe > pb , and hence the flow experiences
additional expansion after leaving the nozzle.

ISTUDY
2.13. GAS DYNAMICS 123

The preceding results can be summarized as follows:


1. For pe1 /p0 < 1, the flow is subsonic at the throat and so the divergent
portion acts as a diffuser. This is the case of flow through a venturi.
2. For pe3 /p0 < 1, the pressure at the throat is p∗ and so M = 1 at the
throat. But as pe3 /pth > 1, the divergent portion acts as a diffuser and
the flow does not become supersonic.
3. For pressure at the exit equal to pec , the flow expands isentropically
in the nozzle and there is shock-free supersonic flow in the divergent
portion of the nozzle.
4. For pe3 /p0 < pe /p0 < pec /p0 , there will be supersonic velocity locally,
but at the exit it cannot be supersonic, and so there will be a jump
in static pressure at some section of the nozzle, i.e. there is a shock.
Therefore, there is a certain back pressure pec , above which there cannot
be supersonic flow at the exit. Only below pec , there can be shock-free
supersonic flow at the exit.

Prandtl–Meyer function
The Prandtl–Meyer function ν is an important parameter to solve supersonic
flow problems involving isentropic expansion or isentropic compression. Ba-
sically, the Prandtl–Meyer function is a similarity parameter. The Prandtl–
Meyer function can be expressed in terms of M as

√ √
γ+1 γ−1
ν= arc tan (M 2 − 1) − arc tan M 2 − 1 (2.97)
γ−1 γ+1

From Eq. (2.97) it is seen that, for a given M , ν is fixed.

EXAMPLE 2.29 A supersonic wind tunnel is designed to generate Mach


2.0 flow in the test-section, with standard sea level conditions. Compute the
ratio of exit area (Ae ) to throat area (A∗ ) and the settling chamber conditions
necessary to achieve the designed conditions.

Solution We know that at sea level, p = 101.32 kPa and T = 288 K. These
values are the required pressure and temperature at the test-section.
From Eqs. (2.88), (2.89) and (2.96), at M = 2.0, we have

T0 /T = 1.8, p0 /p = 7.824, Ae /A∗ = 1.687


Since p = 101.32 kPa and T = 288 K,

p0 = 792.73 kPa , T0 = 518.4 K


Note: This problem may also be solved with isentropic tables instead of using
the above equations.

ISTUDY
124 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

EXAMPLE 2.30 An underexpanded, two-dimensional supersonic nozzle


exhausts into a region where p = 0.75 atm. The flow at the nozzle exit plane
is uniform with p = 1.6 atm and M = 2.0. Determine the flow direction and
Mach number after the initial expansion.

Solution Since the flow is underexpanded, it will expand isentropically


through an expansion fan at the nozzle exit. Let the upstream and down-
stream states of the expansion fan be 1 and 2, respectively.
From the isentropic relations, Eq. (2.89), for M = 2.0, we have

p1 /p01 = 0.1278

Since p01 = p02 for isentropic expansion,

p2 /p02 = (0.1278)(0.75/1.60) = 0.0599

For this pressure ratio, from Eq. (2.89), we get M2 = 2.48. Now, from
Eq. (2.97),
ν1 = 26.38◦ , ν2 = 38.655◦
Therefore, the flow turning angle ν12 = ν2 − ν1 = 12.275◦ .

2.13.4 Normal Shock Relations


The shock may be described as a compression front in a supersonic flow field
and flow process, across which results an abrupt change in fluid properties.
The thickness of the shock is comparable to the mean free path of the gas
molecules in the flow field. When the shock is normal to the flow direction
it is called normal shock and when it is inclined at an angle to the flow it is
termed oblique shock. For a perfect gas, it is known that all the flow property
ratios across a normal shock are unique functions of γ and the upstream Mach
number.
Considering the normal shock shown in Figure 2.44 the following normal
shock relations, assuming the flow to be one-dimensional, can be obtained.

FIGURE 2.44 Flow through a normal shock.

ISTUDY
2.13. GAS DYNAMICS 125

γ−1 2
1+ M1
M22 = 2 (2.98)
γ−1
γM12 −
2
p2 2γ
=1+ (M12 − 1) (2.99)
p1 γ+1

ρ2 V1 (γ + 1)M12
= = (2.100)
ρ1 V2 (γ − 1)M12 + 2

T2 h2 a2 2(γ − 1) (γM12 + 1)
= = 22 = 1 + (M12 − 1) (2.101)
T1 h1 a1 (γ + 1)2 M12
In Eq. (2.101), h1 and h2 are the static enthalpies upstream and downstream
of the shock, respectively. The total pressure ratio on either side of a normal
shock can be expressed as
[ ]−1/(γ−1) [ ]γ/(γ−1)
p02 2γ (γ + 1)M12
= 1+ (M − 1)
2
(2.102)
p01 γ+1 1 (γ − 1)M12 + 2

The change in entropy across the normal shock is given by


p01
s2 − s1 = R ln (2.103)
p02
This equation justifies that although mathematically it is possible for the flow
to decelerate from supersonic to subsonic and vice versa across a normal shock
wave, only the former is physically feasible. From Eq. (2.103), if M1 < 1,
△s < 0; and if M1 > 1, we have △s > 0. Since it is necessary that △s ≥ 0,
therefore, from the second law of thermodynamics, for a process to take place,
M1 must be greater than or equal to 1. When M1 is subsonic, the entropy
across the wave decreases, which is impossible. Therefore, the only physically
feasible flow is for M1 > 1, and from the above results we have M2 < 1,
ρ2 /ρ1 > 1, p2 /p1 > 1, and T2 /T1 > 1.
The changes in flow properties across the shock take place over a very
short distance, of the order of 10−5 cm. Hence, the velocity and temperature
gradients inside the shock structure are very large. These large gradients
result in increase of entropy across the shock. Also, these gradients internal
to the shock provide heat conduction and viscous dissipation that render the
shock process internally irreversible.
There is no heat added or taken away from the flow as it traverses the
shock wave, i.e. the flow across the shock wave is adiabatic. Therefore, the
total temperature remains the same ahead of and behind the shock wave, i.e.

T02 = T01

Thus the flow through a shock is adiabatic but irreversible.

ISTUDY
126 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

It is natural that one may wonder, how is it possible to have no heat


transfer in spite of the presence of a severe temperature gradient (∆T /∆x =
(T2 − T1 )/∆x, where x is the shock thickness, which is of the order of the
mean free path) across the shock. The answer to this doubt is the following.
Even though there is a significantly high level of temperature gradient across
the shock wave, the heat transfer experienced by the flow is negligible, be-
cause the flow passes the thin region of the shock with a rapid speed. Due
to this rapid speed the contact time available for the high temperature fluid
elements to transfer heat to the surrounding becomes extremely small, leading
to insignificant heat transfer. This makes the process approximately adiabatic.

EXAMPLE 2.31 The flow Mach number, pressure and temperature ahead
of a normal shock are 2.0, 0.5 atm and 300 K, respectively. Compute M2 , p2 , T2 ,
and V2 downstream of the wave.

Solution For M1 = 2.0, from Eqs. (2.98), (2.99) and (2.101), we get

M2 = 0.5774 , p2 /p1 = 4.5, T2 /T1 = 1.687

Therefore,

p2 = 4.5 × 0.5 = 2.25 atm

T2 = 1.687 × 300 = 506.1 K


At 506.1 K, the speed of sound is

a2 = γRT2 = 450.94 m/s

Hence,
V2 = M2 a2 = 0.5774 × 450.94 = 260.37 m/s

2.13.5 Oblique Shock Relations


In the previous section, a compression wave normal to the flow direction was
studied. However, in a wide variety of physical situations, a compression wave
inclined at an angle to the flow occurs. Such a wave is called an oblique shock.
It is usually much simpler to view the motion from a frame of reference system
in which the body is stationary and the fluid flows over it. If the relative speed
is supersonic, the disturbance waves cannot propagate with the body. Thus,
in the reference system in which the body is stationary, the wave system is
also stationary; then the correspondence between the wave system and the
flow field is direct.
The normal shock wave is a special case of oblique shock wave. Also, it
can be shown that the superposition of a uniform velocity, which is normal
to the upstream flow, on the flow field of the normal shock will result in
a flow field through an oblique shock wave. Oblique shocks usually occur

ISTUDY
2.13. GAS DYNAMICS 127

when a supersonic flow is turned into itself. The opposite to this, i.e. when
a supersonic flow is turned away from itself, results in the formation of an
expansion fan. These two families of waves play a dominant role in all flow
fields involving supersonic velocities.
Consider the flow through an oblique shock wave, as shown in Figure 2.45.
The flow through a normal shock has been modified to result in flow through
an oblique shock, by superimposing a uniform velocity Vy (parallel to the
normal shock) on the field of the normal√ shock (Figure 2.45). The resultant
velocity upstream of the shock is V1 = 2 + V 2 and is inclined at an angle
Vx1 y
β[= tan−1 (Vx1 /Vy )] to the shock. This angle β is called the shock angle. The
velocity component Vx2 is always less than Vx1 ; therefore, the inclination of
the flow to the shock ahead of the shock and that after the shock are different.
The inclination ahead is always more than that behind the shock wave, i.e.
the flow is turned suddenly at the shock. Since Vx1 is always more than Vx2 ,
the turn is always towards the shock. The angle θ by which the flow turns
towards the shock is called the flow deflection angle θ, and is positive as shown
in Figure 2.45. The rotation of the flow field in Figure 2.45(a) by an angle β
results in the field shown in Figure 2.45(b), with V1 in the horizontal direction.
The shock in that field inclined at an angle β to the incoming supersonic flow
is called the oblique shock.
The relations between the flow parameters upstream and downstream of
the flow field through the oblique shock, illustrated in Figure 2.45(b), can be
obtained from the normal shock relations, since the superposition of uniform
velocity Vy on the normal shock flow field in Figure 2.45(a) does not affect
the flow parameters (e.g. static pressure) defined for the normal shock. The
only change is that in the present case the upstream Mach number is

resultant velocity V1
M1 = =
speed of sound a1

FIGURE 2.45 Flow through an oblique shock.

The component of M1 normal to the shock wave is

Mn1 = M1 sin β

where β is the shock angle.

ISTUDY
128 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

The shock in Figure 2.45 can be visualized as a normal shock with up-
stream Mach number M1 sin β. Thus, the replacement of M1 in normal shock
relations Eqs. (2.98) to (2.102) by M1 sin β results in the corresponding rela-
tions for the oblique shock giving
2
M12 sin2 β +
2 γ−1
Mn2 = (2.104)

M 2 sin2 β − 1
γ−1 1

p2 2γ
=1+ (M 2 sin2 β − 1) (2.105)
p1 γ+1 1

ρ2 (γ + 1)M12 sin2 β
= (2.106)
ρ1 (γ − 1)M12 sin2 β + 2

T2 a2 2(γ − 1) M12 sin2 β − 1 ( )


= 22 = 1 + 2 2 2 γM12 sin2 β + 1 (2.107)
T1 a1 (γ + 1) M1 sin β

p02 2γ ( 2 2 )−1/(γ−1) (γ + 1)M12 sin2 β


=1+ M1 sin β − 1 (2.108)
p01 γ+1 (γ − 1)M12 sin2 β + 2

The entropy change across the oblique shock is given by


p01
s2 − s1 = R ln
p02

Equation (2.104) gives only the normal component of Mach number Mn2
behind the shock. But the Mach number of interest is M2 . It can be obtained
from Eq. (2.104) as follows:
From the geometry of the oblique shock flow field in Figure 2.45, it is seen
that M2 is related to Mn2 by
Mn2
M2 = (2.109)
sin(β − θ)
where θ is the flow turning angle across the shock. Combining Eqs. (2.104)
and (2.109), the Mach number M2 after the shock can be obtained.

EXAMPLE 2.32 An oblique shock is incident on a solid boundary, as


shown in Figure 2.46(a). The boundary is to be turned through such an angle
that there is no reflected wave. Determine the angle θ and the flow Mach
number M2 .

Solution The flow field is as shown in Figure 2.46(b).

ISTUDY
2.13. GAS DYNAMICS 129

FIGURE 2.46(a) Incident oblique shock.

FIGURE 2.46(b) Wave cancellation turning.

For M1 = 3.5 and β = 45◦ , from oblique shock chart, we get θ = 28.158◦ .

Mn1 = M1 sin β = 3.5 × sin 45◦ = 2.47

From Eq. (2.104),


Mn2 = 0.516

For this Mn2 , using Eq. (2.109), we get M2 = 1.78 .

2.13.6 Flow with Friction


In Section 2.13.3 on flow with area change, it was assumed that the changes
in flow properties, for compressible flow of gases in ducts, were brought about
solely by area change. That is, the effects of viscosity have been neglected.
But in practical flow situations like stationary power plants, aircraft engines,
high vacuum technology, transport of natural gas in long pipe lines, transport
of fluids in chemical process plants, and various types of flow systems, the
high-speed flow travels through passages of sufficient length and the effects
of viscosity (friction) cannot be neglected for such flows. In many practical
flow situations, friction can even have a decisive effect on the resultant flow
characteristics.
Consider one-dimensional steady flow of a perfect gas with constant spe-
cific heats through a constant area duct. Assume that there is neither external
heat exchange nor external shaft work and the difference in elevation produces
negligible changes in flow properties compared to frictional effects. The flow
with the above said conditions, namely the adiabatic flow with no external

ISTUDY
130 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

work is called the Fanno line flow. For Fanno line flow, the wall friction (due
to viscosity) is the chief factor bringing about changes in flow properties.

Working formulae for Fanno type flow


Consider the flow of a perfect gas through a constant area duct, choosing an
infinitesimal control volume as shown in Figure 2.47. The relation between
Mach number M and friction factor f can be written as
∫ ∫
Lmax
dx 1
1 − M2
4f = ( ) dM 2 (2.110)
0 D M2 γ−1 2
γM 4 1+ M
2

FIGURE 2.47 Control volume for Fanno flow.

In Eq. (2.110) the integration limits are taken as (a) the section where the
Mach number is M , and where the length x is arbitrarily set equal to zero,
and (b) the section where the Mach number is unity and x is the maximum
possible length of duct, Lmax , and D is the hydraulic diameter, defined as

4(cross-sectional area)
D∼
=
wetted perimeter

On integration, Eq. (2.110) yields


 
Lmax 1 − M2 γ+1  2
 ( (γ + 1)M ) 

4f = + ln   (2.111)
D γM 2 2γ γ − 1
2 1+ M2
2

where f is the mean friction coefficient with respect to duct length, defined
by
∫ Lmax
1
f= f dx
Lmax 0

ISTUDY
2.13. GAS DYNAMICS 131

Likewise, the local flow properties can be found in terms of the local Mach
number. Indicating the properties at M = 1 as superscripted with ‘aster-
isk’ and integrating between the duct sections with M = M and M = 1,
the following relations can be obtained (E. Rathakrishnan, Gas Dynamics,
Prentice-Hall of India, New Delhi, 1995).
 1/2
p 1 
 ( γ+1

= ) (2.112)
p∗ M γ−1 2 
2 1+ M
2
 1/2
V  γ+1 
=M
 ( ) (2.113)
V ∗ γ−1 2 
2 1+ M
2
 
T a2  γ+1 
= ∗2 = 
 ( ) (2.114)
T ∗ a γ−1 2 
2 1+ M
2

 ( ) 1/2
γ−1 2
2 1 + M
ρ V∗ 1 
 2 


= =   (2.115)
ρ V M γ+1

 ( ) (γ+1)/2(γ−1)
γ−1 2
2 1+ M 
p0 1 
 2 
= (2.116)
p∗0 M γ+1 

F 1 + γM 2
F∗
= [ ( )]1/2 (2.117)
γ−1 2
M 2(γ + 1) 1 + M
2

In Eq. (2.117), the parameter F is called the impulse function, defined as

F ∼
= pA + ρAV 2 = pA(a + γM 2 )

The superscript ∗ in the above equations refers to sonic state.

EXAMPLE 2.33 Air enters through an insulated constant area duct of


diameter 0.02 m and length 1.05 m. The average friction coefficient for the
duct is 0.002. If the stagnation conditions of the flow at the entrance are

ISTUDY
132 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

3 MPa and 500 K, determine the maximum mass flow rate of the air possible
through the duct.

Solution For maximum mass flow rate, the duct exit Mach number should
be unity. For the given duct, we have

4f L∗1 4 × 0.002 × 1.05


= = 0.42
D 0.02
where the subscript 1 refers to inlet condition.
For 4f L∗1 /D = 0.42, from Fanno flow table (E. Rathakrishnan, Gas Dy-
namics, Prentice-Hall of India, New Delhi, 1995), we have the inlet Mach
number as M1 = 0.62. For M1 = 0.62, from isentropic table, we have
p1 T1
= 0.7716, = 0.9286
p01 T01
Thus,

p1 = 0.7716 p01
= 0.7716 × 3 = 2.315 MPa
T1 = 0.9286 T01
= 0.9286 × 500 = 464.3 K

The corresponding density is


p1
ρ1 =
R T1
2.315 × 106
= = 17.37 kg/m3
287 × 464.3
Thus, the mass flow rate ṁ is

ṁ = ρ1 A1 V1
π
= 17.37 × × 0.022 × M1 a1
4
π √
= 17.37 × × 0.022 × 0.62 × 1.4 × 287 × 464.3
4

= 1.46 kg/s

EXAMPLE 2.34 Determine the length of the duct required if it has to


decelerate air flow from Mach 3.0 to Mach 1.5, adiabatically. The diameter
of the duct is 30 cm and the average friction coefficient f = 0.005.

ISTUDY
2.13. GAS DYNAMICS 133

Solution Let the station at M = 3.0 be denoted by the subscript 1 and that
at M = 1.5 by subscript 2. From Fanno flow table, for M1 = 3.0, we have

4f L∗1
= 0.5222
D
For M2 = 1.5, we have from Fanno flow table

4f L∗2
= 0.1360
D
Therefore, the distance between the sections 1 and 2 gives the required length
as
D
△L = (0.5222 − 0.1360)
4f

0.3
= (0.3862)
4 × 0.005

= 5.79 m

2.13.7 Flow with Simple T0 -Change


In the section on flow with area change, the process was considered to be
isentropic with frictional and energy effects absent. In Fanno line flow, only
the effect of wall friction was taken into account in the absence of area change
and energy effects. In the present section, the flow process involving change in
the stagnation temperature or the stagnation enthalpy of a gas stream, which
flows in a frictionless constant area duct, is considered. From one-dimensional
point of view, this is yet another effect producing continuous changes in the
state of a flowing stream and this factor is called energy effect such as exter-
nal heat exchange, combustion, or moisture condensation. Though a process
involving a simple stagnation temperature T0 -change is difficult to achieve in
practice, many useful conclusions of practical significance may be drawn by
analysing the process of simple T0 -change. This kind of flow involving only
T0 -change is called Rayleigh type flow.

Working formulae for Rayleigh type flow


Consider the flow of a perfect gas through a constant-area duct without
friction, as shown in Figure 2.48.
Considering a control volume as in Figure 2.48, the normalized expressions
(working formulae) for the flow process involving only heat transfer can be
obtained as (E. Rathakrishnan, Gas Dynamics, Prentice-Hall of India, New
Delhi, 1995).
p γ+1

= (2.118)
p 1 + γM 2

ISTUDY
134 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

FIGURE 2.48 Control volume for Rayleigh flow.

ρ V∗ 1 + γM 2
= = (2.119)
ρ∗ V (γ + 1)M 2

T (γ + 1)2 M 2

= (2.120)
T (1 + γM 2 )2

γ−1 2
T0 2(γ + 1)M 2 (1 + M )
= 2 (2.121)

T0 (1 + γM ) 2 2

 
γ − 1 2 γ/(γ−1)
γ+1  2(1 + M )
p0 2 
∗ =   (2.122)
p0 1 + γM 2 γ+1

where the superscript ∗ refers to sonic state.

EXAMPLE 2.35 Air at 300 K, 101 kPa and Mach 2 flows through a
frictionless constant area pipe. Heat is added to the flow to decelerate it to
Mach 1.2. Determine the resultant temperature, the pressure, and the density.
Solution The given flow is frictionless and the change of state is brought
about solely by heat addition. Thus, it is a Rayleigh flow. Let the subscripts
1 and 2 refer to the initial and final states of the flow, respectively. Therefore,

M1 = 2.0, p1 = 101 kPa, T1 = 300 K

From the Rayleigh flow table (E. Rathakrishnan, Gas Dynamics, Prentice-
Hall of India, 1995), for M1 = 2.0, we have

T1 p1 ρ∗
= 0.52893, = 0.36364, = 1.4545
T∗ p∗ ρ1

ISTUDY
2.14. SUMMARY 135

For M2 = 1.2, we have


T2 p2 ρ∗
= 0.91185, = 0.79576, = 1.1459
T∗ p∗ ρ2
Thus,
T2 T ∗
T2 = T1
T ∗ T1
0.91185
= × 300
0.52893

= 517.19 K

p2 p∗
p2 = p1
p∗ p1
0.79576
= × 101
0.36364

= 221.02 kPa

ρ2 ρ∗
ρ2 = ρ1
ρ∗ ρ1
ρ2 ρ∗ p1
=
ρ∗ ρ1 RT1

1.4545 101 × 103


= ×
1.1459 287 × 300

= 1.49 kg/m3

Note: The final density ρ2 can also be obtained from the calculated p2 and
T2 . Thus,
p2
ρ2 =
RT2
221.02 × 103
=
287 × 517.19

= 1.49 kg/m3

2.14 SUMMARY
Fluid Mechanics is the science of fluid flow where the temperature changes
involved are less than 5 per cent. Just the continuity and momentum equations
are sufficient to solve them. On the other hand, Gas Dynamics is that science

ISTUDY
136 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

which deals with flows involving appreciable change (more than 5 per cent) in
both density and temperature. Hence, in gas dynamic analysis all the three
conservation equations, namely the continuity, the momentum and the energy
equations become essential.
Fluid may be defined as a substance which will continue to change shape
as long as there is a shear stress present, however small it may be.
Pressure may be defined as the force per unit area which acts normal to
the surface of any object which is immersed in a fluid. For a fluid at rest,
at any point the pressure is the same in all directions. The pressure in a
stationary fluid varies only in the vertical direction and is constant in any
horizontal plane. That is, in stationary fluids the pressure increases linearly
with depth. This linear pressure distribution is called the hydrostatic pressure
distribution.
By virtue of their motion the molecules possess kinetic energy, and this
energy is sensed as temperature of the solid, liquid or gas. In the case of a gas
in motion it is called the static temperature.
The total number of molecules in a unit volume is a measure of the density
ρ of the substance. It is expressed as mass per unit volume, say kg/m3 .
Mass is defined as weight divided by acceleration due to gravity. At standard
atmospheric temperature and pressure (288.15 K and 101325 Pa) the density
of dry air is 1.225 kg/m3 .
The property which characterizes the resistance that a fluid offers to
applied shear force is termed viscosity. This resistance, unlike for solids,
does not depend upon the deformation itself but on the rate of deformation.
Viscosity is often regarded as the stickiness of a fluid and its tendency is to
resist sliding between layers.
For air, the Sutherland’s relation can also be expressed as
( )
−6 T 3/2
µ = 1.46 × 10 kg/m · s
T + 111

where T is in kelvin.
The change in volume of a fluid associated with change in pressure is called
compressibility. When a fluid is subjected to pressure it gets compressed and
its volume changes. Bulk modulus of elasticity is a measure of how easily the
fluid may be compressed, and is defined as the ratio of pressure change to
volumetric strain associated with it.
The specific heat is defined as the amount of heat required to raise the
temperature of a unit mass of medium by one degree. The specific heat at
constant volume is defined as
( )
∂u
cv ≡
∂T v
The specific heat at constant pressure is defined as
( )
∂h
cp ≡
∂T p

ISTUDY
2.14. SUMMARY 137

The ratio of specific heats given by

cp
γ=
cv

is an important parameter in the study of compressible flows. This is a mea-


sure of the relative internal complexity of the gas molecules. It has been
determined from the kinetic theory of gases that the ratio of specific heats
can be related to the number of degrees of freedom n of the gas molecules by
the relation
n+2
γ=
n
In general γ varies from 1 to 1.67, i.e.

1 ≤ γ ≤ 1.67

Basically, two treatments are followed for fluid flow analysis. They are
Lagrangian and Eulerian descriptions. The Lagrangian method describes the
motion of each particle of the flow field in a separate and discrete manner.
The relation
DV ∂V
= + (V · ∇)V
Dt ∂t
is known as the Euler’s acceleration formula. This links the particle approach
and the field approach.
The three important concepts for visualizing or describing flow fields are
the pathline, the streakline and the streamline.
Pathline may be defined as a line in the flow field describing the trajectory
of a given fluid particle. Streakline may be defined as the instantaneous locus
of all the fluid elements that have passed the point of injection at some earlier
time.
Streamlines are imaginary lines in a fluid flow, drawn in such a manner
that the flow velocity is always tangential to it.
The pathline, the streamline and the streakline are different in general but
coincide in a steady flow.
In modern fluid flow analysis, yet another graphical representation namely
the timeline is used. When a pulse input is periodically imposed on a line of
tracer source placed normal to a flow, a change in the flow profile can be
observed. The tracer image is generally termed timeline. Timelines are often
generated in the flow field to aid the understanding of flow behaviour such as
velocity and velocity gradient.
In the range of engineering interest, four basic laws must be satisfied for
any continuous medium. They are:
1. Conservation of matter (continuity equation)
2. Newton’s second law (momentum equation)

ISTUDY
138 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

3. Conservation of energy (the first law of thermodynamics)

4. Increase-of-entropy principle (the second law of thermodynamics)

In addition to these primary laws, there are numerous subsidiary laws, some-
times called the constitutive relations, that apply to specific types of media or
flow processes (e.g. equation of state for perfect gas, Newton’s viscosity law
for certain viscous fluids, isentropic process, adiabatic process).
For air at normal temperature and pressure, the density ρ, the pressure p
and the temperature T are connected by the relation

p = ρRT

where R is a constant called the gas constant.


Kinematics is the branch of mechanics that deals with quantities involving
space and time only. It treats variables such as displacement, velocity, accel-
eration, deformation, and rotation of fluid elements without referring to the
forces responsible for such a motion. In other words, kinematics essentially
describes the appearance of a motion. Another branch of mechanics which
deals with the forces due to motion is termed the dynamics.
For steady incompressible flow the continuity equation in differential
form is
∂Vx ∂Vy ∂Vz
+ + =0
∂x ∂y ∂z

The general form of the continuity equation which expresses the differential
form of the conservation of mass principle is given by

∂ρ
+ ∇ · (ρV ) = 0
∂t

For steady incompressible flows the momentum equation can be written


as
( )
∂Vx ∂Vx ∂Vx 1 ∂p ∂ 2 Vx ∂ 2 Vx ∂ 2 Vx
Vx + Vy + Vz =− +ν + +
∂x ∂y ∂z ρ ∂x ∂x2 ∂y 2 ∂z 2
( )
∂Vy ∂Vy ∂Vy 1 ∂p ∂ 2 Vy ∂ 2 Vy ∂ 2 Vy
Vx + Vy + Vz =− +ν + +
∂x ∂y ∂z ρ ∂y ∂x2 ∂y 2 ∂z 2
( )
∂Vz ∂Vz ∂Vz 1 ∂p ∂ 2 Vz ∂ 2 Vz ∂ 2 Vz
Vx + Vy + Vz =− +ν + +
∂x ∂y ∂z ρ ∂z ∂x2 ∂y 2 ∂z 2
These equations are the x, y, z components of the momentum equation,
respectively. These equations are generally known as Navier–Stokes equations.
Boundary layer thickness δ may be defined as the distance from the wall
out to where the fluid velocity is within 1 per cent of the local main stream
velocity.

ISTUDY
2.14. SUMMARY 139

Displacement thickness δ ∗ may be defined as the distance by which the


boundary would have to be displaced if the entire flow were imagined to be
frictionless and the same mass flow maintained at any section.
Circulation is an abstract quantity, defined as the line integral of velocity
vector between any two points in a flow field. By definition, the circulation is
given as
H
Γ ≡ C V · dl

where dl is an elemental length, and the loop through the integral sign signifies
that the contour is closed.
The vorticity ζ at a point equals the circulation per unit area, i.e.

Γ
ζ=
A
In vector form, ζ becomes

ζ = ∇ × V = curl V

For a two-dimensional flow in xy-plane, ζ becomes

∂Vy ∂Vx
ζz = −
∂x ∂y

If the vorticity components are zero, the flow is known as irrotational flow.
Inviscid flows are essentially irrotational flows.
Based on the streamline concept, a mathematical function ψ called the
stream function can be defined. The velocity components of a flow field can
be obtained by differentiating the stream function.
In terms of stream function ψ the velocity components of a two-dimensional
incompressible flow are given as

∂ψ ∂ψ
Vx = , Vy = −
∂y ∂x

If the flow is compressible, the velocity components are expressed as

1 ∂ψ 1 ∂ψ
Vx = , Vy = −
ρ ∂y ρ ∂x

It is important to note that the stream function is defined only for two-
dimensional flows and the definition does not exist for three-dimensional flows.
Even though, some books define ψ for axisymmetric flows, they again prove to
be equivalent to two-dimensional flows. We must realize that the definition of
ψ does not exist for three-dimensional flows. This is because such a definition
demands a single tangent at any point on a streamline, which is possible only
in two-dimensional flows.

ISTUDY
140 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

In cylindrical polar coordinates the continuity equation for a two-dimensional


incompressible flow can be written as

1 ∂(rVr ) 1 ∂Vθ
+ =0
r ∂r r ∂θ
and the velocity components Vr and Vθ can be related to the stream function
ψ(r, θ), as
1 ∂ψ ∂ψ
Vr = , Vθ = −
r ∂θ ∂r
For irrotational flows (the fluid elements in the field are free of angular
motion), there exists a function ϕ called the velocity potential or potential
function. For two-dimensional flows, ϕ must be a function of x, y and t. The
velocity components are given by
∂ϕ ∂ϕ
Vx = , Vy =
∂x ∂y
Also,
∂ψ ∂ϕ ∂ψ ∂ϕ
= , =−
∂y ∂x ∂x ∂y
These relations between ψ and ϕ are the famous Cauchy–Riemann equations
of complex-variable theory. Unlike the stream function, the potential function
exists for three-dimensional flows as well.
Potential flow is based on the concept that the flow field can be represented
by a potential function ϕ such that

∇2 ϕ = 0

This linear partial differential equation is popularly known as Laplace


equation.
For two-dimensional potential flows, we have the vorticity ζz as
∂Vy ∂Vx
ζz = − =0
∂x ∂y
The Bernoulli equation is a statement of conservation of mechanical energy
between any two points along a streamline. The constant in the equation,
known as the Bernoulli constant, in general varies from one streamline to
another but remains constant along a streamline in a steady, frictionless,
incompressible flow. These four assumptions are needed and must be kept
in mind while applying this equation.
Each term in the Bernoulli equation may be interpreted as a form of avail-
able energy. This equation is also referred to as a conservation of mechanical
energy equation. It is essential here to realize that this energy equation was
derived from the conservation of momentum equation. Thus, the Bernoulli
equation is essentially a momentum equation and not an energy equation.

ISTUDY
2.14. SUMMARY 141

A type of flow in which the fluid emanates from the origin and spreads
radially outwards to infinity is called a source. The stream function for a
source of strength q̇ can be obtained as


ψ= θ

A sink is a type of flow in which the fluid at infinity flows radially towards
the origin. The stream function for a sink is


ϕ=− ln r

Vortex motion is a fluid flow in which the streamlines are concentric circles.
The vortex motions encountered in practice may in general be classified as ‘free
vortex or potential vortex’ and ‘forced vortex or flywheel vortex’. The stream
function for a free vortex is


ψ=− ln r

A doublet or a dipole is a potential flow field due to a source and a sink


of equal strength, brought together in such a way that the product of their
strength and the distance between them remain constant. Its potential func-
tion is
m
ϕD = cos θ
2πr

The lateral force experienced by rotating bodies is called the Magnus effect.
The lower critical Reynolds number is that Reynolds number below which
the entire flow is laminar. The Reynolds number above which the entire flow is
turbulent is called the upper critical Reynolds number. The critical Reynolds
number is that at which the flow field is a mixture of laminar and turbulent
flows. It is essential to realize that the transition is not taking place at a fixed
Reynolds number. Flow changes over from laminar to turbulent over a range
of Reynolds numbers—from lower critical to upper critical Reynolds numbers.
The force on the body along the flow direction is called drag. The drag is
essentially a force opposing the motion of the body. Viscosity is responsible for
a part of the drag force, and the body shape generally determines the overall
drag. In the design of transport vehicles, shapes experiencing minimum drag
are considered to keep the power consumption at a minimum. Low drag
shapes are called streamlined bodies and high drag shapes are termed bluff
bodies. Drag arises due to (a) the difference in pressure between the front and
back regions and (b) the friction between the body surface and the fluid. The
drag force at (a) above is known as pressure drag and that at (b) above is
known as skin friction drag or shear drag.

ISTUDY
142 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

Turbulence may be described as a three-dimensional, random phenomenon,


exhibiting multiplicity of scales, possessing vorticity and showing very high dis-
sipation. The turbulence level for any given flow with a mean velocity U is
expressed as the turbulence number n, which is defined as

u′2 + v ′2 + w′2
n = 100
3U
Laminar flow may be described as a well orderly pattern where the fluid
layers are assumed to slide over one another, i.e. in laminar flow the fluid
moves in layers, or laminas, one layer sliding over an adjacent layer with only
a molecular interchange of momentum. Any tendency towards instability and
turbulence is damped out by viscous shear forces that resist relative motion
of adjacent fluid layers. In other words, laminar flow is an orderly flow in
which the fluid elements move in an orderly manner such that the transverse
exchange of momentum is insignificant. On the other hand, the turbulent
flow is a three–dimensional random phenomenon, exhibiting multiplicity of
scales, possessing vorticity, and showing very high dissipation. Turbulent flow
is basically an irregular flow. Turbulent flow has an erratic motion of fluid
particles, with a violent transverse exchange of momentum.
Gas dynamics is the science of fluid flow, where both density and tempera-
ture changes become significant.!The essence of gas dynamics is that the entire
flow field is dominated by Mach waves, expansion waves and shock waves,
when the flow speed is supersonic. Change of flow properties from one state
to another, in a supersonic flow field takes place only through these waves.
Mach number M is a dimensionless velocity, expressed as the ratio between
the magnitudes of local flow velocity and local speed of sound, i.e.
local flow velocity V
M≡ =
local speed of sound a
For isentropic flow of a perfect gas through a duct, the Area–Mach number
relation may be expressed, assuming one-dimensional flow, as

( )2 [ ( )] γ + 1
A 1 2 γ−1 2 γ−1
= 2 1+ M
A∗ M γ+1 2
The Prandtl–Meyer function ν is an important parameter to solve super-
sonic flow problems involving isentropic expansion or isentropic compression.
Basically the Prandtl–Meyer function is a similarity parameter. The Prandtl–
Meyer function can be expressed in terms of M as
√ √
γ+1 γ−1
ν= arc tan (M 2 − 1) − arc tan M 2 − 1
γ−1 γ+1
The shock may be described as a compression front in a supersonic flow
field and flow process, across which results an abrupt change in fluid proper-
ties. When the shock is normal to the flow direction it is called normal shock

ISTUDY
2.15. PROBLEMS 143

and when it is inclined at an angle to the flow it is termed oblique shock. The
changes of flow properties across the shock take place over a very short dis-
tance, of the order of 10−5 cm. Hence, the velocity and temperature gradients
inside the shock structure are very large. These large gradients result in in-
crease of entropy across the shock. Also, these gradients internal to the shock
provide heat conduction and viscous dissipation that render the shock process
internally irreversible. Flow through a shock is adiabatic but irreversible.
For Fanno line flow, the wall friction (due to viscosity) is the chief factor
bringing about changes in flow properties.
Flow involving only the T0 -change is called Rayleigh type flow.

2.15 PROBLEMS
2.1 Air flows through a tube, as shown in Figure P2.1, with a velocity of
0.2 m/s. The temperature gradient in the direction of flow is 2 ◦ C/m. The
temperature of each particle increases at a rate of 0.5 ◦ C/s due to absorption of
thermal radiation. Find the rate of change of the air temperature as recorded
by a stationary temperature probe. Also, find whether the temperature field
is steady.

FIGURE P2.1

[Ans. 0.1 ◦ C/s]


2.2 Flow through a convergent nozzle shown in Figure P2.2 is approximated
as one-dimensional. If the flow is steady, will there be any fluid acceleration?
If there is an acceleration, obtain the expression for it in terms of volumetric
flow rate Q̇ if the area of cross-section is given by A(x) = e−x .

FIGURE P2.2
 ( )2 
Ans. Q̇ 
e−x

ISTUDY
144 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.3 Atmospheric air is cooled by a desert cooler by 18◦ C and sent into a
room. The cooled air then flows through the room and picks up heat from
the room at a rate of 0.15◦ C/s. The air speed in the room is 0.72 m/s. After
some time from switching on, the temperature gradient assumes a value of
0.9◦ C/m in the room. Determine ∂T /∂t at a point 3 m away from the cooler.
[Ans. − 0.498 ◦ C/s]
2.4 For proper functioning, an electronic instrument on board a balloon
should not experience a temperature change of more than ± 0.006 K/s. The
atmospheric temperature is given by
( )( )
T = 288 − 6.5 × 10−3 z 2 − e−0.02t K

where z is the height in metres above the ground and t is the time in hours
after sunrise. Determine the maximum allowable rate of ascent when the
balloon is at the ground at t = 2 h.
[Ans. 1.12 m/s]
2.5 Flow through a tube has a velocity given by
( )
r2
u = umax 1 − 2
R
where R is the tube radius, and umax is the maximum velocity which occurs
at the tube centreline. (a) Find a general expression for the volume flow rate
and that for the average velocity through the tube; (b) compute the volume
flow rate if R = 25 mm and umax = 10 m/s; and (c) compute the mass flow
rate if ρ = 1000 kg/m3 .
1 1
[Ans. (a) umax πR2 , umax ; (b) 0.00982 m3 /s; (c) 9.82 kg/s]
2 2
2.6 A two-dimensional velocity field is given by
( )
V = x − y 2 i + (xy + 2y) j

in arbitrary units. At x = 2 and y = 1, compute (a) the acceleration


components ax and ay , (b) the velocity component in the direction θ = 30◦ ,
and (c) the magnitudes and directions of maximum velocity and maximum
acceleration.
[Ans. (a) –7 m/s2 , 17 m/s2 ; (b) 2.87 units;
(c) V = 4.123 m/s at 75.96◦ from x-axis,
a = 18.385 m/s2 at 292.38◦ from x-axis]
2.7 The velocity at a fixed point in a flow field is given by i + 2tj. Fluid
particles passing through that point continue to move with the velocity they
had at that point. Find (a) the pathline passing through the point at t = 0,
and (b) the pathline of the particle passing through the point at t = 1.
[Ans. (a) At t = 0, V = i, i.e. the pathline is along the x-direction,
(b) At t = 1, V = i + 2j, i.e. the pathline is
at an angle θ = tan−1 (2) to the x-direction]

ISTUDY
2.15. PROBLEMS 145

2.8 The velocity field of a flow is given by

V (x, y, z, t) = 10 x2 i − 20 yx j + 100 t k

Determine the velocity and acceleration of a particle at x = 1 m, y = 2 m,


z = 5 m, and t = 0.1 s.
[Ans. (10 i − 40 j + 10 k) m/s, (200 i + 400 j + 100 k) m/s2 ]
2.9 Oxygen atoms O enter a chamber shown in Figure P2.9 and exit as
O2 . The pressure and temperature of the system at the inlet and exit are the
same. The inlet port diameter is 10 mm and that of the exit is 20 mm. If the
inlet velocity is 12 m/s, obtain the steady-state velocity at the exit assuming
that all O atoms get transfered to O2 . Also, assume that both O and O2 obey
the ideal gas state equation.

FIGURE P2.9

[Ans. 1.5 m/s]


2.10 A rotating device to sprinkle water is shown in Figure P2.10. Water
enters the rotating device at the centre at a rate of 0.03 m3 /s and is then
directed radially through three identical channels of exit area 0.005 m2 each,
perpendicular to the direction of flow relative to the device. The water leaves
at an angle of 30◦ relative to the device as measured from the radial direction,
as shown in the figure. If the device rotates clockwise with a speed of 20 rad/s
relative to the ground, compute the magnitude of the average velocity of the
fluid leaving the vane as seen from the ground.

FIGURE P2.10

[Ans. 9.16 m/s, at an angle of 79◦ with respect to the ground (horizontal)]

ISTUDY
146 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.11 A tank, placed on an elevator as shown in Figure P2.11, starts moving


upwards at time t = 0 with a constant acceleration a. A stationary hose
discharges water into the tank at a constant velocity V1 . Determine the time
required to fill the tank if it is empty at t = 0.

FIGURE P2.11
 √ 
A2
 −V1 ± V12 + 2 aH 
Ans. A1 
 a 

2.12 Show that the general continuity equation


∂ρ
+ ▽. (ρV ) = 0
∂t
can be written in the equivalent form

+ ρ (▽.V ) = 0
dt

2.13 Develop the differential form of the continuity equation for the cylin-
drical polar coordinates shown by taking an infinitesimal control volume as
depicted in Figure P2.13. [ ]
∂ρ 1 ∂(ρrVr ) 1 ∂(ρVθ ) ∂(ρVz )
Ans. + + + =0
∂t r ∂r r ∂θ ∂z
2.14 A flow field is given by V = (3x i + 4y j − 5t k) m/s. (a) Find the
velocity at position (10, 6) at t = 3 s. (b) What is the slope of the streamlines
for this flow at t = 0 s? (c) Determine the equation of the streamlines at t = 0
up to an arbitrary constant. (d) Sketch the streamlines at t = 0.
4y
[Ans. (a) V = (30 i + 24 j − 15 k) m/s, (b) , (c) ln y = 43 ln x + 4 ln c, where
3x
c is an arbitrary constant, (d) At t = 0, the streamlines are straight lines at
an angle of 38.66◦ to the x-axis]
2.15 For the fully-developed two-dimensional flow of water between two
impervious flat plates as shown in Figure P2.15, show that Vy = 0 everywhere.

ISTUDY
2.15. PROBLEMS 147

FIGURE P2.13

FIGURE P2.15

2.16 Obtain the variation of Vr with r for the radial flow in-between two
large discs, as shown in Figure P2.16. Assume the flow to be incompressible.

FIGURE P2.16
[ ]
f (z)
Ans.Vr = , where f (z) is a constant which is a function of z alone .
r
2.17 Obtain the horizontal force acting at the flange AA of the elbow-nozzle
assembly, shown in Figure P2.17. The water issues out as a free jet from the
nozzle. The volume flow rate Q̇ = 1.8 m3 /s.
[Ans. 1.45 × 105 N, acting in the negative x-direction.]

ISTUDY
148 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

FIGURE P2.17

2.18 Water enters section 1 at 200 N/s and exits at 30◦ angle at
section 2, as shown
( in )
Figure P2.18. Section 1 has a laminar velocity pro-
r2
file, u = um1 1 − 2 , while the section 2 has a turbulent profile u =
R
( r )1/7
um2 1 − . If the flow is steady and incompressible(water), find the
R
maximum velocities um1 and um2 in m/s. Assume uav = 0.5 um , for laminar
flow, and uav = 0.82 um , for turbulent flow.

FIGURE P2.18

[Ans. 5.2 m/s, 8.79 m/s]


2.19 A water jet impinges on a stationary plane as shown in Figure P2.19.
Assuming that, V1 = V2 = V3 and the force exerted on the plate acts normal
to the plate, determine the force. Also, determine the volume flow rates Q̇2
and Q̇3 .
[Ans. 61.2 N, 0.0106 m3 /s, 0.00353 m3 /s]
2.20 A tank has a circular re-entrant outlet near its bottom as shown in
Figure P2.20. Such an outlet is called Borda’s mouthpiece. Water issues from
this outlet as a jet with a uniform velocity of 5 m/s. Assuming that the water
jet fills the outlet tube completely, find the area of the jet as a fraction of
the mouthpiece entrance area. Assume that the velocity along the walls is
negligible.
[Ans. 0.39]

ISTUDY
2.15. PROBLEMS 149

FIGURE P2.19

FIGURE P2.20

2.21 A tank with a re-entrant orifice called the Borda’s mouthpiece, is shown
in Figure P2.21. The re-entrant orifice essentially eliminates flows along the
tank walls, so the pressure along the walls is nearly hydrostatic. Water flows
out from the mouth as a free jet with a uniform velocity of Vj m/s. The water
jet is surrounded by atmospheric air. Calculate the contraction coefficient,
Ci = Aj /Ao . Treat the water to be inviscid.

FIGURE P2.21
1
[Ans. ]
2

ISTUDY
150 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.22 Consider a jet of fluid directed at the inclined plate shown in Figure P2.22.
Determine the force necessary to hold the plate in equilibrium against the jet
pressure. Also, determine the volume flow rates Q̇1 and Q̇2 in terms of the
incoming flow rate Q̇0 . Assume that, V0 = V1 = V2 and also that the fluid is
inviscid.

FIGURE P2.22

Q̇0 Q̇0
[Ans. ρV0 Q̇0 sin α, Q̇1 = (1 + cos α), Q̇2 = (1 − cos α)]
2 2
2.23 A vertical plate has a sharp-edged orifice at its centre. A water jet
of speed V strikes the plate concentrically as shown in Figure P2.23. Obtain
an expression for the external force required to hold the plate in place, if the
jet leaving the orifice also has speed V . Estimate the force for V = 3 m/s,
D = 80 mm, and d = 22 mm.

FIGURE P2.23
π
[Ans. ρV 2 (d2 − D2 ), − 41.82 N. The negative sign indicates that the exter-
4
nal force required to hold the plate in place should be applied from right to
left. (Note that the direction is known from common sense.)]

ISTUDY
2.15. PROBLEMS 151

2.24 Experimental measurements are made in a low-speed wind tunnel


(a device to produce a uniform wind stream of desired velocity over a sec-
tion termed test-section) to determine the drag force on a circular cylinder.
Velocity measurements at two sections, where the pressure is uniform and
equal, give the results as shown in Figure P2.24. Obtain the drag coefficient
for the cylinder in the following form. Neglect the viscous force at the inlet
and exit cross-sections of the control volume.
∫ m
force on the cylinder in the flow direction
CD = 1 2 lD
= 4 u∗ (1 − u∗ )dy ∗
2 ρU 0

where l = length of the cylinder, u∗ = u/U and y ∗ = y/D. Assume that the
fluid bleeds through the sides of control volume.

FIGURE P2.24

2.25 A jet of water is directed against a vane as shown in Figure P2.25.


The water leaves the stationary nozzle of 30 mm diameter, with a speed of
25 m/s and reaches the vane tangentially at A. The vane at B makes an angle
θ = 150◦ with the x-direction. Calculate the force (Rx and Ry ) that must be
applied to maintain the vane speed constant at U = 3 m/s. Assume friction
at the wheels to be zero.

FIGURE P2.25

[Ans. Rx = − 638.4 N, Ry = 171.06 N]

ISTUDY
152 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.26 Consider a fully developed laminar flow without body forces through a
long straight pipe of circular cross-section (Poiseuille flow) shown in
Figure P2.26. Apply the momentum equation and show that
p1 − p2 r
τrz =
l 2
Assuming (p1 − p2 )/l = constant, obtain the velocity profile using the relation
( )
dVz
τrz = − µ
dr

FIGURE P2.26
( )
p1 − p2 1 ( 2 )
[Ans. Vz = R − r2 ]
l 4µ
2.27 Water flows through a circular pipe as shown in Figure P2.27. It enters
at A and leaves at C and D. If the velocity at B is 0.8 m/s, and the velocity
at C is 2 m/s, calculate the velocities at A and D, and the volumetric flow
rate. Assume the flow to be inviscid.

FIGURE P2.27

[Ans. VA = 1.42 m/s; VD = 13.33 m/s; 0.1004 m3 /s]


2.28 Water from a large tank flows through an orifice as shown in
Figure P2.28. If p1 = patm , compute the velocity of the jet V when h = 5 m.
[Ans. 9.9 m/s]

ISTUDY
2.15. PROBLEMS 153

FIGURE P2.28

2.29 In the flow system shown in Figure P2.29, a high-speed water jet issuing
from a pipe of area Aj with velocity Vj and the surrounding water flow with
velocity V1 get mixed and the velocity becomes V2 at section 2. Assuming
the velocity profiles to be one-dimensional at sections 1 and 2, neglecting the
viscosity, and assuming the pressure to be uniform across the section 1, and
using the momentum equation show that
( )
Aj Aj 2
p2 − p1 = ρ 1− (V1 − Vj )
Ap Ap

FIGURE P2.29

2.30 A liquid of density ρ and viscosity µ flows down a stationary wall, under
the influence of gravity, forming a thin film of constant thickness h, as shown
in Figure P2.30. An upflow of air next to the film exerts an upward constant
shear stress τ on the surface of the liquid layer, as shown in the figure. The
pressure in the film is uniform. Derive expressions for (a) the film velocity Vy
as a function of y, ρ, µ, h and τ , and (b) the shear stress τ that would result
in a zero net volume flow rate in the film. ( )
x2
ρg hx − − τx
2 2
[Ans. (a) Vy = , (b) τ = ρgh]
µ 3

ISTUDY
154 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

FIGURE P2.30

2.31 For fully developed steady laminar flow of a fluid between two infi-
nite, parallel and stationary flat plates, as shown in Figure P2.31, determine
(a) the velocity distribution Vx (y) across the channel between the two plates,
(b) the maximum and average velocities Vx,max and Vx,av , and (c) the shearing
stress at the wall of the plate and the local frictional coefficient Cf .

FIGURE P2.31

1 dp ( 2 ) h2 dp
[Ans. (a) Vx (y) = y − h2 , (b) Vx,max = − ,
2µ dx 2µ dx
2 Vx,max 6µ ]
Vx,av = Vx,max , (c) (τyx )h = 2µ , Cf =
3 h h Vx,av
2.32 A two-dimensional fluid motion takes the form of concentric horizontal
circular streamlines. Show that the radial pressure gradient is given by

dp ρV 2
= θ
dr r
where ρ is the density, Vθ is the tangential velocity, and r is the radius.
Evaluate the pressure gradient for such a flow, defined by ψ = 2 ln r, where ψ
is the stream function, at a radius of 2 m. The fluid density is 103 kg/m3 .
dp
[Ans. = 500 N/m3 ]
dr

ISTUDY
2.15. PROBLEMS 155

2.33 The velocity V of a steady air flow above a horizontal surface (z = 0)


is given by ( )
V = az − bz 2 i + czj
where the constants a, b and c have the numerical values of a = 1.0 s−1 ,
b = 0.1 m−1 s−1 and c = 2.0 s−1 . Calculate the numerical values of all the
viscous stress components at z = 1 m if µ = 1.82 × 10−5 Pa·s.
[Ans. σxx = 0, σyy = 0, τzx = 1.456×10−5 Pa, τyz = τzy = 3.64×10−5 Pa]

2.34 For the flow of Problem 2.34, calculate the value of ▽p.
[Ans. 0.364 × 10−5 Pa/m]
2.35 For a plane Couette flow of a viscous fluid between two parallel plates
separated by a gap h, shown in Figure P2.35, with the upper plate moving
with a velocity V = Vp , show that the velocity profile is linear with respect
to vertical distance y.

FIGURE P2.35
(y)
[Ans. u = Vp ]
h
2.36 An incompressible flow of a viscous fluid flows parallel to a wide inclined
plate in the downhill direction, as shown in Figure P2.36. Treating the flow
to be laminar and fully developed, show that the pressure within the fully
developed region is a function of y alone. Also obtain the velocity profile and
the volumetric flow rate per unit distance normal to the plane of the flow, i.e.
Q̇/w.

FIGURE P2.36
( )
g sin θ y2 gh3 sin θ
[Ans. Vx = hy − , ]
ν 2 3ν

ISTUDY
156 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.37 A tank weighs 150 N and contains 0.33 m3 of water. A force of 220 N
acts on the tank. What is value of θ when the free surface of the water assumes
a fixed orientation relative to the tank, as shown in Figure P2.37.

FIGURE P2.37

[Ans. − 3.71◦ ]
2.38 Consider the pressure-driven flow between the stationary parallel plates
separated by distance
[ 2h,] as shown in Figure P2.38. The velocity field is given
by u = umax 1 − (y/h)2 , where y is the transverse direction. Evaluate the
rates of linear and angular deformation. Obtain an expression for the vorticity
vector ζ. Also, find the location where the vorticity is maximum.

FIGURE P2.38

[Ans. The rate of linear deformation is zero, the rate of angular deformation
umax 2y umax
in the xy-plane is − 2y 2 , ζ = k, the vorticity is maximum at
h h2
y = ±h]
2.39 Show that the head loss for laminar, fully developed flow in a straight
circular pipe is given by
2
64 L Vav
hl =
Re D 2g
where Re is the Reynolds number defined as ρVav D/µ.
2.40 The pipe shown in Figure P2.40 contains glycerine at 20◦ C flowing at
a rate of 8 m3 /h. Verify that the flow is laminar. Identify whether the flow is
from right to left or left to right. For glycerine at 20◦ C, µ = 1.49 kg/m·s and
ρ = 1260 kg/m3 .
[Ans. Flow is from right to left, since headB > headA ]

ISTUDY
2.15. PROBLEMS 157

FIGURE P2.40

2.41 The flow in the pipe shown in Figure P2.41 is driven by pressurized
air in the tank. What gauge pressure p1 is needed to provide a flow rate,
Q̇ = 50 m3 /h? Take ρ = 998 kg/m3 and µ = 0.001 kg/m·s.

FIGURE P2.41

[Ans. 1890 kPa]


2.42 A jar with a small circular outlet, containing oil on top of water, is
immersed in a pool of oil, as shown in Figure P2.42. Find the velocity of oil
flow at time t = 0? What would be the position of the oil–water interface in
the jar when the flow stops?. The density of the oil is 800 kg/m3 .
[Ans. 6.26 m/s, 8 m]
2.43 Determine the net vertical force acting on the circular plate shown in
Figure P2.43, if the water spreads radially on it. Neglect the weight of the
water on the plate.
[Ans. 7.927 × 104 N]
2.44 The water jet shown in Figure P2.44 strikes normal to a fixed plate.
Neglecting gravity and friction, compute the force F required to hold the plate
fixed.
[Ans. 502.64 N]

ISTUDY
158 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

FIGURE P2.42

FIGURE P2.43

FIGURE P2.44

2.45 The propulsion device of a boat consists of a pump that takes in water
from the river (inlet pipe area A1 ) and forces it out as a jet of area Aj , as
shown in Figure P2.45. The boat experiences a drag force F when it moves
at a speed of V . Show that the mass flow rate ṁ of water through the pump

ISTUDY
2.15. PROBLEMS 159

is given by √
ρ V Aj ρ2 V 2 A2j
ṁ = + + ρ F Aj
2 4
Also, obtain the head developed by the pump. Neglect the hydrostatic pres-
sure forces and viscous losses.

FIGURE P2.45
( )
V 2 A2j − A21
[Ans. ]
2g A2j
2.46 Water exits to the atmosphere (pa = 101 kPa) through a split nozzle as
shown in Figure P2.46. Duct areas are A1 = 0.01 m2 and A2 = A3 = 0.005 m2 .
The flow rate Q̇2 = Q̇3 = 150 m3 /h, and the inlet pressure p1 = 140 kPa.
Compute the force on the flange bolts at section 1.

FIGURE P2.46

[Ans. 411.24 N]
2.47 The fan fixed at the end of a duct (shown in Fig. P2.47) sucks air from
atmosphere. If the volume displacement of the fan is 1 m3 /s, find the power
required to run the fan. What is the maximum length h through which water
will be sucked up a tube by the flowing air? What is the force required to
hold the fan in place?
[Ans. 61 W, 0.62 cm, 12.25 N]
2.48 For water flow up the sloping channel shown in Figure P2.48, h1 =
10 mm, V1 = 3 m/s and head H = 600 mm. Neglecting losses and assuming
uniform flow at 1 and 2, find the downstream depth h2 and show that three
solutions are possible, of which only two are realistic.

ISTUDY
160 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

FIGURE P2.47

FIGURE P2.48

[Ans. 0.459 m]

2.49 Gasoline with density ρ = 680 kg/m3 flows from a 30 cm diameter


pipe in which the pressure is 300 kPa into a 15 cm diameter pipe in which
the pressure is 120 kPa. If the pipes are horizontal and the viscous effects are
negligible, determine the flow rate.
[Ans. 0.420 m3 /s]

2.50 A hand glider soars through standard sea level air with an air speed of
22 m/s. What is the gauge pressure at a stagnation point on the structure?
[Ans. 296.45 Pa]

2.51 Water flows into a large tank at a rate of 0.011 m3 /s, as shown in
Figure P2.51. The water leaves the tank through 20 holes at the bottom of
the tank, each of which produces a stream of 10 mm diameter. Determine the
equilibrium height h, for steady-state flow.
[Ans. 2.5 m]

2.52 Compare the column heights of water and mercury corresponding to a


pressure of 100 kPa. Express your answer in metres.
[Ans. 10.194 m of water, 0.750 m of mercury]

ISTUDY
2.15. PROBLEMS 161

FIGURE P2.51

2.53 During a study of a certain flow system the following equation relating
the pressures p1 and p2 at two points was developed.

f lV
p2 = p1 +
Dg

where V is the velocity, l is the distance between the two points, D is the
diameter, g is the gravitational acceleration, and f is a dimensionless coeffi-
cient. Is the equation dimensionally consistent?
[Ans. No]
2.54 A large airship of volume 90,000 m3 contains helium under standard
atmospheric conditions. Determine the density and total weight of the helium.
[Ans. 0.1686 kg/m3 , 1.4886 × 105 N]
2.55 A large movable plate is located between two fixed plates, as shown in
Figure P2.55. Two newtonian fluids having viscosities indicated in the figure
are contained between the plates. Determine the magnitude and direction of
the shearing stresses that act on the fixed walls when the moving plate has
a velocity of 4 m/s, as shown. Assume the velocity distribution between the
plates to be linear.
[Ans. 13.33 N/m2 on the upper plate, 13.33 N/m2 on the lower plate,
both the stresses act in the direction of the moving plate.]
2.56 The velocity in a certain flow field is given by the equation

V = 3yz 2 i + xz j + y k

Determinetheexpressionsfor the three rectangular components of acceleration.


[Ans. ax = 3xz 3 + 6y 2 z, ay = 3yz 3 + xy, az = xz]

ISTUDY
162 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

FIGURE P2.55

2.57 Determine an expression for the vorticity of the flow field described by

V = x2 y i − xy 2 j
Is the flow irrotational? ( )
[Ans. ζ = ζz k = − x2 + y 2 k. The flow is not irrotational, since the
vorticity is not zero]
2.58 For a certain incompressible, two-dimensional flow field the velocity
component in the y-direction is given by the equation

v = x2 + 2xy

Determine the velocity component in the x-direction so that the continuity


equation is satisfied.
[Ans. u = − x2 + f (y)]
2.59 The velocity potential for a certain flow field is ϕ = 4xy. Determine
the corresponding stream function. ( )
[Ans. ψ = 2 y 2 − x2 + constant]
2.60 In a certain steady, incompressible, inviscid, two-dimensional flow the
x-component of velocity is given by

u = x2 − y

Will the corresponding pressure gradient in the horizontal direction(x-direction)


be a function of y only, x only, or of both x and y?
[Ans. The pressure gradient in the x-direction is a function of x only]
2.61 In a certain viscous, incompressible, steady flow field without body
forces the velocity components are

u = ay − b(cy − y 2 )
v=w=0

ISTUDY
2.15. PROBLEMS 163

where a, b, and c are constants. (a) Using the Navier–Stokes equations, find
an expression for the pressure gradient in the x-direction. (b) For what com-
bination of the constants a, b, and c will the shearing stress, τyx , be zero at
y = 0 where the velocity is zero?
∂p
[Ans. (a) = 2 µb, (b) a = bc]
∂x
2.62 Consider steady, laminar flow of an incompressible fluid through
the horizontal rectangular channel shown in Figure P2.62. Assume that the
velocity components in the x and y directions are zero and the only body
force is the weight. Starting from the Navier–Stokes equations, (a) determine
the appropriate set of differential equations and boundary conditions for this
problem and (b) show that the pressure distribution is hydrostatic at any
given cross-section.

FIGURE P2.62
( 2 )
∂p ∂p ∂p ∂ w ∂2w
[Ans. (a) = 0, = − ρg, = µ + . The boundary
∂x ∂y ∂z ∂x2 ∂y 2
b a
conditions are: at x = ± , w = 0 and at y = ± , w = 0. (b) p = ρgy]
2 2
2.63 An airfoil of chord 2 m is tested in an air stream of velocity 50 m/s
at sea level. (a) Determine the Reynolds number. (b) If the same airfoil were
attached to an airplane flying at the same speed in a standard atmosphere at
an altitude of 3000 m, what would be the Reynolds number?
[Ans. (a) 6.85 × 106 , (b) 5.37 × 106 ]
2.64 An incompressible fluid between two large parallel plates, shown in
Figure P2.64, is set to motion by suddenly moving the bottom plate at a
constant speed U . The governing differential equation describing the fluid
motion is
∂u ∂2u
ρ =µ 2
∂t ∂y
where u is the velocity in the x-direction, and ρ and µ are the fluid density and
viscosity, respectively. Rewrite the equation and the initial and the boundary
conditions in dimensionless form using h and U as reference parameters for
length and velocity, and h2 ρ/µ as a reference parameter for time.
[Ans. The initial condition is: at y ∗ = 0, u∗ = 0. The boundary conditions
are: at y ∗ = 0, u∗ = 1 and at y ∗ = 1, u∗ = 0.]

ISTUDY
164 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

FIGURE P2.64

2.65 A fluid flows through a pipe of radius R with a Reynolds number of


100,000. (a) At what radial location, r/R, does the fluid velocity become
equal to the average velocity? (b) Repeat if the Reynolds number is 1000.
[Ans. (a) 0.758, (b) 0.707]
2.66 Determine the pressure drop per 300 m length of 0.20 m diameter
horizontal cast iron pipe when the average velocity is 1.7 m/s. The equivalent
roughness of the cast iron pipe is 0.26 mm. Assume the viscosity of water to
be 1.12 × 10−3 kg/m · s.
[Ans. 4.86 m]
2.67 Water at 80◦ C flows through a 120 mm diameter pipe with an average
velocity of 2 m/s. If the pipe wall roughness is small enough so that it does
not protrude through the laminar sublayer, the pipe can be considered as
smooth. Approximately, what is the best roughness allowed to classify the
pipe as smooth?
[Ans. 2.28 × 10−5 m]
2.68 Determine the lift and drag coefficients(based on frontal area) for the
triangular two-dimensional object, shown in Figure P2.68. Assume that the
skin friction is negligible.

FIGURE P2.68

[Ans. CL = 0, CD = 1.495]

ISTUDY
2.15. PROBLEMS 165

2.69 A horizontal pipe of length L and diameter D conveys air. Assuming


the air to expand according to the law p/ρ = constant and that acceleration
effects are small, prove that

p1 ( 2 ) 16f Lṁ2
ρ1 − ρ22 =
ρ1 π 2 D5

where ṁ is the mass flow rate of air through the pipe, f is the average fric-
tion coefficient, and 1 and 2 are the inlet and discharge ends of the pipe,
respectively.
2.70 A sink of strength 20 m2 /s is situated 3 m upstream of a source of
strength 40 m2 /s, in a uniform horizontal stream. If it is found that at a
point equidistant from both source and sink and 2.5 m above the line joining
the source and sink, the local velocity is normal to the line joining the source
and sink, find the velocity at this point and the velocity of the undisturbed
stream.
[Ans. 0.9373 m/s, 2.813 m/s]
2.71 A jet of water issues from a nozzle with a velocity of 50 m/s. If the
flow rate is 0.22 m3 /s, what is the power of the jet?
[Ans. 275 kW]
2.72 In the boundary layer over the upper surface of an aeroplane wing, at
a point A near the leading edge, the flow velocity just outside the boundary
layer is 250 km/h. At another point B, which is downstream of A, the velocity
outside the boundary layer is 470 km/h. If the temperature at A is 288 K,
calculate the temperature and Mach number at point B.
[Ans. 281.9 K, 0.388]
2.73 A pump discharges 2 m3 /s of water through a pipeline. If the pressure
difference between the inlet and the outlet of the pump is equivalent to 10 m
of water, how much power will be transmitted to the water from the pump?
[Ans. 196 kW]

2.74 A free vortex flow field is given by v = , for r > 0. If the flow
2πr
density ρ = 10 kg/m and the volume flow rate q̇ = 20π m2 /s, express
3 3

the radial pressure gradient, ∂p/∂r, as a function of radial distance r, and


determine the pressure change between r1 = 1 m and r2 = 2 m.
102 ρ
[Ans. , 37.5 kPa]
r3
2.75 For a forced vortex of angular velocity ω = 10 rad/s, assuming the fluid
to be frictionless with density ρ = 1000 kg/m3 , express the radial pressure
gradient ∂p/∂r, as a function of r. Evaluate the pressure change between
radii r1 = 1 m and r2 = 2 m.
∂p
[Ans. = ρω 2 r, 150 kPa]
∂r

ISTUDY
166 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.76 A viscous incompressible fluid is in a two-dimensional motion in circles


about the origin with tangential velocity
( 2)
1 r
uθ = f
r νt
where ν is the kinematic velocity and t is the time. Find the vorticity (ζ. )
2 ′ r2
[Ans. ζ = f ]
νt νt
2.77 A circular cylinder of radius a is in an otherwise uniform stream of
inviscid fluid but with a positive circulation round the cylinder, as shown in
Figure P2.77. Find the lift and the drag per unit span of the cylinder. Also,
sketch the streamlines around the cylinder if the circulation is subcritical.

FIGURE P2.77

[Ans. lift = ρV∞ Γ, drag = 0]

2.78 A long right circular cylinder of diameter a m is set horizontally in a


steady stream of velocity u m/s and made to rotate at an angular velocity of
ω rad/s. Obtain an expression in terms of ω and u for the ratio of pressure
difference between the top and bottom of the cylinder to the dynamic pressure
of the stream.
−8aω
[Ans. ]
u
2.79 When a circulation of strength Γ is imposed on a circular cylinder placed
in a uniform incompressible flow of velocity U∞ , the cylinder experiences lift.
If the lift coefficient CL = 2, calculate the peak (negative) pressure coefficient
on the cylinder.
[Ans. − 4.375]
2.80 Considering the flow field that results if a vortex with clockwise circu-
lation is superposed on a combination of a doublet and uniform flow, deter-
mine the maximum lift coefficient that can be generated for a circulating flow
around a cylinder, neglecting the cases of strong circulations which will shift
the stagnation points outside the body.
[Ans. 4π]
2.81 A wing with an elliptical plan-form and an elliptical lift distribution
has an aspect ratio 6 and a span of 12 m. The wing loading is 900 N/m2 ,

ISTUDY
2.15. PROBLEMS 167

when flying at a speed of 150 km/h at sea level. Calculate the induced drag
for this wing.
[Ans. 968.3 N]
2.82 The velocity and temperature fields of a fluid are given by
( )
V = x i + 3y + 3t2 y j + 12 k
T = x + y 2 z + 5t

Find the rate of change of temperature recorded by a floating probe


(thermocouple) when it is at 3 i + 5 j + 2 k at time t = 2 units.
[Ans. 1808]
2.83 The working section of a wind tunnel of cross-sectional area 0.6 m2
has a Mach number of 0.80. In the settling chamber of cross-sectional area
4.0 m2 , the pressure, the density, and the temperature are 1.014 × 105 Pa,
1.144 kg/m3 , and 35◦ C, respectively. Calculate the pressure, the density, and
the temperature in the working section, neglecting the effects of viscosity and
treating the flow as one-dimensional.
[Ans. 0.665 × 105 Pa, 0.847 kg/m3 , 273 K]
2.84 Air from a storage tank exhausts through a convergent nozzle of exit
area 0.5 cm2 to an environment at 1 atm pressure. If the tank is at 500 kPa
and 30◦ C, compute the mass flow rate at the beginning of discharge.
[Ans. 0.058 kg/s]
2.85 Air at stagnation state at 700 kPa and 180◦ C enters a nozzle. The
pressure at the nozzle exit is 100 kPa. Flow expands isentropically through
the nozzle. Determine the flow velocity at the nozzle exit if the flow expands
isentropically. Assume the air to be a perfect gas with γ = 1.4.
[Ans. 623.4 m/s]
2.86 Calculate the maximum mass flow possible through a frictionless,
insulated convergent nozzle of exit area 6.5 cm2 operating at sea level, if the
stagnation conditions are 5 bar and 15◦ C. Also, calculate the exit tempera-
ture.
[Ans. 0.774 kg/s, 240.12 K]
2.87 Air at 101 kPa and 20◦ C is drawn isentropically through a convergent–
divergent nozzle of exit area 0.033 m2 . If the pressure at the nozzle exit is
91.4 kPa, determine the mass flow rate through the nozzle. What is the
pressure at the location with area 0.022 m2 ?
[Ans. 4.74 kg/s, 73.5 kPa]
2.88 A rocket nozzle has to generate 9 kN thrust at an altitude of 16 km
above the earth, with its chamber pressure and temperature of 15 atm and
2600◦ C, respectively. Calculate its exit and throat areas and the velocity and
temperature at the nozzle exit. Take γ = 1.4, R = 287 J/kg· K. Assume the
nozzle to operate at the adapted condition.
[Ans. 0.0394 m2 , 0.00374 m2 , 2094.6 m/s, 689.33 K]

ISTUDY
168 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.89 Determine the stagnation pressure p0 , the temperature T0 , and the


density ρ0 of an air stream at a speed of 200 m/s. The air temperature is
15◦ C and the pressure is 101 kPa.
[Ans. 127.8 kPa, 308.2 K, 1.445 kg/m3 ]
2.90 Air flows through a convergent duct. At a station 1 in the duct, A1 =
10 cm2 , p1 = 100 kPa, T1 = 30◦ C, and V1 = 90.5 m/s. Calculate M2 , p2 and
T2 at a station 2 where A2 = 6.9 cm2 . Assume the flow to be one-dimensional
and isentropic.
[Ans. 0.4, 93.86 kPa, 297.72 K]
2.91 An intermittent wind tunnel is operated by expanding atmospheric
air at 15◦ C through the test-section into an evacuated tank. Determine the
static pressure and the pressure that a pitot tube placed in the test-section
can measure, if the Mach number in the test-section is 3.0.
[Ans. 2758 Pa, 33.265 kPa]
2.92 Upstream of a normal shock in air, M1 = 2.5, p1 = 1 atm, ρ1 =
1.225 kg/m3 . Determine p2 , ρ2 , T2 , M2 , V2 , p02 and T02 downstream of it.
[Ans. 7.125 atm, 4.083 kg/m3 , 616.03 K, 0.51299,
255.22 m/s, 8.5261 atm, 648.45 K]
2.93 Nitrogen gas passes through a normal shock with upstream conditions
of p1 = 300 kPa, T1 = 303 K and V1 = 923 m/s. Determine the velocity V2
and the pressure p2 downstream of the shock. If the same deceleration from
V1 to V2 takes place isentropically, what will be the resultant p2 ?
[Ans. 2.316 MPa, 267.64 m/s, 5.036 MPa]
2.94 A blunt-nosed model is placed in a Mach 3 supersonic tunnel test-
section. If the settling chamber pressure and temperature of the tunnel are
10 atm and 315 K, respectively, calculate the pressure, temperature and den-
sity at the nose of the model. Assume the flow to be one-dimensional.
[Ans. 332.69 kPa, 315 K, 3.68 kg/m3 ]
2.95 A normal shock travels with velocity Cs in a still atmosphere at
101 kPa and 330 K. If the pressure just downstream of the shock is
5000 kPa, determine the velocity Cs and the velocity of the field traversed
by the shock, just downstream of it.
[Ans. 2374.15 m/s, 1932.19 m/s]
2.96 There is a normal shock in a uniform air stream. The properties
upstream of the shock are V1 = 412 m/s, p1 = 92 kPa, and T1 = 300 K.
Determine V2 , p2 , T2 , T02 and p02 downstream of the shock. Also calculate
the entropy increase across the shock.
[Ans. 311.98 m/s, 136.66 kPa, 336.51 K, 384.96 K, 218.82 kPa,
1.8173 J/kg·K]
2.97 A convergent–divergent nozzle of exit area 4.0 cm2 is to be designed
to generate Mach 2.5 air stream. The nozzle is correctly expanded and
discharges into atmosphere. The stagnation temperature at the entry is

ISTUDY
2.15. PROBLEMS 169

500 K. Determine the back pressure required to just produce a normal shock
at the nozzle exit plane.
[Ans. 7.125 atm]
2.98 Suppose the back pressure is increased for the nozzle in Problem 2.97
until a normal shock wave is formed in the divergent section where M = 1.5.
Find the back pressure necessary to accomplish this, and the resulting velocity
and temperature at the nozzle exit?
[Ans. 15.27 atm, 108.71 m/s, 490 K]
2.99 Air from a storage tank at 700 kPa and 530 K is expanded through
a frictionless convergent–divergent duct of throat area 5 cm2 and exit area
12.5 cm2 . The back pressure is 350 kPa. There is a normal shock in the
divergent portion and the Mach number just upstream of the shock is 2.32.
Determine (a) the cross-sectional area at the shock location, (b) the exit
Mach number, (c) the back pressure for the flow to be isentropic throughout
the duct.
[Ans. 11.15 cm2 , 0.45, ≤ 44.98 kPa]

2.100 A normal shock wave forms in an air stream at a static temperature


of 22 K, the total temperature being 400 K. Determine the Mach number and
the static temperature behind the shock.
[Ans. 0.3893, 382.8 K]
2.101 The flow properties upstream of a normal shock are 500 m/s, 100 kPa
and 300 K. Determine the velocity, the pressure and the temperature of the
gas downstream of the shock and also the increase in entropy. Take the gas
to be air.
[Ans. 284.25 m/s, 225.25 kPa, 384.21 K, 15.432 J/kg·K]
2.102 For a Prandtl–Meyer expansion, the upstream Mach number is 2 and
the pressure ratio across the fan is 0.5. Determine the angles of the front and
the end Mach lines of the expansion fan relative to the freestream.
[Ans. 30◦ , 12.86◦ ]
2.103 Calculate the ratios of static and total pressures across the shock
wave emanating from the leading edge of a wedge of 5◦ half-angle flying at
Mach 2.2.
[Ans. 1.3397, 0.99726]
2.104 A uniform supersonic flow of air at Mach 3.0 and p1 = 0.05 atm passes
over a cone of semi-vertex angle 8◦ kept in line with the flow. Determine the
shock angle and the static pressure at the cone surface, just behind the shock.
[Ans. 25.61◦ , 9.1 kPa]
2.105 A supersonic stream of air at Mach 3 and 1 atm passes through a
sudden convex corner and a sudden concave corner of turning angle 15◦ each.
Determine the Mach number and the pressure of the flow downstream of the
concave corner.
[Ans. 2.7, 1.019 atm]

ISTUDY
170 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

2.106 For an oblique shock wave with a wave angle of 33◦ and upstream
Mach number 2.4, calculate the flow deflection angle θ, the pressure and the
temperature ratios across the shock wave and the Mach number behind the
wave.
[Ans. 10◦ , 1.8354, 1.1972, 2.00]
2.107 Show that the pressure difference across an oblique shock wave with
wave angle β may be expressed in the form
( )
p2 − p1 4 1
1 = sin2
β −
2
2 ρ1 u1
γ+1 M12

where the subscripts 1 and 2 refer to states upstream and downstream of the
shock.
2.108 Air flow with Mach number 3.0 and pressure 1 atm passes over
a compression corner. If the pressure downstream of the corner is 5 atm,
determine the flow turning angle.
[Ans. 25.5◦ ]
2.109 A supersonic air stream at M = 2 passes over a 10◦ compression
corner. The oblique shock from the corner is reflected from a flat wall which
is parallel to the freestream, as shown in Figure P2.109. Compute the angle
of the reflected shock wave relative to the flat wall, and the Mach number
downstream of the reflected shock.

FIGURE P2.109

[Ans. 39.5◦ , 1.28]


2.110 Air at Mach 2 passes over two compression corners of angles 7◦ and θ,
as shown in Figure P2.110. Determine the value of θ up to which the second
shock will remain attached.

FIGURE P2.110

[Ans. 18◦ ]

ISTUDY
2.15. PROBLEMS 171

2.111 Air flow at Mach 2 is compressed by turning it through 15◦ . For each
of the possible solutions, calculate (a) the shock angle, (b) the Mach number
downstream of the shock, (c) the change in entropy. What is the maximum
deflection angle up to which the shock remains attached?
[Ans. Weak solution: (a) 45.34◦ , (b) 1.45, (c) 13.73 J/kg·K
Strong solution: (a) 79.83◦ , (b) 0.64, (c) 88.25 J/kg·K, 22.97◦ ]
2.112 A supersonic flow at Mach 3 with pressure and temperatures of
1 atm and 200 K, respectively, is deflected at a compression corner through
10◦ . Calculate the Mach number, the static and the stagnation pressures and
temperatures downstream of the corner.
[Ans. 2.5, 2.0551 atm, 248.36 K, 35.37 atm, 560 K]
2.113 Air flows adiabatically through a duct of diameter 20 mm. At a
station 1 in the duct, M1 = 0.2, p1 = 5 atm, and T1 = 300 K. Compute p2 ,
T2 , V2 , and p02 at a station 2 where M2 = 0.5.
[Ans. 198.558 kPa, 288 K, 170.09 m/s, 235.52 kPa]
2.114 Air flows through a perfectly insulated square tube of cross-section
0.1 m by 0.1 m. At a section 1 inside the tube, M1 = 0.2, T1 = 72◦ C, and
p1 = 2 atm. At a downstream section 2, M2 = 0.76. Determine the mass flow
rate through the tube and the drag force acting on the duct between sections
1 and 2.
[Ans. 1.524 kg/s, 1223.43 N]
2.115 Carbon dioxide gas enters an insulated circular tube of length-to-
diameter ratio 50. At the entrance, the flow velocity is 195 m/s and the
temperature is 310 K. If the flow at the tube exit is choked, determine the
average friction factor for the tube.
[Ans. 0.00105]
2.116 Air flows through a pipe of 25 mm diameter and 51 m length. The
conditions at the exit of the pipe are M2 = 0.8, p2 = 1 atm and T2 =
270 K. Assuming adiabatic one-dimensional flow, calculate M1 , p1 and T1
at the pipe entrance. Take the local friction coefficient to be 0.005.
[Ans. 0.13, 6.56 atm, 303.52 K]
2.117 A fuel-air mixture enters a duct combustion chamber at 80 m/s and
325 K. Estimate the amount of heat addition at the combustion chamber
required for this flow to exit as a choked flow. Assume the fuel-air mixture to
be equivalent to air.
[Ans. 1272.7 kJ/kg]
2.118 Determine the heat transfer required to decelerate an air stream at
Mach 3 (which corresponds to a velocity of 800 m/s) to a velocity of 700 m/s,
resulting in Mach 1.5.
[Ans. 291.65 kJ/kg]
2.119 Rayleigh flow of air is taking place through a rectangular tube of cross-
section 0.5 m by 0.2 m. At a section 1 in the tube, p1 = 9.36 kPa(gauge) and

ISTUDY
172 CHAPTER 2. FUNDAMENTALS OF FLUID MECHANICS

T1 = 300◦ C. At the tube exit, the speed is subsonic with pe = 1 atm. If


the mass flow rate is 10 kg/s, determine the heat transferred per kg of air, q,
between section 1 and tube exit. Assume γ = 1.4 and cp = 1004.5 J/kg·K.
[Ans. 291.2 kJ/kg]
2.120 Air at a stagnation state of 300 kPa and 350 K flows through a
convergent nozzle of exit diameter 15 mm, into a constant area frictionless
duct of the same diameter as that of the nozzle exit. Heat is transferred to
the duct at the rate of 175 kJ/kg of air. Determine the maximum mass flow
rate through the duct and the range of back pressure for the mass flow rate
to be maximum. Assume air to be a perfect gas.
[Ans. 0.096 kg/s]

ISTUDY
Chapter 3

Dimensional Analysis and


Similarity

3.1 INTRODUCTION
As we all know, all equations in engineering must be dimensionally homoge-
neous. That is, every term in an equation must have the same units. Also,
we know from experience that units can give a constant headache if they are
not used properly in solving problems. But when carefully used, they can
be of great advantage, e.g. they can be used to check formulae. The power
and the usefulness of dimensional analysis and its role in the experimental
investigation of fluid flow phenomena will be discussed in this chapter.

3.2 DIMENSIONLESS GROUPS


Essentially, the dimensionless groups make the results of a theoretical or
experimental investigation independent of any specific local conditions with
which the study may have been carried out. In other words, any result can
be made universal by using dimensionless groups to express them. Also, the
dimensional aspects of the fluid flow will serve as the basic tool to set forth
the fundamental considerations which are vital for experimental investigation.
We have already seen that it is possible to express a dependent dimen-
sion in terms of a chosen set of basic dimensions. Thus, velocity is given
dimensionally by
[ ]
L
V ≡
T
To give the dimensional representation of a product of quantities, we have
to carry out only the ordinary algebraic operations on the basic dimensions
appearing in the dimensional representation of the quantities.

173

ISTUDY
174 CHAPTER 3. DIMENSIONAL ANALYSIS AND SIMILARITY

EXAMPLE 3.1 Find the dimension of the product of velocity and time.

Solution The basic dimensions of velocity and time have to be multiplied


to get the dimension of their product, i.e.
[ ]
L
Vt≡ [T ] ≡ [L]
T
Dimensionally, therefore, the product of velocity and time is a distance.

EXAMPLE 3.2 A tank is filled with water at normal temperature. If the


volume of the tank V = 3 m3 , determine the amount of mass m in the tank.

Solution Suppose that we are interested in getting the formula that relates
mass to density and volume. From basic dimensions we know that mass has
the unit of kilogram, i.e. whatever calculations we do, it must result in the
unit of kilogram for mass. It is given that, ρ = 1000 kg/m3 and V = 3 m3 .
From these two quantities it is obvious that we can eliminate m3 and end
up with kg by simply multiplying the density and volume. Therefore, the
required formula is m = ρV. Thus,

3
m = (1000 kg/m )(3 m3 ) = 3000 kg
At this stage, we must keep in mind that a formula which is not dimensionally
homogeneous is definitely wrong, but a dimensionally homogeneous formula
is not necessarily correct.
It is clear from the above examples that the dimension of a group of
quantities has to be the resulting dimension obtained by performing algebraic
operations on the basic dimensions appearing in the dimensional representa-
tion of the individual quantities.
A group of quantities having a dimensional representation of unity when
multiplied together is called a dimensionless group. For example, the well-
known product ρV D/µ is a dimensionless group, since

ρV D [M/L3 ][L/T ][L]


= =1
µ [M ]/[LT ]
This dimensionless group is called the Reynolds number, which plays a
dominant role in fluid dynamics.

3.3 DIMENSIONAL HOMOGENEITY


PRINCIPLE
It is well known that the analytically derived equations are valid for any
system of units. Therefore, each group of terms in the equations must have
the same dimensions. This rule is called the principle of homogeneity.

ISTUDY
3.3. DIMENSIONAL HOMOGENEITY PRINCIPLE 175

This rule will be of great use in situations where the variables involved in a
physical phenomenon are known, while the relationship between the variables
is not known. For example, examine the problem of computing the drag force
F of a smooth sphere of diameter D immersed in a viscous, incompressible
fluid of density ρ, moving with velocity V . Let the viscosity of the fluid be µ.
The drag F may be expressed as some unknown function of the variables
involved in the flow process, i.e.
F = f (D, V, ρ, µ) (3.1)
To find this function experimentally would be a difficult task since, in every
experiment, the variables in parentheses can be allowed to vary only one at a
time. This will result in the accumulation of a large volume of data or many
charts. To illustrate the quantum of effort required for an effective description
of the process, let us plot F versus D for different values of V , as illustrated in
Figure 3.1. Each plot is for fixed values of ρ and µ. As seen from the figure, a
large number of plots are required for an effective description of the process.

FIGURE 3.1 Force variation plots.

For the sake of argument, let us say that about 10 experimental points are
required to define a curve. To find the effect of sphere diameter in Eq. (3.1), we
have to run the experiment for 10 diameters. For each D, we require 10 values
each of the other three variables in Eq. (3.1). This combination of parameters
results in a total of 10,000 experiments. Thus, it becomes an extremely ex-
pensive and time-consuming study.

ISTUDY
176 CHAPTER 3. DIMENSIONAL ANALYSIS AND SIMILARITY

Now, let us see the capability of dimensional analysis in solving such prob-
lems. As will be shown presently, with dimensional analysis, the problem
may be formulated as a functional relation between just two nondimensional
groups. Thus,
( ) ( )
F ρV D
= g (3.2)
ρV 2 D2 µ

where the nature of the function g is not known. Each group in Eq. (3.2)
is called a π(Pi). It must be realized that this is not the mathematical π
with value 3.1416...; it simply means the product of variables. Even though
the function g in Eq. (3.2) is not known, by experiment, a single curve may
be established by relating the πs, as shown in Figure 3.2. The single curve
in Figure 3.2 is equivalent to hundreds of curves shown in Figure 3.1. In
other words, the single curve in Figure 3.2 contains the complete quantitative
information of the hundreds of charts of the type presented in Figure 3.1.
For example, suppose that drag is desired for conditions V1 , D1 , ρ1 , µ1 . The
dimensionless group (π2 )1 can straightaway be evaluated as V1 D1 ρ1 /µ1 . The
corresponding (π1 )1 is read off from the chart in Figure 3.2. F1 is then
computed as (V12 D12 ρ1 )(π1 )1 .

FIGURE 3.2 Dimensionless force plot.

To establish such a useful chart, a wind or water tunnel may be used. For
a given sphere (i.e. given D), the values of π2 may be continually changed by
simply varying the freestream velocity, which is a simple and straightforward
control in wind or water tunnels.
The force on the sphere is measured for each velocity setting so that the
corresponding values of π1 may be ascertained. Thus, with less expenditure
and time, a curve of the dimensionless groups is established which, as a result

ISTUDY
3.4. BUCKINGHAM’S π-THEOREM 177

of dimensional analysis, is valid for any fluid or any diameter sphere within
the range of the πs tested.
Now, the task before us is to identify the number of dimensionless πs
to be formed for a group of variables known to be involved in a physical
phenomenon. There are several methods of reducing the dimensional variables
involved in a physical process into a smaller number of nondimensional groups.
The scheme given here is the Buckingham’s π-theorem.

3.4 BUCKINGHAM’S π-THEOREM


The number of independent dimensionless groups that may be employed to
describe a phenomenon known to involve n variables is equal to the number
n − r, where r is usually the number of basic dimensions needed to express
the variables dimensionally.
In the example of drag force on a sphere in a fluid stream, the variables in-
volved were F, V, D, µ, and ρ, making n = 5. There are three basic dimensions
M, L, T , which must be employed to express these variables dimensionally, so
that n − r becomes equal to 2.
It is evident from this example that the dimensionless groups employed
are independent, i.e. not related to each other through algebraic operations,
since F appears only in one group and µ appears only in the second group.
Buckingham’s π-theorem states that there can be no other independent group.

3.5 DIMENSIONLESS GROUPS IN


FLUID DYNAMICS
In most fluid flow phenomena where heat transfer can be neglected, the
following eight variables may be essential for the complete description of these
phenomena:
1. Pressure p
2. Length L
3. Viscosity µ
4. Surface tension σ
5. Speed of sound a
6. Acceleration due to gravity g
7. Density ρ
8. Velocity V
Three basic dimensions are needed to describe these eight variables. There-
fore, 5 (= 8 – 3) independent dimensionless groups can be formed with these
variables, in accordance with the Buckingham’s π-theorem. The dimension-
less groups formed out of these important variables of fluid phenomena are:

ISTUDY
178 CHAPTER 3. DIMENSIONAL ANALYSIS AND SIMILARITY

ρV L
1. Reynolds number (Re)L =
µ
V
2. Mach number M=
a
V
3. Froude number Fr = √
Lg
ρV 2 L
4. Weber number W=
σ
△p
5. Euler number Eu =
ρV 2
It is seen that these dimensionless groups are independent. In many engi-
neering problems, only a few of the variables given above are simultaneously
involved. For example, in aerospace applications, surface tension and grav-
ity are not important enough to warrant consideration, so the dimentionless
groups Fr and W would not be involved. Likewise, it can be shown that in
civil engineering applications, Fr plays a dominant role, and Mach number is
not of any significance. The dimensionless groups listed above are some of
the commonly encountered π numbers. Nevertheless, it is important to note
that there is a large number of dimensionless groups of importance in fluid
flow analysis that must be employed for approximate flow phenomena. For
instance, in low density flows, the π number, namely the Knudsen number,
plays a dominant role.

3.6 CALCULATION OF THE


DIMENSIONLESS GROUPS
The exact number of dimensionless groups associated with a problem can be
determined with Buckingham’s π-theorem. Now, let us see how the groupings
can be done. One way to establish the exact number of independent groups
is by the trial-and-error procedure. However, it will be cumbersome when
a large number of variables are involved in a problem. In such situations,
the following procedure will be effective: To illustrate the procedure, let us
dimensionally investigate the pressure drop in a viscous incompressible flow
through a straight pipe of diameter D and length L. The variables involved
in the problem are the pressure drop ∆p, average velocity V , viscosity µ, fluid
density ρ, pipe diameter D, length of the pipe L, and the pipe roughness
e represented by the average variation of its inner radius. Functionally, the
pressure drop ∆p may be expressed as

∆p = h(ρ, µ, V, L, D, e) (3.3)

Let us replace the right-hand side of Eq. (3.3) by an infinite series, as follows:

∆p = (k1 ρa1 µb1 V c1 Ld1 Df 1 eg1 ) + (k2 ρa2 µb2 V c2 Ld2 Df 2 eg2 ) + ... (3.4)

ISTUDY
3.6. CALCULATION OF THE DIMENSIONLESS GROUPS 179

where k1 , k2 ,... are dimensionless coefficients and a1 , b1 , ..., a2 , b2 , ... are ex-
ponents required by the series. We know that for dimensional homogeneity,
each group in Eq. (3.4) must have the same dimensions. Therefore, we need
to include only one term of the series in the dimensional representation of
Eq. (3.4). Hence, dropping the subscripts of the exponents and expressing
the equation dimensionally, we have
[ ] [ ]a [ ]b [ ]c
M M M L
≡ [L]d [L]f [L]g
LT 2 L3 LT T

For dimensional homogeneity the exponents of the basic dimensions M, L,


and T on both sides of the equation should be equal. Therefore, equating the
exponents of the basic dimensions on either side, we have

1=a+b for M (3.5)


−1 = −3a − b + c + d + f + g for L (3.6)
−2 = −b − c for T (3.7)

Equations (3.5)–(3.7) relate six quantities. Therefore, any three quantities


can be solved in terms of the remaining three. For instance, let us solve a, c,
and f in terms of b, d, and g. The resulting relations are

a=1−b (3.8)
c=2−b (3.9)
f = −b − d − g (3.10)

Therefore, the pressure drop Eq. (3.4) may be expressed by restricting the dis-
cussion to the first term of the series and replacing a, c, and f, by
Eqs. (3.8)–(3.10). Therefore,

∆p = kρ(1−b) µb V (2−b) Ld D(−b−d−g) eg


This can be expressed as
( )b ( )d (
∆p µ L e )g
=k
ρV 2 ρV D D D

Therefore, by grouping together the terms with the same exponents and ex-
tending the results to all the terms of the series, Eq. (3.4) may be expressed
as
( )b1 ( )d1 ( ) ( )b2 ( )d2 ( )
∆p µ L e g1 µ L e g2
2
= k 1 + k 2 + ...
ρV ρV D D D ρV D D D

In the functional form the above equation becomes


[( ) ( ) ( )]
∆p µ L e
2
=h , ,
ρV ρV D D D

ISTUDY
180 CHAPTER 3. DIMENSIONAL ANALYSIS AND SIMILARITY

It is seen that by this procedure the correct number of independent dimension-


less groups has been formed. The function h is unknown and, therefore, the
dimensionless group µ/(ρV D) may be inverted to result in (Re)D . Further,
∆p/(ρV 2 ) is the Euler number and (L/D) and (e/D) are the dimensionless
geometrical ratios.
The pressure through a pipe may then be characterized by the equation

Eu = h((Re)D , L/D, e/D)

The above example illustrates that a set of dimensionless groups, which are
independent and of a number consistent with the Buckingham’s π-theorem,
can be established by dimensional analysis.

3.7 SIMILARITY
Similarity or similitude in a general sense is the indication of a known relation-
ship between two phenomena. In fluid dynamics this is usually the relation
between a full-scale flow and a flow with smaller but geometrically similar
boundaries. We must also note that there are similarity rules in common
use in fluid dynamics involving flows with dissimilar boundaries. Here we
shall restrict our discussion to geometrically similar flows only, i.e. flows with
geometrically similar boundaries.
Two flows with geometrically similar sets of streamlines are called kine-
matically similar flows. But we know that the boundaries will form some
of the streamlines; therefore, the kinematically similar flows must also be
geometrically similar. However, the geometrically similar flows need not be
kinematically similar. It is quite easy to generate kinematically dissimilar
flows despite the presence of geometrically similar boundaries, as shown in
Figure 3.3. Here the lack of similarity between the streamlines in subsonic
and supersonic flows over similar double wedges in two-dimensional flows is
apparent.

FIGURE 3.3 Flow past double wedges.

In fluid dynamics there exists a third similarity called the dynamic


similarity, where the force distribution between the two flows is such that,
at corresponding points in the flows, identical types of forces (such as pres-
sure, shear, etc.) are parallel and have a ratio which has the same value at

ISTUDY
3.8. RELATIONSHIP BETWEEN DIMENSIONAL ANALYSIS... 181

all points of correspondence between the two flows. Further, this ratio must
be common to the various types of forces present. This nature of dynamic
similarity is important in experimental studies, since the forces such as lift
and drag are usually predicted for full-scale operations by measuring the cor-
responding forces on scaled models.
The conditions for dynamic similarity are:
1. The flows must be kinematically similar.
2. The flows must have mass distribution such that the density ratios at
the corresponding points of the flows are of the same ratio for all sets of
points.
Flows satisfying the condition 2 above are described as flows having similar
mass distribution.
The condition of kinematic similarity implies that the velocities and
accelerations at corresponding points are parallel and have a constant ratio
of magnitude between all corresponding sets of points. That is, kinemati-
cally, similar flows with similar mass distribution will satisfy all conditions of
dynamically similar flows.

3.8 RELATIONSHIP BETWEEN


DIMENSIONAL ANALYSIS
AND SIMILARITY
Consider two dynamically similar incompressible viscous flows about spheres,
called model flow and prototype flow as shown in Figure 3.4. Neglecting
the body forces, two types of forces–shear and pressure forces-may be distin-
guished as acting on each particle. Since the fluid particles are in motion, the
force due to inertia will also be acting on each particle.

FIGURE 3.4 Two flows past spheres.

Let Am and Ap be the corresponding points in the flow fields around the
model and the prototype. By dynamic similarity rules the forces acting at
these points are similar. Therefore, the following equations, which are true
for all corresponding points in any two flow fields, may be formed:
(inertia force)m (pressure force)m (friction force)m
= = = const. (3.11)
(inertia force)p (pressure force)p (friction force)p

ISTUDY
182 CHAPTER 3. DIMENSIONAL ANALYSIS AND SIMILARITY

From Eq. (3.11), we may derive the following relations:

(inertia force)m (inertia force)p


= = (const.)1 (3.12)
(friction force)m (friction force)p

(inertia force)m (inertia force)p


= = (const.)2 (3.13)
(pressure force)m (pressure force)p

The evaluation of Eqs. (3.12) and (3.13) in terms of flow variables will result
in Reynolds number and Euler number, respectively. Let us try to express
the force components in the above equations in terms of flow variables.

Viscous or friction force


Examine the infinitesimal fluid element with rectangular sides, at a position
along a streamline, as shown in Figure 3.5. The fluid element has sides of
length ds, dn, and dz. From Figure 3.5, the net shear force due to viscosity on
one pair of sides is ((∂τ /∂n)dndsdz). But, by Newton’s law of viscosity, τ =
µ(∂V /∂n). Therefore, the shear force can be expressed as µ(∂ 2 |V |/∂n2 )dV,
where dV = dndsdz, the volume of the fluid element considered.

Pressure force
The net pressure force on the pair of sides shown in Figure 3.5 is (∂p/∂n)dV.

FIGURE 3.5 Fluid element along a streamline.

Inertia force
The direction of motion of the fluid element under consideration is along the
streamline shown in Figure 3.5. Therefore, for steady flow,

∂|V |
dmaT = (ρdV)|V |
∂s
where dm = ρdV is the mass of the fluid element and aT is the acceleration
along the streamline.

ISTUDY
3.8. RELATIONSHIP BETWEEN DIMENSIONAL ANALYSIS... 183

Now, by the laws of dynamic similarity, the ratios of forces expressed in


Eqs. (3.12) and (3.13) can be replaced with the above components of the
viscous, pressure, and inertia forces. Therefore, from Eqs. (3.12) and (3.13),
we obtain
   
∂|V | ∂|V |
ρ|V | ρ|V |
 ∂s   ∂s 
  =  (3.14)
∂ 2 |V | ∂ 2 |V |
µ µ
∂n2 m ∂n2 p

   
∂|V | ∂|V |
ρ|V | ρ|V |
 ∂s   ∂s 
 ∂p  =
∂p  (3.15)
∂n m ∂n p

From dynamic similarity, we know that at all corresponding points in the


model and prototype flow fields the velocities have the same ratio in magni-
tude. Therefore, the local velocities in Eqs. (3.14) and (3.15) can be replaced
with the respective freestream velocities in the model and prototype flow fields.
Further, dynamic similarity implies that the difference between the velocity
vectors at corresponding points of the two flows are parallel and their ratio is
equal to the magnitude of the ratio between the velocities at corresponding
points, as shown in Figure 3.6.

FIGURE 3.6 Velocities at corresponding points.

Thus, we have

|VA |m |VB |m |VA − VB |m


= = (3.16)
|VA |p |VB |p |VA − VB |p

Taking the points A and B in Figure 3.6 infinitesimally close to each other
and replacing the local velocities by the corresponding freestream velocities,

ISTUDY
184 CHAPTER 3. DIMENSIONAL ANALYSIS AND SIMILARITY

we obtain from Eq. (3.16) the relation

|dV |m |V∞ |m
= (3.17)
|dV |p |V∞ |p

The above argument can be extended to the second derivative of velocity


appearing in Eq. (3.14). It can be easily shown that the freestream pressures
may be substituted for local pressures. Also, from kinematic similarity we
know that the distance between the corresponding positions of the flows must
have the same ratio as that of the corresponding lengths of model and proto-
type boundaries. Thus, dn may be replaced by the diameter of the sphere or
any other characteristic length L. Using the above results in Eqs. (3.14) and
(3.15), we obtain
( 2
) ( 2
)
ρV∞ /L ρV∞ /L
=
µV∞ /L2 m µV∞ /L2 p
( 2 ) ( 2 )
ρV∞ /L ρV∞ /L
=
p∞ /L m p∞ /L p
On simplification, the above two equations result in
( ) ( )
ρV∞ L ρV∞ L
= (3.18)
µ m µ p

( ) ( )
p∞ p∞
2
= 2
(3.19)
ρV∞ m ρV∞ p

From Eqs. (3.18) and (3.19) it is seen that the condition for dynamic similarity
for the flow fields over the model and prototype spheres is the equality of the
Reynolds number and Euler number between the flows. Thus, the dimensional
analysis results in the dimensionless groups whose values must be duplicated
between geometrically similar flows if dynamic similarity is to be attained
between the flows.
Further, in Eqs. (3.18) and (3.19), it can be shown that if the Reynolds
numbers are duplicated, the Euler number will then be automatically
duplicated. Therefore, for dynamic similarity between flows under discus-
sion, it is sufficient if (Re)D only is matched. Hence, (Re)D is called the
similarity parameter for the problem considered.
Since the similarity parameter was arrived at with dimensional analysis
by considering geometric similarity, it may be stated that the dynamically
similar flows must be geometrically similar. However, we must also note that
geometrically similar flows need not be dynamically similar.
Like Reynolds number and Euler number, there are many similarity pa-
rameters of importance in fluid dynamic studies. For instance, Mach number,
Froude number, Weber number, and Strouhal number are some of the impor-
tant similarity parameters commonly employed in fluid flow analysis.

ISTUDY
3.9. SIMILARITY REQUIREMENTS FROM THE EQUATIONS... 185

3.9 SIMILARITY REQUIREMENTS FROM


THE EQUATIONS OF FLOW
The conditions to be satisfied by dynamically similar flows may also be
obtained from the governing equations of the flow along with the bound-
ary conditions. For instance, examine the steady, incompressible flow past
a circular cylinder of radius R as shown in Figure 3.7. Let ρ and µ be the
density and viscosity, respectively. The governing equations for this flow are
the continuity equation

∇·V =0 (3.20)

FIGURE 3.7 Flow past a circular cylinder.

and the momentum equation

ρV · ∇V = −∇p − ρgj + µ∇2 V (3.21)

where j is the unit vector along the y-direction. The boundary conditions are:

1. V → V∞ i at x, y → ± ∞
where i is the unit vector along the x-direction.
2. V = 0 on the cylinder surface, i.e. on x2 + y 2 = R2
3. p = p∞ at x → −∞, y = 0

Let x∗ = x/R, V ∗ = V /V∞ , and p∗ = p/p∞ be the dimensionless forms of


x, V , and p, respectively, and R, V∞ and p∞ be the characteristic parameters
of length, velocity, and pressure.
Substituting these in Eqs. (3.20) and (3.21) and simplifying, we get

∇∗ · V ∗ = 0 (3.22)

p∞ ∗ ∗ gR µ
V ∗ · ∇∗ V ∗ = − 2
∇ p − 2 j+ ∇∗2 V ∗ (3.23)
ρV∞ V∞ ρV∞ R

with the boundary conditions as follows:

1. V ∗ → 1 at x∗ , y ∗ → ±∞

ISTUDY
186 CHAPTER 3. DIMENSIONAL ANALYSIS AND SIMILARITY

2. V ∗ = 0 on x∗2 + y ∗2 = 1
3. p∗ → 1 at y ∗ = 0 as x∗ → −∞

∂ ∂
∇∗ = i +j
∂x∗ ∂y∗
2
In Eq. (3.23), the dimensionless groups are p/ρV∞ = Eu; µ/ρV∞ R = 1/Re
2
and gR/V∞ = 1/Fr. Therefore, for the problem of flow past a cylinder, the
parameters to be matched for dynamic similarity are the Reynolds number,
the Euler number, and the Froude number.

3.10 SCALE FACTORS


The concept of scale factors is very useful, and is derived from the principle
of similarity. This has proved to be extremely useful in solving problems
with similarity principle. We know that dynamically similar flows must be
geometrically as well as kinematically similar and also should have similar
mass distribution. Also, the flow properties at corresponding points, also
referred to as homologous points, will bear the same ratio for all such points.
For geometric similarity, as we know, the corresponding length dimensions
of the prototype and model must be equal. Thus,

Lp Rp Tp
= = ≡ kL
Lm Rm Tm
where L, R, and T stand for the geometrical parameters, namely, the length,
the curvature and the thickness (subscripts p and m represent the prototype
and model, respectively), and kL is a constant, referred to as the length scale
factor. Similarly, we can define the velocity scale factor kV , the time scale
factor kT , etc. Like the basic dimensions M, L, and T , we can make use of
scale factors kM , kL , and kT .

EXAMPLE 3.3 Express mass in terms of scale factors.


Solution We know mass m = ρV, where ρ is density and V is volume. In
terms of scale factors, we can express mass as

kM = kρ (kL )3

Therefore, the scale factors kρ , kL , and kT , may be treated as the basic scale
factors of expressing the parameters governing any flow phenomenon. For
instance, force can be expressed in terms of basic scale factors as

Force F = (mass)(acceleration) = (ρV)(V /T )

kL 1 kρ (kL )4
= kρ (kL )3 =
kT kT (kT )2

ISTUDY
3.10. SCALE FACTORS 187

From the above discussion, it is seen that the scale factor is neither a concept
nor does it help in understanding any concept in a simple fashion. However, it
can serve as an elegant tool in solving problems associated with dimensional
analysis and similarity. At this stage, we must also realize that those who
believe only in the basic dimensions can solve problems in this field without
the aid of scale factors.

EXAMPLE 3.4 It is required to calculate the drag experienced by a small


submarine hull when it is moving far below the surface of water. A 1/10 scale
model is to be tested. What dimensionless groups should be matched between
the model and the prototype and why? If the drag on the prototype at 1 knot
is desired, at what speed should the model be moved to give the drag to be
experienced by the prototype? Would this result be true if the prototype were
to move close to the water surface? Explain.

Solution When the submarine travels far below the water surface, the
viscous forces predominate and, therefore, the Reynolds number must be
matched for dynamic similarity. That is,
( ) ( )
ρV L ρV L
=
µ m µ p

Let the model be tested in the same fluid. Then µp = µm and ρp = ρm .


Therefore,
Lp
Vm = Vp
Lm
Given
Lp 10
Vp = 1 knot, =
Lm 1
It follows that
Vm = 10 knots
The model must be moved at a speed of 10 knots in order to simulate dynam-
ically similar conditions between the model and prototype flow fields.
When the submarine moves close to the water surface, the waves offer
additional resistance and, therefore, the gravity effect must also be consid-
ered in the analysis. The above result would, therefore, not be valid then.

EXAMPLE 3.5 The pressure drop per unit length in a 6 mm diameter


gasoline fuel line is to be determined from a laboratory test using the same
tube but with water as the fluid. The pressure drop at a gasoline velocity of
1 m/s is of interest. (a) Determine the water velocity required. (b) At the
properly scaled velocity from part (a), the pressure drop per unit length was
found to be 65 Pa/m. What is the predicted pressure drop per unit length
for the gasoline line? For gasoline, the density is 680 kg/m3 and the viscosity
µ = 3.1 × 10−4 N · s/m2 ; for water the density is 1000 kg/m3 and the viscosity
µ = 1.12 × 10−3 N · s/m2 .

ISTUDY
188 CHAPTER 3. DIMENSIONAL ANALYSIS AND SIMILARITY

Solution For dynamics similarity between gasoline and water flows, their
Reynolds number should be equal, i.e.
( ) ( )
ρV d ρV d
=
µ g µ w

where the subscripts ‘g’ and ‘w’ refer to gasoline and water, respectively.
Therefore,
ρg µw
Vw = Vg
ρw µg
680 1.12 × 10−3
=1× ×
1000 3.1 × 10−4
= 2.46 m/s

The pressure drop through the water tube per unit length is given by

L 1
∆pw = f ρV 2
D 2
where f is the friction factor. Thus,

65 × (6 × 10−3 ) × 2
f= = 0.000129
1000 × 2.462

The pressure drop per unit length of gasoline line is

0.000129 1
∆pg = × × 680 × 12
6 × 10−3 2

= 7.31 Pa/m

EXAMPLE 3.6 A scaled model of an automobile vehicle of width 2.44 m


and frontal area 7.8 m2 was tested in a wind tunnel to find the aerodynamic
drag that would act on the vehicle when it plies at 100 km/h and to find the
power required to overcome this drag. The scale of the model was 16:1. The
test revealed that the drag coefficient on the model was 0.46 and it becomes
independent of Reynolds number for Reynolds numbers greater than 105 .
Estimate the drag force and the power required for the actual vehicle, if
the test-section air flow conditions were identical to sea level atmospheric
conditions.

Solution At standard sea level state, we have

p = 101325 Pa T = 288 K

ISTUDY
3.10. SCALE FACTORS 189

The corresponding density and viscosity are

p 101, 325
ρ= =
RT 287 × 288
= 1.226 kg/m3

T 3/2
µ = 1.46 × 10−6
T + 111
2883/2
= 1.46 × 10−6
288 + 111
= 1.79 × 10−5 kg/(m s)

Let the subscripts p and m refer to the prototype and model, respectively.
The model width is
wp 2.44
wm = =
16 16
= 0.153 m

The model frontal area is


Ap 7.8
Am = 2
= 2
16 16

= 0.0305 m2

For dynamic similarity between the model and prototype,


( ) ( )
ρV L ρV L
=
µ m µ p

But the density and viscosity are invariants. Thus,

(V L)m = (V L)p
( )
16
Vm = (100/3.6) = 444.4 m/s
1

The corresponding Reynolds number is

1.226 × 444.4 × (2.44/16)


(Re)m =
1.79 × 10−5

= 46.4 × 105

ISTUDY
190 CHAPTER 3. DIMENSIONAL ANALYSIS AND SIMILARITY

The prototype also will have the same Reynolds number, thus the drag
coefficient for the prototype is also 0.46. The aerodynamic drag for the pro-
totype becomes
1
Dp = C D ρ Vp2 Ap
2
( )2
1.226 100
= 0.46 × × × 7.8
2 3.6

= 1697 N

The power required to overcome this drag is


100
Power = Dp Vp = 1697 × = 47139 W
3.6

= 47.139 kW

EXAMPLE 3.7 A 1:50 model boat has a wave resistance of 0.02 N when op-
erating at 1.0 m/s. Find the corresponding prototype wave resistance. Also,
find the horse power requirement for the prototype and the corresponding
velocity.

Solution For dynamic similairity, the Froude numbers of the model and
prototype should be equal. That is,

Frp = Frm

where subscripts p and m, refer to the prototype and model, respectively.

The Froude number is given by

V
Fr = √
gL

Thus, ( ) ( )
V V
√ = √
gL p gL m

This simplifies to
Vp2 V2
= m
Lp Lm
or
Vp2 Lp
2
= = 50
Vm Lm
The force is given by
F = ρL2 V 2

ISTUDY
3.11. SUMMARY 191

Therefore,

Fp ρL2p Vp2
=
Fm ρL2m Vm2

L2p Vp2
=
L2m Vm2
or ( )2 ( ) ( )3
Fp Lp Lp Lp
= =
Fm Lm Lm Lm
Therefore,
( )3
Lp
Fp = Fm
Lm

= 503 × 0.02

= 2500 N

The velocity for the prototype is



Lp
Vp = Vm
Lm

= 50 × 1

= 7.07 m/s

The power required for the prototype is

Pp = Fp Vp

= 2500 × 7.07

= 17675 watts

17675
=
746

= 23.69 hp

3.11 SUMMARY
The principle of homogeneity states that analytically derived equations are
valid for any system of units. This principle is of immense help in identify-
ing the independent dimensionless groups, called the π numbers involved in

ISTUDY
192 CHAPTER 3. DIMENSIONAL ANALYSIS AND SIMILARITY

any flow problem. The scheme proposed by Buckingham to identify these π


numbers is called the Buckingham′ s π-theorem.
The dimensionless groups which are of importance in fluid dynamics study
are: Reynolds number, Mach number, Froude number, Weber number, Euler
number, and so on.
Similarity in a general sense is the indication of a known relationship
between two phenomena. The geometric and kinematic similarities are the
prerequisites for dynamic similarity.
The scale factor is an elegant tool used in solving problems associated
with dimensional analysis and similarity.

3.12 PROBLEMS
3.1 The velocity of sound a of a gas varies with pressure p and density ρ.
Show by dimensional reasoning that the proper form of the sound velocity
should be a = c (p/ρ)1/2 , where c is a constant.
3.2 The drag on a small flying object which will move in air at 20◦ C and at
a speed of 10 m/s is to be found through a model test. If a 1/20 scale model
is tested in water at 20◦ C, to simulate the flying object in air, what should
be the water velocity? At this velocity, if the measured drag on the model
(in water) is 4000 N, what will be the drag on the prototype of the flying
object and the power required to propel it?
[Ans. 13.46 m/s, 8849 W]
3.3 Flow through a heat exchanger tube is to be studied by means of a 1/10
scale model. If the heat exchanger normally carries water, determine the ratio
of pressure losses between the model and the prototype if (a) water is used
in the model and (b) air at normal temperature and pressure is used in the
model. Take ρw = 103 kg/m3 , µw = 10−3 N · s/m2 , ρa = 1.23 kg/m3 , µa =
1.8 × 10−5 N · s/m2 , where the subscripts ‘w’ and ‘a’ refer to water and air,
respectively.
[Ans. (a) 100, (b) 26.34]
3.4 Consider the jet pump shown in Figure P3.4. The functional dependence
of △p is ( )
△p = f ρ, V, d, D, µ, Q̇

Find the number of independent dimensionless groups needed to characterize


the jet pump. Obtain by dimensional analysis the nondimensional groups that
contain the volume flow rate Q̇ and the viscosity coefficient µ.

ISTUDY
3.12. PROBLEMS 193

FIGURE P3.4


[Ans. 4, ]
V d2
3.5 If an airplane travels at a speed of 1100 km/h at an altitude of 15 km,
what is the required speed at an altitude of 8 km to satisfy the Mach number
similarity?
[Ans. 1220 km/h]
3.6 The water velocity at a certain point along a 1:20 scale model of a dam
spillway is 6 m/s. What is the corresponding prototype velocity if the model
and the prototype flow fields are to be similar?
[Ans. 26.8 m/s]
3.7 A 1:13 scale model of a ballistic missile which travels at 380 m/s through
air at 23◦ C and 95 kPa has to be tested in a high-speed wind tunnel. (a) If
the air in the wind tunnel test-section is at −20◦ C and 89 kPa, what must its
velocity be? (b) Determine the drag force on the prototype if the drag force
on the model is 400 N.
[Ans. (a) 351.3 m/s, (b) 79.1 kW]
3.8 A 1:16 model of a bus is tested in a wind tunnel in standard air. The
model is 150 mm wide, 200 mm high, and 760 mm long. The model drag
measured at 26.5 m/s is 6 N. The longitudinal pressure gradient in the wind
tunnel test-section is − 12 Pa/m. (a) Estimate the correction that should be
made to the measured drag force to correct for horizontal buoyancy caused
by the pressure gradient in the test-section. (b) Calculate the drag coefficient
for the model. (c) Evaluate the aerodynamic drag force on the prototype at
100 km/h on a calm day.
[Ans. (a) 0.274 N, (b) 0.444, (c) 1.613 kN]
3.9 A 1:50 scale model of a submarine is tested in a water tunnel. If the
model test at 10 knots yielded a measured drag of 13 N, determine the drag
on the full-scale submarine at 27 knots.
[Ans. 236.93 kN]
3.10 An aircraft flies at a true speed of 350 m/s at an altitude where the
pressure and temperature are 18.25 kPa and 216.5 K, respectively. (a) If a

ISTUDY
194 CHAPTER 3. DIMENSIONAL ANALYSIS AND SIMILARITY

one-fourteenth scale model of the aircraft is to be tested, under dynamically


similar conditions, in a wind tunnel with test-section temperature 288 K,
(a) what should be the pressure in the test-section? Assume the viscosity of
air varies with temperature as T 3/4 approximately. (b) Show that the forces
on the model will be about 10% of the corresponding forces on the prototype.
[Ans. (a) 365.26 kPa]
3.11 An aircraft wing of span 10 m and mean chord 2 m is designed to
develop 45 kN lift at freestream velocity 400 km/h and density 1.2 kg/m3 . A
1/20 scale model of the wing section is tested in a wind tunnel at velocity 500
m/s and density 5.33 kg/m3 . The total drag measured is 400 N. Assuming the
wind tunnel data refer to a section of infinite span, calculate the total drag and
aerodynamic efficiency of the aircraft wing, assuming the load distribution to
be elliptic.
[Ans. D = 2.649 kN, L/D = 16.99]
3.12 A 1:15 model of the water spillway of width 20 m and flow rate 125
m3 /s is to be designed. (a) Determine the width and flow rate for the model.
(b) Also, find the time for the model which corresponds to one day operation of
the prototype. Assume the viscosity and surface tension effects are negligible.
[Ans. (a) 1.33 m, 0.1434 m3 /s, (b) 6.198 hours]

ISTUDY
Chapter 4

Boundary Layer

4.1 INTRODUCTION
The concept of boundary layer conceived by Ludwig Prandtl (1857–1953) in
1904 was one of the greatest inventions in the field of fluid dynamics. The
concept of ‘fluid boundary layer’ laid the foundation for the unification of the
theoretical and experimental aspects of fluid mechanics. To gain an insight
into this concept let us examine the streaming flow of a fluid past a body of
reasonable slender form as shown in Figure 4.1. In the majority of problems
associated with aerodynamics the fluid viscosity is relatively small, so that,
unless the transverse velocity gradients are appreciable, the shearing stresses
developed [given by equation τ = µ (∂u/∂y)] will be very small. For flows,
such as that indicated in Figure 4.1, the transverse velocity gradients are
usually negligibly small throughout the flow field except for thin layers of fluid
immediately adjacent to the solid body boundaries. Within these boundary
layers of fluid, however, large shearing velocities are produced with consequent
shearing stresses of considerable magnitude.

FIGURE 4.1 Viscous flow past an aerofoil.

195

ISTUDY
196 CHAPTER 4. BOUNDARY LAYER

Prandtl pointed out that these boundary layers were usually thin, pro-
vided the body was of streamlined form, at a moderate angle of incidence to
the flow and that the flow Reynolds number was sufficiently large, so that
as a first approximation their presence might be ignored in order to estimate
the pressure field produced about the body. For aerofoil shapes, the pressure
field around them is only slightly modified by the boundary layer flow, since
almost the entire lifting force is produced by normal pressures at the aerofoil
surface. Therefore, it is possible to develop theories for the evaluation of the
lift force by consideration of the flow field outside the boundary layers, where
the flow is essentially inviscid in behaviour. From this we can understand the
importance of the inviscid flow theories. But it is important to realize that no
drag force, other than the induced drag, can be evaluated from inviscid flow
theories. For streamlined bodies like aerofoil, the drag force is essentially due
to shearing stresses at the body surface and therefore the study of bound-
ary layer behaviour is essential for estimating this. Prandtl’s boundary layer
concept aids enormous simplification in the study of the whole problem. The
equations of viscous motion need be considered only in the limited regions
within the boundary layers, where appreciable simplifying assumptions can
reasonably be made. However, in spite of this simplification, the prediction
of boundary layer behaviour is still by no means simple.

4.2 BOUNDARY LAYER DEVELOPMENT


Essentially, there are two types of boundary layer developments encountered
in practice. They are:

• For flow around a body with a sharp leading edge, the boundary layer
will grow from zero thickness at the upstream edge (leading edge) of the
body.

• For a body with a blunt nose, like that of a typical aerofoil, the boundary
layer will develop on top and bottom surfaces from the front stagnation
point. The boundary layer for such bodies will have finite (not zero)
thickness at the leading edge too, unlike a body with a sharp leading
edge for which the boundary layer thickness is zero at the leading edge.

When the flow proceeds downstream along a surface, large shearing gra-
dients and stresses develop adjacent to the surface because of the relatively
large velocities in the mainstream and the condition of no-slip (zero velocity)
at the surface. Initially, this shearing action occurs only at the body surface
and retards the layers of fluid adjacent to the surface, causing the fluid ele-
ments in contact with the surface to come to rest. This is popularly termed
no-slip condition. These elements in turn will interact with the elements in
the layers above them and retard their motion. In this way, as the fluid near
the surface passes downstream, the retarding action penetrates farther away
from the surface and the boundary layer of retarded fluid thickens up.

ISTUDY
4.2. BOUNDARY LAYER DEVELOPMENT 197

4.2.1 Velocity Profile


The flow velocity increases continuously from zero at the surface to nearly the
freestream value at the edge of the boundary layer. Let y be the perpendicular
distance from the surface at any point and let u be the corresponding velocity
parallel to the surface. Consider a small element of fluid of unit depth normal
to the flow plane, having unit length in the direction of motion and a thickness
δy normal to the flow direction, as shown in Figure 4.2.

FIGURE 4.2 A fluid element in a viscous flow.

The shearing stress on the face AB will be τ = µ(∂u/∂y) and that on


the face CD will be τ + (∂τ /∂y)δy, in the direction shown, assuming that u
increases with y. Thus, the resultant shearing force in the x-direction will
be [τ + (∂τ /∂y)δy] − τ = (∂τ /∂y)δy.( But τ )= µ(∂u/∂y), therefore, the net
shear force on the element will be µ ∂ 2 u/∂y 2 δy. Unless µ is zero, it follows
that ∂ 2 u/∂y 2 cannot be infinite and therefore, the rate of change of velocity
gradient in the boundary layer must also be continuous.
Now, it is clear that a smooth curve of the general form shown in
Figure 4.3 will develop if the velocity is plotted versus the distance y. Note
that, at the surface (y = 0) the curve is not tangential to the u-axis as this

FIGURE 4.3 Velocity profile in a boundary layer.

ISTUDY
198 CHAPTER 4. BOUNDARY LAYER

would imply an infinite gradient ∂u/∂y, and hence an infinite shearing stress
at the surface. It is also evident that as the shearing gradient decreases, the
retarding action decreases too, so that at some distance from the surface where
∂u/∂y becomes very small, the shearing stress becomes negligible, although
theoretically a small gradient must exist out to y = ∞.
It will be useful to note that the value of viscosity coefficient µ to be used
in a turbulent boundary layer, will not, in general, be the simple coefficient
of absolute viscosity of the fluid.

4.3 BOUNDARY LAYER THICKNESS


In order to make the concept of boundary layer realistic, an arbitrary decision
must be made about its extent, and therefore the usual convention is that the
boundary layer extends to a distance δ from the surface such that the velocity
u at that distance is 99 per cent of the local mainstream velocity U . Thus,
δ is the physical thickness of the boundary layer. It is a usual practice to
express the freestream velocity as U∞ , but here the subscript ∞ is dropped
for simplicity.
The boundary layer thickness may also be defined as “the distance from
the solid boundary within which the local value of the flow velocity increases
from zero at the wall (y = 0) to 0.99 of the freestream value at the edge of
the layer (y = δ)”, i.e.

at y = 0, u=0
at y = δ, u = 0.99 U

where U is the freestream velocity. The above definition of the boundary layer
is somewhat arbitrary.

4.3.1 Displacement Thickness


A more physically meaningful definition of the boundary layer can be intro-
duced by considering a parameter known as displacement thickness δ ∗ in order
to account for the decrease in the total mass flow rate caused by the boundary
layer. The displacement thickness is defined as “the distance by which the
wall would have to be displaced outwards in a hypothetical frictionless flow
so as to maintain the same mass flux as in the actual flow”.
Examine the velocity profile shown in Figure 4.4. If the boundary layer
is ignored and the flow is treated as potential, the mass flow rate through
the thickness δ would be ρ∞ U∞ (considering the flow to be two-dimensional).
But when the boundary layer is considered, the flow rate through the same
area would decrease from ρ∞ U∞ to some value ρ u. Therefore, to satisfy the
continuity the streamtube cross-sectional area will increase, and the stream-
tube in the mainstream flow will be displaced slightly away from the surface.
The effect on the mainstream flow will then be as if, with no boundary layer

ISTUDY
4.3. BOUNDARY LAYER THICKNESS 199

FIGURE 4.4 Displacement thickness.

present, the solid surface has been displaced a small distance into the stream,
as shown in Figure 4.4.
In mathematical terms, the mass of the fluid which is absent due to the
presence of the boundary layer is ρ∞ U∞ δ ∗ . Equating this to the mass of fluid
which is absent due to the actual boundary layer gives the equation which
defines the displacement thickness, δ ∗ . Thus,
∫ δ

ρ∞ U∞ δ = (ρ∞ U∞ − ρu) dy
0

i.e.
∫ δ ( )
∗ ρu
δ = 1− dy (4.1)
0 ρ∞ U∞

For incompressible flows, the local ρ and the freestream ρ∞ densities cancel
out and Eq. (4.1) reduces to
∫ δ ( )
∗ u
δ = 1− dy (4.2)
0 U∞
The upper limit in Eq. (4.2) may be allowed to extend to infinity because,
u/U∞ → 1 exponentially in y as y → ∞ at the edge of the boundary layer.
The concept of displacement thickness proposed here is purely based on two-
dimensional flow past a flat plate in order to conceive the concept in its
simplest form. The above equations may be used for any two-dimensional
flow without restriction. They also can sensibly be used for three-dimensional
bodies provided the curvature in the plane normal to the freestream direction
is not large, i.e. if the local radius of curvature is much greater than the
boundary layer thickness. When the curvature is large, a displacement thick-
ness may still be defined but the form of Eqs. (4.1) and (4.2) will be slightly
modified. The concept of displacement thickness is used in the design of
ducts, in the intakes of air-breathing engine, and in the wind tunnels, etc.
by first assuming a frictionless flow and then enlarging the passage wall by
the displacement thickness so as to allow the same flow rate. Another use of
displacement thickness is in finding the pressure gradient dp/dx at the edge of

ISTUDY
200 CHAPTER 4. BOUNDARY LAYER

the boundary layer, needed for solving the boundary layer equations. The first
approximation is to neglect the existence of the boundary layer, and calcu-
late the irrotational dp/dx over the body surface. A solution of the boundary
layer equations gives the displacement thickness, using Eq. (4.2). The body
surface is then displaced outwards by this distance and a next approximation
of dp/dx is found from a solution of the irrotational flow, and so on.
Using similar arguments to those given for boundary layer and displace-
ment thicknesses, we can define other thicknesses associated with boundary
layer, namely the momentum and energy thicknesses using the momentum
and energy flow rates, respectively.

4.3.2 Momentum Thickness


Another type of boundary layer thickness which is frequently used is the
momentum thickness, denoted by θ. This is defined based on the momentum
flow rate within the boundary layer. This rate is less than that which would
occur if no boundary layer existed, and the velocity in the vicinity of the sur-
face at the station considered would be constant and equal to the mainstream
velocity U∞ . The momentum thickness is defined as the thickness of a layer
of fluid of velocity U∞ for which the momentum flux is equal to the deficit of
momentum flux through the boundary layer.
For a streamtube of thickness δy within the boundary layer (Figure 4.2)
the rate of momentum deficit (relative to the mainstream) is ρu(U∞ − u)δy.
Note that the actual mass flow rate ρu within the streamtube must be used
here, the momentum deficit of this mass being the difference between its
momentum based on the mainstream velocity and its actual momentum at
position x within the boundary layer.
2
The rate of momentum deficit for the thickness θ is given by ρ∞ U∞ θ.
Thus,
∫ δ
2
ρ∞ U∞ θ = ρ u (U∞ − u) dy
0
or
∫ δ ( )
ρu u
θ= 1− dy (4.3)
0 ρ∞ U∞ U∞

For the case of incompressible flow this thickness reduces to


∫ δ ( )
u u
θ= 1− dy
0 U∞ U∞

The momentum thickness concept is conveniently used in transition problems


and in the calculation of the skin friction loss. From the expression of θ above,
we can infer that the momentum thickness is the distance through which the
surface would have to be displaced in order that, with no boundary layer, the
total flow momentum at the station considered would be the same as that
which would actually occur.

ISTUDY
4.3. BOUNDARY LAYER THICKNESS 201

4.3.3 Kinetic Energy Thickness


The kinetic energy thickness, denoted by ϑ, may be defined as the distance
through which the surface would have to be displaced in order that, with no
boundary layer, the total flow kinetic energy at the station considered would
be the same as that which would actually occur. This quantity is defined
with reference to kinetic energy of the fluid in a manner comparable with the
momentum thickness. The rate of kinetic energy deficit within the boundary
layer at any station x is given by the difference between the kinetic energy
which the element would have at mainstream velocity U∞ and that which it
would actually have at velocity u, thus being equal to
∫ δ
1 ( 2 )
ρu U∞ − u2 dy
0 2

The rate of kinetic energy deficit in the thickness ϑ is 12 ρ∞ U∞


3
ϑ. Thus,

1 δ
1 ( 2 )
3
ρ∞ U∞ ϑ= ρu U∞ − u2 dy
2 0 2
or [
∫ δ ( )2 ]
ρu u
ϑ= 1− dy (4.4)
0 ρ∞ U∞ U∞

The displacement, the momentum, and the energy thicknesses can be ex-
pressed in non-dimensional form by dividing them by the boundary layer
thickness to result in
∫ 1( )
δ∗ ρu
= 1− dy
δ 0 ρ∞ U∞

∫ 1 ( )
θ ρu u
= 1− dy
δ 0 ρ∞ U∞ U∞

∫ [ ( )2 ]
1
ϑ ρu u
= 1− dy
δ 0 ρ∞ U∞ U∞

where y = y/δ. In the above three equations the integrals on the right-hand
side are simply the numbers which may be evaluated readily if the boundary
layer velocity profile is known.

4.3.4 Non-Dimensional Velocity Profile


The expression of velocity profile in non-dimensional form will be useful to
compare boundary layer profiles of different thicknesses. This may be done
by writing u = u/U and y = y/δ, so that the velocity profile shape is given by

ISTUDY
202 CHAPTER 4. BOUNDARY LAYER

u = f (y), where u is the local flow velocity and U is the freestream velocity.
Over the range y = 0 to y = δ, the dimensionless velocity u varies from 0 to
0.99. For convenience when using u values as integration limits, a negligible
error is introduced by using u = 1.0 instead of 0.99, at the edge of the bound-
ary layer, and therefore considerable arithmetic simplification is achieved. The
velocity profile is then plotted as in Figure 4.5.

FIGURE 4.5 Non-dimensional velocity profile.

4.3.5 Types of Boundary Layer


It has been proved experimentally that, as far as pipe flow is concerned, there
are two different flow regimes, namely the laminar flow and the turbulent flow,
which can exist. In laminar flow the layers of fluid elements slide smoothly
over one another and there is little interchange of fluid mass between the
adjacent layers. The shearing traction which is developed due to the velocity
gradient is thus entirely due to the viscosity of the fluid, i.e. the momentum
exchange between the adjacent layers is on a molecular scale only.
In turbulent flow, considerable random motion exists in the form of velocity
fluctuations both along the mean direction of flow and perpendicular to it.
As a result of the latter, there is appreciable transport of mass between the
adjacent layers. If there is a mean velocity gradient in the flow, there will be
corresponding interchanges of streamwise momentum between adjacent layers,
which will result in shearing stress between them. These shearing stresses may
be of much greater magnitude than those which develop as a result of purely
viscous action. The stresses due to turbulence are termed Reynolds stresses.
The velocity profile shape in a turbulent boundary layer is largely controlled
by these Reynolds stresses.
Due to the difference between the laminar and the turbulent flow shearing
stresses, the velocity profiles in the two types of boundary layers are differ-
ent. Typical velocity profiles in laminar and turbulent boundary layers on

ISTUDY
4.3. BOUNDARY LAYER THICKNESS 203

a flat plate, where there is no streamwise pressure gradient, are shown in


Figure 4.6.

FIGURE 4.6 Laminar and turbulent velocity profiles in a flat plate


boundary layer.

In the laminar boundary layer, the energy from the mainstream is trans-
ferred towards the slower moving fluid near the surface through the viscous
action alone, resulting in only a relatively small perturbation. Owing to this,
a considerable portion of the boundary layer flow has a significantly reduced
velocity. Throughout the boundary layer, the shearing stress τ is given by

∂u
τ =µ
∂y

and at the wall the shearing stress becomes


( ) ( )
∂u ∂u
τw = µw = µw (say)
∂y y=0 ∂y w

In the turbulent boundary layer, as has already been noted, large Reynolds
stresses are set up due to mass transport in the direction perpendicular to the
surface, so that the energy from the mainstream may easily penetrate to
fluid layers quite close to the surface. Because of this, these layers have a
velocity which is not much less than that of the mainstream. However, in
layers which are very close to the surface it is not possible for the velocity to
exist perpendicular to the surface, so that in a very thin region immediately
adjacent to the surface, the flow approximates to laminar flow. This thin layer
adjacent to the surface is termed laminar sublayer or viscous sublayer.
In the laminar sublayer the shearing action becomes purely viscous and
the velocity falls very sharply, and almost linearly, within it, to zero at the
surface. Therefore, the wall shear stress now depends only on viscosity, i.e.
τw = µw (∂u/∂y)w . The surface friction stress in a turbulent boundary layer
will be far greater than that in a laminar boundary layer of the same thickness,
since (∂u/∂y)w is much greater for the turbulent boundary layer. It should

ISTUDY
204 CHAPTER 4. BOUNDARY LAYER

be noted that the viscous shear stress relation is employed only in the laminar
sublayer very close to the surface and not throughout the turbulent boundary
layer.

4.4 Boundary Layer Flow


Boundary layer may be defined as that thin layer adjacent to a solid bound-
ary within which the flow velocity increases from zero to 99 per cent of its
freestream value. Boundary layer may also be defined as that fluid layer which
has had its velocity affected by the boundary shear.
Boundary layer theory is a technique to compute the viscous-layer motion
near solid walls, and ‘patch’ it onto the outer inviscid flow. The patching
becomes more successful as the Reynolds number becomes larger. Examine
the boundary layers over a flat plate as shown in Figures 4.7 and 4.8.

FIGURE 4.7 Boundary layer over a flat plate for a laminar, low-Reynolds
number flow.

FIGURE 4.8 Boundary layer over a flat plate for a high-Reynolds


number flow.

It is seen from the comparison of low- and high-Reynolds number flows


over a sharp flat-plate that:

ISTUDY
4.4. BOUNDARY LAYER FLOW 205

• For low-Reynolds numbers, the viscous region is very broad and extends
far ahead and to the sides of the plate.
• The plate retards the oncoming stream greatly, and small changes in
flow parameters cause large changes in the pressure distribution along
the plate.
• Thus, although in principle it should be possible to patch the viscous and
inviscid layers in a mathematical analysis, their interaction is vigorous
and nonlinear.
• There is no simple theory that exists for the external flow analysis, for
the Reynolds number ranges from 1 to about 1000.
• Such thick shear layer flows are typically studied by experimental or
numerical methods.
• A high-Reynolds number flow is much more amenable to boundary layer
patching than a low-Reynolds number flow.
• The viscous layers, either laminar or turbulent, are very thin, as shown
in Figure 4.7.
• It can be shown for flat-plate flow that the boundary layer thickness δ
can be expressed as
δ 5.0
≈ 1/2
for laminar boundary layer
x [(Re)x ]
δ 0.16
≈ 1/7
for turbulent boundary layer
x [(Re)x ]

where (Re)x = ρ∞ U∞ x/µ∞ is called the local Reynolds number. Here


ρ∞ , µ∞ , and U∞ are the freestream density, viscosity and velocity,
respectively and x is the axial distance from the leading edge of the
flat plate.

The turbulent boundary layer thickness relation given above applies for
Reynolds numbers greater than approximately 106 . Some values of δ/x
calculated using the above relations are given in Table 4.1.
The blanks in Table 4.1 indicate that the relation is not applicable for
those cases.

• In all the cases in Table 4.1 the boundary layer is so thin that its dis-
placement effect on the outer inviscid flow region is negligible. Thus,
the pressure distribution along the plate can be computed from inviscid
theory as if the boundary layer is not there.

• For slender bodies such as a flat-plate or an aerofoil kept parallel to


the oncoming stream, the assumption of negligible interaction between
the boundary layer and the outer pressure distribution is an excellent
approximation.

ISTUDY
206 CHAPTER 4. BOUNDARY LAYER

TABLE 4.1 (δ/x) values

(Re)x 104 105 106 107 108

(δ/x)lam 0.050 0.016 0.005 — —

(δ/x)tur — — 0.023 0.016 0.011

• For blunt-bodies, however, even at high Reynolds numbers there is a


discrepancy in the viscous-inviscid patching concept.

Examine the inviscid and actual flow fields over circular cylinders shown in
Figure 4.9.

FIGURE 4.9 (a) Ideal flow past a circular cylinder and (b) actual flow
past a circular cylinder.

• In the idealized flow field there is a thin film of boundary layer about
the body and a narrow sheet of viscous wake in the rear. The patching
would be glorious for this picture, but it is false.

• In the actual flow the boundary layer is thin on the front, or


windward, side of the body where the pressure decreases along the
surface (favourable pressure gradient). But the rear boundary layer
encounters increasing pressure (adverse pressure gradient) and breaks
off, or separates, into a broad, pulsating wake.

• The theory of strong interaction between the blunt-body viscous and


inviscid layers is not well-developed. Flows like the actual ones shown
in Figure 4.9(b) are usually studied experimentally.

4.5 BOUNDARY LAYER SOLUTIONS


For solving external flows, basically three techniques are used. They are:

ISTUDY
4.6. MOMENTUM-INTEGRAL ESTIMATES 207

1. Numerical methods
2. Experimental methods
3. Boundary layer theory

Here we are concerned with only the boundary layer theory. It was first formu-
lated by Prandtl in 1904. Prandtl proposed order-of-magnitude assumptions
to simplify the Navier–Stokes equation to result in boundary layer equations,
which can be solved relatively easily and patched onto the outer inviscid-flow
field. One of the great achievements of the boundary layer theory is its ability
to predict the flow separation illustrated in Figure 4.9(b). However, it should
be realized that this prediction is only approximate and not accurate.

4.6 MOMENTUM-INTEGRAL ESTIMATES


4.6.1 Conservation of Linear Momentum
By Newton’s second law, the momentum equation for flow through a control
volume can be expressed as
dV d
F = ma = m = (mV )
dt dt
where F is the force acting on the control volume and V is the flow velocity
relative to the control volume. By Reynolds transport theorem, the linear-
momentum relation for a deformable control-volume becomes
∑ (∫ ∫ ∫ ) ∫∫
d d
(mV )system = F = V ρ dV + V ρ (Vr · n) dA
dt dt CV CS

where V is fluid velocity relative to an inertial(non-accelerating) coordinate


system, otherwise the Newton’s law ∑must be modified to include the non-
inertial relative acceleration terms. F is the vector sum of all forces acting
on the control-volume material considered as a freebody and Vr is the velocity
of flow relative to the control-volume.
The entire equation is a vector relation, both the integrals are vectors due
to the term V in the integrals. The equation thus has three components. If
we want only, say, the x-component, the equation reduces to
∑ (∫ ∫ ∫ ) ∫∫
d
Fx = uρ dV + uρ (Vr · n) dA
dt CV CS
∑ ∑
and similarly, Fy and Fz would involve v and w, respectively.
For a fixed control-volume, the relative velocity Vr ≡ V , and we can use
the partial derivative to express the momentum equation as

∑ (∫ ∫ ∫ ) ∫∫

F = V ρ dV + V ρ (V · n) dA (4.5)
∂t CV CS

ISTUDY
208 CHAPTER 4. BOUNDARY LAYER

FIGURE 4.10 Boundary layer over a flat plate.

This is called the momentum-integral relation.


Consider a shear layer of unknown thickness that grows along the sharp
flat-plate shown in Figure 4.10.
The drag force on the plate is given by the following momentum integral
across the exit plane.
∫ δ(x)
D(x) = ρ b u (U − u) dy (4.6)
0

where u is the local velocity, U is the freestream velocity, ρ is the local density,
and b is the width of the plate. Thus, the momentum-integral equation finds
application in the boundary layer analysis.

4.6.2 Karman’s Analysis of the Flat Plate


Boundary Layer
Equation (4.6) was derived by Karman in 1921, who wrote it in the convenient
form of the momentum thickness θ, as follows:

D(x) = ρ b U 2 θ

where the momentum thickness θ is given by


∫ δ(x) (
u u)
θ= 1− dy (4.7)
0 U U
Thus, the momentum thickness is a measure of the total drag of the plate.
Karman then noted that the drag is also equal to the integral of wall shear
stress along the plate. Therefore, the drag can also be expressed as
∫ x
D(x) = b τw (x) dx (4.7a)
0
or
dD
= bτw
dx

ISTUDY
4.7. BOUNDARY LAYER EQUATIONS 209

Differentiating the equation D = ρ b U 2 θ, we get

dD dθ
= ρ b U2
dx dx
Comparing the above equation with Eq. (4.7), Karman arrived at the follow-
ing equation which is now called the momentum-integral relation for flat-plate
boundary layer,

dθ d δ
u( u)
τw = ρ U 2
= ρU 2 1− dy (4.8)
dx dx 0 U U

This is the momentum equation, and is valid for both laminar and turbulent
flows.
To get a numerical result for laminar flow, Karman assumed that the
velocity profile had an approximately parabolic profile and expressed that as
( )
2y y 2
u(x, y) ≈ U − 2 0 ≤ y ≤ δ(x) (4.9)
δ δ

4.7 BOUNDARY LAYER EQUATIONS


The boundary layer concept is based on the fact that the layer is thin. For
a flat plate this means that at any location x along the plate length, δ ≪ x,
and θ ≪ x and ϑ ≪ x. The structure and properties of the boundary layer
flow depend on whether the flow is laminar or turbulent. Also, we saw that
the increase in velocity from zero at the wall to the full magnitude of the
freestream value at some distance above the wall takes place in a very thin
layer, termed the boundary layer. In this manner there are two regions to be
considered, even if the division between them is not very sharp.

1. A very thin boundary layer adjacent to the body, in which the velocity
gradient normal to the wall, i.e. ∂u/∂y is very large. In this region the
very small viscosity of the fluid exerts an essential influence insofar as
the shearing stress τ = µ (∂u/∂y) may assume large values.

2. In the region outside the boundary layer no such large velocity gradient
occurs and the influence of viscosity is insignificant. In this region the
flow is almost frictionless and can be treated as potential.

In general the thickness of the boundary layer decreases with viscosity,


i.e. it decreases as the Reynolds number increases. Many exact solutions
of the Navier–Stokes equations reveal that the boundary layer thickness is
proportional to the square root of kinematic viscosity ν, i.e.

δ∼ ν

ISTUDY
210 CHAPTER 4. BOUNDARY LAYER

For simplifying the Navier–Stokes equations it is assumed that the boundary


layer thickness is very small compared to any linear dimension L of the body
under consideration, i.e.
δ≪L
In this way the solutions obtained from the boundary layer equations are
asymptotic and apply only to very large Reynolds numbers.
Now let us simplify the Navier–Stokes equations using an estimate of the
order of magnitude of each term. For a two-dimensional flow, let us be-
gin by assuming the wall to be flat and coinciding with the x-direction, and
y-axis to be perpendicular to it. Let us express the Navier–Stokes equations
in dimensionless form by referring all velocities to the freestream velocity U ,
and by referring all linear dimensions to a characteristic length L of the body,
which is so chosen that the dimensionless derivative ∂u/∂x does not exceed
unity in the region under consideration. Further, the Reynolds number

ρU L UL
Re = =
µ ν

is assumed to be very large. Under these assumptions, and retaining the same
symbols for the dimensionless quantities as for their dimensional counterparts,
we can write the Navier–Stokes equations for plane flow as follows:
x-direction:
( )
∂u ∂u ∂u ∂p 1 ∂2u ∂2u
+u +v =− + + 2 (4.10)
∂t ∂x ∂y ∂x Re ∂x2 ∂y

1 1 1 δ 1/δ δ2 1 1/δ 2
y-direction:

( )
∂v ∂v ∂v ∂p 1 ∂2v ∂2v
+u +v =− + 2
+ 2 (4.11)
∂t ∂x ∂y ∂y Re ∂x ∂y

δ 1 δ δ 1 δ2 δ 1/δ

Continuity:

∂u ∂v
+ =0 (4.12)
∂x ∂y

1 1

The boundary conditions are:

u=v=0 for y=0 (i.e. at the wall)

u=U for y → ∞ (i.e. at the outer edge of the boundary layer)

ISTUDY
4.7. BOUNDARY LAYER EQUATIONS 211

With the assumptions made in this analysis, the dimensionless boundary layer
thickness, δ/L, for which we will retain the symbol δ, is very small compared
to unity, i.e. δ ≪ 1.
Let us now estimate the order of magnitude of each term and drop the
small terms to obtain the desired equations governing the boundary layer flow.
The order of ∂u/∂x is 1, thus in the continuity equation ∂v/∂y is of the order
1. Hence (since v = 0 at the wall), in the boundary layer, v is of the order
δ. Thus ∂v/∂x and ∂ 2 v/∂x2 are also of the order δ. Also, ∂ 2 u/∂x2 is of the
order 1. The orders of magnitudes are shown in Eqs. (4.10) to (4.12) under
each term.
Let us further assume that the unsteady acceleration ∂u/∂t is of the same
order as the convective term u ∂u/∂x which means that very sudden accelera-
tions, such as those occurring in very large pressure waves, are excluded. Also,
some of the viscous terms are of the same order of magnitude as the inertia
terms, at least in the proximity of the wall. Hence some of the second-order
derivatives of velocity such as ∂ 2 u/∂y 2 and ∂ 2 v/∂y 2 must be very large near
the wall. The component of velocity parallel to the wall increases from zero
at the wall to the value 1 in the freestream across the layer of thickness δ.
Thus,

∂u 1 ∂2u 1
∼ and ∼ 2
∂y δ ∂y 2 δ
where ∂v/∂y ∼ δ/δ ∼ 1 and ∂ 2 v/∂y 2 ∼ 1/δ. Inserting these terms in
Eq. (4.10), the viscous forces in the boundary layer can become of the same
order of magnitude as the inertia forces, only if the Reynolds number is of the
order 1/δ 2 , i.e.
1
= δ2 (4.13)
Re
The x-momentum equation can be simplified by neglecting ∂ 2 u/∂x2 with
respect to ∂ 2 u/∂y 2 . In the continuity equation both the terms are of equal
order and remain unaltered. In the y-momentum equation, ∂p/∂y is of the
order of δ. The pressure change across the boundary layer which would be
obtained by integrating the second equation is of the order δ 2 , i.e. it is very
small. Thus the pressure in a direction normal to the boundary layer is
practically invariant, it may be assumed to be equal to that at the outer edge
of the boundary layer where its value is determined by the potential flow. The
pressure in the freestream outside the boundary layer is said to be ‘impressed’
through the boundary layer, and it depends only on the coordinate x (flow
direction) and on time t.
At the outer edge of the boundary layer the parallel component of velocity
u becomes equal to the velocity in the outer flow, U (x, t). Since there is no
large velocity gradient here, the viscous terms in Eq. (4.10) vanish for large
values of Re, and consequently, for the outer flow, we have
∂U ∂U 1 ∂p
+U =− (4.14)
∂t ∂x ρ ∂x

ISTUDY
212 CHAPTER 4. BOUNDARY LAYER

Here again the symbols denote dimensionless quantities.


For a steady flow the equations can be simplified further if the pressure
depends only on x, i.e. if the derivative ∂p/∂x becomes dp/dx, so that
1
p+ ρU 2 = constant (4.15)
2
The boundary conditions for the external flow are nearly the same as for
frictionless flow. The boundary layer thickness is very small and the transverse
velocity component v is also very small at the edge of the boundary layer
(v/U ∼ δ/L). Thus the potential flow about the body under consideration, in
which the perpendicular velocity component is vanishingly small near the wall,
offers a very good approximation to the actual external flow. The pressure
gradient along the x-direction in the boundary layer can be obtained simply
by analyzing the Bernoulli Eq. (4.15) along a streamline at the well known
potential flow.
With the above implications and assumptions, we can write down the
simplified Navier–Stokes equations, known as the Prandtl’s boundary layer
equations. In dimensional form, for a steady, two-dimensional, incompressible,
viscous flow over a flat-plate, as shown in Figure 4.11, we have the boundary
layer equations as

FIGURE 4.11 Boundary layer on a flat plate.

∂u ∂u 1 ∂p ∂2u
u +v =− +ν 2 (4.16)
∂x ∂y ρ ∂x ∂x

∂u ∂v
+ =0 (4.17)
∂x ∂y
with boundary conditions

at y = 0: u = 0, v = 0; at y = ∞: u = U (x) (4.18)

We know that the boundary layer is very thin. Therefore, we can assume that

v ≪ u

ISTUDY
4.7. BOUNDARY LAYER EQUATIONS 213

and
∂ ∂

∂x ∂y

Applying these approximations to the Navier–Stokes equation in the y-direction,


we get
∂p
≈0 (4.19a)
∂y
or
p ≈ p(x) (4.19b)
In other words, the y-momentum equation can be neglected completely,
and the pressure p varies along the boundary layer and not through it. Equa-
tion (4.19) implies that pressure is approximately uniform across the boundary
layer. The pressure at the body surface is therefore equal to that at the edge
of the boundary layer. Therefore, the pressure can be found from the solutions
of the irrotational flow around the body. That is, the pressure in the outer
flow at the edge of the boundary layer is ‘impressed’ through the boundary
layer. This justifies the experimental fact that the observed surface pres-
sure on a model is approximately equal to that calculated with the irrotational
(potential) flow theory. Here it is important to note the assumption that ∂p/∂y
is vanishingly small, and this assumption is not valid if the boundary layer is
separated or about to separate from the wall or if the radius of curvature of
the surface is not large compared to the boundary layer thickness.
The pressure gradient ∂p/∂y in Eq. (4.16) is assumed to be known in
advance from the Bernoulli’s equation applied to the outer inviscid flow and
it is
∂p dp dU
= = −ρU (4.20)
∂x dx dx
where U = U∞ , and the subscript ∞ is dropped for simplicity. Further, in
Eq. (4.16),
∂2u ∂2u

∂x2 ∂y 2
Note that (a) neither term in the continuity equation (4.17) can be neglected
and (b) the continuity equation is always a vital part of any fluid flow analysis.
The governing Eqs. (4.16) and (4.17) reduce to Prandtl’s two boundary
layer equations as follows.
The continuity equation is

∂u ∂v
+ =0 (4.21a)
∂x ∂y

The momentum equation along the wall is

∂u ∂v dU 1 ∂τ
u +v ≈U + (4.21b)
∂x ∂y dx ρ ∂y

ISTUDY
214 CHAPTER 4. BOUNDARY LAYER

where
∂u
τ =µ for laminar flow
∂x
and
∂u
τ =µ − ρ u′ v ′ for turbulent flow
∂x
The additional stress −ρ u′ v ′ in this stress expression is known as the apparent
or virtual stress of turbulent flow or the Reynolds stress. This is due to tur-
bulent fluctuations which are given by the time-mean values of the quadratic
terms in the turbulent components. Since this stress is added to the ordinary
viscous term in laminar flow, it has similar influence on the course of the flow;
it is often said that it is caused by eddy viscosity. In general, the apparent
stresses far outweigh the viscous components and, consequently, the latter
may be omitted in many actual cases with a good degree of approximation.
• These equations are to be solved for u(x, y) and v(x, y), with the
assumption that U (x) is known from the inviscid-flow analysis, exte-
rior of the boundary layer.
• There are two boundary conditions, one on u and one on v. They are:

1. At y = 0 (at the wall), u = v = 0, due to no-slip condition.


2. At y = δ(x) exterior flow(outer stream), u = U (x) by patching.

• Unlike the Navier–Stokes equations which are mathematically elliptic,


the boundary layer equations are parabolic and are solved by beginning
at the leading edge and marching downstream as far as you like, stopping
at the separation point or earlier if it is preferred.
• The boundary layer equations have been solved for many interesting
cases of internal and external flows for both laminar and turbulent con-
ditions, utilizing the inviscid distribution U (x) appropriate to each flow.
Note that the estimation of the boundary layer thickness in Eq. (4.13) shows
that √
δ 1 ν
∼ = (4.22)
L (Re)L UL

The fact that δ ∼ ν, inferred from the exact solution of the Navier–Stokes
equations, is thereby confirmed. The numerical coefficient still missing in
Eq. (4.22), will turn out to be equal to 5 for the case of flat plate at zero
incidence, when L will mean the distance from the leading edge. Let us
have a closer look at one such solution, namely the flat-plate solution, in the
following section.

4.8 FLAT PLATE BOUNDARY LAYER


One of the classic and most often used solutions of boundary layer theory is
the semi-infinite flat plate solution.

ISTUDY
4.8. FLAT PLATE BOUNDARY LAYER 215

4.8.1 Laminar Flow


For laminar flow past a flat-plate, the boundary layer Eqs. (4.16)–(4.18)
can be solved exactly for u and v, assuming that the freestream velocity
U is constant(dU/dx = 0). The solution was given by Prandtl’s student
Blassius in his doctoral dissertation from Gottingen in 1908. With an in-
genious coordinate transformation, Blassius showed that the dimensionless
velocity profile u/U is a function only of the single composite dimensionless
( )1/2
U
variable y and expressed the profile as
νx
( )1/2
u U
= f ′ (η) η=y (4.23)
U νx
where the prime denotes differentiation with respect to η. It is important here
to understand how the dimensionless group for η is arrived at. Basically, the
velocity ratio is a function of ν, x and t and can be expressed as
u
= f (ν, x, t)
U
where the dimensions of the kinematic viscosity coefficient ν is L2 /T
√ . The only
non-dimensional group which can be formed from ν, x, t is x/ νt or some
power thereof. The quantity η, which is a combination of the old independent
variables x and t, is called a similarity variable. It now plays the role of a
single independent variable for the problem. Substitution of Eq. (4.23) into
the boundary layer Eq. (4.16) reduces the problem, after much algebra, to a
single third-order nonlinear ordinary equation of f as
1
f ′′′ + f f ′′ = 0 (4.24)
2
The boundary conditions become
at η=0 f (0) = 0, f ′ (0) = 0 (4.24a)

at η → ∞ f (∞) → 1.0 (4.24b)
In this example, both partial differential equations (4.16) and (4.17) have
been transformed into an ordinary differential equation by the substitution
1/2
η = (U/ν x) . The resulting differential equation is nonlinear and of the
third order. The three boundary conditions in Eq. (4.24) are, therefore, suffi-
cient to determine the solution completely. This is the Blassius equation for
which accurate solutions have been obtained only by numerical integration.
Some tabulated values of the velocity-profile shape f ′ (η) = u/U are given in
Table 4.2.
Since u/U approaches 1.0, only as η → ∞, it is customary to select the
boundary layer thickness δ as that point where u/U = 0.99. From Table 4.2,
it is seen that this occurs at η ≈ 5.0. Therefore, we can write
( )1/2
U
δ ≈ 5.0
νx

ISTUDY
216 CHAPTER 4. BOUNDARY LAYER

or
δ 5.0
=√ (4.25)
x (Re)x

TABLE 4.2 The Blassius velocity profile


√ √
η = y U/(ν x) u/U η = y U/(ν x) u/U
0.0 0.0 2.8 0.81152
0.2 0.06641 3.0 0.84605
0.4 0.13277 3.2 0.87609
0.6 0.19894 3.4 0.90177
0.8 0.26471 3.6 0.92333
1.0 0.32979 3.8 0.94112
1.2 0.39378 4.0 0.95552
1.4 0.45627 4.2 0.96696
1.6 0.51676 4.4 0.97587
1.8 0.57477 4.6 0.98269
2.0 0.62977 4.8 0.98779
2.2 0.68132 5.0 0.99155
2.4 0.72899 ∞ 1.00000
2.6 0.77246

With the velocity profile known, Blassius of course could also compute the
wall shear stress coefficient Cf as

0.664
Cf = 1/2
(4.26)
[(Re)x ]
This is the local skin friction coefficient for flow on one side of the plate.
If the flow remains laminar over a length L of the plate, the total skin friction
coefficient (for flow over both the surfaces of the plate) becomes

1.328
Cf = 1/2
(4.26a)
[(Re)x ]

Here (Re)x = ρ U x/µ denotes the Reynolds number with respect to the length
of the plate and the freestream velocity. This law of friction deduced by
Blassius is valid only in the region of laminar flow, i.e. for (Re)x < 5 × 105 to
106 . In the region of turbulent motion with (Re)x > 106 , the drag becomes
considerably greater than that given by Eq. (4.26a).

4.8.2 Boundary Layer Thickness


It is impossible to indicate the boundary layer thickness precisely, since the
influence of velocity on the boundary layer decreases asymptotically. The

ISTUDY
4.8. FLAT PLATE BOUNDARY LAYER 217

parallel component of velocity u tends asymptotically to the freestream value


U of the potential flow. If it is desired to define the boundary layer thickness
as the distance for which u = 0.99 U , then as seen from Table 4.2, η ≈ 5.0.
Hence the boundary layer thickness, as defined here, becomes

νx
δ≈5
U
A physically meaningful measure of the boundary layer thickness is the
displacement thickness δ ∗ , which was earlier introduced in Eq. (4.1). The
displacement thickness is that distance by which the external potential field
of the flow is displaced outwards as a consequence of the decrease in velocity
in the boundary layer. By Eq. (4.1), we have
∫ ∞ (
u)
δ∗ = 1− dy
y=0 U

With u/U from Eq. (4.23), we can express this as


√ ∫
∗ νx ∞
δ = [1 − f ′ (η)] dη
U η=0

This can be solved to obtain



νx
δ ∗ = 1.721
U
or
δ∗ 1.721
= 1/2
(4.27)
x [(Re)x ]
When Cf is converted into the dimensional form, we have the shear stress
as
1
τw (x) = ρ U 2 Cf (4.28)
2
Substituting for Cf , we can express the shear stress as

0.332 ρ1/2 µ1/2 U 3/2


τw (x) =
x1/2
The shear stress decreases inversely with x1/2 , a result of the thickening of the
boundary layer and the associated decrease in the velocity gradient because of
the boundary layer growth, and varies directly with velocity to the 1.5 power.
This is in contrast to laminar pipe flow, where τw ∝ U and is independent
of x. Note that the wall shear stress at the leading edge is predicted to be
infinite. Clearly, the boundary layer theory breaks down near the leading edge
where the assumption ∂/∂x ≪ ∂/∂y is not valid. The local Reynolds number
(Re)x in the neighbourhood of the leading edge is of the order 1, for which
the boundary layer assumptions are not valid.

ISTUDY
218 CHAPTER 4. BOUNDARY LAYER

If τw is substituted into Eq. (4.7a), the total drag force on one side of a
plate of length x and width b becomes
∫ x
D(x) = b τw (x)dx
0

Substituting for τw , we get the drag as

D(x) = 0.664 b ρ1/2 µ1/2 U 3/2 x1/2 (4.29)

The drag increases only as the square root of the plate length. The drag
coefficient is defined as

2 D(L) 1.328
CD = = = 2 Cf (L) (4.30)
ρ U 2b L [(Re)L ]
1/2

Thus, for the laminar flat plate flow, the drag coefficient is equal to twice the
value of the skin-friction coefficient at the trailing edge. This is the drag on
one side of the plate. Further, Karman pointed out that the drag could be
computed from the momentum relation given by Eq. (4.6). In dimensionless
form, Eq. (4.6) becomes

2 δ
u( u)
CD = 1− dy (4.31)
L 0 U U
This can be written in terms of θ at the trailing edge as
2 θ(L)
CD = (4.32)
L
The computation of θ from the profile u/U or from CD gives
θ 0.664
= 1/2
(laminar flat plate boundary layer) (4.33)
x [(Re)x ]

Since the boundary layer thickness δ is not properly defined, the momentum
thickness θ, being definite, is often used to correlate data taken for a variety
of boundary layers under differing conditions. The ratio of displacement
thickness δ ∗ to momentum thickness θ, called the dimensionless-profile shape
factor H, is also useful in integral theories.
For laminar flat plate boundary layer,
δ∗ 1.721
H= = = 2.59 (4.34)
θ 0.664
As we shall see, a large shape factor implies that the boundary layer separation
is about to occur.
The plot of Blassius velocity profile from Table 4.2 in the form of u/U
versus y/δ, shown in Figure 4.12, will highlight the importance of the simple
guess of the velocity profile in integral theory.

ISTUDY
4.8. FLAT PLATE BOUNDARY LAYER 219

FIGURE 4.12 Comparison of laminar and turbulent flat plate


velocity profiles.

It is seen from the plot in Figure 4.12 that the simple parabolic profile is
not far from the true Blassius profile, hence its momentum thickness is within
10 per cent of the true value. Instead of decreasing monotonically to zero, the
turbulent profiles are flat and then drop off sharply at the wall. As it rightly
be guessed, they follow the logarithmic-law shape and thus can be analyzed
by momentum-integral theory if this shape is properly represented.
The flow in a boundary layer along a wall becomes turbulent when the
external velocity is sufficiently large. Experimental investigation into the tran-
sition from laminar to turbulent flow in a boundary layer revealed that the
transition becomes most clearly discernible by a sudden and large increase in
the boundary layer thickness and in the shearing stress near the wall, when x
is √
replaced by the local length l, the dimensionless boundary layer thickness
δ/ (ν x)/U becomes constant for laminar flow, and is approximately equal
to 5. Boundary layer thickness variation with the Reynolds number based on
the current length x along a plate in parallel flow at zero incidence is shown
in Figure 4.13.
At (Re)x > 3.2 × 105 , a sharp increase in the boundary layer thickness is
clearly seen and an identical increase would also be experienced by the wall
shear stress. The sudden increase in these quantities denotes that the flow
has changed from laminar to turbulent. The (Re)x based on the local length
x is related to the Reynolds number (Re)δ = U δ/ν based on the boundary

ISTUDY
220 CHAPTER 4. BOUNDARY LAYER

FIGURE 4.13 Boundary layer thickness variation in laminar and


turbulent regimes of flow.
(Source: Schlichting, H., Boundary Layer Theory, McGraw-Hill,
New York, 1955).

layer thickness through the equation



(Re)δ = 5 (Re)x

Hence the critical Reynolds number for the plate based on local length x,
( )
Ux
(Re)cri = = 3.2 × 105
ν cri

becomes 2800 when the boundary layer thickness δ is taken as the length in the
Reynolds number expression. The numerical value of (Re)cri depends on the
amount of disturbance in the external flow, and the value (Re)cri = 3.2 × 105
should be regarded as the lower limit. With the exceptionally disturbance-free
external flow, (Re)cri as high as 106 has been obtained.

EXAMPLE 4.1 A flat plate of length 0.8 m and width 1.9 m is kept
in sea-level air stream flowing at a velocity of 5.3 m/s. Assuming a linear
velocity profile for the boundary layer over the plate, develop an expression

ISTUDY
4.8. FLAT PLATE BOUNDARY LAYER 221

for the variation of wall shear stress with distance along the plate. Also,
obtain an expression for the total skin-friction drag on the plate and then
evaluate the skin-friction drag.

Solution The velocity profile is linear. Thus,


u y
= =η
U δ

Hence the momentum thickness becomes


∫ 1
u( u ) (y)
θ=δ 1− d
0 U U δ
∫ 1
=δ η (1 − η) dη
0
[ ]1
η2 η3 1
=δ − = δ
2 3 0 6

Therefore, the wall shear stress becomes

du U d(u/U )
τw = µ y=0

dy δ d(y/δ)
U 1 dδ
=µ = ρ U2
δ 6 dx

Separating the variables and integrating, we get


µ
δ dδ = 6 dx
ρU

δ2 µ
=6 x+c
2 ρU

But δ = 0 at x = 0, thus, c = 0. Therefore,


√ ( )1/2
12 µ x µx
δ= = 3.46
ρU ρU

or

δ 3.46
=√
x (Re)x

ISTUDY
222 CHAPTER 4. BOUNDARY LAYER

Therefore,

U 1 µ1/2 ρ1/2 U 3/2


τw = µ =
δ 3.46 x1/2



= 0.289 (Re)x
x

Thus, the skin friction drag becomes


∫ ∫ L
Df = τw dAs = τw b dx
0

∫ L ∫ L

=b ρ U2 dx = b ρ U 2 dθ
0 dx 0

= ρ U 2 b θL


where θL = δL /6, δL = 3.46 L/ (Re)L , b and L are the width and length of
the plate, respectively. The Reynolds number is

UL 5.3 × 0.8
(Re)L = = = 2.9 × 105
ν 1.46 × 10−5
Thus,

3.46 × 0.8
δL = √ = 5.14 mm
2.9 × 105
δL
θL = = 0.857 mm
6
Df = ρ U 2 b θL = 1.226 × 5.32 × 1.9 × (0.857 × 10−3 )

= 0.0561 N

EXAMPLE 4.2 The velocity profile in a laminar boundary layer flow at


zero pressure gradient is approximated as

u 3 ( y ) 1 ( y )3
= −
U 2 δ 2 δ
δ
Obtain the expressions for and Cf , using the momentum integral equation.
x

ISTUDY
4.8. FLAT PLATE BOUNDARY LAYER 223

Solution The shear stress can be expressed as



dθ d δ
u( u)
τw = ρ U 2 = ρ U2 1− dy
dx dx 0 U U

dδ 1
u( u ) (y)
= ρU 2
1− d
dx 0 U U δ
∫ 1 ( )( )
dδ 3 1 3 1 3
= ρU 2
η − η3 1 − η + η dη
dx 0 2 2 2 2
y
where = η. On integration, this yields
δ
39 dδ
τw = ρ U2
280 dx
The shear stress can also be expressed as

∂u U ∂(u/U )
τw = µ y=0

∂y δ ∂(y/δ)
( )
3 η3
∂ η−
U 2 2

δ ∂η

3 U
= µ
2 δ
Thus, we have
3 U 39 dδ
µ = ρ U2
2 δ 280 dx
280 3 µ
δ dδ = × dx
39 2 ρU

δ2 µ
= 10.77 x+c
2 ρU
But at x = 0, δ = 0, therefore, c = 0. Thus,

δ2 21.54
=
x2 (Re)x
or

δ 4.64
= √
x (Re)x

ISTUDY
224 CHAPTER 4. BOUNDARY LAYER

The skin friction coefficient is


τw µU 1
Cf = 1 =3
2 ρ U2 δ ρ U2

(Re)x 1
=3
4.64 (Re)x

0.647
= √
(Re)x

4.9 TURBULENT BOUNDARY LAYER FOR


INCOMPRESSIBLE FLOW ALONG
A FLAT PLATE
The majority of boundary layers encountered in engineering practice are tur-
bulent over most of their length. Thus, the turbulent boundary layer analysis
is regarded as of greater fundamental importance than that of the laminar
boundary layer. But there is no exact theory available for turbulent flat plate
flow, although there are many elegant computer solutions of the boundary
layer equations using the various empirical models for turbulent eddy viscos-
ity. The most widely accepted result is simply an integral analysis similar to
our study of laminar-profile approximation [Eq. (4.6)].
The momentum Eq. (4.8) may be applied to the turbulent boundary layer
as no limiting assumptions are made in its derivation. However, a new relation
for the velocity profile up through the boundary layer will have to be found
and the shear stress will no longer be obtained simply from the product of
fluid viscosity and the gradient of the velocity profile. Because of the basic
similarity between the development of boundary layer within circular pipes
and over flat plates, Prandtl suggested that the results from pipe can be
applied to the analysis of flat plate turbulent boundary layers. We know that
the boundary layer growth in pipes is limited to the pipe radius R, so that
u = U at r = R, and the mean velocity in turbulent pipe flow is known to be
about 0.8U . The velocity distribution in such a flow is adequately represented
by the Prandtl power law,
u ( y )n
= (4.35)
U δ
where n = 17 for (Re)x < 107 . Obviously, this profile breaks down at the
wall, where y = 0. But the presence of laminar sublayer makes the velocity to
decrease linearly to zero at the wall, this profile being tangential to the power
law.
To develop the analogy between the flat plate and pipe flow, it is necessary
to appreciate that δ = R in the fully-developed region and also to develop

ISTUDY
4.9. TURBULENT BOUNDARY LAYER FOR... 225

some relation for τw to replace


∂u
τw = µ
∂y
which no longer applies.
Blassius proposed that for smooth pipes the shear stress at the wall could
be expressed by
1
τw = f ρ u2 (4.36)
2
where u is the mean velocity of the flow, which is equal to 0.8 U and f is an
empirical constant known as friction factor, which is a function of the flow
Reynolds number based on pipe diameter d and the ratio of wall roughness-
2
to-pipe diameter. Thus, τw = f 21 ρ (0.8U ) and, as Blassius developed the
expression
0.079 0.079
f= 1/4
= 1/4
(Re) (ρ u d/µ)
Thus, for smooth pipes, we have
( )1/4
1 2 µ
τw = ρ (0.8 U ) 0.079
2 ρ 0.8 U 2R
If δ = R, this becomes
( )1/4
2 µ
τw = 0.0225 ρ U (4.37)
ρU δ
As the assumption of zero pressure gradient has been made, Eq. (4.8) can be
applied. Thus,
∫ δ (
2 d u u)
τw = ρ U 1− dη
dx 0 U U
i.e. ∫
dδ 1 ( )
τw = ρ U 2 1 − η 1/7 η 1/7 dη (4.38)
dx 0

where u/U = (y/δ)1/7 = η 1/7 . Therefore,


7 dδ
τw = ρ U2 (4.39)
72 dx
Equating Eqs. (4.37) and (4.39), we get
( )1/4
µ
δ 1/4 dδ = 0.234 dx
ρU
Integrating this, we get
( )1/4
4 5/4 µ
δ = 0.234 x+c
5 ρU

ISTUDY
226 CHAPTER 4. BOUNDARY LAYER

Now, assuming the boundary layer to be turbulent right from the leading edge
of the plate, which is reasonable if the plate is long compared to the length of
the laminar boundary layer, then δ = 0 at x = 0 and c = 0. Hence,
( )1/4
5/4 µ
δ = 0.292 x
ρU
or
0.37x
δ= (4.40)
[(Re)x ]1/5

Thus, the thickness δ of a turbulent boundary layer increases as x4/5 , which is


far faster than the laminar boundary layer thickness which increases as x1/2 .
The shear stress on the flat surface may be determined by eliminating δ
between Eqs. (4.37) and (4.40). Thus, the shear stress becomes
( )1/5
2 µ
τw = 0.029 ρ U
ρ Ux
and the skin-friction force acting on the wall becomes
∫ l
F = τw dx (per unit width)
0

where l is the length of the plate. Now, substituting for τw , we get


∫ l ( )1/5
µ
F = 0.029 ρ U 2
x−1/5 dx
0 ρU

[ ( )1/5 ]l
2 µ x−4/5
= 0.029 ρ U
ρU 4/5
0

( )1/5
2 µ
= 0.036 ρ U l
ρU l
The skin-friction coefficient is
F
Cf = 1 (per unit width)
2 ρ U2 l
i.e.
0.072
Cf = (4.41)
[(Re)l ]1/5
This expression is valid for Reynolds numbers up to 107 , but the experimental
results indicate that a better approximation is given by
0.074
Cf = (4.42)
[(Re)l ]1/5

ISTUDY
4.10. FLOWS WITH PRESSURE GRADIENT 227

Prandtl has suggested subtracting the length of the laminar layer, resulting
in an expression
0.074 1700
Cf = 1/5

[(Re)l ] (Re)l
which can be applied from (Re)l = (5 × 105 ) to 107 .
To extend the Reynolds number range further, Schlichting employed the
logarithmic velocity distribution for a pipe under turbulent flow condition,
resulting in a semi-empirical relation

0.455
Cf = 2.58 (4.43)
[log10 (Re)l ]

Equation (4.43) can be used for Reynolds numbers greater than 107 . For
Re < 107 , Eq. (4.43) gives the values for Cf that are very close to those given
by Eq. (4.42), and consequently engineers commonly use Eq. (4.43) over
the entire range of Reynolds numbers above 500,000. Comparing Eqs.(4.41)
and (4.26a), it is seen that the skin-friction is proportional to the 95 power of
velocity of the main stream and the 54 power of plate length for the turbulent
boundary layer, compared to 32 and 12 powers, respectively, for the laminar
boundary layer.

4.10 FLOWS WITH PRESSURE GRADIENT


In flows with zero pressure gradient, e.g. flat plate flow, the point of in-
flection (PI) is at the wall itself. In flows with adverse pressure gradient, a
point of inflection occurs in the boundary layer and its distance from the wall
increases with the strength of the adverse pressure gradient. The flow profile of
Figure 4.14, usually occurs in sequence as the boundary layer progresses along
the wall of a body.
Examine the flow through a convergent-divergent passage shown in
Figure 4.14.

FIGURE 4.14 Viscous flow through a convergent-divergent duct.

ISTUDY
228 CHAPTER 4. BOUNDARY LAYER

The nozzle flow is a flow with favourable gradient and never separates, nor
does the throat flow where the pressure gradient is approximately zero. But
in diffusers the velocity decreases and the pressure increases, and an adverse
gradient is created. When the diffuser angle is too large and the adverse
pressure gradient is excessive, the boundary layer will separate at one or both
walls, with back flow, increased losses, and poor pressure recovery. In diffuser
literature this condition is called diffuser stall. This is usually referred to as
the boundary layer separation. At this stage we should note that the boundary
layer theory can compute only up to the separation point and after which it
is invalid.

4.11 LAMINAR INTEGRAL THEORY


Both laminar and turbulent theories can be developed from Karman’s general
two-dimensional boundary layer integral relation, which can be expressed in
terms of U (x) as

τw 1 dθ θ dU
= Cf = + (2 + H) (4.44)
ρ U2 2 dx U dx

where θ(x) is the momentum thickness and H(x) = δ ∗ (x)/θ(x) is the shape
factor. Also, we know that negative dU/dx is equivalent to positive dp/dx, i.e.
an adverse pressure gradient. We can integrate Eq. (4.44) to determine θ(x)
for a given U (x) if we correlate Cf and H with momentum thickness. This
has been done by examining typical velocity profiles of laminar and turbulent
boundary layer flows for various pressure gradients.
Consider the velocity profiles shown in Figure 4.15, illustrating that the
shape factor H is a good indicator of pressure gradient.
The higher the H, the stronger the adverse pressure gradient, and separa-
tion occurs approximately at

H ≈ 3.5 for laminar flow

and
H ≈ 2.4 for turbulent flow

From Figure 4.15 it is seen that the laminar profile exhibits the S-shape
and a point of inflection(PI) with an adverse gradient. But in the turbulent
profile the point of inflection is typically buried deep within the thin viscous
sublayer.
For laminar flow, a simple and effective method was developed by Thwaits,
who found that Eq. (4.44) can be correlated by a simple dimensionless
momentum thickness variable λ, defined as

θ2 dU
λ= (4.45)
ν dx

ISTUDY
4.11. LAMINAR INTEGRAL THEORY 229

FIGURE 4.15 Velocity profile with pressure gradients for laminar flow.

Using a straight-line fit to his correlation, Thwaits was able to integrate


Eq. (4.44) in a closed form, to result in

0.45 ν x 5
θ2 = θ02 + U dx (4.46)
U6 0

where θ0 is the momentum thickness at x = 0 (usually taken to be zero).


Separation (Cf = 0) was found to occur at a particular value of λ. This value
is
λ = − 0.09 (separation) (4.47)
τw θ
Finally, Thwaits correlated the values of the dimensionless shear, S = ,
ρU
with λ, and his graphed result can be curve-fit as follows.
τw θ 0.62
S(λ) = ≈ (λ + 0.009) (4.48)
ρU
This parameter is related to skin friction by the identity
1
S≡ Cf (Re)θ (4.49)
2
Equations (4.46) to (4.49) constitute a complete theory for the laminar bound-
ary layer with variable U (x), with an accuracy of ±10% compared with the

ISTUDY
230 CHAPTER 4. BOUNDARY LAYER

exact digital computer solution of the laminar boundary layer Eq. (4.21).

EXAMPLE 4.3 Water flows at a speed of 1 m/s over a flat plate of length
1 m in the flow direction. The boundary layer is tripped to make it turbu-
lent at the leading edge. Assuming 17 power turbulent velocity profile, find
the boundary layer thickness, the displacement thickness, and the wall shear
stress at the trailing edge of the plate. Solve the same problem if the flow
over the plate is laminar. The kinematic viscosity of water is 10−6 m2 /s.

Solution At the trailing of the plate, x = L = 1, the Reynolds number is


UL 1×1
(Re)L = = −6 = 106
ν 10
This value is greater than the flat plate critical Reynolds number of 5×105 and
hence the flow is turbulent. Therefore, by Eq. (4.40), we have the boundary
layer thickness as
0.37 L
δ ≈ 1/5
[(Re)L ]
0.37 × 1
= 1/5
= 0.02335 m
(106 )

= 23.35 mm
The displacement thickness (with u/U = (y/δ)1/7 ) = η 1/7 ) is given by
Eq. (4.2) as
δ 23.35
δ∗ = =
8 8

= 2.92 mm
The wall shear stress is given by
( )1/5
2 µ
τw = 0.029 ρU
ρU x
( )1/5
1
= 0.029 × 10 × 1 ×
3 2
106

= 1.83 N/m2

For laminar flow, by Eq. (4.25), we have


5 5
δL = √ ×L= √
ReL 106

= 5 mm

ISTUDY
4.11. LAMINAR INTEGRAL THEORY 231

By Eq. (4.33), we have


θ 0.664
= 1/2
L [(Re)L ]
By Eq. (4.34), we have

δ ∗ = 2.59 θ

2.59 × 0.664 2.59 × 0.664


= 1/2
= √
[(Re)L ] 106

= 1.72 mm

By Eq. (4.26), we have


0.664
Cf = 1/2
[(Re)L ]
Therefore,
1
τw = Cf ρ U2
2
0.664 1
=√ × × 103 × 12
106 2

= 0.332 N/m2

EXAMPLE 4.4 (a) Determine the friction drag acting on one side of a
smooth flat plate of length 0.5 m and width 0.15 m, placed longitudinally in
an air stream of 1 m/s at sea level. (b) Find the boundary layer thickness
and the shear stress at the trailing edge of the plate.

Solution (a) At standard sea level the atmospheric temperature is 15◦ C =


288 K. Therefore, the viscosity of air becomes

(288)3/2
µ = 1.46 × 10−6
288 + 111

= 17.88 × 10−6 kg/(m s)

The air density is


p 101325
ρ= = = 1.226 kg/m3
RT 287 × 288
The Reynolds number at the plate end is
ρU L 1.226 × 1 × 0.5
Re = = = 34, 284
µ 17.88 × 10−6

ISTUDY
232 CHAPTER 4. BOUNDARY LAYER

This is less than 500,000 and hence the flow is laminar. Therefore, the skin
friction coefficient becomes

1.328 1.328
Cf = √ =√ = 0.00717
Re 34284
The friction drag acting on one side of the plate becomes

1 S
Ff = ρ U 2 Cf
2 2

(0.5 × 0.15)
= 0.5 × 1.226 × 12 × × 0.00717
2

= 1.65 × 10−4 N

where S is the surface area of the plate.

(b) The boundary layer thickness at the trailing edge of the plate is

5L 5 × 0.5
δ=√ =√ = 13.5 mm
(Re)L 34284

The shear stress at the plate end is



ρ µ 3/2
τw = 0.332 U
L

ρU L µU
= 0.332
µ L

√ µU
= 0.332 Re
L
√ 17.88 × 10−6 × 1
= 0.332 34, 284 ×
0.5

= 0.0022 N/m2

EXAMPLE 4.5 A submarine has been approximated to a rectangular box


of length 360 m, width 70 m and height 25 m. If the submarine is travelling at
a speed of 24 km/h in sea water with kinematic viscosity, ν = 1.4 × 10−6 m2 /s
and density 1020 kg/m3 , determine the skin friction drag of the submarine
and the power required to overcome the drag.

ISTUDY
4.11. LAMINAR INTEGRAL THEORY 233

Solution The Reynolds number of the flow over the submarine, based on
its length is

UL (24/3.6) × 360
(Re)L = =
ν 1.4 × 10−6

= 1.714 × 109

Considering each surface of the submarine to be a flat plate, we have the skin
friction coefficient by Eq. (4.43) as

0.455
Cf = 2.58
[log10 (1.714 × 109 )]

= 0.00147

The skin friction drag is given by

1
Df = Cf As ρ U2
2
where As is the surface area. For the given vessel, we have

As = 2 [(360 × 70) + (360 × 25)] = 68400 m2

Note that the surface areas of the face and the base of the box are neglected
here. Thus,

1 2
Df = 0.00147 × 68400 × × 1020 × (24/3.6)
2
= 2.279 MN

The power required to overcome this drag is

P = D U = 2.279 × (24/3.6)

= 15.19 MW

EXAMPLE 4.6 A viscous fluid flows over a flat plate such that the bound-
ary layer thickness at a distance 1.3 m from the leading edge is 12 mm.
Assuming the flow to be laminar, determine the boundary layer thickness at
a distance of (a) 0.2 m, (b) 2.0 m, and (c) 20 m from the leading edge.

Solution For laminar boundary layer, the boundary layer thickness is

5x
δ= 1/2
[(Re)x ]

ISTUDY
234 CHAPTER 4. BOUNDARY LAYER

where x is the distance from the leading edge. Therefore, the Reynolds number
at x = 1.3 m is
52 x2 52 × 1.32
Re = =
δ2 0.0122

= 293, 403
i.e.
ρ U (1.3)
= 293, 403
µ
(a) Assuming ρ, U , and µ to remain constant, at x = 0.2 m, we have
0.2
(Re)0.2 = 293, 403 ×
1.3

= 45, 139
Therefore, the boundary layer thickness at x = 0.2 m becomes
5 × 0.2
δ0.2 = √
45, 139

= 4.71 mm
(b) At x = 2 m, the Reynolds number is
2
(Re)2 = 293, 403 ×
1.3

= 451, 389
Therefore,
5×2
δ2 = √
451, 389

= 14.88 mm
(c) At x = 20 m, the Reynolds number is
20
(Re)20 = 293, 403 ×
1.3

= 4, 513, 892
Therefore,
5 × 20
δ20 = √
4, 513, 892

= 47.07 mm

ISTUDY
4.12. SUMMARY 235

4.12 SUMMARY
Boundary layer may be defined as the layer adjacent to a solid boundary in
which the flow velocity increases from zero to freestream value.
The boundary layer thickness may also be defined as “the distance from
the solid boundary within which the local value of the flow velocity increases
from zero at the wall (y = 0) to 0.99 of the freestream value at the edge of
the layer (y = δ)”.
The displacement thickness δ ∗ is the distance by which the wall would have
to be displaced outwards in a hypothetical frictionless flow so as to maintain
the same mass flux as in the actual flow, i.e.
∫ δ ( )
∗ ρu
δ = 1− dy
0 ρ∞ U∞

The momentum thickness θ is the thickness of a layer of fluid of velocity


U∞ for which the momentum flux is equal to the deficit of momentum flux
through the boundary layer, i.e.
∫ δ ( )
ρu u
θ= 1− dy
0 ρ∞ U∞ U∞

The kinetic energy thickness ϑ is the distance through which the surface
would have to be displaced in order that, with no boundary layer, the total
flow kinetic energy at the station considered would be the same as that which
would actually occur. i.e.
∫ δ [ ( )2 ]
ρu u
ϑ= 1− dy
0 ρ∞ U∞ U∞

In laminar flow the layers of fluid elements slide smoothly over one an-
other and there is little interchange of fluid mass between the adjacent layers.
The shearing traction which is developed due to the velocity gradient is thus
entirely due to the viscosity of the fluid, i.e. the momentum exchange between
the adjacent layers is on a molecular scale only.
In turbulent flow, considerable random motion exists in the form of
velocity fluctuations both along the mean direction of flow and the direction
perpendicular to it. As a result of the latter, there is appreciable transport of
mass between the adjacent layers.
Throughout the boundary layer, the shearing stress τ is given by
∂u
τ =µ
∂y
and at the wall the shearing stress becomes
( ) ( )
∂u ∂u
τw = µw = µw (say)
∂y y=0 ∂y w

ISTUDY
236 CHAPTER 4. BOUNDARY LAYER

For flat-plate flow the boundary layer thickness δ can be expressed as

δ 5.0
≈ 1/2
for laminar boundary layer
x [(Re)x ]
δ 0.16
≈ 1/7
for turbulent boundary layer
x [(Re)x ]

For solving external flows, basically three techniques are used. They are:

1. Numerical methods

2. Experimental methods

3. Boundary layer theory

By Newton’s second law, the momentum equation for flow through a control
volume can be expressed as

dV d
F = ma = m = (mV )
dt dt
By Reynolds transport theorem, the linear-momentum relation for a deformable
control-volume becomes
∑ (∫ ∫ ∫ ) ∫∫
d d
(mV )system = F = V ρ dV + V ρ (Vr · n) dA
dt dt CV CS

For a fixed control-volume, the relative velocity Vr = V , and we can use the
partial derivative to express the momentum equation as

∑ (∫ ∫ ∫ ) ∫∫

F = V ρ dV + V ρ (V · n) dA
∂t CV CS

This is called the momentum-integral relation.


The drag force on the plate is given by the following momentum integral
across the exit plane.
∫ δ(x)
D(x) = ρ b u (U − u) dy
0

In the convenient form of the momentum thickness θ, the drag becomes

D(x) = ρ b U 2 θ

The drag can also be expressed as


∫ x
D(x) = b τw (x) dx
0

ISTUDY
4.12. SUMMARY 237

The momentum-integral relation for the flat-plate boundary layer is



dθ d δ
u( u)
τw = ρ U 2 = ρU 2 1− dy
dx dx 0 U U

This is the momentum equation, and is valid for both laminar and turbulent
flows.
In dimensional form, for a steady, two-dimensional, incompressible, viscous
flow over a flat-plate, the boundary layer equations are

∂u ∂u 1 ∂p ∂2u
u +v =− +ν 2
∂x ∂y ρ ∂x ∂x

∂u ∂v
+ =0
∂x ∂y
with boundary conditions

at y = 0: u = 0, v = 0; at y = ∞: u = U (x)

Unlike the Navier–Stokes equations which are mathematically elliptic, the


boundary layer equations are parabolic and are solved by beginning at the
leading edge and marching downstream as far as we like, stopping at the
separation point or earlier if preferred.
For laminar flow past a flat-plate,

δ 5.0
= 1/2
x [(Re)x ]

and the wall shear stress coefficient is

0.664
Cf = 1/2
[(Re)x ]

This is the local skin friction coefficient for flow on one side of the plate. For
flow over both the surfaces of the plate it becomes

1.328
Cf = 1/2
[(Re)x ]

δ∗ 1.721
= 1/2
x [(Re)x ]

The total drag force on one side of a plate of length x and width b becomes

D(x) = 0.664 b ρ1/2 µ1/2 U 3/2 x1/2

ISTUDY
238 CHAPTER 4. BOUNDARY LAYER

The drag coefficient is

2 D(L) 1.328
CD = 2
= 1/2
= 2 Cf (L)
ρU bL [(Re)L ]

Thus, for the laminar flat plate flow, the drag coefficient is equal to twice the
value of the skin-friction coefficient at the trailing edge. This is the drag on
one side of the plate.
Also,
2 θ(L)
CD =
L

Computation of θ from the profile u/U or from CD gives

θ 0.664
= 1/2
laminar flat plate boundary layer
x [(Re)x ]

The majority of boundary layers encountered in engineering practice are


turbulent over most of their length. Thus, the turbulent boundary layer anal-
ysis is regarded as of greater fundamental importance than that of the laminar
boundary layer. But there is no exact theory available for turbulent flat plate
flow, although there are many elegant computer solutions of the boundary
layer equations using the various empirical models for turbulent eddy viscos-
ity. The most widely accepted result is simply an integral analysis similar to
our study of laminar-profile approximation [Eq. (4.6)].
The velocity distribution in such flow is adequately represented by the
Prandtl power law,
u ( y )n
=
U δ
where n = 71 for (Re)x < 107 .
Blassius proposed that for smooth pipes the shear stress at the wall could
be expressed by
1
τw = f ρ u2
2
where u is the mean velocity of the flow, which is equal to 0.8 U and f is an
empirical constant known as friction factor, which is a function of the flow
Reynolds number based on pipe diameter d and the ratio of wall roughness-
2
to-pipe diameter. Thus, τw = f 21 ρ (0.8U ) and, as Blassius developed the
expression
0.079 0.079
f= 1/4
= 1/4
(Re) (ρ u d/µ)
Thus, for smooth pipes, we have
( )1/4
1 2 µ
τw = ρ (0.8 U ) 0.079
2 ρ 0.8 U 2R

ISTUDY
4.13. PROBLEMS 239

The shear stress on the flat surface may be determined by eliminating δ


between Eqs. (4.37) and (4.40). Thus, the shear stress becomes
( )1/5
2 µ
τw = 0.029 ρ U
ρ Ux

and the skin-friction force acting on the wall becomes


∫ l
F = τw dx (per unit width)
0

The empirical relation skin-friction coefficient for logarithmic velocity distri-


bution for a pipe under turbulent flow condition is

0.455
Cf = 2.58
[log10 (Re)l ]

In flows with zero pressure gradient, e.g. flat plate flow, the point of
inflection is at the wall itself. In flows with adverse pressure gradient, a
point of inflection occurs in the boundary layer and its distance from the wall
increases with the strength of the adverse pressure gradient.
The nozzle flow is a flow with favourable gradient and never separates, nor
does the throat flow where the pressure gradient is approximately zero. But
in diffusers the velocity decreases and the pressure increases, and an adverse
gradient is created. When the diffuser angle is too large and the adverse
pressure gradient is excessive, the boundary layer will separate at one or both
walls, with back flow, increased losses, and poor pressure recovery. In diffuser
literature this condition is called diffuser stall. This is usually referred to as
boundary layer separation. At this stage we should note that the boundary
layer theory can compute only up to the separation point and after which it
is invalid.
What we saw in this chapter is just a glimpse of boundary theory, to
get an idea about this important aspect of fluid flow analysis. For in-depth
information, the reader should consult books specializing on this topic, such
as Boundary Layer Theory by Schlichting.

4.13 PROBLEMS
4.1 Air flows over a flat plate. At a given location along the plate the
boundary layer thickness is δ = 50 mm. At this location, what would be the
boundary layer thickness if it were defined as a distance from the plate where
the velocity is 97 per cent of the freestream velocity rather than 99 per cent?
Assume the flow to be laminar.
[Ans. 42.8 mm]

ISTUDY
240 CHAPTER 4. BOUNDARY LAYER

4.2 Consider an incompressible boundary layer with freestream velocity


U∞ = constant, as shown in Figure P4.2. Assuming a velocity distribution
( )
y y2
u = U∞ 2 − 2 ,
δ δ
determine the wall shearing stress τ0 (x) and the boundary layer thickness
δ(x). The given parabolic velocity profile is an approximation to a laminar
boundary layer profile.

FIGURE P4.2

2 µ 5.484 x
[Ans. τ0 = 0.365 ρU∞ , δ= 1/2
]
ρU∞ x [(Re)x ]
4.3 A laminar boundary layer formed on one side of a flat plate of length
l produces a drag D. How much must the plate be shortened if the drag on
the new plate is to be D/4? Assume that the upstream velocity remains the
same. Explain your answer physically.
[Ans. l/16]
4.4 A hut to serve as temporary housing near a seashore has to be designed.
The hut may be considered to be a closed (without leak) semi-cylinder, whose
radius R is 5 m, mounted on tie-down blocks, as shown in Figure P4.4. The
viscous effects are neglected and the flow over the top of the hut is identical
to the flow over a cylinder for 0 ≤ θ ≤ π. When calculating the flow over
the upper surface of the hut, the presence of the air space under the hut is
neglected. The air under the hut is at rest and the pressure is equal to the
stagnation pressure pt . What are the net lift and drag forces per unit depth of
the hut? The wind speed is 50 m/s and the stagnation freestream properties
are those of the standard sea-level condition. Also, find the lift and drag
coefficients.

[Ans. 40833.33 N/m, 0, 2.67, 0]


4.5 Calculate the skin friction coefficient Cf for a flow over the top surface of
a flat plate at a Reynolds number 2.5 × 105 , if the flow is (a) entirely laminar,
1/7
and (b) turbulent assuming the power law u/U = (y/δ) .

ISTUDY
4.13. PROBLEMS 241

FIGURE P4.4

[Ans. (a) 0.00265, (b) 0.006]


4.6 Water at a speed of 0.5 m/s flows over a flat plate of length 6 m.
Assuming the boundary layer to be laminar, determine the boundary layer
thickness and the wall shear stress (a) at the centre and (b) at the trailing
edge of the plate. Take the viscosity of water as 1.12 × 10−3 kg/m·s.
[Ans. (a) 0.0013 m, 0.0718 N/m2 , (b) 0.0183 m, 0.0508 N/m2 ]
4.7 A thin flat plate of length 0.3 m and width 1 m is installed in a water
tunnel as a splitter. The freestream speed is 1.6 m/s and the boundary layer on
the plate is laminar with a parabolic velocity profile. Neglecting the pressure
drag, determine the total viscous drag force on the plate.
[Ans. 1.6 N]
4.8 A thin flat plate of length L = 0.3 m and width b = 1 m is installed in
a water tunnel as a splitter. The freestream speed is 2 m/s at 20◦ C and the
velocity profile is approximated as

u ( y ) ( y )2
=2 −
U δ δ
and the boundary layer thickness over the plate at this condition is given by
1/2
δ/x = 5.48/[(Re)x ] . Show that the total drag force on one side of the plate
can be expressed as D = ρ U 2 θL b. Calculate (a) the momentum thickness θL
and (b) the drag D.
[Ans. (a) 0.283 mm, (b) 1.132 N]
4.9 An aircraft of wing span 12 m and average chord 2 m flies at a speed of
200 km/h at an altitude where the air density is 1.2 kg/m3 and the kinematic
viscosity is 1.5 × 10−5 m2 /s. Assuming that the wing skin friction is the same
as that on a smooth flat plate of same dimensions, calculate (a) the frictional
drag and (b) the power required to overcome this frictional drag.
[Ans. (a) 280.3 N, (b) 15.57 kW]
4.10 A rectangular flat plate of length 200 mm and width 1800 mm is in
a uniform air stream of speed 30 m/s at sea level condition. (a) Determine
the chordwise distribution of the skin friction coefficient and the displacement
thickness. (b) What is the drag coefficient for the plate?

ISTUDY
242 CHAPTER 4. BOUNDARY LAYER

4.634 × 10−4 √
[Ans. (a) √ , 0.0012 x, (b) 0.004148]
x
4.11 Air at sea level condition with speed ue = 2.4 m/s flows past a flat
plate. Assuming streamwise velocity component for the laminar boundary
layer over the plate as
y
u = ue
δ

where δ = 0.0125 x, find the streamwise distribution of the displacement
thickness and the skin friction coefficient.
√ 1.644 × 10−3
[Ans. 0.00625 x, √ ]
x
4.12 A thin flat plate 55 cm by 110 cm is immersed in a 6 m/s stream of
SAE 10 oil with kinematic viscosity 1.20 × 10−4 m2 /s and density 870 kg/m3 ,
at 20◦ C. Compute the total skin friction drag if the stream is parallel to (a)
the long side and (b) the short side and compare and discuss the magnitudes
of the drag for these orientations.
[Ans. (a) 107.3 N, (b) 151.74 N]

ISTUDY
Chapter 5

Vortex Theory

5.1 INTRODUCTION
Before getting into the dynamics of vortex motion, it is essential to have a
thorough understanding of rotational and irrotational flows. Translation and
rotation are the two types of basic motion in a fluid flow. These two may exist
independently or simultaneously. When they coexist, they may be considered
as one superimposed on the other. It should be emphasized that rotation
refers to the orientation of a fluid element and not the path followed by the
element. Thus, for an irrotational flow, if a pair of small sticks were placed on
a fluid element it can be observed that the orientation is retained even while
the fluid element moves along a circular path, as shown in Figure 5.1(a). In
other words, in an irrotational flow, the fluid elements do not rotate about
their own axes, that is, fluid elements do not spin in an irrotational flow.
But in a rotational flow fluid elements rotate about their axes, as shown in
Figure 5.1(b). Thus, in an irrotational flow, like the one shown in
Figure 5.1(a), the fluid elements move along circular paths but do not ro-
tate about their own axes. Thus, the angular velocity of fluid elements in an
irrotational flow is zero. If the flow field were rotating like a rigid body, then
the fluid elements in the field would experience a rotation about their own
axes, as shown in Figure 5.1(b). This type of motion is termed rotational and
cannot be described with a velocity potential.

(a) (b)

FIGURE 5.1 (a) Irrotational flow, (b) rotational flow.

243

ISTUDY
244 CHAPTER 5. VORTEX THEORY

If the possible distortion of the fluid elements caused by severe viscous


traction is ignored then there are only three possible ways in which a fluid
element can move. They are the following:

(a) Pure translation — the fluid elements are free to move anywhere in
space but continue to keep their axes parallel to the reference axes fixed
in space, as shown in Figure 5.2(a). The flow in the potential flow zone
outside the boundary layer over an aerofoil is substantially this type of
flow.
(b) Pure rotation — the fluid elements rotate about their own axes which
remain fixed in space, as shown in Figure 5.2(b).
(c) The general motion in which translation and rotation are compounded.
Such a motion is found, for example, in the wake of a bluff body.

(a) (b)
FIGURE 5.2 (a) Pure translational motion, (b) pure rotational motion.

A flow in which all the fluid elements behave as in item (a) above is called
potential or irrotational flow. All other flows exhibit, to a greater or lesser
extent, the spinning property of some of the constituent fluid elements, and
are said to possess vorticity, which is the aerodynamic term for elemental spin.
The flow is then termed rotational flow.
From the above descriptions it is evident that, a flow is either rotational,
possessing vorticity, or irrotational, for which vorticity is zero. The rotational
and irrotational nature and the properties of a flow can be examined analyt-
ically, leading to the development of characteristic equations governing the
flow. Using these equations, the nature of any unknown flow can be analyzed.

5.2 VORTICITY EQUATION IN


RECTANGULAR COORDINATES
In a two-dimensional motion, the vorticity at a point P, which is located
perpendicular to the plane, is equal to the limit of the ratio of the circulation
in an infinitesimal circuit embracing P to the area of the circuit.
A flow possesses vorticity if any of its elements are rotating (spinning). It
is a convenient way to investigate the motion of a circular element, treating
it as a solid, at the instant of time considered. Let P (x, y) be the centre of
the circular element and u and v are the velocity components, along x- and
y-directions, respectively, as shown in Figure 5.3.

ISTUDY
5.2. VORTICITY EQUATION IN RECTANGULAR COORDINATES 245

FIGURE 5.3 A fluid element and appropriate coordinates and velocity


components.

Let us assume that, the fluid element consists of numerous fluid particles
of mass ∆m each, such as one at the point Q(x + δx, y + δy). At point Q, the
velocity components, along x- and y-directions, respectively, are
∂u ∂u
u+ δx + δy
∂x ∂y
and
∂v ∂v
v+ δx + δy
∂x ∂y
The moment of momentum (or angular momentum) of the fluid element about
point P (x, y) is the sum of the moments of momentum of all the particles such
as Q about point P. Taking the anti-clockwise moment as positive, we have

Moment of momentum of the element

∑ [( ∂v ∂v
) (
∂u ∂u
) ]
= δx + δy δx − δx + δy δy ∆m
∂x ∂y ∂x ∂y
∑ ∑ ∑ ( )
∂v 2 ∂u 2 ∂v ∂u
= ∆m (δx) − ∆m (δy) + ∆m − δxδy
∂x ∂y ∂y ∂x
For a circular disc, about its centre, we have

∆m δxδy = 0

Therefore, the angular momentum of the disc becomes


∑ ∂v 2
∑ ∂u 2
Angular momentum = ∆m (δx) − ∆m (δy)
∂x ∂y
If the disc were a solid disc, its angular momentum would be Iω, where I is
its polar moment of inertia about P and ω its angular velocity about P. Thus,
assuming the fluid element as a solid disc, we have
∑ { }
2 2
I= ∆m (δx) + (δy)

ISTUDY
246 CHAPTER 5. VORTEX THEORY

and ∑ 2
∑ 2
∆m (δx) = ∆m (δy)
Thus, we have the angular momentum relation as
∑ { } ∑ ∂v ∑ ∂u
2 2 2 2
ω ∆m (δx) + (δy) = ∆m (δx) − ∆m (δy)
∂x ∂y
This gives the angular velocity as
∂v ∂u
2ω = −
∂x ∂y
The quantity 2ω is the elemental spin, also referred to as vorticity, which is
usually denoted as ζ. Thus,

∂v ∂u
ζ= − (5.1)
∂x ∂y

The units of ζ are radian per second. From Eq. (5.1) and the angular velocity
relation, it is seen that ζ = 2ω, i.e., the vorticity is twice the angular velocity.

5.2.1 Vorticity Equation in Polar Coordinates


In the polar coordinates, the vorticity equation can be expressed as

qt ∂qt 1 ∂qn
ζ= + − (5.2)
r ∂r r ∂θ

where r and θ are the polar coordinates and qt and qn are the tangential and
normal components of velocity, respectively. The derivation of Eq. (5.2) is
given in Section 5.3.
If (r, θ, n) are the radial, azimuthal and normal coordinates of a polar
coordinates system, the vorticity expression is given by

ζ = ir ζr + iθ ζθ + in ζn

where ir , iθ and in are the unit vectors in the directions of r, θ and n,


respectively. The vorticity components can be expressed as
1 ∂un ∂uθ
ζr = −
r ∂θ ∂n
∂ur ∂un
ζθ = −
∂n ∂r
1 ∂(ruθ ) 1 ∂ur
ζn = −
r ∂r r ∂θ
where ur , uθ and un are the velocity components along r, θ and n directions,
respectively.

ISTUDY
5.2. VORTICITY EQUATION IN RECTANGULAR COORDINATES 247

EXAMPLE 5.1 Find the vorticity of the following flows:

(a) V = c (x + y)i − c (x + y)j


(b) V = (x + y + z + t)i + 2 (x + y + z + t)j − 3 (x + y + z + t)k
( 2 )
(c) ur = uθ = 0, un = ar 2 − 1

Solution (a) Given, u = c(x + y), v = −c(x + y). This is a two-dimensional


flow in the xy-plane. Therefore, the vorticity component (with rotational axis
in the z-direction, which is normal to xy-plane) is

∂v ∂u
ζz = −
∂x ∂y
= −c − c

= −2c

(b) Given, u = x + y + z + t, v = 2 (x + y + z + t), w = −3 (x + y + z + t).

The vorticity components are:

∂w ∂v
ζx = −
∂y ∂z
= −3 − 2 = −5
∂u ∂w
ζy = −
∂z ∂x
=1+3=4
∂v ∂u
ζz = −
∂x ∂y
=2−1=1

Therefore, the vorticity becomes

ζ = ζx i + ζy j + ζz k

= −5 i + 4 j + k

( )
r2
(c) Given, ur = uθ = 0, un = c a2 −1 .

ISTUDY
248 CHAPTER 5. VORTEX THEORY

The vorticity components are


1 ∂un ∂uθ
ζr = −
r ∂θ ∂n
=0

∂ur ∂un
ζθ = −
∂n ∂r
2cr
=−
a2
1 ∂(ruθ ) 1 ∂ur
ζn = −
r ∂r r ∂θ
=0

Therefore, the vorticity becomes

2cr
ζ= −
a2

5.3 CIRCULATION
Circulation is the line integral of a vector field around a closed plane curve in
a flow field. By definition,
H
Γ= c
V · ds (5.3)

where Γ is circulation, V is flow velocity tangential to the streamline c, encom-


passing the closed curve under consideration, and ds is an elemental length.
If a line AB forms a closed loop or circuit in the flow, as shown in
Figure 5.4, then the line integral of Eq. (5.3) taken round the circuit is
defined as circulation, i.e.,
I
Γ= V cos β · ds
AB

= (u dx + v dy)
AB

where u is the component of V is the x-direction and v is that in the


y-direction. Note that, the circuit is imaginary and does not influence the
flow, i.e., it is not a boundary. In Eq. (5.3), both V and ds are vectors.
Therefore, the dot product of V and ds results in above expression for Γ.
Circulation implies a component of rotation of flow in the system. This
is not to say that there are circular streamlines, or the elements, of the fluid
are actually moving around some closed loop although this is a possible flow

ISTUDY
5.3. CIRCULATION 249

FIGURE 5.4 A loop AB in a flow field.

system. Circulation in a flow means that, the flow system could be resolved
into an uniform irrotational portion and a circulating portion. Figure 5.5
illustrates the concept of circulation.

V +v v

v v +
v
v
v
V −v
(a) Actual flow (b) Sum of circulating and irrotational parts
FIGURE 5.5 Illustration of circulation.

This implies that, if circulation is present in a fluid motion, then vorticity


must be present, even though it may be confined to a restricted space, as
in the case of the circular cylinder with circulation, where the vorticity at
the centre of the cylinder may actually be excluded from the region of flow
considered, namely that outside the cylinder.
An alternative equation for circulation Γ can be obtained by considering
the circuit of integration made up of a large number of rectangular elements
of sides δx and δy, as shown in Figure 5.6.

FIGURE 5.6 A circuit of integration, c, in the flow field.


Γ= (u dx + v dy)
c

ISTUDY
250 CHAPTER 5. VORTEX THEORY

Applying the integral round the element abcd with point P (x, y) at its
centre, where the velocity components are u and v, as shown in Figure 5.6,
we get

( ) ( ) ( ) ( )
∂v δx ∂u δy ∂v δx ∂u δy
∆Γ = v+ δy − u + δx − v − δy + u − δx
∂x 2 ∂y 2 ∂x 2 ∂y 2

Simplification of this results in


( )
∂v ∂u
∆Γ = − (δxδy)
∂x ∂y

The sum of the circulations of all the elemental areas in the circuit consti-
tutes the circulation of the circuit as a whole. As the circulation ∆Γ of each
element is added to the ∆Γ of the neighbouring element, the contributions of
the common sides disappear. Applying this reasoning from an element to the
neighboring element throughout the area, the only sides contributing to the
circulation, when the ∆Γs of all areas are summed together, are those sides
which actually form the circuit itself. This means that, for the total circuit c,
the circulation becomes

∫∫ ( )
∂v ∂u
Γ= − (dxdy)
area ∂x ∂y
I
= (u dx + v dy)
c

and the vorticity ζ is given by

∂v ∂u
ζ= −
∂x ∂y

If the strength of the circulation Γ remains constant whilst the circuit shrinks
to encompass an elemental area, that is, until it shrinks to an area of the size
of a rectangular element, then

Γ = ζ × (δxδy) = ζ × area of element

Therefore,
Γ
Vorticity = lim (5.4)
area→0
area of element
This is a result which enables an easy derivation of the vorticity relation in
polar coordinates.
Let us consider a segment of a fluid element of width δr, subtending angle
δθ, at the origin and width δr, as shown in Figure 5.7. If the segment is located

ISTUDY
5.3. CIRCULATION 251

FIGURE 5.7 A fluid element.

at the point P (r, θ), where the normal and tangential velocity components are
qn and qt , respectively, then the velocities along AB, BC, CD, DA are
( )
∂qn δθ
qn − − along AB
∂θ 2
( )
∂qt δr
qt + − along BC
∂r 2
( )
∂qn δθ
− qn + − along CD
∂θ 2
( )
∂qt δr
− qt − − along DA
∂r 2
where the direction qn is along r-direction and qt is along θ-direction. The
lengths of the sides of the elements are
AB = δr
( )
δr
BC = r + δθ
2
CD = δr
( )
δr
DA = r − δθ
2
The circulation about the element is the line integral of the tangential com-
ponent of flow velocity, i.e.,
( ) ( )( )
∂qn δθ ∂qt δr δr
Γ = qn − δr + qt + r+ δθ
∂θ 2 ∂r 2 2
( ) ( )( )
∂qn δθ ∂qt δr δr
− qn + δr − qt − r− δθ
∂θ 2 ∂r 2 2
This simplifies to
( )
qt ∂qt 1 ∂qn
Γ= + − rδrδθ
r ∂r r ∂θ

ISTUDY
252 CHAPTER 5. VORTEX THEORY

We know that,
Γ = vorticity × area of element

Also, the area of the element under consideration is (r δrδθ). Thus, the vor-
ticity is
( )
qt ∂qt 1 ∂qn
ζ= + −
r ∂r r ∂θ

This is the vorticity expression in polar coordinates.

5.4 LINE (POINT) VORTEX


A line vortex is a string of rotating particles. In a line vortex, a chain of
fluid particles are spinning about their common axis and carrying around
with them a swirl of fluid particles which flow around in circles. A cross-
section of such a string of particles and the associated flow show a spinning
point, outside of which the flow streamlines are concentric circles, as shown in
Figure 5.8.

axis

(a) (b)

FIGURE 5.8 (a) Straightline vortex, (b) cross-section showing


the associated streamlines.

Vortices can commonly be encountered in nature. The difference between


a real (actual) vortex and theoretical vortex is that, the real vortex has a core
of fluid which rotates like a solid, although the associated swirl outside is the
same as the flow outside the point vortex. The streamlines associated with a
line vortex are circular, and, therefore, the particle velocity at any point must
be only tangential.
Stream function of a vortex can easily be obtained as follows. Consider a
vortex of strength Γ, at the origin of a polar coordinate system, as shown in
Figure 5.9.
Let P (r, θ) be a general point and velocity at P is always normal to OP
(tangential). The radial velocity at any point P is zero, i.e.,

1 ∂ψ
=0
r ∂θ

ISTUDY
5.4. LINE (POINT) VORTEX 253

FIGURE 5.9 A vortex at origin.

since in polar coordinates, the radial velocity qr and tangential velocity qθ , in


terms of stream function ψ are
1 ∂ψ
qr =
r ∂θ
∂ψ
qt = −
∂r
For qr = 0, the stream function ψ should be a function of r only. The tan-
gential velocity at any point P 1 is

Γ ∂ψ
qt = =−
2πr ∂r
Therefore, ∫
Γ
ψ= − dr
2πr
Integrating along a convenient boundary, such as from A to P in Figure 5.9,
from radius r0 (radius of streamline, ψ = 0) to P (r, θ), we get the stream
function as
[ ]r
Γ
ψ=− ln r
2π r0

i.e.,
Γ r
ψ=− ln (5.5)
2π r0
This is the stream function for a vortex, and the circulation Γ of a flow is
positive when it is counterclockwise.
We know that, the streamlines of a line vortex are concentric circles.
Therefore, the equipotential lines (which are always orthogonal to the stream-
lines) must be radial lines emanating from the centre of the vortex. Also, for
a vortex, the normal component of velocity qn = 0. Therefore, the potential
function ϕ must be a function of θ only. Thus,
1 Refer to Eq. (2.49).

ISTUDY
254 CHAPTER 5. VORTEX THEORY

1 dϕ Γ
qt = =
r dθ 2π r
Therefore,
Γ
dϕ = dθ

Integrating this, we get
Γ
ϕ= θ + constant

By assigning ϕ = 0 at θ = 0, we obtain

Γ
ϕ= θ (5.6)

This is the potential function for a vortex.


Also, we know that the stream function for a source2 is


ψ=

where m is the strength of the source.
Comparing the stream functions of a vortex and a source, we see that the
streamlines of the source (the radial lines emanating from a point) and the
streamlines of the vortex (the concentric circles) are orthogonal.

5.5 LAWS OF VORTEX MOTION


In Section 5.4, we saw that a point vortex can be considered as a string of
rotating particles surrounded by fluid at large moving irrotationally. Further,
the flow investigation was confined to a plane section normal to the length or
axis of the vortex. A more general definition is that, a vortex is a flow system
in which a finite area in a plane normal to the axis of a vortex contains
vorticity. Figure 5.10 shows a sectional area S in the plane normal to the
axis of a vortex. The axis of the vortex is clearly always normal to the two-
dimensional flow plane considered and the influence of the so-called line vortex
is the influence, in a section plane, of an infinitely long straightline vortex of
vanishingly small area.
The axis of a vortex, in general, is a curve in space, and area S is a
finite size. It is convenient to consider that the area S is made up of several
elemental areas. In other words, a vortex consists of a bundle of elemental
vortex lines or filaments. Such a bundle is termed a vortex tube, being a tube
bounded by vortex filaments.
The vortex axis is a curve winding about within the fluid. Therefore, it
can flexure and influence the flow as a whole. The estimation of its influence
2 Refer to Eq. (2.46).

ISTUDY
5.6. HELMHOLTZ’S THEOREMS 255

FIGURE 5.10 The vorticity of a section of vortex tube.

on the fluid at large is somewhat complex. In our discussions here the vortices
considered are fixed relative to some axes in the system or free to move in a
controlled manner and can be assumed to be linear. Furthermore, the vortices
will not be of infinite length, therefore, the three-dimensional or end influence
must be accounted for.
In spite of the above simplifications, the vortices conform to laws of motion
appropriate to their behaviour. A rigorous treatment of the vortices, without
the simplifications imposed in our treatment can be found in Theoretical Hydro
and Aerodynamics, Vols. I and II by Milne-Thomson and Hydrodynamics by
Lamb.

5.6 HELMHOLTZ’S THEOREMS


The four fundamental theorems governing vortex motion in an inviscid flow
are called Helmholtz’s theorems (named after the author of these theorems).
The first theorem refers to a fluid particle (or element) in general motion
possessing all or some of the following:

1. Linear velocity
2. Vorticity
3. Distortion

This theorem has been discussed in part in Section 5.3, where the vorticity
was explained and its expression in Cartesian or polar coordinates were de-
rived. Helmholtz’s first theorem states that, “the circulation of a vortex-tube
is constant at all cross-sections along the tube”.
The second theorem demonstrates that, “the strength of a vortex tube
(i.e., the circulation) is constant along its length”.
This is sometimes referred to as the equation of vortex continuity. It can
be shown that the strength of a vortex cannot grow or diminish along its axis
or length. The strength of a vortex is the magnitude of the circulation around
it, and is equal to the product of vorticity ζ and area S. Thus,

ISTUDY
256 CHAPTER 5. VORTEX THEORY

Γ=ζS

It follows from the second theorem that, ζ S is constant along the vortex
tube (or filament), so that if the cross-sectional area diminishes, the vorticity
increases and vice versa. Since infinite vorticity is unacceptable, the cross-
sectional area S cannot diminish to zero. In other words, a vortex cannot
end in the fluid. In reality the vortex must form a closed loop, or originate
(or terminate) in a discontinuity in the fluid such as a solid body or a surface
of separation. In a different form it may be stated that a vortex tube cannot
change its strength between two sections unless vortex filaments of equivalent
strength join or leave the vortex tube, as shown in Figure 5.11.

Γ − ∆Γ

Section B
Section A

FIGURE 5.11 Vortex-tube fragmentation.

It is seen that at section A, the vortex tube strength is Γ. Downstream of


section A an opposite vortex filament of strength −∆Γ joins the vortex tube.
Therefore, at section B, the strength of the vortex tube is

Γ = Γ − ∆Γ

as shown in the figure. This is of great importance to the vortex theory of


lift.
The third theorem demonstrates that, a vortex tube consists of the same
particles of fluid, i.e., “there is no fluid interchange between the vortex tube
and surrounding fluid.”
The fourth theorem states that, “the strength of a vortex remains constant
in time.”

5.7 VORTEX THEOREMS


Now let us have a closer look at the theorems governing vortex motion. Con-
sider the circulation of a closed material line. By definition ([Eq. (5.3)], we
have the circulation as
I
Γ= V · ds
c

ISTUDY
5.7. VORTEX THEOREMS 257

The time rate of change of Γ can be expressed as


I I ∫
DΓ D DV
= V · ds = · ds + V · dV (5.7)
Dt Dt c c Dt

since ds/dt = V , where V is the velocity, s is length(and t ) is time. The


V
second integral in Eq. (5.7) vanishes, since V · dV = d V · is the total
2
differential of a single valued function, and the starting point of integration
coincides with the end point.
By Euler equation, we have

DV ▽p
= FB −
Dt ρ
where FB is the body force. From Eq. (5.7) and the Euler equation, we obtain
the rate of change of the line integral over the velocity vector in the form
∫ ∫
DΓ ▽p
= FB · ds − · ds (5.8)
Dt c c ρ

In Eq. (5.8), DΓ/Dt vanishes if (FB · ds) and ▽p/ρ can be written as total
differentials. When the body force FB has a potential (i.e., when the body
force is a conservative force field); implying that the work done by the weight
in taking a body from a point P to another point Q is independent of the
path taken from P to Q, and depends only on the potential, the first closed
integral in Eq. (5.8) becomes zero because

FB · ds = − ▽ ψ · ds = − dψ (5.9)
For a homogeneous density field or in barotropic flow, the density depends
only on pressure, that is ρ = f (p). For such a flow, the second term on the
right-hand side of Eq. (5.8), can be expressed as

▽p dp
· ds = = dp (5.10)
ρ ρ(p)

Therefore, for barotropic fluids, the second integral also vanishes in Eq. (5.8).
Equations (5.8) to (5.10) form the content of Thompson’s vortex theorem
or Kelvin’s circulation theorem. This theorem states that, “in a flow of inviscid
and barotropic fluid, with conservative body forces, the circulation around a
closed curve (material line) moving with the fluid remains constant with time,
if the motion is observed from a non-rotating frame.”
The vortex theorem can be interpreted as follows:

“The position of a curve c in a flow field, at any instant of time, can be located
by following the motion of all the fluid elements on the curve”, i.e., Kelvin’s
circulation theorem states that, “the circulation around the curve c at the two
locations is the same.” In other words,

ISTUDY
258 CHAPTER 5. VORTEX THEORY


=0 (5.11)
Dt
where D/Dt(≡ ∂/∂t + ▽ · ) has been used to emphasize that the circulation
is calculated around a material contour moving with the fluid.
With Kelvin’s theorem as the starting point, we can explain the famous
Helmholtz’s vortex theorem, which allows a vivid interpretation of vortex
motions which are of fundamental importance in aerodynamics. Before ven-
turing to explain Helmholtz’s vortex theorems, it would be beneficial if we
consider the origin of the circulation around an aerofoil, in a two-dimensional
potential flow, because Kelvin’s theorem seems to contradict the formulation
of this circulation.
It is well known that, the force on an aerofoil in a two-dimensional
potential flow is proportional to the circulation. Also, the lift, namely the
force perpendicular to the undisturbed incident flow direction, experienced
by the aerofoil is directly proportional to the circulation Γ, around the aero-
foil. The lift per unit span of an aerofoil can be expressed as
L = ρV Γ
where the ρ and V , respectively, are the density and velocity of the freestream
flow.
Now let us examine the flow around a symmetrical and an unsymmetri-
cal aerofoil in identical flow fields, as shown in Figure 5.12. As seen from
Figure 5.12(a), the flow around the symmetrical aerofoil at zero angle of inci-
dence is also symmetric. Therefore, there is no net force perpendicular to the
incident flow direction. The contribution of the line integral of velocity about
the upper-half of the aerofoil to the circulation has exactly the same magni-
tude as the contribution of the line integral of velocity about the lower-half,
but with opposite sign. Therefore, the total circulation around the symmetric
aerofoil is zero.
Γ Γ

u
u

u u

(a) (b)
FIGURE 5.12 (a) Symmetrical and (b) unsymmetrical aerofoil in
uniform flow.

The flow around the unsymmetrical aerofoil, as shown in Figure 5.12(b),


is asymmetric. The contribution of the line integral of velocity about upper-
half of the aerofoil has an absolute value larger than that of the contribution

ISTUDY
5.7. VORTEX THEOREMS 259

about the lower-half. Therefore, the circulation around the unsymmetrical


aerofoil is nonzero. By Bernoulli theorem it can be inferred that, the velocity
along a streamline which runs along the upper side of the aerofoil is larger
on the whole than the velocity on the lower side. Therefore, the pressure on
the upper side is less than the pressure on the lower side. Thus there is a net
upward force acting on the aerofoil.

FIGURE 5.13 Velocity on either side of separation surface behind the


aerofoil.

For an unsymmetrical aerofoil the flow velocity over the upper and lower
surfaces are different even when it is at zero angle of incidence to the freestream
flow. Because of this, the pressure on either side of the dividing streamline,
shown in Figure 5.13, are different. Also, the velocities on either side of the
separation surface are different, as shown in the figure. This implies that, the
pressure on either side of the separation surface are different. It is well known
that, the separation surface which is also called slipstream can not be stable
when the pressures on either side are different3 The slipstream will assume a
shape in such a manner to have equal pressure on either side of it. Here the
pressure at the lower side is higher than that at the upper side. Thus, the
slipstream bends up, as shown in Figure 5.14(a).

FIGURE 5.14 Flow past an aerofoil at start-up.

At the first instant of start-up, the flow around the trailing edge of the
aerofoil is at very high velocities. Also, the flow becomes separated from the
3 Applied Gas Dynamics by Ethirajan Rathakrishnan, John Wiley, NJ, 2010.

ISTUDY
260 CHAPTER 5. VORTEX THEORY

upper surface. Flow field around an aerofoil at different phases of start-up is


shown in Figure 5.14. The separation at the upper surface is caused by the
very large deceleration of the flow from the maximum thickness location to
the separation point S, which is formed on the upper surface since the flow is
still circulation-free flow [Figure 5.14(a)]. This flow separates from the upper
surface even with very little viscosity (i.e., µ → 0) and forms the wake, which
becomes the discontinuity surface in the limiting case of µ = 0. The flow
is irrotational everywhere except the wake region. Soon after start-up, the
separation point is dipped to the trailing edge, as per Kutta hypothesis, and
the slipstream rolls-up as shown in Figure 5.14(b). The vortex thus formed is
pushed downstream and positioned at a location behind the aerofoil, as shown
in Figure 5.14(c). This vortex is called starting vortex. The starting vortex
is essentially a free vortex because it is formed by the kinematics of the flow
and not by the viscous effect.
By Kelvin’s circulation theorem, a closed curve which surrounds the aero-
foil and the vortex still has zero circulation. In other words, the circulation of
the starting vortex and the bound vortex (this is due to the boundary layer at
the surface of the aerofoil in viscous flow) are of equal magnitude, as shown
in Figure 5.15.

Bound vortex Free vortex

−|Γ| +|Γ|

FIGURE 5.15 Circulation of starting and bound vortices.

A closed line which surrounds only the vortex has a fixed circulation and
must necessarily cross the discontinuity surface. Therefore, Kelvin’s circula-
tion theorem does not hold for this line. A curve which surrounds the aerofoil
only has the same circulation as the free vortex, but with opposite sign and,
therefore, the aerofoil experiences a lift. The circulation about the aerofoil
with a vortex lying over the aerofoil, due to the boundary layer at the surface,
is called the bound vortex.
In the above discussion, we used the obvious law that the circulation of
a closed loop is equal to the sum of the circulation of the meshed network
bounded by the curve, as shown in Figure 5.16.

Γclosed loop = Γi (5.12)
that is, the sum of the circulations of all the areas is the neighbouring circu-
lation of the circuit as a whole. This is because, as the ∆Γ of each element is
added to the ∆Γ of the neighbouring element, the contribution of the common
sides (Figure 5.16) disappears. Applying this argument from one element to
the neighbouring element throughout the area, the only sides contributing to

ISTUDY
5.7. VORTEX THEOREMS 261

Γi
FIGURE 5.16 Circulation of meshed network.

the circulation when the ∆Γs of all elemental areas are summed together are
those sides which actually form the circuit itself. This means, that for the
circuit as a whole, the circulation is
∫∫ ( ) I
∂v ∂u
Γ= − dxdy = (u dx + v dy)
∂x ∂y

In this relation, the surface integral implies that the integration is over the
area of the meshed network, and the cyclic integral implies that the integration
is around the circuit of the meshed network.
For discussing the physics of Helmholtz’s theorem, we need to make use
of Stokes’ integral theorem.

5.7.1 Stokes’ Theorem


Stokes’ theorem relates the surface integral over an open surface to a line
integral along the bounded curve. Let S be a simply connected surface, which
is otherwise of arbitrary shape, whose boundary is c, and let u be any arbitrary
vector. Also, we know that any arbitrary closed curve on an arbitrary shape
can be shrunk to a∫ single point. The Stokes’ integral theorem states that,
“the line∫∫integral u · dx about the closed curve c is equal to the surface
integral (▽ × u) · n ds over any surface of arbitrary shape which has c as
its boundary.” i.e., the surface integral of a vector field u is equal to the line
integral of u along the bounding curve.
I ∫∫
u · dx = (curl u) · n ds (5.13)
c s

where dx is an elemental length on c, and n is unit vector normal to any


elemental area on ds, as shown in Figure 5.17.
Stokes’ integral theorem allows a line integral to be changed to a surface
integral. The direction of integration is positive counter-clockwise as seen
from the side of the surface, as shown in the figure.
Helmholtz’s first vortex theorem states that, “the circulation of a vortex
tube is constant along the tube.”
A vortex tube is a tube made up of vortex lines which are tangential
lines to the vorticity vector field, namely curl u (or ζ). A vortex-tube is

ISTUDY
262 CHAPTER 5. VORTEX THEORY

Line of sight

ds

FIGURE 5.17 Sign convention for integration in Stokes’ integral theorem.

Γ2 curl u

Γ1

Γ1 = Γ2

FIGURE 5.18 A vortex tube.

shown in Figure 5.18. From the definition of vortex tube it is evident that,
it is analogous to the streamtube, where the flow velocity is tangential to the
streamlines constituting the streamtube. A vortex line is, therefore, related to
the vorticity vector in the same way the streamline is related to the velocity
vector. If ζx , ζy and ζz are the Cartesian components of the vorticity vector
ζ, along x-, y- and z-directions, respectively, then the orientation of a vortex
line satisfies the equation,
dx dy dz
= =
ζx ζy ζz
which is analogous to,
dx dy dz
= =
u v w
along a streamline. In an irrotational vortex (free vortex), the only vortex
line in the flow field is the axis of the vortex. In a forced vortex (solid-body
rotation), all lines perpendicular to the plane of flow are vortex lines.
Now consider two closed curves c1 and c2 in a vortex tube, as shown in
Figure 5.19.
According to Stokes’ theorem, the two line integrals over the closed curves
in Figure 5.19 vanish, because the integrand on the right-hand side of
Eq. (5.13) is zero, since curl u is, by definition, perpendicular to n. The
contribution to the integral from the infinitely close segments c3 and c4 of the
curve cancel each other, leading to the equation

ISTUDY
5.7. VORTEX THEOREMS 263

n
c4
c2

c1
c3

FIGURE 5.19 Two loops on a vortex tube.

∫ ∫
u · dx + u · dx = 0 (5.14)
c1 c2
since the distance between the segments c3 and c4 are infinitesimally small,
we ignore that and treat c1 and c2 to be closed curves. By changing the
direction of integration over c2 , thus changing the sign of the second integral in
Eq. (5.14), we obtain Helmholtz’s first vortex theorem.
I I
u · dx = u · dx (5.15)
c1 c2

Derivation of this equation clearly demonstrates the kinematic nature of


Helmholtz’s first vortex theorem. Another approach to the physical expla-
nation of this theorem stems from the fact that, the divergence of the vor-
ticity vector vanishes. That is, the vorticity vector field curl u can be con-
sidered as analogous to an incompressible flow (for which the divergence of
velocity is zero). In other words, the vortex tube becomes the streamtube
of the new∫∫ field. Now applying the continuity equation in its integral form
(that is, s ρui ni ds = 0) to a part of this streamtube, and at the same time
replacing u by curl u, we get
∫∫
ρ (curl u) · n ds = 0
s
Since ρ is a constant, we can write this as
∫∫
(curl u) · n ds = 0 (5.16)
s

i.e., for every closed surface s, the flux of the vorticity is zero. Applying
Eq. (5.16) to a part of the vortex tube whose closed surface consists of the
surface of the tube and two arbitrarily oriented cross-sections A1 and A2 , we
obtain ∫∫ ∫∫
(curl u) · n ds + (curl u) · n ds = 0 (5.17)
A1 A2
since the integral over the tube surface vanishes. The integral
∫∫
(curl u) · n ds

ISTUDY
264 CHAPTER 5. VORTEX THEORY

is called the vortex strength. It is identical to the circulation. Form Eq. (5.17)
it is evident that, “the vortex strength of a vortex-tube is constant.”
Noting the sense of integration of the line integral, Stokes’ theorem trans-
forms Eq. (5.17) into Helmholtz’s first theorem [Eq. (5.15)]. From this
representation, it is obvious that just like the streamtube, the vortex tube
also cannot come to an end within the fluid, since the amount of fluid which
flows through the tube (in unit time) cannot simply vanish at the end of the
tube. The tube must either reach out to infinity (i.e., should extend to infin-
ity), or end at the boundaries of the fluid, or close around into itself and, in
the case of a vortex tube, form a vortex ring.
A very thin vortex tube is referred to as a vortex filament. The vortex
filaments are of particular importance in aerodynamics. For a vortex filament
the integrand of the surface integral in Stokes’ theorem [Eq.(5.13)]
I ∫∫
u · dx = (curl u) · nds = Γ (5.18)
c ∆s

can be taken in front of the integral to obtain

(curl u) · n∆s = Γ (5.19)

or
2ω · n ∆s = 2ω ∆s = constant (5.20)
where ω is the angular velocity. From this it is evident that, the angular
velocity increases with decreasing cross-section of the vortex filament.
It is an usual practice to idealize a vortex tube of infinitesimally small cross
section into a vortex filament. Under this idealization, the angular velocity of
the vortex, given by Eq. (5.20), becomes infinitely large. From the relation

ω ∆s = constant (5.21)
we have ω → ∞, for ∆s → 0.
The flow field outside the vortex filament is irrotational. Therefore, for a
vortex of strength Γ at a particular position, the spatial distribution of curl u
is fixed. In addition, if div u is also given (e.g., div u = 0 in an incompress-
ible flow), then according to the fundamental theorem of vortex analysis, the
velocity field u (which may extend to infinity) is uniquely determined provided
the normal component of velocity vanishes asymptotically sufficiently fast at
infinity and no internal boundaries exist.
The fundamental theorem of vector analysis is also essentially purely kine-
matic in nature. Therefore, it is valid for both viscous and inviscid flows, and
not restricted to inviscid flows only. Let us split the velocity vector u into two
parts, namely due to potential flow and rotational flow. Therefore,

u = uIR + uR (5.22)
where uIR is velocity of irrotational flow field and uR is velocity of rotational
flow field. Thus, uIR is velocity of an irrotational flow field, i.e.,

ISTUDY
5.8. CALCULATION OF UR , THE VELOCITY ... 265

curl uIR = ▽ × uIR = 0 (5.23)


The second is a solenoidal (coil like shape) flow field, thus,

div uR = ▽ · uR = 0 (5.24)

Note that, Eq. (5.23) is the statement that “the vorticity of a potential flow is
zero” and Eq. (5.24) is the statement of continuity equation of incompressible
flow.
The combined field is, therefore, neither irrotational nor solenoidal. The
field uIR is a potential flow and, thus, in terms of potential function ϕ, we
have uIR = ▽ ϕ. Let us assume that, the divergence u to be a given function
g(x). Thus,
div u = ▽ · uIR + ▽ · uR = g(x)
i.e.,
div u = ▽ · uIR = g(x) (5.25)
since ▽ · uR = 0. Also, uIR = ▽ ϕ. Therefore,

∂2ϕ
▽ · ▽ϕ = = g(x)
∂x∂x

▽2 ϕ = g(x) (5.26)

This is an inhomogeneous Laplace equation, also called Poisson’s equation.


The theory of this partial differential equation is the subject of potential the-
ory which plays an important role in many branches of physics as in fluid
mechanics. It is well known from the results of potential theory that the
solution of Eq. (5.26) is given by
∫∫∫
1 g(x′ )
ϕ(x) = − dV (5.27)
4π ∞ |x − x′ |
where x is the place where the potential ϕ is calculated, and x′ is the
′ ′ ′ ′ ′ ′
abbreviation for the integration variables x1 , x2 and x3 , and dV = (dx1 dx2 dx3 )
is a differential volume. The domain ∞ implies that the integration is to be
carried out over the entire space.

5.8 CALCULATION OF uR , THE VELOCITY


DUE TO ROTATIONAL FLOW
We see that Eq. (5.24) is satisfied if uR is represented as the curl of a new,
yet unknown, vector field a. Thus,

uR = curl a = ▽ × a (5.28)

ISTUDY
266 CHAPTER 5. VORTEX THEORY

We know that the divergence of the curl always vanishes4 . Therefore,

▽ · (▽ × a) = ▽ · uR = 0 (5.29)

Now let us form the curl of u and, from Eq. (5.23), obtain the equation

▽ × (u) = ▽ × (▽ × a) (5.30)

But using the vector identity

∆u = ▽ (▽ · u) − ▽ × (▽ × u)

We can express Eq. (5.30) as

▽ × u = ▽ (▽ · a) − ∆a (5.31)

Up to now the only condition on vector a is to satisfy Eq. (5.28). But this
condition does not uniquely determine this vector, because we can always add
the gradient of some other function f to a without changing Equation (5.28),
since ▽ × ▽f ≡ 0. If, in addition, we want the divergence of a to vanish (that
is, ▽ · a = 0), we obtain from Eq. (5.31) the simpler equation

▽ × u = − ∆a (5.32)
In this equation, let us consider ▽ × u as a given vector function b(x), which
is determined by the choice of the vector filament and its strength (i.e., circu-
lation). Thus, the Cartesian component form of the vector Eq. (5.32) leads
to three Poisson’s equations, namely

∆ ai = − bi ; i = 1, 2, 3 (5.33)

For each of these component equations, we can apply the solution [Eq. (5.27)]
of Poisson’s equation. Now, vectorially combining the result, we can write the
solution for a, from Eq. (5.32), in short as
∫∫∫
1 b(x′ )
a=+ ′
dV (5.34)
4π ∞ |x − x |

Thus, calculation of the velocity field u(x) for a given distribution g(x) ≡ div u
and b(x) = curl u is reduced to the following integration processes, which may
have to be done numerically,
4 In deed, this is true for any vector, e.g., if a and b are vectors,
a · (a × b) = [aab] = 0
Therefore, in general, it can be expressed,
[▽ ▽ a] = 0
where ▽ and a are vectors. The representation “[ ]” is termed “box” notation in vector
algebra.

ISTUDY
5.8. CALCULATION OF UR , THE VELOCITY ... 267

[ ∫∫∫ ] [ ∫∫∫ ]
1 div u(x′ ) 1 curl u(x′ )
u(x) = − ▽ dV +▽× dV (5.35)
4π ∞ |x − x′ | 4π ′
∞ |x − x |

Now, let us calculate the solenoidal term of the velocity uR , using Eq. (5.35).
This is the only term in incompressible flow without internal boundaries.
Consider a field which is irrotational outside the vortex filament, shown in
Figure 5.20.

Γ
x
curl
r u
′ n
x

FIGURE 5.20 A vortex filament.

The velocity field outside the filament is given by


[ ∫∫∫ ]
1 curl u(x′ )
uR (x) = ▽ × ′
dV (5.36)
4π ∞ |x − x |

The integration is carried out only over the volume of the vortex filament,
whose volume element is
dV = ds n · dx′ (5.37)
where dx′ = n ds′ , is the vectorial element of the vortex filament, ds is the
cross-sectional area and n is the unit vector. Also, the unit vector n can be
expresses as
curl u
n=
|curl u|
Therefore, Eq. (5.37) becomes

curl u
dV = ds · nds′
|curl u|
or
ds′
dV = curl u · n ds
|curl u|
Substituting this into Eq. (5.36), we get
[ ∫∫∫ ]
1 (curl u(x′ )) · n ds ′
uR (x) = ▽ × dx (5.38)
4π filament |x − x′ |

ISTUDY
268 CHAPTER 5. VORTEX THEORY

since
curl u ds′
= n ds′ = dx′
|curl u|
First let us integrate over a small cross-section surface ∆S. For ∆S → 0,
the change of the vector x′ over this surface can be neglected. Thus, taking
1
in front of the surface integral, we obtain
|x − x′ |
[ ∫ (∫ ∫ ) ]
1 1 ′ ′
uR (x) = ▽ × (curl u(x )) · n ds dx (5.39)
4π |x − x′ |

From Stokes’ theorem, the surface integral is equal to the circulation Γ. By


Helmholtz’s first vortex theorem, Γ is constant along the vortex filament and,
therefore, independent of x′ . Thus, from Eq. (5.39) we get

Γ dx′
uR (x) = ▽× (5.40)
4π |x − x′ |
In index notation, the right-hand side of Eq. (5.40) can be written as

Γ ∂ 1 ′
ϵijk dx
4π ∂xj r k


The operator5 ϵijk can directly be taken into the integral. The term
∂xj
∂ ( − 1)
r (with ri = (xi − x′i ) and r = |r|) becomes
∂xj

∂ ( − 1) 1 ∂r 1 ( ) 1
r =− 2 = − 2 xj − x′j = − rj r − 3
∂xj r ∂xj r r

In vector form, this is simply


∂ ( − 1) r
r =− 3
∂xj r

Therefore, substituting the above into Eq. (5.40), we get the famous Biot–
Savart law, ∫
Γ dx′ × r
uR (x) = (5.41)
4 π filament r3
where r = (x−x′ ). The Biot–Savart law is an useful relation in aerodynamics.
5ϵ is a tensor and is positive when the subscripts i, j, k are expressed in cyclic order.
ijk
i.e.,
ϵijk = ϵjki = ϵkij
But, for non-cyclic orders of i, j, k
ϵijk = −ϵjki
and so on.

ISTUDY
5.9. BIOT–SAVART LAW 269

5.9 BIOT–SAVART LAW


Biot–Savart law relates the intensity of magnitude of magnetic field close to
an electric current carrying conductor to the magnitude of the current. It
is mathematically identical to the concept of relating intensity of flow in the
fluid close to a vorticity-carrying vortex tube to the strength of the vortex
tube. It is a pure kinematic law, which was originally discovered through
experiments in electrodynamics. The vortex filament corresponds there to a
conducting wire, the vortex strength to the current, and the velocity field to
the magnetic field. The aerodynamic terminology namely, “induced velocity”
stems from the origin of this law.
Now let us calculate the induced velocity at a point in the field of an
elementary length δs of a vortex of strength Γ. Assume that a vortex-tube of
strength Γ, consisting of an infinite number of vortex filaments, to terminate
in some point P , as shown in Figure 5.21.

FIGURE 5.21 Vortex tube discharging into a sphere.

The total strength of the vortex tube will be spread over the surface of
a spherical boundary of radius R. The vorticity in the spherical surface will
thus have the total strength of Γ. Because of symmetry the velocity of flow at
the surface of the sphere will be tangential to the circular line of intersection
of the sphere with a plane normal to the axis of the vortex-tube. Such plane
will be a circle ABC of radius r subtending a conical angle 2θ at P , as shown
in Figure 5.22.
If the velocity on the sphere at (R, θ) from P is v, then the circulation
round the circuit ABC is Γ′ , where
Γ′ = 2πR sin θ v
The radius of the circuit is r = R sin θ, therefore, we have
Γ′ = 2πr v (5.42)
But the circulation round the circuit is equal to the strength of the vorticity
in the contained area. This is on the cap ABCD of the sphere. Since the

ISTUDY
270 CHAPTER 5. VORTEX THEORY

FIGURE 5.22 Vortex tube discharged into an imaginary sphere.

distribution of the vorticity is constant over the surface, we have

Surface area of the cap 2πR2 (1 − cos θ)


Γ′ = Γ= Γ
Surface area of the sphere 4πR2

i.e.,
Γ
Γ′ = (1 − cos θ) (5.43)
2
From Eqs. (5.42) and (5.43), we obtain the induced velocity as

Γ
v= (1 − cos θ) (5.44)
4πr

FIGURE 5.23 A short vortex tube discharged into an imaginary sphere.

Now, assume that the length of the vortex decreases until it becomes very
short, as shown (P1 P ) in Figure 5.23. The circle ABC is influenced by the
opposite end P1 also (i.e., both the ends P and P ′ of the vortex influence the
circle). Now the vortex elements entering the sphere are congregating on P1 .
Thus, the sign of the vorticity is reversed on the sphere of radius R1 . The
velocity induced at P1 becomes

Γ
v1 = − (1 − cos θ1 ) (5.45)
4πr

ISTUDY
5.9. BIOT–SAVART LAW 271

The net velocity on the circuit ABC is the sum of Eqs. (5.44) and (5.45),
therefore, we have
[ ]
Γ
v − v1 = (1 − cos θ) − (1 − cos θ1 )
4πr
Γ
= (cos θ1 − cos θ)
4πr
As the point P1 approaches P ,

cos θ → cos (θ − δθ) = cos θ + sin θ δθ


and
(v − v1 ) → δv
Thus, at the limiting case of P1 approaching P , we have the net velocity as
Γ
δv = sin θ δθ (5.46)
4πr
This is the velocity induced by an elementary length δs of a vortex of strength
Γ which subtends an angle δθ at point P located by the ordinate (R, θ) from
the element. Also, r = R sin θ and R δθ = δs sin θ, thus, we have

Γ
δv = sin θ δs (5.47)
4πR2
It is evident from Eq. (5.47) that, to obtain the velocity induced by a vortex
this equation has to be integrated. This treatment of integration varies with
the length and shape of the finite vortex being studied. In our study here,
for applying Biot–Savart law, the vortices of interest are all nearly linear.
Therefore, there is no complexity due to vortex shape. The vortices will vary
only in their overall length.

5.9.1 A Linear Vortex of Finite Length


Examine the linear vortex of finite length AB, shown in Figure 5.24. Let P
be an adjacent point located by the angular displacements α and β from A
and B respectively. Also, the point P has coordinates r and θ with respect to
an elemental length δs of AB. Further, h is the height of the perpendicular
from P to AB and the foot of the perpendicular is at a distance s from δs.
The velocity induced at P by the element of length δs, by Eq. (5.47), is

Γ
δv = sin θ δs (5.48)
4πr2
The induced velocity is in the direction normal to the plane ABP , shown in
Figure 5.246 .
6 The induced velocity for the circulation shown, that is, clockwise, when viewed from

right to left, is into the page. When the circulation direction is reversed (i.e., counter-
clockwise) the induced velocity will be from the page to upwards.

ISTUDY
272 CHAPTER 5. VORTEX THEORY

FIGURE 5.24 A linear vortex of finite length.

The velocity at P due to the length AB is the sum of induced velocities


due to all elements, such as δs. However, all the variables in Eq. (5.48) must
be expressed in terms of a single variable before integrating to get the effective
velocity. A variable such as ϕ, shown in Figure 5.24 may be chosen for this
purpose. The limits of integration are
(π ) (π )
ϕA = − − α to ϕB = + −β
2 2
since ϕ passes through zero while integrating from A to B. Here we have
sin θ = cos ϕ
r2 = h2 sec2 ϕ
ds = d (h tan ϕ) = h sec2 ϕ dϕ
Thus, we have the induced velocity at P due to vortex AB, by Eq. (5.48), as
∫ 2 −β )
+(π
Γ
v= cos ϕ dϕ
−( 2 −α
π
) 4πh

[ (π ) (π )]
Γ
= sin − β + sin −α
4πh 2 2

Γ
v= (cos α + cos β) (5.49)
4πh
This is an important result of vortex dynamics. Form this result, we obtain
the following specific results of velocity in the vicinity of the line vortex.

5.9.2 Semi-infinite Vortex


A vortex is termed semi-infinite vortex when one of its ends stretches to
infinity. In our case, let the end B in Figure 5.24 stretches to infinity. There-

ISTUDY
5.9. BIOT–SAVART LAW 273

fore, β = 0 and cos β = 1, thus, from Eq. (5.49), we have the velocity induced
by a semi-infinite vortex at a point P as

Γ
v= (cos α + 1) (5.50)
4πh

5.9.3 Infinite Vortex


An infinite vortex is that with both ends stretching to infinity. For this case,
we have α = β = 0. Thus, the induced velocity due to an infinite vortex
becomes
Γ
v= (5.51)
2πh
For a specific case of point P just opposite to one of the ends of the vortex,
say A, we have α = π/2 and cos α = 0. Thus, the induced velocity at P
becomes
Γ
v= (5.52)
4πh
This amounts to precisely half of the value for the infinitely long vortex fila-
ment [Eq. (5.51)], as we would expect because of symmetry.
While discussing Figure 5.15, we saw that the circulation about an aerofoil
in two-dimensional flow can be represented by a bound vortex. We can as-
sume these bound vortices to be straight and infinitely long vortex filaments
(potential vortices). As far as the lift is concerned, we can think of the whole
aerofoil as being replaced by the straight vortex filament. The velocity field
close to the aerofoil is of course different from the field about a vortex filament
in cross flow, but both fields become more similar when the distance of the
vortex from the aerofoil becomes large.
In the same manner, a starting vortex can be assumed to be a straight
vortex filament which is attached to the bound vortex at plus and minus
infinity. The circulation of the vortex determines the lift, and the lift formula
which gives the relation between circulation Γ, and lift per unit width l in
inviscid potential flow is the Kutta–Joukowski theorem, namely

l = − ρ ΓU∞ (5.53)

where l is the lift per unit span of the wing, Γ is circulation around the wing,
U∞ is the freestream velocity and ρ is the density of the flow.
It is important to note that, the lift force on a wing section in inviscid
(potential) flow is perpendicular to the direction of the undisturbed stream
and thus, an aerofoil experiences only lift and no drag. This result is, of
course, contrary to the actual situation where the wing experiences drag also.
This is because here in the present approach the viscosity of air is ignored,
whereas in reality air is a viscous fluid. The Kutta–Joukowski theorem in the
form of Eq. (5.53) with constant Γ holds only for wing sections in a two-
dimensional plane flow. In reality, all wings are of finite span and, hence, the
flow essentially becomes three-dimensional. But as long as the span is much

ISTUDY
274 CHAPTER 5. VORTEX THEORY

larger than the chord of the wing section, the lift can be estimated assuming
constant circulation Γ along the span. Thus, the lift of the whole wing span
2b is given by
L = −ρ ΓU∞ 2b (5.54)
But in reality there is flow communication from the bottom to the top at
the wing tips, owing to higher pressure on the lower surface of the wing than
the upper surface. Therefore, by Euler equation, the fluid flows from lower to
upper side of the wing under the influence of the pressure gradient, in order to
even out the pressure difference. In this way, the magnitude of the circulation
on the wing tips tends to become zero. Therefore, the circulation over the
wing span varies and the lift is given by
∫ +b
L = −ρ U∞ Γ(x) dx (5.55)
−b

where the origin is at the middle of the wing, x is measured along the span,
and b is the semi-span of the wing.
According to Helmholtz’s first vortex theorem, being purely kinematic,
the above relations for lift are valid for the bound vortex also. Thus, isolated
pieces of a vortex filament cannot exist. Also, it can not continue to be straight
along into infinity, where the wing has not cut through the fluid and, thus,
no discontinuity surface has been generated as is necessary for the formation
of circulation. Therefore, free vortices Γt , which are carried away by the flow
must be attached at the wing tips. Together with the bound vortex Γb , and
the starting vortex Γs , they (the tip vortices) form a closed vortex ring frame
in the fluid region cut by the wing, as shown in Figure 5.25. If a long time has
passed since start-up, the starting vortex is at infinity (far downstream of the
wing), and the bound vortex and the tip vortices together form a horseshoe
vortex.

Γt
Tip vortex

2b Γb
Bound vortex

Γs
Starting vortex

Γt

FIGURE 5.25 Simplified vortex system of a finite wing.

Even though the horseshoe vortex system represents only a very rough
model of a finite wing, it can provide a qualitative explanation for how a
wing experiences a drag in inviscid flow, as already mentioned. The velocity

ISTUDY
5.9. BIOT–SAVART LAW 275

w induced at the middle of the wing by the two tip vortices accounts to
double the velocity induced by a semi-infinite vortex filament at distance b.
Therefore, by Eq. (5.50), we have
Γ Γ
w= (1 + 0) = (5.56)
4πb 4πb
This velocity is directed downwards and, hence, termed induced downwash.
Thus, the middle of the wing experiences not only the freestream velocity
U∞ , but also a velocity u, which arises from the superposition of U∞ and
downwash velocity w, as shown in Figure 5.26.
Di

L A

U∞ U∞
w
u

FIGURE 5.26 Illustration of induced drag.

In inviscid flow, the force vector is perpendicular to the actual approach


direction of the flow stream, and therefore has a component parallel to the
undisturbed flow, as shown in Figure 5.26, which manifests itself as the
induced drag Di , given by
w
Di = A (5.57)
U∞
It is important to note that, Eq. (5.57) holds if the induced downwash from
both vortices is constant over the span of the wing. However, the downwash
does change since at a distance x from the wing centre, one vortex induces a
downwash of
Γ
4 π (b + x)
whereas the other vortex induces
Γ
4 π (b − x)
Both the downwash are in the same direction, therefore, adding them, we get
the effective downwash as
Γ Γ
w= +
4 π (b + x) 4 π (b − x)
Γ 2b
=
4 π b2 − x2
Γ b
=
2 π b − x2
2

ISTUDY
276 CHAPTER 5. VORTEX THEORY

From this, it can be concluded that, the downwash is the smallest at the centre
of the wing [i.e., Eq. (5.57) underestimates the induced drag] and tends to
infinity at the wing tips. The unrealistic value there (at wing tips) does not
appear if the circulation distribution decreases towards the wing tips, as in
deed it has to. For a semi-elliptical circulation distribution over the span
of the wing, the downwash distribution becomes constant and Eq. (5.57) is
applicable. Helmholtz first vortex theorem stipulates that, for an infinitesimal
change in the circulation in the x-direction of


dΓ = dx
dx
and a free vortex of the same infinitesimal strength must leave the trail-
ing edge. This process leads to an improved vortex system, as shown in
Figure 5.27.

Γb Γt

Γt

FIGURE 5.27 Simplified vortex system of a wing.

The free vortices form a discontinuity surface in the velocity components


parallel to the trailing edge, which rolls them into the kind of vortices, as
shown in Figure 5.28. These vortices must be continuously renewed as the
wing moves forward. This calls for continuous replenishment of kinetic energy
in the vortex. The power needed to do this is the work done per unit time by
the induced drag.

FIGURE 5.28 Vortices formation due to rollup of the discontinuity.

The manifestation of Helmholtz’s first theorem can be encountered in daily


life. Recall the dimples formed at the free surface of coffee in a cup when a

ISTUDY
5.9. BIOT–SAVART LAW 277

spoon is suddenly dipped into it. The formation process of dimples looks like
that shown schematically in Figure 5.29. As the fluid flows together from the
front and back, a surface of discontinuity forms along the rim of the spoon.
The discontinuity surface rolls itself into a bow-shaped vortex whose endpoints
form the dimples on the free surface, as shown in the figure.

Surface Dimple

Spoon

FIGURE 5.29 Vortex formation due to dipping of a spoon.

The flow outside the vortex filament is a potential flow. Thus, by incom-
pressible Bernoulli equation, we have
1
p+ 2 ρ u2 + ρ g z = constant

This is valid both along a streamline and between any two points in the flow
field7 . Also, at the free surface the pressure is equal to the ambient pressure
pa . Further, at some distance away from the vortex the velocity is zero and
there is no dimple at the free surface and, hence, z = 0. Thus, the Bernoulli
constant is equal to pa and we have
1
ρ u2 + ρ g z = 0
2
Near the end points of the vortex the velocity increases by the formula given
by Eq. (5.52) and, therefore, z must be negative, that is, a depression of the
free surface. In reality, the cross-sectional surface of the vortex filament is not
infinitesimally small, therefore we cannot take the limit h → 0 in Eq. (5.52),
for which the velocity becomes infinite. However, the induced velocity due to
the vortex filament is so large that it causes a noticeable formation of dimples.
It should be noted that an infinitesimally thin filament cannot appear
in actual flow because the velocity gradient of the potential vortex tends to
infinity for h → 0, so that the viscous stresses cannot be ignored even for
very small viscosity. Also, it is well known that the viscous stresses make no
contribution to particle acceleration in incompressible potential flow, but they
do deformation work and thus provide a contribution to the dissipation. The
energy dissipated in heat stems from the kinetic energy of the vortex.
7 It would be of value to note that, for a steady, incompressible viscous flow, the Bernoulli

equation can be applied between any two points along a streamline only. But for a steady,
incompressible and inviscid (i.e., potential) flow, the Bernoulli equation can be applied
between any two points, in the entire flow field. That is, the two points between which the
Bernoulli equation is applied need not lie on a streamline.

ISTUDY
278 CHAPTER 5. VORTEX THEORY

5.9.4 Helmholtz’s Second Vortex Theorem


The second vortex theorem of Helmholtz’s states that, “a vortex-tube is al-
ways made up of the same fluid particles.” In other words, a vortex tube
is essentially a material tube. This characteristic of a vortex tube can be
represented as a direct consequence of Kelvin’s circulation theorem. Let us
consider a vortex tube and an arbitrary closed curve c on its surface at time
t0 , as shown in Figure 5.30. By Stokes’ integral theorem, the circulation of the
closed curve c is zero (i.e., DΓ/Dt = 0). The circulation of the curve, which
is made up of the same material particles, still has the same (zero) value of
circulation at a latter instant of time t.

c1

x (t )
c

Time t0 n

FIGURE 5.30 A closed curve on a vortex ring at times t0 and t.

By inverting the above reasoning it follows from Stokes’ integral theorem


that these material particles must be on the outer surface of the vortex tube.
If we examine smoke rings, it can be seen that the vortex tubes are material
tubes. The smoke will remain in the vortex ring and will be transported with
it, so that it is the smoke itself which carries the vorticity. This statement
holds under the restrictions of barotropy [i.e., ρ = ρ(p), the density is a
function of pressure only] and zero viscosity. The slow disintegration seen in
smoke rings is due to friction and diffusion. A vortex ring which consists of
an infinitesimally thin vortex filament induces an infinitely large velocity on
itself (similar to the horseshoe vortex), so that the ring would move forward
with infinitely large velocity. The induced velocity at the centre of the ring
remains finite (as in horseshoe vortex). From Biot–Savart law, the induced
velocity becomes
∫ 2π
Γ h2 dϕ Γ
u= =
4π 0 h3 2h
This velocity becomes infinitely large (i.e., unrealistic) when the cross-section
of the vortex ring is assumed to be infinitesimally small. For finite cross-
section, the velocity induced by the ring on itself, that is, the velocity with
which the ring moves forward remains finite. But in reality, the actual cross-
section of the ring is not known, and probably depends on how the ring was
formed.

ISTUDY
5.9. BIOT–SAVART LAW 279

In practice we notice that the ring moves forward with a velocity which
is slower than the induced velocity in the centre. Also, it is well known
that two rings moving in the same direction continually overtake each other
whereby one slips through the another in front. This phenomenon, illustrated
in Figure 5.31, is explained by mutually induced velocities on the rings and
formula given above for the velocity at the centre of the ring.

t = t1 t = t2 t = t3 t = t4

FIGURE 5.31 Two vortex rings passing through one another.

In a similar manner, it can be explained why a vortex ring towards a wall


becomes larger in diameter and at the same time its velocity gets reduced.
Also, the diameter decreases and the velocity increases when a vortex ring
moves away from a wall, as illustrated in Figure 5.32.

FIGURE 5.32 Kinematics of a vortex ring near a wall.

To work out the motion of vortex rings, the cross-section of vortex must
be known. Further, for infinitesimally thin rings, the calculation fails because
vortex rings, such as curved vortex filaments, induce large velocities on them-
selves. However, for straight vortex filaments, that is, for vortex filaments
in two-dimensional flows, a simple description of the “vortex dynamics” for
infinitesimally thin filaments is possible, since for such a case the self induced
translational velocity vanishes. We know that vortex filaments are material
lines, therefore it is sufficient to calculate the paths of the fluid particles which
carry the rotation in xy-plane perpendicular to the filaments, using

dx dxi
= u(x, t) or = ui (xi , t)
dt dt

ISTUDY
280 CHAPTER 5. VORTEX THEORY

i.e., to determine the paths of the vortex centres. The induced velocity which
a straight vortex filament at position xi induces at position x is known from
Eq. (5.49), i.e.,

Γ
v= (cos α + cos β)
4πh
As we have seen, the induced velocity is perpendicular to the vector hi = ri =
hi
(x − xi ) and, therefore, has the direction ez × , so that the vectorial form
|hi |
of Eq. (5.41) reads as
Γ x − xi
uR = ez ×
2π |x − xi |2
For x → xi , the velocity tends to infinity, but because of symmetry, the vortex
cannot be moved by its own velocity field, that is, the induced translational
velocity is zero. The induced velocity of n vortices with the circulation Γi (i =
1, 2, ....n) is
1 ∑ x − xi
uR = Γi ez ×
2π i |x − xi |2

If there are no internal boundaries, or if the boundary conditions are satisfied


by reflection, as in Figure 5.32, the last equation describes the entire velocity
field, and using dx/dt = u(x, t) or dxi /dt = ui (xi , t), the “equation of motion”
of the k th vortex becomes
dxk 1 ∑ x − xi
= Γ i ez × (5.58)
dt 2π |x − xi |2
i(i̸=k)

For i = k, the induced translational velocity becomes zero, owing to symmetry,


and, hence, excluded from the summation. Equation (5.58) gives the 2n
relations for the path coordinates.
The dynamics of vortex motion has invariants which are analogous to
the invariants of a point mass system on which no external forces act. ∑ The
conservation of strengths of the vortices by Helmholtz’s theorem ( Γk =
constant) corresponds to mass conservation of total mass of the point mass
system. When Eq. (5.58) is multiplied by Γk , summed over k and expanded,
we get

∑ dxk dx1 dx2 dx3


Γk = Γ1 + Γ2 + Γ3 + ...
dt dt dt dt
[
1 x1 − x2 x1 − x3
= ez × Γ1 Γ2 + Γ1 Γ 3 + ...
2π |x1 − x2 |2 |x1 − x3 |2
]
x2 − x1 x2 − x3
+ Γ2 Γ1 + Γ2 Γ3 + ...
|x2 − x1 |2 |x2 − x3 |2

ISTUDY
5.9. BIOT–SAVART LAW 281

In the above equation, the terms on the right-hand side cancel out in pairs,
and the equation reduces to
∑ dxk
Γk =0
dt
k

On integration, this results in


∑ ∑
Γk xk = xg Γk (5.59)
k k

The integration constants are written as xg , which is like a centre of gravity


coordinate (this is done here for dimensional homogeneity). Equation (5.59)
states that “the centre of gravity of the strengths of the vortices is conserved.”
For a point mass system, by conservation of momentum, we have the
corresponding law, namely, “the velocity of the centre of gravity is a conserved
quantity in the absence of external forces.”

For Γk = 0, the centre of gravity lies at infinity, so that, for example,
two vortices with Γ1 = − Γ2 must take a turn about a centre of gravity point
Pg which is at a finite distance, as shown in Figure 5.33.

Pg

Γ1 > 0
Γ1
Γ1 > 0

Pg
Γ2 = Γ1

Γ2 > 0
Γ2 > 0

FIGURE 5.33 Pathlines of a pair of straight vortices.

The paths of the vortex pairs are determined by numerical integration of


Eq. (5.58). The paths will look like as those shown in Figure 5.34.

Γ1 Γ2 = Γ1

t = t1 t = t2 t = t3
FIGURE 5.34 Pathlines of two straight vortex pairs.

5.9.5 Helmholtz’s Third Vortex Theorem


The third vortex theorem of Helmholtz’s states that “the circulation of a
vortex-tube remains constant in time.”

ISTUDY
282 CHAPTER 5. VORTEX THEORY

Using Helmholtz’s second theorem and Kelvin’s circulation theorem, the


above statement can be interpreted as “a closed line generating the vortex
tube is a material line whose circulation remains constant.”
Helmholtz’s second and third theorems hold only for inviscid and barotropic
fluids.

5.9.6 Helmholtz’s Fourth Vortex Theorem


The fourth theorem states that “the strength of a vortex remains constant in
time”.
This is similar to the fact that the mass flow rate through a streamtube
is invariant as the tube moves in the flow field. In other words, the circu-
lation distribution gets adjusted with the area of the vortex tube. That is
the circulation per unit area (that is vorticity) increases with decrease in the
cross-sectional area of the vortex tube and vice versa.

5.10 VORTEX MOTION


Vortex is a fluid flow in which the streamlines are concentric circles. The
vortex motions encountered in practice may in general be classified as free
vortex or potential vortex and forced vortex or flywheel vortex. The streamline
pattern for a vortex may be represented as concentric circles, as shown in
Figure 5.35.


r
θ
x

Streamline

Potential line
FIGURE 5.35 Streamline pattern of a vortex.

When a fluid flow is along a curved path, as in a vortex, the velocity of the
fluid elements along any streamline will undergo a change due to its change
of direction, irrespective of any change in magnitude of the fluid stream.
Consider a streamtube shown in Figure 5.36.
As the fluid flows round the curve, there will be a rate of change of
velocity, that is, an acceleration, towards the curvature of the streamtube.
The consequent rate of change of momentum of the fluid must be due to a

ISTUDY
5.10. VORTEX MOTION 283

FIGURE 5.36 Fluid element in a vortex.

force acting radially across the streamlines resulting from the difference of
pressure between the sides BC and AD of the streamline element, as per
Newton’s second law. The control volume ABCD in Figure 5.36 subtends
an angle δθ at the centre of curvature O and has length δs in the direction
of flow. Let the thickness of ABCD perpendicular to the plane of diagram
be ‘b’. For the streamline AD, the radius of curvature is r and that for BC
is (r + δr). The pressure and velocity at AD and BC are p, V, (p + δp) and
(V + δV ), as shown in Figure 5.36. Thus, the change of pressure in the radial
direction is δp.
The change of velocity in the radial direction (as shown in the velocity
diagram in Figure 5.36) is
δV = V δθ
But δθ = δs/r. Thus, the radial change of velocity between AB and CD is

δs
δV = V
r

Mass flow rate through the streamtube = density × area × velocity


= ρ(b × δr)V

Rate of change of momentum in radial direction


= mass per unit time × radial change of velocity
V δs
= (ρbδr V )
r (5.60)

This rate of change of momentum is produced by the force due to the


pressure difference between BC and AD of the control volume, given by

F = [(p + δp) − p] bδs (5.61)

ISTUDY
284 CHAPTER 5. VORTEX THEORY

According to Newton’s second law, Eq. (5.60) = Eq. (5.61). Thus,

V 2 δs
δpbδs = ρbδr
r
or
δp ρV 2
= (5.62)
δr r
For an incompressible fluid, density ρ is constant and Eq. (5.62) can be
expressed in terms of pressure head h. Pressure is given by

p = ρgh

Therefore,
δp = ρgδh
Substituting this into Eq. (5.62), we get

δh ρV 2
ρg =
δr r
or
δh V2
=
δr gr
In the limit δr → 0, this gives the rate of change of pressure head in the radial
direction as
dh V2
= (5.63)
dr gr
The curved flow shown in Figure 5.36 will be possible only when there is a
change of pressure head in a radial direction, as seen from Eq. (5.63). How-
ever, since the velocity V along streamline AD is different from the velocity
(V + δV ) along BC, there will also be a change in the velocity head from one
streamline to another. Such a change of velocity head in the radial direction
is given by

Rate of change of velocity head in the radial direction


[ ]
(V + δV )2 − V 2
=
(2 g δr)

V δV
= (neglecting the products of small quantities)
g δr
V dV
= (as δr → 0) (5.64)
g dr

For a planar flow (say in the horizontal plane), the changes in potential head
is zero. Therefore, the change of total head H, that is, the total pressure
energy per unit weight, in a radial direction, δH/δr, is given by

ISTUDY
5.11. FORCED VORTEX 285

δH
= change of pressure head + change of velocity head
δr
Substituting Eqs. (5.63) and (5.64) into the results of the above equation, we
get

dH V2 V dV
= +
dr gr g dr
( )
V V dV
= + (5.65)
g r dr
( )
V dV
The term + is also known as vorticity of the flow.
r dr
In obtaining Eq. (5.65), it is assumed that the streamlines are horizontal.
But this equation also applies to cases where the streamlines are inclined to
horizontal, since the fluid in a control volume is in effect weightless, being
supported vertically by the surrounding fluid.

5.11 FORCED VORTEX


Forced vortex is a rotational flow field in which the fluid rotates as a solid
body with a constant angular velocity ω, and the streamlines form a set of
concentric circles. Because the fluid in a forced vortex rotates like a rigid
body, the forced vortex is also called flywheel vortex. The change of total
energy per unit weight in a vortex motion is governed by Eq. (5.65).
The velocity at any radius r is given by

V = ωr

From this, we have


dV

dr
and
V

r
Substituting dV /dr and V /r into Eq. (5.65), we get

dH ωr
= (ω + ω)
dr g

2ω 2 r
=
g

Integrating this, we get


ω 2 r2
H= +c (5.66)
g

ISTUDY
286 CHAPTER 5. VORTEX THEORY

where c is a constant. By Bernoulli equation, at any point in the fluid, we


have

p V2
H= + +z
ρg 2g

Note that, in the above equation and Eq. (5.66), the unit of the total head is
meters. Substitution of this into Eq. (5.66), results in

p ω 2 r2 ω 2 r2
+ +z = +c
ρg 2g g

p ω 2 r2
+z = +c
ρg 2g

If the rotating fluid has a free surface, the pressure at the surface will be
atmospheric; therefore, the pressure at the free-surface will be zero.
p
Replacing with 0 in the above equation, the profile of the free surface
ρg
is obtained as

ω 2 r2
z= +c (5.67)
2g

Thus, the free surface of a forced vortex is in the form of a paraboloid.


Similarly, for any horizontal plane, for which z will be constant, the pres-
sure distribution will be given by

p ω 2 r2
= + (c − z) (5.68)
ρg 2g
Axis of rotation

Velocity
e
ac

variation
rf
su

r V ∼r
ee
Fr

Streamline
Datum
(a) (b)
FIGURE 5.37 Forced vortex (a) shape of free surface, (b) velocity
variation.

The typical shape of the free surface and the velocity variation along a
radial direction of a forced vortex are shown in Figure 5.37.

ISTUDY
5.12. FREE VORTEX 287

5.12 FREE VORTEX


Free vortex is an irrotational flow field in which the streamlines are concentric
circles, but the variation of velocity with radius is such that there is no change
of total energy per unit weight with radius, so that dH/dr = 0. Since the flow
field is potential, the free vortex is also called potential vortex.
For a free vortex, from Eq. (5.65), we have
( )
V V dV
0= +
g r dr

dV dr
+ =0
V r
Integrating, we get
ln V + ln r = constant
or
Vr =c
where c is a constant known as the strength of the vortex at any radius r.
The tangential velocity becomes
c
V = (5.69)
r
This shows that, in the flow around a vortex core the velocity is inversely
proportional to the radius (see Section 5.4). When the core is small, or as-
sumed concentrated on a line axis, it is apparent from the relation V = c/r
that when r is small V can be very large. However, within the core the fluid
behaves as though it were a solid cylinder and rotates at an uniform angular
velocity. Figure 5.38 shows the variation of velocity with radius for a typical
free vortex. The solid line represents the idealized case, but in reality it is not
so precise, and the velocity peak is rounded off, as shown by the dashed lines.

Γ
V =
2πr

The core
FIGURE 5.38 Velocity distribution in a free vortex core.

At any point in the flow field, by Bernoulli equation, we have


p V2
+ +z =H
ρg 2g

ISTUDY
288 CHAPTER 5. VORTEX THEORY

Substituting Eq. (5.69), we get

p c2
+ +z =H
ρg 2gr2
p
At the free surface, = 0. Thus, the profile of the free surface is given by
ρg

c2
H −z = (5.70)
2gr2

This is a hyperbola asymptotic to the axis of rotation and to the horizontal


plane through z = H, as shown in Figure 5.39.

Velocity
variation
H
z r V ∝ r1

(a) (b)

FIGURE 5.39 Free vortex: (a) shape of free surface, (b) velocity
variation.

For any horizontal plane, z is constant and the pressure variation is given
by
p c2
= (H − z) − (5.71)
ρg 2gr2
Thus, in a free vortex, pressure decreases and circumferential velocity
increases as we move towards the centre, as shown in Figure 5.39.
The free vortex discussed above is essentially a free cylindrical vortex. The
fluid moves along streamlines that are horizontal concentric circles; there is
no variation of total energy with radius. Combination of a free cylindrical
vortex and radial flow will result in a free spiral vortex.

5.12.1 Free Spiral Vortex


A free spiral vortex is essentially the combination of a free cylindrical vortex
and a radial flow. Before getting in to the physics of free spiral vortex, let us
see what is a radial flow.

ISTUDY
5.12. FREE VORTEX 289

Radial Flow
Examine the flow between two parallel planes as shown in Figure 5.40. In
the flow, the streamlines will be radial straight lines and the streamtube will
be in the form of sectors. This kind of flow in which the fluid flows radially
inwards, or outwards from a centre is called a radial flow. The area of the
flow will therefore increase as the radius increases, causing the velocity to
decrease. The flow pattern is symmetrical and therefore, the total energy per
unit weight H will be the same for all streamlines and for all points along
each streamline if we assume that there is no loss of energy.

R r
R1
p2 p1 p2
V1 V1 b
V2 V2

p2 pr
p1

R2 R1
r
r
R2
δr

(a) (b)
FIGURE 5.40 A radial flow: (a) streamlines, (b) pressure variation.

If Vr is the radial velocity and p is the pressure at any radius r, then the
total energy per unit weight H becomes

p V2
H= + r = constant (5.72)
ρg 2g

Assuming the flow to be incompressible (as would be the case of a liquid), by


continuity, we have the volume flow rate Q̇ as

Q̇ = area × velocity

= 2πrb × Vr

where b is the distance between the planes. Thus,


Vr = (5.73)
2πrb

ISTUDY
290 CHAPTER 5. VORTEX THEORY

Substituting this into Eq. (5.72), we get

p Q̇2
+ 2 2 2 =H
ρ g 8π r b g
[ ( ) ( )]
Q̇2 1
p = ρg H − ×
8π 2 b2 g r2

The plot of pressure p at any radius will be, as shown in Figure 5.40(b),
parabolic and is some times referred to as Barlow’s curve.
If the radial flow discharges to atmosphere at the periphery, the pres-
sure at any point between the two plates will be below atmospheric (i.e.,
subatmospheric); there will be a force tending to bring the plates together
and so shut-off the flow. This phenomenon can be observed in the case of a
disc valve. Radial flow under the disc will cause the disc to be drawn onto the
valve seating. This will return to atmospheric and the static pressure of the
fluid on the upstream side of the disc will push it off its seating again. The
disc will tend to vibrate on the seating and the flow will be intermittent.
Now, let us find an expression for the pressure difference between two
points on the same horizontal plane in a free vortex. For a free cylindrical
vortex, the streamlines are concentric circles and there is no variation of the
total energy with radius, i.e.,

dH
=0
dr
Also, by Eq. (5.69), we have
c
V =
r
Let p1 and p2 be the pressures in two concentric streamlines of radius r1 and
r2 which have velocities V1 and V2 . Since there is no change of total energy
with radius, for the same horizontal plane, by Bernoulli equation

p1 V2 p2 V2
+ 1 = + 2
ρg 2g ρg 2g

p1 − p2 V 2 − V12
= 2
ρg 2g

But V1 = c/r1 and V2 = c/r2 . Thus,


( )
p1 − p2 c2 1 1
= − 2 (5.74)
ρg 2g r22 r1

Now, let us obtain an expression for the pressure difference between two
points at radii R1 and R2 , on a radial line, when a fluid flows radially inward
or outward from a centre, neglecting friction. Flow is radial and, therefore,

ISTUDY
5.12. FREE VORTEX 291

in straight line, so that r the radius of curvature of the streamlines is infinite,


dH/dr = 0, and for all streamlines

p V2
H= + = constant
ρg 2g
If p1 and p2 are the pressures at radii R1 and R2 , respectively, where the
velocities are V1 and V2 ,
p1 − p2 V 2 − V12
= 2
ρg 2g
By volume conservation,

Q̇ = 2πR1 V1 t = 2πR2 V2 t

where Q̇ is the volume flow rate and t is the distance between the radial
passage boundaries, i.e.,


V1 =
2πR1 t


V2 =
2πR2 t
Thus,
( )
p1 − p2 Q̇2 1 1
= − 2 (5.75)
ρg 8π 2 t2 g R22 R1
It is evident from Eqs. (5.73) and (5.74) that the relation between pressure
and radius and between velocity and radius is similar for both free vortex and
radial flow. Both types of motion may therefore occur together. The fluid
rotates and flows radially forming a free spiral vortex in which a fluid element
will follow a spiral path, as shown in Figure 5.41.

α
v

Path of a particle
FIGURE 5.41 A free spiral vortex.

ISTUDY
292 CHAPTER 5. VORTEX THEORY

5.13 COMPOUND VORTEX


In the free vortex, v = c/r and thus, theoretically, the velocity becomes
infinite at the centre. The velocities near the axis would be very high and,
skin friction losses vary as the square of the velocity, they will cease to be
negligible. Also, the assumption that the total head H remains constant will
cease to be true. The portion of fluid around the axis tends to rotate as a
solid body. Thus, the central portion essentially forms a forced vortex. The
free surface profile of such a compound vortex and the pressure variation with
radius on any horizontal plane in the vortex is shown in Figure 5.42.

FIGURE 5.42 Compound vortex.

The velocity at the common radius R must be the same for the two vortices.
For the free vortex, if y1 = depression of the free surface at radius R below
the level of the surface at infinity, then
c2
y1 =
2gR2

v2
=
2g

ω 2 R2
=
2g
For the forced vortex, if y2 is the height of the free surface at radius R above
the centre of the depression,
v2
y2 =
2g

c2
=
2gR2

ISTUDY
5.14. PHYSICAL MEANING OF CIRCULATION 293

Thus, the total depression is

c2 ω 2 R2
y1 + y2 = 2
= (5.76)
gR g

For the forced vortex, the velocity at radius R is ωR, while for the free vortex,
from Equation (5.69), the velocity at radius R is c/R. Therefore, the common
radius at which these two velocity will be the same is given by
c
ωR =
R

c
R=
ω

5.14 PHYSICAL MEANING OF


CIRCULATION
In Equation (5.3), the vector field is taken as velocity V . But the vector field
need not be the flow velocity alone. The vector field can be force, mass flow
rate of a fluid, etc. Therefore, in general, the circulation may be defined as
the line integral of a vector field around a Hclosed plane curve in a flow field.
If the vector is a force F , then the integral c F · ds is equal to the work done
by the force. Taking the vector as ρV , the mass flow rate per unit area, we
can get a physical meaning of the circulation in the following way. Imagine a
tiny paddle-wheel probe is placed [Figure 5.43(c)] in any of the flow patterns
shown in Figure 5.43.

(a) (b)

Paddle-wheel
probe

(c) (d)
FIGURE 5.43 Streamlines in: (a) a vortex, (b) parallel flow with constant
velocity, (c) parallel flow with variable velocity, (d) Flow around a corner.

ISTUDY
294 CHAPTER 5. VORTEX THEORY

When the flow velocity on one side is greater than the other side, as in
Figure 5.43(c), the wheel will turn. If the mass flow rate per unit area ρV is
larger on one side of the wheel than the other, then the circulation is different
from zero, but if ρV is the same on both sides as in Figure 5.43(b), then the
circulation is zero. We shall show that the component of curl ṁ along the
axis of the paddle wheel equals
I
1
lim ṁ · ds (5.77)
dA→0 dA

where dA is the area enclosed by the curve along which we calculate circula-
tion. The paddle wheel then acts as a “curl meter” to measure curl ṁ; when
the wheel does not rotate, curl ṁ = 0. In Figure 5.43(c), curl ṁ ̸= 0 in spite
of the fact that the streamlines are parallel. In Figure 5.43(d), it is possible
to have curl ṁ = 0 even though the streamlines go around a corner. In fact,
for the flow of water around a corner curl ṁ = 0. We must realize that the
value of curl ṁ at a point depends upon the circulation in the neighbourhood
of the point and not on the overall flow pattern.

EXAMPLE 5.2 A cylindrical tank of 1 m diameter and 0.7 m height is


filled with a liquid of relative density 0.9 up to 0.5 m from the bottom of the
tank and the rest of the tank contains atmospheric air. The tank revolves
about its vertical axis at a speed such that the liquid begins to uncover the
base. (a) Calculate the speed of rotation and (b) the upward force on the
cover.

Solution The flow field described is shown in Figure 5.44.

FIGURE 5.44 Flow field in a rotating tank.

(a) When the tank is static the volume of oil is

πR2 z1

ISTUDY
5.14. PHYSICAL MEANING OF CIRCULATION 295

While rotating a forced vortex is formed and the free surface will be a
paraboloid CGD.
Volume of oil = Volume of cylinder ABFE − Volume of paraboloid CGD

1
= πR2 z − πr12 z
2
Since the volume of the paraboloid is equal to the half the volume of the
circumscribing cylinder.
No oil is spilled out from the cylinder, therefore,
1
πR2 z1 = πR2 z − πr12 z
2
( z1 )
r12 = 2 R2 1 −
z
√ ( )
0.5
r1 = R 2 1 −
0.75

= 0.816 R

= 0.816 × 0.5 = 0.408 m

= 408 mm

For the free surface, by Eq. (5.67), we have

ω2 r2
z= + constant
2g
Between C and G, taking G is the datum level, we have

zA = 0, when r = 0

and
zC = z, when r = r1
Thus,

ω 2 r2
z=
2g

2gz
ω=
r12

2 × 9.81 × 0.75
=
4082
= 9.4 rad/s

ISTUDY
296 CHAPTER 5. VORTEX THEORY

9.4 × 60
=

= 89.76 rpm

(b) The top cover annular area from r = r1 to r = R is in contact with the
oil. If p is the pressure at any radius r, the force on an annular of radius r
and width dr is given by

dF = p × 2πrdr
Integrating from r = r1 to r = R, we get the force F acting on the top cover
as ∫ R
F = 2π prdr
r1

The pressure, given by Equation (5.68) is

p ω 2 r2
= +c
ρg 2g

since the pressure at r1 is atmospheric, p = 0 (gauge), where r = r1 , thus,

−ω 2 r12
c=
2g
Therefore, [ ]
ω2 r2 ω 2 r12 ρω 2 ( 2 )
p = ρg − = r − r12
2g 2g 2
Substituting this in the force equation above, we have

ρω 2 R ( )
F = 2π r2 − r12 rdr
2 r1

[ ]R
r4 r2 2
= ρω 2 π − r
4 2 1 r1

ρω 2 π [ 2 ]2
= R − r12
4
π ( ) ( )2
= 0.9 × 103 × 9.42 0.52 − 0.4082
4

= 435.85 N

EXAMPLE 5.3 Show that a free vortex is an irrotational motion. A hollow


cylinder of diameter 1 m, open at the top, spins about its axis which is vertical,
thus generating a forced vortex motion of the liquid contained in it. Calculate

ISTUDY
5.14. PHYSICAL MEANING OF CIRCULATION 297

the height of the vessel so that the liquid just reaches the top when the mini-
mum depth of the free surface of the liquid (from the top) is 25 cm at 200 rpm.

Solution The tangential velocity in a free vortex (excluding the core) is


given by Eq. (5.69) is
c
Vθ =
r
where c is a constant. This field is potential if the vorticity content in the
field is zero. The vorticity ζ in polar coordinates is [Eq. (5.2)]

qt ∂qt 1 ∂qn
ζ= + −
r ∂r r ∂θ
In free vortex, the normal component of velocity qn = 0. Thus,

qt ∂qt
ζ= +
r ∂r
Therefore,
qt Vθ c2
= = 2
r r r
and
∂qt ∂ (c) c
= =− 2
∂r ∂r r r
Hence,
c c
ζ= 2
− 2
r r
=0

Thus, the motion in a free vortex is irrotational.


The spinning cylinder described in the problem is shown schematically in
Figure 5.45.

FIGURE 5.45 A spinning cylinder containing water.

ISTUDY
298 CHAPTER 5. VORTEX THEORY

For a forced vortex, the free surface ABC is a paraboloid, as shown in the
figure. The shape is given by Eq. (5.67) as

ω 2 r2
z= +c
2g
At r = 0 on the free surface, zB = 25 cm = 0.25 m. Thus

c = 0.25 m

At r = R, zA = h. Therefore, we have the expression for free surface as


ω 2 R2
h= + 0.25
2g
Given, ω = 200 rpm = 200/60 rps = 200/60 × 2π rad/s. Therefore,

(20 π)2 0.52


h= + 0.25
32 2 × 9.81

= 5.84 m

5.15 RECTILINEAR VORTICES


A rectilinear vortex is a vortex tube whose generators are perpendicular to
the plane of motion. Now, let us have a closer look at some aspects of two-
dimensional vortex motion. We know that the circulation in an infinitesimal
plane circuit is proportional to the area of the circuit. In a two-dimensional
motion, the vorticity vector ζ at any point P which is perpendicular to the
plane of motion and whose magnitude is equal to the limit of the ratio of the
circulation is an infinitesimal circuit embracing P to the area of the circuit.
That is, the vorticity vector is by definition perpendicular to the plane of the
motion, so that the vortex lines are straight and parallel. All vortex tubes
are therefore cylinders whose generators are perpendicular to the plane of
motion. Such vortices are known as rectilinear vortices. For our discussions
in this section, let us consider the fluid between parallel plates at unit distance
apart and parallel to the plane of the motion, which is half-way between them.

5.15.1 Circular Vortex


A circular vortex is that with cross-section normal to its axis is a circle.
For example a single cylindrical vortex tube, whose cross-section is a circle
of radius ‘a’, surrounded by unbounded fluid, as shown in Figure 5.46 is a
circular vortex.
The section of this cylindrical vortex by the plane of motion is a circle, as
shown in Figure 5.47.
Let us assume that the vorticity ζ over the area of the circle is a constant.
Outside the circle the vorticity is zero. Consider concentric circles of radii r′

ISTUDY
5.15. RECTILINEAR VORTICES 299

FIGURE 5.46 A cylindrical vortex tube.

V′
ω o r′

FIGURE 5.47 Section of cylindrical vortex tube.

and r, where r′ < a < r, as shown in Figure 5.47. Let the tangential speeds
of the fluid motion on the circles of radii r′ and r be V ′ and V , respectively.
We know that, “the circulation in a closed circuit is the line integral of
the tangential component of the velocity taken round the circuit in the sense
in which the arc length (elemental length along the circuit) increases”. Thus,
the circulation around the circles of radii r and r′ , respectively, are
I I
V ds and V ′ ds

where ds is an elemental arc length. Also, V and V ′ are constants. Therefore,


I
V ds = ωπa2 , r > a
I
V ′ ds = ωπr′2 r < a

where ω is the angular velocity. Thus,


2πrV = ωπa2

ωa2
V = , r>a
2r
2πr′ V ′ = ωπr′2

ωr′ ′
V = , r <a
2

ISTUDY
300 CHAPTER 5. VORTEX THEORY

when r′ = r = a, we have V ′ = V = 12 aω, so that the velocity is continuous


as we pass through the circle.
From the above discussions, it can be inferred that the existence of a
vortex implies the co-existence of certain distribution of velocity field. This
velocity field which co-exists with the vortex is known as the induced velocity
field and the velocity at any point of it is called the induced velocity. It is
important to note that, it is customary to refer to the velocity at a point in
the field as the velocity induced by the vortex, but it is merely a convenient
abbreviation of complete statement that were the vortex alone in the otherwise
undisturbed field the velocity at the point would have the value in question. In
other words, when several vortices are present in the field, each will contribute
to the velocity at a point.
For circular vortex, the induced velocity at the extremity of any radius
vector r joining the centre of the vortex to a point of the fluid external to the
vortex is of magnitude inversely proportional to r and is perpendicular to r.
Thus the induced velocity tends to zero at great distances. The fluid inside
the vortex will have velocity of magnitude proportional to r and therefore the
fluid composing the vortex moves like a rigid body rotating about the centre
O with angular velocity ω = 21 ζ. The velocity at the centre is zero, i.e., “a
circular vortex induces zero velocity at its centre.”
Thus, a circular vortex alone in the otherwise undisturbed fluid will not
tend to move.
The velocities at the extremities of oppositely directed radii are of the
same magnitude but of opposite sense so that the mean velocity of the fluid
within the vortex is zero. Thus, if a circular vortex of small radius be “placed”
at a point in a flow field where the velocity is u, the mean velocity at its centre
will still be u and the fluid composing the vortex will move with velocity u,
i.e., the vortex will move with the stream carrying its vorticity with it.
Naturally occurring tropical cyclone, hurricane and typhoon which attain
a diameter of from 150 to 800 km, and travel at a speed seldom exceeding
25 km/h are circular vortices on a large scale. Within the area the wind can
reach hurricane force, while there is a central region termed “the eye of the
storm”, of diameter about 15 to 30 km where conditions may be completely
calm.

5.16 VELOCITY DISTRIBUTION


Consider a circular vortex of strength γ, defined by

2πγ = circulation = πa2 ζ

Thus,
1 2
γ= a ζ
2

ISTUDY
5.16. VELOCITY DISTRIBUTION 301

Therefore, the velocity induced by this circular vortex is

r′
V′ =γ , r′ < a
a2

γ
V = , r>a
r
The velocities at all points of a diameter are perpendicular to that diameter,
hence the extremities of the velocity vectors at the different points of the
diameter will lie on a curve which gives the velocity distribution as we go
along the diameter from −∞ to +∞, as shown in Figure 5.48.

FIGURE 5.48 Velocity distribution along a diameter of a circular vortex.

From points between C and D the velocity variation (along a diameter)


is a straight line EAF , the variation from C to −∞ and D to +∞ is part
of a rectangular hyperbola whose asymptotes are the diameter CD and the
perpendicular diameter through the centre A. The ordinate DE, CF each
represent the velocity γ/a. Thus, if for a circular vortex of constant strength
γ, as the radius a decreases, DE will increase. Therefore, in the limit when
a → 0 the velocity variation will consists of the rectangular hyperbola with
the asymptote perpendicular to CD.
Now let us study the field of two identical circular vortices of radius a
but with opposite vorticity (ζ and −ζ) at a finite distance apart, as shown
in Figure 5.49. If the distance between their centre BA is sufficiently large
compared to a, as a first approximation, we can suppose that the vortices do
not interfere and remain circular. For such a case their velocity fields may
be compounded by the ordinary law of vector addition. The effect on the
velocity distribution plot of A will reduce all the velocities at points near A
on the diameter CD (refer Figure 5.48) by approximately v = γ/BA. The
general shape of the velocity distribution plot for the pair of vortices will be
as shown in Figure 5.49.
It can be seen that the centre of each vortex is now in the field of velocity
induced by the other and would therefore move with velocity v in the direction
perpendicular to AB. Thus the vortices are no longer at rest, but move with
equal uniform velocity, remaining at a constant distance apart. This is an
application of the theorem that “a vortex induces no velocity on itself”. The

ISTUDY
302 CHAPTER 5. VORTEX THEORY

FIGURE 5.49 General shape of the velocity distribution for a pair of


vortices.

vortices and velocity field shown in Figure 5.49 has its application to the study
of the induced velocity due to the wake of a monoplane aerofoil at a distance
behind the trailing edge.

5.17 SIZE OF A CIRCULAR VORTEX


It can be shown that the pressure at the centre of a circular vortex of strength
γ and radius a is the lowest pressure in the field of the vortex and the value is
(p∞ − γρ2 /a2 ), where p∞ is the pressure at infinity and ρ is the local density.
Therefore, if the pressure in the field to be positive everywhere, a2 ≤ γ 2 ρ/p∞ .
That is, the radius of the vortex should be greater than or equal to γ 2 ρ/p∞ .
But in many occasions we are concerned with the case of a → 0. In such a
case the resulting point vortex must be regarded as an abstraction. However,
we can make a as small as we wish by making γ small enough, or p∞ large
enough, but we shall still have a circular vortex and the induced velocity will
be everywhere finite. The apparently infinite velocities which occur are due
to the over-simplification of taking a = 0. Note that a similar lower limit is
encountered for the size of a point source in two-dimensional motion, and is
given by the same relation if γ is taken as the strength of the source.

5.18 POINT RECTILINEAR VORTEX


It is the limiting case of a circular vortex of constant strength γ with radius
a tending to zero. We have seen that any point outside a circular vortex of
strength γ, at distance r from the centre, the velocity γ/r is at right angles
to r. If we let the radius a of the vortex tend to zero, the circle shrinks to a
point. This limiting vortex of zero radius is called a point rectilinear vortex,
or simply a point vortex of strength γ. When the radius tends to zero the
cylindrical vortex tube shown in Figure 5.46 shrinks to a straight line and the
vortex becomes a single rectilinear vortex represented by a point in the plane
of motion, as shown in Figure 5.50.
If we take the origin at the vortex, the velocity at the point P (r, θ) is
represented by a complex number

γ i(θ+ 1 π) ik
e 2 = −iθ
r re

ISTUDY
5.18. POINT RECTILINEAR VORTEX 303

γ/r

P
γ
r
θ θ+ π
2
o

FIGURE 5.50 Point rectilinear vortex at origin in xy-plane.

We can relate this to the complex potential w = ϕ + iψ as follows:


dw dw ∂z
=
dx dz ∂x

But z = x + iy, therefore,


dw ∂ϕ ∂ψ
= +i
dx ∂x ∂x
= −u + iv
since u = ∂ϕ/∂x, v = −∂ψ/∂x, or
dw
− = u − iv
dz
Thus,
dw
− = u − iv
dz
−ik −ik
= =
reiθ z
Integrating this, we get the complex potential w as
w = ik ln z + constant
Here the constant is irrelevant and hence can be ignored, then
w = ik ln z
Note that the motion is irrotational except at the origin O where the vortex
is positioned and so a complex potential exists, with a logarithmic singularity
at the vortex.
If the vortex were at the point z0 instead of at the origin, the complex
potential would be
w = ik ln (z − z0 )
It is essential to note that the velocity derived from the complex potential is
the velocity induced by the vortex.

ISTUDY
304 CHAPTER 5. VORTEX THEORY

5.19 VORTEX PAIR


Two vortices of equal strength γ but opposite nature (one rotating clockwise
and the other rotating counterclockwise) from a vortex pair, as shown in
Figure 5.51.

FIGURE 5.51 A vortex pair.

Each vortex in the pair induces a velocity γ/AB on the other, in the direc-
tion perpendicular to AB and in the same sense. Thus the vortex pair moves
in the direction perpendicular to AB, remaining at the constant distance AB
apart. The fluid velocity at O, the mid-point of AB, is

2γ 2γ 4γ
+ =
AB AB AB

which is four times the velocity of each vortex (refer Figure 5.49).
Taking O as the origin and the x-axis along OA, if AB = 2a, we have the
complex potential, at the instant when the vortices are on the x-axis, as

w = iγ ln (z − a) − iγ ln (z + a) (5.78)

Thus,
( )
1 1
u − iv = iγ −
z−a z+a

With y = 0, this gives the velocity distribution along the x-axis as

2aiγ
u − iv =
x2 − a2

Thus u = 0, v = −2aγ/(x2 − a2 ). The plot of v against x is shown in


Figure 5.52.
The curve is as per the equation v(x2 −a2 ) = −2aγ, so that the asymptotes
are the straight portions of Figure 5.52 go over into the asymptotes x ± a and
thus the velocity of vortex A cannot be reached in Figure 5.52, although this
is still one-quarter of the velocity at O.

ISTUDY
5.20. IMAGE OF A VORTEX IN A PLANE 305

5.20 IMAGE OF A VORTEX IN A PLANE


For a vortex shown in Figure 5.51, because of symmetry, there will not be any
flux across yy ′ the perpendicular bisector of AB. Thus, yy ′ can be regarded
as a streamline and could therefore be replaced by a rigid boundary. Hence
the motion due to a vortex at A in the presence of this boundary is the same
as the motion that would result if the boundary were removed and an equal
vortex of opposite rotation were placed at B. The vortex at B is called the
image of the actual vortex at A with respect to the plane boundary and the
complex potential is still given by Eq. (5.78).

γ γ
x
B o A

FIGURE 5.52 Variation of v with x.

5.21 VORTEX BETWEEN PARALLEL PLATES


Let us consider a vortex of strength γ midway between the planes y = ±a/2
and at the origin, as shown in Figure 5.53.

y η

a/2 γ +i ξ
x
a/2 γ −i
γ

(a) (b)
FIGURE 5.53 (a) A vortex between two plates, (b) a vortex and its
image.

The transformation ζ = ieπz/a would map the strip between the planes on
the upper half of the ζ-plane (the thick and thin lines in Figures 5.53(a) and
5.53(b) indicate which parts of the boundaries correspond) as follows:
π
ζ = ψ + iη = ie a (x+iy)
[ π π ]
= i cos (x + iy) + sin (x + iy)
a a

ISTUDY
306 CHAPTER 5. VORTEX THEORY

For z = x + iy = 0, ζ = i. That is,

ψ = 0 and η = i

Thus, by the image system, we have vortices of strength γ at ζ = i and −γ


at ζ = −i, as illustrated in Figure 5.53(b). Therefore, by Eq. (5.78),
ζ −i
w = iγ ln
ζ +1
πz
e a −1
= iγ πz
e a +1
( πz )
= iγ ln tanh
2a
But
dw
= u − iv
dz
Thus, we have
π 1 ( πz )
u − iv = iγ ( πz ) × coth
2a cosh2 2a
2a
π 1
= iγ ( )
a sinh πz
a
Thus when y = 0,
γπ 1
u = 0, v = − ( )
a sinh πx
a
and the velocity at different points on the x-axis is given by this relation.
When there are no walls present, on the x-axis, v = γ/x. Thus,
v πx 1
=− × ( πx ) < 1
v0 a sinh
a
Therefore, the walls reduce the velocity v at points on the x-axis. For example,
for x = a,
v π
=−
v0 sinh π
π
=−
(eπ − e−π )/2
π
=−
11.549
= −0.272

ISTUDY
5.22. FORCE ON A VORTEX 307

FIGURE 5.54 Streamlines of a vortex at origin, bounded by two parallel


planes.

The streamlines of a vortex at the origin between two parallel plates would
be as shown in Figure 5.54.
Note that the walls increase the velocity component u when x = 0 and
decrease v when y = 0. In other words, the walls make the vortex to stretch
along the x-direction and shrink along the y-direction.

5.22 FORCE ON A VORTEX


A rectilinear vortex may be regarded as the limit of a circular vortex which
rotates about its centre as if rigid. Consider a circular vortex inserted in a
steady flow field as shown in Figure 5.55, so that its centre is at the point
whose velocity is (u0 , v0 ) before the vortex is inserted. The vortex would then
move with the fluid with velocity (u0 , v0 ) soon after inserting, so that the flow
motion would no longer be steady. Let us imagine the vortex to be held fixed
by the application of a suitable force (in the form of pressure distribution).
This force would be equal but opposite to that exerted by the fluid on the
vortex.

FIGURE 5.55 Circular vortex in a flow field.

When the motion is steady, the force exerted by the fluid is the Kutta-
Joukowski lift which is independent of the size and shape of the vortex. This
force, being independent of the size, is also the force exerted by the fluid on
a point vortex. The direction of the force (shown in Figure 5.55) is obtained
by rotating the velocity vector through a right angle in the opposite to that
of the circulation (vorticity).

ISTUDY
308 CHAPTER 5. VORTEX THEORY

5.23 MUTUAL ACTION OF TWO VORTICES


Consider two vortices of strength γ and γ ′ located at (0, 0) and (0, h), as
shown in Figure 5.56.

γ
γ/h

γ ′ /h
γ
FIGURE 5.56 Two vortices at a finite distance between them.

These two vortices repel one another if γ and γ ′ have the same sign, and
attract if the signs are opposite. This result has its application to the action
between the vortices shed by the wings of a biplane.

5.24 ENERGY DUE TO A PAIR OF


VORTICES
Consider two circular vortices of equal radius a and equal strength γ placed
as shown in Figure 5.57 with the distance 2b between their centres very large
compared to a, so that their circular form is preserved.

FIGURE 5.57 Two small circular vortices at a finite distance apart.

Neglecting the interaction between them, we can write the vorticity as


ζ = iγ ln (z − b) − iγ ln (z + b)
The stream function is (
)
r1
ψ = γ ln
r2
where r1 , r2 are the distance of the point z from the vortices, as shown in
Figure 5.57.
For the region external to the vortices, the kinetic energy of the fluid is
∫∫
1
KEo = ρ (u2 + v 2 ) dxdy
2

ISTUDY
5.24. ENERGY DUE TO A PAIR OF VORTICES 309

Now in terms of stream function ψ,


∂ψ ∂ψ
u2 + v 2 = u −v
∂y ∂x

∂(uψ) ∂(vψ) ∂u ∂v
= − −ψ +ψ
∂y ∂x ∂y ∂x
( )
∂(uψ) ∂(vψ) ∂v ∂u
= − +ψ −
∂y ∂x ∂x ∂y
But the region outside the vortices is irrotational and hence vorticity
∂v ∂u
ζ= − =0
∂x ∂y
Thus,

∂(uψ) ∂(vψ)
u2 + v 2 = −
∂y ∂x
Therefore, we have
∫∫ ( )
1 ∂(uψ) ∂(vψ)
KEo = ρ − dxdy
2 ∂y ∂x
By Stokes’ theorem,
∫∫ ( ) ∫
∂(uψ) ∂(vψ)
− dxdy = − uψ dx + vψ dy
∂y ∂x c

Thus, ∫
1
KEo = ρ×2 −(uψ dx + vψ dy)
2 c
The integration is taken positively (in the counterclockwise direction) round
c, the circumference of the vortex at z = b. The factor 2 is to account for the
two vortices contributing the same amount to the energy.
Now,
u dx + v dy = Vs ds
where Vs is the speed tangential to contour c and ds is arc length along c.
Therefore, ∫
Vs ds = 2πγ, the circulation
c
Also, on c, r1 = a, and r2 = 2b (approximately), so that we may express the
KEo as
(a)
KEo = −ρ × 2πγ × γ ln
2b
( )
2b
= 2πργ 2 ln
a

ISTUDY
310 CHAPTER 5. VORTEX THEORY

The fluid inside the contour c is rotating (Figure 5.48) with angular velocity
γ/a2 and moving as a whole with velocity γ/2b induced by the other vortex.
Thus the KE inside c is
( )
1 γ2 1 a2 γ 2
KEi = πa2 ρ + ×
2 4b2 2 2 a4

where the first term is the contribution due to the whole motion and the
second term is due to the angular velocity (rω). But a2 /b2 is small and hence
can be neglected. Hence,
1
KEi = πργ 2
4
Thus, the total KE is

KE = KEo + 2KEi
 ( )
1 
2b 
= 2πργ 2 
 + ln 
4 a

5.25 LINE VORTEX


Consider a continuous distribution of vortices on a straight line AB stretching
from (−c/2, 0) to (c/2, 0), as shown in Figure 5.58.

FIGURE 5.58 Continuous line of vortices.

Let the elements dξ of the line at point (ξ, 0) behave like a point rectilinear
vortex of strength γdξ, where γ may be constant or a function of ξ. This
element taken by itself will induce at the point P(x, 0) a velocity of dvx , in
the negative y-direction, as shown in Figure 5.58, given by
γdξ
dvx =
ξ−x
Thus the whole line of vortices will induce at P the velocity
∫ c/2
γdξ
vx = (5.79)
−c/2 ξ−x

Note that in Eq. (5.79) ξ is a variable and x is fixed. When ξ = x the integrand
is infinite. On the other hand, using the principle that a vortex induces no
velocity at its own centre, the point x must be omitted from the range of

ISTUDY
5.26. SUMMARY 311

variation of ξ. To do this we define the “improper” integral Eq. (5.79) by its


“principal value”, namely
∫ x−ϵ ∫ c/2 

 γdξ γdξ 
vx = lim  + 
 (5.80)
ξ→∞ −c/2 ξ − x x+ϵ ξ − x

In this way the point (x, 0) is always the centre of the omitted portion between
(x − ϵ) and (x + ϵ).
In the theory of aerofoil, the type of integral [Eq. (5.80)] in which we shall
be interested is that for which ξ = − 21 c cos ϕ and γ = γn sin nϕ, where γn is
independent of ϕ.
Now let x = − 12 c cos θ, where θ, like x, is fixed. We get from Eq. (5.79)

∫ π
sin nϕ sin ϕdϕ
vx = γ n
0 cos θ − cos ϕ

1 π
[cos (n − 1) ϕ − cos (n + 1) ϕ] dϕ
= γn
2 0 cos θ − cos ϕ
In this relation, we have integral of the type
∫ π
cos nϕ
In = dϕ
0 cos θ − cos ϕ
It can be shown that
π sin nθ
In =
sin θ
Therefore,
1
vx = γn [In+1 − In−1 ]
2
1 sin (n + 1)θ − sin (n − 1)θ
= πγn
2 sin θ
= πγn cos nθ

5.26 SUMMARY
Translation and rotation are the two types of basic motion in a fluid flow.
These two may exist independently or simultaneously. When they coexist,
they may be considered as one superimposed on the other. In an irrotational
flow the fluid elements do not rotate about their own axes, i.e., fluid elements
do not spin in an irrotational flow. But in a rotational flow, fluid elements
rotate about their axes.
(a) Pure translation — the fluid elements are free to move anywhere in
space but continue to keep their axes parallel to the reference axes fixed
in space.

ISTUDY
312 CHAPTER 5. VORTEX THEORY

(b) Pure rotation — the fluid elements rotate about their own axes which
remain fixed in space.
(c) The general motion — in which translation and rotation are compounded.

A flow in which all the fluid elements behave as in item (a) above is called
potential or irrotational flow. All other flows exhibit, to a greater or lesser
extent, the spinning property of some of the constituent fluid elements, and
are said to possess, vorticity, which is the aerodynamic term for elemental
spin. The flow is then termed rotational flow.
The angular velocity is
∂v ∂u
2ω = −
∂x ∂y
The quantity 2ω is the elemental spin, also referred to as vorticity, which is
usually denoted as ζ. Thus,

∂v ∂u
ζ= −
∂x ∂y

The units of ζ are radian per second.


In the polar coordinates, the vorticity equation can be expressed as

qt ∂qt 1 ∂qn
ζ= + −
r ∂r r ∂θ

where r and θ are the polar coordinates and qt and qn are the tangential and
normal components of velocity, respectively.
Circulation is the line integral of a vector field around a closed plane curve
in a flow field. By definition,
H
Γ= c
V · ds

where Γ is circulation, V is flow velocity tangential to the streamline c, encom-


passing the closed curve under consideration, and ds is an elemental length.
A line vortex is a string of rotating particles. In a line vortex, a chain of
fluid particles are spinning about their common axis and carrying around with
them a swirl of fluid particles which flow around in circles. The streamlines
associated with a line vortex are circular, and therefore, the particle velocity
at any point must be only tangential.
Stream function of a vortex is

Γ r
ψ=− ln
2π r0

This is the stream function for a vortex, and the circulation Γ of a flow is
positive when it is counter clockwise.

ISTUDY
5.26. SUMMARY 313

This potential function for a vortex is

Γ
ϕ= θ

We saw that a point vortex can be considered as a string of rotating


particles surrounded by fluid at large moving irrotationally. Further, the flow
investigation was confined to a plane section normal to the length or axis of
the vortex. A more general definition is that a vortex is a flow system in which
a finite area in a plane normal to the axis of a vortex contains vorticity.
A vortex consists of a bundle of elemental vortex lines or filaments. Such
a bundle is termed a vortex tube, being a tube bounded by vortex filaments.
The four fundamental theorems governing vortex motion in an inviscid
flow are called Helmholtz’s theorems.
Helmholtz’s first theorem states that, “the circulation of a vortex-tube is
constant at all cross-sections along the tube.”
The second theorem demonstrates that, “the strength of a vortex tube
(i.e., the circulation) is constant along its length.”
This is sometimes referred to as the equation of vortex continuity.
The third theorem demonstrates that, a vortex tube consists of the same
particles of fluid, i.e., “there is no fluid interchange between the vortex tube
and surrounding fluid.”
The fourth theorem states that, “the strength of a vortex remains constant
in time.”
Thompson’s vortex theorem or Kelvin’s circulation theorem states that
“in a flow of inviscid and barotropic fluid, with conservative body forces, the
circulation around a closed curve (material line) moving with the fluid remains
constant with time.”, if the motion is observed from a non-rotating frame.
The force on an aerofoil in a two-dimensional potential flow is propor-
tional to the circulation. Also, the lift, namely the force perpendicular to
the undisturbed incident flow direction, experienced by the aerofoil is directly
proportional to the circulation, Γ, around the aerofoil. The lift per unit span
of an aerofoil can be expressed as

L = ρV Γ

where ρ and V , respectively, are the density and velocity of the freestream
flow.
Stokes’ theorem relates the surface integral over an open surface to a line
integral along the ∫bounded curve. The Stokes’ integral theorem states that,
“the line∫∫integral u · dx about the closed curve c is equal to the surface
integral (▽ × u) · n ds over any surface of arbitrary shape which has c as
its boundary.”
That is, the surface integral of a vector field u is equal to the line integral
of u along the bounding curve.

ISTUDY
314 CHAPTER 5. VORTEX THEORY

I ∫∫
u · dx = (curl u) · n ds
c s

where dx is an elemental length on c and n is a unit vector normal to any


elemental area on ds.
A vortex tube is a tube made up of vortex-lines which are tangential lines
to the vorticity vector field, namely curl u (or ζ).
In an irrotational vortex (free vortex), the only vortex line in the flow field
is the axis of the vortex. In a forced vortex (solid-body rotation), all lines
perpendicular to the plane of flow are vortex lines.
The integral ∫∫
(curl u) · n ds

is called the vortex strength. It is identical to the circulation.


A very thin vortex-tube is referred to as a vortex filament.
Biot–Savart law relates the intensity of magnitude of magnetic field close
to an electric current carrying conductor to the magnitude of the current. It
is mathematically identical to the concept of relating intensity of flow in the
fluid close to a vorticity-carrying vortex tube to the strength of the vortex
tube. It is a pure kinematic law, which was originally discovered through
experiments in electrodynamics.
Velocity induced by a vortex of finite length, by Eq. (5.48), is

Γ
v= (cos α + cos β)
4πh

A vortex is termed semi-infinite vortex when one of its ends stretches to


infinity.
Therefore, the velocity induced by a semi-infinite vortex at a point P
becomes
Γ
v= (cos α + 1)
4πh
An infinite vortex is that with both ends stretching to infinity. For this
case we have α = β = 0. Thus, the induced velocity due to an infinite vortex
becomes
Γ
v=
2πh
The circulation of the vortex determines the lift, and the lift formula which
gives the relation between circulation, Γ, and lift per unit width l, in inviscid
potential flow is the Kutta–Joukowski theorem, namely

l = − ρ ΓU∞

where l is the lift per unit span of the wing, Γ is circulation around the wing,
U∞ is the freestream velocity and ρ is the density of the flow.

ISTUDY
5.26. SUMMARY 315

The circulation over the wing span varies and the lift is given by
∫ +b
L = − ρ U∞ Γ(x) dx
−b

where the origin is at the middle of the wing, x is measured along the span,
and b is the semi-span of the wing.
The starting vortex is at infinity (far downstream of the wing), and the
bound vortex and the tip vortices together form a horseshoe vortex.
The velocity w induced at the middle of the wing by the two tip vortices
accounts to double the velocity induced by a semi-infinite vortex filament at
distance b, therefore,

Γ Γ
w= (1 + 0) =
4πb 4πb
This velocity is directed downwards and, hence, termed induced downwash.
The induced drag Di , given by
w
Di = A
U∞
It is important to note that, Eq. (5.57) holds if the induced downwash from
both vortices is constant over the span of the wing. However, the downwash
does change since at a distance x from the wing centre, one vortex induces a
downwash of

Γ
4 π (b + x)
whereas the other vortex induces
Γ
4 π (b − x)

Both the downwash are in the same direction, therefore adding them, we get
the effective downwash as
Γ b
w=
2 π b − x2
2

The flow outside the vortex filament is a potential flow. Thus, by incompress-
ible Bernoulli equation, we have

1
p+ 2 ρ u2 + ρ g z = constant

This is valid both along a streamline and between any two points in the flow
field.
The second vortex theorem of Helmholtz’s states that “a vortex tube is
always made up of the same fluid particles.”

ISTUDY
316 CHAPTER 5. VORTEX THEORY

In other words, a vortex tube is essentially a material tube. This charac-


teristic of a vortex tube can be represented as a direct consequence of Kelvin’s
circulation theorem.
The induced velocity at the centre of the ring remains finite (as in horse-
shoe vortex), from Biot–Savart law, becomes

∫ 2π
Γ h2 dϕ Γ
u= 3
=
4π 0 h 2h

The third vortex theorem of Helmholtz’s states that “the circulation of a


vortex tube remains constant in time.”
The fourth theorem states that “the strength of a vortex remains constant
in time”.
The vortex motions encountered in practice may in general be classified
as free vortex or potential vortex and forced vortex or flywheel vortex. The
streamline pattern for a vortex may be represented as concentric circles,
For a planar flow (say in the horizontal plane), the changes in potential
head is zero.
Forced vortex is a rotational flow field in which the fluid rotates as a solid
body with a constant angular velocity ω, and the streamlines form a set of
concentric circles. Because the fluid in a forced vortex rotates like a rigid
body, the forced vortex is also called flywheel vortex.
Free vortex is an irrotational flow field in which the streamlines are con-
centric circles, but the variation of velocity with radius is such that there is
no change of total energy per unit weight with radius, so that dH/dr = 0.
Since the flow field is potential, the free vortex is also called potential vortex.
In the flow around a vortex core the velocity is inversely proportional to the
radius.
A free spiral vortex is essentially the combination of a free cylindrical
vortex and a radial flow.
In the free vortex, v = c/r and, thus, theoretically, the velocity becomes
infinite at the centre. The velocities near the axis would be very high and,
skin friction losses vary as the square of the velocity, they will cease to be
negligible. Also, the assumption that the total head H remains constant will
cease to be true. The portion of fluid around the axis tends to rotate as a
solid body. Thus, the central portion essentially forms a forced vortex.
A rectilinear vortex is a vortex tube whose generators are perpendicular
to the plane of motion.
A circular vortex is that with cross-section normal to its axis is a circle.
A point rectilinear vortex is the limiting case of a circular vortex of con-
stant strength γ with radius a tending to zero.
Two vortices of equal strength γ but opposite nature (one rotating clock-
wise and the other rotating counterclockwise) from a vortex pair.

ISTUDY
5.27. PROBLEMS 317

5.27 PROBLEMS
5.1 Evaluate the vorticity of the following two-dimensional flow.

(i) u = 2xy, v = x2 .
(ii) u = x2 , v = −2xy.
(iii) ur = 0, uθ = r.
(iv) ur = 0, uθ = 1r .

[Ans. (i) 0, (ii) −2(x + y), (iii) 2, (iv) 0]


5.2 If the velocity induced by a rectilinear vortex filament of length 2 m, at
a point equidistance from the extremities of the filament and 0.4 m above the
filament is 2 m/s, determine the circulation around the vortex filament.
[Ans. 5.414 m2 /s]

5.3 A point P in the plane of a horseshoe vortex is between the arms and
equidistant from all the filaments. Prove that the induced velocity at P is

Γ(1 + 2)
v=
πAB
where Γ is the intensity and AB is the length of the finite side of the horseshoe.

5.4 If the velocity induced by an infinite line vortex of intensity 100 m2 /s,
at a point above the vortex is 40 m/s, determine the height of the point above
the vortex line.
[Ans. 0.398 m]

5.5 If a wing of span 18 m has a constant circulation of 150 m2 /s around


it while flying at 400 km/h, at sea level, determine the lift generated by the
wing.
[Ans. 367.5 kN]

5.6 If the tangential velocity at a point at radial distance of 1.5 m from the
centre of a circular vortex is 35 m/s, determine (a) the intensity of the vortex
and (b) the potential function of the vortex flow.
[Ans. (a) 329.87 m2 /s, (b) 52.5 θ]

5.7 Show that a circular vortex ring of intensity Γ induces an axial velocity
Γ
2R at the centre of the ring, where R is the radius for the vortex.

5.8 If a circulation of 2.4 m2 /s is superposed around a circular cylinder of


diameter 300 mm immersed in water flowing at 3.5 m/s perpendicular to the
cylinder axis. Find (a) the location of the stagnation points and (b) the lift
on a 10 m length of cylinder.
[Ans. (a) −21.33◦ , (b) 84 kN]

ISTUDY
318 CHAPTER 5. VORTEX THEORY

5.9 A wing of rectangular planform has 10 m span and 1.2 m chord. In


straight and level flight at 240 km/h the total aerodynamic force acting on
the wing is 20 kN. If the aerodynamic efficiency of the wing is 10 calculate
the lift coefficient. Assume air density to be 1.2 kg/m3 .
[Ans. CL = 0.625]

5.10 Determine the circulation around the circle (x − 1)2 + (y − 6)2 = 4, in


a two-dimensional flow field given by u = 3x + y and v = 2x − 3y.
[Ans. −4π]

ISTUDY
Chapter 6

Flow Through Pipes

6.1 INTRODUCTION
The general features of flow through circular pipes and loss associated with
such a flow were discussed briefly in Chapter 2. Now let us look into the
experimental results relating to flow in pipes and fittings, considering fluids
of constant viscosity and density.

6.2 FLOW IN CIRCULAR PIPES


We defined the laminar flow as an orderly flow in which the transverse ex-
change of momentum in negligibly small, and described the turbulence as a
random three-dimensional phenomenon, exhibiting multiplicity of scales, pos-
sessing vorticity, and showing very high dissipation, in chapter 2. Now let
us look at the different features of laminar and turbulent flow through pipes.
We have seen that, the nature of the flow is dictated by the magnitude of the
Reynolds number. For a circular pipe, the diameter, d, is usually taken as the
characteristic dimension or the linear measurement representative of the flow
pattern. Thus the Reynolds number for flow through a circular pipe may be
expressed as
ρud
Red =
µ

where ρ is the flow density, u is the mean velocity, d is the pipe diameter and
µ is the viscosity coefficient. But the ratio of viscosity coefficient to density
µ/ρ is the kinematic viscosity, ν. Thus, in terms of ν, the Reynolds number
becomes
ud
Red =
ν
The Reynolds number is the ratio of inertia to viscous forces. Laminar flow
occurs at low velocities, and therefore at low Reynolds numbers, whereas

319

ISTUDY
320 CHAPTER 6. FLOW THROUGH PIPES

turbulent flow takes place at high Reynolds numbers. Thus, in a laminar flow
the viscous forces are predominant and in a turbulent flow the inertia forces
are predominant. Also, when the velocity is increased beyond certain limiting
value, eddies begin to form suddenly. This aspect of the sudden formation
of eddies indicates that laminar flow is unstable, and because of this even a
slight perturbation is sufficient to make the laminar flow to become turbulent.
The velocity at which the transition from laminar to turbulent nature begins
is termed critical velocity. The critical velocity is very sensitive to any initial
disturbance, such as vibration of the apparatus, insufficient stilling of the fluid
before it enters the pipe, an insufficiently smooth bell-mouth at the pipe inlet,
or sudden change in the flow cross-section (such as partial closing of valve).
Almost all these disturbances tend to reduce the critical velocity.
The Reynolds number at which the flow transits from laminar to turbulent
is termed critical Reynolds number. For circular pipes the theoretical value
of critical Reynolds number is 2300, based on the diameter. But in practical
(or actual) conditions, the transition from laminar to turbulent flow occurs at
Reynolds numbers between 2000 to 4000. But by taking extreme precautions
to avoid disturbances of any kind, laminar flow in pipes at values of Reynolds
number much higher than 4000 can be achieved. This implies that, apparently
there is no precise upper limit to the value of Reynolds number at which the
change from laminar to turbulent flow occurs. However, there is a definite
lower limit for the critical Reynolds number. When Reynolds number is less
than the lower limit any disturbance in the flow is damped out by the viscous
forces.

6.3 LAMINAR TO TURBULENT


TRANSITION
Flow transition from laminar to turbulent was demonstrated by Osborne
Reynolds in the early 1880s, with a simple visualization, popularly known
as Reynolds experiment. The apparatus used by Reynolds was as shown in
Figure 6.1.

Dye Tank

Water

Glass tube

Jet Dye streak

Water flow
FIGURE 6.1 Reynolds apparatus.

ISTUDY
6.3. LAMINAR TO TURBULENT TRANSITION 321

A straight circular glass tube with a smooth rounded, flared inlet was
placed in a large glass-walled tank filled with water. The other end of the
tube passed through the end of the tank, as illustrated in Figure 6.1. The
water flow rate through the tube was controlled by a valve at the outlet end.
A fine nozzle connected to a small reservoir of a liquid dye discharged a color
filament into the inlet of the glass tube. By observing the behavior of the dye
streak, Reynolds could able to study the way in which the water was flowing
along the glass tube.
If the velocity of the water remained low and especially if the water in the
tank has previously be allowed to settle for some time so as to eliminate all
disturbances as far as possible, the filament of dye would pass down the tube
without mixing with water, and often so steadily as illustrated in Figure 6.2(a).

(a) Steady flow

(b) Wavy flow

(c) Mixed flow

FIGURE 6.2 Illustration of steady, wavier and mixed flows.

As the flow velocity is increased progressively, a stage will reach at which


the dye streak will become wavier as shown in Figure 6.2(b). Further increase
of water velocity will make the fluctuations in the dye streaks to become more
intense, particularly towards the outlet of the tube, until a state is reached,
quite steadily, in which the dye mixed more or less completely with the water
in the tube, as shown in Figure 6.2(c). Thus, except for a region near the
inlet, the water in the tube becomes evenly colored by the dye. Still further
increase of water velocity will not cause any further alteration in the type
of flow, but the dye would mix more readily with the water, and complete
mixing will be achieved nearer to the inlet of the tube. The original type of
flow, in which the dye remained as a distinct streak, as in Figure 6.2(a), could
be restored by reducing the velocity.
It is of particular interest that the disturbed flow always begins far from
the inlet (in Reynolds tests, usually at a length from the inlet equal to about
30 times the diameter of the tube); also the complete mixing is found to occur
suddenly.

ISTUDY
322 CHAPTER 6. FLOW THROUGH PIPES

Although Reynolds used water in his experiments, subsequent experiments


have shown that the phenomenon is exhibited by all fluids, gases as well as
liquids. Moreover, the two types of flow are to be found whatever the shape
of the solid boundaries.
• In the first kind of flow, shown in Figure 6.2(a), the fluid elements are
moving in a straight line even though the velocity with which elements
move along one line is not necessarily the same as that along another line.
Since the fluid may be considered as moving in layers or laminae, this
kind of flow is termed laminar flow. In other words, laminar flow may
be defined as the flow in which the transverse exchange of momentum is
negligible.
• The second kind of flow is known as turbulent flow. As shown in Figure
6.2(c), the paths of individual elements of the fluid are no longer straight
but sinuous, intersecting and crossing one another in a disorderly man-
ner so that a thorough mixing of the fluid takes place. When turbulent
flow occurs in a cylindrical tube, for example, only the average motion of
the fluid is parallel to the axis of the tube. A single fluid element would
thus follow an erratic path involving movements in three directions.
Incorporating all the physical features of the laminar and turbulent flows,
in Chapter 2, we defined the laminar flow as an orderly flow in which the trans-
verse exchange of momentum is negligibly small and described the turbulent
flow as a random three-dimensional phenomenon, exhibiting multiplicity of
scales, possessing vorticity and showing very high dissipation. In engineering
practice, most fluid flows are nearly always turbulent. There are, however,
some important instances of mostly laminar flows, for example in lubrication.
Also, there are many instances in which part of the flow is laminar.
As we discussed in Chapter 2, the laminar and turbulent nature of the
flow depends on the magnitude of the Reynolds number. For flow in a circular
pipe, the characteristic length used in Reynolds number expression is the pipe
diameter d and the representative velocity is the mean velocity over the cross-
section. Under normal practical conditions, flow through pipe at a Reynolds
number (ρud/µ) below 2000 may be regarded as laminar, and for Reynolds
number more than 4000 may be taken as turbulent.

EXAMPLE 6.1 Air at atmospheric pressure and 22◦ C flows through a tube
of diameter 10 mm at 12 m/s. (a) Determine whether the flow is laminar or
turbulent. (b) What is the limiting maximum velocity up to which the flow
would remain laminar?

Solution Given, T = 22 + 273.15 = 295.15 K, d = 10 mm, u = 12 m/s and


p = 101325 Pa.
The viscosity coefficient is

( ) T 3/2
µ = 1.46 × 10−6
T + 111

ISTUDY
6.3. LAMINAR TO TURBULENT TRANSITION 323

( ) 295.153/2
= 1.46 × 10−6
295.15 + 111

= 1.823 × 10−5 kg/(m s)

The flow density is


p
ρ=
RT
101325
=
287 × 295.15

= 1.196 kg/m3

(a) The Reynolds number of the flow is

ρud
Re =
µ

1.196 × 12 × 0.010
=
1.823 × 10−5

= 7872.73

This value of Re implies that the flow through the tube is turbulent.
(b) For Reynolds number up to 2000 the flow will be laminar. Thus

ρud
Re = 2000 =
µ

This gives the velocity as

2000 µ
u=
ρd

2000 × (1.823 × 10−5 )


=
1.196 × 0.01

= 3.05 m/s

Thus the limiting maximum value of velocity up to which the flow will remain
laminar is 3.05 m/s.
As the flow rate is reduced, the technique using a streak of dye (as in
Reynolds experiment) is not suitable for determining the change from turbu-
lent to laminar flow. The phenomenon governing the resistance to fluid flow,
which is different for the two kinds of flow, provides an alternative approach.
For visualizing this kind of flow, the apparatus shown in Figure 6.3 is suitable.

ISTUDY
324 CHAPTER 6. FLOW THROUGH PIPES

1 2

Pressure port Pressure port

∆p Manometer

FIGURE 6.3 Water flow through a pipe.

The velocity of water through a horizontal pipe is controlled by a valve at


its downstream end, and the various values of the mean velocity, the pressure
drop over a given length of the pipe, is measured by a differential manometer.
The first pressure port is at a location, from the pipe entrance, where any
disturbance from the inlet is removed and a fully developed flow state is
reached. The results of the experiments in the apparatus shown in Figure 6.3
is illustrated in Figure 6.4.

1.7 - 2.0
1
∆p/l (on log scale)

A B

1
1
Laminar Turbulent

u (on log scale)


FIGURE 6.4 Pressure drop variation with flow velocity.

The pressure drop per unit length, ∆p/l, is plotted against the mean
velocity, u, in Figure 6.4. For laminar flow, pressure drop with velocity is
linear with a slope of one (that is, unity), as shown in Figure 6.4. This shows
that for laminar flow the pressure drop per unit length is directly proportional
to the velocity, as predicted by the theory of laminar flow.

ISTUDY
6.4. FULLY DEVELOPED LAMINAR FLOW IN A PIPE 325

As the velocity is increased, an abrupt increase in the pressure drop is


found (point B in Figure 6.4). This point corresponds to the breakdown of
laminar flow. There is a region in which it is difficult to discern any simple
law connecting ∆p/l and u, and the variation of ∆p/l with u is given by line
CD, which has a slope greater than the first line. The slope of line CD varies
from test to test. For smooth pipes the slope of CD may be as low as 1.7
and for rough pipes the slope approaches a maximum value of 2.0. When
the velocity is carefully reduced, the previous route is not exactly retraced.
The line DC is followed but, instead of the path CB being taken, the line
CA is traced out. Then, from point A, the line corresponding to laminar
flow is retraced. This kind of process which causes a turbulent flow (in a
pipe) to become laminar when the velocity is gradually decreased is termed
relaminarization. These results show that there are two critical velocities: one,
corresponding to the change from laminar to turbulent flow, is known as the
upper or higher critical velocity (point B), the other (point A) is known as the
lower critical velocity, when the flow changes from turbulent to laminar. The
critical velocities may be determined from the graph and the corresponding
critical Reynolds numbers can be calculated. Here it is essential to note that,
in Chapter 2 the lower critical Reynolds number is defined as that below
which the entire flow is laminar and the upper critical Reynolds number is
that above which the entire flow is turbulent. The lower critical value of
the Reynolds number is of greater interest because above that point laminar
flow becomes unstable. It is usual to refer to this as the lower critical value
or simply the critical Reynolds number. The experiments of Reynolds, and
later, more detailed ones of Ludwig Schiller (1882 - 1961) have shown that,
for very smooth, straight, uniform circular pipes, the lower critical value of
Reynolds number is about 2300. This value is slightly lower for pipes with the
usual degree of roughness of the walls, and for ordinary purposes it is usual
to take it as 2000.

6.4 FULLY DEVELOPED LAMINAR FLOW


IN A PIPE
Let us consider fully developed laminar flow in a circular pipe, illustrated in
Figure 6.5. For axisymmetric flows, such as this, it is convenient to work in
cylindrical coordinate system.
For fully developed steady flow, the x-component of the momentum equa-
tion is
Fx = 0

The pressure force on the left-end of the pipe is


( )
∂p dx
p− 2πrdr
∂x 2

ISTUDY
326 CHAPTER 6. FLOW THROUGH PIPES

Control volume
dr
r Control volume
R
Flow
r
x p, τrx
dr
dx

Circular pipe

FIGURE 6.5 Fully developed laminar flow in a circular pipe and control
volumes along x and r directions.

The pressure force on the right-end of the pipe is


( )
∂p dx
− p+ 2πrdr
∂x 2

where p is the pressure at the centre of the annular control volume (CV) of
radius r and thickness dr, shown in Figure 6.5.
The shear force on the inner surface of the control volume is
( ) ( )
dτrx dr dr
− τrx − 2π r − dx
dr 2 2

The shear force on the outer surface of the control volume is


( ) ( )
dτrx dr dr
τrx + 2π r + dx
dr 2 2

where τrx is the shear stress at the centre of the annular control volume.
The sum of the x-component of the force acting on the control volume
must be zero. Thus
∂p dτrx
− 2πr dr dx + τrx 2π dr dx + 2πr dr dx = 0
∂x dr
∂p
Dividing by 2πr dr dx and solving for ∂x , we get

∂p τrx dτrx
= +
∂x r dr
or
∂p 1 d(r τrx )
= (6.1)
∂x r dr
Since τrx is a function of r only, Equation (6.1) is valid for all r and x, iff each
side of the equation is constant. Thus this equation can be written as

1 d(r τrx ) ∂p
= = constant
r dr ∂x

ISTUDY
6.4. FULLY DEVELOPED LAMINAR FLOW IN A PIPE 327

or
d(r τrx ) ∂p
=r
dr ∂x
Integrating, we get
( )
r2 ∂p
rτrx = + c1
2 ∂x
or ( )
r ∂p c1
τrx = +
2 ∂x r
But
du
τrx = µ
dr
Therefore, ( )
du r ∂p c1
µ = +
dr 2 ∂x r
Solving this we get the velocity as
( )
r2 ∂p c1
u= + ln r + c2 (6.2)
4µ ∂x µ

The two integration constants c1 and c2 are to be evaluated. But we have


only one boundary condition: u = 0 at r = R. This can be handled as follows.
Although the velocity at the pipe centreline (r = 0) is not known, it is
known that the velocity must be finite at r = 0. This leads to the condition
that c1 = 0. Hence ( )
r2 ∂p
u= + c2
4µ ∂x
To evaluate c2 , the boundary condition; at r = R, u = 0, can be used. Thus
( )
R2 ∂p
0= + c2
4µ ∂x

This gives
( )
R2 ∂p
c2 = −
4µ ∂x
Hence ( ) ( )
r2 ∂p R2 ∂p
u= −
4µ ∂x 4µ ∂x
Now the velocity profile can be expressed as
( )[ ( r )2 ]
R2 ∂p
u=− 1− (6.3)
4µ ∂x R

ISTUDY
328 CHAPTER 6. FLOW THROUGH PIPES

6.4.1 Shear Stress Distribution


The shear stress is given by
du
τrx = µ
dr
Differentiating Equation (6.3) and substituting, we get the shear stress as
( )
r ∂p
τrx = (6.4)
2 ∂x

6.4.2 Volume Flow Rate


The volume flow rate through the pipe is


Q̇ = V · dA
A

∫ R
= u 2πr dr
0

∫ R ( )
1 ∂p
= (r2 − R2 ) 2πr dr
0 4µ ∂x

Integrating, we get the volume flow rate as


( )
πR4 ∂p
Q̇ = − (6.5)
8µ ∂x

The volume flow rate can also be expressed as follows.

In fully developed flow, the pressure gradient is, ∂p/∂x, is constant. Therefore,

∂p (p2 − p1 ) ∆p
= =−
∂x L L

Substituting this into Equation (6.5), we get the volume flow rate as
[ ]
πR4 ∆p
Q̇ = − −
8µ L

Replacing the radius R with the diameter (D = 2R), we have the volume flow
rate as
π∆pD4
Q̇ = (6.5a)
128 µL
where D is the pipe diameter and L its length.

ISTUDY
6.4. FULLY DEVELOPED LAMINAR FLOW IN A PIPE 329

6.4.3 Average Velocity


The average velocity, V , can be obtained by dividing the volume flow rate, Q̇,
with the cross-sectional area, A, of the pipe. That is,


V =
A

Substituting Equation (6.5), we have


( )
R2 ∂p
V =− (6.6)
8µ ∂x

6.4.4 Point of Maximum Velocity


At the point of maximum velocity, du/dr = 0. From Equation (6.3), we have
( )
du 1 ∂p
= r
dr 2µ ∂x

Thus
du
= 0 at r = 0
dr

At r = 0,
( )
R2 ∂p
u = umax =U =− = 2V (6.7)
4µ ∂x

The velocity profile, may be expressed in terms of U as

u ( r )2
=1− (6.8)
U R

EXAMPLE 6.2 Sea level air enters a tube of 300 mm diameter. (a) If
the volume flow rate is 0.033 m3 /s, determine whether the flow is laminar
or turbulent. (b) Estimate the entrance length required to establish fully
developed flow.

Solution (a) Given, d = 0.3 m, Q̇ = 0.033 m3 /s.

For sea-level air, the density and viscosity, respectively, are

ρ = 1.225 kg/m3 , µ = 1.81 × 10−5 kg/(m s)

ISTUDY
330 CHAPTER 6. FLOW THROUGH PIPES

The average velocity is


u=
A

4Q̇
=
πd2
4 × 0.033
=
π × 0.32

= 0.467 m/s

The Reynolds number is

ρud
Re =
µ

1.225 × 0.467 × 0.3


=
1.81 × 10−5

= 9481.9

Thus the flow is turbulent.

(b) For turbulent flow, by Equation (2.73b), the entrance length required for
the flow to become fully developed is

Le = 4.4 d (Re)1/6

= 4.4 × 0.3 × (9481.9)1/6

= 6.07 m

EXAMPLE 6.3 For laminar flow of water in a 12.7 mm diameter tube,


(a) determine the maximum allowable volume flow rate and (b) the maximum
average velocity if the fluid is air. What is the corresponding entrance length
for the flow of air to become fully developed. For water, take the density
and viscosity, respectively, as 103 kg/m3 and 10−3 kg/(m s), and for air the
density and viscosity, respectively, as 1.225 kg/m3 and 1.8 × 10−5 kg/(m s).

Solution (a) Given, d = 12.7 mm, ρ = 103 kg/m3 , µ = 10−3 kg/(m s).

For the flow to remain laminar, the maximum allowable Reynolds number is
2300.

ISTUDY
6.4. FULLY DEVELOPED LAMINAR FLOW IN A PIPE 331

Thus,

ρud
Re =
µ

Reµ
u=
ρd

2300 × 10−3
=
103 × 0.0127

= 0.181 m/s

Thus the maximum allowable volume flow rate is

Q̇ = Au

πd2
= u
4

π × 0.01272
= × 0.181
4

= 2.29 × 10−5 m3 /s

(b) For air, the maximum velocity up to which the flow will remain laminar
is

Re µ
u=
ρd

2300 × (1.8 × 10−5 )


=
1.225 × 0.0127

= 2.66 m

For laminar flow, by Equation (2.73a), the entrance length required for the
flow to become fully developed is

Le = 0.06 Re d

= 0.06 × 2300 × 0.0127

= 1.753 m

ISTUDY
332 CHAPTER 6. FLOW THROUGH PIPES

6.5 SHEAR STRESS DISTRIBUTION IN FULLY


DEVELOPED PIPE FLOW
For fully developed steady flow in a horizontal pipe the pressure drop is bal-
anced only by the shear forces at the pipe wall. This is true for both laminar
and turbulent flow. This can be verified by applying the momentum equation
to a cylindrical volume (CV) in the flow, shown in Figure 6.6.

CV τrx 2πr dx
r
R
r
Flow
x p  
∂p dx
πr 2
 
p− ∂p dx
πr 2 p+ ∂x 2
∂x 2
dx

FIGURE 6.6 Fully developed laminar flow in a circular pipe and control
volume (CV).

For steady, incompressible, fully developed flow in a horizontal pipe, the


x-component of the momentum equation is
Fsx = 0
where Fsx is the surface force in the x-direction. The surface forces acting on
the control volume are the pressure force at the left and right faces and the
shear force over the surface area, as shown in Figure 6.6. Thus
( ) ( )
∂p dx ∂p dx
Fsx = p − πr − p +
2
πr2 + τrx 2πr dx = 0
∂x 2 ∂x 2
or
∂p

dxπr2 + τrx 2πr dx = 0
∂x
Simplifying, we get the shear stress as
r ∂p
τrx = (6.9)
2 ∂x
This shows that the shear stress on the fluid varies linearly across the pipe,
from zero at the centreline to a maximum at the pipe wall. That is, the wall
shear stress τw at the surface of the pipe is
R ∂p
τw = − [τrx ]r=R = − (6.10)
2 ∂x
Equation (6.10) relates the wall shear stress to the axial pressure gradient.
Also, this equation is valid for both and laminar and turbulent fully developed
pipe flow.

ISTUDY
6.6. HEAD LOSS DUE TO FRICTION 333

6.6 HEAD LOSS DUE TO FRICTION


The pressure difference, or difference of piezometric head, required to induce
fluid flow at any required steady rate through a pipe is of practical impor-
tance. Therefore, large number of researchers focused attention on this aspect.
Among them was the French engineer Hendry Darcy (1803 - 1858) who studied
the flow of water, under turbulent conditions, in long, unobstructed, straight
pipes of uniform diameter. The fall of piezometric head in the direction of
flow is caused by the dissipation of energy by fluid friction. If the pipe is of
uniform cross-section and roughness, and the flow is fully developed, that is,
if it is sufficiently far from the inlet of the pipe, for flow conditions to have
become settled, the piezometric head falls uniformly. Darcy’s results suggest
the formula, termed Darcy relation, expressed as
( )
∆p fL u2
hf = = (6.11)
ρg d 2g

where hf is the head loss due to friction, corresponding (in steady flow) to
the drop ∆p of piezometric pressure over length L of the pipe of diameter d, ρ
is the fluid density, u the mean velocity, f is a constant, g is the gravitational
acceleration. The unit for hf in Equation (6.11) is meter. Indeed, to make
the unit as meter, the relation is divided by g, as seen in the relation. But
the head loss due to friction can also be expressed as

∆p f L u2
hf = = (6.11a)
ρ d 2

Now the unit for hf as become m2 /s2 .


It is useful to note that the Darcy relation, Equation (6.11), is valid for
both smooth-walled and rough-walled circular pipe, flowing full, irrespective
of whether the flow is laminar or turbulent.
The power required to compensate for this pressure loss can be expressed
as
Power = ∆p × Q̇

where Q̇ is the volume flow rate through the pipe.


The coefficient f in Equation (6.11) is known as the friction factor or
friction coefficient1 .
1 The pipe friction coefficient f , in Equation (6.11) is independent of the system of mea-

surement used. The term (u2 )/2 is the kinetic energy per unit weight of fluid corresponding
to the mean velocity and it is a length measurement, or head of the fluid motion. Note that
it is not the mean kinetic energy of the fluid ( in the
) 2pipe. For a circular pipe of diameter
4f ′ L
d the equation is expressed as hf = ∆p ρg
= d
u
2g
in European texts. Some confusion
is caused by the figure 4 because most text-books prefer to incorporate it into the coeffi-
cient f . It should always be made clear whether the general coefficient (as f in Equation
(6.11) is meant, or the particular value f applicable to circular pipes (i.e. f = 4f ′ and so
f L u2
hf = d 2
)

ISTUDY
334 CHAPTER 6. FLOW THROUGH PIPES

The friction coefficient depends on the Reynolds number of the flow and
the roughness of the pipe surface. The value of f is related to the shear stress
τw at the wall of the pipe. Over a short length δx the change in head is δhf .
For fully developed flow, the head falls linearly with x, therefore,

δhf dhf hf
= =
δx dx L

It can be shown that, the shear stress τw and dp/dx are related by the ex-
pression

A dp
τw =
P dx

where A is the cross-sectional area and P is the perimeter of the pipe. Thus

πd2 /4 dp
τw =
πd dx
d dp
=
4 dx

The drop in piezometric pressure is −δp = ρg δhf , therefore,

δhf 1 dp
=− (6.12)
δx ρg dx

Thus
d δhf
τw = − ρg
4 δx
From Equation (6.11), we have the friction factor, f , as

d 2g
f= hf
4L u2
( ) ( )
Note that, the factor fdL in Equation (6.11) is expressed as 4fdL , in terms
of hydraulic diameter, instead of the circular pipe diameter, in order to make
the relation valid for cross-sections other than circular too. But

hf dhf
=
L dx

Therefore,
d 2g dhf
f=
4 u2 dx

ISTUDY
6.7. POISEUILLE’S LAW 335

Substituting Equation (6.12), we get


( )
d 2g 1 dp
f= −
4 u2 ρg dx
( )
2 d dp
=− 2
ρu 4 dx
( )
1 d dp
=−1 2
2 ρu
4 dx

But
d dp
= τw
4 dx
Therefore,
τw
f =−1 2
(6.13)
2 ρu

The negative sign in Equation (6.13) implies that the shear stress at the pipe
wall acts opposite to the flow direction.
We saw that the friction factor, f , depends on the wall roughness of the
pipe. The irregularity of the surface usually vary greatly in shape, size and
spacing, therefore, quantitatively specifying the roughness is extremely dif-
ficult. Nevertheless, one feature that may be expected to influence the flow
appreciably is the average height of the ‘bumps’ on the surface. However, it
is not the absolute size of the bumps that is important, but their size com-
pared to some other characteristic length of the system. In other words, it is
the relative roughness that affects the flow. For a circular pipe, the relative
roughness may be suitably expressed as the ratio of the average height, ε, of
the surface irregularities to the pipe diameter, d.
Dimensional analysis may be used to show that the friction factor, f , is a
function of Reynolds number and the relative roughness, ε/d. Other features
of the roughness, such as the spacing of the bumps, may also influence the
flow. Yet the use of ε/d as a simple measure of the relative roughness had
made possible notable progress towards the solution and understanding of a
very complex problem.

6.7 POISEUILLE’S LAW


Poiseuille’s law states that, “the velocity of a liquid flowing through a capillary
is directly proportional to the pressure of the liquid and the fourth power of
the radius of the capillary and is inversely proportional to the viscosity of the
liquid and the length of the capillary”.
In the case of smooth flow (laminar flow), the volume flow rate, Q̇, is given
by the pressure difference divided by the viscous resistance. This resistance
depends linearly upon the viscosity and the length, but the fourth power

ISTUDY
336 CHAPTER 6. FLOW THROUGH PIPES

dependence upon the radius is dramatically different. Poiseuille’s law is found


to be in reasonable agreement with experiment for uniform liquids (called
Newtonian fluids) in cases where there is no appreciable turbulence.

π (pressure difference)(radius)4
Q̇ =
8 (viscosity)(length)
Poiseuille’s law can be used to calculate volume flow rate only in the case of
laminar flow.

6.8 FRICTION FACTOR VARIATION


Friction factor variation with Reynolds number and relative roughness ε/d
may be represented as shown in Figure 6.7. This plot of f variation with
Reynolds number and relative roughness is based on the results obtained
by the German engineer John Nikuradse using circular pipes that had been
artificially roughed using sand grains.
The friction factor variation with Reynolds number for laminar flow with
Reynolds number less than 2000 is linear as shown in Figure 6.7. In this range
the flow is laminar and the volume flow rate is given by Poiseuille’s equation

πR4
Q̇ = (p1 − p2 ) (6.14)
8µL

where R is the pipe radius, L is the pipe length, µ is viscosity coefficient and
p1 and p2 are the pressures at the beginning and end of the pipe segment of
length L.
This gives the pressure drop as

8µLQ̇
(p1 − p2 ) =
πR4

Substituting this into Equation (6.11), we get the head loss as

8µLQ̇ 1
hf =
πR4 ρg

f L u2
=
d 2g

That is,
8µL Q̇ f L u2
4
=
πR ρ d 2
Note that u is written as u for simplicity.

ISTUDY
6.8. FRICTION FACTOR VARIATION 337

0.1

Laminar
0.08
Rough zone
Transition zone
Friction factor f

0.06 ε/d = 0.0333

0.0163

0.04
0.00833
0.00397
0.00198
0.02 0.000985

Smooth zone
0.01
5 103 2 5 104 2 5 105 2 5 106
Reynolds number

FIGURE 6.7 Friction factor variation with Reynolds number and relative
roughness.

The volume flow rate Q̇ = πR2 u, therefore,

8µL πR2 u f L u2
4
=
πR ρ d 2

8µL u f L u2
=
R2 ρ d 2

64µL u fL 2
= u
d2 ρ d
or
64µ
flaminar =
ρud
This can be expressed as

64
flaminar = (6.15)
Re
where
ρud
Re =
µ

Equation (6.15) is represented by a straight line in Figure 6.7. Experimental


results confirm this equation and the fact that laminar flow is independent of
the roughness of the pipe wall.

ISTUDY
338 CHAPTER 6. FLOW THROUGH PIPES

For turbulent flow with Reynolds number greater than 2000 also the flow
depends on the roughness of the pipe wall, and different curves are obtained
for different values of the relative roughness (Figure 6.7). In Nikuradse’s ex-
periments, grains of sand of uniform size were glued to the wall of the pipes
of various diameters, which were initially smooth. Thus a value of the rela-
tive roughness was readily deduced since ε could be said to correspond to the
diameter of the sand grains. Values of the relative roughness in Nikuradse’s
experiments are marked against each curve in Figure 6.7. It is seen that, close
to the critical Reynolds number, all the curves coincide, but for successively
higher Reynolds numbers the curves separate in sequence from the curve for
smooth pipes, and greater the relative roughness, ε/d, the sooner the corre-
sponding curve branches off. Eventually, each curve flattens out to a straight
line parallel to the Reynolds number axis, indicating that the friction factor,
f , becomes independent of Reynolds number.
Nikuradse’s results confirm the significance of relative roughness (ε/d)
rather than absolute roughness (ε): the same value of ϵ/d was obtained with
different values of ε and d individually, and yet points for the same value of
ε/d lie on a single curve.
For moderate degree of roughness a pipe acts as a smooth pipe up to that
value of Reynolds number at which its curve of f versus Re separates from
the smooth-pipe line. The region in which the curve is coincident with the
smooth-pipe line is known as the turbulent smooth zone flow. When the curve
becomes horizontal, exhibiting that the friction coefficient f is independent
of Reynolds number, the flow is said to be in the turbulent rough zone and
the region between the two is known as the transition zone. The position
and extent of these zones depend on the relative roughness of the pipe. This
behavior may be explained by reference to the viscous sublayer (see Section 4
of Chapter 4). The random movement of fluid particles perpendicular to the
pipe axis, which occurs in turbulent flow, must die out as the wall of the pipe
is approached, so even for the most highly turbulent flow there is inevitably
a very thin layer, immediately adjacent to the wall, in which these random
motions are negligible. The higher the Reynolds number, the more intense
are the secondary motions that constitute the turbulence, and closer they
approach to the boundary. This causes the very small thickness of the viscous
sublayer to become smaller even when the Reynolds number increases.
In the turbulent smooth zone of flow, the viscous sublayer is thick enough
to completely cover the irregularities of the surface. Consequently the size
of the irregularities has no effect on the main flow (just as when the flow
is laminar) and all the curves for the smooth zone coincide. However, with
increasing Reynolds number the thickness of the viscous sublayer decreases
and so the surface bumps can protrude through it. The rougher the pipe, the
lower the value of Reynolds number at which this occurs. In the turbulent
rough zone of the flow, the thickness of the viscous sublayer is negligible
compared to the height of the surface irregularities. The turbulent flow around
each bump then generates a wake of eddies giving rise to a resistance force
known as form drag. Energy is dissipated by the continuous production of

ISTUDY
6.8. FRICTION FACTOR VARIATION 339

these eddies. The form drag is proportional to the square of the mean velocity
of the flow. In the complete turbulence of the rough zone of flow surface,
viscous effects are negligible and so the head loss is proportional to the square
of the mean velocity (hf ∝ u2 ) and the friction factor, f , is a constant. In
the transition zone the surface bumps partly protrude through the viscous
sublayer. Thus both form drag and viscous effects are present to some extent.
In the forgoing simplified explanation, it was assumed that the demarca-
tion between the viscous sublayer and the rest of the flow is precise. More-
over, a bump on the surface can affect the flow to some degree before the
peak emerges from the sublayer. Nevertheless, this idealized picture of the
way in which the surface irregularities influence the flow provides a useful
quantitative explanation of the phenomena. It also permits the definition of
a fluid-dynamically smooth surface as one on which the protuberances are so
far submerged in the viscous sublayer as to have no effect on the flow. Thus a
surface that is smooth at low Reynolds number may be rough at high Reynolds
numbers.
Nikuradse’s results were obtained for uniform roughness and not for that
of pipes encountered in practice, which have a roughness elements of varying
height and distribution. Even though the average height of the irregularities
on commercial pipe surfaces may be determined, Nikuradse’s diagram (Figure
6.7) is not suitable for actual pipes. Protrusion on the surface of commercial
pipes are of various heights and because of this they begin to protrude through
the viscous sublayer at different values of Reynolds number, and the transi-
tion zones of Nikuradse’s curves do not correspond at all well to those for
commercial pipes. However, at a high enough Reynolds number the friction
factors of many industrial pipes become independent of Reynolds number,
and under these conditions a comparison of the value of f with Nikuradse’s
results enables an equivalent uniform size of sand grain ε to be specified for
the pipe. Using these equivalent grain sizes, the American engineer Lewis F.
Moody (1880 - 1953) prepared a modified diagram, shown in Figure 6.8, for
use with ordinary commercial pipe.
Even though Moody’s diagram is widely used for predicting the value of the
friction factor f applicable for commercial pipes, the concept of an equivalent
grain size is open to serious objection. This concept implies that only the
height of surface irregularities significantly affects the flow. But in the rough
zone of flow, the space of the irregularities is also of great importance. If the
irregularities are far apart, the wake of eddies formed by one bump may die
away before the fluid encounters the next bump. When the bumps are closer,
however, the wake from one bump may interfere with the flow around the
next. If the bumps are exceptionally close together the flow may largely skim
over the peaks while the eddies may be trapped in the valleys. In Nikuradse’s
experiments the sand grains were closely packed, and so the spacings may
be assumed to be approximately equal to the grain diameter. His results
could therefore just be validly be taken to demonstrate that f depends on
s/d, where s is the average spacing of the grains. The equivalent grain size
do not account for the shape of the irregularities. Another factor that may

ISTUDY
340 CHAPTER 6. FLOW THROUGH PIPES

ISTUDY
6.8. FRICTION FACTOR VARIATION 341

appreciably affect the value of f in large pipes is the waviness of the surface,
that is, the presence of transverse ridges on a large scale than the normal
roughness.
Large number of empirical relations have been proposed to describe certain
parts of Figure 6.8. For example, Blasius expressed the friction factor f for
the turbulent smooth-pipe curve as

− 14
f = 0.316 (Re) (6.16)

This agrees closely with experiments results for Reynolds number between
4000 to 105 . Many relations have been proposed so that f can be calculated
directly for the entire range of ε/d and Re. The one among them which is
considered to be the best is the following proposed by S.E. Haaland
[ ]
1 6.9 ( ε )1.11
√ = −1.8 log10 + (6.16a)
f Re 3.71d

This relation predicts f which is accurate within ±1.5%.

EXAMPLE 6.4 A liquid, of viscosity 1.74 × 10−3 kg/(m s), flows through
a capillary tube of diameter 0.5 mm, at the rate of 880 mm3 /s. If the pressure
drop over the tube length is 1 MPa, determine the length of the tube, assuming
the flow as laminar.

Solution Given, D = 0.5 mm, µ = 1.74 × 10−3 kg/(m s), Q̇ = 880 mm3 /s,
∆p = (p1 − p2 ) = 1 MPa.

By Equation (6.5a),
π∆pD4
Q̇ =
128 µL
Thus

πD4
L= × (p1 − p2 )
128 Q̇ µ

π (0.5 × 10−3 )4 × 106


= ×
128 (880 × 10−9 ) × (1.74 × 10−3 )

= 1.00 m

EXAMPLE 6.5 Water with viscosity 1.14 × 10−3 kg/(m s) flows through
a 40 mm diameter tube of length 750 m. If the flow rate is 68 × 10−6 m3 /s,
determine the head loss due to friction and the power required.

Solution Given d = 0.04 m, L = 750 m, µ = 1.14 × 10−3 kg/(m s) and


Q̇ = 68 × 10−6 m3 /s.

ISTUDY
342 CHAPTER 6. FLOW THROUGH PIPES

Therefore,


u=
A

4Q̇
=
πd2

4 × (68 × 10−6 )
=
π × 0.042

= 54.1 × 10−3 m/s

The Reynolds number is

ρud
Re =
µ

103 × (54.1 × 10−3 ) × 0.04


=
1.14 × 10−3

= 1898

Thus the flow is laminar. For laminar flow,

64
f=
Re
64
=
1898

= 0.0337

By Equation (6.11),

L u2
hf = f
d 2g

750 (54.1 × 10−3 )2


= 0.0337 × ×
0.04 2 × 9.81

= 32.2 × (54.1 × 10−3 )2

= 0.0942 m

ISTUDY
6.9. VELOCITY DISTRIBUTION IN TURBULENT FLOW 343

The power required is

Power = ∆p × Q̇

= hf ρg × Q̇

= 0.0942 × 103 × 9.81 × (68 × 10−6 )

= 0.0628 W

Aliter: The power can also be calculated as follows.

By Equation (6.5a),

128µLQ̇
∆p =
πd4

128 × (1.14 × 10−3 ) × 750 × (68 × 10−6 )


=
π × 0.044

= 925 Pa

Therefore, the power is

Power = 925 × (68 × 10−6 )

= 0.0629 W

6.9 VELOCITY DISTRIBUTION IN


TURBULENT FLOW
Ludwieg Prandtl was the first to deduce an acceptable expression for the
variation of velocity in turbulent flow past a flat plate and in a circular pipe.
He used his mixing-length concept together with some intuitive assumptions
about its variation with distance from the boundary. Using somewhat different
assumptions, von Karman and others obtained the same basic result. These
expressions are all semi-empirical in that the values of the constant terms have
to be determined by experiment, but the forms of the expressions are derived
theoretically.
The same results may also be obtained from dimensional analysis, without
the hypothesis of mixing length and only with less far-reaching assumptions.
Let us use this more general method based on dimensional analysis. For
fully developed turbulent flow in a circular pipe we require primarily to know
the way in which the time-averaged value of the velocity varies with position
(along the radius) in the cross-section. If the flow is steady - with the time-
average value at a given point does not change with time - then considerations

ISTUDY
344 CHAPTER 6. FLOW THROUGH PIPES

of symmetry indicate that this velocity u is the same at all points at the same
distance from the pipe axis. The independent variables that affect the value
of u are the density of the flow, ρ, and the dynamic viscosity, µ, of the fluid,
radius, R, of the pipe, the position of the point (distance y from the pipe wall,
or the radial distance r = (R−y) from the axis), the roughness of the pipe wall
surface ε, and the shear stress τw at the wall. The pressure drop divided by
the length could be used as a variable in place of τw , as it is related to τw - here
taken as positive - is more convenient for our discussions here. Application of
the principle of dimensional analysis suggests the following relation, for the
velocity distribution
[ ( )1/2 ]
u R τw y ε
1/2
= ϕ1 , , (6.17)
(τw /ρ) ν ρ R R

1/2
where ν = µ/ρ is the kinematic viscosity. The term (τw /ρ) has the same
dimensions as velocity, therefore it is convenient to make the substitution
( )1/2
τw
uτ =
ρ

where uτ is known as the friction velocity. Using the friction velocity, the
velocity distribution equations can be expressed in simpler and more compact
forms. We know that the effect of viscosity in the flow is appreciable only near
the wall, where the velocity gradient is large (that is, within the boundary
layer), and that its effect near the centreline of the pipe will be only marginal.
1/2
Similarly, although the wall roughness affects the value of (τ0 /ρ) for a given
rate of flow, it has little effect on the flow near the pipe centreline. Therefore,
it may be inferred that for the flow around the pipe axis the flow velocity
defect; that is, the difference between the maximum velocity um at the pipe
centre and the local velocity u depends on y/R only, that is,

um − u
1/2
= ϕ2 (η) (6.18)
(τw /ρ)

where η = y/R. Equation (6.18) is known as the velocity defect law.


This hypothesis was first corroborated experimentally by Sir Thomas E.
Stanton in 1914, and later Nikuradse and others obtained curves of u(τ0 /ρ)−1/2
against η, for a wide range of wall roughness. When the points of maximum
velocity were superposed the curves coincide for most of the pipe cross-section.
Only for a narrow region close to the wall the curves differ. Equation (6.18)
may therefore be regarded as the universal relation valid for η = ∆1 , say, to
the centre of the pipe, where η = 1.
Note that, for smaller the value of friction factor f (that is, for small τ0 ),
(um − u) is small. This leads to nearly a uniform velocity over the section.

ISTUDY
6.10. VELOCITY DISTRIBUTION IN SMOOTH PIPES 345

6.10 VELOCITY DISTRIBUTION IN SMOOTH


PIPES
A smooth pipe is that for which the projection on the wall do not affect the
flow. For a smooth pipe, Equation (6.18) becomes
u
= ϕ1 (ξ, η) (6.19)

√ √
where uτ = τ0 /ρ, ξ = (R/ν) τ0 /ρ = Ruτ /ν. Equation (6.19) is valid for
all values of η.
For η = 1,
um
= ϕ1 (ξ, 1) = ϕ3 (ξ) (6.20)

Combining Equations (6.18) and (6.19), we get
um
= ϕ3 (ξ) = ϕ1 (ξ, η) + ϕ2 (η) (6.21)

This is valid for ∆1 < η ≤ 1.
Let us now consider a very thin layer of the flow close to the wall. Since
the radius of the pipe is large compared to the thickness of this layer the flow
in it will be randomly affected by the fact that the wall of a circular pipe
is curved. In other words, the radius of the pipe will be of negligible effect
on the velocity near the wall. This assumption that the velocity near the
wall depends on the distance from the wall but not on the pipe radius was
put forward by Prandtl and is well supported by experimental evidence. For
positions close to the wall, that is, for 0 < η < ∆2 , say, the function, ϕ1 (ξ, η)
in Equation (6.19) becomes ϕ4 (ξη), since ξ η = y uτ /ν, which is independent
of R. Therefore
u
= ϕ4 (ξη) (6.22)

Equation (6.22) is termed the law of the wall or the inner velocity law and is
valid for 0 < η < ∆2 .
Experimental results suggest that ∆2 > ∆1 , that is, there is a region where
Equations (6.18) and (6.19) are equally valid. This region is known as the
overlap region. For the region where ∆1 < η < ∆2 and both equations apply,
they may be added to give
um
ϕ2 (η) + ϕ4 (ξη) =

But, by Equation (6.19), we have
um
= ϕ3 (ξ)

Therefore,
um
ϕ2 (η) + ϕ4 (ξη) = = ϕ3 (ξ) (6.23)

ISTUDY
346 CHAPTER 6. FLOW THROUGH PIPES

Differentiating this with respect to ξ, we get

η ϕ′4 (ξη) = ϕ′3 (ξ) (6.24)

where ϕ′ (x) means (∂ϕ(x)/∂x).


Differentiation of Equation (6.24) with respect to η gives

ϕ′4 (ξη) + ξ η ϕ′′4 (ξη) = 0 (6.25)

Note that, Equation (6.25) involves only the combined variable ξη. Integration
with respect to ξη yields

ξ η ϕ′4 (ξη) = constant = A

Hence,
A
ϕ′4 (ξη) =
(ξη)
Integration of this gives

ϕ4 (ξη) = A ln (ξη) + B

where B is a constant. Therefore, Equation (6.22) becomes


u
= A ln (ξη) + B

But ξ = Rur /ν and η = y/R, therefore,
( )
u Ruτ y
= A ln +B
uτ ν R
or ( yu )
u τ
= A ln +B (6.26)
uτ ν
Similar arguments may be used in considering turbulent flow between two
parallel smooth plates, separated by a distance 2h. The expressions obtained
differ from those above only in having h in place of R, the final result [Equa-
tion (6.26)] is independent of either h or R. Equation (6.26) applies also to
turbulent flow over a single smooth flat plate (that is, h = ∞).
Equation (6.26) is popularly known as the logarithmic law. If the local
velocity u is measured(at√various
) distance y from a flat plate, for example, a
y τ0 /ρ
graph of u against ln ν is a straight line for a considerable range of
values. However, from the derivation of the relation, it can be inferred that
it is applicable only over a certain range of values of y. Also, it should be
noted that in practice the velocity does not increase indefinitely with y, as the
equation suggests, but tends to the velocity of the main stream. At very small
values of y it is difficult to obtain reliable experimental values of u. Further,
the equation clearly fails when y → 0, since it predicts an infinite negative

ISTUDY
6.10. VELOCITY DISTRIBUTION IN SMOOTH PIPES 347

velocity at the boundary layer. This failure is natural in view of the existence
of the viscous sublayer: as the turbulence is suppressed immediately next to
the boundary the equation can apply only outside the sublayer.
For a circular pipe,
( √ when)experimental results are plotted as a graph of
√ y τ0 /ρ
u/ τ0 /ρ against ln ν the graph, as shown in Figure 6.9, a straight
( √ )
y τ0 /ρ
line is obtained over a wide range of ln ν .

30
Equation (6.26)
25
Combined viscous
turbulent effects
20 Viscous
sublayer
u (τ0 /ρ)−1/2

15

10 Turbulent core
u = τ0 y/µ Experimental results
5 follow the dashed curve

0
1 2 3 4 5 6 7 8 9 10
"   #
1/2
y τ0
y

τ0
1/2
y

τ0
1/2 ln
=8 = 30 ν ρ
ν ρ ν ρ ( √ )
√ y τ0 /ρ
FIGURE 6.9 Variation of u/ τ0 /ρ with ln ν .

The logarithmic law cannot be expected to be valid as far as the centre


of the pipe, because of the hypothesis that the velocity is independent of
the pipe radius breaks down there. Moreover, symmetry requires ∂u/∂y
[= R−1 (τ0 /ρ) ∂ϕ1 /∂η] to be zero at the axis and this condition is not
1/2

met.
From Figure 6.9, it is seen that for large Reynolds numbers, and therefore
for large values of ξ, Equation (6.26) is valid everywhere, except close to
the pipe axis and close to the wall. This equation is not valid for very small
values of η because of the presence of the viscous sublayer. Within the viscous
sublayer, y/R = η is very small and due to that τ differs negligibly from τ0 .
That is, the shear stress in the viscous sublayer may be considered constant.
Integrating the laminar stress equation,
( )
∂u
τ =µ
∂y
and setting the integration constant to zero, so that u = 0 when y = 0, we
have
τ0 y
u=
µ

ISTUDY
348 CHAPTER 6. FLOW THROUGH PIPES

Rearranging, this becomes


τ0 ρ
u= y
ρ µ
( )
y τ0
=
µ/ρ ρ
y 2
= u
ν τ


since uτ = τ0 /ρ.
or
u yuτ
= (6.26a)
uτ ν
This relation is plotted at the left-hand side of Figure 6.9. The viscous sub-
layer, however, has no definite edge where viscous effects end and appreciable
turbulence begins, and so the experimental points make a gradual transition
between the two curves of Figure 6.9.
Nikuradse made many detailed measurements of the velocity distribution
in turbulent flow for a wide range of Reynolds number, and his results suggest
that the constants A and B in Equation (6.26) should be A = 2.5 and B = 5.5.
Substituting these into Equation (6.26), we have

u ( yu )
τ
= 2.5 ln + 5.5 (6.26b)
uτ ν

or
u ( yu )
τ
= 5.75 log10 + 5.5
uτ ν
This equation clearly represents the velocity distribution in smooth pipe at
fairly high Reynolds number. The distribution follows the viscous line (Figure
6.9) up to
yuτ
≈8
ν
or

y=


=√
τw /ρ

But
τw
f= 1 2
2 ρu

ISTUDY
6.10. VELOCITY DISTRIBUTION IN SMOOTH PIPES 349

Thus
τw f
= u2
ρ 2
√ √
τw f
= u
ρ 2
Therefore,
y 8ν
= √
d f
ud
2
or

δl 8ν 2
= √
d ud f

8 2µ
= √
d u fρ

8 2
= √
(ρu d)/µ f

i.e.

δl 8 2
=√ (6.26c)
d f Re
where d is the pipe diameter, δl is the thickness of the viscous sublayer, u =
Q̇/(πd2 /4) and Re = ud/ν.

6.10.1 Friction Factor for Smooth Pipes


The parameter
1
uτ = (τ0 /ρ) 2 (6.26d)
is related to the friction factor f , by Equation (6.13), as
τ0
f= 1 2
2 ρu

1 τ0
f= 2
2 ρu

(τ0 /ρ)
=
u2

u2τ
=
u2

ISTUDY
350 CHAPTER 6. FLOW THROUGH PIPES

Therefore,

u 2
= (6.26e)
uτ f

We therefore use the relation for velocity distribution, given by Equation


(6.26), to obtain an expression for the mean velocity u. It is seen that Equation
(6.26) is not valid over the entire cross-section, but experimental results such
as those represented in Figure 6.9 show that ∆1 , the value of η at which the
equation ceases to apply, is small, especially at high Reynolds numbers. As
an approximation, we therefore, suppose that Equation (6.26) is valid for all
values of η. Multiplying the equation by 2πrdr, integrating between the limits
r = 0 and r = R, and then dividing the result by πR2 we obtain

u u
=
τ0 /ρ1/2 uτ
∫ R [ ]
1
= A ln (ξη) + B 2πrdr
πR2 0

∫ 0 [ ]
= A ln (ξη) + B 2(1 − η)(−dη)
1

or
u 3
= A ln η − A + B (6.27)
τ0 /ρ1/2 2
This integration neglects the fact that Equation (6.27) does not correctly
describe the velocity profile close to the axis, but as this inaccuracy is appre-
ciable only as η → 1 and (1 − η) is then small, the effect on the value of the
1/2
integral is not significant. Then substituting for u (τ0 /ρ) from Equation
(6.13), we have
( )1/2
2 3
= A ln ξ − A + B
f 2
Noting that
( )1/2 ( )1/2 ( )1/2
R τ0 d f f
ξ= = u = Re
ν ρ 2ν 2 8

we have [ ]
1 1 ( √ ) √ 3
√ = √ A ln Re f − A ln 8 − A + B
f 2 2
Substituting Nikuradse’s values of A and B and converting ln to log10 this
gives
1 ( √ )
√ = 4.07 log10 Re f − 0.6
f

ISTUDY
6.10. VELOCITY DISTRIBUTION IN SMOOTH PIPES 351

In view of the approximation made it would be surprising if this equation


exactly represents experimental results, especially as the last term is a differ-
ence between relatively large quantities. Good agreement with experiment is
obtained if the coefficients are adjusted to give

1 ( √ )
√ = 2 log10 Re f − 0.8 (6.28)
f

This expression has been validated by Nikuradse for Reynolds numbers from
5000 to 3 × 106 . For Reynolds numbers up to 105 , however, Blasius equation
[Equation (6.16)] provides results of equal accuracy and is easier to use since
Equation (6.28) has to be solved for f by trial and error.

EXAMPLE 6.6 Air at 20◦ C flows through a 140 mm diameter tube under
fully developed condition. If the centreline velocity is 5 m/s, estimate (a) the
frictional velocity, uτ , (b) the wall shear stress and (c) the average velocity.

Solution Given, T = 20 + 273 = 293 K, d = 0.14 m.

Therefore, the viscosity, µ and density ρ are


( )
−6 2933/2
µ = (1.46 × 10 ) ×
293 + 111

= 18.12 × 10−6 kg/(m s)

p
ρ=
RT
101325
=
287 × 293

= 1.205 kg/m3

(a) By Equation (6.26b),

u ( yu )
τ
= 2.5 ln + 5.5
uτ ν
At y = R = 0.07, u = 5 m/s, therefore,
( )
5 0.07 uτ
= 2.5 ln + 5.5
uτ (18.12 × 10−6 )/1.205

5
= 2.5 ln (4655 uτ ) + 5.5

ISTUDY
352 CHAPTER 6. FLOW THROUGH PIPES

Let us solve this by trial and error.

Trial 1: Let uτ = 0.23 m/s. Therefore,

5
= 2.5 ln (4655 × 0.23) + 5.5
0.23

21.7 = 22.9

LHS < RHS.

Trial 2: Let uτ = 0.22 m/s. Therefore,

5
= 2.5 ln (4655 × 0.22) + 5.5
0.22

22.7 = 22.8

LHS ≈ RHS, thus,


uτ = 0.22 m/s

(b) The wall shear stress in terms of uτ , by Equation (6.26d) is

ρu2τ = τ0 = τw

Therefore,

τw = 1.205 × 0.222

= 0.0583 Pa

(c) The average velocity can be found by integrating the logarithmic law of
velocity distribution given by Equation (6.26),
u ( yu )
τ
= A ln +B
uτ ν

where y = (R − r). The average velocity is given by


u=
A
∫ R
1
= u[2πrdr]
πR2 0

From Equation (6.26),


( yu )
τ
u = A ln +B
ν

ISTUDY
6.10. VELOCITY DISTRIBUTION IN SMOOTH PIPES 353

Therefore,
∫ R [ ( yu ) ]
1 τ
u= uτ A ln + B 2πrdr
πR2 0 ν
∫ [ ( ) ]
1 R
(R − r) uτ
= uτ A ln + B 2πrdr
πR2 0 ν
[ ( ) ]
1 Ruτ
= uτ 2A ln + 2B − 3A
2 ν
Inserting A = 2.5 and B = 5.5, we have
[ ( )]
u 2 × 2.5 Ruτ 3
= 2A ln + 5.5 − × 2.5
uτ 2 ν 2
( )
Ruτ
= 2.5 ln + 1.75
ν
( )
0.07 × 0.22
= 2.5 ln + 1.75
(18.12 × 10−6 )/1.205

= 2.5 × 6.93 + 1.75

= 19.075

u = 19.075 × 0.22

= 4.196 m/s

Note: We used the turbulent velocity profile given by Equation (6.26), assum-
ing that the flow is turbulent. Now let us check whether our assumption is
right.

The Reynolds number is


ρud
Re =
µ

1.205 × 4.196 × 0.14


=
18.12 × 10−6

= 39, 065

The Reynolds number is well above the critical value, thus our assumption is
right.
EXAMPLE 6.7 Water flows through a smooth pipe of diameter 100 mm,
at a rate of 0.08 m3 /s. If the kinematic viscosity and density of water are

ISTUDY
354 CHAPTER 6. FLOW THROUGH PIPES

10−6 m2 /s and 103 kg/m3 , respectively, determine the average velocity, Reynolds
number, friction coefficient, friction velocity and the thickness of the viscous
sublayer.

Solution Given, d = 0.1 m, Q̇ = 0.08 m3 /s, ν = 10−6 m2 /s and ρ = 103


kg/m3 .

The average velocity is


u=
A

4Q̇
=
πd2
4 × 0.08
=
π × 0.12

= 10.19 m/s

The Reynolds number is

ud
Re =
ν
10.19 × 0.1
=
10−6

= 1.019 × 106

Thus the flow is turbulent. For turbulent flow, by Equation (6.28), we have
the relation for friction coefficient in terms of Reynolds number as

1 ( √ )
√ = 2 log10 Re f − 0.8
f

Therefore,
1 [ √ ]
√ = 2 log10 (1.019 × 106 ) f − 0.8
f
Solving this we get the friction coefficient as

f = 0.0118

The shear velocity in terms of u and f is



u 2
=
uτ f

ISTUDY
6.10. VELOCITY DISTRIBUTION IN SMOOTH PIPES 355

Therefore,

f
uτ = u
2

0.0118
= 10.19 ×
2

= 0.783 m/s

The relation for the thickness of viscous sublayer is



δl 8 2
=√
d f Re

Therefore,

8 2
δl = √ d
f Re

8 2
=√ × 0.1
0.0118 × (1.019 × 106 )

= 0.01022 mm

6.10.2 Velocity Distribution and Friction Factor for Rough


Pipes
We saw that a pipe roughness projections on the wall are small enough to
be submerged within the viscous sublayer and have no influence on the flow
outside the layer is termed fluid-dynamically smooth.[ Also, the sublayer
] thick-
1/2
ness is determined by the value of the parameter y (τ0 /ρ) /ν , therefore
the maximum height ε of [roughness elements
] which will not affect the flow is
1/2
governed by the value of ε (τ0 /ρ) /ν . Nikuradse’s results for pipe artifi-
cially roughened with uniform grains of sand indicate that the roughness has
no effect on the friction factor when
[ ]
1/2
ε (τ0 /ρ) /ν < 4 (approximately)

On the other hand, if


1/2
ε (τ0 /ρ) > 70 (approximately)

the friction coefficient, f , becomes independent of Reynolds number, and this


suggests that the effect of roughness projections overshadows viscous effect.

ISTUDY
356 CHAPTER 6. FLOW THROUGH PIPES

For this rough zone of flow the sublayer is completely disrupted and viscous
effects are negligible. Equation (6.17) may then be written as
(y ε) ( )
u u R
1/2
= =ϕ , = ϕ5 η,
(τ0 /ρ) uτ R R ε

Here ε may be regarded as some magnitude suitably characterizing the size


and it need not necessarily be the average height of the protrusions.
By similar reasoning to that used for smooth pipes (except that R/ε here
takes the place of ξ) the equation for the velocity distribution may be shown
to take the form
u (y )
= A ln +c (6.29)
uτ ε
A similar expression may be deduced for flow over rough plates; η then repre-
sents y/δ. The constant A in Equation (6.29) is the same as that for smooth
pipes. The constant c, however, differs from B. Equation (6.29) fails to give
∂u/∂y = 0 on the pipe axis, but over most of the cross-section, experimental
results are well described by the equation
u (y)
= 5.75 log10 + 8.48 (6.30)
uτ ε
It is seen that this equation is independent of the relative roughness ε/R;
experimental points for a wide range of ϵ/R all conform to straight line when
u/uτ is plotted against log10 (y/ε).
Reasoning similar to that used in the previous section on friction factor
for smooth pipes leads to a relation for the friction factor as
( )
1 R
√ = 2.035 log10 + 1.675 (6.31a)
f ε
Slight adjustment of the coefficients to achieve better agreement with experi-
ment gives
( )
1 d
√ = 2 log10 + 1.14 (6.31b)
f ε
where d is the diameter of the pipe.
All these relations apply only to fully developed turbulent flows and to
Reynolds numbers sufficiently high for f to be independent of Reynolds num-
bers, that is, these relations are not applicable to entrance length of a pipe.
1/2
Equation (6.31) is valid for values of ε (τ0 /ρ) /ν > 70, whereas the
1/2
approximate equation for smooth pipes [Eq. (6.28)] is valid for ε (τ0 /ρ) /ν <
1/2
4. For 4 < ε (τ0 /ρ) /ν < 70, friction factor f is dependent on both Reynolds
number and the relative roughness. If 2 log10 (d/ε) is subtracted from both
sides of Equations (6.28) and (6.31b), we obtain
( ) ( )
1 d 1
√ − 2 log10 = 2 log10 Re f 2 ε/d − 0.8
f ε

ISTUDY
6.10. VELOCITY DISTRIBUTION IN SMOOTH PIPES 357

for smooth pipes and


( )
1 d
√ − 2 log10 = 1.14
f ε
for rough pipes. ( )
( ) 1
When √1f − 2 log10 dε is plotted against log10 Re f 2 ε/d , as shown in
Figure 6.10, two straight lines result - one for smooth pipes and one for rough
pipes.
4
Uniform roughness (Nikuradse)
8)]
 

Rough [Equation (6.31b)]


6.2
d
ǫ

n(

2
√ − 2 log10

tio
qua

Equation (6.32)
[E
th
f
1

oo

0
Sm

−2
0 1 2 3
log10 Re f 1/2 ǫ/d

[ ]
d [ ( √ )]
FIGURE 6.10 Variation of √1
f
− 2 log10 ( ) with log10 Re f ε/d
ε

For commercial pipes, in which the roughness in not uniform, an equiva-


lent roughness size ε may be deduced from the (constant) value of f at high
Reynolds numbers. C. F. Colebrook measured values of f for new commercial
pipes and found that when the results were plotted in the form of Figure 6.10,
the points were closely clustered about a curve joining the two straight lines
having the equation
[ ]
1 ε 2.51
√ = −2 log10 + √ (6.32)
f 3.71 d Re f

For ε → 0 this expression approaches the smooth pipe equation [Equation


(6.28)], and for Re → ∞ it approaches rough pipe equation [Equation (6.31)].
The fact that the points for the commercial pipes all quite closely fit a
single curve suggests that their random roughness may be described by a
single parameter ε. The curve, however, significantly differ from that for
uniform sand roughness.
It is seen that the friction factor f appears on both sides of Colebrook’s
equation [Equation (6.32)]. To overcome this situation of dealing with this
kind of relation, L. F. Moody constructed his chart (Figure 6.8). Although

ISTUDY
358 CHAPTER 6. FLOW THROUGH PIPES

Equation (6.32) and Moody’s chart permit determination of f when ε and


Re are known, there remains the problem of specifying the sand roughness
size ε for a particular pipe. In principle this can be done by conducting
an experiment on the pipe at a sufficiently high Reynolds number that the
constant value of f may be determined and ε then be deduced from Equation
(6.31b). √
For very rough pipes such that ε/d ≫ 1/(Re f ), the friction depends only
upon ε/d and not upon Reynolds number. This occurs when the laminar sub-
layer is completely disrupted by the surface roughness, and the fluid viscosity
plays no role in the pressure drop.
The experimental values of wall roughness ε for some common pipes de-
termined from flow tests are listed in Table 6.1.

TABLE 6.1 Wall roughness ε of some pipe materials


Material ε, m
Concrete 0.3 × 10−3 – 3.0 × 10−3
Cast iron 3.0 × 10−4
Galvanized iron 1.5 × 10−4
Commercial steel 5 × 10−5
Drawn Tubing 1.5 × 10−5
Glass, Plastic 0

EXAMPLE 6.8 A liquid of density 1184 kg/m3 and viscosity 1.29 × 10−5
kg/(m s), flows at a rate of 4.3 m3 /s through a pipe of diameter 1 m. The
pressure taps at locations 1 and 2, 10 m apart, are connected to the tubes of
a U-tube mercury manometer reads 42 mm. Determine the average relative
roughness of the pipe surface.

Solution Given, ρ = 1184 kg/m3 , µ = 1.29 × 10−5 kg/(m s), Q̇ = 4.3 m3 /s,
d = 1 m, L = 10 m, ∆h = 42 mm.

The pressure drop over 10 m length of the pipe is

∆p = ρHg g∆h

= 13600 × 9.81 × (42 × 10−3 )

= 5603.47 Pa

The average velocity is


u=
A

4Q̇
=
πd2

ISTUDY
6.10. VELOCITY DISTRIBUTION IN SMOOTH PIPES 359

4 × 4.3
=
π × 12

= 5.47 m/s

The head loss due to friction, by Equation (6.11), is


∆p
hf =
ρg

5603.47
=
1184 × 9.81

= 0.482 m

The head loss is also given by


L u2
hf = f
d 2g
Therefore,
2hf dg
f=
Lu2
2 × 0.482 × 1 × 9.82
=
10 × 5.472

= 0.0316

The Reynolds number is


ρud
Re =
µ

1184 × 5.47 × 1
=
1.29 × 10−5

= 502.05 × 106

For Re = 502.05 × 106 and f = 0.0316, from Moody’s chart, Figure 6.8, we
have ε/d = 0.004 .

Note that the above values of Re, ε and d agrees well with the plot of Equation
(6.31), in Figure 6.10.

6.10.3 Universal Features of the Velocity Distribution in


Turbulent Flow
In the velocity expression in Equation (6.29) the factor A is considered to
be a universal constant. Although A = 2.5 is the universally accepted value,

ISTUDY
360 CHAPTER 6. FLOW THROUGH PIPES

different researchers have obtained somewhat different figures, and the value
could well depend on the Reynolds number. For flow over flat plate, the
values A = 2.4 and B = 5.84 in Equation (6.26) have been found to give
better agreement with experiment.
The velocity defect law is independent of both Reynolds number and
roughness. Thus, if as in the integration leading to Equation (6.27), this
law is assumed to apply with sufficient accuracy over the entire cross-section,
integration of the appropriate expressions based on the law enables values of
the kinetic energy correction factor α and the momentum correction factor β
to be evaluated. Hence we obtain
( )3/2
15 2 9 3 f
α=1+ A f − A
8 4 2
and
5
β = 1 + A2 f
8
With A = 2.5, we have

α = 1 + 11.72 f − 12.43 f 3/2 (6.33)

β = 1 + 3.91 f (6.34)
for all values of roughness and Reynolds number. For fully developed turbu-
lent flow the maximum values of α and β in practice are thus about 1.1 and
1.04, respectively, but considerably higher values may be realized if the flow
is subjected to other disturbances.
If transfer of heat takes place across the boundary surface, variations of
temperature and other properties, especially viscosity, with distance from the
boundary may affect the results appreciably.

6.11 ENERGY CONSIDERATIONS


So far, the analysis of the flow through pipe was based on continuity and
momentum equations only. But by using the energy equation and its con-
servation law, additional insight into the nature of pressure loss in internal
viscous flows can be gained. For this let us consider steady flow through the
piping system with a reducing elbow, shown in Figure 6.11.
Flow enters through the pipe at 1 and leaves at 2 of the control volume (cv)
shown in Figure 6.11 with control surface (cs). The general form of energy
equation given by the first law of thermodynamics is

∫ ∫

Q̇ − Ẇs − Ẇshear − Ẇother = eρdV + (e + pv)ρV · dA (6.35)
∂t cv cs
where Q̇ is the volume flow rate through the control volume, Ẇs is shaft work,
Ẇshear is the shear work, Ẇother is other forms of work, p is the pressure,

ISTUDY
6.11. ENERGY CONSIDERATIONS 361

z
1

y
Flow
cv x

FIGURE 6.11 Flow through the pipe with a reducing elbow.

v is the specific volume, ρ is the density, dV is the elemental volume, dA is


elemental area and e is the specific energy, given by

V2
e=u+ + gz
2
where u is the internal energy, V is flow velocity, z is the elevation above a
specified reference and g is the gravitational acceleration.
Assuming that the flow is steady and incompressible with Ẇs = 0, Ẇother =
0, and u, p and ρ are uniform at 1 and 2, the energy equation can be simplified
as ∫ ∫

Q̇ − Ẇshear = eρdV + (e + pv)ρV · dA
∂t cv cs
In this relation, the work due to shear stress also becomes zero, since even
though the shear stresses are present at the walls of the elbow, the velocity at
the walls are zero, due to no-slip condition. Also, the first term on the right-
hand side is zero for steady flow. Thus the above energy relation reduces
to
Z Z
V22 V12
Q̇ = ṁ(u2 − u1 ) + ṁ (p2 v2 − p1 v1 ) + ṁg(z2 − z1 ) + ρV2 dA2 − ρV1 dA1
A2 2 A1 2
But v = 1/ρ, thus
  R R
V22 V12
Q̇ = ṁ(u2 − u1 ) + ṁ p2
ρ
− p1
ρ
+ ṁg(z2 − z1 ) + A2 2
ρV2 dA2 − A1 2
ρV1 dA1

(6.36)

It is essential to note that in the derivation of Equation (6.36), the velocity


at stations 1 and 2 were assumed to be uniform. But in viscous flows the
velocity at a section is not uniform. However, using the average velocity into
Equation (6.36) the integrals in the equation can be eliminated. To do this
let us define a kinetic energy coefficient in the subsection to follow.

ISTUDY
362 CHAPTER 6. FLOW THROUGH PIPES

6.11.1 Kinetic Energy Coefficient


The kinetic energy coefficient, α, is defined such that
∫ ∫ 2 2
V2 V V
ρV dA = α ρV dA = α ṁ
A 2 A 2 2
Thus ∫
A
ρV 3 dA
α= 2 (6.37)
ṁV

6.11.2 Head Loss


Using the definition of kinetic energy coefficient, the energy equation (6.36)
can be written as
( ) ( 2 2
)
p2 p1 α2 V 2 α1 V 1
Q̇ = ṁ(u2 − u1 ) + ṁ − + ṁg(z2 − z1 ) + ṁ −
ρ ρ 2 2

Diving by ṁ, we have


2 2
δQ p2 p1 α2 V 2 α1 V 1
= u2 − u1 + − + gz2 − gz1 + −
dm ρ ρ 2 2
or ( 2
) ( 2
)
p1 α1 V 1 p2 α2 V 2 δQ
+ − + = (u2 − u1 ) − (6.38)
ρ 2 ρ 2 dm
In Equation (6.38), the term
( 2
)
p αV
+ + gz
ρ 2

represents the mechanical energy per unit mass at a cross section. The group
of terms
δQ
(u2 − u1 ) −
dm
is equal to the difference in mechanical energy per unit mass between sections
1 and 2. This represents the irreversible conversion of mechanical energy
at section 1 to unusable thermal energy (also known as the internal energy)
(u2 − u1 ) and loss of energy via heat transfer − dm
δQ
. This group of terms is
the head loss, h. Thus
( 2
) ( 2
)
p1 α1 V 1 p2 α2 V 2
+ + gz1 − + + gz2 =h (6.39)
ρ 2 ρ 2

This equation can be used to calculate the pressure difference between any
two points in a piping system, provided the head loss, which is the sum of the

ISTUDY
6.11. ENERGY CONSIDERATIONS 363

major loss h due to frictional effects in fully developed flow in constant-area


tubes, and minor losses, due to the entrances, fittings, area change and so on.

EXAMPLE 6.9 A 75 cm diameter water pipe carries 0.66 m3 /s. At point


1 the pressure is 175 kPa (gauge) and the elevation is 36 m. At point 2, 1500
m from point 1, the pressure is 310 kPa (absolute) and the elevation is 30 m.
Determine the friction coefficient of the pipe.

Solution Given, d = 0.75 m, L = 1500 m, Q̇ = 0.66 m3 /s, p1 = 175 + 101 =


276 kPa, p2 = 310 kPa, z1 = 36 m, z2 = 30 m.

For water, ρ = 103 kg/m3 , µ = 10−3 kg/(m s).

The head at 1 and 2 are

p1
hf 1 = z1 +
ρg

276 × 103
= 36 +
103 × 9.81

= 36 + 28

= 64 m
p2
hf 2 = z2 +
ρg

310 × 103
= 30 +
103 × 9.81

= 30 + 31.6

= 61.6 m

Thus the net head loss is

hf = hf 1 − hf 2

= 64 − 61.6

= 2.4 m

By Equation (6.11),

L u2
hf = f
d 2g

ISTUDY
364 CHAPTER 6. FLOW THROUGH PIPES

The average velocity u is


u=
A

4Q̇
=
πd2
4 × 0.66
=
π × 0.752

= 1.49 m/s

Therefore,
1500 1.492
hf = × f
0.75 2 × 9.81

= 226.31 f

hf
f=
226.31
2.4
=
226.31

= 0.0106

Note: The friction coefficient f determined can be validated with Equation


(6.28),
1 ( √ )
√ = 2 log10 Re f − 0.8
f
The Reynolds number is
ρud
Re =
µ

103 × 1.49 × 0.75


=
10−3

= 1.1 × 106

Substituting Re and f in Equation (6.28), we have


1 [ √ ]
√ = 2 log10 (1.1 × 106 ) × 0.0106
0.0106

9.7 = 9.3

The LHS ≈ RHS, thus the value of f found is correct.

ISTUDY
6.12. MINOR LOSSES 365

6.12 MINOR LOSSES


In many circuits, the pipe flow may be required to pass through fittings, such
as bends, nozzles and diffusers involving abrupt changes in area. Because of
these passages, additional losses are encountered. The primary loss associated
in these passages is separation losses, in which energy is dissipated due to the
violent mixing in the separation zones. Usually these losses are considerably
smaller than the frictional loss and hence termed minor losses. The Minor
head loss may be expressed as
2
V
hminor = K (6.40)
2
where K is the loss coefficient, which has to be determined experimentally
and V is the mean velocity in the pipe.
Minor head loss may also be expressed as
2
Le V
hminor =f (6.41)
d 2
where Le is an equivalent length of straight pipe.
For flow through pipe bends and fittings, the loss coefficient, K, is found
to vary with pipe diameter. Consequently, the length, Le /d, tends toward a
constant for different sizes of a given type of fitting.

6.12.1 Inlets and Exits


If not designed properly, a inlet to a pipe can cause considerable head loss.
If the inlet has sharp corners, the flow at the corners would separate and
form vena contracta. The flow would be accelerated while passing through
the throat of the vena contracta. When the flow deceleration after the vena
contracta would experience unconfined mixing. This mixing would result is
loss of mechanical energy. Some inlet geometries and their loss coefficients are
shown in Figure 6.12. It is seen that the entrance loss is almost negligible for
a well-rounded inlet with r/d ≥ 0.15.
The loss coefficient for sudden area changes depends on the area ratio of
the expansion or contraction. Variation of loss coefficient with area ratio of
suddenly expanded and suddenly contracted ducts as a function or area ratio
is shown in Figure 6.13.
2
The kinetic energy per unit mass, αV /2, is completely dissipated by
mixing when flow discharges from a duct into a large reservoir. The situation
corresponds to flow through an abrupt (or sudden) expansion with AR = 0
(that is A2 → ∞), as shown in Figure 6.13.

ISTUDY
366 CHAPTER 6. FLOW THROUGH PIPES

Reentrant, K = 0.78

Square-edged, K = 0.5

d
r
Rounded-entrance, r/d 0.02 0.06 ≥ 0.15
K 0.28 0.15 0.04

FIGURE 6.12 Three types of pipe entrance and their loss coefficients,
2
based on hminor = K V2 .

FIGURE 6.13 Loss coefficient variation with area ratio.

EXAMPLE 6.10 A smooth pipe of diameter 75 mm and length 100 m


is attached to a large water tank. If the pipe has to discharge at a constant
rate of 0.008 m3 /s of water to atmosphere, what should be the height of
water above the centreline of the pipe. The inlet of the pipe is square-edged.
The density and viscosity to water are 1000 kg/m3 and 1 × 10−3 kg/(m s),
respectively.

Solution Given, Q̇ = 0.008 m3 /s, D = 0.075 m, L = 100 m.

The free-surface of water in the tank is open to atmosphere. Let subscripts


1 and 2, respectively, refer to the free-surface of water and centreline of the
pipe.

ISTUDY
6.12. MINOR LOSSES 367

The mean velocity of water in the pipe is


V =
A

4Q̇
=
πD2
4 × 0.008
=
π × 0.0752

= 1.81 m/s

The Reynolds number is


ρV D
Re =
µ

1000 × 1.81 × 0.075


=
10−3

= 1.36 × 105

From Figure 6.8, for Re = 1.36 × 105 , we have f = 0.017.

By Equation (6.39),
( 2
) ( 2
)
p1 α1 V 1 p2 α2 V 2
+ + gz1 − + + gz2 = h = hf + hminor
ρ 2 ρ 2

where
2 2
LV V
hf = f , and hminor = K
D 2 2
Also, p1 = p2 = patm , V1 = 0, V2 = V , and α ≈ 1.

Assuming z2 = 0, we have
2 2 2
V LV V
gz1 − =f +K
2 D 2 2
2 [ ]
V L
z1 = f +K +1
2g D
From Figure 6.12, K = 0.5, therefore,
[ ]
1.812 100
z1 = 0.017 × + 0.5 + 1
2 × 9.81 0.075

= 4.04 m

ISTUDY
368 CHAPTER 6. FLOW THROUGH PIPES

6.12.2 Contractions and Enlargements


Minor loss coefficients for sudden contractions and expansions in circular
2
ducts, based on the larger V /2, are given in Figure 6.13. Thus looses for
2
sudden contraction are based on V2 /2, and for sudden expansion the looses
2
are based on V1 /2.
Losses due to area change can be reduced by installing a nozzle or diffuser
between sections of straight pipe. Loss coefficients (K) for gradual contrac-
tion, shown in Figure 6.14. The K values listed in Table 6.2 are valid for both
round and rectangular ducts.

Flow
θ◦ A1
A2

FIGURE 6.14 Contraction in a pipe.

TABLE 6.2 Loss coefficient (K) for gradual contraction


θ (degrees)
A1 /A2 10 15-40 50-60 90 120 150 180
0.50 0.05 0.05 0.06 0.12 0.18 0.24 0.26
0.25 0.05 0.04 0.07 0.17 0.27 0.35 0.41
0.10 0.05 0.05 0.08 0.19 0.29 0.37 0.43

Losses in diffusers are usually presented in terms of pressure recovery co-


efficient, Crecov , defined as the ratio of static pressure rise to the inlet dynamic
pressure
p2 − p1
Crecov = 2 (6.42)
1
2 ρV 1

where p1 and p1 , respectively, are the static pressure at the exit and inlet of
the diffuser and V1 is the average velocity at the diffuser inlet.
The pressure recovery coefficient for conical diffusers with fully devel-
oped turbulent pipe flow at the inlet, as a function of geometry is shown in
Figure 6.15. From Figure 6.15 it is seen that, for every AR there is an N/R1
above which there is no increase in pressure recovery. Similarly, for a given
N/R1 there is an optimum AR for maximum pressure recovery. Diffuser
pressure recovery is found to be independent of Reynolds number for inlet
Reynolds numbers greater than 7.5 × 104 [McDonald, A.T., and R.W. Fox,
“An Experimental Investigation of Incompressible Flow in Conical Diffusers”,
International Journal of Mechanical Sciences, Vol. 8, No. 2, 1966, pp. 125–
139.]. Diffuser pressure recovery with uniform inlet flow is somewhat better
than that for fully developed inlet flow. Performance maps for plane wall,
conical, and annular diffusers for a variety of inlet flow conditions are given

ISTUDY
6.12. MINOR LOSSES 369

by Runstadler [Runstadler, P.W., Jr., Diffuser Data Book, Technical Note


186, Hanover, NH: Creare, Inc., 1975.].

4.0
18 16 14 12 8 6
2
3.0

N
AR = 1+ R1 tan φ
4
Area ratio, AR = A2 /A1

φ
0.70
R1 p1 p2

20 ◦
2.0

15 ◦ φ =
N 0.60
1.8
2

5◦
=
=


1.6

0.50
10 ◦

1.4 0.45
=

Crecov 0.40
1.3
0.35

1.2 0.30

18
16
14 12 8 6 4

0.5 0.8 1.0 1.5 2.0 3.0 4.0 5.0 8.0 10.0
Nondimensional length, N/R1
FIGURE 6.15 Pressure recovery coefficient for conical diffusers.

The static pressure rise in the direction of flow in a diffuser may cause
flow separation. For some geometries, the outlet flow is distorted and this
distortion may lead to pulsations. For wide angle diffusers, vanes or splitters
can be used to suppress stall and improve pressure recovery [Reneau, L.R.,
J.P. Johnston, and S.J. Kline, “Performance and Design of Straight,
Two-Dimensional Diffusers”, Transactions of the ASME, Journal of Basic
Engineering, Vol. 89D, No. 1, March 1967, pp. 141–150.]. The pressure
recovery coefficient, Crecov , can be related to head loss, as follows. Using the
definition of kinetic energy coefficient, α, the energy equation (6.36) can be
written as
( ) ( 2 2
)
p2 p1 α2 V 2 α1 V 1
Q̇ = ṁ(u2 − u1 ) + ṁ − + ṁg(z2 − z1 ) + ṁ −
ρ ρ 2 2
Differentiating with respect to ṁ, we have
2 2
∂Q p2 p1 α2 V 2 α1 V 1
= u2 − u1 + − + gz2 − gz1 + −
dm ρ ρ 2 2

ISTUDY
370 CHAPTER 6. FLOW THROUGH PIPES

or
( 2
) ( 2
)
p1 α1 V 1 p2 α2 V 2 ∂Q
+ + gz1 − + + gz2 = (u2 − u1 ) − (6.43)
ρ 2 ρ 2 dm

The term ( 2
)
p αV
+ + gz
ρ 2
represents the mechanical energy per unit mass at a cross section. The term
∂Q
(u2 − u1 ) −
dm
gives the difference in mechanical energy per unit mass between sections 1 and
2. It represents the irreversible conversion of mechanical energy at section 1
to unwanted thermal energy (u2 − u1 ) and loss of energy via heat transfer
(−∂Q/dm). The group (u2 − u1 ) − ∂Q/dm is termed as the total head loss,
ht , thus ( ) ( )
2 2
p1 α1 V 1 p2 α2 V 2
+ + gz1 − + + gz2 = ht (6.44)
ρ 2 ρ 2
If the gravity is neglected, and α1 = α2 = 0, Equation (6.43) reduces to
( 2
) ( 2
)
p1 V1 p2 V2
+ − + = ht = h
ρ 2 ρ 2
Thus
2 2
V V p2 − p1
h= 1 − 2 −
2 2 ρ
2
[( 2
) ]
V1 V2 p2 − p1
= 1− 2 − 2
2 V 1
ρV 1 2 1

2
[( 2
) ]
V V2
= 1 1− 2 − Crecov
2 V1
By continuity,
A1 V 1 = A2 V 2
Therefore, [ ]
2 ( )2
V1 A1
h= 1− − Crecov
2 A2
or
2 [ ]
V1 1
h= 1− − Crecov (6.45)
2 (AR)2

ISTUDY
6.12. MINOR LOSSES 371

For frictionless flow, h = 0; for this case Equation (6.45) gives the ideal
pressure recovery coefficient, Cideal , as

1
Cideal = 1 − (6.46)
(AR)2

This result can be obtained by applying Bernoulli equation, together with the
continuity equation, to frictionless flow through the diffuser. Thus the head
loss for flow through an actual diffuser may be written as
2
V
h = (Cideal − Crecov ) 1 (6.47)
2

EXAMPLE 6.11 In a water pipeline there is an abrupt change in diameter


from 100 mm to 200 mm. If the head lost due to separation when the flow is
from the smaller to larger pipe is 0.6 m greater then the head lost when the
same flow is reversed. Determine the volume flow rate.

Solution Given, d1 = 100 mm and d2 = 200 mm. Therefore,


( )2
A1 100
= = 0.25
A2 200
( )2
A2 200
= =4
A1 100

A1 A2
From Figure 6.13, for A2 = 0.25, K12 = 0.35 and for A1 = 4, K21 = 0.48.

Since the flow rate is the same in both the pipes and u1 is the average speed
in smaller pipe when the flow is from smaller pipe to larger pipe and u2 is the
average speed once again in the smaller pipe when the flow is reversed (i.e.
from the larger pipe to smaller pipe), we have

u1 = u2 = u

Also, given that


h21 − h12 = 0.6

From Equation (6.40),

u1 2 u2
h12 = K12 = 0.35
2g 2g

and
u2 2 u2
h21 = K21 = 0.48
2g 2g

ISTUDY
372 CHAPTER 6. FLOW THROUGH PIPES

Therefore,

u2
h21 − h12 = (0.48 − 0.35)
2g

u2
0.6 = (0.48 − 0.35)
2g

2 × 9.81 × 0.6
u2 =
0.48 − 0.35

= 90.55

u = 9.512 m/s

The volume flow rate is

Q̇ = Au

πd21
= × 9.512
4

π × 0.12
= × 9.512
4

= 0.0747 m3 /s

6.12.3 Loss in Bends


The head loss associated with bends is larger than the loss experiences by
fully developed flow through a straight section of equal length. The loss in
a bend can be represented in terms of an equivalent length of straight pipe.
The equivalent length depends on the relative radius of curvature of the bend.

6.12.4 Valves and Fittings


Losses associated with values and fittings also can be expressed in terms of
an equivalent length of straight pipe. Some representative data is given in
Table 6.2. All resistances are given for fully open valves. Loses increase
considerably when the valves are partially open.
Fittings in a pipe system may be threaded, flanged, or welded connections.
For small diameters, threaded joints are common. Large pipe systems usually
have flanged or welded joints. The minor losses caused by the bends, valves
and fittings can be a large fraction of the overall system loss. The equivalent
lengths for some valves and fittings are listed in Table 6.3.

ISTUDY
6.12. MINOR LOSSES 373

TABLE 6.3 Dimensionless equivalent lengths (Le /d) for valves and
2
fittings, hminor = f Lde V2 .

Fitting type Le /d
Gate valve 8
Globe valve 340
Angle valve 150
Ball valve 3
Lift check valve:
globe lift 600
angle lift 55
Foot valve with strainer:
poppet disk 420
hinged disk 75
Standard elbow:
90◦ 30
45◦ 16
Return bend, closed pattern 50
Standard tee:
flow through run 20
flow through branch 60

EXAMPLE 6.12 A smooth-walled pipeline 3500 m long is to connect


two water reservoirs 7 m different in elevation. Entry is abrupt and outlet
is a conical diffuser. There are two globe valves and five 60◦ bends are also
required in the pipeline. If gravity is to cause the flow, what should be the
diameter so that the discharge is 0.03 m3 /s. Take water temperature as 20◦ C.
Assume the equivalent length for conical diffuser as 6 diameters of the pipe
and that for each bend as 25 diameters of the pipe.

Solution First of all let us make a guess for the Reynolds number. Con-
sidering the length of the pipe and elevation, it is proper to guess the flow as
turbulent.

Trial 1: Let Re = 104 . Based on this Re let us find the friction factor f , from
Figure 6.8.

For Re = 104 , from Figure 6.8, we have f = 0.0305.

By Equation (6.11),
L u2
hf = f
d 2g

Also,
Q̇ Q̇
u= =
A πd2 /4

ISTUDY
374 CHAPTER 6. FLOW THROUGH PIPES

Therefore,
( )2
L 4Q̇
hf = f
2gd πd2

8f LQ̇2
d5 =
π 2 ghf
8 × 0.0305 × 3500 × 0.032
=
π 2 × 9.81 × 7
= 1.134 × 10−3

d = (1.134 × 10−3 )1/5


= 0.258 m

The average velocity is


u=
πd2 /4
4 × 0.03
=
π × 0.2582
= 0.574 m/s

The Reynolds number is


ρud
Re =
µ
103 × 0.574 × 0.256
=
10−3
= 1.47 × 105

Thus, the first trial of Re is underestimated.


Trial 2: Let Re = 1.5 × 105 .
For this Re, from Figure 6.8, f = 0.158. This gives
0.0158
d5 = × (1.134 × 10−3 )
0.0305

= 0.587 × 10−3

d = (0.587 × 10−3 )1/5

= 0.226 m

ISTUDY
6.12. MINOR LOSSES 375

Therefore,
4 × 0.03
u=
π × 0.2262

= 0.748 m/s
The Reynolds number is
103 × 0.748 × 0.226
Re =
10−3

= 1.69 × 105
This estimate is better than the first trial. This may be taken as the good
enough for engineering applications.

The minor losses are the following.

(1) For abrupt change,


u2
hminor = K
2
where K = 0.5, from Figure 6.12. Thus,
u2
hminor = K
2

L u2
=f
d 2g

L Kg
=
d f

0.5 × 9.81
=
0.0158

= 310

L = 310 × 0.226

= 70 m
(2) For diffuser inlet
L = 6d = 6 × 0.226

= 1.356 m
(3) For globe valve, from Table 6.3
L = 340d = 340 × 0.226

= 76.84 m

ISTUDY
376 CHAPTER 6. FLOW THROUGH PIPES

(4) For bend,


L = 25d = 25 × 0.226

= 5.65 m
The total additional length due to the fittings is
L = 70 + 1.356 + (2 × 76) + (5 × 5.65)

= 251.6 m
Thus the effective length is (3500 + 251.6) = 3751.6 m.

Assuming the 2nd approximation,


3751.6
d5 = (0.587 × 10−3 ) ×
3500

= 0.629 × 10−3
vd = (0.629 × 10−3 )1/5

= 0.229 m

Thus, d = 0.229 m is the diameter required.

EXAMPLE 6.13 Water flows through a 0.5 m diameter tube of average


roughness 0.05 mm. If the frictional loss gradient (hf /L) is 0.005, determine
the flow rate.

Solution Given, d = 0.5 m, hf /L = 0.005, and ε = 0.05 mm.

By Equation (6.11),
L u2
hf = f
d 2g
Therefore,
hf f u2
=
L d 2g

f u2
0.005 =
0.5 × 2 × 9.81

= f u2 × 0.102

0.005
u=
0.102 f

0.221
= √
f

ISTUDY
6.13. NONCIRCULAR DUCTS 377

ε 0.05
= = 1 × 10−4
d 500
For ε
d = 1 × 10−4 , from Figure 6.8,

fmin ≈ 0.0119

Therefore,
0.221
u= √
0.0119

= 2.03 m/s

The Reynolds number is


ud
Re =
ν
2.03 × 0.5
=
10−6

= 1.015 × 106

For Re = 1.015 × 106 and dε = 1 × 10−4 , the value of f from Figure 6.8, agrees
closely with the guessed value of f = 0.0119.

The flow rate is

Q̇ = Au

πd2
= u
4

π × 0.52
= × 2.03
4

= 0.399 m3 /s

6.13 NONCIRCULAR DUCTS


The empirical correlations of pipe flow may also be used for noncircular ducts,
provided their cross sections are not too exaggerated. Thus ducts of square
or rectangular cross section may be treated if the ratio of height to width is
less than 4.
The correlations for turbulent pipe flow are extended for noncircular ducts,
using the hydraulic diameter, dh , defined as
4 × the cross-sectional area
dh =
Perimeter

ISTUDY
378 CHAPTER 6. FLOW THROUGH PIPES

that is,
4A
dh = (6.48)
P
instead of the diameter, d. In Equation (6.48), A is the cross-sectional area,
and P is the wetted perimeter, the length of wall in contact with the flowing
fluid at any cross section. The factor 4 is introduced so that the hydraulic
diameter will equal the duct diameter for a circular geometry. For a circular
duct, A = πd2 /4 and P = πd, so that
4A
dh =
P
4 × (πd2 /4)
=
πd
=d
For a rectangular duct with b and height h, A = bh and P = 2(b + h), so that
bh
dh =
2(b + h)
If the aspect ratio, AR, is defined as AR = h/b, then
2h
dh =
1 + AR
The hydraulic diameter concept can be applied in the approximate range
1
4 < AR < 4.
Losses due to secondary flows increase rapidly for more extreme geome-
tries, so that the correlations are not valid for wide, flat ducts, or to ducts of
triangular or other irregular shapes.
EXAMPLE 6.14 A 30 m long steel annulus is connected to a water tank, as
shown in Figure E6.13. If the level of water has to be 4 m above the annulus,
to maintain a flow rate of 0.01 m3 /s, determine the friction coefficient and
the average surface roughness of the annulus surface. Neglect the entrance
effect and take the density and kinematic viscosity of water as 1000 kg/m3
and 1.02 × 10−6 m2 /s, respectively.
1

2
Water
60 mm

30 mm

FIGURE E6.13 Flow through an annulus.

ISTUDY
6.13. NONCIRCULAR DUCTS 379

Solution Given, l = 30 m, h = 4 m, Q̇ = 0.01 m3 /s, ρ = 1000 kg/m3 and


ν = 1.02 × 10−6 m2 /s.

The flow area is

A = π × (0.052 − 0.032 )

= 5.0265 × 10−3 m2

The average velocity is


u=
A
0.01
=
5.0265 × 10−3

= 1.99 m/s

By steady-flow energy equation, Equation (6.39), we have

( 2
) ( 2
)
p1 α1 V 1 p2 α2 V 2
+ + gz1 − + + gz2 = ghf
ρ 2 ρ 2

But p1 = p2 = pa , V1 ≈ 0, V2 = u in the pipe and α2 = 0. Therefore,

L u2 u2
hf = f = (z1 − z2 ) −
dh 2g 2g

where (z1 − z2 ) = h and dh is the hydraulic diameter, given by

4×A
dh =
P

P = π(0.03 + 0.06)

4 × π4 (0.062 − 0.032 )
dh =
π(0.03 + 0.06)

0.062 − 0.032
=
2 × 0.045

= 0.03 m

ISTUDY
380 CHAPTER 6. FLOW THROUGH PIPES

Therefore,
( )
u2 L
h= 1+f
2g dh

L 2g h
f = 2 −1
dh u
( )
dh 2g h
f= −1
L u2
( )
0.03 2 × 9.81 × 4
= −1
30 1.992

= 0.0188

The Reynolds number is

u 2 dh
Re =
ν
1.99 × 0.03
=
1.02 × 10−6

= 5.85 × 104

For Re = 5.85 × 104 and f = 0.0188, from Moody’s chart,


ε
= 0.0009
d
Therefore, the average surface roughness of the annulus is

ε = 0.0009 × 40

= 0.036 mm

6.14 PIPE FLOW SOLUTION


Once the head loss is calculated, pipe flow problems can be solved using energy
equation.

6.14.1 Single-Path Systems


For single-path systems Equation (6.39) is the computing equation. The pres-
sure drop through a pipe system is a function of volume flow rate, elevation
change, and the total head loss. The total head loss consists of major losses

ISTUDY
6.15. MULTIPLE-PATH SYSTEMS 381

due to friction in constant-area sections and minor losses due to fittings, area
changes, etc. The pressure drop in the functional form is
∆p = f1 (L, Q, d, e, ∆z, ρ, µ, system configuration) (6.49)
For incompressible flow through pipes, the fluid properties are constant. The
roughness, elevation change, and system configuration depend on the pipe
system layout. For a given system and fluid, once the roughness and the
elevation change are known, the pressure drop reduces to
∆p = f2 (L, Q, d) (6.50)
Equation (6.50) relates four variables. Any one of these may be the unknown
quantity in a practical flow situation. Thus there are four possibilities;

(a) L, Q, d known and ∆p unknown.


(b) ∆p, Q, d known and L unknown.
(c) ∆p, L, d known and Q unknown.
(d) ∆p, L, Q known and d unknown.

Cases (a) and (b) can be solved directly, using the continuity and energy
equations and loss data. Cases (c) and (d) can also be solved with continu-
ity and energy equations and loss data, but direct solution is not possible.
Iteration is required to solve (c) and (d).

6.15 MULTIPLE-PATH SYSTEMS


In practical systems such as water supply lines, complex pipe networks need
to be analyzed. A representative multi-pipe system, with two nodes (node
A and node B) and three branches, is shown in Figure 6.16. The total flow
rate into the system must be distributed among the branches. But the flow
rate through each branch is not known. However, the pressure drop for each
branch is the same, (pA − pB ). This data is sufficient for iterative solution for
the flow rate through each branch.

Branch 1

Flow
Branch 2

Node A Node B
Branch 3

FIGURE 6.16 A multi-pipe system.

6.16 SUMMARY
For a circular pipe, the diameter, d, is usually taken as the characteris-
tic dimension or the linear measurement representative of the flow pattern.

ISTUDY
382 CHAPTER 6. FLOW THROUGH PIPES

Reynolds number for flow through a circular pipe is

ρud
Red =
µ

In terms of ν, the Reynolds number becomes


ud
Red =
ν
The Reynolds number is the ratio of inertia to viscous forces.
In a laminar flow the viscous forces are predominant and in a turbulent
flow the inertia forces are predominant.
The velocity at which the transition from laminar to turbulent nature
begins is termed critical velocity.
The Reynolds number at which the flow transits from laminar to turbulent
is termed critical Reynolds number. For circular pipes the theoretical value
of critical Reynolds number is 2300, based on the diameter. But in practical
(or actual) conditions, the transition from laminar to turbulent flow occurs at
Reynolds numbers between 2000 to 4000.
This implies that, apparently there is no precise upper limit to the value of
Reynolds number at which the change from laminar to turbulent flow occurs.
However, there is a definite lower limit for the critical Reynolds number.
When Reynolds number is less than the lower limit any disturbance in the
flow is damped out by the viscous forces.
For fully developed steady flow, the x-component of the momentum equa-
tion is
Fx = 0
The velocity profile can be expressed as
( )[ ( r )2 ]
R2 ∂p
u=− 1−
4µ ∂x R

The volume for rate in a pipe is given by


π∆pD4
Q̇ =
128 µL
where D is the pipe diameter and L its length.
The average velocity, V is

V =
A
u ( r )2
=1−
U R
Shear stress τw at the surface of the pipe is
R ∂p
τw = − [τrx ]r=R = −
2 ∂x

ISTUDY
6.16. SUMMARY 383

The head loss due to friction can also be expressed as

∆p f L u2
hf = =
ρ d 2

The power required to compensate for this pressure loss can be expressed
as
Power = ∆p × Q̇

The shear stress τw and dp/dx are related by the expression

A dp
τw =
P dx
Poiseuille’s Law law states that, “the velocity of a liquid flowing through
a capillary is directly proportional to the pressure of the liquid and the fourth
power of the radius of the capillary and is inversely proportional to the vis-
cosity of the liquid and the length of the capillary”.
In the case of smooth flow (laminar flow), the volume flow rate, Q̇, is given
by the pressure difference divided by the viscous resistance. This resistance
depends linearly upon the viscosity and the length, but the fourth power
dependence upon the radius is dramatically different. Poiseuille’s law is found
to be in reasonable agreement with experiment for uniform liquids (called
Newtonian fluids) in cases where there is no appreciable turbulence.
For turbulent flow with Reynolds number greater than 2000 also the flow
depends on the roughness, ε/d, of the pipe wall.
Moody’s diagram is widely used for predicting the value of the friction
factor f applicable for commercial pipes.
Blasius expressed the friction factor f for the turbulent smooth-pipe curve
as
− 14
f = 0.316 (Re)

This agrees closely with experiments results for Reynolds number between
4000 to 105 . Many relations have been proposed so that f can be calculated
directly for the entire range of ε/d and Re. The one among them which is
considered to be the best is the following proposed by S.E. Haaland
[ ]
1 6.9 ( ε )1.11
√ = −1.8 log10 +
f Re 3.71d

This relation predicts f which is accurate within ±1.5%.


Ludwieg Prandtl was the first to deduce an acceptable expression for the
variation of velocity in turbulent flow past a flat plate and in a circular pipe.
He used his mixing-length concept together with some intuitive assumptions
about its variation with distance from the boundary. Using somewhat different
assumptions, von Karman and others obtained the same basic result. These
expressions are all semi-empirical in that the values of the constant terms have

ISTUDY
384 CHAPTER 6. FLOW THROUGH PIPES

to be determined by experiment, but the forms of the expressions are derived


theoretically.
Application of the principle of dimensional analysis suggests the following
relation, for the velocity distribution
[ ( ) ]
1/2
u R τw y ε
1/2
= ϕ1 , ,
(τw /ρ) ν ρ R R

where ν = µ/ρ is the kinematic viscosity.


The difference between the maximum velocity um at the pipe centre and
the local velocity u depends on y/R only, that is,

um − u
1/2
= ϕ2 (η)
(τw /ρ)

where η = y/R. Equation (6.18) is known as the velocity defect law.


The parameter
1
uτ = (τ0 /ρ) 2

is related to the friction factor f , by Equation (6.13), as



u 2
=
uτ f

Friction coefficient in terms of Reynolds number is

1 ( √ )
√ = 2 log10 Re f − 0.8
f

This expression has been validated by Nikuradse for Reynolds numbers from
5000 to 3 × 106 .
Friction factor for smooth pipes is
( )
1 d
√ = 2 log10 + 1.14
f ε
where d is the diameter of the pipe.
For commercial pipes,
[ ]
1 ε 2.51
√ = −2 log10 + √
f 3.71 d Re f

The kinetic energy coefficient, α, is defined such that


∫ ∫ 2 2
V2 V V
ρV dA = α ρV dA = α ṁ
A 2 A 2 2

ISTUDY
6.16. SUMMARY 385

Thus

A
ρV 3 dA
α= 2
ṁV
In many circuits, the pipe flow may be required to pass through fittings,
such as bends, nozzles and diffusers involving abrupt changes in area. Be-
cause of these passages, additional losses are encountered. The primary loss
associated in these passages is separation losses, in which energy is dissipated
due to the violent mixing in the separation zones. Usually these losses are
considerably smaller than the frictional loss and hence termed minor losses.
The Minor head loss may be expressed as
2
V
hminor =K
2
where K is the loss coefficient, which has to be determined experimentally
and V is the mean velocity in the pipe.
Minor head loss may also be expressed as
2
Le V
hminor = f
d 2
where Le is an equivalent length of straight pipe.
The head loss associated with bends is larger than the loss experiences by
fully developed flow through a straight section of equal length. The loss in
a bend can be represented in terms of an equivalent length of straight pipe.
The equivalent length depends on the relative radius of curvature of the bend.
Losses associated with values and fittings also can be expressed in terms
of an equivalent length of straight pipe.
The correlations for turbulent pipe flow are extended for noncircular ducts,
using the hydraulic diameter, dh , defined as

4 × the cross-sectional area


dh =
Perimeter
that is,
4A
dh =
P
instead of the diameter, d. In Equation (6.48), A is the cross-sectional area,
and P is the wetted perimeter, the length of wall in contact with the flowing
fluid at any cross section. The factor 4 is introduced so that the hydraulic
diameter will equal the duct diameter for a circular geometry.
For single-path systems Equation (6.39) is the computing equation. The
pressure drop through a pipe system is a function of volume flow rate, elevation
change, and the total head loss. The total head loss consists of major losses
due to friction in constant-area sections and minor losses due to fittings, area
changes, etc.

ISTUDY
386 CHAPTER 6. FLOW THROUGH PIPES

In practical systems such as water supply lines, complex pipe networks


need to be analyzed. A representative multi-pipe system, with two nodes
(node A and node B) and three branches, is shown in Figure 6.16. The total
flow rate into the system must be distributed among the branches. But the
flow rate through each branch is not known. However, the pressure drop for
each branch is the same, (pA −pB ). This data is sufficient for iterative solution
for the flow rate through each branch.

6.17 PROBLEMS
6.1 For fully developed laminar flow in a pipe, determine the radial distance
from the pipe centreline at which the velocity equals the average velocity.
[Ans. r = 0.707R]
6.2 Oil with kinematic viscosity of 1.85 × 10−5 m2 /s flows through a circular
pipe of 150 mm diameter. Find the maximum velocity up to which the flow
will remain laminar. [Ans. 0.284 m/s]
6.3 A steel pipe of diameter 150 mm and length 1000 m carries water at a
rate of 0.13 m3 /s. (a) Calculate the pressure drop and (b) the power required
to maintain the flow.
[Ans. (a) 2.796 × 106 Pa, (b) 0.3635 MW]
6.4 A flow of 420 l/min of oil is pumped through a 7.5 cm diameter pipe of
length 62 m, whose outlet is 3m higher than the inlet. The oil density is 0.91
g/cm3 and viscosity is 1.24 poise. Estimate the power required.
[Ans. 671.14 W]
6.5 An oil of density 900 kg/m3 and kinematic viscosity 10−5 m2 /s flows
through a cast-iron pipe of diameter 200 mm with a flow rate of 0.2 m3 /s. If
the pipe is 500 m long, determine the head loss.
[Ans. 117.88 m]
6.6 For an incompressible flow through a circular pipe, (a) derive general
expression for Reynolds number in terms of volume flow rate. (b) If the
Reynolds number is 2000 at a section where the diameter is 10 mm, find the
diameter of the section at which the Reynolds number is 5000.
4Q̇
[Ans. (a) πdν , (b) 4 mm]
6.7 If an oil of viscosity 0.048 kg/(m s) and density 1.4 × 103 kg/m3 , flows
through a 45 m long pipe, experiences a pressure drop of 64 kPa, when the
flow rate is 7.634 × 10−5 m3 /s, determine the pipe diameter.
[Ans. 18 mm]
6.8 A liquid of density 1.18 × 103 kg/m3 and viscosity 0.0045 (N s)/m2 flows
through a tube. If the flow rate per unit length is 12 × 10−6 m3 /s and the
pressure drop is 8.6 kPa, assuming laminar flow, determine the tube diameter.
Also, verify whether the assumption made is justified.
[Ans. 4 mm, Justified]

ISTUDY
6.17. PROBLEMS 387

6.9 (a) Show that for Poiseuille flow, of a Newtonian fluid with viscosity µ,
in a tube of radius R, the wall shear stress, τw , can be obtained from

4µLQ̇
τw =
πR3

(b) Determine the magnitude of τw if the viscosity is 0.003 (N s)/m2 , the


average velocity is 100 mm/s and the tube diameter is 2 mm.
[Ans. 1.2 Pa]
6.10 Oil of viscosity 0.048 kg/(m s) flows through a pipe at the rate of
7.634 × 10−5 m3 /s. If the pressure drop over the pipe length of 45 m is 64
kPa, assuming the flow as fully developed, determine the pipe diameter.
[Ans. 18 mm]
6.11 An oil of density 800 kg/m3 and viscosity 0.001865 kg/(m s) is pumped
through a 6.3 km long pipeline, at the rate of 0.22 m3 /s. If the power required
to drive the pump with overall efficiency is 75 percent is 592 kW, determine
the friction coefficient and the average roughness of the inner surface of the
pipeline.
[Ans. 0.0248, 0.75 mm]
6.12 Water flows through a galvanized tube of diameter 150 mm and length
300 m, at a rate of 50 liters/second. If the head loss due to friction in 17
m, determine the friction coefficient of the tube surface, through (a) the head
loss equation and (b) using moody’s chart. (c) Verify the f calculated using
Equation (6.31b).
[Ans. 0.0208]
6.13 An oil of kinematic viscosity 10−5 m2 /s flows through a cast-iron pipe
of 100 mm diameter and length 120 m. If the head loss due to friction is 5 m,
determine the flow rate and the Reynolds number.
[Ans. 0.014 m3 /s, 1.773 × 104 ]
6.14 A 900 mm long pipe of diameter 75 mm carries an oil of specific gravity
0.85 and kinematic viscosity 0.0033 m2 /s, at the rate o 47 kg/hour. (a) Check
whether the flow is laminar, and (b) find the power required to overcome the
friction in the pipe.
[Ans. (a) The flow is laminar, (b) 551.3 W]
6.15 Determine the diameter of galvanized needed to carry water, with kine-
matic viscosity 1.14 × 10−6 m2 /s, at 85 liters/second, if the length is 180 m
and the head loss due to friction is 9 m.
[Ans. 200 mm]
6.16 A 150 mm diameter pipe reduces abruptly to 10 mm. (a) If the pipe
carries water at 30 liters/second, calculate the pressure loss across the con-
traction. (a) Express the percentage of pressure loss to be expected if the flow
were reversed. Take the coefficient of contraction as 0.6.
[Ans. (a) 3.1 kPa, (b) 139%]

ISTUDY
388 CHAPTER 6. FLOW THROUGH PIPES

6.17 Water flows at the rate of 0.02 m3 /s through a cast-iron pipe of diameter
200 mm and length 500 m. Determine (a) the head loss due to friction and
(b) the pressure drop if the pipe slopes down at 10◦ in the flow direction.
[Ans. (a) 113.75 m, (b) 264.2 kPa]
6.18 Water flows through a 150 mm diameter pipe with surface roughness
0.015 mm. (a) If the mean velocity is 5 m/s, what is the thickness of the
viscous sublayer, δl . (b) What will δl be if the velocity is increased to 6.5
m/s. Assume the density and viscosity of water be 1000 kg/m3 and 0.001
kg/(m s), respectively.
[Ans. (a) 0.0193 mm, (b) 0.0151 mm]
6.19 Gasoline with kinematic viscosity 5 × 10−7 m2 /s flows in a 200 mm
diameter pipe. If the friction head is 0.43 per 100 m, determine the flow rate.
[Ans. 0.0331 m3 /s]
6.20 (a) Determine the maximum volume flow rate at which the flow will
be laminar in a water injection line made of smooth capillary tube of inside
diameter 0.25 mm. (b) Evaluate the pressure drop associated with flow rate
through a section of tubing with length 0.75 m.
[Ans. (a) 0.45 × 10−6 m3 /s, (b) 3.52 × 106 Pa]
6.21 Calculate the diameter of a pipe 800 m long to convey a gas of density
0.7 kg/m3 , at 600 m3 /hour, from a gas-holder to a power station. Delivery is
15 m above the entrance of the pipe, pressure at the holder is 10 cm of water
and at station in 5 cm of water. The friction coefficient of the pipe is 0.011.
[Ans. 0.205 m]

ISTUDY
Chapter 7

Flow with Free Surface

7.1 INTRODUCTION
In our discussions on flow through pipes, in Chapter 6, the flowing fluid has
been assumed to be bounded on all sides by solid surfaces. But in many
situations flow may take place with the uppermost boundary as the free surface
of the fluid itself. A typical example of flow with free surface is the flow of
liquid in a channel. The flow cross-section is not then determined entirely
by the solid boundaries, but is free to change. As a result, the conditions
controlling the flow are different from those governing flow that is entirely
enclosed. Indeed, the flow of a liquid with a free surface is, in general, more
complicated than flow in pipes and other closed passages.
When the liquid is bounded by side walls - such as the banks of a river
or canal–the flow is said to take place in an open channel. The free surface
is usually subjected only to atmospheric pressure and, since this pressure is
constant, the flow is caused by the weight of the fluid. As in pipes, uniform
flow is accompanied by a drop in piezometric pressure, (p + ρgz), but for an
open channel it is only the second term, ρgz, that is significant, and uniform
flow in an open channel is always accompanied by a fall in the level of the
surface.
Natural streams and rivers, artificial canals, irrigation canals are typical
examples of open channels. But pipe-lines or tunnels that are not completely
full of liquid also have the essential features of open channels. Water is the
liquid usually involved, and practically all the experimental data for open
channels relate to water at normal temperature.
Complete solutions of open-channel flow problems are usually more
difficult to obtain than those for flow in pipes, even when the flow is
assumed to be steady and uniform. Unlike pipes which are mostly of circular
cross-section, open cross-section channels may have cross-section ranging from
simple geometrical shapes to the quite irregular sections of natural streams.
Furthermore, the state of the boundary surfaces also vary much more widely
- from smooth to rough. The choice of a suitable friction factor for an open

389

ISTUDY
390 CHAPTER 7. FLOW WITH FREE SURFACE

channel is thus likely to be much more uncertain than a similar choice for a
pipe.
A open channel flow may be uniform or non-uniform, steady or unsteady.
A channel flow is termed uniform when the magnitude and direction of
velocity of the liquid remains invariant throughout the channel. This con-
dition is achieved only when the cross-section of the flow does not change
along the length of the channel, and thus the depth of liquid must be un-
changed. Consequently, uniform flow is characterized by the liquid surface
being parallel to the base of the channel. This implies that for a channel flow
to be uniform it is sufficient if the velocity profile is the same at all cross-
sections and the velocity across any one section of the channel need not be
constant.
Flow in which the liquid surface is not parallel to the base of the channel
is said to be non-uniform. In a non-uniform channel flow the depth of the
liquid continuously varies from one section to another. The change in depth
may be rapid or gradual. Flow with rapid and gradual change in depth are
referred to as rapidly varied flow and gradually varied flow, respectively. It
is important to note that the rapid and gradual terms here refer only to the
variations from section to section along the channel, and not to variation with
time. Uniform flow of course can exist in one part of the channel while varied
flow exits in another part.
When the velocity, and hence the depth, at a particular point in the chan-
nel varies with time the flow is termed unsteady. When the velocity remains
invariant with time the flow is termed steady. In most problems concerned
with open channels the flow is steady - at least approximately. But if there
is a surge wave, for example, depth at a particular point changes suddenly as
the wave passes by and the flow becomes unsteady.
A steady uniform flow, in which the depth of the liquid changes neither
with distance along the channel nor with time, can easily be solved analyti-
cally. The various types of open channel flow with free surface are illustrated
in Figure 7.1. In these diagrams, the slope of the channel is exaggerated.
Most open channels have a gentle slope, of the order of 1 in 1000. In practice,
non-uniform flow is encountered more frequently than uniform flow. This is
especially true for short channels because a certain length of channel is re-
quired for the establishment of uniform flow. But much of the theory of flow
in open channel is based on the behavior of the liquid in uniform flow, owing
to its simplicity.
Like pipe flow, the flow in an open-channel may be either laminar or
turbulent, depending on the relative magnitude of viscous and inertia forces,
the Reynolds number ul/ν, which is used as the criterion for laminar and
turbulent flow for channel flow also. For the characteristic length, l, it is
customary to use the hydraulic mean depth, m, defined as the ratio of the
flow cross-sectional area, A, to the perimeter, P . The lower critical Reynolds
number for open channels seldom occurs in problems of practical interest.
Indeed, in channels of engineering interest, completely turbulent flow may
invariably be assumed because the surface of the flowing liquid occasionally

ISTUDY
7.2. STEADY-FLOW ENERGY EQUATION FOR OPEN CHANNELS 391

Constant depth

Vari
a ble d
epth h
Uniform
flow Variable dept
Non-un
iform fl
ow
Non-un
iform fl
Rapidly varied ow
flow
Gradually varied
flow
Wier
Rapidly varied
A flow
B
Non-uniform C
flow
Unifor
m flow D E
Varied flow Uniform flow
FIGURE 7.1 Various types of open channel flow.

appears smooth and glossy is no indication that turbulent flow does not exist
underneath. The inertia forces usually far outweigh the viscous forces. Thus
it is not ordinarily necessary to consider the effect of Reynolds number, in
detail, for the flow in a channel.
Another important classification of open-channel flow is derived from the
magnitude of the Froude number of the flow. When the velocity of the liquid is
small it is possible for a small disturbance in the flow to travel against the√flow
and thus affect the condition upstream. The Froude number (Fr = V / gh)
is then less than 1.0, and the flow is termed tranquil. When the velocity of
the flow is high that a small disturbance cannot propagate upstream (but is
washed downstream), the Froude number becomes greater than 10 and the
flow is said to be rapid. When the Froude number is equal to 1.0 the flow is
termed critical. Thus, a open-channel flow with a free surface would always
consists of the following four characteristics.

(i) Uniform or non-uniform.


(ii) Steady or unsteady.
(iii) Laminar or turbulent.
(iv) Tranquil or rapid.

7.2 STEADY-FLOW ENERGY EQUATION FOR


OPEN CHANNELS
Let us consider a steady flow of constant density in a open-channel in which
the temperature changes are negligible. Therefore at any particular point the
mechanical energy divided by weight is represented by the sum of the three

ISTUDY
392 CHAPTER 7. FLOW WITH FREE SURFACE

terms:
p u2
+ +z
ρg 2g
If the streamlines are straight and parallel, the pressure at any point in the
flow stream will be governed only by its depth below the free-surface.
When the pressure variation is hydrostatic, a point at which the (gauge)
pressure p is at a depth p/(ρg) below the surface, and so the sum p/(ρg)+z, as
illustrated in Figure 7.2, represents the height of the surface above the datum
level.

Free s
p
urface
ρg

Stream
line

Chann
el bed

z
Arbitrary horizontal datum

FIGURE 7.2 Pressure variation in free-surface flow.

The expression for the mechanical energy divided by weight is thus sim-
plified to
u2
Height of the surface above datum + (7.1)
2g
It is seen that, the height of the individual streamlines above the datum has no
place in the expression. If it is further assumed that at the section considered
the velocity is the same along all streamlines, then Equation (7.1) has the
value of the mechanical energy for the entire system. But in reality, it is
not possible to achieve a uniform distribution of velocity over a section. The
actual velocity distribution in an open channel is influenced both by the solid
boundaries (as in the case of pipes) and by the free surface. Bends in the
channel and irregularities in the boundaries also have an effect.
The irregularities in the boundaries of open channel are usually so large,
and occur in such a random manner, that each channel has its own particular
pattern of velocity distribution. Nevertheless, it may in general be said that
the maximum velocity usually occurs at a point slightly below the free surface
(from 0.05 to 0.25 times the free depth) and that the average velocity, which
is usually about 85% of the velocity at the surface, occurs at about 0.6 of
the full depth below the surface. A typical constant velocity contours in a
rectangular channel, with free surface, is shown in Figure 7.3. The numerical
values on the contours are proportional to mean velocity.
Because of this lack of uniformity of velocity over a cross section, the
velocity head u2 /(2g), representing the kinetic energy of the fluid divided by
weight, has two low a value if calculated from the average velocity u. To

ISTUDY
7.2. STEADY-FLOW ENERGY EQUATION FOR OPEN CHANNELS 393

Free surface

1.2

1.1

1.0
0.9 0.8

FIGURE 7.3 Constant velocity contours in a rectangular channel.

compensate for the error, α u2 /(2g) may be used in place of u2 /(2g), where
α is the kinetic energy correction factor. Experimental results show that the
value of α varies from 1.03 to as much as 1.6 in irregular natural streams, the
higher values generally being found in small channels.
Momentum of the streams is also affected by a nonuniform distribution of
velocity. The rate at which momentum is carried by the fluid past a particular
cross-section is given by β Q̇ρu, where Q̇ represents the volume flow rate, ρ
the density of the liquid and β the momentum correction factor. The value β
typically varies from 1.01 to about 1.2.
In straight channels of rectangular cross-section, however, the effects of
a non-uniform velocity distribution of the calculated velocity head and mo-
mentum flow rate are not of importance. Indeed, other uncertainties in the
numerical data are usually of greater consequence.

7.2.1 Energy Gradient


Friction causes mechanical energy of the flow to be converted into heat, and
thus lost to the environment through heat convection. If the energy loss
divided by weight is denoted by hf , then for steady flow between two sections
1 and 2,
u21 u2
(Height of surface)1 + − hf = (Height of surface)2 + 2
2g 2g
With reference to Figure 7.4, this equation can be written as
u21 u2
h1 + z1 + − hf = h2 + z2 + 2 (7.2)
2g 2g
To account for the non-uniformity of velocity over the cross-section, we may
write
u2 u2
h1 + z1 + α1 1 − hf = h2 + z2 + α2 2 (7.3)
2g 2g
The rate at which mechanical energy is lost to friction may be expressed as
hf /l, where l represents the length of channel over which the head loss hf
takes place. This quantity hf /l may be termed energy gradient. In the case of
uniform flow, u1 = u2 , α1 = α2 and h1 = h2 . Substituting this into Equation
(7.3), we get,
hf = z1 − z2

ISTUDY
394 CHAPTER 7. FLOW WITH FREE SURFACE

Free su
rface
p
ρg
Streamlin
e
h1
Chann h2
el bed

z1 l
z2
Arbitrary horizontal datum
FIGURE 7.4 Flow in a standard channel.

The energy gradient is thus the same as the actual, geometrical, gradient of
the channel bed and of the liquid surface. This is true only for uniform flow
in open channel.

7.3 STEADY UNIFORM FLOW


This is the simplest of open-channel flow, although in practice it is not easy
to achieve. Uniform conditions over a length of the channel are possible only
when there are no influences to cause a change of depth, thus no alteration
of the cross-section of the stream and there is no variation in the surface
roughness of the solid boundaries. Indeed, strictly uniform flow is rarely
achieved in practice, and even approximately uniform conditions are more an
exception than the rule. Nevertheless, when uniform flow is obtained the free
surface becomes parallel to the bed of the channel and the depth from the
surface to the bed is then termed the normal depth.
The basic formula describing uniform flow is due to the French engineer
Antoine de chézy. He deduced the equation from the experimental results on
flow in canals and river Seine, in the year 1769.
In normal (that is, steady uniform) flow there is no change of momentum,
and thus the net force on the liquid is zero. A stretch of a channel in which
these conditions are satisfied is as shown in Figure 7.5. The slope of the

Horizontal
Energy lin
e hf
Free surface
u2 /2g
u

W sin α
F1
F2
1 τ0
α
2
α
l
W
FIGURE 7.5 Flow through a stretch of a slanted channel.

ISTUDY
7.3. STEADY UNIFORM FLOW 395

channel, α, is constant, the length of the channel between planes 1 and 2 is


l and the cross-section area A is constant, and the flow in the stretch of the
channel considered is assumed to be sufficiently far from the inlet and the flow
is fully developed.
Let forces F1 and F2 act at the ends of the control volume of liquid between
sections 1 and 2. Since the cross-sections at 1 and 2 are identical, F1 and F2
are equal in magnitude and have the same line of action; they thus balance
and have no effect on the motion of the liquid. Hydrostatic forces acting on
the sides and bottom of the control volume are perpendicular to the motion,
and so they too have no effect. The only group of forces need to be considered
are those due to gravity and the resistance exerted by the bottom and sides of
the channel. If the average stress at the boundaries is τ0 , the total resistance
force is given by the product of τ0 and the area over which it acts, that is, by
τ0 P l, where P is the wetted area perimeter, as illustrated in Figure 7.6.

A D

B C

FIGURE 7.6 Wetted area perimeter at a section of the channel.

Note that the perimeter, P , in Figure 7.6 does not represent the total perime-
ter of the cross section, since the free surface is not included. Only that part
of the perimeter where the liquid is in contact with the solid boundary is
relevant here, that is the only part where resistance to flow can be exerted.
Further, the effect of the air at the free surface on the resistance is negligible
compared with that of the sides and bottom of the channel.
For the net force in the direction of motion to be zero, the total resistance
must exactly balance the component of the weight W . That is
τ0 P l = W sin α = A l ρg sin α
This gives
A
τ0 = ρg sin α (7.4)
P
For uniform flow, sin α = hf /l, the energy gradient. Denoting this by i (=
sin α), we may write
A
τ0 = ρg i
P
where τ0 is the shear stress at the boundary. In almost all cases of practical
interest, the Reynolds number of the flow in an open channel is high enough
for the flow to become turbulent in which the shear stress at the boundary is
proportional to the square of the mean velocity. Therefore, we can express τ0
as
1
τ0 = ρu2 f
2

ISTUDY
396 CHAPTER 7. FLOW WITH FREE SURFACE

where the friction coefficient, f , is independent of velocity u. Substituting for


τ0 in Equation (7.4), we have

1 2 A
ρu f = ρg i
2 P
This gives
2g A 2g
u2 = i= mi (7.5)
f P f
where m = A/P . The quantity m is termed the hydraulic mean depth or
hydraulic diameter. For the channel shown in Figure 7.6, m is the cross-
sectional area ABCD divided by wetted perimeter, that is, the length ABCD.
Equation (7.5) can be arranged as

2g
u= mi
f

Expressing √
2g/f = C (7.6)
we have

u=C mi (7.7)
This is known as Chézy’s equation.
The discharge through the channel is given by

Q = Au = AC m i (7.8)

The factor C is known as Chézy’s coefficient and its units are L 2 T−1 . There-
1

fore the expression for the magnitude of C depends on the system of units
adapted. We know that, the friction factor, f , depends on the Reynolds
number and the relative roughness, ϵ/d. Thus Chézy’s coefficient C may be
expressed to depend on Re and ϵ/m (the hydraulic mean depth m is used
here as the relevant characteristic length of the system), although for the
flow conditions usually encountered in open channels the dependence on Re
is less significant than the roughness factor ϵ/m. Although open channels
vary widely in the shape of their cross-sections, the use of the hydraulic mean
depth m largely accounts for the differences of shape.
For predicting the value of Chézy’s coefficient C, for a particular channel,
many correlations have been proposed based on experimental data. Among
them the simplest and widely used correlation is by the Irish engineer Robert
Manning, which gives
1
C = m 6 /n
Using Equation (7.7), Manning’s expression can be casted as

m2/3 i1/2
u= (7.9)
n

ISTUDY
7.3. STEADY UNIFORM FLOW 397

The units of u in this equation is m/s. The n in this equation is known


as Manning’s roughness coefficient. Note that, for Equation (7.9) to be di-
mensionally homogeneous n should have dimensions of [TL−1/3 ]. However,
it is not proper that an expression for the roughness of the surface should
involve dimensions in respect to time, and it may be seen from comparison
with Equation (7.6) that the relation would be more logically written as
( )
N g 1/2
u= m2/3 i1/2
n

If N is regarded as a numeric, then n takes the dimensional formula [L1/6 ].


However, because g is not explicitly included, Equation (7.9) must be regarded
as a numerical formula, suitable only for a particular set of units. With the
figures usually quoted for n, the units are those based on the meter and the
second. The expression may be adapted for use with foot-second units by
changing numeric N from 1.0 to 1.49. Then
( )
1.49
u= m2/3 i1/2 (7.9a)
n

This change allows the same numbers for n to be used in either Equation (7.9)
or Equation (7.9a). Table 7.1 lists a few representative values on n, but it
should be noted that they are subject to considerable variation. The values
of n listed in Table 7.1 are valid only for Equations (7.9) and (7.9a).

TABLE 7.1 Manning’s roughness constant n for straight, uniform


channels

Channel type n

Smooth cement, planed timber 0.010


Rough timber, canvas 0.012
Cast iron, good ashlar masonry, brick work 0.013
Vitrified clay, asphalt, good concrete 0.015
Rubble masonry 0.018
Firm gravel 0.020
Canals and rivers in good condition 0.025
Canals and rivers in bad condition 0.035

At this stage, it is essential to note that the flow in channel of small size
or at an unusually small velocity might well have a Reynolds number lower
than that for which the formula is truly applicable. Also, the formulas such as
Chézy’s, that account for the friction in an open channel have no connection
with the Froude number and are thus applicable only to tranquil or rapid
flow, which are steady or uniform.

ISTUDY
398 CHAPTER 7. FLOW WITH FREE SURFACE

Like pipe flow, flow in open channels is also, in addition to frictional losses,
subject to additional losses resulting from the presence of abrupt changes
of cross-section, bends or other disturbances to the flow. However, these
additional energy losses are usually negligible compared to the frictional losses.

7.4 BOUNDARY LAYER IN OPEN


CHANNELS
We saw that the friction factor, f , in pipe flow is closely related to the relative
roughness, ϵ/d. Therefore, it is natural to expect that Chézy’s coefficient C
is also closely related to the roughness of the boundaries of open channels.
For circular pipe the diameter is usually taken as the significant linear dimen-
sion and the hydraulic mean depth is regarded as the characteristic√ length
parameter for open channel. Using the Chézy’s coefficient C = 2g/f , there-
fore, it is possible to re-plot Nikuradse’s data (Figure 6.7) in the manner of
Figure 7.7, with 4m in place of d, because the hydraulic mean depth of the
circular section (with which Nikuradse was concerned) is d/4.
The dependence of C on Reynolds number and relative roughness ϵ/m is
plotted in Figure 7.7. In the rough zone of the graph, C is constant for a
particular value of ϵ/m. But the shape of open channel cross-section lacks
the axial symmetry. Therefore, Figure 7.7 can be expected to represent open-
channel flow only in a qualitative manner. Nevertheless, for the rough zone,
it is of interest to combine Equation (7.6) with Equation (6.31), which gives
the friction factor for turbulent flow in rough pipes:
C 1
√ = √ = 2 log10 (d/ϵ) + 1.14
2g f
log10 (1/ f ) = log10 (C/ 2g)

oth ǫ/m small


Smo
inar

ǫ/m large
Rough
Lam

Transition

log (um/ν)
FIGURE 7.7 Friction factor variation with dimensionless m.
Replacing C with m1/6 /n, as in Equation (7.9) and d with 4m, we get

m1/6
= 2 log10 (4m/ϵ) + 1.14
(19.62)1/2 n

ISTUDY
7.4. BOUNDARY LAYER IN OPEN CHANNELS 399

where
0.0564 m1/6
n=
log10 (14.86 m/ϵ)

Though the numerical factors in this expression are accurate and reliable,
owing to the lack of axial symmetry of open channels, the expression suggests
that, with the logarithmic type of velocity profile to be expected in turbulent
flow, n is not very sensitive to change in ϵ, and even less so to changes in m.
It is not easy to the theories developed for turbulent flow in pipes to solve
open channel flow. This is because the effects of the free surface and of non-
uniform shear stress round the wetted perimeter are uncertain. However, the
qualitative conclusions drawn from Figure 7.7 about the relation of C to the
Reynolds number and the roughness size are valid, at least for rigid channel
boundaries.
In an alluvial channel, with its surfaces composed of movable sand or
gravel, the roughness elements are not permanent, but depend on the flow.
Ripples and dunes may form in the boundary material, and the spacing of
these humps may be much larger than the spacing of the irregularities on
the walls of pipes to or rigid boundaries of open channels. Under such con-
ditions as Nikuradse investigated (where roughness projections are closer to
each other) the wake behind the projections interferes with the flow around
those immediately downstream. The larger irregularities formed in alluvial
channels, however, have a different kind of effect, which, in turn, results in a
much larger frictional loss than the size of the individual particles alone would
cause.
Chézy formula and Manning formula show that, for any given value of
slope, surface roughness and cross-sectional area, the discharge Q increases
with increase in the hydraulic mean depth m. Therefore the discharge is a
maximum when m is a maximum, that is, when, for a given area, the wetted
perimeter is a minimum (since m = A/P ). A cross-section having such a
slope that the wetted perimeter is a minimum is thus, from a hydraulic point
of view, the most efficient.
It may be seen that, of all sections whose sides do not slope inwards towards
the top, the semicircle has the maximum hydraulic mean depth. But, though
semicircular channels are built from prefabricated sections, for other forms
of construction the semicircular shape is impractical. Trapezoidal sections
are very popular, but when the sides are made of loose granular material its
angle of repose may limit the angle of the sides. The most efficient section
will be that with the maximum discharge for a given area and, conversely, the
minimum area for a given discharge.
When the hydraulic efficiency is the primary concern, determining the
most efficient shape of section for a given area is simply a matter of obtaining
an expression for the hydraulic mean depth, differentiating it and equating to
zero to obtain the condition for the maximum. For a channel section in the
form of a symmetrical trapezium with horizontal base, as shown in Figure 7.8,

ISTUDY
400 CHAPTER 7. FLOW WITH FREE SURFACE

the area A and the wetted perimeter P are given by

A = bh + h2 cot α
P = b + 2 h cosec α

But
A
b=− h cot α
h
therefore, the hydraulic mean depth becomes
A A
m= =
P (A/h) − h cot α + 2 h cosec α

h
α h cosec α
b
FIGURE 7.8 A channel with symmetrical trapezium cross-section and
horizontal base.
For a given value of cross-sectional area A, this expression is a maximum
when its denominator is a minimum, that is, when
A
− − cot α + 2 cosec α = 0
h2
The second derivative 2A/h3 is clearly positive and so the condition is indeed
that for a minimum. Thus

A = h2 (2 cosec α − cot α) (7.10)

Substituting this A, we get the maximum value of hydraulic mean depth as

h2 (2 cosec α − cot α)
mmax =
h (2 cosec α − cot α) + h (2 cosec α − cot α)
h
=
2
This shows that for maximum efficiency a trapezoidal channel should be so
proportioned that its hydraulic mean depth is half of the central depth of
flow. For rectangular section (a special case of a trapezium with α = 90◦ )
also the optimum proportions are given by m = h/2. Taking A = bh = 2h2
[from Equation (7.10)] we get b = 2h.
If α can be varied, a minimum perimeter and therefore maximum m is
obtained for α = 60◦ . This shows that the most efficient of all trapezoidal
sections is half a regular hexagon. The concept of most efficient sections con-
sidered here applies only to channels with rigid boundaries. For channels with

ISTUDY
7.4. BOUNDARY LAYER IN OPEN CHANNELS 401

erodible boundaries, such as a sand boundary, the design must take account
of the maximum shear stress τ0 on the boundary.

EXAMPLE 7.1 A concrete channel with slope 0.001 is to be designed to


carry a flow, assuming the manning roughness constant as 0.015. (a) If the
cross-section of the channel is square with sides 2 m determine the flow rate.
(b) If the cross-section of the channel is semi-circular with diameter 2 m, what
will be the flow rate?

Solution Given, i = 0.001, n = 0.015.

(a) For the square channel with b = 2 m, the perimeter, P , and cross-sectional
area, A, are

P = 3b = 3 × 2

= 6m

A=2×2

= 4 m2

The hydraulic mean depth is

A
m=
P
4
=
6
= 0.67 m

The flow velocity, by Equation (7.9), is

m2/3 i1/2
u=
n

(0.67)2/3 × 0.001
=
0.015
= 1.614 m/s

Therefore,

Q = Au

= 4 × 1.614

= 6.456 m3 /s

ISTUDY
402 CHAPTER 7. FLOW WITH FREE SURFACE

(b) For semi-circular section with d = 2 m,

πd
P =
2
π×2
=
2
= 3.14 m

πr2
A=
2
π × 12
=
2
= 1.57 m2

Therefore,
A
m=
P
1.57
=
3.14
= 0.5

The flow velocity is

m2/3 i1/2
u=
n

0.52/3 0.001
=
0.015
= 1.328 m/s

The volume flow rate is

Q = Au

= 1.57 × 1.328

= 2.085 m3 /s

7.4.1 Partially Filled Flow in Closed Conduits


Partially filled flow through closed conduits are encountered in civil engineer-
ing practice. Flow in sewerage and drainage passages are typical examples
of partially filled flow. Because of the free surface of the liquid, the flow is

ISTUDY
7.4. BOUNDARY LAYER IN OPEN CHANNELS 403

governed by the same principles as if it were in a channel completely open at


the top. However, there are some special features resulting owing to the con-
vergence of the boundaries towards the top. For partially filled flow through
circular conduits, shown in Figure 7.9, the area of the cross-section of the
liquid is ( ) ( )
1 1
r2 θ − 2 r sin θ r cos θ = r2 θ − sin 2θ
2 2
and the wetted perimeter is 2rθ. From an equation such as Manning’s [Equa-
tion (7.9)] the mean velocity and the discharge may then be calculated for
any value of θ and hence h.

d
r
θ

FIGURE 7.9 Partially filled flow through a circular conduit.


Variation of the velocity, volume flow rate and depth of partially filled flow
in a channel, such as that shown in Figure 7.9, will be as shown in Figure 7.10.
It will be seen that both the maximum discharge and the maximum mean
velocity are greater than the value for the full conduit. Although it may
be desirable to design such a conduit to operate under the conditions giving
maximum discharge, the corresponding value of h/d is so near unity that in
practice the shortest obstacle or increase in frictional resistance beyond the
design value would cause the conduit to flow completely full.

h/d
1.0

Q/Qfull

u/ufull

0 1.0 u/ufull or Q/Qfull

FIGURE 7.10 Variation of velocity and volume flow rate with depth in a
partially filled flow.

When large fluctuations in discharge are encountered, oval or egg-shaped


sections are commonly used. Thus at low discharges a velocity which is high

ISTUDY
404 CHAPTER 7. FLOW WITH FREE SURFACE

enough to prevent deposition of sediment is maintained. On the other hand,


too large a velocity at full discharge is undesirable as this could lead to ex-
cessive scouring of the lining material.

EXAMPLE 7.2 A 900 mm diameter pipe flows just full when it carries
0.8 m3 /s. What will be the depth and discharge when the velocity is 0.6 m/s?

Solution Given, d = 0.9 m, Q = 0.8 m3 /s and u = 0.6 m/s.

Therefore,

Q
ufull =
A
0.8
=
πd2 /4
4 × 0.8
=
π × 0.92
= 1.26 m/s

Thus,

u 0.6
=
ufull 1.26
= 0.476

u
For ufull = 0.476, from Figure 7.10,

h
= 0.2
d
h = 0.2 × 0.9

= 0.18 m

h
For d = 0.2, from Figure 7.10,

Q
= 0.08
Qfull

Q = 0.08Qfull

= 0.08 × 0.8

= 0.064 m3 /s

ISTUDY
7.5. WAVES AND SURGES IN OPEN CHANNELS 405

7.5 WAVES AND SURGES IN OPEN


CHANNELS
Flow in open channels may be modified by waves and surges of various kinds
that produce unsteady conditions. Any temporary disturbance of the free
surface produces waves. For example, a stone dropped into a water pond
causes a series of small surface waves that travel radially outwards from the
point of disturbance. If the flow along a channel is increased or decreased by
removal or insertion of an obstruction, such as sudden opening or closing of a
sluice gate, surge waves are formed and propagated upstream and downstream
of the obstruction. In certain circumstances, tidal action may cause a surge,
known as a bore, in large rivers. A positive wave is one that results in an
increase in the depth of the stream and a negative one causes a decrease in
depth.
Let us examine a simple positive surface illustrated in Figure 7.11. To
avoid complexity, let us assume the channel to be straight with rectangular
cross-section, uniform width and horizontal base. Also, let the slope of the
bed be zero, to ensure that the weight of the liquid has no component in the
direction of flow.

u1 u2
h2
h1

FIGURE 7.11 A simple positive surface in a channel with flat bed.

Uniform flow of velocity u1 and depth h1 , as shown in Figure 7.11, is


disturbed by, say, closing a gate downstream so that a positive surge travels
upstream, with constant velocity C, relative to the bed of the channel. A
short distance downstream of the wave the flow has again become uniform
with velocity u2 and depth h2 , as illustrated in Figure 7.11.
The change of velocity from u1 to u2 caused by the passage of the wave is
the result of a net force on the fluid, the magnitude of which is given by the
momentum equation. To apply the steady-flow momentum equation, let us
assume the coordinate axes to move with the wave. The wave then appears
stationary, conditions at any point fixed with respect to those axes do not
change with time, and the velocities are as shown in Figure 7.12.
The net force acting on the fluid in the control volume indicated is the
difference between the horizontal thrusts at sections 1 and 2. These sections
are sufficiently near each other for friction at the boundaries of the fluid to be
negligible. If the streamlines at these two sections are substantially straight
and parallel then the variation of pressure is hydrostatic and the total thrust

ISTUDY
406 CHAPTER 7. FLOW WITH FREE SURFACE

Control
volume

u1 + C u2 + C
h2
h1

1 2

FIGURE 7.12 Steady flow velocity and height at sections 1 and 2


of the wave.

on a vertical plane divided by the width of the channel is therefore,


( )
h ρgh2
ρg h=
2 2
For sufficiently uniform velocity at cross-sections 1 and 2, the momen-
tum correction factor is nearly unity and the steady-flow momentum equation
yields
ρgh21 ρgh22
Net force on fluid in CV = −
2 2
= Rate of increase of momentum

or
Net force on fluid in CV = ρQ (u2 − u1 ) (7.11)
By continuity,
Q = (u1 + C)h1 = (u2 + C)h2
Therefore,
h1
u2 = (u1 + C) −C (7.12)
h2
Substituting for Q and u2 into Equation (7.11), we get
[ ]
ρg ( 2 ) h1
h1 − h2 = ρ(u1 + C)h1 (u1 + C)
2
− C − u1
2 h2
h1
= ρ(u1 + C)2 (h1 − h2 )
h2
This simplifies to
( )1/2
√ 1 + h2 /h1
u1 + C = gh2 (7.13)
2
For waves of small height, that is, h2 ≈ h1 ≈ h, Equation (7.13) reduces to

u1 + C = gh (7.13a)

ISTUDY
7.5. WAVES AND SURGES IN OPEN CHANNELS 407

Equation (7.13a) shows that for waves of small √ height, the velocity of the wave
relative to the undisturbed liquid ahead of it is gh. Note that, no restriction
was imposed on the relative values of u1 and u2 : either may be zero to even
negative, but the analysis is still valid. This analysis applies only to waves
propagated in rectangular channels. For a channel of any other shape it √ may
be shown that the velocity of propagation of a small surface wave is gh
relative to the undisturbed liquid, where h is the mean depth, defined, with
reference to Figure 7.13, as
Area of cross-section A
=
Width of liquid surface B

Cross-sectional
area = A

FIGURE 7.13 Flow in a channel of arbitrary cross-section.



Note that, though gh represents the velocity of propagation of a very small
surge wave, a larger positive wave will be propagated with a higher velocity, as
shown in Figure 7.12. Also, the height of the wave does not remain constant
over appreciable distance; frictional effects, which in the above analysis were
justifiably assumed negligible in the short distance between sections 1 and 2,
gradually reduce the height of the wave.
For a positive wave which is stable, the front of the wave will be more or less
vertical, If the wave originally had a sloping front, as shown in
Figure 7.14, it could be regarded as the superposition of a number of waves
of smaller height.

FIGURE 7.14 A free wave with a sloping front.



The velocity of propagation of smaller waves is given by gh, therefore,
the velocity of the uppermost elements would be somewhat greater than that
of the lower ones. Thus the sloping front of the waves would tend to become
vertical. However, the opposite is true for a negative wave, illustrated in
Figure 7.15. Since the upper elements of the wave move rapidly, the wave
flattens and the wave front soon degenerates into a train of tiny wavelets.
Small disturbances in a liquid may cause waves of the shape shown in
Figure 7.16. This sort of wave may simply be regarded as a positive wave

ISTUDY
408 CHAPTER 7. FLOW WITH FREE SURFACE

FIGURE 7.15 A negative wave.

followed by a negative one√and, provided that it is only of small height, its


velocity of propagation is gh, as for the small surge wave.

FIGURE 7.16 Wave caused by a small disturbance.

In the lab experimental studies with small scale models, another type of
wave termed capillary wave, formed due to the influence of surface tension is
of significance. In general, the velocity of propagation of any surface wave is
governed by both gravity forces and surface tension forces. However, for single
surge wave in a large open channel of the size concerned in civil engineering
projects, the effect of surface tension is negligible.
For waves on the surface of deep water, such as the sea, different consider-
ations apply. This is mainly due to the assumption of a uniform distribution
of velocity over the entire depth is not valid for large depths. Furthermore,
ocean waves appear as a succession or train of waves in which waves follow
one another closely, so an individual wave cannot be considered separately
from those next to it.

EXAMPLE 7.3 Find the downstream height of a hydraulic jump occurring


on a level bed when the upstream depth and velocity are 1 m and 10 m/s,
respectively.

Solution Given, h1 = 1 m and u1 = 10 m/s.

By Equation (7.13) we have


( )1/2
√ 1 + h2 /h1
u1 = gh2
2

where h2 is the height of the hydraulic jump.

ISTUDY
7.5. WAVES AND SURGES IN OPEN CHANNELS 409

Therefore,
√ √ √ 1/2
2u1 = g h2 (1 + h2 /h1 )

2 × 10 ( )1/2
√ = h2 + h22
9.81
( )1/2
4.5 = h2 + h22
h22 + h2 = 4.52
h22 + h2 − 20.25 = 0

−1 ±1 + 4 × 20.25
h2 =
2
−1 ± 9.055
h2 =
2
= 4.028 m

Note that, the negative value of h2 is not practical.


EXAMPLE 7.4 A rectangular channel of width 3 m and depth 2 m,
discharging 18 m3 /s, suddenly has the discharge reduced to 12 m3 /s at the
downstream end. Determine the height and speed of the surge wave caused
by this reduction in discharge rate.

Solution Let subscripts 1 and 2 refer to the initial state and the state at
the downstream end.

Given, b = 3 m, h1 = 2 m, Q1 = 18 m3 /s and Q2 = 18 m3 /s.

Therefore,
Q1
u1 =
bh1
18
=
3×2
= 3 m/s

Q2
u2 h2 =
b
12
=
3
= 4 m2 /s

By Equation (7.11),
(u1 + C)h1 = (u2 + C)h2

ISTUDY
410 CHAPTER 7. FLOW WITH FREE SURFACE

where C is the speed of the surge wave. Therefore,

u1 h1 + Ch1 = u2 h2 + Ch2
3 × 2 + Ch1 = 4 + Ch2
6 − 4 = C(h2 − h1 )
2 = C(h2 − 2)
2
C=
h2 − 2

By Equation (7.13),
( )1/2
√ 1 + h2 /h1
u1 + C = gh2
2
( )
2 1 + h2 /h1
(u1 + C) = gh2
2
9.81
= × (h1 + h2 )h2
2×2
( )2
2
3+ = 2.45 × (h1 + h2 )h2
h2 − 2

Solving for h2 , we get


h2 = 2.75 m

The velocity u2 and speed of the surge wave, C, are

4
u2 =
h2
4
=
2.75
= 1.45 m/s

2
C=
h2 − 2
2
=
2.75 − 2
= 2.67 m/s

ISTUDY
7.6. SPECIFIC ENERGY AND ALTERNATIVE DEPTH OF FLOW 411

7.6 SPECIFIC ENERGY AND


ALTERNATIVE DEPTH OF FLOW
We know that the total mechanical energy divided by weight of liquid, termed
the total head is given by
p u2
+ +z
ρg 2g
where z is the height of the point considered above arbitrary horizontal datum
plane. If the channel shape is small and the streamlines are straight and
parallel,( so that) the variation of pressure with depth is hydrostatic then the
p
sum of ρg + z is equivalent to the height of the free surface above the same
datum. If, at a particular position, the datum level coincides with the bed
of the
( channel,) then the local value of the energy divided by weight is given
u2
by h + 2g , where h is the depth of flow at that position. The quantity
( 2
)
h + u2g is termed specific energy e or specific head. Note that the specific
energy is defined assuming the bed of the channel as horizontal. But the bed
of the channel may not be precisely horizontal, then the specific energy should
be regarded as a local parameter, applied over a short length of the channel
in which any change of bed level is negligible.
The mean velocity is
Q
u=
A
where A is the cross-sectional area of the stream at the position considered
and Q is the volume flow rate. Therefore, the specific energy can be expressed
as ( )2
1 Q
e=h+ (7.14)
2g A
Now, let us consider a wide channel of rectangular cross-section to sim-
plify the analysis. If the width of the rectangular section is b, then from
Equation (7.14)
( )2
1 Q
e=h+
2g bh
or
1 ( q )2
e=h+ (7.15)
2g h
where q = Q/b. Equation (7.15) relates the specific energy e, the depth h and
the discharge per unit width q.
For constant q the specific energy variation with depth will be as shown
in Figure 7.17.
For constant specific energy e, the flow discharge per unit depth q will
vary with depth as shown in Figure 7.18.
The plot of e versus h, shown in Figure 7.17, is called specific-energy
diagram. For q constant, a small value of depth h corresponds to a high

ISTUDY
412 CHAPTER 7. FLOW WITH FREE SURFACE

Tranquil

Rapid
emin
hc

FIGURE 7.17 Specific energy variation with depth.

Tranquil

Rapid

hc
qmax

q
FIGURE 7.18 Discharge variation with depth.

velocity and thus, as h tends to zero, u2 /2g tends to infinity and so also does
e. Hence the specific energy curve is asymptotic to e-axis. Conversely, as h
increases, the velocity becomes smaller, the u2 /2g term becomes insignificant
compared to h, and e tends to h. The depth at which e minimum occurs is
known as the critical depth hc .
For all values of e other than minimum there are two possible values of h,
one greater and one less than hc (although Equation (7.15) is a cubic in h,
the third root is always negative and therefore physically impossible). These
two values of h are known as alternative depths.
The critical depth may be obtained by differentiating Equation (7.15) with
respect to h and equating to zero. Differentiating Equation (7.15), we have
( )
∂e q2 2
=1+ − 3
∂h 2g h
Equating this to zero, we get
( )1/3
q2
hc = (7.16)
g

ISTUDY
7.6. SPECIFIC ENERGY AND ALTERNATIVE DEPTH OF FLOW 413

The corresponding minimum value of e becomes

gh3c
emin = hc +
2gh2c

that is,
3
emin = hc (7.17)
2
It is essential to note that Equations (7.16) and (7.17) are valid only for
channels of rectangular cross-section.
The situation in which the specific energy is held constant while h and q
vary, as shown in Figure 7.18, is of interest in many applications. As shown
in Figure 7.18, this curve shows that q reaches a maximum for a particular
value of h. Equation (7.15) may be arranged to give

q 2 = 2gh2 (e − h)

Differentiating with respect to h, we have

∂q
2q = 2g(2eh − 3h2 )
∂h
∂q
Now, = 0 when
∂h
2
h= e (7.18)
3
Note that this is identical to Equation (7.17), therefore, it may be said that
at the critical depth the discharge is a maximum for a given specific energy,
or that the specific energy is a maximum for a given discharge. Thus, if in a
particular channel the discharge is the maximum obtainable then somewhere
along its length the conditions must be critical.
We know that
Q q
u= =
bh h
Therefore, the velocity corresponding to the critical depth may be obtained
from Equation (7.16), as

q gh3c √
uc = = = ghc (7.19)
hc hc
velocity uc is known as the critical velocity. It is important to note that the
velocity uc has no connection with the critical velocity at which turbulent flow
becomes laminar.
For uniform velocity the specific energy at a particular section, by Equa-
tion (7.14), is
Q2
e=h+
2gA2

ISTUDY
414 CHAPTER 7. FLOW WITH FREE SURFACE

Differentiating with respect to h, we have


( )
∂e Q2 2 ∂A
=1+ − 3
∂h 2g A ∂h
Let us consider free-surface flow through a passage, as shown in Figure 7.19.

δh

h Area of
cross-section = A

FIGURE 7.19 Free-surface flow through a channel.

From Figure 7.19, it is seen that the small increase of area δA, which corre-
sponds to a small increase δh in the depth, is given by Bδh. Hence, as δh → 0,
δA/δh → dA/dh = B. Therefore,

∂e Q2
=1− B (7.20)
∂h gA3

which is zero when Q2 = gA3 /B, that is, when

gA Q2
= 2 = u2
B A
If A/B is regarded as the mean depth of the section and is represented by h
then √ √
gA
uc = = gh (7.20a)
B
Note that the mean depth h is different from the hydraulic mean depth A/P .
The important significance of the critical conditions is that they separate
two distinct types of flows, namely, in which the velocity is less than the
critical value and that √ in which the velocity exceeds the critical value. The
critical velocity uc = gh corresponds to the velocity of propagation (relative
to the undisturbed liquid) of a small surface wave in shallow liquid.
When the flow velocity is less than uc , it is possible for a small surface wave
to propagate upstream as well as downstream, as the sound wave propagates
in a subsonic gas. Any small disturbance to the flow can cause a formation of
small surface wave, which can carry the information about the disturbance.
That is, for flow velocity less than the critical velocity, any disturbance down-
stream can be propagated to the liquid upstream also. This flow situation is
identical to subsonic flow in gases. When the flow velocity is greater than the
critical velocity, the liquid travels downstream faster than a small wave can be

ISTUDY
7.6. SPECIFIC ENERGY AND ALTERNATIVE DEPTH OF FLOW 415

propagated upstream. Due to this information about events downstream can


not be transmitted to the liquid upstream, and so the behavior of the liquid
is not controlled by the downstream conditions. This flow process is identical
to that of a supersonic gaseous flow.
When the flow velocity is equal to critical velocity, a relatively large change
of depth causes only a small change of specific energy, as seen in Figure 7.17.
Consequently small undulations on the surface are easily formed under these
conditions. Further, a small wave that attempts to travel upstream makes no
progress, and so is known as a steady wave.
From the above discussions, it is evident that the behavior of liquid flowing
in an open channel depends on the relative magnitude of the velocity compared
to the critical velocity. The ratio of the
√ mean flow velocity to the velocity of
propagation of a small disturbance u/ gh is of the same form as the Froude
number. If the mean depth h is used as the characteristic length, and the
mean velocity is used as the characteristic velocity, then

u
√ = Fr
gh

Thus, for critical condition



u= gh and Fr = 1

That is, critical Froude number is unity.


In a long open channel with suitable slope, uniform flow critical velocity
may be produced. The value of the slope for which critical uniform flow is
achieved is known as the critical slope. Using the Chézy formula, we obtain
the critical velocity as
√ √
uc = gh = C m i

In uniform flow the energy gradient i and the slope of the bed s are equal and
therefore the critical slope sc is defined by


gh = C msc

A slope less than the critical slope sc is known as mild slope, and a slope
greater than the critical slope is termed steep slope.
Flow with velocity less than the critical velocity is referred to as tran-
quil, flow with velocity greater than the critical velocity is called rapid or
shooting. The tranquil and rapid flows are also referred to as subcritical and
supercritical flows, respectively. It is important to note that sub-critical veloc-
ity corresponds to a depth greater than the critical depth, and supercritical
velocity corresponds to a depth smaller than the critical depth.
When the velocity is uniform over the cross-section, the relations governing
the critical conditions of flow involve only the volume rate of flow and the

ISTUDY
416 CHAPTER 7. FLOW WITH FREE SURFACE

shape of the cross-section. They do not depend on the roughness of the


boundaries.

EXAMPLE 7.5 A flood of volume flow rate 400 m3 /s flows through a


river. At a particular location if the width of the river is 25 m, determine the
critical depth at that location.
Solution Given, Q = 400 m3 /s, b = 25 m.
By Equation (7.16),
q 2 = gh3c
where
Q
q=
b
400
=
25
= 16
Therefore, the critical depth becomes
( 2 )1/3
q
hc =
g
( 2 )1/3
16
=
9.81

= 2.97 m
EXAMPLE 7.6 Water flows through a rectangular channel with a velocity
of 1.5 m/s and depth 1.2 m. Determine (a) the specific energy of the flow, (b)
the critical depth and (c) the maximum discharge under the critical condition,
if the channel is 3 m wide.

Solution Given, u = 1.5 m/s, h = 1.2 m.

(a) Flow rate per unit width is


Q
q=
b
Au
=
b
(b h) × u
=
b
=h×u
= 1.2 × 1.5
= 1.8 m2 /s

ISTUDY
7.6. SPECIFIC ENERGY AND ALTERNATIVE DEPTH OF FLOW 417

The specific energy, by Equation (7.15), is


1 ( q )2
e=h+
2g h
( )2
1 1.8
= 1.2 + ×
2 × 9.81 1.2
= 1.2 + 0.1147

= 1.315 m

(b) The critical, by Equation (7.18), is


2
hc = e
3
2
= × 1.315
3
= 0.877 m

(c) Given, b = 3 m. The critical velocity, by Equation (7.19), is



uc = ghc

= 9.81 × 0.877
= 2.93 m/s

Therefore, the maximum discharge is

Q = Auc

= (hc × b) × uc

= (0.877 × 3) × 2.93

= 7.71 m3 /s

EXAMPLE 7.7 The triangle channel shown in Figure E7.7 has a flow rate
of 16 m3 /s. Determine (a) the critical depth, (b) critical velocity and (c) the
energy slope, i, if n = 0.018.
Solution Given, Q = 16 m3 /s, n = 0.018.
(a) By Equation (7.14), the specific energy is

Q2
e=h+
2gA2

ISTUDY
418 CHAPTER 7. FLOW WITH FREE SURFACE

y cot 50◦

y cosec 50◦
y

50◦

Figure E7.7 A triangular channel.

Differentiating with respect to h,


( )
de d Q2
=1+
dh dh 2gA2
de
For critical state, dh= 0, therefore,
( 2 )
d Q
= −1
dh 2gA2
dA Q2
−2A−3 = −1
dh 2g
dA gA3
= 2
dh Q
But
dA = b0 dh
where b0 is the width at the free surface. Therefore,
gA3
b0 =
Q2
This gives,
( )1/3
b0 Q2
Ac =
g
But Ac = h2c cot 50◦ and b0 = 2hc cot 50◦ . Therefore,
2hc cot 50◦ Q2
(h2c cot 50◦ )3 =
g
2Q2
h5c =
g(cot 50◦ )2
2 × 162
=
9.81 × (cot 50◦ )2
= 74.127
hc = (74.127)1/5

= 2.37 m

ISTUDY
7.6. SPECIFIC ENERGY AND ALTERNATIVE DEPTH OF FLOW 419

(b) The critical velocity is

Q
uc =
Ac
Q
=
h2c cot 50◦
16
=
2.372 × cot 50◦
16
=
4.713
= 3.39 m/s

(c) The cross-sectional area, A and the perimeter, P , are

A = h2c cot 50◦

P = 2hc cosec 50◦

Therefore, the hydraulic mean depth, m, become


A
m=
P
h2c cot 50◦
=
2hc cosec 50◦
hc
= cos 50◦
2
2.37
= × cos 50◦
2
= 0.762 m

By Equation (7.8), √
Q = AC m i
where C is the Chézy’s coefficient, given by

m1/6
C=
n
0.7621/6
=
0.018
= 53 m1/2 /s

Therefore, √ √
Q = AC m i = AC m i

ISTUDY
420 CHAPTER 7. FLOW WITH FREE SURFACE

This gives

u2c
i=
C 2m
3.392
=
532 × 0.762
= 0.00537

7.6.1 Dimensionless Forms of Specific-Energy Curves


The specific-energy curve shown in Figure 7.17 is for a particular value of
flow rate q. A general curve, applicable to any value of q may be obtained
by Equation (7.15) in dimensionless form. Let us consider the cross-section
of the channel as rectangular. Dividing Equation (7.15) by the critical depth
hc , we have
e h 1 ( q )2 1
= +
hc hc 2g h hc
Substituting q 2 = gh3c , from Equation (7.16), this becomes

( )2
e h 1 hc
= + (7.21)
hc hc 2 h

This is the dimensionless form of specific-energy relation. Figure 7.20 shows


the curve representing Equation (7.21). From this we may determine the two
possible values of h/hc (one greater than unity and one less than unity). These
depths corresponds to points A and B in Figure 7.20.

h
hc

A
Tranquil

1.0
Rapid
B

1.5 e
hc
FIGURE 7.20 Dimensionless forms of specific-energy curves.

The relation between the rate of flow and the depth for a given specific
energy, shown in Figure 7.18, may also be re-plotted in dimensionless form.

ISTUDY
7.6. SPECIFIC ENERGY AND ALTERNATIVE DEPTH OF FLOW 421

2
Dividing Equation (7.15) by qmax = gh3c , we have

e h 1 ( q )2 1
= +
gh3c gh3c 2g h 2
qmax
( )2 ( )3
2e h h q2
=2 + 2
hc hc hc qmax

For a given e, from Equation (7.17), 2e = 3hc , therefore,

( )2 ( )2 ( )3
q h h
=3 −2 (7.22)
qmax hc hc

EXAMPLE 7.8 A trapezoidal channel with bottom 3 m and side slopes 1


horizontal and 2 vertical carries 10 m3 /s. Determine the critical depth.

Solution Given, b = 3 m and Q = 10 m3 /s. The trapezoidal is as shown in


Figure E7.8.

h/2 h/2
2
h 1

3m
Figure E7.8 A trapezoidal channel.

The cross-sectional area, A, is

A = bh + h2 cot α
h2
= 3h +
2

The surface width B is

h
B =b+2
2
=3+h

For critical condition, by Equation (7.20),

Q2 B
=1
gA3

ISTUDY
422 CHAPTER 7. FLOW WITH FREE SURFACE

Therefore,

Q2 B
A3 =
g
102
= × (3 + h)
9.81
( )3
h2
3h + = 10.19 × (3 + h)
2
( )3
h2
3h + = 30.57 + 10.19h
2

Solving, we get the critical depth as

hc = 0.984

EXAMPLE 7.9 Water flows under a sluice gate. The head and velocity
ahead of the gate are 1.5 m and 0.2 m/s, respectively. (a) Find the depth of
flow and velocity downstream of the gate, at this flow rate. (b) Also, find the
maximum possible flow rate.

Solution Let subscripts 1 and 2 refer to the state at upstream and down-
stream of the sluice gate and subscript 0 refers to the total or stagnation
state.

(a) Given, h1 = 1.5 m, u1 = 0.2 m/s.

The total head is


p0 p1 u2
= + 1
ρg ρg 2g
0.22
= 1.5 +
2 × 9.81
= 1.502 m

The specific energy at ahead of the gate is

u21
e1 = h1 +
2g
= 1.502 m

By Equation (7.15)

1 ( q )2
e=h+
2g h

ISTUDY
7.6. SPECIFIC ENERGY AND ALTERNATIVE DEPTH OF FLOW 423

The flow rate per unit width is


q = h1 u1
= 1.5 × 0.2
= 0.3 m2 /s
The specific energy at behind the gate is
u22
e2 = h2 +
2g
( )2
1 q
= h2 +
2g h2
( )2
1 0.3
= h2 +
2 × 9.81 h2
0.0046
1.502 = h2 +
h22

This gives,
h32 − 1.502h22 + 0.0046 = 0
Solving, we get the depth behind the gate as h2 = 0.056 .

Therefore, the velocity behind the gate is


q
u2 =
h2
0.3
=
0.056
= 5.36 m/s

(b) The height corresponding to a discharge maximum for a given specific


energy, by Equation (7.18), is
2
h= e
3
2
= × 1.502
3
= 1m
The velocity, by Equation (7.19), is

u = gh

= 9.81 × 1
= 3.13 m/s

ISTUDY
424 CHAPTER 7. FLOW WITH FREE SURFACE

Therefore, the maximum flow rate is

qmax = u h
= 3.13 × 1

= 3.13 m2 /s

7.7 THE HYDRAULIC JUMP


A stationary surge wave, through which the depth of flow increases is known
as hydraulic jump. The essential features of a hydraulic jump can be discerned
from Equation (7.13), which relates the velocity of a positive wave traveling
upstream in a horizontal channel of rectangular cross-section, and the depths
of flow before and after the wave.
Substituting c = 0 in Equation (7.13), we get
( )1/2
√ 1 + h2 /h1
u1 = gh2
2

This can be arranged, by dividing and multiplying the first group on right-
hand side with h1 , as
√ ( )1/2
√ h2 1 + h2 /h1
u1 = gh1 (7.23)
h1 2

This gives
[( )( )]1/2
u1 h2 1 + h2 /h1
Fr1 = √ =
gh1 h1 2

or [ ]1/2
2
(h2 /h1 ) + (h2 /h1 )
Fr1 = (7.23a)
2

where Fr1 is the Froude number of the flow ahead of the hydraulic jump.
Since h2 > h1 , this expression is greater than unity. In other words, the
Froude number ahead of the hydraulic jump is always greater than unity and
the flow is rapid. That is, hydraulic jump will occur only in rapid flow.
Now, from continuity relation,

u1 bh1 = u2 bh2

or
u1 h1
u2 =
h2

ISTUDY
7.7. THE HYDRAULIC JUMP 425

Substituting Equation (7.23), we have


( )1/2 ( )1/2 ( )
√ h2 1 + h2 /h1 h2
u2 = gh1
h1 2 h1
( )1/2 ( )
√ 1 + h2 /h1 h2
= gh2
2 h1
or [ ]1/2
u2 (h1 /h2 )2 + (h1 /h2 )
Fr2 = √ = (7.23b)
gh2 2
where Fr2 is the Froude number behind the hydraulic jump. Since h2 > h1 ,
Fr2 is less than 1. Hence the flow behind a hydraulic jump is always tranquil.
A hydraulic jump is essentially an abrupt change from rapid to tranquil
flow. The depth of the liquid is less than the critical depth, hc , before the
jump and greater than hc after it. The rapid flow upstream of the hydraulic
jump may rise in a number of ways. For example, jump can be created by
releasing water at high velocity through a sluice gate into a channel. But the
rapid flow thus produced cannot persist indefinitely in a channel where the
slope of the bed is insufficient to sustain it. For a given flow rate, the depth
corresponding to uniform flow in the channel is determined by the roughness
of the boundaries and the slope of the bed. For a mild slope this depth is
greater than the critical depth, that is, uniform flow has to be tranquil.
A gradual transition from rapid to tranquil flow is not possible. As the
depth of rapid flow increases, the lower limit of the specific-energy diagram
(Figure 7.20) is followed from right to left, that is, the specific-energy de-
creases. If the increase has to continue to the critical value, an increase in
specific-energy would be required for a further increase of depth to the value
corresponding to uniform flow diagram. However, in such circumstances an
increase in specific energy is impossible. In uniform flow the specific energy
remains constant and the total energy decreases at a rate exactly correspond-
ing to the slope of the bed. For any depth less than that for uniform flow
the velocity is higher, due to friction consumes energy at a greater rate than
the loss of gravitational energy. That is, the energy gradient is greater than
the slope of the channel bed, therefore, as shown in Figure 7.21, the specific
energy must decrease.

Energy
line

Free surface

E1 E2 < E1

FIGURE 7.21 Energy gradient in a channel.

ISTUDY
426 CHAPTER 7. FLOW WITH FREE SURFACE

The depth of the tranquil flow after the jump is determined by the re-
sistance offered to the flow, either by some obstruction such as a weir or by
the friction forces in a long channel. The jump causes eddy formation and
turbulence. Thus there is an appreciable loss of mechanical energy, and both
the total and specific energy behind the jump are less than those ahead of the
jump.
For determining the change in depth at a hydraulic jump, let us consider
a channel of uniform rectangular cross-section. Though hydraulic jump can
be formed in a channel of any shape of cross-section, rectangular cross-section
is considered here to simplify the mathematical complexity.
Equation (7.23) can be arranged as
2u21 h1
h22 + h1 h2 − =0 (7.24)
g
Now, substituting u1 = q/h1 , where q is the discharge divided by width, we
get √( )
h1 h21 2q 2
h2 = − ± +
2 4 gh1
There are two solutions for h2 . But the solution with negative sign for the
second term on the right-hand-side is physically impossible, since h2 cannot
be negative. Hence,
√( )
h1 h21 2q 2
h2 = − + + (7.25)
2 4 gh1

Note that, Equation (7.24) is symmetrical in respect of h1 and h2 , therefore


a similar solution for h1 in terms of h2 may be obtained by interchanging the
suffixes. These depths h1 and h2 ahead of and behind a hydraulic jump are
termed conjugate depths for the jump.
The loss of mechanical energy associated with the jump may be determined
from the energy equation. The head loss across a jump hj can be expressed
as
( ) ( )
u21 u22
hj = h1 + − h2 +
2g 2g
1 ( 2 )
= (h1 − h2 ) + u − u22
2g 1
But the volume flow per unit width is
q = u1 h1 = u2 h2
Thus
( )
q2 1 1
hj = (h1 − h2 ) + − 2
2g h21 h2

ISTUDY
7.7. THE HYDRAULIC JUMP 427

From Equation (7.25),


q2 h1 h22 + h21 h2
=
2g 4
Substituting this, we have the head loss across a jump as
( )( )
h1 h22 + h21 h2 1 1
hj = (h1 − h2 ) + −
4 h21 h22
or
3
(h2 − h1 )
hj = (7.26)
4h1 h2
A typical hydraulic jump and the head loss across it are illustrated in Figure
7.22.
The head loss, hj , across a jump is primarily due to the turbulence in the
wave because friction at the boundaries makes a negligible contribution to the
energy loss. The frictional forces in the wave are in the form of innumerable
pairs of action and reaction and so, by Newton’s third law, annul each other
in the net force on the control volume considered in deriving Equations (7.13)
and (7.23).
h

Tranquil flow
u2
Rapid flow hj (head lost in jump)
h2
hc u1
u2
E =h+ 2g
h1
(a) (b)
FIGURE 7.22 A hydraulic jump and the head loss across it.

The energy dissipation across a jump is always finite. Indeed, the hydraulic
jump is an effective means of reducing unwanted energy in a stream. For
example, if water from steep spillway is fed into a channel, severe scouring of
the bed may occur if the rapid flow is allowed to continue. A hydraulic jump
arranged to occur at the foot of the spillway will dissipate much of the surplus
energy, and the stream may then be safely discharged as tranquil flow. Due
to the dissipation of energy, the temperature of the liquid is raised by a small
amount.
The position at which the jump occurs is always such that the momentum
relation is satisfied: the value of h2 is determined by conditions downstream
of the jump, and the rapid flow continues until h1 has reached the value that
fits Equation (7.25).
From Equation (7.25), it is seen that a hydraulic jump is possible only from
rapid to tranquil and not vice versa. This is because, if h2 were less than h1
then h1 would be negative, that is, there would be a gain of energy, and this

ISTUDY
428 CHAPTER 7. FLOW WITH FREE SURFACE

kind of process will violate the increase of entropy principle (the Second law of
thermodynamics) and hence such a process is physically impossible. In other
words, it can be stated that flow from tranquil to rapid across a hydraulic
jump will result is the violation of increase of entropy principle, hence such
a flow process is impossible, because no process which violates the Second
law of thermodynamics is physically feasible. Thus the flow process across a
hydraulic jump is an irreversible process.
Equation (7.25) can be expressed in dimensionless form as
√( )
h2 1 1 2u21
=− + +
h1 2 4 gh1

or √
h2 1 1 2
=− + + 2 (Fr1 ) (7.27)
h1 2 4
This shows that, the ratio of the conjugate depths h2 /h1 is a function of the
initial Froude number only and that the larger the initial Froude number the
larger the ratio of the depths. For Fr = 1, h2 /h1 = 1 and the jump becomes
a standing wave of infinitesimal height.
Jumps
√ causing h2 /h1 < 1 are termed small jumps. This corresponds to
Fr1 < 3, for rectangular sections. For small jumps the surface does not rise
abruptly but passes through a series of undulations gradually diminishing in
size. Such a jump is known as an undular jump. Hydraulic jumps associated
with some specific range of Fr1 are illustrated in Figure 7.23.

FIGURE 7.23 Illustration of undular oscillations, weak, steady and


strong jumps.

ISTUDY
7.7. THE HYDRAULIC JUMP 429

For larger values of h2 /h1 and Fr1 , the jump is termed direct and across a
direct jump the surface would rise abruptly.
At the wave front the fluid moves like a roller. For example, an ocean wave
about to break on the shore resembles a roller. This results from the upper
layer of the wave tending to spread over the oncoming rapid stream. The
frictional drag of the rapid stream penetrating underneath, and the transfer
of momentum from the lower layers by eddies, however, prevent the upper
layers moving upstream.
For values of h2 /h1 , in the range from 3.0 to 5.5, oscillations may be caused
that result in the formation of irregular waves which travel downstream. For
h2 /h1 from 5.5 to 12 the jump is stable and a good dissipator of energy. The
length of the jump; defined as the horizontal distance between the front of the
jump and a point just downstream of the rollers, is usually of the order of five
times its height. At this stage, it is important to note that the above discussed
ranges of h2 /h1 refer only to channels of rectangular sections. For channels
of other cross-sections the jump becomes complicated owing to cross-currents.

EXAMPLE 7.10 A hydraulic jump occurs downstream of a sluice gate. If


the flow depth ahead of the jump is 1.5 m and Froude number downstream of
the jump is 0.29, (a) determine the flow depth after the jump, (b) the Froude
number ahead of the jump. (c) If the gate is 15 m wide, determine the power
dissipated in the jump.

Solution Let subscripts 1 and 2, respectively, refer to the state upstream


and downstream of the jump.

Given, h1 = 1.5 m and Fr2 = 0.29.

By Equation (7.23b),
[ ]1/2
(h1 /h2 )2 + (h1 /h2 )
Fr2 =
2

Therefore,
[ ]1/2
(h1 /h2 )2 + (h1 /h2 )
= 0.29
2
(h1 /h2 )2 + (h1 /h2 ) = 0.292 × 2
(h1 /h2 )2 + (h1 /h2 ) − 0.1682 = 0

h1 −1 ± 1 + 4 × 0.1682
=
h2 2
−1 ± 1.293
=
2
= 0.1465

ISTUDY
430 CHAPTER 7. FLOW WITH FREE SURFACE

h1
h2 =
0.1465
1.5
=
0.1465
= 10.24 m

(b) We have
h2 10.24
=
h1 1.5
= 6.83
By Equation (7.23a),
[ ]1/2
(h2 /h1 )2 + (h2 /h1 )
Fr1 =
2
(6.832 + 6.83)
=
2
= 5.17

(c) The velocity upstream of the jump is



u1 = Fr1 × gh1

= 5.17 × 9.81 × 1.5
= 19.83 m/s
The head loss across the jump, by Equation (7.26), is
(h2 − h1 )3
hj =
4h1 h2
(10.24 − 1.5)3
=
4 × 1.5 × 10.24
= 10.87 m
The power dissipated is
Power = ρQhj
= ρ(Au1 )hj
= ρ([bh1 ]u1 )hj
= 103 × (15 × 1.5 × 19.83) × 10.87

= 4.85 MW

ISTUDY
7.7. THE HYDRAULIC JUMP 431

EXAMPLE 7.11 A jet, 1.3 m deep, with a total head 10 m, issues from a
sluice gate 6 m wide and then flows along a horizontal frictionless channel to
a narrow culvert. Determine the width of the culvert so that the sluice is not
drowned.

Solution Let subscripts 1 and 2, respectively, refer to the flow upstream


and downstream of the jump which is caused by the narrow culvert.

Given, h1 = 1.3 m, h0 = 10 m and b1 = 6 m.

Therefore, the velocity and Froude number upstream of the jump are,

u1 = 2g(h0 − h1 )

= 2 × 9.81 × (10 − 1.3)
= 13.06 m/s
u1
Fr1 = √
gh1
13.06
=√
9.81 × 1.3
= 3.66

By Equation (7.27),

h2 1 1 2
=− + + 2 (Fr1 )
h1 2 4

1 1
=− + + 2 × 3.662
2 4
= 4.7
h2 = 4.7h1
= 4.67 × 1.3
= 6.11 m

Also,

A1 = b1 h1
= 6 × 1.3
= 7.8 m2
P1 = b1 + 2h1
= 6 + 2 × 1.3
= 8.6 m

ISTUDY
432 CHAPTER 7. FLOW WITH FREE SURFACE

The hydraulic mean depth is


A1
m1 =
P1
7.8
=
8.6
= 0.907

The cross-section area and perimeter at 2 are

A2 = b2 h2
= 6.11b2
P = b2 + 2h2
= b2 + 2 × 6.11
= 12.22 + b2

Therefore, the hydraulic mean depth at 2 is


A2
m2 =
P2
6.11b2
=
12.22 + b2
But m2 should be equal to m1 , for the frictionless flow on a horizontal channel,
therefore,
6.11b2
= 0.907
12.22 + b2
6.11b2 = 0.907(12.22 + b2 )
b2 (6.11 − 0.907) = 0.907 × 12.22
11.08
b2 =
5.203
= 2.13 m

7.7.1 Force on an Obstacle in a Stream


When an obstacle such as a rectangular block is placed in a flow, as shown
in Figure 7.24, the obstacle would exert a force F on the liquid, as shown
in the figure. Assuming the bed as horizontal and the flow be of steady and
uniform velocity over the cross-section, and uniform depth across the width
and negligible boundary friction, the momentum equation can be written as
ρgh1 ρgh2
h1 − h2 − F = ρg(u2 − u1 ) (7.28)
2 2

ISTUDY
7.7. THE HYDRAULIC JUMP 433

for unit width of uniform rectangular channel. This equation is true whatever
the values of h1 and h2 in relation to the critical depth. It can be shown
that, for tranquil flow the effect of the applied force is to reduce the depth of
the stream (h2 < h1 ), although a limit is set at the critical depth hc because
of energy is then a maximum. A further increase in the obstructing force F
beyond the value giving h2 = hc merely raises the upstream level. For rapid
flow, the depth is increased by force F (h2 > h1 ). For a given flow, measuring
h1 and h2 on either side of the obstacle, the force F can be calculated using
Equation (7.28).

h2 u2
u1
h1 F
Weir

FIGURE 7.24 An obstacle in a steady uniform flow.

EXAMPLE 7.12 Find the height and Froude number downstream of a hy-
draulic jump occurring on a level bed, when the upstream depth and velocity
are 0.9 m and 10.5 m/s, respectively. Also, find the nature of this jump.

Solution Given, h1 = 0.9 m, u1 = 10.5 m/s.

By Equation (7.23), we have

( )1/2
√ 1 + h2 /h1
u1 = gh2
2

where h2 is the height downstream of the jump. Therefore,

( )
1 + h2 /h1
u21 = gh2
2
( )
h2 2u2
h2 1 + = 1
h1 g
2u21 h1
h2 (h1 + h2 ) =
g
2 × 10.52 × 0.9
h2 + 0.9h2 =
9.81
h2 + 0.9h2 = 20.23
h2 + 0.9h2 − 20.23 = 0

ISTUDY
434 CHAPTER 7. FLOW WITH FREE SURFACE

−0.9 ±0.92 + 4 × 20.23
h2 =
2
−0.9 ± 9.04
=
2
= 4.07 m

The Froude number downstream of the jump, By Equation (7.23b), is


[ ]1/2
(h1 /h2 )2 + (h1 /h2 )
Fr2 =
2
[ ]1/2
(0.9/4.07)2 + (0.9/4.07)
=
2

= 0.367

The Froude number upstream of the jump is


u1
Fr1 = √
gh1
10.5
=√
9.81 × 0.9
= 3.53

The Froude number upstream of the jump is in the range 2.5 - 4.5, therefore,
the jump is oscillating jump (Refer Figure 7.23).

Aliter: Depth downstream of the jump can also be calculated using Equation
(7.27), which relates h2 to h1 and Fr1 , as

h2 1 1
=− + + 2[Fr1 ]2
h1 2 4
Therefore,

h2 1 1
=− + + 2 × 3.532
h1 2 4

= −0.5 + 25.1718
= −0.5 + 5.017
= 4.517
h2 = 0.9 × 4.517

= 4.07

Note: The answers can be checked as follows.

ISTUDY
7.8. CRITICAL FLOW 435

The velocity downstream of the jump is



u2 = Fr2 × gh2

= 0.367 × 9.81 × 4.07

= 2.32 m/s

For volume conservation,


h1 u1 = h2 u2
Thus

0.9 × 10.5 = 4.07 × 2.32

9.45 = 9.44

This validates the results obtained.

7.8 CRITICAL FLOW


For flow in open channels the onset of critical flow condition imposes a lim-
itation on the discharge, as illustrated in 7.18. As we saw, critical condition
is experienced by the flow while changing from tranquil to rapid, as shown in
Figure 7.25.

hc

s < sc hc

s > sc

FIGURE 7.25 Illustration of flow transition from tranquil to rapid,


in a long channel.

A long channel of mild slope (that is, s < sc ) is connected to a long


channel of steep slope (that is, s > sc ). At a sufficiently large distance from
the junction the depth in each channel is the normal depth corresponding to
the particular slope and rate of flow; that is, in the channel of mild slope there
is uniform tranquil flow, and in the other channel there is uniform rapid flow.
But these trenches of the flow changes from uniform flow to non-uniform flow,
therefore at some position the depth must pass through the critical value.
This position is close to the junction of the slope. If both the channels have
the same constant slope the critical depth remains unchanged throughout,
as shown by the dashed line in Figure 7.25. When the change of slope is

ISTUDY
436 CHAPTER 7. FLOW WITH FREE SURFACE

abrupt, there is appreciable curvature of the streamlines near the junction.


The assumption of hydrostatic variation of pressure with depth (on which the
concept of specific energy is based) is then not justified and the specific energy
is only approximately√ valid. In these circumstances the section at which the
velocity is given by gh is not exactly at the junction of the two slopes, but
slightly upstream from it.
The discharge of a liquid from a long channel of steep slope to a long
channel of mild slope requires the flow to change from rapid to tranquil,
as illustrated in Figure 7.26. The rapid flow may persist for some distance
downstream of the junction, and the change to tranquil flow then takes place
abruptly at a hydraulic jump.

Hydraulic jump

hc

s > sc

s < sc
FIGURE 7.26 Transition from rapid to tranquil at a jump.

7.9 THE BROAD-CRESTED WEIR


Weir is an obstruction which extends across the full width of the stream.
In other words, a weir is a raised portion of the bed, which runs across the
complete width of the stream. Figure 7.27 illustrates a weir with a broad
horizontal crest, at a sufficient height above the channel bed for the cross-
sectional area of the flow approaching the weir to be large compared to the
cross-sectional area of the flow over the top of it.

Movable gate
Energy for all flows h
A
B
Free surface hA hB
C C
H = constant
h B
A Q
hc Weir

(a) (b)

FIGURE 7.27 A weir with a broad horizontal crest.

ISTUDY
7.9. THE BROAD-CRESTED WEIR 437

Let us assume that the approaching flow is tranquil. The upstream edge
of the weir is well rounded so that there is no undue formation of eddies
and thus there is no loss of energy. Further let us assume that there is no
obstruction downstream of the weir and the flow volume comes from the
reservoir is sufficiently large for the surface level upstream of the weir to be
constant.
Over the top of the weir the surface level fall to give a depth h there, as
shown in Figure 7.27. Moreover, for the channel of constant width considered,
if the crest is sufficiently broad (in the flow direction) and friction is negligible,
the liquid surface becomes parallel to the crest. For determining the depth h,
let us assume that downstream of the weir the flow is controlled by a movable
sluice gate. If, initially, the gate is completely closed, the liquid is stationary
and the surface level above the crest of the weir is the same as the level in the
reservoir. This level corresponds to the energy available and h represents the
specific energy for the liquid on the crest of the weir. This state is represented
by point A in Figure 7.27(a). If the gate is raised slightly to a position B, a
small rate of flow QB takes place. This state is shown as B in Figure 7.27(b).
Further raising of the gate results in the maximum flow rate Qc and since the
gate is then just clear of the surface of the liquid, no additional raising has
any effect on the flow rate. That is, the flow is critical over the crest of the
weir and h = hc . Even if it were possible to reduce h below the critical depth,
the plot of h verses Q in Figure 7.27(b) shows that the rate of flow would be
less than the critical flow rate Qc .
For a channel of rectangular cross-section, by Equation (7.16),
( )1/3
q2
hc =
g

where q is the discharge divided by the width (Q/b). Thus

q = g 1/2 h3/2
c

From this equation it is evident that the rate of flow could be calculated
simply from a measurement of hc . But it is not easy to directly measure hc
accurately. When critical flow occurs there are usually many ripples on the
surface and so the depth seldom has a steady value. For a rectangular section,
from Equation (7.18),
2
hc = e
3
Therefore,
( )3/2
2
q = g 1/2 e
3
that is,
( )3/2
1/2 2
Q=g b e (7.29)
3

ISTUDY
438 CHAPTER 7. FLOW WITH FREE SURFACE

The specific energy e in Equation (7.29) is that over the crest of the weir,
the crest itself being the datum. If the velocity u of the liquid upstream is so
small that u2 /(2g) is negligible, and if the friction at the approach to the weir
is also negligible, the specific energy for the weir is simply given by the H. It
is to be noted that H must be measured above the crest as datum. It is not
the full depth in the channel upstream. For a negligible velocity of approach,
Equation (7.29) becomes
( )3/2
2
Q = g 1/2 b H (7.30)
3

In the above analysis, friction has been ignored, but friction is always present
in some degree. The effect of friction over the crest, and the curvature of the
streamlines, is to reduce the discharge given by Equation (7.30). That is, the
discharge given by Equation (7.30) is ideal one, which does not account for the
losses. In the practical weirs will cause some losses due to their geometry, such
as edges, etc. therefore, the loss due to these aspects should by accounted for.
This is usually done by introducing discharge coefficient, Cd , in the expression
for Q, and expressed as
( )3/2
2
Q = Cd g 1/2 b H (7.30a)
3

If the crest of the weir is insufficiently broad in comparison to the head H,


the liquid surface may not become parallel to the crest before the downstream
edge is reached, as illustrated in Figure 7.28.

Weir

FIGURE 7.28 Parallel flow ahead of crest.

Although critical flow may still be obtained, it takes place at a section


where the streamlines have appreciable curvature, therefore the specific-energy
u2
e=h+
2g
does not strictly apply. In this situation Equation (7.29) underestimates the
actual rate of flow. Even when the crest is very broad, so that reasonably
parallel flow is obtained resulting in truly critical conditions, because of fric-
tion, the critical condition which occurs only at the downstream edge, where

ISTUDY
7.9. THE BROAD-CRESTED WEIR 439

the streamlines are curved. This causes an appreciable departure from the
value of discharge given by Equation (7.30). This is true especially for weirs
with cross-section different from the more or less rectangular one shown in
Figure 7.27(a). Therefore, for reliable determination of the rate of flow over
a broad-crested weir, it is essential to calibrate the flow.
Sometimes a weir is used where the velocity of the approaching flow is not
negligible. In this situation the height of the upstream surface level above the
2
crest of the weir may be less than the specific-energy (E = h − u2g ). However,
the measured H may be used as a first approximation to E, and a value
of Q calculated using this value of H. The approach velocity may then be
estimated from this value of Q and the cross-section area of the channel at the
section where the measurement of H is made. From this value of velocity (say,
2
u1 ) in terms of u21 /(2g) may be calculated, and then H + u2g may be taken as
an approximation to the specific-energy so that another, more accurate, value
of Q may be determined.
As a measuring device the broad-crested weir has the following advantages.

(i) It is simple to construct and has no sharp edge that can wear and thus
alter the discharge characteristics with time.
(ii) It does not cause any appreciable raising of the surface level upstream,
and the results are not affected by conditions downstream provided that
critical flow occurs over the crest.

In spite of these advantages, the limitation of the theoretical


analysis makes it unwise to use it, for any accurate measurement of flow
rate. Thus broad-base weir are suitable only for approximate measurement of
flow rate.

EXAMPLE 7.13 Water flows across a broad-crested weir in a 400 mm


wide rectangular channel. The depth of water just upstream of the weir is
70 mm and the crest of weir is 40 mm above the channel bed. Calculate the
fall of the surface level and the corresponding discharge, assuming that the
velocity of approach is negligible.

Solution Given, b = 400 mm, H = 70 − 40 = 30 mm.

The volume flow rate, by Equation (7.30), is


( )3/2
√ 2
Q = gb H
3
( )3/2
√ 2
= 9.81 × 0.4 × × 0.03
3

= 3.54 × 10−3 m3 /s

ISTUDY
440 CHAPTER 7. FLOW WITH FREE SURFACE

The discharge per unit width is


Q
q=
b
3.54 × 10−3
=
0.4
= 8.85 × 10−3 m2 /s
The discharge per unit width is in terms of hc is

q = gh3/2
c

Therefore,
q
h3/2
c =√
g
8.85 × 10−3
= √
9.81
= 2.826 × 10−3
hc = (2.826 × 10−3 )2/3
= 0.02 m
Therefore, fall in surface level becomes
Fall in surface level = H − hc
= 30 − 20

= 10 mm
EXAMPLE 7.14 A faired-crest weir is to be designed to control and
measure water flow rate, in the range from 45 m3 /hour to 70 m3 /hour. If
the level change between these flow rates is to be held 30 mm, determine the
width of the weir.

Solution Let subscripts 1 and 2 refer to the lower and higher flow rates,
respectively.

Given,
45
Q1 =
3600
= 0.0125 m3 /s
70
Q2 =
3600
= 0.0194 m3 /s

ISTUDY
7.9. THE BROAD-CRESTED WEIR 441

By Equation (7.30),

( )3/2
√ 2
Q= gb H
3

where H is the depth above the weir top-surface. Therefore,

( )3/2
√ 2
Q= 9.81 bH 3/2
3

= 1.705 bH 3/2
( )2/3
Q1
b2/3 H1 =
1.705
( )2/3
0.0125
=
1.705

= 0.0377
( )2/3
2/3 Q2
b H2 =
1.94
( )2/3
0.0194
=
1.705

= 0.0506

(H2 − H1 )b2/3 = 0.0506 − 0.0377

0.0506 − 0.0377
b2/3 =
(H2 − H1 )
( )3/2
0.0129
b=
0.03

= 0.282 m

EXAMPLE 7.15 Find the discharge over a flat-topped broad-crested weir


30 m wide, when the upstream level is 0.67 m above the crest and the discharge
coefficient is 0.66.

Solution Given, b = 30 m, H = 0.67 m and Cd = 0.66.

ISTUDY
442 CHAPTER 7. FLOW WITH FREE SURFACE

By Equation (7.30a), the discharge is


( )3/2
√ 2
Q = Cd g b H
3
( )3/2
√ 2
= 0.66 × 9.81 × 30 × × 0.67
3

= 18.513 m3 /s

EXAMPLE 7.16 Water flows across a broad-crested weir in a 0.4 m wide


rectangular channel at 0.00354 m3 /s. If the depth of water just upstream of
the weir is 70 mm, determine the height of the weir above the channel bed.

Solution Given, b = 0.4 m, h = 70 mm and Q = 0.00354 m3 /s

By Equation (7.30), the flow rate Q is


( )3/2
√ 2
Q = gb H
3

where H = h − z. Here z is the height of the weir above the channel bed.
Thus,
( )3/2
2 Q
H =√
3 gb

( )2/3
3 Q
H= √
2 gb
( )2/3
3 0.00354
= √
2 9.81 × 0.4

= 0.03 m

Therefore,

z =h−H

= 0.07 − 0.03

= 0.04 m

= 40 mm

ISTUDY
7.9. THE BROAD-CRESTED WEIR 443

7.9.1 Drowned Weir and Free Outfall


Critical flow or maximum discharge over a weir in a channel takes place only
in the absence of greater restrictions downstream. If the depth downstream
of the weir is sufficiently increased due to the presence of an obstruction
downstream, as shown in Figure 7.29, the flow over the weir may not be
critical, and the weir is then said to be drowned.

Energy line

Free surface hc

Critical depth line hc

hc Weir hc

FIGURE 7.29 A drowned weir in a channel.

Drowning of a weir may also be caused when the channel discharge into a
reservoir, as shown in Figure 7.30, in which the flow level is high enough to
maintain a depth greater than the critical depth over the weir.

Weir

Reservoir

FIGURE 7.30 A weir drowned due to high level in downstream reservoir.

When a broad-crested weir is drowned a depression of the surface over the


crest still occurs, but it is not sufficient for the critical depth to be reached.
But when a weir over which critical flow prevails (Figure 7.31(a)) is raised still
further above the channel bed, the liquid surface would rise by a corresponding
amount so that the depth over the crest would continue to be critical depth,
as illustrated in Figure 7.31(b).
A free outfall from a broad-crested weir in a long channel of mild slope
is shown in Figure 7.32. The discharge is a maximum for the specific-energy
available, and the flow must pass through the critical conditions in the vicinity
of the brink. At the brink the streamlines have a pronounced curvature.
Because of this large curvature the specific energy relation
u2
e=h+
2g

ISTUDY
444 CHAPTER 7. FLOW WITH FREE SURFACE

hc
hc

(a) (b)
FIGURE 7.31 A critical weir will continue to be critical even
when it is raised.
is invalid. The depth just behind the brink is less than the critical depth value
given by Equation (7.16).
True critical
depth here
Normal (q 2 /g)1/3
depth

s < sc

3 to 4 times (q 2 /g)1/3
FIGURE 7.32 Free outfall from a weir.

EXAMPLE 7.17 Calculate the discharge over a 15 m wide broad-crested


weir when the upstream level is 0.6 m above the crest and the coefficient of
discharge is 0.6.

Solution Given, b = 15 m, H = 0.6 m and Cd = 0.6.

By Equation (7.30a), we have the discharge as


( )3/2
√ 2
Q = Cd g b H
3
( )3/2
√ 2
= 0.6 × 9.81 × 15 × × 0.6
3

= 7.13 m3 /s

7.10 RAPID FLOW APPROACHING


A WEIR OR OTHER OBSTRUCTION
In our discussions on broad-based weir and drowned weir, the flow approaching
the weir was considered tranquil. But in actual situations even in a channel of

ISTUDY
7.10. RAPID FLOW APPROACHING ... 445

mild slope, rapid flow may be produced. This may occur when the flow enters
the channel down a spillway or from under a sluice gate. For a sufficient head
upstream of the sluice gate, the discharge through the aperture may be great
enough for the flow to be rapid. For example, in a channel of rectangular
cross-section if the discharge divided by width is q, the critical depth is given
by (q 2 /g)1/3 , and if this exceeds the height of the aperture, the flow becomes
rapid. Usually the velocity head upstream of sluice is negligible, therefore the
discharge is determined by the difference in head (h0 − h1 ) across the sluice
opening, shown in Figure 7.33. Thus

q = Cd h1 [2g(h0 − h1 )] (7.31)

where Cd is the coefficient of discharge and is approximately unity for well-


rounded aperture.

u20 /2g (negligible)


Energy line

u22 /2g
h0

h2 hc
h1 hc z

FIGURE 7.33 Free outfall from a weir.

If the friction is negligible, the total energy of the liquid remains un-
changed. But when the rapid flow reaches the raised part of the bed the
specific energy is reduced by an amount equal to the height of the weir, z.
For a moderate value of z, the approximate point in the specific energy dia-
gram, shown in Figure 7.34, moves from position 1 to 2. For this process the
flow remains rapid but the depth approaches the critical depth hc .

Tranquil

2 Rapid
hc 1

z E

FIGURE 7.34 Specific energy diagram.

ISTUDY
446 CHAPTER 7. FLOW WITH FREE SURFACE

For a large value of z the specific energy may be reduced to the minimum
value and the conditions over the weir are then critical. But the flow is not
controlled by the weir as it was when the approaching flow was tranquil. Rapid
flow can never be controlled by conditions downstream.
Increase of z beyond the limiting value at which the condition over the weir
has become critical, the specific energy remains invariant. This leads to the
increase of depth upstream of the weir. The flow is therefore tranquil instead
of rapid. However, the depth on the crest of the weir continues to be critical.
If the sluice is drowned, discharging liquid into the slower moving liquid at a
depth greater than the height of the aperture, as shown in Figure 7.35, the
weir will exercise control over the discharge, since the approaching flow is at a
depth greater than the critical. Eddy motion of high level turbulence develops
where the liquid discharged from the sluice enters the slowly moving liquid
ahead of it. Thus there is an appreciable reduction in the specific energy
over this part of the flow. Over the crest of the weir, however, there is little
production of eddies since the flow here is converging before becoming parallel.

Energy drops here


owing to losses
caused by eddies
Energy line

h0
hc
hc

FIGURE 7.35 A drowned sluice discharging liquid.

EXAMPLE 7.18 A sluice discharges 7.2 m3 /s into a frictionless channel.


When the head upstream of the sluice is 6 m (a) find the flow depth in the
channel which is 6 m wide. (b) If an obstacle exerts a resistance of 10 kN to
the flow, find the depth beyond the obstacle.

Solution Given, Q = 7.2 m3 /s, b = 6 m, h0 = 6 m and F = 10 kN.

By Equation (7.31), with Cd = 1,


2g(h0 − h1 )
q = h1
72 √
= h1 2g(h0 − h1 )
6
[ ]
h1 2g(h0 − h1 ) = 122
2

122
h21 (h0 − h1 ) =
2 × 9.81

ISTUDY
7.10. RAPID FLOW APPROACHING ... 447

h21 (6 − h1 ) = 7.34
h31 − 6h21 + 7.34 = 0

Solving, we get
h1 = 1.24 m

(b) Refer to Figure 7.24. By Equation (7.28),

ρgh1 ρgh2
h1 − h2 − F = ρg(u2 − u1 )
2 2
The velocity u1 and U2 are
Q
u1 =
h1 b
7.2
=
1.24 × 0.6
= 9.7 m/s
Q
u2 =
h2 b
12
=
h2
Therefore,
2F
h21 − h22 − = 2(u2 − u1 )
ρg
2 × (10 × 103 ) ( 12 )
1.242 − h22 − =2 − 9.7
10 × 9.81
3 h2
24
−h22 = − 19.9
h2
h32 − 19.9h2 + 24 = 0

Solving, we get
h2 ≈ 1.32 m

7.10.1 Flow through Venturi


Tranquil flow in an open channel may become critical not only by passing
through over a raised portion of the bed but also by passing through a reduced
width. Flow through a section of a channel with reduced width is illustrated
in Figure 7.36. The constriction shown in Figure 7.36 is called a Venturi.
Measurement of flow rate with Venturi has the following advantages over
a broad-crested weir.

ISTUDY
448 CHAPTER 7. FLOW WITH FREE SURFACE

Venturi

b1 b2

2
1
FIGURE 7.36 Flow through a Venturi in a channel.

• The loss of head experienced by the liquid in passing through the Venturi
is significantly less than that encountered while passing over a broad-
crested weir.
• Operation of the Venturi is not affected by the deposition of silt.
The mid-section of the Venturi with minimum cross-sectional area is known
as the throat, and the discharge per unit area of the throat is the maximum.
The flow at the throat is therefore critical. For a rectangular cross-section in
which the√streamlines are straight and parallel the velocity at the throat is
given by gh2 and the discharge becomes

Q = b2 h2 gh2 (7.32)

where b2 is the width at the throat and h2 is the corresponding depth.


Neglecting the frictional loss and the slope of the bed, we can write
u21 u2
h1 + = h2 + 2
2g 2g

But the velocity u2 = gh2 , therefore,

u21 h2
h1 + = h2 +
2g 2
or
u21 3
h1 + = h2 (7.33)
2g 2
where h1 and u1 , respectively, are the depth and velocity upstream of the
constriction. Also,
Q
u1 =
b1 h1
Substituting Equation (7.32), we get

b2 h2 gh2
u1 = (7.32a)
b1 h1
Substituting this into Equation (7.33), we get
( )2 3
1 b2 h2 3
h1 + 2 g = h2
2g b1 h1 2

ISTUDY
7.10. RAPID FLOW APPROACHING ... 449

that is,
( )3 ( )2 ( )2
h1 1 b2 3 h1
+ = (7.34)
h2 2 b1 2 h2
From Equation (7.34), it is seen that, for a given value of b2 /b1 , the ratio of
depths h1 /h2 is constant whatever the rate of flow.
Equation (7.32) may be expressed as
3/2
Q = b2 g 1/2 h2 (7.34a)
3/2
Dividing and multiplying the right-hand side by h1 , we have
3/2
b2 g 1/2 h1
Q= (7.35)
r3/2
where r = h1 /h2 , and the rate of flow through a given Venturi may be de-
termined by measuring the upstream depth h1 . The value of r in Equation
(7.35) has three roots and out of them the only one meeting the requirement
of r > 1 is ( )

r = 0.5 + cos
3
where 0 ≤ θ ≤ 90◦ and sin θ = b2 /b1 . If b1 is large compared to b2 , r becomes
equal to 1.5 and Equation (7.35) may be expressed as
3/2
b2 g 1/2 h1
Q= (7.36)
(1.5)3/2

Substituting Equation (7.33) into Equation (7.34a), the flow discharge may
be expressed as
( )3/2 ( )3/2
1/2 2 u21
Q = b2 g h1 + (7.37)
3 2g
where b1 is very large, u1 (= Q/(b1 h1 )) is small and u21 /(2g) may be neglected.
Note that, the discharge expressions in Equations (7.36) and (7.37) are
derived assuming that the friction is absent. But is actual flow, friction is
finite. Therefore, the volume flow rate obtained with these relations should be
corrected for frictional losses. Usually a correction factor known as discharge
coefficient Cd is used to account for the friction between the inlet and throat
of the Venturi. In practice, value of Cd lies between 0.95 and 0.99. The
discharge may be written as

3/2
b2 g 1/2 h1
Q = Cd (7.37a)
(1.5)3/2

Equations (7.37) and (7.37a) apply to flow through Venturi with critical flow
at the throat. A Venturi with critical condition at the throat is said to be un-
der free discharge condition. For free discharge the liquid surface downstream

ISTUDY
450 CHAPTER 7. FLOW WITH FREE SURFACE

of the Venturi is not maintained at too high a level. The level at the outlet
from the Venturi continues to fall and thus rapid flow exists where the width
of the passage again increases. If the velocity downstream of the Venturi is
greater than the critical velocity (that is, greater than the velocity at the
throat) the surface of the liquid issuing from the Venturi gradually merges
into the normal depth of the flow in the downstream, and any excess energy
possessed by the liquid emerging from the flume is dissipated by friction. On
the other hand, if the downstream velocity is less than the critical velocity, the
flow has to change from rapid to tranquil. The change from rapid to tranquil
normally takes place through a hydraulic jump, as illustrated in Figure 7.37.

Hydraulic jump

hc

FIGURE 7.37 Transition of flow from tranquil to rapid through a


hydraulic jump.
For a rectangular channel, the hydraulic jump occurs at a point where
the depth of the rapid flow is such as to give the correct depth of subsequent
tranquil flow, in accordance with Equation (7.25). A limiting position is that
in which the jump resulting in zero height occurs at the throat itself. If the
downstream level is raised further the velocity at the throat no longer reaches
the critical velocity and the flow through the Venturi is termed drowned. In
these situations the ratio h1 /h2 is no longer constant and therefore the rate of
flow cannot be determined from a measurement of the upstream depth only.
Both h1 and h2 (on either side of the throat) need be measured to calculate
the flow rate.
A properly designed Venturi under ideal conditions the loss of mechanical
energy may be kept as low as 10%. Like the broad-crested weir, the Venturi
is also suitable for measurement only when the flow is tranquil.

EXAMPLE 7.19 A Venturi in a horizontal channel of rectangular cross-


section 1.4 m wide by constricting the width to 0.9 m and raising the floor
level in the constricted section by 0.25 m above that of the channel. If the
difference in level of the free surface between the throat and upstream is 30
mm and both upstream and downstream depth are 0.6 m, calculate the volume
flow rate.

Solution Given, b1 = 1.4 m, b2 = 0.9 m, h1 = 0.6 m, h2 = h1 + 0.03 =


0.6 + 0.03 = 0.63 m.

By Equation (7.32a), the upstream velocity is



b2 h2 gh2
u1 =
b1 h1

ISTUDY
7.11. GRADUALLY VARIED FLOW 451

Thus,

0.9 × 0.63 × 9.81 × 0.63
u1 =
1.4 × 0.6
= 1.678 m/s

The volume flow rate is

Q = A1 u1
= (1.4 × 0.6) × 1.678

= 1.41 m3 /s

Aliter: The volume flow rate can also be calculated using Equation (7.35),
which expresses Q as
3/2
b2 g 1/2 h1
Q=
r3/2
where
h1
r=
h2
0.6
=
0.63
= 0.952

Thus

0.9 × 9.81 × 0.63/2
Q=
(0.952)3/2

= 1.41 m3 /s

7.11 GRADUALLY VARIED FLOW


Uniform flow discussed in the foregoing sections of this chapter is usually
experienced only in artificial channels because the condition required for uni-
form flow is encountered in cross-sections of constant shape and area. That is,
for uniform flow the liquid surface must be parallel to the bed of the channel
and this in turn demands that the slope of the bed be constant. In a natural
stream, such as a river, the shape and size of cross-section and the slope of
the bed usually vary appreciably, and truly uniform flow is extremely rare.
Achieving the condition required for uniform flow is not practically possible.
Therefore, the equations for uniform flow give only approximate results and
not the correct ones.
For a particular shape, there is only one depth at which uniform flow can
take place. This depth is known as the normal depth. But there are number

ISTUDY
452 CHAPTER 7. FLOW WITH FREE SURFACE

of ways in which the same steady rate of flow can take place along the same
channel in non-uniform flow. For non-uniform flow the liquid surface will not
be parallel to the bed and will take the form of a curve.
Among the steady and non-uniform flows, there are two kinds in each
category. In one kind of non-uniform flow the changes of depth and velocity
takes place over a long distance. Such flow is termed gradually varied flow.
In the second type the changes of depth and velocity take place within a
short distance and may also be quite abrupt, as in a hydraulic jump. This
kind of non-uniform flow is termed rapidly varied flow. It is usual to assume
the gradually varied flow as that in which the changes occur slow enough for
the effects of the acceleration of the liquid to be negligible. Because of this
assumption, the relation based on the gradually varied flow is not applicable
to rapidly varied flow.
Gradually varied flow may be caused by a change in the geometry of the
channel; such as the change in the shape of the cross-section and change of
slope or an obstruction, or by a change in the frictional forces at the bound-
aries. It can occur both in tranquil and rapid flow. In a tranquil flow, if the
depth is increased upstream of an obstruction the resulting curve of the liquid
surface, as illustrated in Figure 7.38(a), is usually known as a backwater curve.
The curvature effect, such as a fall in the surface as the liquid approaches a
free outfall from the end of the channel, as shown in Figure 7.38(b), is termed
a drown-drop or drawn-down curve. Both curves are asymptotic to the surface
of uniform flow.
du
dl
<0

(a) Backwater curve

du
dl
>0

(b) Downdrop curve

FIGURE 7.38 Illustration of backwater and down-drop curves.

7.12 GOVERNING EQUATIONS OF


GRADUALLY VARIED FLOW
The equation representing the surface profile is required for estimating the
depth of a stream at a particular point or to determine the distribution over
which the effects of an obstruction such as a weir are transmitted upstream.

ISTUDY
7.12. GOVERNING EQUATIONS OF GRADUALLY VARIED FLOW 453

The depth at a particular section determines the area of the cross-section


and hence the mean velocity. Thus it is necessary to investigate the way in
which the depth changes with distance along the channel. But the changes in
gradually varied flow occupy a considerable distance, therefore, the effects of
boundary friction are important. Note that, this is directly opposite to rapidly
varied flow, where the boundary friction is usually neglected in comparison
with the other forces involved.
For developing the governing equation for gradually varied flow, it is es-
sential to distinguish between the slope i of the energy line and the slope s of
the channel bed. In uniform flow i and s are identical.
Let us examine the gradually varied flow illustrated in Figure 7.39. It is
seen that over a short distance δl of the channel the flow is steady. Over this
length the bed falls by an amount s δl, the depth of the flow increases from h
to (h + δh) and the mean velocity increases from u to (u + δu). Let us assume
that the increase of level δh is constant across the width of the channel.
u2 /2g
i δl

Energy line
(u + δu) 2/2
g
u u + δu
h h + δh

sδl Datum line


δl
FIGURE 7.39 Gradually varied flow.

Now, assuming the slope of the flow to be straight and parallel and the
slope of the bed is small so that the variation of pressure with depth is hy-
drostatic, the total head above the datum level can be expressed as
α u2
h + s δl +
2g
where α is the kinetic energy correction factor accounting for the non-uniformity
of velocity over the cross-section. For simplifying the analysis, let us assume
that α is only slightly different from unity, so that it may be omitted, without
introducing any appreciable error.
The total head above datum at the second section is
(u + δu)2
(h + δh) +
2g
If the loss of head due to friction, divided by the length along the channel,
termed the head loss gradient, is i, then
u2 (u + δu)2
h + s δl + − i δl = (h + δh) + (7.38)
2g 2g

ISTUDY
454 CHAPTER 7. FLOW WITH FREE SURFACE

Expanding the right-hand side and neglecting the term (δu)2 and simplifying,
we get
u δu
δh = (s − i) δl −
g
In the limit as δl → 0, this becomes
dh u δu
= (s − i) − (7.39)
dl g dl
From continuity equation, we have

Au = constant (7.40)

where A is the cross-sectional area of the flow. Differentiating with respect to


l, we have
du dA
A +u =0
dl dl
or
du u dA
=− (7.41)
dl A dl
To evaluate dA/dl, let us assume that the channel is prismatic, that is, its
shape and alignment do not vary with l. Thus the cross-sectional area of the
flow, A, changes only as a result of a change in h. In Figure 7.19, it is seen
that
δA = B δh
where B denotes the surface width of the cross-section. Therefore,
dA dh
=B
dl dl
Substituting this into Equation (7.41), we have
du u dh
=− B
dl A dl
Now, substituting this into Equation (7.39), we obtain

dh u2 B dh
= (s − i) +
dl g A dl
Thus
dh (s − i)
=
1 − ug AB
2
dl
or
dh (s − i)
= (7.42)
dl 1 − gh
u2

where h = A/B is the mean depth. This relation represents the slope of the
free surface with respect to the channel bed.

ISTUDY
7.12. GOVERNING EQUATIONS OF GRADUALLY VARIED FLOW 455

Equation (7.42) shows that when i = √ s, dh/dl = 0 and the flow is uni-
form. The term u2 /(gh) is Fr2 (Fr = u/ gh), therefore for critical flow, the
denominator of the right-hand side of Equation (7.42) is zero. It is important
to note that dh/dl becomes infinite for Fr = 1, leading the surface of the flow
perpendicular to the bed. Such a conclusion is not valid for the assumption of
gradually varied flow. Therefore, for situation where the liquid surface is per-
pendicular to the bed Equation (7.42) is not valid. Nevertheless the equation
is valid for critical flow if the numerator
√ is also zero.
For tranquil flow with u < gh, the denominator√of Equation (7.42) is
positive. On the other hand, for rapid flow with u > gh, the denominator
is negative. Thus if the slope of the bed is less than that corresponding to
the rate of dissipation of energy by friction, that is, if s < i, then dh/dl is
negative for tranquil flow and the depth decreases in the direction of flow. For
rapid flow, in these situation, dh/dl is positive and at critical condition dh/dl
changes sign, and this change of sign leads to unstable depth of the liquid flow
at critical state. However, wavy surface and exactly same conditions cannot
be maintained over a finite length of a channel.
In Equation (7.42), i is a function of u and the equation cannot be inter-
preted directly. To simplify this state let us assume that the value of i at a
particular section is the same as it would be for uniform flow having the same
velocity and hydraulic mean depth. This assumption gives results of accept-
able accuracy. Any error introduced by this assumption is small compared
to the uncertainties in selecting suitable values of Chézy’s coefficient or man-
ning’s roughness coefficient. For a given steady discharge Q, u is a function of
h only and so the integration of Equation (7.42) is possible, and even then the
result is complicated. Therefore, normally the solution is obtained by either
a numerical or graphical integration.
Usually we are interested in determining the position in the channel at
which a particular depth is reached, that is, l is required for a particular value
of h. Therefore, Equation (7.42) may be rewritten as
∫ h2
1 − u2 /(gh)
l= dh (7.42a)
h1 s−i

EXAMPLE 7.20 Find the slope of the water surface in millimeter per
kilometer at a section in a rectangular channel at which the discharge is 2.8
m3 /s, width 15 m, depth 3 m, bed slope 1 in 6700 and the Chézy’s coefficient
is 61 m1/2 /s.

Solution Given, Q = 20 m3 /s, b = 15 m, h = 3 m, s = 1/6700 = 1.49×10−4


and C = 61 m1/2 /s.

The mean velocity is

Q
u=
bh

ISTUDY
456 CHAPTER 7. FLOW WITH FREE SURFACE

28
=
15 × 3
= 0.622 m/s

The cross-sectional area, A and perimeter, P are

A = bh
= 15 × 3
= 45 m2
P = b + 2h
= 15 + 2 × 3
= 21 m

Therefore, the mean hydraulic depth becomes

A
m=
P
45
=
21
= 2.143 m

By Equation (7.7), we have √


u = C mi
where i is the energy gradient. Therefore,

u2
i=
C 2m
0.6222
= 2
61 × 2.143
= 4.852 × 10−5

By Equation (7.42), the slope of the water surface is given by

dh s−i
=
dl 1 − gh
u2

1.49 × 10−4 − 0.4852 × 10−4


=
1 − 9.81×3
0.6222

1.0048 × 10−4
=
0.987
= 1.018 × 10−4 m/m
= 101.8 mm/km

ISTUDY
7.12. GOVERNING EQUATIONS OF GRADUALLY VARIED FLOW 457

EXAMPLE 7.21 Water flows through a 6 m wide rectangular channel


with a slope of 1 in 2000. If the depth is 0.9 m and flow rate is 8.5 m3 /s, (a)
determine the value of Chézy’s coefficient. (b) In order to raise the depth of
water a dam is built downstream. If the depth of water in the dam is 1.8 m,
how far upstream the depth will be 1.5 m.

Solution (a) Given, Q = 8.5 m3 /s, b = 6 m, h = 0.9 m and bed slope


i = 1/2000 = 0.0005.

The average velocity is


Q
u=
A
Q
=
bh
8.5
=
6 × 0.9
= 1.57 m/s

The cross-sectional area, A, and perimeter, P are

A = bh
= 6 × 0.9
= 5.4 m2
P = b + 2h
= 6 + 2 × 0.9
= 7.8 m

The hydraulic mean depth is


A
m=
P
5.4
=
7.8
= 0.692 m

The Chézy’s constant, C, by Equation (7.7), is


u
C=√
mi
1.57
=√
0.692 × 0.0005
= 84.4

ISTUDY
458 CHAPTER 7. FLOW WITH FREE SURFACE

(b) Let subscripts 1 and 2 refer to location with water depth 1.8 and 1.5,
respectively.

The mean depth is


h1 + h2
h=
2
1.8 + 1.5
=
2
= 1.65 m

Therefore, the cross-sectional area, A, and perimeter, P are

A = bh
= 6 × 1.65
= 9.9 m2
P = b + 2h
= 6 + 2 × 1.65
= 9.3 m

The hydraulic mean depth is


A
m=
P
9.9
=
9.3
= 1.065 m

The mean velocity is


Q
u=
A
8.5
=
9.9
= 0.86 m/s

The value of i, by Equation (7.7), is

u2
i=
C 2m
0.862
=
84.42 × 1.065
= 0.975 × 10−4

ISTUDY
7.13. CLASSIFICATION OF SURFACE PROFILE 459

By Equation (7.42),

dh s−i
=
1− u
2
dl
g(h)

0.0005 − 0.0000975
=
1 − 9.81×1.65
0.862

= 4.22 × 10−4

Therefore,
dh
dl =
4.22 × 10−4
h1 − h2
=
4.22 × 10−4
1.8 − 1.5
=
4.22 × 10−4
= 710.9 m

Thus the distance between the locations of height 1.8 m and 1.5 m is 710.9 m.

7.13 CLASSIFICATION OF SURFACE


PROFILE
The surface profile may have a variety of forms, depending on how the flow
is controlled by weirs and other obstructions, changes in bed slope and so on.
These circumstances affect the relative magnitude of the quantities appearing
in the equation and dh/dl is accordingly positive of negative. Any problem
may be studied by considering separate lengths of the channel, in each of
which the flow corresponds to one of the various types of surface profile. The
separate lengths may be studied one by one until the entire problem has been
covered.
The slope of the bed may be adverse or uphill (A), zero or horizontal (H)
with s = 0, mild (M ) with s < sc , critical (C) with s = sc and steep (S) with
s > sc . The profiles are further classified according to the depth of the stream.
If the depth is greater than the normal depth h0 and the critical depth hc (that
is, h > h0 and h > hc ), the profile is of type 1; for h0 < h < hc , the profile
is of type 2; and if h < h0 and h < hc , the profile is of type 3. The types of
non-uniform flow possible are illustrated in Figure 7.40. It may appear that
there are 15 different types of profiles possible with 5 of them represented by
letters A, H, M , C and S and 3 of them represented by numbers 1, 2 and 3.
But normal flow is impossible on either an adverse or a zero slope, therefore,
A1 and H1 do not exist. Also, when the slope is critical with h0 = hc , C2
curve cannot occur.

ISTUDY
460 CHAPTER 7. FLOW WITH FREE SURFACE

When the flow velocity is progressively reduced with increasing depth,


all the type 1 curves approach a horizontal asymptotically. Moreover, all
curves, except C curves, that approach the normal depth line h = h0 do so
asymptotically. This is because uniform flow takes place at sections which
are free from any disturbances. The curves that cross the critical depth line
h = hc perpendicularly because at critical conditions the denominator of
Equation (7.42) is zero and dh/dl becomes infinite. For C curves h0 is identical
to hc and due to this a C curve cannot approach a line asymptotically and
perpendicularly at the same time.
On a mild slope the normal, uniform flow is tranquil and therefore h0 > hc .
For M1 curve h > h0 , i < s and since the flow is tranquil the denominator
of Equation (7.42) is positive. Therefore, dh/dl is positive and the depth
increases in the direction of flow. The M1 curve is the usual backwater curve
caused by a dam or weir. As h approaches h0 , i approaches s and dh/dl → 0;
thus the normal depth is asymptotic at the upstream end of the curve. For
M2 curve, h0 > h > hc . The numerator of Equation (7.42) is negative,
although the denominator is still positive. Therefore dh/dl is negative. This
curve represents accelerated flow and is the down drop curve approaching a
free outfall at the end of a channel of mild slope. For the M3 curve h < hc ,
the flow is rapid, the denominator becomes negative and so dh/dl is positive;
thus the depth increases in the downstream direction, as shown in Figure 7.40.
This curve results from an upstream control such as a sluice gate. Since the

FIGURE 7.40 Types of non-uniform flow.

ISTUDY
7.13. CLASSIFICATION OF SURFACE PROFILE 461

bed slope is not sufficient to sustain rapid flow a hydraulic jump will form at
a point where the equation for the jump can be satisfied.
The slope of a channel may be classified as mild for one rate of flow,
critical for another and steep for a third. The quantitative analysis of surface
profiles is applicable to channels of any slope and roughness, provided that
local variations of slope, shape, roughness and so on are properly taken into
account.
Flows with a free surface may distort into a variety of shapes. Here we
consider disturbances moving over the free surface of a liquid in a periodic
manner, such as wind-generated ocean waves. At a position some liquid rises
above the mean level and then subsides below it, and in doing so appears to
travel over the surface. But what actually travels over the surface is simply
the form of the disturbances. There is practically no net movement of the
liquid itself. For example, an object floating on the surface moves forward
with the crest of the wave but, in the succeeding trough, returns almost to
its original position. As wind-generated ocean waves move into shallow water
close to the shore, the regular wave motion is modified, but even so there is
no net flow of water towards the shore.

7.13.1 The Basic Equations of Motion


Wind-generated ocean waves are found to travel unchanged in size and shape
over long distances and this indicates that effects of viscosity are very small.
In the study of this kind of waves, the liquid can be assumed to be inviscid,
and if we further assume that the motion was originally generated from rest,
it can be regarded as irrotational.
A two-dimensional flow in xz-plane is illustrated in Figure 7.41. In this
field, fluid in the x-direction is unlimited. Let us assume that the horizontal
bed is rigid, and any other fixed boundaries are vertical planes in the x direc-
tion. The effect of any boundary layer is assumed to be negligible. Shear stress
between the liquid and the atmosphere above it are also negligible, as the den-
sity of air is very small compared to the liquid density. Furthermore, the effect
of air being set in motion by the wave is also ignored. The wavelength, λ, is
defined as the distance between corresponding points on consecutive waves.

λ
z
η
x

FIGURE 7.41 Two-dimensional flow in xz-plane.

For constant wave shape, the velocity C of the wave in the x-direction is
constant, hence for an observer traveling along x-direction with velocity C the

ISTUDY
462 CHAPTER 7. FLOW WITH FREE SURFACE

flow pattern would appear steady. For this steady flow pattern, the wave can
be assumed to be sinusoidal. For any point in the liquid the stream function
ψ may be expressed as

ψ = Cz + f (z) sin mx (7.43)

where z is the coordinate in the vertical direction, f (z) is a function of z, and


m = 2π/λ. The term Cz accounts for the uniform velocity.
For irrotational motion, ψ must satisfy Laplace equation

∂2ψ ∂2ψ
+ =0
∂x2 ∂z 2
Substituting from Equation (7.43), we have

d2 f
−m2 f + =0
dz 2
The solution of this equation is

f = A sinh (B + mz)

where A and B are constants. Thus Equation (7.43) becomes

ψ = Cz + A sinh (B + mz) sin mx

The free surface, where z = η (say), must be composed of streamlines and so

ψ = Cη + A sinh (B + mη) sin mx = constant (7.44)

This implies that dη/dx = 0, at the crest or trough, and the position x = 0
must be midway between a trough and a crest. If the free-surface streamline
is represented by ψ = 0, from Equation (7.44) it is seen that streamline with
ψ = 0 must cross x = 0 where z = 0, as shown in Figure 7.41. Also, because
the bed at z = −h consists of streamlines, ψ at the bed is independent of x,
hence
sinh [(B + m(−h)] = 0
that is, B = mh. Therefore, in general,

ψ = Cη + A sinh [m(h + z)] sin mx (7.45)

In particular, at the free surface,

0 = Cη + A sinh [m(h + η)] sin mx (7.46)

Hence, if, as is usual for ocean waves (except when close to the shore), η is
small compared with both h and λ, then the free surface has a sinusoidal form
given by ( )
A
η ≈ − sinh mh sin mx [for η ≪ h, η ≪ λ] (7.47)
C

ISTUDY
7.13. CLASSIFICATION OF SURFACE PROFILE 463

For steady flow, the elevation η of the free surface can be related to veloc-
ity, using Bernoulli’s equation. The velocity V at any location (x, z) can be
expressed as
( )2 ( )2
∂ψ ∂ψ
V2 = + −
∂x ∂z
Differentiating ψ given by Equation (7.44) with respect to x and z and sub-
stituting into this, we have

V 2 = A2 m2 sinh2 [m(h + z)] cos2 mx + [−C − A mcosh (m[h + z]) sin mx]2
(7.48)
For z = η, Equation (7.48) gives the velocity at the free surface. Substituting
for A from Equation (7.46) and neglecting terms in m2 η 2 [because η ≪ h], we
obtain
2
Vsurface = C 2 [1 − 2mη coth (m[h + η])]
or
2
Vsurface ≈ C 2 (1 − 2mη coth [mh]) [for η ≪ h, η ≪ λ] (7.49)

The pressure above the freestream is atmospheric but, because the surface
is not plane, the pressure in the liquid is, in general, modified by surface
tension (σ). The surface tension force divided by the distance perpendicular
to the plane in Figure 7.42 is σ and at the position P its vertical component
is σ sin θ.

Free surface
δη

P θ R
σ δx
η

FIGURE 7.42 Surface tension at a free surface.

At point Q the vertical component is

d
σ sin θ + (σ sin θ) δx
dx
Hence the net upward surface tension force on P Q is

d
(σ sin θ) δx
dx

ISTUDY
464 CHAPTER 7. FLOW WITH FREE SURFACE

If a mean gauge pressure p acts over P R, the total upward force on the fluid
in the control volume P QR is
d
(σ sin θ) δx + p δx
dx
This is the rate of increase of vertical momentum of the fluid through PQR.
This momentum term, is proportional to δx and δη. Therefore, after dividing
the equation by δx and letting δη → 0, the momentum term vanishes resulting
in the pressure p, immediately below the free surface as
d
p=− (σ sin θ)
dx
or [ ]
d dη/dx
p = −σ (7.50)
dx [1 + (dη/dx)2 ]1/2
For a wave in which η is small compared to both h and λ, from Equation
(7.47), we have

dη A
= − m sinh (mh) cos (mx) = m η cot (mx) (7.51)
dx C
But dη/dx is small compared to unity, therefore Equation (7.50) reduces to

d2 η
p = −σ
dx2
Substituting for d2 η/dx2 , from Equation (7.51), this gives

p = σm2 η

Applying Bernoulli’s equation between the points η = η and η = 0 on the


free-surface streamline, and using Equation (7.49), we obtain
1 1
σm2 η + ρC 2 (1 − 2mη coth (mh)) + ρgη = ρC 2 (7.52)
2 2
where
( ) ( ) ( )
σm g 2πσ gλ 2πh
C2 = + tanh (mh) = + tanh (7.53)
ρ m ρλ 2π λ

Equations (7.52) and (7.53) are valid only for η ≪ h and η ≪ λ.


The velocity C is known as the phase velocity because points of the same
phase (that is, of equal η) move at this velocity, regardless of the shape of the
wave. Wave propagation is described as dispersive when the phase velocity
depends on the wave length λ. This is because if a wave of general shape were
split into components of different wavelengths, the components would move
at different velocities and thus become separated, that is, dispersed.

ISTUDY
7.14. GRAVITY WAVES 465

Equation (7.53) shows that the effect of surface tension on the phase ve-
locity is negligible if
2πσ gλ

ρλ 2π
that is, if
( )1/2
σ
λ ≫ 2π
ρg

σ
For water 2π ρg is about 17 mm, therefore the effect of surface tension can
be neglected.
Deep-water waves are those with h > λ/2. Therefore, tanh (2πh/λ) differs
from unity by less than 0.004 and we can take

2πσ gλ
C2 = + (7.53a)
ρλ 2π

This has a minimum value where



σ
λ = 2π
ρg

For water
Cmin ≈ 0.23 m/s

7.14 GRAVITY WAVES


Waves whose properties are determined by gravity effects are referred to as
gravity waves. Surface tension effects can therefore be ignored and Equation
(7.53) reduces to ( )
gλ 2hπ
C2 = tanh (7.53b)
2π λ
( )
For long waves, with λ ≫ h, tanh 2hπλ → 2hπ
λ , resulting in


C= gh (7.53c)

The long wave with λ ≫ h is usually termed shallow-water wave. However


the term long wave is more appropriate, because it more accurately reflects
the fact that waves of long wavelengths are included, irrespective of whether
the water is considered deep or shallow.
When surface tension effects are neglected, for deep-water waves tanh (2πh/λ)
is approximately equal to 1, and Equation (7.53b) simplifies to


C2 = (7.53d)

ISTUDY
466 CHAPTER 7. FLOW WITH FREE SURFACE

7.15 CAPILLARY WAVES


Waves whose characteristics are governed mainly by surface tension are known
as capillary waves. Waves of very short length are termed ripples. For ripples
the term in Equation (7.53) involving g becomes negligible. For ripples on
water with λ = 3 mm (say), gravity affects C by only 1.5%. When λ is
small compared √ to h then tan (2πh/λ) → 1 and the velocity of capillary
waves become 2πσ/(ρλ). The frequency, namely the number of wave crests
passing a given point divided by time interval, is C/λ, which for λ = 3 mm on
deep water, is 131 Hz. Thus capillary waves can be generated by tuning fork
held in a liquid, however, they rapidly decay and cannot be seen for more than
a few centimeters. But the measurements of the length of waves produced by
a tuning fork of known frequency is the basis of one method of determining
surface tension.

7.16 MOTION OF INDIVIDUAL PARTICLES


Imposing a negative velocity (−C) with respect to stationary axes, the flow
may be rendered steady. As the conditions at a point will then depend also
on time t, we should use (x − Ct) in place of x.
For any position (x, z), using Equation (7.45), the absolute horizontal
velocity of particle can be expressed as
∂ψ
C− = −Am cosh [m(h + z)] sin [m(x − Ct)] (7.54)
∂z
Absolute vertical velocity of particle is
∂ψ
= Am sinh [m(h + z)] cos [m(x − Ct)] (7.55)
∂x
The displacements X, Z, from the mean positions, can be obtained by inte-
grating these expressions with respect to t. If the displacement are small, as
in the case of waves of small amplitude, x and z may be approximated by
their mean values x and z, thus
A
X=− cosh [m(h + z] cos [m(x − Ct)]
C
A
Z=− sinh [m(h + z] sin [m(x − Ct)]
C
Also,
X2 Z2
( A )2 + ( )2 =1 (7.56)
C cosh2 [m(h + z] A
C sinh2 [m(h + z]
This is the equation of an ellipse with semi-axes
A
cosh2 [m(h + z)] horizontally
C

ISTUDY
7.16. MOTION OF INDIVIDUAL PARTICLES 467

A
sinh2 [m(h + z)] vertically
C
The elliptical orbits of particles in shallow liquid is illustrated in Figure 7.43.

FIGURE 7.43 Instantaneous positions of particles, for a wave traveling


from left to right (the particles move clockwise).

When h ≫ λ, tanh (mh) → 1, that is, cosh (mh) → sinh (mh) and the
orbit near the free surface (that is, at z → 0) becomes circles. For h ≪ λ,
m(h+z) becomes much less than 1; thus cosh [m(h+z)] → 1 and the horizontal
axis of an orbit is practically independent of z.
For waves with finite amplitude, the paths followed by individual particles
would not be closed curves and the particles would slowly move with the wave.

7.16.1 Wave Energy


The energy of a wave is the sum of the contribution from gravitational energy,
kinetic energy and free-surface energy.
The gravitational energy of a fluid element (δx × δz) per unit width per-
pendicular to x-direction, relative to the equilibrium level, z = 0, is
(ρgδxδz) z
Therefore, the total gravitational energy of a sinusoidal wave, for a complete
wavelength is
∫ λ (∫ η ) ∫ λ
1
ρgzdz dx = ρg η 2 dx
0 0 2 0
∫ λ
1
= ρg a2 sin2 [m(x − Ct)] dx
2 0

1
= ρg a2 λ
4
where a is the amplitude of the wave, the maximum value of a is η, and
m = 2π/λ.

ISTUDY
468 CHAPTER 7. FLOW WITH FREE SURFACE

The kinetic energy per unit width, ke, for a complete wavelength, calcu-
lated from the velocity component for a particle, from Equations (7.54) and
(7.55), is
∫ λ( ∫ η [
1
ke = ρ A2 m2 cosh2 [m(h + z)] sin2 [m(x − Ct)]
2 0 −h
] )
+ A m sinh2 [m(h + z)] cos2 [m(x − Ct)] dz dx
2 2

∫ λ( ∫ η [ ] )
1
= ρA2 m2 cosh [m(h + z)] − cos [m(x − Ct)] dz dx
2 2
2 0 −h
∫ λ( )
1 2 2 1 (h + η)
= ρA m sinh [2m(h + η)] − cos [2m(x − Ct)] dx
2 0 4m 2
Assuming η ≪ h and mη ≪ 1, we obtain
1
ke = ρA2 mλ sinh (2mh)
8
From Equation (7.47),
A = −C a cosech (mh)
therefore,
1[ ]2
ke = − C a cosech (mh) mλ sinh (2mh)
8
This simplifies to
1 2
ρ a mλC 2
ke =
4
Substituting [from Equation (7.53a)],
2πσ gλ
C2 = +
ρλ 2π
we get the kinetic energy as
( )
1 2πσ gλ
ke = ρ a2 mλ +
4 ρλ 2π
But 2π/λ = m, thus
( )
1 mσ g
ke = ρ a2 mλ +
4 ρ m
1 2 ( )
a λ ρ g + σm2
=
4
When the surface tension effect, which is the work done in stretching the
surface in one wavelength when the wave is formed, is significant, its effect
also should be added. The length of the surface is
∫ λ[ ( )2 ]1/2

1+ dx
0 dx

ISTUDY
7.17. ENERGY TRANSMISSION RATE 469

For small amplitude, this gives


∫ λ [ ]1/2
1 + a2 m2 cos2 [m(x − Ct)] dx
0
∫ λ [ ]
1
≈ 1 + a2 m2 cos2 [m(x − Ct)] dx
0 2
1 2 2
= λ+ a m λ
4
1
This shows that the free surface energy over one wavelength is 4 σa2 m2 λ.

Total energy = Gravitational energy + Kinetic energy + Free-surface en-


ergy

Thus the total energy divided by the width of the wave, over one wavelength
becomes
1 2 ( )
a λ ρ g + σm2 (7.57)
2
This is valid for n ≪ h, η ≪ λ.
It should be noted that, the assumption of horizontal bed (that is, uniform
h) is always reasonable, since variations of h have no appreciable effect on the
result provided that h exceeds λ/2.

7.17 ENERGY TRANSMISSION RATE


The shape of the wave moves with velocity C. But all the elements of a liquid
carrying energy do not move at velocity C. The quantity
1
p∗ + ρ(u2 + v 2 )
2
is a measure of the amount of energy carried by a small element of liquid
divided by the volume of the element. Here p∗ is the piezometric pressure and
u, v are the horizontal and vertical components of velocity, respectively. In
any fixed vertical plane perpendicular to the x direction the volume flow rate
through a small element of height δz and unit breath is uδz and so the rate
at which the total amount of energy is transferred across that plane is given
by ∫ η [ ]
1
p∗ + ρ(u2 + v 2 ) u dz (7.58)
−h 2
Since the flow is assumed to be irrotational, Bernoulli’s equation may be
applied between any two points in steady flow. One of these points may be
taken as the free surface at η = 0. Therefore, for steady motion of waves of
small amplitude,
1 [ ] 1
p∗ + ρ (u − C)2 + v 2 = ρC 2 (7.59)
2 2

ISTUDY
470 CHAPTER 7. FLOW WITH FREE SURFACE

This simplifies to
1
p∗ + ρ(u2 + v 2 ) = ρ u C
2
Substituting this into Equation (7.58), we obtain
∫ η
ρC u2 dz
−h

Substituting the expression for u (that is ∂ψ/∂y) from Equation (7.54), we


get the rate of energy transfer across the plane as
1 ( 1 )
ρCA2 m2 h + sinh (2mh) sin2 [m(x − Ct)] (7.60)
2 2m
where η has been neglected in comparison with h.
Integration of Equation (7.60) with respect to t for the passage of one
wavelength through the fixed plane, and then dividing by the corresponding
period λ/C, the mean rate of energy transfer, per unit breath, can be obtained.
But, from Equation (7.47),

A = −C a cosech (mh)

From Equation (7.53d),



C2 =

Therefore, the mean rate of energy transfer per unit breath becomes
1
ρC a2 (1 + 2mh cosech 2mh)(ρg + σm2 )
4
The mean total energy divided by the plan area is
1 2
a (ρg + σm2 )
2
Diving the above mean rate of energy transfer by this factor, we get the
velocity of energy transmission as
1
C(1 + 2mh cosech 2mh) (7.61)
2

7.18 GROUP VELOCITY


When a still liquid is disturbed at a particular position, a group of waves is
produced that moves away from the point of disturbance. A typical example
for this is the waves produced when a stone is thrown into a pond. The waves
generated this way advance as a group at a velocity less than that of the
individual waves within it. Individual waves appear to move forward through
the group, grow to a maximum amplitude, and then diminish before vanishing

ISTUDY
7.18. GROUP VELOCITY 471

entirely at the front of the group. This is due to the process that the group
has components that in general are of slightly different wavelength (hence of
slightly different velocity). For example, the two train of waves A and B with
velocity C and (C +δC) shown in Figure 7.44, combine together as illustrated,
forming a wave group.
Wave trains A and B differ slightly in wavelength, owing to this the total
wave form resulting from their addition has a slowly varying amplitude, giving
the appearance of group waves alternating with intervals of almost still liquid.
The beats produced by the conjunction of two trains of sound waves of nearly
equal wavelength, or the amplitude modulation of radio waves is a typical
example of this kind of wave motion.

a′ a
λ
A C

b′ b
λ + δλ

B C + δC

Wave group (A + B)

FIGURE 7.44 Illustration of wave group formation.

When the crests a and b coincide, they form a maximum combined ampli-
tude for the group at that position. A little later the faster among the waves
A and B would gain a distance δλ relative to the slower wave and then the
crests a′ and b′ would coincide. Since the relative velocity between trains A
and B is δC, this takes a time δλ/δC. However, while the point where the
crests coincide has moved back a distance λ (relative to A), A itself has the
time interval δλ/δC moved forward a distance C × (δλ/δC). Thus the maxi-
mum combined amplitude has moved forward a net distance (C δλ/δC − λ).
The velocity with which it does so is therefore
C δλ/δC − λ δC
=C −λ
δλ/δC δλ
as both δC and δλ tend to zero, this becomes
δC dC
C −λ =C +m (7.62)
δλ dm

ISTUDY
472 CHAPTER 7. FLOW WITH FREE SURFACE

This is known as the group velocity Cg , and clearly, unless the individual wave
velocity C is independent of the wavelength, the group velocity differs from
C. When there are only slight variations between C and λ, the result is for
number of waves of any type. For small surface waves Equation (7.62) shows

dC
Cg = C + m
dm
[ ]
C m d(C 2 )
= 2+ 2
2 C dm

that is,
[ ]
C 3 σm2 + ρg 2 mh
Cg = + (7.63)
2 σm2 + ρg sinh 2mh

When the surface tension effect is negligible, the first term within the bracket
is unity, leading to Cg to become the same as the velocity of energy trans-
mission [Equation (7.61)] and it varies between C/2 (when mh → ∞) to C
(when mh → 0). For small waves in which surface tension effect is significant,
the group velocity differs from the velocity of energy transmission and may
exceed the velocity of individual waves.

7.19 WAVES MOVING INTO SHALLOWER


LIQUID
Let us examine the motion of waves towards a gradually slopping beach. For
these waves the wave period T cannot change, even though the depth of liquid
decreases. This is because the number of waves passing a fixed position in
unit time interval is 1/T , and if this varies from one position to another
then the number of waves entering a given region would increase or decrease
indefinitely. But in the basic relation, between the wavelength λ, velocity C
and wave time T ,
λ=CT (7.64)
wavelength λ can be eliminated from Equation (7.53), and for constant T ,
dC/dh can be shown to be positive. Therefore, a decrease of depth h leads
to a reduction of velocity C. This reduction of wave velocity may produce
refraction effects similar to those in optics. For example, a uniform train
of waves approaching a beach obliquely is deflected so that the wave crests
become more nearly parallel to the contour lines of the bed, as shown in Figure
7.45.
If refraction causes horizontal lines perpendicular to the wave crests to
converge, the wave energy is constricted to a passage of decreasing width and
thus the wave amplitude increases. Due to this large waves are often found at
headlands at the sides of a bay. When the depth becomes little more than the
amplitude, the Airy theory ceases to hold and the wave profile is increasingly

ISTUDY
7.19. WAVES MOVING INTO SHALLOWER LIQUID 473

Wave crests

Bed contours

Shallower water
FIGURE 7.45 An uniform train of waves approaching a beach.

distorted. The crests become sharper and the troughs flatter. Further, the
velocity of propagation of the upper part of the profile is greater than that
of the lower part, owing to this the crests curl forwards and finally break.
Breaking usually occurs when a ≈ 34 h.

7.19.1 Standing Waves


When two trains of wave of equal amplitude, wavelength and period, but
traveling in opposite directions, are combined, the result is a set of standing
or stationary waves. This can happen when a series of waves is reflected by
a fixed solid boundary perpendicular to the direction of propagation; the two
individual wave trains are then formed of the incident waves traveling with
velocity C and the reflected waves with velocity −C. For example, if the
individual waves are small-amplitude sine waves, the amplitude of the free
surface for the resulting standing wave is
η = a sin [m(x − Ct)] + a sin [m(x + Ct)]
or
η = 2a sin mx cos mCt (7.65)
That is, any instant the free surface is a sine curve but its amplitude 2 a cos mCt
varies continuously with time. The values of x that give η = 0 are indepen-
dent of t. Thus the wave profile does not travel over the surface, instead it
simply rises and falls as indicated by the dotted lines in Figure 7.46.
The position with η = 0 are called nodes and those with maximum vertical
motion are called antinodes.
From Equation (7.46),
A = −C a cosech [m(h + a)]
where a denotes the maximum of η. Substituting this into Equation (7.54),
we get the horizontal velocity of an element in a single moving wave is
m C a cosech [m(h + a)] cosech [m(h + z)] sin [m(x − Ct)]

ISTUDY
474 CHAPTER 7. FLOW WITH FREE SURFACE

Antinode

Node
Vertical wall

FIGURE 7.46 Rise and fall of waves.

Adding the value for a wave of equal amplitude traveling with velocity −C,
to this, we get the velocity of a particle in a standing wave as
m C a cosech [m(h + a)] cosech [m(h + z)] × (sin [m(x − Ct)] − sin [m(x + Ct)])
= −2 m C a cosech [m(h + a)] cosech [m(h + z)] cos mx sin mCt
Similarly the vertical velocity can be expressed as
−2 m C a cosech [m(h + a)] sinh [m(h + z)] sin mx sin mCt
As the ratio of horizontal and vertical velocity components is independent
of t, all the particles which are close to the mean position (x, z) move to
and fro in a straight line, the direction of which varies from vertical beneath
the antinodes (where cos mx = 0) to horizontal beneath the node (where
sin mx = 0).
When the liquid is confined in a channel with completely closed vertical
ends, then only certain values of wavelength are possible. The zero velocity
at the closed ends implies that each end coincides with an antinode and so
the length of the channel must be an integral number of half-wavelengths. If
one end of the channel is closed and the other end is connected to an infinite
expanse of liquid, then the open end must be a node (that is, no vertical
moment) and the length of the channel must be an odd number of quarter-
wavelengths. A bay may behave like a channel open at one end if waves or
tides from the open sea arrive with a frequency equal to that at which standing
waves oscillate. The water in the bay then set into resonance and very high
amplitudes may occur at the inner shore even though there is only moderate
vertical movement at the mouth of the bay.
Lakes or harbors may behave like channels with closed ends. For example,
wind action may move water towards one end, and, after the wind drops, the
water may oscillate for a considerable time. Such standing waves are usually
termed seiches. On the surface the horizontal movement at the nodes may be
several times the vertical movement at the antinodes.

ISTUDY
7.20. SUMMARY 475

Cyclical or wave motion in the sea is established in three different ways.


Gravitational forces have a strong influence on all three, but they are dis-
tinguished by the fact that certain features of the motion differ by orders of
magnitude. The motion of the tides in the sea is regular, causing raising and
lowering of sea level, with periodic time of about 12 hours, caused mainly
by the gravitational pull of the moon, with the sun and planets also making
smaller contributions. Also, there are surface waves, which are mainly gen-
erated by the movement of wind over the sea surface, although the motion
of ships and other surface distributions also contribute to this form of wave
motion, The third kind of wave motion is termed tsunami. The tides and
wind-generated waves are always present, but tsunami occurs only rarely.
Tsunami is a Japanese word, meaning harbor wave. A tsunami is a natural
phenomenon, consisting of a series of waves generated at sea as a result of a
triggering event. Once these waves have been set in motion they are very
persistent, eventually arriving at a coastline, and it is there the full effect of
a tsunami is manifested. A small tsunami, triggered by an earth tremor, is
scarcely persistable and is only detectable with sensitive measuring equipment.
A tsunami can cause severe damage to humans as well as properties. For
example, the tsunami caused by the strong earth quake (of magnitude 9 in a
10 point scale) off the east coast of Japan on March 11, 2011, wiped off a vast
area in the east coast, completely damaging everything.
Tsunami can be caused by the following four reasons; (i) the earthquakes,
(ii) a major landslide, (iii) eruption of a volcano and (iv) the decent of a
huge object, such as a meteorite or asteroid, into the sea. A tsunami caused
by an earthquake owing to the relative movement of two tectonic plates along
a fault line, one plate passing beneath the other in a process known as sub-
duction. This movement causes the sudden release of an enormous amount
of strain energy, resulting in a substantial deformation on the sea bed, some
parts elevating, others subsiding. As a consequence, a large volume of water is
impulsively displaced and the local equilibrium of the surface of the sea is de-
stroyed. The water mass displaced, due to the lifting of sea floor, accumulates
potential energy under the influence of gravity. Now the process to establish a
new equilibrium configuration, the potential energy is transferred into kinetic
energy and a system of wave is set in motion. A tsunami is associated with
the movement of a large mass of water. The wave front created in the early
stages of a tsunami has, in general, a complex geometry. For an earthquake
concentrated at a single point beneath the surface of the sea, the waves would
radiate away from the source across the surface of the sea in circles.

7.20 SUMMARY
When the liquid is bounded by side walls - such as the banks of a river or
canal - the flow is said to take place in an open channel. The free surface
is usually subjected only to atmospheric pressure and, since this pressure is
constant, the flow is caused by the weight of the fluid. As in pipes, uniform

ISTUDY
476 CHAPTER 7. FLOW WITH FREE SURFACE

flow is accompanied by a drop in piezometric pressure, (p + ρgz), but for an


open channel it is only the second term, ρgz, that is significant, and uniform
flow in an open channel is always accompanied by a fall in the level of the
surface.
Natural streams and rivers, artificial canals, irrigation canals are typical
examples of open channels. But pipe-lines or tunnels that are not completely
full of liquid also have the essential features of open channels. Water is the
liquid usually involved, and practically all the experimental data for open
channels relate to water at normal temperature.
A open channel flow may be uniform or non-uniform, steady or unsteady.
A channel flow is termed uniform when the magnitude and direction of velocity
of the liquid remains invariant throughout the channel.
Flow in which the liquid surface is not parallel to the base of the channel
is said to be non-uniform. Flow with rapid and gradual change in depth are
referred to as rapidly varied flow and gradually varied flow, respectively.
When the velocity, and hence the depth, at a particular point in the chan-
nel varies with time the flow is termed unsteady. When the velocity remains
invariant with time the flow is termed steady.
A steady uniform flow, in which the depth of the liquid changes neither
with distance along the channel nor with time, can easily be solved analyti-
cally.
In practice, non-uniform flow is encountered more frequently than uniform
flow. This is especially true for short channels because a certain length of
channel is required for the establishment of uniform flow.
Like pipe flow, the flow in an open-channel may be either laminar or
turbulent, depending on the relative magnitude of viscous and inertia forces,
the Reynolds number ul/ν. The lower critical Reynolds number for open
channels seldom occurs in problems of practical interest. Indeed, in channels
of engineering interest, completely turbulent flow may invariably be assumed
because the surface of the flowing liquid occasionally appears smooth and
glossy is no indication that turbulent flow does not exist underneath.
Another important classification of open-channel flow is derived from the
magnitude
√ of the Froude number of the flow. The Froude number (Fr =
V / gh) is then less than 1.0, and the flow is termed tranquil. When the
velocity of the flow is high that a small disturbance cannot propagate upstream
(but is washed downstream), the Froude number becomes greater than 10 and
the flow is said to be rapid. When the Froude number is equal to 1.0 the flow
is termed critical.


u=C mi

is known as Chézy’s equation.


The discharge through the channel is given by

Q = Au = AC m i

ISTUDY
7.20. SUMMARY 477

Manning’s expression for u is

m2/3 i1/2
u=
n

The units of u in this equation is m/s. The n in this equation is known as


Manning’s roughness coefficient.
In an alluvial channel, with its surfaces composed of movable sand or
gravel, the roughness elements are not permanent, but depend on the flow.
Ripples and dunes may form in the boundary material, and the spacing of
these humps may be much larger than the spacing of the irregularities on the
walls of pipes to or rigid boundaries of open channels.
Chézy formula and Manning formula show that, for any given value of
slope, surface roughness and cross-sectional area, the discharge Q increases
with increase in the hydraulic mean depth m.
It may be seen that, of all sections whose sides do not slope inwards towards
the top, the semicircle has the maximum hydraulic mean depth. But, though
semicircular channels are built from prefabricated sections, for other forms
of construction the semicircular shape is impractical. Trapezoidal sections
are very popular, but when the sides are made of loose granular material its
angle of repose may limit the angle of the sides. The most efficient section
will be that with the maximum discharge for a given area and, conversely, the
minimum area for a given discharge.
Partially filled flow through closed conduits are encountered in civil engi-
neering practice. For partially filled flow through circular conduits the area
of the cross-section of the liquid is
( ) ( )
1 1
r2 θ − 2 r sin θ r cos θ = r2 θ − sin 2θ
2 2

Flow in open channels may be modified by waves and surges of various


kinds that produce unsteady conditions. If the flow along a channel is in-
creased or decreased by removal or insertion of an obstruction, such as sudden
opening or closing of a sluice gate, surge waves are formed and propagated
upstream and downstream of the obstruction. In certain circumstances, tidal
action may cause a surge, known as a bore, in large rivers.
In the lab experimental studies with small scale models, another type of
wave termed capillary wave, formed due to the influence of surface tension is
of significance.
For waves on the surface of deep water, such as the sea, different consider-
ations apply. This is mainly due to the assumption of a uniform distribution
of velocity over the entire depth is not valid for large depths.
The total head is given by

p u2
+ +z
ρg 2g

ISTUDY
478 CHAPTER 7. FLOW WITH FREE SURFACE

where z is the height of the point considered above arbitrary horizontal datum
plane.
The ratio of the √ mean flow velocity to the velocity of propagation of a
small disturbance u/ gh is of the same form as the Froude number. If the
mean depth h is used as the characteristic length, and the mean velocity is
used as the characteristic velocity, then

u
√ = Fr
gh

Thus, for critical condition



u= gh and Fr = 1

That is, critical Froude number is unity.


The value of the slope for which critical uniform flow is achieved is known
as the critical slope. Using the Chézy formula, we obtain the critical velocity
as √ √
uc = gh = C m i
In uniform flow the energy gradient i and the slope of the bed s are equal and
therefore the critical slope sc is defined by


gh = C msc

A slope less than the critical slope sc is known as mild slope, and a slope
greater than the critical slope is termed steep slope.
Flow with velocity less than the critical velocity is referred to as tran-
quil, flow with velocity greater than the critical velocity is called rapid or
shooting. The tranquil and rapid flows are also referred to as subcritical and
supercritical flows, respectively. It is important to note that sub-critical veloc-
ity corresponds to a depth greater than the critical depth, and supercritical
velocity corresponds to a depth smaller than the critical depth.
( )2
e h 1 hc
= +
hc hc 2 h

This is the dimensionless form of specific-energy relation.


A stationary surge wave, through which the depth of flow increases is
known as hydraulic jump. A hydraulic jump is essentially an abrupt change
from rapid to tranquil flow. The depth of the liquid is less than the critical
depth, hc , before the jump and greater than hc after it.
A gradual transition from rapid to tranquil flow is not possible. As the
depth of rapid flow increases, the lower limit of the specific-energy diagram
(Figure 7.20) is followed from right to left, that is, the specific-energy de-
creases. If the increase has to continue to the critical value, an increase in

ISTUDY
7.20. SUMMARY 479

specific-energy would be required for a further increase of depth to the value


corresponding to uniform flow diagram. In uniform flow the specific energy
remains constant and the total energy decreases at a rate exactly correspond-
ing to the slope of the bed. For any depth less than that for uniform flow the
velocity is higher, due to friction consumes energy at a greater rate than the
loss of gravitational energy.
There are two solutions for h2 . But the solution with negative sign for the
second term on the right-hand-side is physically impossible, since h2 cannot
be negative. Hence,
√( )
h1 h21 2q 2
h2 = − + +
2 4 gh1

The depths h1 and h2 ahead of and behind a hydraulic jump are termed
conjugate depths for the jump.
The head loss across a jump hj can be expressed as

3
(h2 − h1 )
hj =
4h1 h2

The head loss, hj , across a jump is primarily due to the turbulence in the
wave because friction at the boundaries makes a negligible contribution to
the energy loss.
The energy dissipation across a jump is always finite. Indeed, the hydraulic
jump is an effective means of reducing unwanted energy in a stream.
The flow process across a hydraulic jump is an irreversible process. Jumps

causing h2 /h1 < 1 are termed small jumps. This corresponds to Fr1 < 3,
for rectangular sections. For small jumps the surface does not rise abruptly
but passes through a series of undulations gradually diminishing in size. Such
a jump is known as an undular jump. For larger values of h2 /h1 and Fr1 , the
jump is termed direct and across a direct jump the surface would rise abruptly.
Assuming the bed as horizontal and the flow be of steady and uniform
velocity over the cross-section, and uniform depth across the width and neg-
ligible boundary friction, the momentum equation can be written as
ρgh1 ρgh2
h1 − h2 − F = ρg(u2 − u1 )
2 2
for unit width of uniform rectangular channel. This equation is true whatever
the values of h1 and h2 in relation to the critical depth.
Weir is an obstruction which extends across the full width of the stream.
In other words, a weir is a raised portion of the bed, which runs across the
complete width of the stream.
As a measuring device the broad-crested weir has the following advantages.
(i) It is simple to construct and has no sharp edge that can wear and thus
alter the discharge characteristics with time.

ISTUDY
480 CHAPTER 7. FLOW WITH FREE SURFACE

(ii) It does not cause any appreciable raising of the surface level upstream, and
the results are not affected by conditions downstream provided that critical
flow occurs over the crest.
In spite of these advantages, the limitation of the theoretical analysis
makes it unwise to use it, for any accurate measurement of flow rate. Thus
broad-base weir are suitable only for approximate measurement of flow rate.
Critical flow or maximum discharge over a weir in a channel takes place
only in the absence of greater restrictions downstream. If the depth down-
stream of the weir is sufficiently increased due to the presence of an obstruction
downstream, the flow over the weir may not be critical, and the weir is then
said to be drowned. When a broad-crested weir is drowned a depression of
the surface over the crest still occurs, but it is not sufficient for the critical
depth to be reached.
Measurement of flow rate with Venturi has the following advantages over
a broad-crested weir.
• The loss of head experienced by the liquid in passing through the Venturi
is significantly less than that encountered while passing over a broad-
crested weir.
• Operation of the Venturi is not affected by the deposition of silt.
The mid-section of the Venturi with minimum cross-sectional area is known
as the throat, and the discharge per unit area of the throat is the maximum.
The flow at the throat is therefore critical. For a rectangular cross-section in
which the√streamlines are straight and parallel the velocity at the throat is
given by gh2 and the discharge becomes

Q = b2 h2 gh2
The discharge may be expressed as
( )3/2 ( )3/2
1/2 2 u21
Q = b2 g h1 +
3 2g
where b1 is very large, u1 (= Q/(b1 h1 )) is small and u21 /(2g) may be neglected.
Note that, the discharge expressions in Equations (7.36) and (7.37) are
derived assuming that the friction is absent. But is actual flow, friction is
finite. Therefore, the volume flow rate obtained with these relations should be
corrected for frictional losses. Usually a correction factor known as discharge
coefficient Cd is used to account for the friction between the inlet and throat
of the Venturi. In practice, value of Cd lies between 0.95 and 0.99. The
discharge may be written as
3/2
b2 g 1/2 h1
Q = Cd
(1.5)3/2

For a rectangular channel, the hydraulic jump occurs at a point where


the depth of the rapid flow is such as to give the correct depth of subsequent

ISTUDY
7.20. SUMMARY 481

tranquil flow. A limiting position is that in which the jump resulting in zero
height occurs at the throat itself. If the downstream level is raised further
the velocity at the throat no longer reaches the critical velocity and the flow
through the Venturi is termed drowned.
A properly designed Venturi under ideal conditions the loss of mechanical
energy may be kept as low as 10%. Like the broad-crested weir, the Venturi
is also suitable for measurement only when the flow is tranquil.
For a particular shape, there is only one depth at which uniform flow can
take place. This depth is known as the normal depth. But there are number
of ways in which the same steady rate of flow can take place along the same
channel in non-uniform flow.
Among the steady and non-uniform flows, there are two kinds in each
category. In one kind of non-uniform flow the changes of depth and velocity
takes place over a long distance. Such flow is termed gradually varied flow. In
the second type the changes of depth and velocity take place within a short
distance and may also be quite abrupt, as in a hydraulic jump. This kind of
non-uniform flow is termed rapidly varied flow.
The surface profile may have a variety of forms, depending on how the
flow is controlled by weirs and other obstructions, changes in bed slope and
so on. The slope of the bed may be adverse or uphill (A), zero or horizontal
(H) with s = 0, mild (M ) with s < sc , critical (C) with s = sc and steep
(S) with s > sc . The profiles are further classified according to the depth of
the stream. If the depth is greater than the normal depth h0 and the critical
depth hc (that is, h > h0 and h > hc ), the profile is of type 1; for h0 < h < hc ,
the profile is of type 2; and if h < h0 and h < hc , the profile is of type 3. The
types of non-uniform flow possible are illustrated in Figure 7.40. The slope
of a channel may be classified as mild for one rate of flow, critical for another
and steep for a third.
For any point in the liquid the stream function ψ may be expressed as
ψ = Cz + f (z) sin mx
where z is the coordinate in the vertical direction, f (z) is a function of z, and
m = 2π/λ. The term Cz accounts for the uniform velocity.
For irrotational motion, ψ must satisfy Laplace equation
∂2ψ ∂2ψ
+ =0
∂x2 ∂z 2
Deep-water waves are those with h > λ/2.
Waves whose properties are determined by gravity effects are referred to
as gravity waves.
The long wave with λ ≫ h is usually termed shallow-water wave. However
the term long wave is more appropriate, because it more accurately reflects
the fact that waves of long wavelengths are included, irrespective of whether
the water is considered deep or shallow.
Waves whose characteristics are governed mainly by surface tension are
known as capillary waves. Waves of very short length are termed ripples.

ISTUDY
482 CHAPTER 7. FLOW WITH FREE SURFACE

The energy of a wave is the sum of the contribution from gravitational


energy, kinetic energy and free-surface energy.
The total energy divided by the width of the wave, over one wavelength is

1 2 ( )
a λ ρ g + σm2
2
This is valid for n ≪ h, η ≪ λ.
The shape of the wave moves with velocity C. But all the elements of a
liquid carrying energy do not move at velocity C. The quantity

1
p∗ + ρ(u2 + v 2 )
2
is a measure of the amount of energy carried by a small element of liquid
divided by the volume of the element. Here p∗ is the piezometric pressure and
u, v are the horizontal and vertical components of velocity, respectively.
When two trains of wave of equal amplitude, wavelength and period, but
traveling in opposite directions, are combined, the result is a set of standing
or stationary waves.
Lakes or harbors may behave like channels with closed ends. For example,
wind action may move water towards one end, and, after the wind drops, the
water may oscillate for a considerable time. Such standing waves are usually
termed seiches. On the surface the horizontal movement at the nodes may be
several times the vertical movement at the antinodes.
Tsunami is a Japanese word, meaning harbor wave. A tsunami is a natural
phenomenon, consisting of a series of waves generated at sea as a result of a
triggering event.
Tsunami can be caused by the following four reasons; (i) the earthquakes,
(ii) a major landslide, (iii) eruption of a volcano and (iv) the decent of a huge
object, such as a meteorite or asteroid, into the sea.

7.21 PROBLEMS
7.1 Water flows through a rectangular channel of width 2.5 m, energy gradient
0.0028, Manning’s roughness coefficient 0.014. If the flow depth is 300 mm,
find the flow rate.
[Ans. 1.101 m3 /s]

7.2 Water flows through a rectangular flume of width 1.5 m, made of timber
with n = 0.013. Find the necessary channel slope if the water has to flow
uniformly at depth of 0.6 m with velocity 4.5 m/s
[Ans. 0.01483]

7.3 If the critical velocity of water in a channel of width 4 m is 3 m/s,


determine the cross-sectional area of the channel and the volume flow rate of
water.
[Ans. 3.67 m2 , 11 m3 /s]

ISTUDY
7.21. PROBLEMS 483

7.4 A smooth-concrete channel 2.4 m wide has a slope of 5◦ and a weir depth
of 1.2 m. If the manning factor for the channel is 0.015, determine the volume
flow rate.
[Ans. 40.31 m3 /s]

7.5 A sluice gate across a 6 m wide channel, discharges a stream 1.2 m deep.
What will be the flow rate if the upstream depth is 6 m?
[Ans. 69.9 m3 /s]

7.6 A Venturi flume with a level bed is 12 m wide and 1.5 m deep upstream
with a throat width of 6 m. Assuming that a standing wave forms downstream
calculate the rate of flow of water.
[Ans. 185.4 m3 /s]

7.7 A rectangular channel has to be designed to carry water at the rate of 2


m3 /s with a slope of 0.002. Determine the width and height if the height to
width ratio has to be 0.5, assuming the manning roughness constant as 0.013.
[Ans. Width = 1.496 m, Height = 0.748 m]

7.8 Water flows through a 5 m wide rectangular channel with Manning’s


roughness coefficient 0.015. The channel has a constant slope of 0.02. At one
cross section the speed is 4 m/s and depth is 1.5 m. Determine the distance
between the locations with depths 1.5 m and 0.9 m.
[Ans. 53.7 m]

7.9 In a rectangular channel the specific energy of flow is 1.8 J/N. Determine
the critical depth and the rate of discharge per unit width.
[Ans. 1.2 m, 4.12 m2 /s]

7.10 A rectangular channel of width 3.6 m carries water at 2.5 m3 /s. (a)
What will be the critical depth. (b) If the depth is reduced to half of the
critical depth check whether a hydraulic jump will be formed. (c) If so find
the depth after the jump.
[Ans. (a) 0.366 m, (b) hydraulic jump will be formed, (c) 0.647 m]

7.11 Water flows in a channel of trapezoidal section with a velocity of 4.5


m/s and a depth of 0.6 m at the centre. If the base width is 1.2 m and the
side slopes are 1 vertical and 2 horizontal, determine (a) the specific energy
and (b) the critical depth.
[Ans. (a) 1.632 J/N, (b) 1.26 m]

7.12 Water flow in a wide channel approaches a 100 mm high bump at 1.5
m/s and a depth of 1 m. (a) Find whether the flow is supercritical. Estimate
(b) the water depth over the bump and (c) the bump height which will cause
the crest flow to be critical.
[Ans. (a) No, (b) 0.93 m, (c) 0.612 m]

ISTUDY
484 CHAPTER 7. FLOW WITH FREE SURFACE

7.13 Water flows through a rectangular channel at a flow rate 20 m3 /s and


velocity 5 m/s. If the depth of this steady flow is 3 m, determine the critical
velocity.
[Ans. 5.29 m/s]

7.14 In a rectangular channel of width 4 m, water flows at a rate of 5


m3 /s. The slope of the channel is 0.0001 and the friction coefficient is 0.005.
Determine the depth of this steady flow.
[Ans. 2 m]

7.15 Water flows at a rate of 20 m3 /s in a 4 m wide rectangular channel with


manning roughness coefficient of 0.012. If the flow depth is 1.8 m, determine
the slope of the channel.
[Ans. 0.0012]

7.16 A hydraulic jump is encountered by a uniform flow of depth 2 m. If the


Froude number downstream of the jump is 0.4, determine the Froude number
upstream of the jump.
[Ans. 3.105]

7.17 A hydraulic jump occurs on a horizontal bed. If the velocity and


depth upstream of the jump are 5 m/s and 0.06 m, determine (a) the depth
downstream of the jump and (b) the head loss across the jump.
[Ans. (a) 0.524 m, (b) 0.794 m]

7.18 A 3.2 m wide rectangular channel has a horizontal bed. It is desired


to measure the flow through it by a Venturi. If the depth upstream is 1.8 m,
when the flow rate is 8 m3 /s, what should be the width of the Venturi throat?
[Ans. 1.8 m]

7.19 Water flows in a wide channel at 10 m2 /s, per unit width, undergoes a
hydraulic jump. If the flow depth upstream of the jump is 1.25 m. Determine
(a) depth and Froude number downstream of the jump and (b) the power
dissipated per unit width.
[Ans. (a) 3.463 m, 0.496, (b) 6.26 kW/m]

7.20 A hydraulic jump occurs in a rectangular channel carrying water at 8


m3 /s with a depth of 1.8 m. If the power loss associated with the jump is 9
kW, determine the depth downstream of the jump.
[Ans. 5.3 m]

7.21 A weir in a horizontal channel is 1 m high and 4 m wide. The water


depth upstream is 1.6 m. If the discharge is 2.53 m3 /s, determine the discharge
coefficient of the weir.
[Ans. 0.8]

7.22 Water flows through a 4 m wide rectangular channel with energy gra-
dient 0.0009 and manning’s roughness coefficient 0.018. If the flow rate is 7
m3 /s, determine the depth of flow.
[Ans. 1.25 m]

ISTUDY
7.21. PROBLEMS 485

7.23 The sides of a horizontal channel are contracted to form a Venturi flume.
The width of the channel is 1.2 m and the throat is 0.6 m. The hydraulic
jump on the downstream side ensures that the flow has maximum value. The
depth of the water on the upstream side is 0.6 m. Determine the discharge
and the depth of flow at the throat.
[Ans. 0.505 m3 /s, 0.416 m]

7.24 If 12 m2 /s of water per unit width flows down a spillway into a horizontal
channel as a uniform flow of velocity 20 m/s, determine the downstream depth
required to cause a hydraulic jump and loss of power caused by the jump.
[Ans. 6.7 m, 1662.2 kW]

7.25 A 6 m wide rectangular channel carries 22.5 m3 /s of water. Determine


the necessary slope depth to achieve uniform flow at (a) a depth of 3 m and
(b) the critical depth, assuming the manning roughness coefficient as 0.012.
[Ans. (a) 1/7631, (b) 1.127 m]

7.26 In a 0.6 m wide rectangular channel, a jump occurs where the Froude
number is 3. The depth after the jump is 0.6 m. Find (a) the total loss and
(b) the power dissipated in the jump.
[Ans. (a) 0.225 m, (b) 0.79 kW]

7.27 A Venturi flume with 12 m level bed and throat 6 m wide carries 18.54
m3 /s. If the flow upstream of the throat is 1.5 m deep find the velocity and
Froude number at the throat.
[Ans. 3.1 m/s, 3.11]

7.28 A stream issuing beneath a vertical sluice gate is 0.3 m deep at the vena
contracta. Its mean velocity is 6 m/s. A standing wave is created on the level
bed below the sluice gate. Find (a) the height of the jump, (b) the loss of
head and (c) the power dissipated per unit width.
[Ans. (a) 1.04 m, (b) 0.7 m, (c) 12.36 kW]

7.29 A sluice across a 6 m wide channel discharges a stream 1.2 m deep.


What will be the flow rate if the upstream depth is 6 m?
[Ans. 69.9 m3 /s]

7.30 A frictionless channel of constant width sloping downwards at 1 in 20


is fed by a sluice as broad as the channel and 1.6 m open. If the discharge
coefficient is 0.6, determine the depth (a) at 15 m and (b) 50 m, from the
sluice when the upstream stagnation level is 6.5 m.
[Ans. (a) 0.9 m, (b) 1.29 m]

7.31 A long wide channel of bed slope 1:2000 has a Chézy’s coefficient of 45
m1/2 /s. A flow of 5 m3 /s per meter is maintained at normal depth by a sluice
gate at the downward end. If the Manning’s coefficient of the bed is 0.042,
determine the opening of the gate.
[Ans. 0.735 m]

ISTUDY
486 CHAPTER 7. FLOW WITH FREE SURFACE

7.32 Determine the maximum depth in a 3 m wide rectangular channel if


the flow is to be supercritical with a flow rate of 60 m3 /s.
[Ans. 3.44 m]

7.33 A trout jumps, producing waves on the surface of a 0.8 m mountain


stream. If it is observed that the waves do not travel upstream, what is the
minimum velocity of the current.
[Ans. 2.8 m/s]

7.34 A 5 m wide
√ rectangular channel carries water at a depth of 1 m and
Froude number 10. If the flow undergoes a hydraulic jump, determine
Froude number after the jump and the energy dissipation associated with
the jump.
[Ans. 0.395, 182.17 kW]

7.35 Water flows up a 150 mm tall ramp in a constant width rectangular


channel at 0.5 m2 /s. If the upstream depth is 700 mm, determine the elevation
of water surface down of the ramp, neglecting the viscous effects.
[Ans. 0.68 m]

ISTUDY
Chapter 8

Hydraulic Machinery:
Pumps and Turbines

8.1 INTRODUCTION
Hydraulic machinery are essentially devices to convert one form of fluid energy
to another. There are various types of fluid machinery. They are broadly
classified as pumps, turbines, blowers, fans and compressors. The devices
which convert mechanical energy to fluid energy are called pumps. The devices
meant for converting fluid energy to mechanical energy are termed turbines.
The conversion of mechanical energy to fluid energy is accomplished by pumps
for the case of incompressible fluids, and by blowers, fans, and compressors
for the case of compressible fluids.
Hydraulic machines fall into two categories: (i) positive displacement pis-
ton and cylinder machines which are not suitable for handling large quantities
of fluid but are important in hydraulic control systems and (ii) turbines or
rotodynamic machines. The common factor in all rotodynamic machines is
that the fluid is fed to the runner or rotating element continuously in such a
way that it has a tangential velocity component (or velocity of whirl) about
the axis of the shaft as it enters the runner and emerges radially or axially
having lost its tangential momentum and exerted a torque on the runner in
the process.

8.2 PUMPS
Pumps are broadly classified into centrifugal pumps and axial-flow pumps. The
rotating element of a centrifugal pump is called the impeller. The impeller may
be shaped to force water outward in a plane at right angles to its axis (radial
flow), to give the water an axial as well as radial velocity (mixed flow), or to
induce a spacial flow on coaxial cylinders in an axial direction (axial flow).
Radial-flow and mixed-flow machines are commonly referred to as centrifugal

487

ISTUDY
488 CHAPTER 8. HYDRAULIC MACHINERY

pumps. Radial-flow and mixed-flow impellers may be either open or closed.


The open impeller consists of a hub to which vanes are attached, while the
closed impeller has plates (or shrouds) on each side of the vanes. Though the
efficiency of open impeller is less than the efficiency of the closed impeller,
the open impeller is less likely to become clogged and hence is suitable for
handling liquid containing solid particles (that is, liquid contaminated with
impurities in the form of solid particles).
Radial-flow pumps are provided with a spacial casing, often referred to
as a volute casing, as shown in Figure 8.1, which guides the flow from the
impeller to the discharge pipe.
The continuously increasing cross-section around the casing tends to main-
tain a constant velocity within the casing. This helps to provide relatively
smooth flow conditions at exit of the impeller. Some pumps have diffuser vanes
instead of a volute casing. Such pumps are known as turbine pumps. Some
radial pumps are of the double-suction type. They have identical, mirror-
image impellers placed back to back. Water enters the pump from both sides
and is discharged into a volute casing or diffuser vanes. The advantage of the
double-suction pump is the reduced mechanical friction that results because
the thrust on the bearing is balanced.

r1

V1
w1 vn1
α1
β1 α1 vn1 w1 r2
u1
Vt1

Runner blade
exit

FIGURE 8.1 Radial-flow centrifugal pump with volute casing.

The velocity polygon shown in Figure 8.1 introduced the blade, flow ve-
locity and angles. In the idealized situation, the flow relative to the rotor
is assumed to enter and leave tangential to the blade profile at each section.
This idealized inlet condition is called shockless entry flow. Blade angles, β,
are measured relative to the circumferential direction. The inlet blade angle,
β1 , fixes the direction of the relative inlet velocity.
The runner speed is u1 = ω r1 , where ω is the machine operating speed
and r1 is the inlet radius. The absolute velocity of the flow, V1 , is the vector
sum of impeller velocity, u1 , and the flow velocity relative to the blade, w1 .
The angle of absolute velocity, α1 , is measured from the normal direction, as
shown in Figure 8.1.

ISTUDY
8.2. PUMPS 489

The tangential component of velocity, Vt1 , and the component of velocity


normal to the flow area, Vn1 , are also shown in Figure 8.1. Note that, at
exit section the normal component of absolute velocity, Vn and the normal
component of velocity relative to the blade, w, are equal.
When the inlet flow is swirl-free, the absolute inlet velocity will be purely
radial. The inlet blade angle may be specified for the design flow rate and
pump speed to provide shockless entry flow. If there is swirl in the flow, the
absolute inflow direction will not be in the radial direction.
Velocity polygon at the outlet section may be controlled in a similar man-
ner to the inlet section. The runner speed at the outlet is u2 = ω r2 .
Typical centrifugal-flow and axial-flow pump installations are shown in
Figure 8.2. Pumps can be of single-stage or multistage. A single-stage pump
has one impeller, while a multistage pump has two or more impellers arranged
in such a way that the discharge from one impeller enters the eye of the next
impeller.

Motor

Pump Check valve Shaft


Gate valve

Guide vane

Propeller blade

Foot valve

Strainer

(a) (b)
FIGURE 8.2 Typical (a) centrifugal and (b) axial-flow pump
installations.

Deep-well pump, shown in Figure 8.3, a type of turbine pump, are usually
multistage, having several impellers on a vertical shaft suspended from a prime
mover, usually an electric motor, located at the ground surface. Each impeller
discharges into a fixed-vane diffuser, or bowl, coaxial with the drive shaft,
which directs water to the next impeller.
For best efficiency of a centrifugal pump, proper arrangement of the suction
and discharge piping is necessary. For economy, the diameter of the pump
casing at suction and discharge is often smaller than that of the pipe to which
it is attached. If there is a horizontal reducer between the suction and the
pump, an eccentric reducer, shown in Figure 8.2(a), has to be used to prevent
air accumulation. A foot valve (check valve) can be installed in the suction
pipe to prevent water from leaving the pump when it is stopped. The discharge
pipe is usually provided with a check valve and a gate value. The check valve
prevents back-flow through the pump when there is a power failure. Suction

ISTUDY
490 CHAPTER 8. HYDRAULIC MACHINERY

pipes taking water from a sump or reservoir are usually provided with a screen
to prevent entry of debris that might clog the pump.

FIGURE 8.3 Deep-well multistage mixed-flow turbine pump.

Axial-flow pumps, shown in Figure 8.2(b), usually have only two or four
blades and hence, large unobstructed passages that permit handling of water
containing debris without clogging. The blades of some large axial-flow pumps
are adjustable to permit the setting of pitch for the best efficiency under
existing conditions.

8.3 HEAD DEVELOPED BY A PUMP


The pressure head, h, developed by a pump is determined by measuring the
pressures on the suction and discharge sides of the pump, computing the
velocities by dividing the measured discharge by the respective cross-sectional
areas, and noting the difference in elevation between the suction and discharge
sides. The head, h, delivered by the pump to the fluid is
( ) ( )
pd V2 ps V2
h = Hd − Hs = + d + zd − + s + zs (8.1)
ρg 2g ρg 2g

where the subscript d and s, respectively, refer to the discharge and suction
sides of the pump, as illustrated in Figure 8.4. Usually the intake is larger
than the discharge pipe.
Flow conditions at the discharge flange are usually too irregular for accu-
rate pressure measurement, but the pressure can be measured reliably at 10 or
more pipe diameter away from the pump and the friction head for that length
of pipe can be estimated and added to the measured pressure. On the intake
side, pre-rotation some times exits in the pipe near the pump, and this will

ISTUDY
8.3. HEAD DEVELOPED BY A PUMP 491

Vd2 /2g
EL
HGL
Vd2 /2g
pd
pd γ
γ

d
s

zs Suction ps
zd
lift γ
Datum
EL
HGL

Vs2 /2g

FIGURE 8.4 Head developed by a pump (in this case ps /(ρ g)) is
negative, in this figure HGL is hydraulic grade line and EL is energy line.

cause the pressure reaching in a gauge to be different from the true average
pressure at that section.
Usually V1 and V2 are about the same, z2 − z1 is no more than a meter or
so, and the net pump head is essentially equal to the change in pressure head
p2 − p1 ∆p
H≈ = (8.2)
ρg ρg
The power delivered to the fluid simply equals the specific weight times the
discharge times the net head change
Pw = ρgQH (8.3)
This is traditionally called the water horsepower. The power required to drive
the pump is the brake horsepower.
bhp = ωT (8.4)
where ω is the shaft angular velocity and T the shaft torque. If there were no
losses, Pw and brake horsepower would be equal, but of course Pw is actually
less, and the efficiency, η, of the pump is defined as
Pw ρgQH
η= = (8.5)
bhp ωT
The aim of the pump designer is to make the efficiency as high as possible
over as broad a range of discharge Q as possible.
The efficiency is basically composed of three parts: volumetric, hydraulic,
and mechanical. The volumetric efficiency is
Q
ηv = (8.6)
Q + QL

ISTUDY
492 CHAPTER 8. HYDRAULIC MACHINERY

where QL is the loss of fluid due to leakage in the impeller-casing clearances.


The hydraulic efficiency is
hf
ηh = 1 − (8.7)
hs
where hf has three parts: (1) shock loss at the eye due to imperfect match
between inlet flow and the blade entrances, (2) frictional losses in the blade
passages, and (3) circulation loss due to imperfect match at the exit side of
the blades.
The mechanical efficiency is
Pf
ηm = 1 − (8.8)
bhp
where Pf is the power loss due to mechanical friction in the bearings, packing
glands, and other contact points in the machine.
The total efficiency is given by

η = ηv ηh ηm (8.9)

The elementary theory of pump performance is developed assuming the


flow as one-dimensional and combining the idealized fluid-velocity vectors
through the impeller with the angular-momentum theorem for a control vol-
ume.
The idealized velocity diagrams are shown in Figure 8.5. The fluid is
assumed to enter the impeller at r = r1 with velocity component w1 tangent
to the blade angle β1 plus the circumferential speed u1 = ωr1 matching the
tip speed of the impeller. Its absolute entrance velocity is thus the vector sum
of w1 and u1 , shown as V1 . Similarly, the flow exits at r = r2 with component
w2 parallel to the blade angle β2 plus tip speed u2 = ωr2 , with resultant
velocity V2 .
Applying the angular-momentum theorem, it can be shown that the ap-
plied torque T is given by

T = ρQ(r2 Vt2 − r1 Vt1 ) (8.10)

where Vt1 and Vt2 are the absolute circumferential velocity components of the
flow.
The power delivered to the fluid is thus

Pw = ωT = ρQ(u2 Vt2 − u1 Vt1 )

The power can also be expressed as

Pw = ρgQH

where H is the head without loss, called Euler head. Thus,


Pw 1
H= = (u2 Vt2 − u1 Vt1 ) (8.11)
ρgQ g

ISTUDY
8.3. HEAD DEVELOPED BY A PUMP 493

V2 Vn2

u2 = ωr2
w2 α2
β2
V t2

Blade w1
V1
r2 β1
α1 Vn1

Vt1 u1 = ωr1

r1

FIGURE 8.5 Inlet and exit velocity diagrams for an idealized pump
impeller.

These are the Euler turbomachine equations, showing that the torque, power,
and ideal head are functions only of the rotor-tip velocities u1 , u2 and the
absolute fluid tangential velocities Vt1 , Vt2 , independent of the axial velocities
(if any) through the machine.
Additional insight is gained by rewriting these relations in another form.
From Figure 8.5, we have
V 2 = u2 + w2 − 2uw cos β
But
w cos β = u − Vt
Therefore,
V 2 = u2 + w2 − 2u(u − Vt )
= u2 + w2 − 2u2 + 2uVt
= w2 − u2 + 2uVt
that is,
1 2
uVt = (V + u2 − w2 ) (8.12)
2
Substituting this into Equation (8.11), we get
1 [ 2 ]
H= (V2 − V12 ) + (u22 − u21 ) − (w22 − w12 ) (8.13)
2g
Thus the ideal head relates to the absolute plus the relative kinetic-energy
change of the fluid minus the rotor-tip kinetic-energy change. Finally, substi-
tuting for H from its definition in Equation (8.1) and rearranging, we obtain
the classic relation

ISTUDY
494 CHAPTER 8. HYDRAULIC MACHINERY

p w2 r2 ω 2
+z+ − = constant (8.14)
ρg 2g 2g
This is the Bernoulli equation in rotating coordinates and applies to either
two- or three-dimensional ideal incompressible flow.
For a centrifugal pump, the power can be related to the radial velocity
Vn = Vt tan α and the continuity relation

Pw = ρQ (u2 Vn2 cot α2 − u1 Vn1 cot α1 ) (8.15)

where
Q Q
Vn1 = , Vn2 =
2πr1 b1 2πr2 b2
and where b1 and b2 are the blade widths at inlet and exit, respectively.
With the pump parameters r1 , r2 , β1 , β2 and ω known, Equation (8.11) or
Equation (8.15) is used to compute idealized power and head versus discharge.
The ‘design’ flow rate Q∗ is commonly estimated by assuming that the flow
enters exactly normal to the impeller

α1 = 90◦ , Vn1 = V1 (8.16)

8.3.1 Effect of Blade Angle on Pump Head


The simple theory above can be used to predict an important blade-angle
effect. If we neglect inlet angular momentum, the theoretical water horsepower
is

Pw = ρQu2 Vt2 (8.17)


where
Q
Vt2 = u2 − Vn2 cot β2 , Vn2 =
2πr2 b2
The theoretical head then becomes (from Equation (8.11))

u22 u2 cot β2
H≈ − Q (8.18)
g 2πr2 b2 g

The head varies linearly with discharge Q, having a shutoff value u22 /g, where
u2 is the exit blade-tip speed. The slope is negative if β2 < 90◦ (backward-
curved blades) and positive for β2 > 90◦ (forward-curved blades). This effect
is shown in Figure 8.6 and is accurate only at low flow rates.
The positive-slope condition in Figure 8.6 can be unstable and can cause
pump surge, an oscillatory condition where the pump ‘hunts’ for the proper
operating point. Surge may cause only rough operation in a liquid pump, but
it can be a major problem in gas-compressor operation. For this reason a
backward-curved or radial blade design is generally preferred.

ISTUDY
8.3. HEAD DEVELOPED BY A PUMP 495

β2 > 90◦

Head H
(forward-curved)
β2 = 90◦
(radial blades)
β2 < 90◦
(backward-curved)

Discharge Q
FIGURE 8.6 Effect of blade exit angle on pump head versus discharge
pump impeller.

EXAMPLE 8.1 An axial-flow fan of hub diameter 1.5 m and tip diameter
2 m pumps 5 m3 /s while rotating at 18 radians/second. Determine the blade
outlet and inlet angles at the hub and tip.

Solution The fan and the velocity triangles at the inlet an outlet of the
blade are shown in Figure S8.1.

A A ω1
A
r2 β1
V1 u2
Flow r1 u1 β2
A V2
ω2
Inlet Exit

FIGURE S8.1 The axial-flow fan and the velocity triangles at the entry
and exit of the runner.

Given, d1 = 1.5 m, d2 = 2 m, Q = 5 m3 /s, ω = 18 radians/second.

The flow velocity is

Q
V1 =
A
Q
=
π(r22− r12 )
5
=
π(12 − 0.752 )

ISTUDY
496 CHAPTER 8. HYDRAULIC MACHINERY

5
=
0.4375 × π
= 3.64 m/s

Blade velocity at the tip is

u2 = ωr2
= 18 × 1
= 18 m/s

Blade velocity at the hub is

u1 = ωr1
= 18 × 0.75
= 13.5 m/s

From the velocity diagrams at the hub (1) and tip (2), we have

V1
tan β1 =
u1
( )
−1 3.64
β1 = tan
13.5

= 15.1◦
V2
tan β2 =
u2
( )
3.64
β2 = tan−1
18

= 11.43◦

EXAMPLE 8.2 A centrifugal pump with a 700 mm diameter impeller runs


at 1800 rpm. The water enters without whirl, and leaves at 60◦ . The actual
head produced by the pump is 17 m. If the exit velocity is 6 m/s, find the
hydraulic efficiency.

Solution Given, d2 = 0.7 m (refer Figure 8.5), ω = 1800 rev/min, α2 = 60◦ ,


h = 17 m and V2 = 6 m/s.

The head which will be developed, if there are no losses, termed Euler head,
by Equation (8.11), is
1
H = (u2 Vt2 − u1 Vt1 )
g

ISTUDY
8.3. HEAD DEVELOPED BY A PUMP 497

where the circumferential velocity u1 = 0 and u2 is

u2 = ωr2
= (1800/60) × 2π × 0.35
= 65.97 m/s
Vt2 = V2 cos α2
= 6 × cos 60◦
= 3 m/s

Thus
u2 Vt2
H=
g
65.97 × 3
=
9.81
= 20.17 m

By Equation (8.7), the hydraulic efficiency is

hf
ηh = 1 −
hs

where hf head lost by friction and hs is the Euler head. Therefore,

h = hs − hf

Hence
hs − hf
ηh =
hs
17
=
20.17
= 0.8427

= 84.27%

EXAMPLE 8.3 Water is pumped at 227 m3 /hour through a 150 mm


diameter pipe, by a pump. If the elevation and pressure of the suction point
are 300 mm and −39 kPa (gauge) and those for the delivery point are 900
mm and 236 kPa (gauge), determine the head developed by the pump and
the power to be supplied to the pump.

Solution Given, Q = 227/3600 = 0.063 m3 /s, ρ = 103 kg/m3 , d = 0.15 m,


zs = 0.3 m, ps = −39 kPa (gauge), zd = 0.9 m and pd = 236 kPa (gauge).

ISTUDY
498 CHAPTER 8. HYDRAULIC MACHINERY

The velocity through the pipe is

Q
u=
A
4Q
=
πd2
4 × 0.063
=
π × 0.152
= 3.57 m/s

Thus, Vs = Vd = 3.5 m/s.


The pressures at the suction and delivery points are

ps = −39 + 101
= 62 kPa
pd = 236 + 101
= 337 kPa

By Equation (8.1), the head is


( ) ( )
pd V2 ps V2
h = Hd − Hs = + d + zd − + s + zs
ρg 2g ρg 2g

Thus,
( ) ( )
337 × 103 Vd2 62 × 103 Vs2
h= + + 0.9 − + + 0.3
103 × 9.81 2g 103 × 9.81 2g
= (34.35 + 0.9) − (6.3 + 0.3)

= 28.65 m

Therefore, the pressure head becomes

∆p = ρgh
= 103 × 9.81 × 28.65
= 281.06 kPa

Thus the power to be supplied to the pump becomes

Power = Q × ∆p
= 0.063 × 281.06

= 17.706 kW

ISTUDY
8.4. EFFICIENCY OF PUMP 499

8.4 EFFICIENCY OF PUMP


Only part of the energy, of the liquid flowing through a pump, imparted to the
shaft of the impeller is transferred to the flowing liquid. There is friction in the
bearing and packings, not all liquid passing through the pump is effectively
acted upon by the impeller, and there is substantial loss of energy due to fluid
friction which has a number of components including shock loss at entry to
the impeller1 , fluid friction as the fluid passes through the space between the
vanes or blades, and head loss as the fluid leaves the impeller. The efficiency of
the pump is strongly influenced by the conditions under which it is operated.
The power input to the pump, delivered to the pump shaft by the motor, is
called the shaft power or the break power. The power output from the pump,
delivered to the fluid, usually water, is called the fluid power or the water
power. The efficiency of a pump is given by

Output power Fluid power ρ gQh


η = Input power = = (8.19)
Shaft power Tω
where ρ is water density, g is gravitational acceleration, Q is discharge of
water, h is pressure head, T is the torque exerted on the pump shaft by the
motor driving the shaft, and ω is the rate of rotation of the shaft in radians
per second.
As we saw, the efficiency is basically composed of three parts: volumetric,
hydraulic, and mechanical. The volumetric efficiency is
Q
ηv =
Q + QL
where QL is the loss of fluid due to leakage in the impeller-casing clearances.
The hydraulic efficiency is
hf
ηh = 1 −
hs
where hf has three parts: (i) shock loss at the eye due to imperfect match
between inlet flow and the blade entrances, (ii) friction losses in the blade
passages, and (iii) circulation loss due to imperfect match at the exit side of
the blades.
The mechanical efficiency is
Pf
ηm = 1 −
bhp
where Pf is the power loss due to mechanical friction in the bearings, packing
glands, and other contact points in the machine.
By definition, the total efficiency is simply the product of its three parts

η = ηv ηh ηm
1 Shock loss occurs when the flow does not enter the impeller smoothly. This results in

separation of the flow from the impeller blade.

ISTUDY
500 CHAPTER 8. HYDRAULIC MACHINERY

EXAMPLE 8.4 If the efficiency of the pump in Example 8.3 is 78.4% and
the torque developed by the motor running the pump is 4870 N-m, determine
the rpm of the motor.

Solution Given, η = 0.784 and T = 4870 N-m.

The Fluid power, from Example 8.3, is 17.7 kW.

The efficiency, by Equation (8.5), is

ρ gQh
η=

Therefore,

17.7
0.784 =

17.7
Tω =
0.784
= 22.576 kW
= 22576 W
22576
ω=
4870
= 4.64 radians/second
4.64
= revolutions/second

= 0.7384 revolutions/second
= 0.7384 × 60

= 44.3 rpm

EXAMPLE 8.5 Water is to be transferred from tank 1 to tank 2, shown in


Figure E8.5. The pipe diameter is 150 mm and the total length is 60 m. The
friction coefficient of the pipe is 0.02 and the minor loss coefficients at the
pipe entrance, exit and the bend are 0.5, 1.0 and 1.5, respectively. If the flow
to be pumped with a pump of efficiency 84% is 6.1 m3 /minute, what should
be the power required to run the pump?

Solution Given, d = 150 mm, L = 60 m, f = 0.02, K1 = 0.5, K2 = 1.0,


Kb = 1.5, η = 0.84 and Q = 6.1 m3 /minute.

By Equation (8.1), the energy relation is

p1 V2 p2 V2
+ 1 + z1 + h = + 2 + z2 + hf + hm
ρg 2g ρg 2g

ISTUDY
8.4. EFFICIENCY OF PUMP 501

where hf is the head loss due to pipe friction, thus

LV2
hf = f
d 2g
Here V is the flow velocity in the pipe, given by
Q
V =
A
6.1/60
=
(π × 0.152 )/4
= 5.75 m/s

Therefore,
2
Water
Kb = 1.5 K2 = 1.0
Pump

3m
d = 150 mm
Total length of the pipe = 60 m

K1 = 0.5

FIGURE E8.5 Pump and pipeline connecting the tanks.

60 5.752
hf = 0.02 × ×
0.15 2 × 9.81
= 13.48 m

The minor losses are due to the state at 1, at the bend and at 2. Thus, the
minor loss, hm , is

hm = hm1 + hmb + hm2


V2 V2 V2
= K1 + Kb + K2
2g 2g 2g

V2
= (K1 + Kb + K2 )
2g
5.752
= (0.5 + 1.0 + 1.5) ×
2 × 9.81
= 5m

ISTUDY
502 CHAPTER 8. HYDRAULIC MACHINERY

Also, p1 = p2 = 0, V1 = V2 = 0, for large tanks as the given ones with their


free surface open to atmosphere. Thus the energy equation reduces to

z1 + h = z2 + hf + hm
h = (z2 − z1 ) + hf + hm

where z2 − z1 = 3 m. Therefore, the pump head h becomes

h = 3 + 13.48 + 5
= 21.48 m

For the ideal pump with 100% efficiency, the power is

P = ρghQ

6.1
= 103 × 9.81 × 21.48 ×
60
= 21.42 kW

Thus the power required to run the pump with 84% efficiency is
P
Pactual =
η
21.42
=
0.84
= 25.5 kW

25.5 × 103
=
746

≈ 34 hp

EXAMPLE 8.6 An axial-flow fan of hub diameter 0.8 m and blade tip
diameter 1.1 m, with blade inlet and exit angles 30◦ and 60◦ , respectively,
runs at 1200 rpm. The air flow enters at an angle of 30◦ . There is no change
in the axial component of velocity across the rotor. Assuming that the flow
enters and leaves the rotor at geometric angle of the blade and using the
properties at the mean bade radius determine (a) the volume flow rate of air,
(b) angle with which the flow leaves the rotor, and (c) the minimum power
required to run the fan.

Solution The velocity diagrams at the inlet (state 1) and exit (state 2) are
shown in Figure S8.6.

ISTUDY
8.4. EFFICIENCY OF PUMP 503

β2

V2
2
Vn
u2

α2
w2
V t1 β1
β2

1
V
V n1
Blade u1
w1
α1
β1
V t1
2
r

r1

FIGURE S8.6 Velocity triangles at the entry and exit of the runner.

Given, d1 = 0.8 m, d2 = 1.1 m, ω = 1200 rpm, β1 = 30◦ , β2 = 60◦ , α1 = 30◦ .

The mean diameter is


d1 + d2
dm =
2
0.8 + 1.1
=
2
= 0.95 m

The rotational speed is


1200
ω= × 2π
60
= 125.66 rad/s

The tangential speed at entrance is

u1 = rm ω
0.95
= × 125.66
2
= 59.7 m/s

The resultant velocity at 1 is

V1 = u1 sin α1
= 59.7 × sin 30◦
= 29.85 m/s

ISTUDY
504 CHAPTER 8. HYDRAULIC MACHINERY

Therefore,
Vt1 = V1 cos α1
= 29.85 × cos 30◦
= 25.85 m/s
Also,
u1 = Vt1 + Vn1 tan (90 − β1 )
= Vn1 cot α1 + Vn1 tan (90 − β1 )
= Vn1 (cot α1 + cot β1 )
u1
Vn1 =
cot α1 + cot β1
59.7
=
cot 30◦ + cot 30◦
59.7
=
3.46
= 17.23 m/s
(a) The volume flow rate is
Q = A Vn1
π
= (1.12 − 0.82 ) × 17.23
4
= 7.71 m3 /s

(b) The tangential component of velocity at the exit is


Vt2 = u2 − Vn1 tan 30◦
= 59.7 − 17.23 × tan 30◦
= 59.7 − 9.95
= 49.75 m/s
Therefore,
vn2
tan α2 =
u2 − Vt2
17.23
=
59.7 − 49.75
= 1.732
α = tan−1 (1.732)

= 60◦

ISTUDY
8.5. SIMILARITY LAWS FOR PUMPS 505

(c) The torque T , by Equation (8.10), is

T = ρQ(r2 Vt2 − r1 Vt1 )


( )
1.1 0.8
= 1.225 × 7.71 × × 49.95 − × 25.85
2 2
= 1.225 × 7.71 × 17.1325
= 161.81 N-m

Thus the power is

P = ωT
= 125.66 × 161.81

= 20.33 kW

8.5 SIMILARITY LAWS FOR PUMPS


Similarity laws (Chapter 3) enables us to predict the performance of a pro-
totype pump (or turbine) from the test of a scaled model. Moreover, and of
particular value in pump selection, these laws make it possible to predict the
performance of a given machine under different conditions of operation from
those under which it has been tested.
Similarity laws are based on the concept that two geometrically similar
machines with similar velocity diagrams, at entrance and exit, from the ro-
tating elements are homologous. This means that their behavior will bear a
resemblance to one another.
The similarity laws can be derived by dimensional analysis. The most
significant variable2 affecting the operation of a turbomachine are the head
h, discharge Q, the relative speed n, the diameter of the rotor D, and the
acceleration due to gravity g. Thus, from the Buckingham Π-theorem, since
there are five dimensional variables and two fundamental dimensions (L and
T), there will be three dimensionless groups. We have

f (h, Q, n, D, g) = 0

where h is the head developed by the pump, Q is the discharge, n is the


rotational speed (the rotational speed is also expressed as ω), D is the diameter
and g is the gravitational acceleration.
Upon grouping these variables into dimensionless quantities, we get
( )
Q g h
f1 , , =0
nD3 n2 D D
2 If we wish to relate the operation of one pump to another with different fluids in each

then kinematic viscosity is a significant variable.

ISTUDY
506 CHAPTER 8. HYDRAULIC MACHINERY

Laboratory tests on turbomachine have shown that the dimensionless group


g h
2
is inversely proportional to . These can be combined to give
n D D
g D
=K
n2 D h
that is,
gh
K=
n2 D 2
Thus ( )
Q gh
f2 3
, 2 2 =0 (8.20)
nD n D
From Equation (8.3), we get

Q ∝ nD3 or Q = Kq nD3 (8.21)

and, assuming that g is constant, we have

h ∝ n2 D 2 or h = Kh n2 D2 (8.22)

But power P ∝ Q h, therefore,

P ∝ n3 D 5 or P = KP n3 D5 (8.23)

While designing a pump, the coefficients KQ , Kh and KP can be evaluated,


from the test data, and then used to predict the performance of homologous
pumps. Similarity laws are of great practical value, but care must be exercised
when applying them. Thus, in comparing two machines of different sizes, the
two must be homologous and the variation in the values of h, D, and n
should not be too large. For example, a machine that operates satisfactorily
at low speeds may cavitate at high speeds. The values of KQ , Kh and KP in
Equations (8.21)-(8.23) change somewhat as h, D and n are varied, because
the efficiencies of homologous machines are not identical. Large machines are
usually more efficient than smaller ones, because their flow passages are larger.
Also, efficiency usually increases with speed of rotation, because power output
varies with the cube of the speed while mechanical losses increase only as the
square of the speed. As a result, Equations (8.21)-(8.23) are appropriate.
An empirical equation suggested by Moody [J.H.T. Sun, Hydraulic Ma-
chinery, Davi’s Handbook of Applied Hydraulics, by V.J. Zipparrp and H.
Hasen (Eds.), 4th ed., McGraw-Hill, New York, 1993, pp. 21-26.] that gives
fairly reasonable results for estimating the efficiency of a prototype pump from
the test of a geometrically similar (model) pump is
( )1/5
1 − ηp Dm
≈ (8.24)
1 − ηm Dp

where the subscripts p and m refer to the prototype and model, respectively.

ISTUDY
8.6. PERFORMANCE CHARACTERISTICS OF PUMPS ... 507

The effect of rotational speed on the performance (Q, h, P ) of a given


pump are known as affinity laws. Equations (8.21)-(8.23) govern these effects,
which generally change pump efficiency little. The effects of trimming the
impeller outside diameter on the same performances are also known as affinity
laws. In this case the same equations do not govern, instead Q ∝ D, hQ ∝ D2 ,
P ∝ D3 . Because trimming increases losses through larger clearances, it can
reduce efficiency appreciably and thereby affect performance, so we should
use only these relations for small diameter changes.

8.6 PERFORMANCE CHARACTERISTICS


OF PUMPS AT CONSTANT SPEED
The efficiency of a pump depends on the condition under which it operates.
Because of this, for selecting a pump for a given operating condition, it is
necessary to have information about the performance of various pumps among
which the selection is to be made. The pump manufacturer usually has this
type of information, as determined by laboratory tests, for what are called
shelf items, that is, standard pumps. However, large-capacity pumps are
sometimes custom-made. Usually a model of such a pump is made and tested
before design of the prototype pump.
Though some centrifugal pumps are driven by variable-speed motors, the
usual mode of operation of a pump is at constant speed. Typical speeds of
some constant-speed electric motors are given in Table 8.1.

TABLE 8.1 Operating speeds of constant-speed electric motors


60-cycle 50-cycle
Pairs of Synchronous Inductiona Synchronous Inductiona
poles rpm rpm rpm rpm
1 3600 3500 3000 2900
2 1800 1750 1500 1450
3 1200 1160 1000 960
4 900 870 750 720
5 720 695 600 575
6 600 580 500 480
8 450 435 375 360
10 360 350 200 290
12 300 290 250 240

a These values for the operating speed of induction motors are approximate. The speed of
induction motors is usually 2–3% lower than that of synchronous motors.

ISTUDY
508 CHAPTER 8. HYDRAULIC MACHINERY

The plot of head versus capacity is called pump characteristic curve. The
pump characteristic curve and other performance curves for a typical mixed-
flow centrifugal pump are shown in Figure 8.7.

FIGURE 8.7 Characteristic curves for a mixed-flow centrifugal pump.

The capacity of a pump is called normal or rated capacity when its oper-
ating point corresponds to the point of optimum efficiency or best efficiency
power (BEP), as shown in Figure 8.5. Let us denote the rotational speed at
the BEP by a subscript e. Characteristic curves for a typical axial-flow pump
are shown in Figure 8.8. Curves such as those shown in Figures 8.7 and 8.8
are usually determined by pump manufacturers through laboratory tests.
By inspecting these two figures we see the remarkable difference in the
characteristics of these pumps. We observe that the efficiency of both pumps
drop rather rapidly when the flow rate at which they are pumping exceeds
the optimum. This is particularly true in the case of the axial-flow pump.
The shape of the impellers and vanes and their relationship to the pump
casing cause variations in the intensity of shock loss, fluid friction, and turbu-
lence. These vary with head and flow rate, and are responsible for the wide
variation in pump characteristics. The shutoff head is that which is developed
when there is no flow. In the case of the mixed-flow centrifugal pump (Figure
8.7), the shutoff head is usually about 10% greater than the normal head,
that which occurs at the point of optimum efficiency, while in the case of the
axial-flow pump (Figure 8.8), the shutoff head may be as much as three times
the normal head.

ISTUDY
8.7. PERFORMANCE CHARACTERISTICS AT DIFFERENT ... 509

12 100

h vs Q
Efficiency
9 75

and shaft horsepower ×10−1


Efficiency, %
Head h, m

6 50

BEP

3 Shaft horsepower 25

ne = 450 rpm
0 0
0 1.5 3.0 4.5 6.0
Capacity Q, m3 /s
FIGURE 8.8 Characteristic curves for a typical axial-flow pump.

8.7 PERFORMANCE CHARACTERISTICS


AT DIFFERENT SPEEDS AND SIZES
The choice of a pump for a given situation will depend on the rotational speed
of the motor used to drive the pump. If the characteristic curve for a pump
at a given rotational speed is known, the relation between head and capacity
at different rotational speeds can be derived approximately through use of
Equations (8.21) and (8.22). For example, in Figure 8.9, assume the charac-
teristic curve (Curve 1) of a pump when operating at n1 rpm is given. We can
approximately derive the characteristic curves at different speeds of operation
such as n2 and n3 by transferring points on Curve 1 to corresponding points
Curves 2 and 3, respectively. Thus, from Equations (8.21) and (8.22), we have
for Curve 2, Q2 = Q1 (n2 /n1 ) and h2 = h1 (n2 /n1 )2 . Similar expressions can
be used to develop Curve 3.
Superimposed on Figure 8.9 are lines of equal efficiency as determined by
test. It is seen that the efficiency drops off rather rapidly as one moves away
from the BEP. We can also approximately derive characteristic performance
curves for pumps of different size, all operating at the same speed, using
Equations (8.21) and (8.22). Such curves are shown in Figure 8.8. Corre-
sponding points were transferred through the use of Q2 = Q1 (D2 /D1 )3 and
h2 = h1 (D2 /D1 )2 . Curves of equal efficiencies have also been superimposed
on Figure 8.10.
We can approximately develop characteristic performance curves for pumps
having similar geometric shape, of both different size and operating at various

ISTUDY
510 CHAPTER 8. HYDRAULIC MACHINERY

36
n2
56%
66%
30
74%
80%
n1
24
80%
Head h, m

18 BEP
n3

12
74%
66%
56%
6
n1 = 1450 rpm

0
0 12 24 36 48 60 72
Capacity Q, thousands of litres/minute
FIGURE 8.9 Characteristic performance curves of a typical mixed-flow
centrifugal pump (Figure 8.7) at various speeds of rotation with contours
of equal efficiency.

constant speeds, using Equations (8.21) and (8.22), in which case

Q2 = (n2 /n1 )(D2 /D1 )3

and
h1 = (n2 /n1 )2 (D2 /D1 )2
But the accuracy of the calculations using these curves will drop off with large
variations in D and n.

8.8 OPERATING POINT OF A PUMP


How a pump operates depends on both the performance characteristics of the
pump and the characteristics of the system that it will operate in. The pump
operating characteristics (h versus Q) for a selected speed of operation of a
pump will be as shown in Figure 8.11.
The pump head, h, required for desired flow rate, Q, known as the system
characteristics is also shown in the figure. For the present case, the pump is
delivering liquid through a piping system with a static lift as ∆z. The head
that the pump used to develop is equal to the static lift plus the total head
loss in the piping systems, which is approximately proportional to Q2 . The
pump-operating head, hr , and flow rate are given by the intersection of the
two curves.

ISTUDY
8.8. OPERATING POINT OF A PUMP 511

30

D2 56%
66%
74%
24 D1

80%

80%
D3 BEP
18
Head h, m

74%
66%

56%
12

6
For n = n1 = 1450 rpm

0
0 12 24 36 48 60 72
Capacity Q, thousands of litres/minute

FIGURE 8.10 Characteristic performance curves with contours of equal


efficiencies for typical homologous mixed-flow centrifugal pumps having
impellers of different size (D1 is the same as that for pump in Figure 8.7).

h
Pump Operating point

hr hL (head loss)

line
Pipe

∆z (static lift)

Qd Q

FIGURE 8.11 Graphical method for finding the operating point of a


pump and pipeline.

The values of h and Q given by the intersection point may or may not be
those for the maximum efficiency. In Figures 8.9 and 8.10 it is seen that, the
efficiency drops off as we move away from the BEP. Therefore, it is important
to select a pump such that the pump performance and system characteristics
curves intersect near BEP.

ISTUDY
512 CHAPTER 8. HYDRAULIC MACHINERY

8.9 SPECIFIC SPEED OF PUMPS


Specific speed is a number that defines the type of pump, such as radial-flow,
axial-flow or mixed-flow. The specific speed Ns of a pump can be expressed
as
( √ )
ω Q
Ns = (8.25)
(gh)3/4 BEP

where ω is the the rotational speed (revolutions per second), Q is the volume
flow rate (m3 /s), h (m) is the head corresponding to the point of optimum
operating efficiency (BEP) and g is the gravitational acceleration.
Figure 8.12 shows several typical impellers and their corresponding specific
speeds. Radial-flow impellers generally have specific speeds between 0.2 to 2.0,
from 1.5 to 3.7 for mixed-flow, and from 3.3 to 5.5 for axial-flow pumps.

ω Q
(Ns )SI =
(gh)3/4
0.2 0.3 0.5 1.0 1.5 2.0 3.0 4.0
100

90
η

80
Efficiency η, %

Radial-flow Mixed-flow Axial-flow


70 2.5

2.0

60 1.5
φe φe
1.0

50 0.5

45 0.0
500

1000

1500

2000

3000
4000
5000

10,000

15,000


n gpm
Ns =
h3/4
FIGURE 8.12 Optimum efficiency values of ϕe for water pumps as a
function of Specific speed.
Equation (8.25) indicates that pumping against high heads requires a low-
specific speed pump. For very high heads and low discharges, the required

ISTUDY
8.9. SPECIFIC SPEED OF PUMPS 513

specific speed may fall below the value for normal design and result in a pump
with low efficiency. To overcome this problem, the head can be distributed
among a number of pumps in series, or a multistage pump can be used. A
multistage pump is usually less expensive than a series of individual pumps,
but it has the disadvantage of developing very high pressures. The excessive
pressure in the system can be avoided by spacing pumps more or less uni-
formly along a pipeline.

EXAMPLE 8.7 A centrifugal pump running at 2950 revolution per minute


under test at peak efficiency with effective head of 75 m of water and specific
speed 0.0779. If the overall efficiency is 76%, determine (a) the flow rate and
(b) the power required to run the pump.

Solution Given, ω = 2950 rev/min, h = 75 m, Ns = 0.0779 and η = 0.76.

(a) By Equation (8.25), the specific speed is



ω Q
Ns =
(gh)3/4

Therefore,
√ Ns (gh)3/4
Q=
ω
0.0779 × (9.81 × 75)3/4
=
(2950/60)
= 0.2238
Q = (0.2238)2

= 0.05 m3 /s

(b) The power consumed by the pump, by Equation (8.11), is

ρgQh
P=
η
103 × 9.81 × 0.05 × 75
=
0.76
= 48.4 kW

EXAMPLE 8.8 A prototype test of a mixed-flow pump with a 2 m diameter


discharge opening, operating at 225 rpm, discharges 9.769 m3 /s with a head
of 13.72 m running at the best efficiency of 88%. What side and synchronous
speed of homologous pump should be used to discharge 5.66 m3 /s at 18.3 m
head at the point of best efficiency?

ISTUDY
514 CHAPTER 8. HYDRAULIC MACHINERY

Solution Let subscripts 1 and 2 refer to the prototype type and homologous
pump, respectively.

Given, d1 = 2 m, n1 = 225 rpm, h1 = 13.72 m, Q1 = 9.769 m3 /s, Q2 = 5.66


m3 /s and h2 = 18.3 m.

For homologous pumps, by Equation (8.20),

Q1 Q2 h1 h2
= and = 2 2
n1 d31 n2 d32 n21 d21 n2 d2

Therefore,

Q2
n2 d32 = n1 d31
Q1

5.66
= × 225 × 23
9.769

= 1042.9

h2 2 2
n22 d22 = n d
h1 1 1

18.3
= × 2252 × 22
13.72

= 270098.4

Thus,

n2 d32 1042.9
2 2 =
n2 d 2 270098.4

= 3.86 × 10−3

d2
= 3.86 × 10−3
n2

d2 = n2 × 3.86 × 10−3

Also,

n2 d32 = 1042.9

1042.9
n2 =
d32

ISTUDY
8.10. PERIPHERAL-VELOCITY FACTOR 515

Therefore,
( )
1042.9
d2 = × (3.86 × 10−3 )
d32
d42 = 4.026
d2 = (4.026)1/4

= 1.417 m
The speed of the synchronous pump is
1042.9
n2 =
d32
1042.9
=
1.4173
= 366 rpm
Note that the efficiency of pump 1 is given as 0.88. Therefore, if the calcula-
tions are correct, the efficiency of the synchronous pump has to be 0.88. This
can be checked as follows.

By Equation (8.24), we have


( )1/5
1 − η1 d2
=
1 − η2 d1
Therefore,
( )1/5
1 − η1 1.417
=
1 − η2 2
= 0.9334
1 − η1
1 − η2 =
0.9334
1 − 0.88
=
0.9334
= 0.1285
η2 = 1 − 0.1285
= 0.8715
It is seen that η1 ≈ η2 .

8.10 PERIPHERAL-VELOCITY FACTOR


The peripheral-velocity
√ factor for a pump impeller is the ratio of the peripheral
velocity to 2gh, usually denoted by ϕ. Thus, the peripheral speed of the

ISTUDY
516 CHAPTER 8. HYDRAULIC MACHINERY

impeller of a pump is √
u2 = ϕ 2gh (8.26)
For an axial-pump u2 is the vane-tip speed.
The peripheral velocity for a machine may vary from zero up to some max-
imum under a given head, depending on the operating speed, hence ϕ would
vary through a wide range. But the speed that is of practical significance
is that corresponds to maximum efficiency. Let ϕe corresponds to maximum
efficiency. At maximum efficiency
2π r ne
u2 =
60
π D ne
=
60

= ϕe 2gh

Thus
√ √
60 2g ϕe h
D= (8.27)
π ne

8.11 CAVITATION IN PUMPS


Avoiding cavitation is essential for efficient operation of a pump without im-
peller damage. As liquid passes through the impeller of a pump, there is a
change in pressure. If the absolute pressure of the liquid drops to the va-
por pressure, cavitation will occur. The region of vaporization hinders the
flow and places a limit on the pump capacity. As the liquid moves further
into a region of higher pressure, the bubble collapses and the implosion3 of
the bubbles may cause pitting on the impeller. The most probable location
where cavitation is likely to occur is near the point of discharge (periphery) of
radial-flow and mixed-flow impellers, where the velocities are the highest. It
may also occur on the suction side of the impeller, where the pressures are the
lowest. In the case of an axial-flow pump, the blade-tip is the most vulnerable
to cavitation.
For pumps, a cavitation parameter, σ, is defined as

(p)abs V2 pv
+ s −
ρg 2g ρg NPSH
σ= = (8.28)
h h

where subscript s refers to values at the pump intake (that is, suction side
of the pump), h is the head developed by the pump, and pv is the vapor
pressure. As pv is given as absolute, ps also should be absolute pressure.
NPSH is referred to as the net positive suction head.
3 a process in which objects are destroyed by collapsing (or being squeezed in) on them-

selves.

ISTUDY
8.11. CAVITATION IN PUMPS 517

Note that, the unit all terms in the numerator and denominator of Equa-
tion (8.28) are meters. Thus, the cavitation parameter, σ is dimensionless.
The value σ, below which there is a drop in pump efficiency is called the
critical cavitation parameter, σc . This is an indication of the onset of cavi-
tation. The value of σc depends on the type of pump and the conditions of
operation. For a given pump, the value of σc varies with flow rate. Approx-
imate values of σc for centrifugal pumps operating under normal conditions
near optimum efficiency are shown in Figure 8.13.

ωe Q
(Ns )SI =
(gh)3/4
0 1 2 3 4

1.6

1.4

1.2
c


1.0
NPSH
h

0.8

σc =

0.6

0.4

0.2

0 2000 4000 6000 8000 10,000 12,000

gpm
r
Specific speed based on gpm, Ns = ωe
h3/4
FIGURE 8.13 Variation of σc as a function of specific speed.

The value of σc for important installations should be determined exper-


imentally in a model study. For modeling purposes, a parameter called the
suction specific speed S, is defined as

ω Q
S= (8.29)
NPSH3/4
If the model and prototype are operated at identical values of Ns and S,
similarity of flow and cavitation is achieved provided the model and prototype
are geometrically similar to one another.
Occurrence of cavitation can be prevented by operating the pump with the
cavitation parameter σ larger than σc . This can be accomplished by selecting
the proper type and size of the pump and speed of operation and by setting

ISTUDY
518 CHAPTER 8. HYDRAULIC MACHINERY

the pump at the proper point of elevation in the system. From Equation
(8.28), it is seen that small values of σ result in large values of h, the head
developed by the pump. Hence, for any given situation, there is a limiting
value of h above which σ will be less than σc , resulting in cavitation.
For the pump shown in Figure 8.4, the energy equation can be expressed
as
(p0 )abs (ps )abs V2
− hL = zs + + s
ρg ρg 2g
or
(ps )abs V2 (p0 )abs
+ s = − hL − zs
ρg 2g ρg
where (p0 )abs /(ρ g) is the pressure on the surface of the reservoir. If the
liquid is drawn from a closed reservoir, (p0 )abs /(ρ g) could be either greater
or less than the atmospheric pressure. Usually, the reservoir is open to the
atmosphere and p0 /(ρ g) = patm /(ρ g).
Substituting the above expressions into Equation (8.28), we get
(p0 )abs /(ρ g) − hL − zs − pv /(ρ g)
σ= (8.30)
h
This expression indicates that σ will tend to be small, leading to a possibility
of cavitation, for (a) high head, (b) low atmospheric pressure, that is, high
elevation, (c) large head loss between source reservoir and the pump, (d)
large value of zs , that is, pump at a relatively high elevation compared to the
reservoir water surface, and (e) large value of vapor pressure. With (a), the
required head can be limited by using multistage pumps. For a given liquid
to be pumped at a certain elevation and temperature, we have no control
over items (b) and (e). The tendency toward cavitation can be reduced by
minimizing hL [(c)], by placing the pump close to the source reservoir, and
by setting the pump at a low elevation [(d)] relative to the reservoir water
surface.
In Equation (8.30), when σ = σc then zs = (zs )max , the highest elevation
at which we can safely set a pump and guard against cavitation. Rearranging,
we have
(ps )abs pv
(zs )max = − − σc h − hL (8.31)
ρg ρg
As long as zs is less than (zs )max , there should be no problem with cavitation.

EXAMPLE 8.9 Tests of a pump model indicate critical cavitation param-


eter of 0.10. A homologous unit installed at a location where the atmospheric
pressure and vapour pressures are 90 kPa and 3.5 kPa, respectively, has to
pump water against a head of 25 m. The head loss from the suction to the
pump impeller is 0.35 N-m/N. What is the maximum permissible suction
head?

Solution Given, σc = 0.10, pa = 90 kPa, pv = 3.5 kPa, h = 25 m and


hL = 0.35 N-m/N.

ISTUDY
8.11. CAVITATION IN PUMPS 519

By Equation (8.28),
pa V2 pv
+ s −
ρg 2g ρg
σc =
h
Therefore,
( )
pa pv
Vs2 = 2 − − σc hg
ρ ρ
( )
90 × 103 3.5 × 103
=2 − − 0.10 × 9.81 × 25
103 103
= 2 × 61.975
= 123.95

The suction speed in terms of pressure head is



2∆p
Vs =
ρ

Thus
2∆p
= Vs2
ρ
ρVs2
∆p =
2
103 × 123.95
=
2
= 61975
ρgh = 61975
61975
h=
103 × 9.81
= 6.32 m

This is the suction head without loss. Therefore, the maximum possible suc-
tion head becomes

hmax = h − hL
= 6.32 − 0.35

= 5.97 m

EXAMPLE 8.10 The available NPSH provided by a pump manufacturer


at a flow of 0.06 m3 /s is 3.5 m. Determine the height of the pump above the
suction reservoir,zs . The water temperature is 25◦ C (with pv /(ρg) = 0.33 m),
atmospheric pressure is 101 kPa and the head loss from the reservoir to pump
is 0.3 m-N/N.

ISTUDY
520 CHAPTER 8. HYDRAULIC MACHINERY

Solution Given, NPSH = 3.5 m, Q = 0.06 m3 /s, pv /(ρg) = 0.33 m, pa =


101 kPa and hL = 0.3 m-N/N.

By Equation (8.28),
NPSH
σ=
h
Therefore,
σ h = 3.5 m

By Equation (8.31),

pa pv
zs = − − σ h − hL
ρg ρg
101 × 103
= − 0.33 − 3.5 − 0.3
103 × 9.81
= 10.296 − 0.33 − 3.5 − 0.3

= 6.166 m

8.12 SELECTION OF PUMPS


For selecting a pump for a given application, a variety of pumps are available
for choice, based on the performance information of the pump, such as that in
Figures 8.7 and 8.8. The users task is to select the pump (or pumps) that is
most suitable for the situation. Alternatives to be investigated include specific
speed Ns , size D of the impeller, and speed of operation n. Other alternatives
include the use of multiple pumps, pumps in series, pumps in parallel etc.
The possibility of cavitation must be carefully investigated in all cases.
Indeed cavitation should be avoided. Cavitation can be eliminated by limiting
the head that the pump must develop, by selecting the proper type of pump,
and by selecting the pump at a low enough elevation.
Selection of a pump with operating point close to the BEP will optimize
the pump efficiency, resulting in a minimization of energy required to run
the pump. The operating point can be shifted by changing the pump char-
acteristic curve, by changing the system characteristic curve, or by shifting
both. The pump characteristic curve can be changed by changing the speed
of operation of a given pump or by selecting a different pump with different
performance characteristics, as illustrated in Figure 8.14(a). In some situa-
tions it is helpful to trim the impeller, that is, reduce its diameter, say by
grinding it down. The smaller impeller is installed in the original casing. The
system characteristic curve can be changed by changing the pipe size or by
throttling the flow, as illustrated in Figure 8.14(b).

ISTUDY
8.13. PUMPS OPERATING IN SERIES AND IN PARALLEL 521

h h
Higher speed
3
Smaller pipe
2 1

3
3
1
Low speed Large pipe

1 Given pump 2
2 Trim the impeller 1 Given system
3 Change the speed 3 2 Use a throttling valve
4 Use a different pump 4 3 Use a different size pipe

Q Q

(a) (b)

FIGURE 8.14 Changing the operating point by (a) changing the pump
characteristic curve or (b) changing the system characteristic curve.

8.13 PUMPS OPERATING IN SERIES AND


IN PARALLEL
The performance characteristics of different pumps with different characteris-
tics, a particular pump at different speeds, homologous pump of different size,
two identical pumps in series, and two identical pumps in parallel, as shown
in Figure 8.15.
For series or parallel installation of pumps, it is important to ensure that
they are reasonably similar, or with identical head-capacity characteristics
throughout their range of operation, otherwise, one pump will carry most
of the load and, under certain conditions, all of the load, with the other
pump acting as hindrance rather than a help. In parallel installation, if the
performance characteristics of the pumps are quite different, a condition of
back-flow can occur in one of the pumps. Finally, we must always be sure
that the selected pump will not encounter cavitation problems over the full
range of operating conditions.
The best mode of operation for any pump is determined by plotting the
pump characteristics and the pipe system characteristics on the same diagram
(Figure 8.11). The point at which the two curves intersect indicates best
operating point.

8.14 HYDRAULIC TURBINES


Hydraulic turbines are machines used primarily for the development of hydro-
electric energy. These turbines extract energy from flowing water and convert
it to mechanical energy to drive electric generators. Impulse turbine and reac-
tion turbine are two basic types of hydraulic turbines. In the impulse turbine

ISTUDY
522 CHAPTER 8. HYDRAULIC MACHINERY

h h D = const. h n = const.
High Ns 1.2n
Q∝n 1.1D Q ∝ D3
h ∝ n2 h ∝ D2
Med. Ns n D

Low Ns 0.8n 0.9D

Q Q Q
(a) (b) (c)

h h

h2 = h1

h1
Q1 Q2 = Q1

Q Q

(d) (e)

FIGURE 8.15 Variation of Q with h for (a) different pumps with


different characteristics, (b) a particular of pump at different speeds,
(c) homologous pump of different size, (d) two identical pumps in series,
(e) two identical pumps in parallel.

a free jet of water impinges on the revolving element of the machine which
is exposed to atmospheric pressure. In the reaction turbine, flow takes place
under pressure in a closed chamber. The energy delivered to an impulse tur-
bine is all kinetic energy and the energy supplied to a reaction turbine is the
combination of pressure energy and kinetic energy. But in both turbines the
action of the machine depends on a change in the momentum of the water
so that a dynamic force is exerted on the rotating element, or runner. The
runner of a reaction turbine is similar in design, but not identical to a pump
impeller.
Turbines are operated at constant speed. If the electric current generated
has to be of 60-cycle (that is, 60 cycle/second or Hz), the relative speed, n,
of a turbine in revolution per minute is given by N = 7200/n, where N is
the the number of poles in the generator and must be an even integer. Most
60 Hz generators have 12 or 96 poles. In many parts of the world, 50-cycle
current is used, in which case N = 6000/n.

8.15 IMPULSE TURBINE


The impulse turbine is also called Pelton wheel, in honor of Lester A. Pelton
(1829 - 1908), who contributed to much of its development in the early gold-

ISTUDY
8.15. IMPULSE TURBINE 523

mining days in California. A typical impulse turbine and flow system are
illustrated in Figure 8.16.
The buckets may be individually cast and bolted to the central spider, or,
more commonly, the entire runner is cast as a single unit. When the jet strikes
the dividing ridge of the bucket, it is split into two parts that discharge at
both sides of the bucket, as illustrated in Figure 8.17.
2
VB
2g

EL hL
HGL
Forebay
Net head
Penstock on wheel
Static head y
pB
ρg Gross head
at plant

Datum
z

Tailwater

FIGURE 8.16 A typical impulse turbine system.

FIGURE 8.17 Velocity vector diagrams at (a) entrance to the bucket and
(b) at the exit from the bucket.

Each split bucket has a notch that enables the bucket to attain position
nearly tangent to the direction of the jet before the bucket lip intercepts the
jet. Only one jet is used on small turbines, but two or more jets impinging at
different points around the wheel are often used in large units. The jets are
usually produced by a needle nozzle, with velocities some times even exceeding
150 m/s.
Since water hammer (that is, abrupt decrease of the velocity of a liquid
in a pipeline) might occur in the supply pipe if the nozzle where closed so

ISTUDY
524 CHAPTER 8. HYDRAULIC MACHINERY

rapidly, some nozzles are provided with a bypass valve that opens whenever
the needle valve is closed quickly. The same effect can be obtained with a
jet deflector, whose position can be adjusted to deflect the jet away from the
wheel when the liquid drops. A governor is required to actuate the nozzle and
bypass or deflector units.
Owing to the risk of greatly increased pressures and stresses in the supply
pipe due to water hammer, in addition to the protective measures, these
pipes are specially designed to have enough strength to take these pressures
and stress. Such a specially designed supply pipe, for an impulse turbine or
a higher-head reaction turbine, is called a penstock.
The generator rotor is usually mounted on a horizontal shaft between two
bearings with the runner installed on the projecting end of the shaft. This is
known as a single-overhang to equalize the bearing loads. Impulse turbines
are provided with housing to prevent splashing, but the air within the housing
is substantially at atmospheric pressure. Some modern wheels are mounted
on a vertical axis below the generator and are driven by jets from several
nozzles spaced uniformly around the periphery of the wheel.
For high efficiency the width of the bucket should be 3-4 times the jet di-
ameter, and the wheel diameter is usually 15-20 times the jet diameter. The
wheel diameter, also known as the pitch diameter, is the diameter of the pitch
circle, the circle to which the centreline of the jet is tangent. The diameter
of the impulse turbines range up to about 5 m. Theoretically maximum effi-
ciency would result if a bucket completely reverses the relative velocity of the
jet. However, this is not possible because the water must be deflected to one
side to avoid interfering with the following bucket, and the bucket exit angle
β2 (Figure 8.17) is usually about 165◦ .

EXAMPLE 8.11 A turbine has a velocity of 6 m/s at the entrance to the


draft tube and a velocity of 1.2 m/s at the exit. If the friction loss is 0.1 m
and the tail water below the entrance to the draft is 5 m, find the pressure
head at the entrance.

Solution Let subscripts 1 and 2 refer to the entrance and exit. With
reference to Figure 8.16, we have

p1 V2 p2 V2
+ 1 = + 2 + hf + z
ρ 2g ρ 2g

Given, V1 = 6 m/s, V2 = 1.2 m/s, hf = 0.1 m, z = −5 m. Assuming that the


turbine delivers to atmosphere, we have p2 = 0. Thus

p1 V2 V2
= − 1 + 2 + hf + z
ρ 2g 2g
62 1.22
=− + + hf + z
2 × 9.81 2 × 9.81

ISTUDY
8.16. IMPULSE TURBINE OPERATION 525

= −1.835 + 0.0734 + 0.1 − 5

= −6.66 m

8.16 IMPULSE TURBINE OPERATION


In the impulse turbine the energy of the fluid supplied to the machine is
converted by one or more nozzles into kinetic energy. The jet from the nozzle
strikes a series of buckets on the circumference of the wheel and is turned
through an angle β (usually around 165◦ ) thus producing a force on the bucket
and a torque on the wheel. The interior of the casing of the wheel is at
atmospheric pressure and not filled with water. The wheel must be placed
above the tailwater level so that the water leaving the buckets falls clear of
the wheel.
A plan view of one of the buckets of an impulse turbine along with the
velocity vector diagrams are shown in Figure 8.17. Position A represents the
inlet entry of the water into the bucket, and position B indicates the moment
the water leaves the bucket. The velocity u is at the centreline of the buckets.
The true path of the water is shown, and the velocity changes from V1 at
entry to V2 at exit. A vector diagram of the velocities at the entry is shown
in Figure 8.17(a). The vector v represents the velocity of the water relative
to the bucket. Assuming friction to be negligible, the magnitude of v of the
water velocity relative to the bucket remains constant, but its direction at
discharge must be tangential to the bucket, as shown in Figure 8.17(b). The
force exerted by the water in the bucket in the direction of motor in terms of
absolute velocity is

F = ρ Q∆V = ρ Q (V1 − V2 cos α2 ) (8.32)


Before casting this equation in terms of relative velocities, let us briefly review
the relation between absolute and relative velocities. The absolute velocity V
of a body (Figure 8.18) is its velocity relative to earth. The relative velocity
v of a body is its velocity relative to a body, which may in turn be in motion
with absolute velocity u relative to earth.
Vu Vu vu
vu

V v V v v
β β
V
α β α α
u u Vu u
vu
V n = vn
V n = vn V n = vn

FIGURE 8.18 Relative (dashed) and absolute velocity relations, with


their components.

The absolute velocity V of the first body is the vector sum of its velocity
v relative to the second body and the absolute velocity u of the latter. The

ISTUDY
526 CHAPTER 8. HYDRAULIC MACHINERY

relation of the three is thus


V =u+v
Let α and β are the angles made by the absolute and relative velocities of
a fluid, respectively, with positive direction of the linear velocity u at some
point on a solid body. From Figure 8.18, it is seen that, whatever the shape
of the velocity vector triangle, the velocity components parallel and normal
to u are always given by

Vu = u + vu = V cos α
= u + v cos β
Vn = vn = V sin α
= v sin β

where subscripts u and n refer to components parallel and normal to the


direction of velocity u.
In terms of relative velocities, Equation (8.32) becomes

F = ρ Q ∆v
= ρ Q (v − v cos β2 )

that is,
F = ρ Q (V1 − u)(1 − cos β2 ) (8.33)
since the relative velocity v = V1 − u.
The power transmitted to the buckets from the water is the product of the
force and the velocity of the body on which the force is acting. Hence,

P = F u = ρ Q (V1 − u)(1 − cos β2 ) u (8.34)

From Equation (8.34), it is seen that no power is developed when u = 0 or


when u = V1 . For a given turbine and jet the maximum power occurs at an
intermediate u, which can be found by differentiating Equation (8.34), with
respect to u, and equating to zero. Thus

dP
= ρQ (1 − cos β2 )(V1 − u − u) = 0 (8.35)
du
This gives
u = V1 /2
Thus the greatest hydraulic efficiency (neglecting fluid friction) occurs when
the peripheral speed of the wheel is half of the jet velocity. Tests of impulse
turbines show that, because of energy loss, the best operating conditions occur
when u/V1 is between 0.43 and 0.48.

ISTUDY
8.16. IMPULSE TURBINE OPERATION 527

In our analysis here the buckets are assumed to have the shape with a split
in such a manner that the bucket angle at entrance is 0◦ , that is, the water
enters tangentially to the bucket. In the actual situation, there is usually a
small bucket angle of less than 10◦ , in which case V1 in Equation (8.34) should
be replaced by assuming α1 = 0◦ . We assume the fluid friction to be zero.
But because of fluid friction, the velocity v of the water relative to the buckets
gets smaller as the water flows through the buckets. The velocity v2 of the
water relative to the buckets at exit is usually between 0.8 v1 and 0.9 v1 . In
the development of Equations (8.32)-(8.35) we assumed v = v1 = v2 . This
assumption leads to substantial error. To achieve accurate results, the drop
in v as it passes through the buckets should be taken into consideration.

EXAMPLE 8.12 An impulse with a pitch diameter 3 m and bucket angle


160◦ , is run by a 50 mm diameter jet issuing at 60 m/s. If the speed of runner
is 240 rpm, determine (a) the force on the buckets, (b) the toque on the bucket
and (c) the power transferred to the runner. Assume the velocity relative to
the bucket do not change.

Solution Given, dj = 50 mm, Vj = 60 m/s, β = 160◦ , ω = 240 rpm and


r = 1.5 m. Therefore, the flow rate is

πd2j
Q = Aj Vj = Vj
4
π × 0.052
= × 60
4
= 0.118 m3 /s

The velocity triangles at the entry and exit if the runner are as shown in
Figure S8.12.

V1 u1

u1 v1 β2
V2 v2

At the inlet At the exit


FIGURE S8.12 Velocity triangles at the entry and exit of the runner.

The peripheral speed is

u1 = ωr
( )
240
= × 2π × 1.5
60
= 37.7 m/s

ISTUDY
528 CHAPTER 8. HYDRAULIC MACHINERY

The relative velocity at the entry is

v1 = V1 − u1
= 60 − 37.7
= 22.3 m/s

Also, given that v1 = v2 . Thus the relative velocity, in the horizontal direction,
at the exit is

v2u = v2 cos 160◦


= 22.3 × cos 160◦
= −20.96 m/s

Hence

∆vv = v2u − v1u


= −20.96 − v1
= −20.96 − 22.3
= −43.26 m/s

Also,

V2u = u + v2 cos β
= 37.7 + 22.3 × cos 160◦
= 37.7 − 20.96
= 16.74 m/s

Thus,

∆Vu = V2u − V1u


= 16.74 − 60
= −43.26 m/s

Thus ∆Vu = ∆vv .

The force transferred to the runner is

F = ρQ∆v
= 103 × 0.118 × 43.26

= 5.1 kN

ISTUDY
8.16. IMPULSE TURBINE OPERATION 529

(b) The torque on the runner is

T = Fr
= 5.1 × 1.5
= 7.65 kJ

(c) The power transferred to runner is


P =
746
( ) ( )
7.65 × 103 240
= × × 2π
746 60

= 257.7 hp

EXAMPLE 8.13 A Pelton wheel is supplied with water under a head of


30 m at a rate of 41 m3 /minute. The buckets deflect the jet through an angle
of 160◦ and the mean speed of the bucket is 12 m/s. Calculate the power and
the hydraulic efficiency of the machine.

Solution The velocity triangles at the inlet and outlet are as shown in
Figure S8.13.

(Inlet velocity triangle)


Plane of wheel
V1

u1 v1 160◦

v2
α β V2
u2
(Outlet velocity triangle)

FIGURE S8.13 Velocity triangles at the entrance and exit of the wheel.

The absolute velocity of the jet at entry to the bucket is V1 .

The mean bucket speed is u and v1 is the velocity of jet relative to bucket at
entry.

In the outlet triangle,

V2 is the absolute velocity of water leaving the bucket.

u2 = u is the mean bucket speed.

v2 is the relative velocity of water leaving the bucket.

ISTUDY
530 CHAPTER 8. HYDRAULIC MACHINERY

For the given head of h = 30 m, the absolute velocity of the jet at entry to
the bucket is

V1 = 2gh

= 2 × 9.81 × 30
= 24.26 m/s

Component of final absolute velocity in this direction is V2 cos β.

From the outlet velocity triangle,

V2 cos β = u − v2 cos α
= u − v2 cos (180◦ − 160◦ )

If there is no friction on the surface of the bucket, the water enters and leaves
with the same relative velocity. Therefore,

v1 = v2 = (V1 − u)

Thus

V2 cos β = u − (V1 − u) cos (180◦ − 160◦ )


= 12 − (24.26 − 12) cos 20◦
= 0.479

The force exerted on the bucket, from Equation (8.32), is

F = ρQ(V1 − V2 cos β)
41
= 103 × × (24.26 − 0.479)
60
= 16.25 × 103 N

Thus the power output is

P = Fu
= (16.25 × 103 ) × 12

= 195 kW

Power supplied to the nozzle is

Pin = ρghQ
41
= 103 × 9.81 × 30 ×
60
= 201.1 kW

ISTUDY
8.16. IMPULSE TURBINE OPERATION 531

The hydraulic efficiency is


Power output
ηh =
Power input
195
=
201.1
= 0.97

= 97%

EXAMPLE 8.14 A Pelton wheel is to be selected to drive a generator


at 600 rpm. The water jet is 75 mm in diameter and has a velocity of 100
m/s. If the blade angle is 170◦ and the ratio of vane speed to initial speed
is 0.47, neglecting losses, determine (a) the diameter to the centreline of the
bucket (vanes), (b) the power developed, and (c) kinetic energy per unit mass,
remaining in the fluid.

Solution Referring to Figure 8.17, it is given that, ω = 600 revolutions per


minute, V1 = 100 m/s, djet = 75 mm, β = 170◦ .

Also,
Vane speed
= 0.47
V1
(a) Therefore, the vane speed is

u = 0.47V1
ω r = 0.47 × 100
= 47

The vane speed is


( )
600 d
ωr = × 2π ×
60 2
( )
600 d
= × 2π ×
60 2
d
= 62.83 ×
2
d
47 = 62.83 ×
2

2 × 47
d=
62.83
= 1.496 m

ISTUDY
532 CHAPTER 8. HYDRAULIC MACHINERY

(b) The flow rate is


Q = AV1
π × 0.0752
= × 100
4
= 0.442 m3 /s
By Equation (8.34), the power is
P = ρQ(V1 − u)(1 − cos β)u
= 103 × 0.442 × (100 − 47) × (1 − cos 170◦ ) × 47
= 103 × 1101.022 × (1 − (−0.9848))

= 2185.3 kW

(c) The absolute velocity components leading to vane are


Vx = v1 cos β + u
= (V1 − u) × cos 170◦ + u
= (100 − 47) × cos 170◦ + 47
= −5.2 m/s
Vy = v1 sin β
= (V1 − u) × sin β
= (100 − 47) × sin 170◦
= 9.2 m/s
Thus the kinetic energy remaining in the flow per unit mass is
1
ke = mV 2
2
1 √
= × 1 × vx2 + Vy2
2
1 √
= × 1 × (−5.2)2 + 9.22
2
= 5.28 J

8.17 HEAD AND EFFICIENCY OF


IMPULSE TURBINE
The gross head for a power plant is the difference in elevation between head-
water and tailwater, or (y + z) in Figure 8.16. The pressure within the case

ISTUDY
8.17. HEAD AND EFFICIENCY OF IMPULSE TURBINE 533

of an impulse wheel is atmospheric, and consequently this is the pressure at


which the jet is discharged. Thus the portion z is unavoidable; so that part
of the gross head available to the wheel itself is y only. This is also called
the static head. Setting the wheel too near the surface of the tailwater is not
practical because it might be submerged with a rise in tailwater level.
We consider the nozzle an integral part of the impulse turbine. Hence the
net head, or effective head h on the turbine (including nozzle), that is, the
head at the base B of the nozzle in Figure 8.16, is the static head minus the
pipe friction losses. Thus, for net head, we have

pB V2
h= + B = y − hL (8.36)
ρ 2g
The head, or energy, supplied at the base of the nozzle is expended in
four ways. Some energy is lost in fluid friction in the nozzle, known as the
nozzle loss; a portion is expended in fluid friction over the buckets, causing v2
less than v1 ; kinetic energy is carried away in the water discharged from the
buckets (V22 /2g); and the rest is available to the buckets. Thus
( )[ ( )2 ] 2
1 Aj Vj v2 V2
h= 2
−1 1− + k 2 + 2 + hb (8.37)
Cv AB 2g 2g 2g

where Cv is the coefficient of velocity of the nozzle, defined as

V
Cv =
Vi

where Vi is the velocity that would be attained in the jet if the friction is
absent, termed ideal velocity and V is the actual velocity and V is always less
than Vi , Aj is the cross-sectional flow area of the jet, AB is the cross-sectional
flow area of the pipe upstream of the nozzle, Vj is the jet velocity, v2 is the
velocity of water relative to bucket at exit, V2 is the absolute velocity of water
leaving the bucket, hb is the head directly available to the buckets.
The nozzle loss is usually expressed as
kn Vj2 /2g
Typical values of kn , the bucket friction loss coefficient, vary from 0.2 to about
0.6. The greater part of the energy delivered to the buckets is used in driving
the generator, but some of it is used in overcoming mechanical friction in the
bearing and air friction loss. Some energy may also be lost if not all the water
issuing from a nozzle acts effectively on the bucket, known as volumetric loss.
The hydraulic efficiency ηh of impulse turbine is the ratio of the power
transferred directly to the turbine buckets to the power in the flow at the
base of nozzles. Thus, for impulse turbines
ρ QhL hL
ηh = = (8.38)
ρ Qh h

ISTUDY
534 CHAPTER 8. HYDRAULIC MACHINERY

where hL is the effective head available, after accounting for the losses in the
pipeline and nozzle and h is the gross head. The overall efficiency η of an
impulse turbine is less than the hydraulic efficiency ηh because of that part of
energy delivered to the buckets that is lost in the mechanical friction in the
bearings and the windage, further reducing the energy delivered to the output
shaft. The efficiency of an impulse turbine is given by

Output (shaft) power Tω


η= Input power = (8.39)
ρ Qh
where T is torque, ω is angular velocity, ρ is density, Q is volumetric flow rate
and h is net head.

EXAMPLE 8.15 Demonstrate that when taking account of fluid friction


in the bucket of an impeller wheel by kv22 /2g with the values of k ranging
from 0.2 to 0.6, this is nearly equivalent to v2 = 0.8 to 0.9 times v1 .

Solution Let the relative velocity at the entry and exit are v1 and v2 ,
respectively.

By Bernoulli’s equation,
v12 v2
= 2 + hL
2g 2g
v2
where hL = k 2g2 . Therefore,

v22 + kv22 = v12


v22 (1 + k) = v12
v1
v2 = √
1+k
For k = 0.2,
v1
v2 = √
1 + 0.2
= 0.913 v1

For k = 0.6,
v1
v2 = √
1 + 0.6
= 0.79 v1

EXAMPLE 8.16 A single-jet Pelton wheel with a head over the nozzle of
210 m has its buckets on a circle of 0.9 m diameter. Find the best speed and
the hydraulic efficiency of the runner, if the Cv for the nozzle is 0.975.

Solution Given, h = 210 m, d = 0.9 m and Cv = 0.975.

ISTUDY
8.17. HEAD AND EFFICIENCY OF IMPULSE TURBINE 535

The gross velocity of the jet is



Vj = 2gh

= 2 × 9.81 × 210
= 64 m/s
The velocity coefficient is 0.975. Therefore, the actual velocity of the jet
becomes
V = Cv Vj
= 0.975 × 62.4
= 62.4 m/s
Thus the effective head hL becomes
V2
hL =
2g
62.42
=
2 × 9.81
= 198.46 m
The hydraulic efficiency, by Equation (8.38), is
hL
ηh =
h
198.46
=
210
= 0.945

= 94.5%
For the best speed of the runner,
V = 2u
where u is the peripheral speed, given by
u = ωr
Therefore,
(ω )
62.4 = 2 × × 2π × r
60
62.4 × 60
ω=
4π × 0.45
= 662 rpm

ISTUDY
536 CHAPTER 8. HYDRAULIC MACHINERY

EXAMPLE 8.17 A pelton wheel has bucket diameter 2 m and the deflecting
angle of the bucket is 160◦ . The jet is 165 mm in diameter and the pressure
behind the nozzle is 1 MPa and the wheel rotates at 320 rpm. Neglecting
friction, find the power developed by the wheel and the hydraulic efficiency.

Solution Refer to Figure 8.17. Given, d = 2 m, β = 160◦ , dj = 0.165 m,


h0 = 1 MPa and ω = 320 rpm.

The jet velocity is



2 × 106
Vj =
103
= 44.7 m/s

The flow rate is

Q = Aj Vj
πd2j
= × Vj
4
π × 0.1652
= × 44.7
4
= 0.956 m3 /s

The peripheral speed of the wheel is

u = ωr
( )
320
= × 2π × 1
60
= 33.5 m/s

The power developed, by Equation (8.34), is

P = ρQ(V1 − u)(1 − cos β2 )u

where V1 = Vj , therefore,

P = 103 × 0.956 × (44.7 − 33.5) × (1 − cos 160◦ ) × 33.5

= 695.75 kW

By Equation (8.38), the hydraulic efficiency is

hL
ηh =
h

ISTUDY
8.17. HEAD AND EFFICIENCY OF IMPULSE TURBINE 537

where hL is the head corresponding to the head developed and h is the head
behind the nozzle. Thus,
P
hL =
ρgQ
695.75 × 103
=
103 × 9.81 × 0.956
= 74.64 m
p0
h=
ρg
106
=
103 × 9.81
= 101.94 m
Therefore,
74.64
ηh =
101.94
= 0.732

= 73.2 %
EXAMPLE 8.18 An impulse wheel to drive a generator for 60 Hz power
is to be designed. If the head is 100 m and the discharge is 40 liters/sec,
determine (a) the number of pairs of poles, the diameters of the jet and the
wheel at the centreline of the bucket and the speed of the wheel. Assume,
Cv = 0.98, efficiency 80 percent, rpm 900 and peripheral velocity factor 0.45.

Solution Given, h = 100 m, Q = 40/1000 = 0.04 m3 /s, Cv = 0.98, η = 0.8,


n = 900 rpm and ϕ = 0.45.

The power input is


P = ρgQhη
= 103 × 9.81 × 0.04 × 100 × 0.8
= 31.39 kW
The relative speed, n, in rpm is
7200
N=
n
where N is the number of poles of the generator. Therefore,
7200
N=
900
=8

ISTUDY
538 CHAPTER 8. HYDRAULIC MACHINERY

Thus the number of pairs of poles is 4 .

The jet velocity is



V = Cv 2gh

= 0.98 × 2 × 9.81 × 100
= 43.4 m/s

The volume flow rate is

πd2
Q = AV = V = 0.4
4

Therefore, the diameter of the jet, d, is



4Q
d=
πV

4 × 0.04
=
π × 43.4
= 0.271 m

The peripheral speed is



u1 = ϕ 2gh

= 0.45 × 2 × 9.81 × 100
= 19.93 m/s

The peripheral speed in terms of ω is

u1 = ωR

where R is the radius of the bucket. Thus


ωD
u1 =
2
2u1
D=
ω
2 × 19.93
=
(900/60) × 2π
= 0.423 m

= 423 mm

ISTUDY
8.18. NOZZLES FOR IMPULSE TURBINES 539

8.18 NOZZLES FOR IMPULSE TURBINES


The rotational speed of a turbine is maintained by varying the flow rate to
match the load on the machine. This is accomplished by varying the position
of the needle nozzle. The shape of both nozzle tip and nozzle should be such
as to cause a minimum friction loss for all positions of the needle and also
such as to avoid cavitation damage to the needle at any position.
For attaining efficiency in an impulse wheel, the jet has to be uniform with
all fluid elements moving in parallel lines with equal velocities and with no
spreading at the jet. Air friction retards the water at the outer surface of
the jet, and the needle cause the velocity in the centre to be slightly reduced.
Value of Cv for needle nozzle varies from about 0.95, when partly closed to
permit one-half of maximum flow, to 0.99, at the fully opened position.
For a given pipeline there is a unique jet diameter that will deliver max-
imum power to a jet. The power of a jet issuing from a nozzle (Figure 8.16)
may be expressed as

VJ2
Pjet = ρ Q (8.40)
2g

where Vj is the jet velocity (Vj = V1 of Equation (8.32)). As the size of the
nozzle opening is increased, the flow rate Q gets larger, while the jet velocity
Vj gets smaller.

8.19 REACTION TURBINE


A reaction turbine is that in which flow takes place in a closed chamber under
pressure. The flow through a reaction turbine may be radially inward, axial,
or mixed (that is partially radial and partially axial). There are two types
of reaction turbine in general use, the Francis Turbine and the axial-flow (or
propeller) turbine. All inward-flow turbines are known as Francis Turbine.
In the reaction turbine or pressure turbine the fluid is fed to the runner all
around the circumference from a volute casing through a ring of stationary
guide vanes, which produce a velocity of whirl. The fluid in the runner is
still under pressure which is converted to kinetic energy in the runner pas-
sages producing a reaction on the runner. Because the water in the runner
is under pressure a reaction turbine must always run full. It need not be
submerged but can be fitted with a draft tube. Since at tailrace level the
pressure is atmospheric the pressure at the exit from the runner will be below
atmospheric.
In the Francis Turbine water enters the scroll case and moves into the
runner through a series of guide vanes with contracting passages that convert
pressure head to velocity head. These vanes, known as wicket gates, are ad-
justable so that the quantity and direction of flow can be controlled. Constant
rotational speed of the runner under varying load is achieved by a governor

ISTUDY
540 CHAPTER 8. HYDRAULIC MACHINERY

that actuates a mechanism that regulates the gate opening. A relief valve or a
surge tank is generally necessary to prevent serious water hammer pressures.
Flow through the Francis runner is at first inward in the radial direction,
gradually changing to axial. Therefore, this kind of turbines are also called
mixed-flow turbines. They are usually mounted on a vertical axis. The scroll
case of a Francis turbine is designed to decrease the cross-sectional area in
proportion to the decreasing flow rate passing a given section of the casing.
This is done to maintain constant velocity in the casing so that flow enters the
guide vanes uniformly around the periphery of the vanes. Large turbines are
sometimes provided with an outside guide-vane assembly known as the stay-
ring. This consists of vanes fixed in position known as stay vanes. The vanes
of stay-ring serve as columns to aid supporting the weight of the generator
above, and also they direct the flow smoothly to the inner guide-vane assembly.
For heads below 12 m, Francis turbines are often used with an open flume
setting, as illustrated in Figure 8.19, without a scroll case.

Generator

Slide gate

Trashrack

Headwater level Overflow to


control level

Turbine

Tailwater

Draft tube

FIGURE 8.19 Open flume setting for a reaction turbine at low heads.

The propeller turbine, an axial-flow machine with its runner confined in


a closed conduit, is usually set on a vertical axis. But it may also be set
on a horizontal or slightly inclined axis, as shown in Figure 8.20. The usual
runner has four to eight blades mounted on a hub, with very little clearance
between the blades and the conduit wall. There blades have free outer ends
like a marine propeller. Adjustable gates upstream, of the runner are used to
maintain constant speed.
A Kaplan turbine is a propeller turbine with movable blades whose pitch
can be adjusted to suit existing operating conditions. The adjustment is done
by a mechanism in the runner hub, which is actuated hydraulically by the
governor in synchronization with guide-vane adjustments. With an axial-flow

ISTUDY
8.20. PERFORMANCE OF REACTION TURBINES 541

Gate

Headwater

Generator

Turbine

Tailwater

FIGURE 8.20 A propeller turbine.

turbine, the generator may be set outside the water passageway as shown
in Figure 8.19, or in a streamlined watertight steel housing mounted in the
centre of the passageway.

8.20 PERFORMANCE OF REACTION


TURBINES
The power generated by a Francis turbine can be obtained by considering
the flow trough a radial-flow runner, shown in Figure 8.20. Usually the ra-
dius of fluid flow varies as it flows through a rotor, there it is preferable to
compute torque rather than force. The resultant torque is the sum of the
torques produced by all the elementary forces, but it has been shown by re-
searchers that this sum may be considered as equivalent to two single forces,
one concentrated at the entrance to and the other at the exit from any device.
For steady flow, these equivalent forces are equal to ρ QV1 and ρ QV2 . The
resultant torque can be expressed as

T = ρ Q (r1 V1 cos α1 − r2 V2 cos α2 ) (8.41)


where α1 and α2 are angles between u1 and V1 and u2 and V2 , respectively.
Now the power is given by

P = T ω = ρ Q ω (r1 V1 cos α1 − r2 V2 cos α2 ) (8.42)


This is the power transferred to the shaft of the runner and is termed shaft
power. At this state it is essential to note that, all the water passing through
does not effectively act on the blades, and friction in the runner bearings
slightly reduce the actual power delivered to the runner shaft and generator.
The analysis for mixed-flow runner in which the direction of flow changes
from radial to axial, should be modified to account for the fact that the in-
flow and outflow vectors do not lie in a plane perpendicular to the axis of the
runner. This makes the analysis of a propeller turbine complex.

ISTUDY
542 CHAPTER 8. HYDRAULIC MACHINERY

EXAMPLE 8.19 A radial-flow turbine has 18 blades of each 5 mm thick.


The inlet radius is 250 mm and the exit radius is 150 mm. The angle between
the relative velocity, v, and the tangential velocity of blade, at the inlet and
exit are 65◦ and 122◦ , respectively. The depth of the flow passage between
the two sides of the turbine is 100 mm. The flow rate of water is 0.212 m3 /s
when rotating at 180 rpm. (a) Draw the velocity triangles at the inlet and
exit, (b) find the torque exerted by the water and the horsepower delivered
to the shaft and (c) find the head converted into mechanical work. Assume
that water enters and leaves the blades without shock.
Solution Given, n = 18, t = 5 mm, r1 = 250 mm, r2 = 150 mm, β1 = 65◦ ,
β2 = 122◦ , b = 100 mm, ω = 180 rpm, Q = 0.212 m3 /s.

(a) The velocity diagrams at the inlet (state 1) and exit (state 2) are shown
in Figure S8.19. By continuity,
Q = Ac1 Vn1 = Ac2 Vn2
where Ac1 and Ac2 are the circumferential area and Vn1 and Vn2 are the radial
component of velocity at radii r1 and r2 .
u1 vu1

α1 β1
r1 V V1 v1 vn1
n1

Vu1
(Vn1 = vn1 )

Entrance

u2
α2
r2 β2

V2
v2 vn2
Vn2
(Vn2 = vn2 )
Vu2 vu2
Exit
FIGURE S8.19 Velocity triangles at the entry and exit of the turbine.

The vanes occupy some of the space, therefore the effective area becomes less
that the gross area. Thus the effective area is
Ac = mA = m × (2πrb)
where m is the factor by which the circumferential area reduces and b is the
depth of the flow passage between the sides of the turbine. The reduction

ISTUDY
8.20. PERFORMANCE OF REACTION TURBINES 543

factor is given by
nt
m=1−
2πr

where n is the number of blades and t is the blade thickness.

The reduction factor at the inlet and exit are

18 × 0.005
m1 = 1 −
2π × 0.25
= 1 − 0.0573
= 0.943
18 × 0.005
m2 = 1 −
2π × 0.15
= 0.905

The effective area at the entrance and exit are

Ac1 = m1 2πr1 b
= 0.943 × 2π × 0.25 × 0.1
= 0.148 m2
Ac2 = m2 2πr2 b
= 0.905 × 2π × 0.15 × 0.1
= 0.0853 m2

The radial component of flow velocity at the entrance and exit are

Q
Vn1 =
Ac1
0.212
=
0.148
= 1.43 m/s
Q
Vn2 =
Ac2
0.212
=
0.0853
= 2.48 m/s

ISTUDY
544 CHAPTER 8. HYDRAULIC MACHINERY

Tangential component of relative velocity at the entrance and exit are

Vn1
vu1 =
tan β1
1.43
=
tan 65
= 0.67 m/s
Vn2
vu2 =
tan β2
2.48
=
tan 122
= −1.55 m/s

The rotational speed is

180
ω= × 2π
60
= 18.85 radians/second

The tangential speed of the blades at the entrance and exit are

u1 = r1 ω
= 0.25 × 18.85

= 4.71 m/s
u2 = r2 ω
= 0.15 × 18.85
= 2.83 m/s

The tangential component of velocity at the entrance and exit are

Vt1 = Vu1 = u1 + vu1

= 4.71 + 0.67
= 5.38 m/s

Vt2 = Vu2 = u2 + vu2


= 2.83 − 1.55
= 1.288 m/s

ISTUDY
8.20. PERFORMANCE OF REACTION TURBINES 545

The angle between the absolute velocity V of the fluid and its tangential
component, Vu1 , at the entrance and exit are
( )
Vn1
α1 = tan−1
Vu1
( )
1.43
= tan−1
5.38
= 14.88◦
( )
−1 Vn2
α2 = tan
Vu2
( )
−1 2.48
= tan
1.28
= 62.7◦

The absolute velocity at the entrance and exit are


Vu1
V1 =
cos α1
5.38
=
cos 14.88
= 5.57 m/s
Vu2
V2 =
cos α2
1.28
=
cos 62.7
= 2.79 m/s

(b) The torque exerted by water, by Equation (8.41), is

T = ρQ(r1 Vu1 − r2 Vu2 )


= 103 × 0.212 × (0.25 × 5.38 − 0.15 × 1.28)

= 244.44 N-m

The power delivered to the shaft, by Equation (8.42), is

P = Tω
244.44 × 18.85
=
746
= 6.18 hp

(c) Power is given by


P = ρQgh

ISTUDY
546 CHAPTER 8. HYDRAULIC MACHINERY

where h is the head. Therefore,


P
h=
ρQg
6.18 × 746
=
103 × 0.212 × 9.81
= 2.216 m
EXAMPLE 8.20 A vertical inward flow reaction turbine develops 12500
kW power, using a water flow of 12.3 m3 /s with a head of 115 m. The runner
is of diameter 1.5 m and rotates at 430 rpm. Water enters the runner without
shock with a velocity of 9.6 m/s and passes to the draft tube without whirl
with a velocity of 7.2 m/s. Determine (a) the velocity and direction of the
water entering the runner from the fixed guide blades and (b) the entry angle
of the runner blades.

Solution The velocity diagrams at the inlet (state 1) and exit (state 2) are
shown in Figure S8.20.

u1
w1 α1
Vn1
β1
V1
β2

Vn2
α2 V2
w2
u2
FIGURE S8.20 Velocity triangles at the entry and exit of the runner.

Given, P = 12500 kW, Q = 12.3 m3 /s, h = 115 m, d = 1.5 m, ω = 430 rpm,


Vn1 = 9.6 m/s, Vn2 = 7.2 m/s, w1 = 0, w2 = 0.
u1 = rω
( )
1.5 430
= × × 2π
2 60
= 33.7 m/s
(a) By Equation (8.42), the power is
P = ρQω(r1 V1 cos α1 − r2 V2 cos α2 )

ISTUDY
8.20. PERFORMANCE OF REACTION TURBINES 547

There is no swirl at the exit, therefore, r2 V2 cos α2 = 0. Thus

P = ρQωr1 V1 cos α1
= ρQωr1 Vt1
P
Vt1 =
ρQωr1
12500 × 103
=
103 × 12.3 × ( 430
60 × 2π) ×
1.5
2

= 30.1 m/s

Thus, the velocity at the runner entrance is



V1 = Vn1 2 +V2
t1

= 9.62 + 30.12

= 31.6 m/s

The angle α1 is

Vn1
tan α1 =
Vt1
9.6
=
30.1
= 0.319
α1 = tan−1 (0.319)

= 17.69◦

(b) The blade angle β1 can be determined as follows.

u1 − Vt1
tan (90 − β1 ) =
Vn1
33.77 − 30.1
=
9.6
= 0.382
90 − β1 = tan−1 (0.382)
= 20.9◦
β1 = 90 − 20.9
= 69.1◦

ISTUDY
548 CHAPTER 8. HYDRAULIC MACHINERY

Thus, the entry angle to the runner blade becomes

β2 = 180 − β1
= 180 − 69.1

= 110.9◦

8.21 DRAFT TUBES AND EFFECTIVE


HEAD ON REACTION TURBINES
For proper operation, reaction tubes must have a submerged discharge. The
water, after passing through the runner, enters the draft tube which directs
the water to the discharge point 2, shown in Figure 8.21. The draft tube is
an integral part of a reacting turbine. The draft tube enables the turbine
to be set above tailwater level without losing any head thereby. A reduced
pressure is produced at the upper end, marked point 1 in Figure 8.21, of the
draft tube, which stipulates a limit to the height z1 above tailwater at which
the turbine runner can be set.
VB2
A EL hL 2g

HGL

Generator pB
ρg
Net Gross
head head

VC2
2g
B 1 zB
z1 C
Turbine
Tailrace

Draft tube

FIGURE 8.21 Illustration of net head h and draft head zB on a


reaction turbine.

The draft tube also reduces the head loss at submerged discharge and
thereby increases the net head available to the turbine runner. This is accom-
plished by using a gradually diverging tube whose cross-sectional area at dis-
charge is considerably larger than the cross-sectional area at the tube entrance.
The absolute pressure head at the entry to the draft tube
(Figure 8.21) is

(p1 )abs patm V2 V2


= − z1 − 1 + hl + 2 (8.43)
ρ ρ 2g 2g

ISTUDY
8.22. EFFICIENCY OF TURBINES 549

where z1 is the elevation of the entrance of the draft tube above the surface
of the water in the tailrace, patm is the atmospheric pressure, hl is the head
V2
loss in the diverging tube, and 2g2 is the kinetic energy head at the exit of the
tube. The head loss in the diverging tube can be estimated using the relation
2
(Vi − Ve )
hl = K ′ (8.44)
2g
where Vi , Ve are the velocity at the inlet and exit of the tube, K ′ is the loss
coefficient, which is a function of cone angle, and g is gravitational accelera-
tion. To prevent cavitation, the distance z1 (Figure 8.21) from the tailwater
to the draft tube inlet should be limited so that no point within the draft
tube or turbine will be the absolute pressure drop to the vapor pressure of
the water.
The net head h for a reaction turbine is the difference between the energy
level just upstream of the turbine and that of the tailrace. Thus in Figure
8.21 the net head on the turbine is

h = hB − hC
or
( )
pB V2 V2
h= zB + + B − c (8.45)
ρ 2g 2g

where zB is the draft head and VC is the velocity in the tailrace. In most
V2
instances 2gc is very small and may be neglected.
The effective head heff that is available to act on the runner of a reaction
turbine is [ 2 ]
2
′ (V1 − V2 ) V2 VC2
heff = h − K − − (8.46)
2g 2g 2g
where h is the net head, and the other two terms refer to the head loss in the
draft tube Equation (8.44) and the loss at submerged discharge from the tube
(the last group in Equation (8.46)).

8.22 EFFICIENCY OF TURBINES


The efficiency of turbines is defined as

Power delivered to the shaft (brake power) Tω


η= = (8.47)
Power taken from the water ρ Qh

where T is the torque delivered to the shaft by the turbine, ω is the rotational
speed in radians per second, Q is the flow rate, and h is the net head on the
turbine.

ISTUDY
550 CHAPTER 8. HYDRAULIC MACHINERY

The efficiency of various types of turbine with power output is shown in


Figure 8.22. It is seen that the impulse turbine maintains high efficiency over
a wide range of power output (loads), with significant decrease in efficiency
occurring when the load drops below about 30%. The efficiency of propeller
turbine is very sensitive to load, and a significant drop in efficiency is experi-
enced as the load falls below normal load, The Kaplan turbine maintains high
efficiency over a wide range of load.

100

80 1
4 2
Efficiency, %

3
60

1 Impulse turbine
2 Francis turbine
40 3 Propeller turbine
4 Kaplan turbine

20

0
0 20 40 60 80 100
Percent of rated power output
FIGURE 8.22 Variation of efficiency with power output.

Large turbines have slightly higher efficiency than small ones, because
mechanical friction and windage losses do not increase at the same rate as
hydraulic losses. The effect of size (that is, diameter D) on turbine efficiency is
of importance in transferring test results on small models to their prototypes.
For turbines, this can be accomplished with reasonable accuracy by the Moody
step-up formula,

( )1/5
1 − ηm Dp
≈ (8.48)
1 − ηp Dm

where the subscripts m and p refer to the model and prototype, respectively,
D is the diameter of the runner at flow entrance, for Francis turbine and
represents the blade-tip diameter for the axial-flow turbine.

ISTUDY
8.23. SIMILARITY LAW FOR REACTION TURBINE 551

8.23 SIMILARITY LAW FOR REACTION


TURBINE
For a turbine, operation under a certain head, which is fixed by the natural
characteristics of the site at which it is located, is of interest. Therefore, when
dealing with turbines, it is convenient to have h in the similarity expressions.
From Equation (8.22), it is seen that the head is proportional to the rpm and
diameter, thus
h ∝ n2 D2 or h = kn n2 D2
This can be arranged as
√ √
h h
n∝ or n = kn (8.49)
D D
Substituting Equation (8.49) into Equation (8.21), we get
(√ )
h √
Q ∝ nD ∝ 3
D3 ∝ h D2
D

Hence, for homologous reaction turbines



Q = kq h D 2 (8.50)

The power is given by


P ∝ ρ Qh

From Equation (8.50), Q ∝ h D2 , therefore, for a given value of ρ we get

P ∝ h D2 h ∝ h3/2 D2

Thus, for homologous reaction turbines with a given value of ρ, the power is

P = kP h ∝ h3/2 D2 (8.51)

where kP is a constant.

EXAMPLE 8.21 (a) Derive an expression for the specific speed of a turbine
in terms of its rotational speed n, output power P and head h. (b) At a new
hydro electric station the available head is to be 60 m when the water flow
rate is 32.3 m3 /s. Francis turbines of a specific speed 190 are to be installed
and are to run at 500 rpm with an overall efficiency of 82 percent. Determine
the maximum power available from the turbines and the number of turbine
units required.

ISTUDY
552 CHAPTER 8. HYDRAULIC MACHINERY

Solution

(a) From Equation (8.51) the power is


P ∝ h3/2 D2
where h is the head and D is the diameter. Thus,
P 1/2
D∝
h3/4
The runner speed in rpm will depend on the velocity of flow and the runner
diameter. By Equation (8.49),

peripheral speed h
n= ∝
runner diameter D
Substituting for D, we have
√ 3/4
hh
n∝
P 1/2
h5/4

P 1/2

The specific speed is defined as the value of n when h = 1 m and P = 1 kW,


therefore,
h5/4
n = ns 1/2
P
Thus the specific speed is
n P 1/2
ns =
h5/4

(b) Given, η = 0.82, Q = 32.3 m3 /s, ns = 190, n = 500 and h = 60 m.

The total power available is


Pavai = ηQρgh
= 0.82 × 32.3 × 103 × 9.81 × 60
= 15589.7 kW
Power per one unit, from the specific speed relation above is
( n )2
s
P = h5/2
n
( )2
190
= × 605/2
500
= 4026.66 kW

ISTUDY
8.24. PERIPHERAL-VELOCITY FACTOR AND SPECIFIC ... 553

Thus the number of turbine units requires is


Pavai
Number of units =
P
15589.7
=
4026.66
≈ 4

8.24 PERIPHERAL-VELOCITY FACTOR


AND SPECIFIC SPEED OF TURBINES
For turbines, the peripheral-velocity is given by

u1 = ϕ 2gh (8.52)
where u1 is the tangential velocity of a point on the periphery of the rotating
element and ϕ is the peripheral-velocity factor. The value of ϕ varies over a
wide range during different conditions of operation.
From Equation (8.27), we have
√ √
60 2g ϕe h
D=
π ne
This equation is applicable to reaction turbines also, and we can use it to
estimate the diameter of the runners for Francis and Propeller-type turbines,
but not for impulse turbines.
Specific speed of a turbine is the value of the rotational speed (say in rpm),
when the head h is 1 m and the power delivered in 1 kW.
The runner speed in rpm is given by Equation (8.49) is

peripheral speed h
n= ∝
runner diameter D
Also, from Equation (8.51) the power is
P ∝ h3/2 D2
where h is the head and D is the diameter. Thus

P
D ∝ 3/4
h
Substituting for D, the runner speed becomes
√ 3/4
hh
n∝ √
P
h5/4
∝ √
P

ISTUDY
554 CHAPTER 8. HYDRAULIC MACHINERY

This the specific speed can be expressed as

h5/4
n = ns √
P

8.25 CAVITATION IN TURBINES


Cavitation can result in pitting, mechanical vibration, and loss of efficiency.
The high heads on impulse turbines create extremely high velocities of flow.
If the nozzle and buckets are not properly shaped for the particular flow con-
ditions under which the turbine is operating, separation of the flow from the
boundaries may cause low pressure and result in cavitation. Low-specific-
speed impulse wheels (Ns = 2) can operate at heads as high as 600 m without
cavitation, whereas high-specific-speed impulse wheels (Ns = 8) will experi-
ence cavitation at heads in the vicinity of 120 m.
In the case of reaction turbines, the cavitation is most likely to take place
on the back sides of the runner blades near their trailing edges. Cavitation
may be avoided by operating a turbine in such a manner that at no point the
local absolute pressure drops to the vapor pressure of water.
A convenient parameter to compare the cavitation characteristics of reac-
tion turbine, is the cavitation parameter σ, defined as
patm pv
− − zB
ρ ρ
σ= (8.53)
h
where h and zB are as illustrated in Figure 8.21. The term
( )
patm pv

ρ ρ
represents the height to which water will rise in a water barometer. At sea
level condition, ( )
patm pv
− ≈ 10 m
ρ ρ
The critical or minimum value σc of the cavitation parameter is the value of
σ below which cavitation will occur. From Equation (8.53), it is seen that σ
tends to decrease with decrease of patm , increase of pv , large values of draft
head zB , and high values of net head h. Thus

patm
ρ − pv
ρ
hatm = (8.54)
σc
or
patm pv
(zB )atm = − − σc h (8.55)
ρ ρ

ISTUDY
8.26. PUMP TURBINES 555

Equation (8.54) gives the maximum net head up to which a reaction turbine
can operate without cavitation, and Equation (8.55) gives the maximum draft
head up to which a reaction turbine can operate without cavitation.

8.26 PUMP TURBINES


Pump-turbine hydraulic machines are similar in design and construction to
the Francis turbine. When water enters the rotor at the periphery and flows
inward the machine acts as a turbine. With water entering at the centre (that
is, eyes) and flowing outward, the machine acts as a pump. The direction of
rotation is opposite in the two cases. The pump turbine is connected to a
motor generator, which acts as either a motor or generator depending on the
direction of rotation.
The pump turbine is used at pumped-storage hydroelectric plants, which
pump water from a lower reservoir to an upper reservoir during off-peak load
periods so that water is available to drive the machine as a turbine during the
time when peak power generation is required.

8.27 SUMMARY
Hydraulic machinery are devices to convert one form of fluid energy to an-
other. They are broadly classified as pumps, turbines, blowers, fans and com-
pressors. The devices which convert mechanical energy to fluid energy are
called pumps. The devices meant for converting fluid energy to mechanical
energy are termed turbines. The conversion of mechanical energy to fluid en-
ergy is accomplished by pumps for the case of incompressible fluids, and by
blowers, fans, and compressors for the case of compressible fluids.
Pumps are broadly classified into centrifugal pumps and axial-flow pumps.
The rotating element of a centrifugal pump is called the impeller. The impeller
may be shaped to force water outward in a plane at right angles to its axis
(radial flow), to give the water an axial as well as radial velocity (mixed
flow), or to induce a spacial flow on coaxial cylinders in an axial direction
(axial flow). Radial-flow and mixed-flow machines are commonly referred to
as centrifugal pumps.
Pumps can be of single-stage or multistage. A single-stage pump has one
impeller, while a multistage pump has two or more impellers arranged in such
a way that the discharge from one impeller enters the eye of the next impeller.
Deep-well pumps are usually multistage, having several impellers on a
vertical shaft suspended from a prime mover, usually an electric motor, located
at the ground surface. Each impeller discharges into a fixed-vane diffuser, or
bowl, coaxial with the drive shaft, which directs water to the next impeller.
The head, h, delivered by the pump to the fluid is
( ) ( )
pd Vd2 ps Vs2
h = Hd − Hs = + + zd − + + zs
ρg 2g ρg 2g

ISTUDY
556 CHAPTER 8. HYDRAULIC MACHINERY

where the subscript d and s, respectively, refer to the discharge and suction
sides of the pump.
The power required to drive the pump is the brake horsepower.

bhp = ωT

The efficiency, η, of the pump is defined as


Pw ρgQH
η= =
bhp ωT
The efficiency is basically composed of three parts: volumetric, hydraulic, and
mechanical. The volumetric efficiency is
Q
ηv =
Q + QL
where QL is the loss of fluid due to leakage in the impeller-casing clearances.
The hydraulic efficiency is
hf
ηh = 1 −
hs
The mechanical efficiency is
Pf
ηm = 1 −
bhp
The total efficiency is given by

η = ηv ηh ηm

The Bernoulli equation in rotating coordinates is

p w2 r2 ω 2
+z+ − = constant
ρg 2g 2g
The power input to the pump, delivered to the pump shaft by the motor, is
called the shaft power or the break power. The power output from the pump,
delivered to the fluid, usually water, is called the fluid power or the water
power. The efficiency of a pump is given by

Output power Fluid power ρ gQh


η = Input power = =
Shaft power Tω
Similarity laws are based on the concept that two geometrically similar
machines with similar velocity diagrams, at entrance and exit, from the ro-
tating elements are homologous. This means that their behavior will bear a
resemblance to one another.
The effect of rotational speed on the performance (Q, h, P ) of a given
pump are known as affinity laws.
The efficiency of a pump depends on the condition under which it operates.

ISTUDY
8.27. SUMMARY 557

The capacity of a pump is called normal or rated capacity when its oper-
ating point corresponds to the point of optimum efficiency or best efficiency
power (BEP).
The choice of a pump for a given situation will depend on the rotational
speed of the motor used to drive the pump.
Specific speed is a number that defines the type of pump, such as radial-
flow, axial-flow or mixed-flow. The specific speed Ns of a pump can be ex-
pressed as
( √ )
ω Q
Ns =
(gh)3/4 BEP

The peripheral-velocity
√ factor for a pump impeller is the ratio of the pe-
ripheral velocity to 2gh, usually denoted by ϕ. Thus, the peripheral speed
of the impeller of a pump is

u2 = ϕ 2gh

For an axial-pump u2 is the vane-tip speed.


For pumps, a cavitation parameter, σ, is defined as

(p)abs V2 pv
+ s −
ρg 2g ρg NPSH
σ= =
h h

NPSH is referred to as the net positive suction head.


The value σ, below which there is a drop in pump efficiency is called the
critical cavitation parameter, σc .
For modeling purposes, a parameter called the suction specific speed S, is
defined as √
ω Q
S=
NPSH3/4
If the model and prototype are operated at identical values of Ns and S,
similarity of flow and cavitation is achieved provided the model and prototype
are geometrically similar to one another.
For selecting a pump for a given application, a variety of pumps are avail-
able for choice, based on the performance information of the pump, such as
that in Figures 8.7 and 8.8.
Selection of a pump with operating point close to the BEP will optimize
the pump efficiency, resulting in a minimization of energy required to run the
pump.
Hydraulic turbines are machines used primarily for the development of
hydroelectric energy. These turbines extract energy from flowing water and
convert it to mechanical energy to drive electric generators. Impulse turbine
and reaction turbine are two basic types of hydraulic turbines.
The impulse turbine is also called Pelton wheel. In the impulse turbine
the energy of the fluid supplied to the machine is converted by one or more

ISTUDY
558 CHAPTER 8. HYDRAULIC MACHINERY

nozzles into kinetic energy. The force exerted by the water in the bucket in
the direction of motor in terms of absolute velocity is

F = ρ Q∆V = ρ Q (V1 − V2 cos α2 )

The power transmitted to the buckets is

P = F u = ρ Q (V1 − u)(1 − cos β2 ) u

The gross head for a power plant is the difference in elevation between
headwater and tailwater.
The net head is given by,

pB V2
h= + B = y − hL
ρ 2g

The head, or energy, supplied at the base of the nozzle is expended in


four ways. Some energy is lost in fluid friction in the nozzle, known as the
nozzle loss; a portion is expended in fluid friction over the buckets, causing v2
less than v1 ; kinetic energy is carried away in the water discharged from the
buckets (V22 /2g); and the rest is available to the buckets. Thus

( )[ ( )2 ] 2
1 Aj Vj v22 V22
h= − 1 1 − + k + + hb
Cv2 AB 2g 2g 2g

where Cv is the coefficient of velocity of the nozzle, defined as

V
Cv =
Vi

The nozzle loss is usually expressed as

kn Vj2 /2g

Typical values of kn , the bucket friction loss coefficient, vary from 0.2 to about
0.6.
The hydraulic efficiency ηh of impulse turbine is

ρ QhL hL
ηh = =
ρ Qh h

where hL is the effective head available, after accounting for the losses in the
pipeline and nozzle and h is the gross head.
The overall efficiency η of an impulse turbine is given by

Output (shaft) power Tω


η= Input power =
ρ Qh

ISTUDY
8.27. SUMMARY 559

where T is torque, ω is angular velocity, ρ is density, Q is volumetric flow rate


and h is net head.
A reaction turbine is that in which flow takes place in a closed chamber
under pressure. The flow through a reaction turbine may be radially inward,
axial, or mixed (that is partially radial and partially axial). There are two
types of reaction turbine in general use, the Francis Turbine and the axial-flow
(or propeller) turbine. All inward-flow turbines are known as Francis Turbine.
In the reaction turbine or pressure turbine the fluid is fed to the runner all
around the circumference from a volute casing through a ring of stationary
guide vanes, which produce a velocity of whirl.
In the Francis Turbine water enters the scroll case and moves into the
runner through a series of guide vanes with contracting passages that convert
pressure head to velocity head. These vanes, known as wicket gates, are
adjustable so that the quantity and direction of flow can be controlled.
Flow through the Francis runner is at first inward in the radial direction,
gradually changing to axial. Therefore, this kind of turbines are also called
mixed-flow turbines.
Large turbines are sometimes provided with an outside guide-vane assem-
bly known as the stay-ring. This consists of vanes fixed in position known as
stay vanes.
A Kaplan turbine is a propeller turbine with movable blades whose pitch
can be adjusted to suit existing operating conditions.
The power is given by

P = T ω = ρ Q ω (r1 V1 cos α1 − r2 V2 cos α2 )

This is the power transferred to the shaft of the runner and is termed shaft
power.
The net head h for a reaction turbine is the difference between the energy
level just upstream of the turbine and that of the tailrace. The net head on
the turbine is
( )
pB V2 V2
h = zB + + B − c
ρ 2g 2g

The effective head heff that is available to act on the runner of a reaction
turbine is [ 2 ]
2
′ (V1 − V2 ) V2 VC2
heff = h − K − −
2g 2g 2g
The efficiency of turbines is defined as

Power delivered to the shaft (brake power) Tω


η= =
Power taken from the water ρ Qh
where T is the torque delivered to the shaft by the turbine, ω is the rotational
speed in radians per second, Q is the flow rate, and h is the net head on the
turbine.

ISTUDY
560 CHAPTER 8. HYDRAULIC MACHINERY

Moody step-up formula for turbines is,


( )1/5
1 − ηm Dp

1 − ηp Dm

where the subscripts m and p refer to the model and prototype, respectively,
D is the diameter of the runner at flow entrance, for Francis turbine and
represents the blade-tip diameter for the axial-flow turbine.
For turbines, the peripheral-velocity is given by

u1 = ϕ 2gh

where u1 is the tangential velocity of a point on the periphery of the rotating


element and ϕ is the peripheral-velocity factor.
Specific speed of a turbine is the value of the rotational speed (say in rpm),
when the head h is 1 m and the power delivered in 1 kW.
A convenient parameter to compare the cavitation characteristics of reac-
tion turbine, is the cavitation parameter σ, defined as
patm pv
− − zB
ρ ρ
σ=
h
where h and zB are as illustrated in Figure 8.21.
Pump-turbine hydraulic machines are similar in design and construction
to the Francis turbine. The pump turbine is used at pumped-storage hydro-
electric plants, which pump water from a lower reservoir to an upper reservoir
during off-peak load periods so that water is available to drive the machine
as a turbine during the time when peak power generation is required.

8.28 PROBLEMS
8.1 At the best efficiency point, a centrifugal pump, produces 6.68 m head at
70 m3 /hour. If the specific speed is 0.742, determine the rpm of the impeller.
[Ans. 1172 rpm]

8.2 A centrifugal pump operating at 1170 rpm delivers 75 m3 /hour. If it


operates at 1750 rpm, what will be the flow rate?
[Ans. 112.18 m3 /hour]

8.3 A centrifugal pump provides a flow rate of 0.03 m3 /s when rotating at


1750 rpm against 50 m head. Determine the flow rate of the pump and the
head developed if the pump speed is raised to 3500 rpm.
[Ans. 0.06 m3 /s, 200 m]

8.4 Water at 35 m3 /hour enters a centrifugal pump impeller axially through


a 32 mm diameter inlet. The inlet velocity is axial and uniform. The out-
let diameter is 100 mm. Flow leaves the impeller at 3 m/s relative to the

ISTUDY
8.28. PROBLEMS 561

radial blades. If the impeller rpm is 3450, determine the exit width and the
maximum torque input to the impeller.
[Ans. 10.3 mm, 8.76 N-m]

8.5 Water enters radially an impeller of a pump and discharged with a velocity
whose radial component is 1.5 m/s. The diameter of the impeller is 1.2 m
and peripheral speed is 9 m/s. The vanes are curved backwards at the exit
and make an angle of 30◦ with the periphery. If the pump discharges 3.4
m3 /minute, what will be the turning moment on the shaft.
[Ans. 217.73 N-m]

8.6 A centrifugal water pump (Figure 8.5) has r2 = 300 mm, r1 = 100
mm, β1 = 20◦ , β2 = 10◦ . The impeller is 50 mm wide at r = r1 , and 20
mm at r = r2 . If the rpm is 1800, neglecting the looses and vane thickness,
determine (a) the discharge at shock-less entrance with α1 = 90◦ , (b) α2
and the theoretical head h, (c) the horsepower required and (d) the pressure
through the impeller.
[Ans. (a) 0.2155 m3 /s, (b) 13.35◦ , 139 m, (c) 394 hp, (d) 1.08 MPa]
8.7 A centrifugal pump is to be placed above a large, open water tank and
is to pump water at 0.014 m3 /s. At this flow rate the required net positive
suction head, NPSH, is 4.6 m. If the water temperature is 10◦ C (pv = 2.34
kPa) and atmospheric pressure is 101 kPa, determine the maximum height
that the pump can be located above the water surface without cavitation.
Assume that the major head loss between the tank and pump inlet is due to
a filter at the pump inlet having minor loss coefficient, K = 20. Other losses
are neglected. The pipe on the suction side of the pump has a diameter of
100 mm. [Ans. 2.231 m]
8.8 Water is pumped at 5.32 m3 /minute through a centrifugal pump oper-
ating at 1750 rpm. The impeller has a uniform blade height of 50 mm, with
r1 = 45 mm, r2 = 180 mm and the exit blade angle is 23◦ (see Figure 8.5).
Assuming ideal flow conditions and that the tangential velocity component
Vt1 of the water entering the blade is zero (α1 = 90◦ ), determine (a) the tan-
gential velocity component Vt2 at the exit, (b) the ideal head rise and (c) the
power transferred to the fluid. [Ans. (a) 29.3 m/s, (b) 98.57 m, (c) 115 hp]

8.9 Water enters a centrifugal impeller axially through a 32 mm diameter


inlet. The flow leaves the impeller of outlet diameter is 100 mm at 3 m/s
relative to the blade angle. If the flow rate is 35 m3 /hour and the impeller
rotates at 3460 rpm, determine (b) the exit width of the impeller and (b) the
minimum torque input to the impeller. [Ans. (a) 10.3 mm, (b) 8.8 N-m]
8.10 A centrifugal pump delivers water at 57.2 m3 /hour when working at
2875 rpm. If the head developed in 42.1 m, determine (a) the specific speed
of the pump and (b) the power input to the pump.
[Ans. (a) 0.414, (b) 6.56 kW]

ISTUDY
562 CHAPTER 8. HYDRAULIC MACHINERY

8.11 An axial flow turbine develops 10000 hp when operating with 12 m


head. Determine the rpm if the maximum efficiency is 88%.
[Ans. 60 rpm]
8.12 A centrifugal pump operates at 300 rpm, delivering water at 6 m3 /s
against 100 m head. Laboratory model for this facility has a maximum flow
rate of 0.28 m3 /s and maximum power of 225 kW. If the efficiencies of the
model and prototype are the same, find (a) the speed of the model and (b)
the scale ratio. (c) Also, calculate the specific speed.
[Ans. (a) 1203 rpm, (b) Dp /Dm = 4.43, (c) 0.07]
8.13 A centrifugal pump, running at 1750 rpm, delivers 0.0036 m3 /s. If the
power to run the pump is 5 hp and the efficiency is 70%, determine the specific
speed.
[Ans. 7.3 × 10−3 ]
8.14 At maximum efficiency a Pelton wheel develops 410 kW. If the flow rate
is reduced by 20 percent by means of a throttle value before the nozzle, what
will be the power output.
[Ans. 196.8 kW]
8.15 A Pelton wheel of bucket diameter 0.9 m, run by a jet of 75 mm diameter,
deflects the water from the bucket at 160◦ . Neglecting the friction find the
power developed by the wheel and the hydraulic efficiency when the speed is
300 rpm and the power behind the nozzle is 690 kPa.
[Ans. 102.2 kW, 91.3%]
8.16 The bucket circle of a Pelton wheel is 1.8 m in diameter. The jet is
100 mm in diameter and the head over the nozzle is 135 h. If the hydraulic
efficiency is 90% while running at 250 rpm, find the flow deflection angle at
the bucket.
[Ans. β2 = 160◦ ]
8.17 A Pelton wheel has a mean bucket speed of 12 m/s is supplied with
water at 0.68 m3 /s under the head of 30 m. If the buckets deflect the jet
through and angle of 160◦ find the power and the efficiency of the wheel.
[Ans. 194 kW, 97%]
8.18 A multi-stage centrifugal pump is required to lift 1.8 m3 /minute of
water from a mine. The net positive suction head required is 750 m. If the
speed of the pump is 2900 rpm, find the least number of stages if the specific
suction speed is 0.323.
[Ans. 10]
8.19 A Pelton wheel is run by a jet from a nozzle at the end of a 1650 long
pipeline which discharges water from a reservoir in which the level of water
is 375 m above that of the wheel. The turbine runs at 500 rpm and develops
5000 kW. If the pipeline losses 10% of the gross head and friction coefficient
for the pipe is 0.005, calculate (a) the diameter of the pipe, (b) the cross-
sectional area of the jet and (c) the mean diameter of the wheel. Assume
bucket speed = 0.46 of the jet speed and efficiency of the turbine is 86%.
[Ans. (a) 0.562 m, (b) 0.0216 m2 , (c) 1.43 m]

ISTUDY
8.28. PROBLEMS 563

8.20 An axial flow turbine of mean diameter 2 m rotates at 145 rpm, when
water at 35 m head is supplied. Water leaves the guide vanes at 30◦ to the
direction of runner rotation. Determine the blade angle at the inlet (at mean
radius).
[Ans. 75◦ ]
8.21 Consider an impulse turbine with a pitch diameter of 3.3 m and bucket
angle 160◦ . If the jet velocity is 65 m/s, jet diameter 50 mm and rotating
speed 240 rpm, find (a) the force on the buckets, (b) the torque on the runner
and (c) the power transferred to the runner. Assume that the velocity relative
to the bucket does not change.
[Ans. (a) 5.824 kN, (b) 9.6 kN-m, (c) 241 hp]

8.22 The impeller of a centrifugal pump has a diameter of 0.1 m and ax-
ial width at outlet of 15 mm. The blade angle are swept backward at 25◦
to the tangent of the periphery. The flow rate through the impeller is 8.5
m3 /hour when it rotates at 750 rpm. Assuming the flow to be ideal and one-
dimensional, calculate (a) the head developed by the pump and (b) the power
supplied to the pump.
[Ans. (a) 1.146 m of water, (b) 26.5 W]

8.23 A centrifugal fan delivering 2 m3 /s of air runs at 960 rpm. The outside
and inside diameters of the impeller is 70 cm and 48 cm, respectively. The im-
peller width at inlet is 16 cm and is designed for constant radial flow velocity.
The blades are backward inclined making an angle of 22.5◦ and 50◦ with the
tangents at inlet and outlet, respectively. Draw the inlet and outlet velocity
triangles and determine the theoretical head produced by the impeller.
[Ans. 84 m of air]

8.24 A Francis turbine operating at 300 rpm develops a shaft horsepower of


1125 hp. Find the torque transmitted from the flowing water to the shaft.
[Ans. 26714 N-m]

8.25 A Pelton wheel has a diameter of 2 m and develops 500 kW when


rotating at 180 rpm. (a) What is the average force of the water against the
blades? (b) If the turbine is operating at the maximum efficiency, determine
the speed of water jet from the nozzle and the mass flow rate.
[Ans. (a) 26.53 kN, (b) 37.7 m/s, 703.71 kg/s]

8.26 A Francis turbine develops a shaft power of 1125 hp when the flow rate
is 3.4 m3 /s. What is the net head on the runner if the turbine efficiency is 85
percent?
[Ans. 29.6 m]

8.27 At maximum efficiency a turbine discharges 10 m3 /s and develops a


shaft power of 2600 kW. If the net head is 30 m, determine the efficiency.
[Ans. 88.34%]

ISTUDY
ISTUDY
Bibliography
Douglas, J.F., Gesiorek, J.M., and Swaffield, J.A., Fluid Mechanics,
4th ed., Pearson Education, 2002.
Finnemore, E.J., and Franzini, J.B., Fluid Mechanics with Engineering
Application, 10th ed., McGraw-Hill, 2002.
Fox, R.W., and McDonald, A.J., Introduction to Fluid Mechanics,
5th ed., John Wiley, 1994.
Francis, J.R.D., Fluid Mechanics for Engineering Students, 5th ed.,
Edward Arnold, 1975.
Hinze, J.O., Turbulence, 2nd ed., McGraw-Hill, New York, 1975.
Hoerner, S.F., Fluid Dynamic Drag, Published by the Author, 1965.
Kundu, P.K., and Cohen, I.M., Fluid Mechanics, 2nd ed., Academic
Press, 2002.
Ligget, J.A., Fluid Mechanics, McGraw-Hill, 1994.
McDonald, A.T., and R.W. Fox, “An Experimental Investigation of
Incompressible Flow in Conical Diffusers’’, International Journal of
Mechanical Sciences, Vol. 8, No. 2, 1966, pp. 125-139.
Rathakrishnan, E., Fundamentals of Engineering Thermodynamics,
2nd ed., Prentice-Hall of India, 2005.
Rathakrishnan, E., Gas Dynamics, Prentice-Hall of India, New Delhi, 1995.
Reneau, L.R., J.P. Johnston, and S.J. Kline, “Performance and Design
of Straight, Two-Dimensional Diffusers’’, Transactions of the ASME,
Journal of Basic Engineering, Vol. 89D, No. 1, March 1967, pp. 141–150.
Runstadler, P.W., Jr., Diffuser Data Book, Technical Note 186, Hanover,
NH: Creare, Inc., 1975.
Schlichting., Boundary Layer Theory, 5th ed., McGraw-Hill, New York,
1960.
Shames, H., Mechanics of Fluids, 3rd ed., McGraw-Hill, New York, 1962.
Streeter, V.L., Benjamin Wylie, E., and Bedford, K.W., Fluid Mechanics,
9th ed., McGraw-Hill, 1998.
White, F.M., Fluid Mechanics, 2nd ed., McGraw-Hill, 1979.

565

ISTUDY
ISTUDY
Index

Absolute pressure, 6 Brake horsepower, 491


Acceleration, 3, 14 Break power, 556
convective, 44 Broad-crested weir, 436
Adverse pressure gradient, 72 Buckingham’s p-theorem, 177
Aerofoil, 51, 58, 101 Buoyancy, 17
Affinity laws, 507, 556 Buoyant force, 17
Alternative depths, 412
Antinodes, 473–482
Area-Mach number relation, 120 Capillary
Atmosphere, isothermal zone of, 24 action, 40
wave, 466
Cauchy-Riemann equations, 66, 140
Back flow, 98, 228 Cavitation parameter, 516, 554
Basic dimensions, 5 Chézy’s coefficient, 396–398
Basic laws, 46 Chézy’s equation, 396
conservation of matter, 46, 49 Choked flow, 121
first law of thermodynamics, 46 Circular cylinder, 88–93
momentum equation, 46 Circular vortex, 298
second law of thermodynamics, 46 size of, 300
Bernoulli equation, 71 velocity distribution of, 300
cautions for use of, 72 Circulation, 58–60, 248–251
compressible, 73 physical meaning of, 293
Biot–Savart law, 268 theorem, 257
Blassius equation, 215 Coefficient of discharge, 438, 445, 449
Bluff body, 97, 101 Colebrook’s equation, 357
Body forces, 10 Compound vortex, 292
Bore, 405 Compressibility, 37
Bound vortex, 260, 273 Conjugate depths, 426, 428
Boundary layer, 51–54, 94 Continuity equation, 46, 48–50, 63
displacement thickness of, 52, 54, 198 Control
energy thickness of, 53, 54, 200 mass, 47
equations, 209 surface, 47
flat plate, 212, 215 volume, 47
in open channels, 398 Convergent-divergent
laminar, 52, 202, 215 nozzle, 120–122
momentum thickness of, 53, 201 passage, 227
solutions, 206 Critical
thickness, 51, 196, 197, 215 depth, 412, 443–446, 448
turbulent, 203, 205, 226 flow, 435, 390

567

ISTUDY
568 INDEX

Reynolds number, 95, 96 Equation of state, 6, 47, 117


slope, 415 Euler’s
speed of sound, 119 acceleration formula, 43
state, 119 equations, 70
velocity, 320, 413 Euler number, 178
lower, 325 Eulerian description, 41
upper, 325 Expansion
Cylinder isentropic, 121, 123
drag of, 94 waves, 116, 122
flow field around, 89, 206
lift force on, 93
Fanno line flow, 117, 130, 132
Field approach, 41
Darcy equation, 110, 329 Flat plate boundary layer, 52–54, 214
Deep-water waves, 465 Flow analysis, 41
d’Alembert’s paradox, 90 Flow(s)
Density, 32 classification of, 8
Differential analysis, 47 fully developed, 108–110, 325–328
Dimensional analysis, 173 with pressure gradient, 227
Dimensionless groups, 173, 177 rate, 74
definition of, 174 Flow through pipes, 108, 319
Dimensions and units, 5 Flow through Venturi, 447–450
Doublet, 78 free discharge condition for, 446
Drag, 94, 97 Fluid, 30
coefficient of, 101 compressibility, 37
Drown-drop or drawn-down curve, 452 as a continuum, 3, 4
Drowned Venturi, 450 definition of, 30
Drowned weir and free outfall, 443 mechanics, definition of, 1
Dynamic analysis, 3 newtonian, 2
Dynamic viscosity, 34 power, 491
statics, 9
basic equation of, 14
Effective head, 533 Fluid-dynamically smooth surface, 339
Efficiency Force on a vortex, 307
hydraulic, 399, 495, 496 Forced vortex, 74, 285
impulse turbine, 533 Form drag, 339
mechanical, 492, 499 Francis turbine, 539
of pump, 499 Free molecular flow, 8
turbines, 546 Free spiral vortex, 288
volumetric, 491, 499 Free vortex, 75, 288
Energy Friction
gradient, 395 coefficient, 101
internal, 38 drag, 101
kinetic, 32, 105, 201 factor, 110
line, 453 Friction factor, 333
mechanical, 70 for rough pipes, 355
transmission rate, 465 for smooth pipes, 349
Energy due to a pair of vortices, 308 variation of, 336
Enthalpy, 38 Friction velocity, 344
Entropy, 125 Froude number, 178

ISTUDY
INDEX 569

Gas constant, 6 Kinetic energy correction factor, 360,


Gas dynamics, 8, 116–131 393, 453
definition of, 1, 8 Kinematic viscosity, 36
hypersonic, 8 Kinematics, 3, 48
rarefied, 8 Knudsen number, 8, 178
subsonic, 8 Kutta–Joukowski theorem, 273
supersonic, 8
General motion, 244
Gradually varied flow, 451 Lagrangian description, 41, 46
Gravity waves, 465 Laminar
Gross head, 523, 532 flow, 107, 215
Group velocity, 470 integral theory, 228
separation, 54
sublayer, 52
Head loss across a jump, 426, 427 Laminar to turbulent transition, 316
Head loss due to friction, 110, 333 Laplace equation, 69–71
Half-body, 82–84 Lapse rate, 23
Head loss, 110, 362 Laval nozzle, 120
Helmholtz’s theorems, 255 Law of dimensional homogeneity, 5, 281
Hydraulic Laws of vortex motion, 254
diameter, 130 Line vortex, 252, 310
grade line (HGL), 491 Linear vortex of finite length, 271
jump, 424 Logarithmic law, 219, 346
mean depth, 396
Hydrostatic pressure distribution, 14, 31
Hypersonic flow, 8 Mach number, 8, 118
definition of, 37
Magnetofluidmechanics, 9
Ideal fluid flow, 7 Magnus effect, 94
Ideal gas, 6 Manometry, 14
Image of a vortex, 305 Mass flow rate, 121
Impact pressure, 31 Material description, 41
Impulse function, 131 Mean free path, 4, 8, 32
Impulse turbine, 521–524 Mild slope, 415
Incompressible flow, 8 Molar mass, 6, 7
Inertia force, 50, 181–183 Momentum correction factor, 360, 393
Infinite vortex, 272 Momentum equation, 46, 50
Integral analysis, 47 Momentum-integral
Internal energy, 38 estimates, 207
Irrotational relation, 208
flow, 59, 66, 70, 82 Momentum thickness, 53, 200
motion, 58 Moody’s diagram, 339, 340
Isentropic Moody step-up formula, 550, 560
flow, 117 Mutual action of two vortices, 308
index, 117
process equation, 73
relations, 118 Navier–Stokes equations, 51, 138
stagnation state, 118 Net positive suction head, 516
Newtonian fluid, 2, 34
Newton’s second law, 46, 50
Kaplan turbine, 540 Newton’s viscosity law, 34
Kelvin’s circulation theorem, 257–260 Nodes, 473

ISTUDY
570 INDEX

Non-newtonian fluid, 2, 34 Rankine’s theorem, 83


Normal depth, 394 Rapid flow, 444–446
Normal shock Rapid velocity, 390, 412
definition of, 123 Rapidly varied flow, 390, 453
relations, 124 Rayleigh flow, 117, 133
total pressure ratio across, 125 Real gas, 6
Nozzle Laval, 120 Relaminarization, 325
convergent-divergent, 119, 122 Rectilinear vortex, 298
Relative roughness, 335–338
Reynolds number, 94
Oblique shock critical, 95, 96
definition of, 125 lower critical, 95
relations, 126 upper critical, 96
Overexpanded flow, 122 lower limit for, 320
Overlap region, 345 upper limit for, 320
Reynolds stress, 106
Rotational flow, 243, 244
Partially filled flow Rotational motion, 58
in closed conduits, 402–403
Particle description, 41
Pathline, 45–46 Scalar quantities, 10
Pelton wheel, 523 Scale factors, 186
Penstock, 523–524 Seiches, 475
Perfect gas, 6, 117 Semi-infinite vortex, 273
state equation, 6 Separation, 99
Peripheral-velocity Separation point, 54
factor, 515, 553 Shaft power, 499
Phase velocity, 464 Shape factor, 218, 228
Pipe flow, 108–111, 319 Shear stress, 30, 34
Plasma, 9 Shooting velocity, 415, 478
Poisson’s equation, 265 Shutoff head, 494, 508
Poiseuille’s equation, 336 Similarity, 180, 181
Poiseuille’s Law, 335 dynamic, 180, 181
Point rectilinear vortex, 302 geometric, 180, 184
Potential flow, 51, 69 kinematic, 181
Potential function, 66 parameter, 184
Prandtl’s boundary layer equations, 212 Single-overhang, 524
Prandtl power law, 224, 383 Sink, 74, 78, 80
Prandtl–Meyer function, 123 Skin friction drag, 98, 100
Pressure, 31 Slip flow, 8
coefficient, 83 Slope
drag, 98 critical, 415
force, 11, 182 mild, 415, 478
gauge, 14 steep, 415, 478
impact, 31 Small jumps, 428
Pure rotation, 58, 244 Specific head, 411
Pure translation, 58, 244 Specific speed
of pump, 512, 517
Source, 74, 78, 80
Radial flow, 288 Specific heat, 38
Rankine’s half-body, 83 Specific heats ratio, 38

ISTUDY
INDEX 571

Specific volume, 6 FPS, 5


Speed of sound, 117, 118 MKS, 5
Stagnation SI, 5
pressure, 73 Undular jump, 428
state, 118 Universal gas constant, 6
Standing waves, 473 Unsteady flow, 41
Starting vortex, 260, 274, 315
State equation, 6, 7, 47
Static pressure, 31 Vector quantities, 10
Statics, 3 Velocity
Steady flow, 41 friction, 344
Steady uniform flow, 394 lower critical, 325
Steady wave, 414 upper critical, 325
Stoke, 36 Velocity defect law, 344
Stokes’ theorem, 261 Velocity distribution
Stratosphere, 23 in turbulent flow, 343
Streakline, 45 in smooth pipes, 345
Stream function, 62 for rough pipes, 355
Streamline, 46 Velocity potential, 66, 74
Streamlined body, 97, 101 Velocity of sound, 117, 118
Subcritical flow, 415 Viscosity, 33
Subsidiary laws, 46 coefficient of, 33–36
Suction specific speed, 517 Viscous flows, 94
Supercritical flow, 415 Viscous forces, 182
Supersonic flow, 8 Vortex, 75
Surface forces, 10 circular, 298
Surface tension, 39 free, 75
Systems, 47 forced, 75
Vortex between parallel plates, 305
Vortex motion, 75, 279
Temperature, 32
Vortex pair, 304
sea level, 32
Vortex theorems, 254–258
Tensor quantities, 10
Vorticity, 59
Total pressure, 31
Vorticity equation, 244
Transition flow, 8
Transition point, 54 in polar coordinates, 246
Transition zone, 337–339, 398 in rectangular coordinates, 244
Transonic flow, 8
Tranquil flow, 391
Water hammer, 523, 540
Troposphere, 23
Tsunami, 475 Water horsepower, 491, 494
Turbulence, 102 Wake, 99
description of, 103 Wave
number, 104 expansion, 116, 122
Turbulent flow, 102, 107 shock, 116, 122
Turbulent rough zone, 338 energy, 467
Turbulent smooth zone flow, 338 Mach, 116
Waves and surges in
open channels, 405
Underexpanded flow, 122 Weber number, 178
Units, 5 Whirlpool, 76
CGS, 5 Wicket gate, 539

ISTUDY
Fourth Edition

FLUID MECHANICS
An Introduction
Ethirajan Rathakrishnan
The Fourth Edi on of this easy-to-understand text con nues to provide students with a sound
understanding of the fundamental concepts of various physical phenomena of science of fluid
mechanics. The third edi on of this book, developed to serve as text for a course in fluid
mechanics at the introductory level for undergraduate course and for an advanced level course
at graduate level, was well received all over the world, because of its completeness and proper
balance of theore cal and applica on aspects of this science.
Over the years, the feedback received from the faculty and students made the author to realize
the need for adding following material to serve as text for students of all branches of engineering.
¦ Three new chapters on:
ð Pipe Flows
ð Flow with Free Surface
ð Hydraulics Machinery
¦ Large number of solved examples in all the chapters to enable the user to gain an insight in to
the theory and applica on aspects of the concepts introduced.
¦ A Solu on Manual that contains solu ons to all the end-of-chapter problems for instructors.

THE AUTHOR
ETHIRAJAN RATHAKRISHNAN, Ph.D., is Professor of Aerospace Engineering at the Indian Ins tute
of Technology Kanpur. He is well-known interna onally for his research in the area of high-speed
jets. The limit for the passive control of jets, called Rathakrishnan Limit, is the contribu on to the
field of jet research. The concept of breathing blunt nose (BBN), which reduces the posi ve
pressure at the nose and increases the posi ve pressure at the base simultaneously, is his
contribu on to drag reduc on at hypersonic speeds. Recently he received a ‘Life me
Achievement Award’ from Interna onal Sustainable Avia on and Energy Research Society
(SARES) for his novel scien fic contribu ons to the area of Sustainable Avia on and Energy. He
has published a large number of research ar cles in many reputed interna onal journals. He is a
Fellow of many professional socie es, including the Royal Aeronau cal Society.
Professor Rathakrishnan serves as the Editor-in-Chief of the Interna onal Review of Aerospace
Engineering (IREASE) and Interna onal Review of Mechanical Engineering (IREME) journals. He
has authored eleven books including Fundamentals of Engineering Thermodynamics (2nd ed),
Helicopter Aerodynamics and Gas Dynamics (7th ed), published by PHI Learning, Delhi.

ISBN:978-93-89347-92-0

9 789389 347920
www.phindia.com
ISTUDY

You might also like