You are on page 1of 168

THE WYKEHAM ENGINEERING

AND TECHNOLOGY SERIES

General Editors:
PROFESSOR SIR NEVILL MOTT, F.R.S.
Emeritus Cavendish Professor of Physics
University of Cambridge

G. R. NOAKES
Formerly Senior Physics Master
Uppingham School

Series Editor:
DR. A. T.FULLER
Assistant Director of Research
Engineering Department
University of Cambridge

Tms BOOK is one of a series in which authorities will attempt to


introduce young people to various technologies. Primarily these
books are aimed at young graduates or apprentices who are about to
begin a career in industry or who have just started work. For this
reason more attention has been paid to giving a broad view of each
subject than to ' necessary and sufficient ' mathematical proofs.
Indeed the use of mathematics has been sparing.
Another feature of the Technology series is the mixture of
disciplines. For example, when micro-circuits are being discussed, the
physical basis of semi-conduction phenomena is inextricably mixed up
with the chemistry of almost pure single crystal formation and with
the engineering necessary to produce a monolithic circuit. This
admixture of disciplines is characteristic of industrial research and
development.
It is hoped both by the authors and by the publishers that this series
will open up some new horizons to the young scientist and engineer in
industry.
INTRODUCTION
TO TRIBOLOGY

J. Halling
University of Salford

g WYKEHAM PUBLICATIONS (LONDON) LTD


(A MEMBER OF THE TAYLOR & FRANCIS GROUP)
LONDON AND WINCHESTER
1976
First published 1976 by Wykeham Publications (London) Ltd.

© 1976 J. Halling. All rights reserved. No part of this publication


may be reproduced, stored in a retrieval system, or transmitted, in any
form or by any means, electronic, mechanical, photocopying, recording
or otherwise, without the prior permission of the copyright owner.

Cover illustration-Techniques for studying surfaces (see Chapter 2).

ISBN O 85109 061 3

Printed in Great Britain by Taylor & Francis (Printers) Ltd.


Rankine Road, Basingstoke, Hants.

Distribution and Representative:


UNITED KINGDOM, EUROPE AND AFRICA
Chapman & Hall Ltd. (a member of Associated Book Publishers
Ltd.), 11 North Way, Andover, Hampshire SPl0 5BE.
THE WESTERN HEMISPHERE
Springer-Verlag New York Inc., 175 Fifth Avenue, New York,
New York, 10010.
AUSTRALIA, NEW ZEALAND AND FAR EAST (EXCLUDING JAPAN)
Australia and New Zealand Book Co. Pty. Ltd.,
P.O. Box 459, Brookvale, N.S.W. 2100.
INDIA, BANGLADESH, SRI LANKA AND BURMA
Arnold-Heinemann Publishers (India) Pvt. Ltd.,
AB-9, First Floor, Safjardang Enclave, New Delhi 11016.
GREECE, TURKEY, THE MIDDLE EAST (EXCLUDING ISRAEL)
Anthony Rudkin, The Old School, Speen, Aylesbury,
Buckinghamshire HPl 7 0SL.
ALL OTHER TERRI TORIES
Taylor & Francis Ltd., 10-14 Macklin Street, London WC2B SNF.
PREFACE

TRIBOLOGY is the new name for any problem concerned with the
carrying of load across interfaces in relative motion. Thus, although
the word is new, the subject concerns itself with such well known
topics as friction, wear and lubrication. Although this subject is
important in all industrial machinery, it also has a much wider signi-
ficance. The cleaning of our teeth or the operation of human joints
are obvious examples of tribological phenomena. It is an inter-
disciplinary subject that concerns itself with physics, chemistry,
materials science, engineering and mathematics, often all at the same
time.
This book outlines the basic principles upon which the subject
rests and shows how the various branches of science and engineering
have contributed to our understanding of these principles. The main
theme concentrates on physical principles. The mathematics
included is only given in outline and it is hoped that students will
understand the basic physical arguments even though they may not
wish to become involved in the mathematical detail.
At the end of each chapter I have indicated the types of project
activity which could enlarge our appreciation of the subject matter.
The level is generally that of a first-year undergraduate in science or
technology or a science sixth former embarking on specialization in
science or technology at a university. As with most books, the reward
which any individual will receive from this book will depend on the
time he is prepared to spend not only on understanding what is
written but also on obtaining a deeper feeling for the subject by
carrying out the projects. The book is not intended as a superficial
introduction to the subject.
I must acknowledge that this book depends considerably on the
contents of an earlier book, Principles of Tribology, written by my
colleagues and myself. The project activities have also been inspired
by the booklet Tribology Projects for Schools written by a Committee
on Tribology Panel under my chairmanship. I am very grateful to
Mr. W. E. W. Smith, of Calday Grange Grammar School, for his
many helpful suggestions during the preparation of this book, which
I am sure will enhance its value to schools. Finally, I acknowledge
the enormous help in producing the manuscript by Mrs. L. M.
Chadderton who probably deserves as much credit as I for the final
achievement.
V
CONTENTS

Preface v

Chapter l INTRODUCTION
1. 1 Tribology 1
1.2 The occurrence of tribology 1
1.3 Historical 3
1.4 Tribological solutions 7
1.5 Economic considerations 11
1.6 Conclusion 13
1. 7 Project suggestions 13

Chapter 2 THE NATURE OF SURFACES AND THEIR


CONTACT
2.1 Introduction 16
2.2 The nature of surfaces 16
2.3 The properties of surfaces 18
2.4 Methods of studying surfaces 19
2.5 Some geometrical properties of surfaces 25
2.6 The loading of solids 29
2.7 Contact of rough surfaces 33
2.8 Thermal effects 36
2.9 Project suggestions 38

Chapter 3 FRICTION
3.1 Introduction 41
3.2 Possible causes of friction 42
3.3 The simple adhesion theory 43
3.4 Extensions to the simple theory 46
3.5 Friction between elastic bodies 49
3.6 Rolling friction 50
3.7 The experimental measurement of friction 57
3.8 Project suggestions 60
vu
Chapter 4 WEAR AND THE PROPERTIES OF
MATERIALS
4.1 Introduction 63
4.2 Types of wear 64
4.3 Mechanisms of wear 65
4.4 Factors affecting wear behaviour 72
4.5 Tribological properties of plastics 75
4.6 The measurement of wear 77
4. 7 Project suggestions 81

Chapter 5 PROPERTIES AND TESTING OF


LUBRICANTS
5 .1 Introduction 83
5.2 Viscosity 83
5.3 Effect of temperature on viscosity 84
5 .4 Effect of pressure on viscosity 86
5.5 The measurement of viscosity 87
5.6 Other properties of lubricants 93
5.7 Greases 94
5.8 Project suggestions 96

Chapter 6 EXTERNALLY PRESSURIZED BEARINGS


6.1 Introduction 97
6.2 The simple pad bearing 99
6.3 The characteristics of capillary-compensated
bearings 105
6.4 Hydrostatic journal bearings 107
6.5 Project suggestions 110

Chapter 7 HYDRODYNAMIC LUBRICATION


7.1 Introduction 111
7.2 Basic theoretical considerations 113
7.3 Gas-lubricated bearings 129
7.4 Elastohydrodynamic lubrication (E.H.L.) 130
7.5 Project suggestions 137
V111
Chapter 8 THE SELECTION OF TRIBOLOGICAL
SOLUTIONS
8.1 Introduction 139
8.2 Environment 140
8.3 Load 141
8.4 Speed 144
8.5 Tribological limits of load and speed 145
8.6 The use of tribological limits 150
8. 7 Conclusions 153
Index 155

IX
CHAPTER 1
introduction

1.1. Tribology
THE name ' tribology ' is derived from the Greek word tribos meaning
' rubbing ', so that a literal translation would be ' the science of
rubbing '. This interpretation is a little too narrow, so that the word
is actually defined as: ' The science and technology of interacting
surfaces in relative motion and of related subjects and practices ',
and this definition, whilst embracing the literal translation, also
includes many other aspects of solids in contact. In fact the word
was specifically created to bring together the interests of chemists,
engineers, materials scientists, physicists and others in the many
aspects of such subjects as contact mechanics, friction, lubrication,
wear, etc. This wide-ranging interest in the subject underlines its
most significant academic feature; it is an interdisciplinary subject.
Although the name is a new one, the subject deals with some of the
oldest problems with which mankind has been concerned. The
wheel, an obvious tribological device, was invented to reduce the
resistance to translatory motion and its invention certainly predates
recorded history. In almost every aspect of our daily lives we are
meeting some manifestation of tribology and not only in the context
of the machinery which is such an essential part of an industrial
society.

1.2. The occurrence of tribology


Let us consider a typical day in our lives. As we lie in bed the
different feel of cotton or nylon sheets is a tribological phenomenon,
whilst the early morning stretching of our limbs often provides an
acute consciousness that our ' tribological joints ' are stiff. Indeed
we owe much to the tribologist for the modern techniques which
provide relief from such diseases of the joints as arthritis.
Gripping something, perhaps striking a match, and skidding on
the soap are yet more manifestations of our contact with tribology.
Cleaning our teeth is a controlled wear process, while the lubrication
afforded by soap and the edge on our razor blades show how tribology,
properly applied, contributes to our comfort. As we proceed
through the day gripping, holding, sliding, brushing, all demonstrates
the impact of tribology, whilst every piece of machinery in the home,
1
the office or at work, is teeming with tribological contacts. The car
we ride in has several thousand different tribological contacts and
the contact of tyre and road and the contact of brakes hold our very
life in their proper functioning.
Even our leisure pursuits are each beset with tribological pheno-
mena. Friction is vital to the rock climber and generally a scourge
to the skier or skater, although even here controlled friction is rather
desirable, particularly if one wishes to stop. \Ve use friction
intelligently in all ball games to induce spin and flight. It is interes-
ting to speculate how mankind might have evolved if walking had not
been a practical mode of transit on the earth's surface. Walking
either with low friction, as on ice, or with large friction, as in thick
mud, soon makes us recognize how our bodies have adapted to actual
friction conditions of the earth's surface.
In the evolutionary process the ability of animals to translate
themselves over the earth's surface must have been an important
criterion. We shall therefore examine the ratio of the translational
resistance to the weight of the animal, i.e. drag /weight ratio, as
evolution proceeded. We start with the primeval reptiles sliding
from the oceans and the development of crawlers, apes and man, and
find that their drag /weight ratio has slowly reduced against the time
scale of history, fig. 1.1. Interpreting the logarithmic scales of this

•Ml

r
/El

10-1
I /
0
,/E2

..... ---....-E3
o' E 1 First reptiles
-0
E~------ E4 E2 Crawlers

~,£
e
Cl
.Q'
;
10- 2
I
•M3 E3 Apes
E4 Man
•M4 E5 Athlete
,< 10- 3 ./ Ml Sliding
M2 Lubricant
• - - - - M5
M6
M3 Early wheel
M4 Spoked wheel
10-4
M5 Railway
M6 Modern railway

Years before present

Fig. 1.1. The reduction in the drag/weight ratio over the timescale of the
Earth's history.
2
figure we see a thirty-fold decrease in the drag /weight ratio during the
past thousand million years-the E line-arising from the evolutionary
development of mankind.
About 10 000 years ago modern man could move large bodies
only by sliding them around. Since then, by the development of
primitive lubricants based on animal fat and later a whole series of
wheel systems, he has vastly increased his ability to translate objects.
Thus in fig. 1.1 we note a reduction of some thousand-fold in the
drag /weight ratio during this 10 000 year period-the M line.

1.3. Historical
The invention of the wheel probably took place at least 6000 years
ago, evidence of its use being found in the earliest known historical
records. It is interesting to note that the highly developed Inca
civilization did not discover or exploit the principle of the wheel.
Other tribological applications stemming from the earliest days of
man's history were the use of friction to create fire, the development
of drills fitted with bearings made from bones and antlers, and the
use of elementary grinding wheels for grinding cereals and the like.

Fig. 1.2 .. The bow drill and the grinding stones used by man for the last
4000 years.
3
Fig. 1.2 shows a simple bow drill, a type which is still used in India,
and a very simple cereal-grinding stone. Evidence of the use of
lubricants has also been recorded by recent archaeological excavations.
A chariot taken from an Egyptian tomb still had wheel bearings
containing some animal fat lubricant contaminated with debris in the
form of quartz sand, compounds of aluminium, iron and lime,
presumably picked up during its service life.
The monumental task of building the Pyramids involved the
Egyptians in many tribological developments. Surviving illustrations
in the form of bas-reliefs show the use of rollers and sledges for the
transportation of large stone blocks and statues. Fig. 1.3 shows one
example of such a process, where 172 slaves are seen to be dragging
a large statue of mass about 60 Mg and weight about 6 x 10 5 N along a
wooden track, the statue being supported on a wooden sled. Closer
examination of this illustration shows one man standing at the front
of the sled pouring some form of lubricant into the path of the sled;
surely one of our earliest tribologists. It is interesting to estimate the
coefficient of friction which applied in this situation. If we consider
that each slave pulls with a force of about 800 N, the coefficient of
friction is seen to be
= no. of slaves x force per slave= 172 x 800 = _ .
0 23
µ, weight of statue 6 x 10 5
This is just about the value we would expect for a poorly lubricated
wooden slide and it leads to an interesting conclusion. This picture
must be a true record of what happened, although the artist of this
time clearly had no ability to produce a drawing according to perspec-
tive. VVe see that in a sense this picture is as faithful a contemporary
record as the photographs used by today's media.
Many other examples of tribological practice in the ancient world
could be cited but the science of tribology, as opposed to its practice,
stems from the 15th century. Leonardo da Vinci (1452-1519)
clearly recognized that the friction force is proportional to the load and
independent of the nominal area of contact. He also made the very
perspicacious statement about rolling motion that ' This kind of
friction is caused not by rubbing, but by contact in what might be
described as progress by infinitely small steps '.
The first two laws of friction are always credited to Amontons,
who stated them in 1699 and was undoubtedly unaware of Leonardo
da Vinci's work, and in 1780 the third law of friction was provided
by Coulomb. He stated that friction between two surfaces is indepen-
dent of their relative velocity, although we now know that this law is
less universally true than the first two statements, viz. that the friction
force is proportional to load and that the friction force is independent
of the nominal area of contact. About this time many other tribo-
4
V1

Fig. 1.3. Transporting an Egyptian colossus ca 1880 Il.C. From a drawing in a grotto at El Bersheh.
(Reproduced from Sir A.H. Layard, Discoveries in Nineveh and Babylon.)
logical developments occurred because of the stimulus of growing
industrialization, and this progress has continued up to the present.
The use of lubricants stems from very early times, but Newton
was the first person to study the laws of viscous flow. Theoretical
understanding of lubrication in bearings came towards the end of the
last century, and the work of Beauchamp Tower (1899), of Osborne
Reynolds (1886), and of Stokes and Petroff about the same time, is
particularly noteworthy. Since then there has been an enormous
amount of work in both the application and understanding of lubri-
cated systems.

Fig. 1.4. Detail of a fragment of the Lake Nemi Bearing (reproduced from
Vrelli Le Novi di Nemi, Libreria Delio Statto, Rome).

The earliest recorded use of what we would now call a rolling-


contact bearing is shown by relics found in Lake Nimi near Rome in
1928. These relics date from about 40 A.D. and were some form of
rolling-contact thrust bearing probably used to allow the easy rotation
of large statues during the sculptor's work (see fig. 1.4). Bearings
of this type appear to have been introduced into machinery about
1700 and are now extensively used in machines of all types. They
are surprisingly cheap and are available in a wide range of sizes and
types, and they owe their popularity to these factors.
The earliest lubricants were undoubtedly derived from animal fat,
but oil rapidly came to be recognized as a plentiful and more efficient
replacement. Chemists have by now developed an extensive range
of lubricants for industrial purposes. These meet the designer's
demand for devices to operate at higher loads and speeds, and often
in very hostile environmental conditions such as high temperature,
high humidity, high vacuum and chemical vapours of all types. In
many cases the environmental conditions have required completely
new approaches, as for example in the problems associated with
nuclear reactor machinery and with space vehicles. And over the
6
past 25 years we find that the life of a car engine has increased by
some 300 per cent, while the engine itself has become more efficient,
achievements which owe more to the development of good tribological
practice than to any other single cause.

1.4. Tribological solutions


Although the remainder of this book will elaborate the details, we
can, even at this stage, catalogue the principal tools available for the
solution of tribological problems. Before we do this, we must
identify the nature of the problem. We shall almost invariably wish
to minimize wear, since it represents both loss of efficiency and money.
In the case of friction we find an interesting paradox. In many cases
we shall wish to minimize friction to prevent the expenditure of energy
in overcoming it. However, in many situations we rely on friction
to perform specific functions. Imagine the difficulties of driving or
braking a car without using friction! And even the humble nut and
bolt only stays tightly locked because of friction between the nut and
the bolt.
In what follows we shall confine ourselves to those problems where
it is implicit that we are trying to minimize both friction and wear.
At this point a word of warning is necessary. Although friction and
wear occur at the same place, there is no simple relation between the
two effects. Thus high friction does not necessarily imply high wear.
This is illustrated by the results of table 1.1. We note two interesting
facts in this table. Firstly, the lowest value ofµ, is not accompanied
by the lowest wear rate, and, secondly, while all the values ofµ are
covered by a spread of about a factor of three, there is a spread of
78 000 times between the lowest and the highest wear rate values.
Let us now consider the problem of controlling both the wear and
the friction at a rubbing interface carrying a load W (that is, with a
force of W acting orthogonally to the two surfaces).

Table 1.1. Friction and wear results for various material combinations.

Materials Wear rate


cm 3 /cm x 10-12

Mild steel on mild steel 0·62 157 000


60 /40 Brass on tool steel 0·24 24 000
PTFE on tool steel 0·18 2000
Stellite on tool steel 0·60 320
Stainless steel on tool steel 0·53 270
Polythene on tool steel 0·65 30
Tungsten carbide on itself 0·35 2

B 7
(a) Choice of materials-fig. 1.5 a
Bearing in mind the results quoted in table 1.1 we might hrst
explore the possibility of meeting our requirements by the correct
choice of our materials. Thus we note in table 1.1 that PTFE
(polytetrafluoroethylene) gives low values ofµ, and modest wear when
rubbing on hardened tool steel. For this reason such materials are
often used in so-called ' dry bearings ' and various designs which
embody them are now commercially available. Unfortunately these
materials suffer from two disadvantages. First, they are relatively
soft and we must therefore limit the size of load applied to such
contacts. Second, they are subject to failure as the temperature
rises, and since all friction generates heat, we cannot use them at
high speeds where temperature rises may become too large (see
Chapters 2 and 4). These disadvantages may in part be overcome by
using metals rather than plastics, and recent work suggests that some
forms of crystal structure in metals are particularly useful in meeting
our requirements. Alloys based on metals such as cobalt appear to
offer possible solutions in this context.

(b) Use of surface films-fig. 1.5 b


A surface film adhering to the solid surfaces offers a barrier to the
contact of the solids. Sliding then occurs by shearing through the
film, which reduces both the friction and the wear at the interface.
We may identify three types of surface film which are useful, though
such films never completely eliminate contact between the solid
surfaces, since they are extremely thin and readily disrupted.
One form of film is produced by chemical reaction with the metals
in contact. Perhaps the best known examples are the E.P. (Extreme
Pressure) additives in which such compounds as chloride or sulphide
surface films are formed on the metal surfaces. Such films are
desirable because of their thermal stability at the high temperatures
that are reached in the rubbing process. They can be disrupted by
the rubbing, and the thicker the film the greater is its resistance to
disruption. Fig. 1.6 shows the behaviour of a chemical film subjected
to increasing numbers of rubbing cycles. Even a monomolecular
layer gives low values of µ, but after one traverse such a film is
destroyed. Increasing the thickness of the film gives it an increased
life.
A second type of surface film uses materials such as graphite or
molybdenum disulphide. These so-called lamellar solids are useful
because of their layer structure, which might be likened to a pack of
playing cards. With such a structure we can appreciate that they
can carry load in one direction orthogonal to the layers but give low
friction as one layer slides freely over the next (see Chapter 4 ).
8
Finally, some situations are best met by surface films in the form
of soft metals such as lead. Here again we rely on the relatively low
shear strength of the soft metal to reduce the friction force whilst
the underlying strong base metal still gives high load-carrying capacity
in the orthogonal direction (see Chapter 3).

CJ~
__ ==---_-~Pressure P
1////,//1///,//,////7/1///;;////,////////

(a) (d)

(b)
1-- W///~!Wi ~ (e)
Elastomer

w
iw
-- D::::,ga,t;o
1111111111111111111111111
1;,ld

(c) (f)

Figs. 1.5. Methods of solution of tribological problems. (a) Choice of


materials, (b) Use of surface films, (c) Use of rolling elements, (d) Use
of pressurized fluid films, (e) Use of elastomers, (j) Use of magnetic
force fields.

(c) Rolling contacts-fig. 1.5 c


One solution to our problem is to separate the two bodies by rolling
elements. Depending on the particular design requirement these
may be balls, cylinders or any other geometry capable of sustained
rolling, i.e. they must have circular geometry in the plane of motion.
At first sight we may appear to have thereby achieved a complete
elimination of friction and wear. Unfortunately this is not the case
9
since some sliding always occurs in such systems (see Chapter 3).
Nevertheless this solution is widely employed and it gives rise to
much less friction and wear as compared to (a) and (b) above.

(d) Pressurized lubricant films-fig. 1.5 d


Another solution to our problem is complete separation of the
surfaces by a relatively thick film of lubricant (typically some 100 µm
thick rather than the surface films of 1 µm or so discussed in (b)
above). If we applied a load to such a system the film would dis-
appear unless the fluid were under a pressure such that (pressure x
area) equals the applied load. One way of proceeding is to supply
the fluid from a pressurized source; this is the principle of the hydro-
static or externally pressurized bearing (Chapter 5).
A second and more ingenious solution is to generate the pressure
by using the relative sliding motion involved. If we arrange the
moving surfaces to be slightly convergent rather than parallel the
fluid is drawn into a channel of reducing cross-sectional area. When
this happens to a viscous fluid a pressure is developed hydro-
dynamically and this is the principle of the hydrodynamic or self-
acting lubricated bearing (see Chapter 6).

(e) Elastic solutions-fig. 1.5 e


In some situations the extent of sliding between components is
restricted, as in small-amplitude reciprocating systems. The suspen-
sion system of a car is an example. In these cases one attractive
solution is to bond an elastomer to the two surfaces so that the motion
is accommodated by the deformation of this elastomer. The load-
carrying capacity is reduced by such a system, particularly if the
thickness of the elastomer is increased so as to allow greater amplitude
of oscillation. Alternatively the elastomers can be replaced by a
system of flexible metallic strips, which operate on the same principle.

(f) Miscellaneous solutions-e.g. fig. 1.5 f


The essential problem is to try to separate the solid surfaces while
at the same time retaining their capacity to carry normal load and to
allow transverse motion. Any technique which does these things is
therefore worthy of consideration, and the use of electrostatic or
magnetic fields is clearly attractive. In particular, magnetic systems
are receiving considerable attention and the common electricity
supply meter uses a magnetic bearing.
It is often both possible and desirable to combine some of the above
solutions in a particular case. Thus, although hydrodynamic
principles are widely employed (as in many of the bearings in the car
engine) during the starting and stopping process the hydrodynamic
effect is not operative since the sliding speeds are too small. To
10
avoid too much damage during these periods we therefore use
additives to provide the surface films discussed in (b) above. Like-
wise some slip and wear does occur in rolling-contact bearings so
that such bearings are provided with lubricants to mitigate the effects
of wear.

t
µ

Many layers

Number of traverses ~

Fig. 1.6. The effect of increasing the number of mono-molecular layers of


lubricant on the life of surface films.

l.5. Economic considerations


Unwanted friction in any machine leads to an expenditure of energy
(which costs money) to do work against it, while wear means an
expenditure of capital to replace worn-out machines. Poor tribo-
logical practices can be expensive in terms of money and, perhaps
more important, in terms of scarce resources. So in a very real
sense tribology is a conservation science. In an industrialized society
we spend perhaps 30 per cent of our Gross National Product in
replacing the ravages of wear and corrosion. In this we include the
wear of machinery, shoes, clothes, furniture, pots and pans, and
innumerable other situations (see section 1.2).
In industry, as well as the primary costs which have already been
discussed, there are inevitable secondary costs due to tribological
failure. Thus a production line may stop due to a bearing failure
in a conveyor. The cost of the bearing replacement represents only
a small part of the total cost of such a breakdown in a highly organized
production line. In 1966* a study of the cost to this country arising
from poor tribological practice suggested that some £500 million per
annum could be saved by improved application of existing knowledge
(fig. 1.7). It will be noted that the secondary effects are the most
significant.
* H.M.S.O. Lubrication (Tribology) Education and Research, 1966.
11
Although in terms of conservation wear is always undesirable,
there are two categories in which it is encouraged rather than avoided.
The first is in the so-called ' running-in ' process, when we allow
initial wear to occur so that mating surfaces may adjust to each other
to provide a smooth running combination. Even here the wear is
not really desirable, we simply use it because of our present inability
to make properly mated components in our factories. In a sense
such running-in is the final machining process leading to proper
functioning. In the case of cars this requirement has been reduced
as manufacturing expertise has increased during the past 50 years.

E
:::,
C
C
0
515
505
495
473
- Manpower savings
Lubricant savings
Investment savings
Less friction losses
'--
<l) 445
Cl.
V)
C
.'2 - - Longer machine I ife
.E
""
C 34,5
>-
a,
0
0 - - Fewer breakdowns
..0
E
'--
l':'
a:; 230
..0
E
.....2
V)
a, Less maintenance
s>
0
and replacement
V)

~
..0
·;,;
V)
0
Q

0
Fig. 1.7. The saYings achiernb]e in Britain each year due to better tribological
practice.

The second category is in ' planned obsolescence '. Herc wear


is used to ensure a finite life for machines, and under our present
organization of society provides a continuing demand for products.
This approach does of course ensure periodic renewal of industrial
equipment and therefore a more rapid adoption of new technological
developments which are incorporated in the replacement machines.
In a sense:we find ourselves on the horns of a dilemma, as we may
12
see by considering the case of the motor car. The question we need
to answer is what is the ideal lifetime of a car if we are both to conserve
raw materials whilst at the same time being able to benefit from
technological advances in the design of cars, particularly with regard
to safety? The answer to this question is clearly a matter of individual
choice, but it obviously offers a useful debating point.

1.6. Conclusion
In view of the industrial importance of tribology, this aspect of the
subject usually receives the most consideration. Nonetheless, the
contribution of tribology to other aspects of our life can be very
profound, for example, in our understanding of human joints and,
in particular, to the alleviation of suffering from diseases such as
arthritis.
I hope that the reader has by now been stimulated to find out more
about the details of the various topics which have been mentioned.
It has so far been tacitly assumed that surfaces are simply planes
separating solids from their environments. This is not the case.
All surfaces are rough rather than smooth and (although this roughness
is on a microscopic scale) it adds to the complexity and interest of the
subject of tribology.

1. 7. Project suggestions
1.7.1. Historical projects
Throughout history mankind has been concerned with movmg
objects relative to each other. One can therefore devise many
historical projects in which developments in tribology may be related
to the development of man towards an industrialized society. A few
examples are:
(1) Consider tribology in the context of the earliest civilizations
leading to the achievements of the Egyptian civilization.
Examine the records of the time and establish the scientific
validity of such records, e.g. in bas-reliefs showing slaves
moving large masses; do the number of slaves and the estimated
weight of the body being moved give reasonable values of the
coefficient of friction for the tribological method being
employed?
(2) It is reported that the Inca civilization did not discover the
principles of the wheel. How did they overcome their tribo-
logical problems and to what extent were they inhibited by not
having wheels? How is the lack of wheeled transport reflected
in their civil engineering and architecture?
13
(3) Rolling motion is a well established method of solving tribo-
logical problems and has led to the development of a wide range
of ball and roller bearings. Study the history and development
of this type of bearing and evaluate its importance to society at
various stages of its development.

(4) Transportation has always been a primary requirement of


mankind and represents a very important area of tribological
development. Consider the tribological solutions to this
particular problem over either the full range from pre-history
to the present day or any particular historical period.
(5) What were the methods used to prevent wear up to the end of
the Middle Ages?
(6) Find the oldest piece of industrial machinery in the area and
consider how its designer overcame tribological problems with
the knowledge available to him at the time. Havv have advances
in tribology been used to improve machinery?
These and many other similar historical projects can be readily
created and provide an important area of project activity. Many
ideas and much useful information can be obtained from the books
Engineering Heritage, Volumes 1 and 2, published by the Institution
of Mechanical Engineers.

1. 7.2. Economic and social problems


From shaving in the morning, throughout our work and leisure
until we clean our teeth at night, we are involved with tribology.
Some of these tribological effects are of primary economic significance
to ourselves, our schools, our city and our country, while others have
major social implications. It is, therefore, hardly surprising that
many exciting tribological projects can be devised in this area, for
example:
(1) What are the important contributions which tribology can
make in athletic achievement? Some examples may be the
resistance of skis and the effect of various surface treatments,
and the value of brush spikes on tartan tracks.

(2) What is the economic significance to your education authority


of different types of floor covering, bearing in mind such factors
as wear resistance, skid resistance, safety and hygiene? Some
elementary test programmes should be easy to arrange in a
school environment to ensure statistically accurate results.
14
(3) Bearing in mind the information contained in the 1966 Report,
Lubrication (Tribology) Education and Research, can you arrange
to study the economic effects of tribology on any local industrial
plants? Consider the tribological shortcomings and try to
evaluate the total economic cost in terms of down-time, etc.
which they create.
(4) l\Iany aspects of shoe design are of tribological interest besides
the obvious ones of the friction of shoe soles, their wear
resistance and their resistance to lubrication by water. Ladies'
shoes in particular involve a balance of practicality against
appearance, fashion and economic factors. Leather was, for
hundreds of years, the material of which shoes were made-how
is this practice changing and are any tribological compromises
involved? The stiletto heel was a very effective indentation
hardness tester for soft floors--why was this?-and can such a
design be justified on any scientific or engineering grounds?

15
CHAPTER 2
the nature of surfaces and their contact

2.1. Introduction
TRADITIONALLY we might think of the surface of a solid as the
geometrical boundary between the solid and its environment. For
our purposes such a definition is too limited and we must probe a
good deal into the body of the solid if we are to understand how its
surface behaves. We shall therefore consider our surfaces in depth
including the nature of the surface layers and the sub-surface behaviour
of the material. We shall be interested in the physical properties of
the surface layers as well as in the geometry of the actual surface
profile. Not surprisingly these characteristics depend on the bulk
properties of the solid material, the methods by which the surface
has been produced, and the nature of the environment around the
surface. Thus the optical behaviour of a metal surface depends on
the smoothness of the surface, the character of the surface layers and
the extent to which the air has reacted with the metal to form an
oxide film.
By virtue of their high strength metals are extensively used in
tribological applications. This book will therefore chiefly be con-
cerned with metallic solids, although many of the arguments apply
equally well to other materials such as plastics.

2.2. The nature of surf aces


The geometry of the Earth's surface is characterized by roughness,
sometimes severe as amongst mountains, but more generally undula-
ting with slopes of only a few degrees. The surface layers consist of
deformed rock strata, surmounted by a finer texture such as soil and
sand. The outermost surface is covered by a layer of vegetation
arising from the interaction of the surface material with the environ-
ment with the help of solar energy.
The examination of a metallic surface in depth reveals the same
kinds of feature, although the scale of these features is some hundred
million times smaller. A typical metallic surface might appear
somewhat like that shown in fig. 2.1. On top of the normal crystalline
structure lies a layer of deformed material created by the processes
used in the manufacture of the surface. This layer is often overlaid
16
by a microcrystalline layer which is also produced in the manufac-
turing process. In such processes the outermost molecular layers
are melted and smeared over the underlying material. The sudden
cooling of this molten layer produces a structure of very fine crystals
which is harder than the underlying material. This hard thin
layer can be very important in producing high resistance of the
surface to wear. There is also an outermost layer produced by
chemical reaction of the surface with its environment. \Vith steels
in air this will be an oxide layer (rust) and it is of considerable signifi-
cance in tribology. In a sense it acts as a barrier between metallic
surfaces that are apparently in contact, and so helps to reduce the
friction between them Finally, the surface will usually be covered
with dust, wear debris and possibly lubricant, and as fig. 2.1 shows,
such particles are of a similar size to the intrinsic roughnesses of the
surface boundary.

Oxide
layer

~Quenched
layer

Base material

Fig. 2.1. A schematic diagram showing the structure of metallic surfaces.

We might pause to consider the meaning of the phrase ' a clean


surface '. To a chemist this almost certainly means a surface from
which one has removed all surface contamination such as dirt, debris,
traces of lubricant and even the chemical films such as oxides. Such
surfaces, which have to be kept under ultra-high vacuum conditions
to avoid their re-contamination, readily adhere to each other. In
everyday terms a surface is often said to be clean after it has been
washed with solvents to remove grease and dirt, but it will still retain
its oxide films and in such cases the adhesion of one surface to another
is negligible. The difference between these two definitions of' clean '
is very important in tribology.
17
2.3. The properties of surfaces
As far as physical properties are concerned we shall be interested
in such features as the crystallographic structure of the surface
layers, the chemical reactivity which leads to formation of surface
films and the general mechanical properties of the surface material.
\Ve are therefore intrinsically interested in the stress (load intensity,
that is, load per unit area) and strain (deformation characteristics) that
occur in the surface and in its ability to sustain such stresses without
failure. This information can be determined if we know the elastic
constants of the material, its yield strength which defines the stress
at which elastic behaviour changes to plastic deformation, and the
hardness of the surface material which is closely related to its yield
strength. These aspects of surface behaviour are dealt with later
in this chapter.

Waviness

Roughness
Fig. 2.2. The constituent geometric components of a solid surface.

The geometrical characteristics fall into three categories (fig. 2.2):


( 1) Errors of form, in which the surface deviates from the desired
shape due to errors inherent in the manufacturing process,
for example the tapering of an ostensibly cylindrical bar.
(2) Waviness or macro-texture, which takes the form of relatively
long wavelength variations in the surface profile and is often
associated with the unwanted vibrations which always occur
in machine tool systems.
(3) Roughness or micro-texture, which is the small scale roughness
of the surface associated with the actual cutting and/or polishing
process during its production, e.g. by the action of the grits in
abrasive processes. From the standpoint of tribology this is
the geometrical variation which is generally of greatest interest
and we shall return to it later.
18
2.4. Methods of studying surf aces
We are fortunate in having a very wide range of instruments for
the study of surface properties. No detailed descriptions are possible
within the limits of this text, but it will be helpful to mention some of
the most commonly used procedures.
(a) Optical microscopy
This is the best known method for observing surfaces, and with
sufficient magnification it reveals many of the finer features.
Unfortunately it suffers from a number of drawbacks. The use
of visible light restricts the resolution of the instrument in so
far as light is unable to discriminate features which are smaller
than 0·25 µ,m, which is about half the wavelength of green light.
Furthermore, even at the highest resolution one obtains a
picture of only a very small part of the surface-and one would
hardly expect to learn much about the general character of a
mountain range by observations limited to only one or two of
the mountains. The optical microscope also lacks depth of
focus so that it tends to emphasize the spacing of features
rather than their actual height, which is of greater interest to
the tribologist. It is also difficult to obtain quantitative values
of the size of surface features using optical microscopy.
(b) lnterf erometry
Light from a common monochromatic source is reflected by a
beam-splitting device from the observed surface and from a
standard plane reference surface, fig. 2.3. The combination
of these two beams gives rise to a pattern of interference fringes
(rather like Newton's rings), which are in effect contour lines
which indicate the profile of the surface. Such instruments
give a good representation of the surface texture and are
self-calibrating since we know that the vertical scale is such
that the distance between adjacent fringes represents one half
the wavelength of the light used, say about 0·25 µ,m. Un-
fortunately this technique still only allows one to view a very
small, and perhaps unrepresentative, sample of the surface.
These optical techniques are described in detail in Introduction
to }Ylodern Microscopy, by H. N. Southworth (Wykeham Science
Series No. 34).
(c) Electron microscopy
The electron microscope uses a beam of monoenergetic
electrons to produce an image on a fluorescent screen, and gives
a much finer resolution of detail than the optical microscope,
about 10 A (1 nm), which is about 250 times finer than that of
the best optical microscope. Electron microscopes may be
used by reflecting the electron beam from the surface, when by
19
a suitable choice of angle of incidence the surface features are
made to cast shadows (fig. 2.4) so that their size may be
calculated. The electron microscope may also be used to
study the structure of material by using a transmitted beam of
electrons but this technique can only be used for relatively
thin films of the material.

A
I
I
I
I

~ I I
I I
1 / Eyepiece
I I
I I
I/
11

(al Micro-interferogrom
1,ii
I~
+ B .
1 1 eom splitter ~
Light source ---~- -nt-}L
*========-=----li.:=hl- ----VJ---u-
Reference
surface

! \
~~ Identical
I
~ objectives
I I
I I

~
'M.~..:..:Z~~'<...«:,,.."-7<;!

Test surface
Fig. 2.3. A schematic layout of a surface interferometer.

Like all the preceding methods the electron microscope only


examines a very small part of the surface, and may therefore
misrepresent its general character. This deficiency is over-
come in the scanning electron microscope which produces a
composite picture of a larger surface area by allowing the beam
to scan in a controlled sequence. Such systems also include
stereoscopic appreciation of the surface and are therefore an
exceedingly useful method of surface examination. Fig. 2.5
is an example of a scanning electron micrograph.
More information on these and related methods can be
found in Electron Microscopy and Analysis, by P. J. Goodhew
(Wykeham Science Series No. 33).
20
Fig. 2.4. A shadow electron micrograph of a metal surface 2500 x .

(d) Taper sectioning


The only method for the direct observation of the surface
geometry is by the relatively simple process of taper sectioning,
fig. 2.6. The surface to be studied is plated with another
metal, e.g. nickel, which locks the surface roughness so that the
composite block may be carefully machined and polished at
some taper angle 0. Such sectioning reveals an interface which
follows the pattern of the original surface profile and enhances
the vertical magnification by a factor cot 0. A typical taper
section is shown in fig. 2.6 b.

(e) Profilometry
The most usual method for the study of surface geometry
is profilometry. In the profilometer a very fine diamond
stylus (tip radius 2 µm or less) is drawn over the surface
21
Fig. 2.5. · A scanning electron micrograph of the surface shown in fig. 2.4,
2500 X.

irregularities (fig. 2.7 a). The vertical movement of the stylus


as it traverses the profile is measured and amplified, usually
electronically, so that the recorded output provides a picture
of the actual surface. One of the best known instruments of
this kind is Talysurf (Rank Taylor Hobson), which produces
records of the type shown in fig. 2.7 b.
One of the most attractive features of this instrument is its
flexibility in controlling the horizontal and vertical magnifica-
tion independently. The horizontal magnification is controlled
by the speed of traversing and the speed of the paper on which
the profile record is produced. This magnification is typically
100 x. The vertical magnification is controlled electronically
and may be varied from 500 x to 10 000 x according to the
precision required.
22
Plated layer
Surface profile

Specimen

Observed heightof asperities

Vect;calhe;ght~
of aspent1es
e
Magnification =
cot 8
e = 5° this gives 11 x magnification
(a)

(b)

Fig. 2.6. The method of taper sectioning and a typical example of the method.
C 23
Stylus

Traversing
motor

Amplifier

Surface
trace

(al

5000

L100

(b)

Fig. 2.7. The basis of a profilometer and a typical profilometer trace.

The difference in the vertical and horizontal magnification,


although useful in giving greater emphasis to the height
characteristics of the surface, does mean that the resulting
record is distorted. Thus in fig. 2.8 a typical Talysurf record
is shown together with a record from the same surface when
both the vertical and horizontal magnifications are the same.
The latter, which is a true picture of the profile, shows that the
surface asperities are undulations rather than the sharp peaks
shown in the distorted profile. The appreciation of the actual
gentle slopes occurring on machined surfaces is of paramount
importance in tribology. Fig. 2.8 also shows the size of the
measuring stylus to the same scale as the record it produces so
that it is seen that some errors must occur due to the finite
size of this stylus.
24
(f) X-ray diffraction
By studying the diffraction patterns which are obtained when
X-rays pass through the surface material at low angles of
incidence one is able to ascertain the crystal structure of surfaces.
In particular it is important to know the orientation of the slip
planes of the material in relation to the geometric contact
surface.

Stylus tip radius

~ I
/
', ___ .,,.,,.,,
\
1µmL ........ 1µml__,
lµm IOOµm

~
Actual profile Distorted record

Fig. 2.8. The actual profile of a surface compared with the surface record
obtained from a profilometer.

(g) Electron micro-probe analyser


This instrument scans a surface with a 1 µm cross-section
electron beam and records the chemical constituents of the
surface. When a surface is in contact with another of a
different material this instrument therefore enables one to
determine the degree of transfer of material from one surface
to the other. A typical result is illustrated in fig. 2.9.

(h) Microhardness tester


The hardness of a surface is defined as its resistance to penetra-
tion, usually of a diamond indenter, and is expressed in terms
of the load applied divided by the area of the indentation. This
instrument, which is associated with a high resolution optical
microscope, enables one to determine the hardness of the
material in the outermost layers of a solid. It is therefore of
great importance in studying a mechanical property of the
surface which is very important in tribology, although its
resolution is inadequate for really detailed study of the hardness
in the very thin surface layers in which the tribologist is most
interested.

2.5. Some geometrical properties of surfaces


Consider the surface profile obtained by a profilometer as shown in
fig. 2.10. It is possible (by a process called Fourier analysis) to
25
C d

Fig. 2.9. A typical electron microprobe record. (a) Electron image, (b) S,
Ka: X-ray image, (c) Mn, Ka: X-ray image, (d) Pb, Ka: X-ray image.

---tS~-JS
Horizontal intercepts

----- - --------

A-LLLr~+ - l - A

Sample
intervals
-------------- - --- ---
Port of profile considered Distribution Bearing area
of ordinates curve

Fig. 2.10. The conversion of a surface profile to an ordinate distribution and


a bearing area curve.

26
break down this complicated shape into a long series of sinusoidal
waveforms which when superposed would produce the same shape.
Unfortunately this leads to an expression with a large number of
coefficients and is therefore of little practical use in defining this
profile reasonably simply. For this reason attempts are made to
represent it by utilizing its statistical properties, as outlined below.
Assume some datum line such as AA in fig. 2.10 and measure the
vertical ordinates of the profile from this datum along a representative
length of the profile, which is assumed to contain all the essential
features of the whole surface, remembering that the profile is only a
sample taken over a small length of the total surface. From such
measurements we can state how many heights of a given ordinate
value z occur at any given horizontal level in the profile. We can then
construct a distribution curve for the profile height over the given
length of the profile.
The horizontal line passing through the centre of area of the
distribution curve is defined as the Centre Line of the profile. The
areas generated by the surface profile above and below this centre
line are equal. The departure of the profile from this centre line
(i.e. from a smooth surface) may be identified by the parameters r.m.s.
(root mean square) and c.l.a. (centre line average) which are defined
as
1 n
r.m.s. = [ - _L (z;) 2
]l/2
n,=1
1 n
c.l.a. = -L jz; I•
ni=l

For most engineering surfaces the c.l.a. and r.m.s. values are very
similar and the r.m.s. value is simply the standard deviation a of
the ordinate distribution curve shown in fig. 2.10. Both these para-
meters are readily obtained by applying the voltage signal from the
profilometer to a meter; directly for the r.m.s. value and after
full-wave rectification for the c.l.a. value. For this reason these
parameters are extensively used in engineering practice. Their
limitations are perhaps best illustrated by the idealized profiles shown
in fig. 2.11, each of which although markedly different would give the
same value of c.l.a. For this reason the tribologist seeks surface
parameters which are rather more informative.
Consider the ordinate distribution curve shown in fig. 2.10. If we
adjust the scale of this diagram so that the area enclosed by the distri-
bution curve is unity we have converted the curve into a probability
density curve, i.e. the size of the horizontal at any given value of z may
be interpreted as the probability of an ordinate of this value of z
occurring. Before leaving fig. 2.10 one other distribution is of
27
interest. By summing the horizontal intercepts produced by the
profile at any given value of z we effectively obtain the area of solid at
this level. Thus repeating this procedure for all values of z produces
a curve which is known as the Bearing Area Curve. Mathematically
it is in fact the cumulative distribution curve of the ordinates, i.e. the
integral giving the area under the ordinate distribution curve, but its
physical interpretation as the amount of area at each depth level is
more informative to the tribologist. The use of a line tracing to
predict area is of course only valid where the surface characteristics
are similar for all orientations of such a trace.

Fig. 2.11. Geometric profiles haYing the same c.l.a. value.

Distribution curves and hence probability density curves may be


produced for many other features of the surface, some of the more
important ones being described below.
Any surface profile clearly consists of a system of peaks and·valleys
and fig. 2.12 a shows how one may represent the occurrence of peaks
and valleys at various values of z by their distribution curves. In an
analogous way one may compute the slope of the profile at all values
of z and again plot distribution curves for the slopes (fig. 2.12 b).
Having obtained the distribution curves for those features which
are of most interest we can deal with them mathematically by repre-
sentations of their shape. Fortuitously it is often found that these
distributions follow a Gaussian law and in such cases their shape is
28
completely defined by the standard deviation a (the r.m.s. value),
fig. 2.13, since the expression for the Gaussian distribution is
f(z) = {1 /ay(21r)} exp ( -z2 /2a 2 ).
Although tribology is concerned with many other aspects of the
statistics of surfaces the foregoing will be enough for our purposes.
In due course we shall see the importance of being able to specify
the probability of the occurrence of a peak at a given value of z.

i--
(a) Peak distribution

1---
~

-......
-_____ ::..+- I
✓-,
\------ ---
i
Max.
- ---
height -4
------
\
Valley distribution Part of profile considered Ordinate
distribution

(bl
,,,- Slope_ "-0
dz

profile
Mean ordinate Mean slope
I I

Ordinate distribution Slope distribution

Fig. 2.12. The peak, valley, ordinate and slope distributions of a typical
surface profile.

2.6. The loading of solids


When a solid is subjected to load, stresses are produced in the solid
which increase as the load is increased. These stresses produce
29
Histogram for a
ground surface

Equivalent
>, Gaussian distribution
u
C
(l)
:::,
CT
(l)

it

0·25 0

Vertical distance z from


centre line in µm

Fig. 2.13. A typical ordinate distribution curve for a ground surface.

deformations which are defined by the strains. Unique relationships


exist between the stresses and the corresponding strains. The most
easily recognized of such relationships is the elastic behaviour of
solids where the stress and strain are linearly related, the constant of
proportionality being an elastic constant. Thus a simple tensile
load applied to a bar produces a stress a 1 and a strain E1 where

a1= load
. . axia
= stress m . l d'irect10n
.
cross-sect10na1 area

E1 = change
. . in length = stram
. . . d' .
1n axia 1 irect10n
ongmal 1ength
and
E= a 1 /E1 ,
E being the elastic constant, called the Young Modulus. Although no
stress acts transversely to the axial direction there will nevertheless he
dimensional changes in the transverse direction, for as a bar extends
axially it contracts transversely. The transverse strains Ez are related
to the axial strains E1 by the Poisson ratio v such that

where the negative sign simply means that the transverse deformation
will he in the opposite sense to the axial deformation. Under more
30
complicated conditions of loading and geometry the equations
defining stress and strain are inevitably more complex and their
treatment is beyond the scope of this book.
As the load increases, elastic behaviour is replaced by plastic
behaviour in which the material is permanently deformed; after
removal of the load the material does not return to its original shape.
The stress state at which the transition from elastic behaviour to
plastic behaviour occurs is known as the yield stress and has a definite
value for a given material at a given temperature. In tribology we are
interested in tv,o bodies which are in contact because of the applied
loading, and here the nature of the transition from elastic to plastic
behaviour is of considerable importance.

(al Actual contact ( bl Pressure distribution (cl Stress pattern


(not to scale)
---c--

~~
H t HH

--c--- w Mean pressure w


= cl
I

,//
(//
~
i111Tu
~
't
~zb- 1/ " "~
'~
Mean pressure = .J:!_ 23
2bl / 1 Max
stress
'
r

Mean pressure= 1r 02
w

Fig. 2.14. The contact of a flat, a cylinder and a sphere and the resulting
pressure and stress distributions.

Consider the three problems shown in fig. 2.14 a, assuming that the
materials behave elastically. The nature of the contact pressure
distribution and contact geometry is as shown in fig. 2.14 b. From
elasticity theory we can show that the state of stress in the lower body
would be as shown in fig. 2.14 c for each case, each stress line being
drawn such that the stress state along it has a constant value; the
increasing numbers indicate increasing stress values. For the contact
of the cylinder and the sphere, as the load increases the actual size
of the contact zone increases due to increased flattening of the curved
31
surfaces. The actual relationships for such contact sizes, which are
usually referred to as the Hertzian values, are:
For a cylinder on a plane
b= I 8WR(l -v2) 11;2
TTLE '
pressure distribution

p= !!~ ( -~r12_
l
For a sphere on a plane
a= 13 WR~ --v2) I1/3,
pressure distribution
- 3w
P- -TT2., 1
(. - R2)
2
l /2
'
a" a

where vis the Poisson ratio and Eis the Young modulus, and W, R and
L are as shown in fig. 2.14.
The nature of the stress distribution is of considerable interest in
these three cases. For the uniform pressure contact it is seen that
the stress increases as the surface is approached and is a maximum at
the surface. For the cylindrical and spherical contact the maximum
stress, rather surprisingly, occurs at a small distance below the surface;
see fig. 2.14 c. Thus as the load W increases and the elastic behaviour
changes to plastic behaviour we can see that for both the cylindrical
and spherical contacts plasticity first occurs below the surface rather
than at the surface as in the first example.
This is a very important result since it means that, even though the
load has created a plastic zone, catastrophic failure does not occur
since the plastic material is totally enclosed by elastic material. As
loads are increased further the volume of the sub-surface plastic zone
increases until it spreads to the surface. At this load the cylinder
or ball will begin to penetrate the surface creating a noticeable
permanent indentation.
It is conventional to define the hardness of a material by the ratio
of the load to the surface area of the permanent indentation, i.e. by the
mean contact pressure between say a ball and a plane, where hardness
H = W / TTa 2 • Because of the constraining effects of the elastic material,
plastic indentation does not materialize until the mean contact pressure
has a value of three times the yield stress Y of the material. Thus,

mean contact pressure -


w = 3 Y = H.
2
TTa

32
From this result we may conclude that no significant permanent
indentation of the surface will occur when spherical surfaces are
loaded against a plane until the mean contact pressure is at least three
times the yield strength of the material. This allows us to apply
much greater loads to such contacts than we might otherwise expect,
and is one of the reasons why rolling-contact bearings continue to
operate satisfactorily even when slightly overloaded.

2.7. Contact of rough surfaces


We have already seen that a surface might be considered as an
array of spherical asperities (peaks) whose heights follow some
particular distribution law. Consider such an array of asperities, all
of the same height, in contact with a smooth rigid plane (fig. 2.15).
Any load W applied to such a system will be equally divided between
the asperities each of which has a load W1 such that
W= 'I,W1 .
At small loads the asperities will deform elastically and their behaviour
is defined by the type of argument discussed in section 2.6. The
total real area of contact A is the sum of each of the discrete areas A;
and A; will be given by 1ra;2 where a; is the radius of each circular
contact spot.

Fig. 2.15. The contact of a smooth surface and a rough surface having
asperities of the same height.

Since a; rx JV/1 3

A1 = 1Ta12rx lV1213

1.e. A 1 = K1 W 1 2 13 , where K; = 1r[3R(~- v


2
)]
213
(see section 2.6).

Hence A= nK1 TV/1 3 = K. w2 13, since W; = W/n,


where n is the number of spherical asperities.
For complete plastic behaviour the mean contact pressure of each
spherical contact is the material hardness.
33
Thus

or

Thus

From these two results we note that for elastic deformation the total
real area of contact is proportional to W 2 l3 , whereas for plastic
deformation it is proportional to W.

(a)

W increased

(bl

Fig. 2.16. The contact of a smooth surface and a rough surface having
asperities of varying heights.

For most real surfaces the asperities will be of different heights so


that at the lightest loads contact will only occur at the very highest
ones (fig. 2.16). As the load is increased the discrete areas of these
initial contacts will increase and new contacts will be created at the
lower asperities. From our knowledge of the height distribution of
the asperities we can define the actual number of asperities in contact
at any load, that is at any degree of compression of the surface texture
which has occurred. Although the method of treating this problem
is beyond the scope of this book, the conclusions of such a theory are
essential to our physical understanding of contact. The following
are the most important observations. Since many surfaces have
34
asperity-height distributions which are nearly Gaussian we shall state
some conclusions for such surfaces.

(a) As the load increases and the smooth surface compresses the
rough surface the mean area of a contact spot remains constant,
i.e. the total real area of contact A divided by the number of
contacts is constant, although it will be appreciated that both
A and n increase with increasing load.

(b) The total real area of contact is linearly proportional to the


applied load no matter whether the surface is deforming
elastically or plastically. This is an important result and
markedly different from that obtained with the array of asperi-
ties of the same height.

,...,__ _ _ _ _ _ _ _ _ A

0
--------►,

. 1 A
This material becomes plastic when A= 3 0

Fig. 2.17. The eon tact of a smooth surface and a rough surface at the point
of macroscopic plastic deformation,

(c) At the loads normally used in engineering practice (usually


expressed in terms of nominal pressure, i.e. load divided by the
apparent area of contact), the surface roughness will only be
compressed by about ten per cent. At loads greater than this
value the underlying bulk material reaches the yield stress and
deforms plastically. This may seem surprising unless one
recalls the strength of spherical contacts as discussed in section
2.6. The physical meaning of this argument is shown in
fig. 2.17, and it explains why the normal contact of bodies does
not compress the surface asperities out of existence. The
applied load W produces a mean contact pressure of value 3 Y
at the asperity contacts, if they are assumed to be plastic. The
total real contact area is A, so that at these contacts
W=3YA.
35
At the level of the bulk material the same W would produce a
stress Y when this material becomes plastic so that
W=YAa
where Aa is the apparent, i.e. the total geometric, area of
contact. Clearly if this condition applies, A= }Aa, which for
many engineering surfaces will occur when the asperities have
been compressed only by about 10 per cent of the total depth
of the texture.
Most contacts will, of course, be between bodies both of which
have rough surfaces. Although this complicates the mathematical
solution, the general conclusions stated above are still found to be
true.

w
G)
Heat- source
----u

Fig. 2.18. Body 1 is subjected to a stationary heat source and body 2 to a


moving heat source.

2.8. Thermal effects


The work done against friction in the contact between two solids
is dissipated as heat, that is, converted to internal energy in the first
instance, and therefore results in a temperature rise at the surface
of both bodies. The rate of heat release Q due to friction is given
(in watts) by
Q =µ,WU= Work done against friction
where U is the velocity of sliding. We now wish to consider the
resulting temperature rise in the two bodies. One thing is immediately
apparent. Recognizing that the heat is released at the interface we see
that the heat source does not move relative to the sliding body, but
that it is travelling along the stationary body, fig. 2.18.
A second problem arises in deciding how much of Q goes to body 1
and how much to body 2. The answer to this question clearly
depends on the heat transfer paths through the two bodies, but for
simplicity we may assume that Q is equally divided between the two
bodies.
36
From our understanding of heat transfer we are then able to show
that the temperature rises in the two bodies are given by

el =µ,WU
Sao,
=0·125 (µ,W) ( u)
a ex
from the stationary heat source theory applied to the moving body 1,
and

8=0·159 (µ,W)
2
a
(£)
aexpc
1 2
'

from the moving heat source theory applied to the stationary body 2,
where a defines the area of contact; ex is the thermal conductivity;
pis the density, and c is the specific heat capacity. From these results
it is seen that the moving body temperature is the one that is more
dependent on the sliding velocity.

t Melting point
---
·-·- - · - · - - - - - -·-.:..:·-;;.-·

0 u-
Fig. 2.19. The variation of surface temperature with sliding velocity.

If we substitute typical values of the constants for metals in the


above equation using the apparent area of contact for a we find
that the surface temperatures would be of the order of 100°C. If,
however, we recognize, as in the preceding section, that the real
area of contact is only a very small part of the apparent area of
contact, it is seen that the temperature at the asperity contacts may
be 1000°C or more. This means that sliding may often cause local
melting of the tips of the asperities, an effect of considerable signifi-
cance when considering the friction and wear of such surfaces. This
type of effect is indicated by the surface temperatures measured in
frictional contacts at increasing values of sliding velocity. The curves
tend to a steady value of temperature at the higher speeds which turns
out to be the melting point of the particular materials (fig. 2.19).
37
2.9. Project suggestions
A surf ace profilometer for machined surf aces
Machined surfaces have, on a microscopic scale, the same sort of
geometry that one finds on the Earth's surface in an area such as the
Lake District. The study of such geometry can readily be made with
fairly simple apparatus. Thus, if an ordinary needle is drawn across
a machined surface and a record of the vertical movements of the
needle are kept, the surface profile is revealed.
The major problem, of course, arises from the very small size of
these vertical movements, and we must therefore be able to obtain
considerable magnification of the vertical movement of the needle.
This can be achieved by mechanical, optical or electrical means and
consideration may be given to possible pieces of equipment which can
be manufactured which will perform this function.
A simple mechanical profilometer identical in principle to some
early industrial instruments such as the Tomlinson recorder is shown
in fig. 2.20.

<-~ 35~7n~

Fig. 2.20. A simple profilometer.

A block of wood (a) having a smooth lower surface forms the basis
of the arrangement. Two needles (b) of identical diameter act as
low friction bearings for the pointer spindle (c) and a spacer (d) cut
from a needle of the same diameter. Needle (e), acting as the stylus,
is loaded against (c) and (d) by elastic bands so that vertical displace-
ment of (e) causes the pointer spindle to rotate and gives rise to a
proportional displacement at the tip of the counterbalanced pointer
38
(g). A lightly smoked glass screen gently touching the tip (lz)
provides a permanent record of its movement. The profilometer
is drawn across the surface to be examined at slow speed by a weight
and pulley arrangement.
Assuming that the vertical movement of (e) is small and that no
sliding occurs between (e) and (c), the height of an asperity, y, can
be deduced from the height of the trace mark, Y, from the equation
R
y=IY
where R = radius of the needle (c)
L = length of the pointer (g).
The surface profile on the smoked glass plate can be further
magnified by projecting an image using a standard slide projector.
The image on the screen may then be recorded to provide a permanent
large magnification of the surface profile.
This equipment, or any alternative design using other magnifica-
tion principles, may then be used to explore the geometrical character
of a wide variety of surfaces. These may be produced by a wide
range of machining processes, varying from abrasive processes using
emery paper to turning in a lathe. It is interesting then to consider
ways in which one may use the profiles of the surface in order to give
numerical assessment of the quality. One simple method ,vould be,
for instance, to determine the distribution of the ordinates and then
obtain the standard deviation of such a distribution. This is usually
referred to as the root mean square value of the profile. Other
measurements of this kind will no doubt suggest themselves or may be
obtained by reading the literature in the field of surface measurement.

Taper sectioning
Another method of examining surfaces is to use a taper section.
Ideally the specimen surface is plated and then sectioned at a small
angle (fig. 2.6). The sectioning must be carried out carefully to
avoid distortion of the sectioned surface. The plated layer helps to
avoid such surface distortion during the sectioning and also provides
a contrasting line of demarcation to reveal the surface profile when
viewed through a microscope.

Hardness testers
The most common way of measuring the hardness of a surface is by
indentation-pressing a hard ball or diamond pyramid into the surface
with a known force and measuring the size of the impression made.
In many projects an absolute knowledge of hardness is not required;
comparative measurements will suffice.
D 39
Two ' home-made' hardness testers are described in School Science
and Technology edited by Dr. T. Kelly, and published by Schools
Council Project Technology on behalf on the Schools Science and
Technology Committee, 1969. One of them is based on a commercial
Rockwell hardness tester with a diamond penetrator, the other is a
portable tester using precompressed springs to apply known forces
and a drill bit as indentor. By suitable choice of spring, measurements
can be made on a wide variety of materials. This type of equipment
is also useful in the study of the nature of the contact between balls,
cylinders and planes using the methods described by K. L. Johnson
(Bulletin of Mechanical Engineering Education, 6, 245, 1967).

40
CHAPTER 3
friction

3 .1. Introduction
FRICTION, as the resistance to motion when bodies slide along each
other, is part of everyday experience. The force system at the contact
can be resolved into a normal ( or load) force together with a tangential
friction force (fig. 3 .1 ). It is also usual to define two states of friction,
namely, the static friction force, which is the force to initiate sliding,
and kinetic or dynamic friction force, the force to maintain steady
state slip.
Although Leonardo da Vinci and Newton had indicated the nature
of the 'Laws of friction', it was not until some time later, in 1699,
that the first two were expressly stated by the French engineer
Amontons, whose name they now bear. The third law is due to
Coulomb.

µ=F/W

Fig. 3.1. The definition of the coefficient of frictionµ.

These laws are entirely empirical and, although they are, surpri-
singly, valid for most situations, exceptions exist. Of the three laws,
the first two are the most important in terms of their range of applica-
bility, the third being most frequently contravened. These three
laws of friction are:
(i) That friction is independent of the apparent area of contact.
Thus the friction between a brick and a plane does not depend
on whether it stands on its small end face or its larger faces.
The reasons for our interest in real areas of contact discussed in
Chapter 2 are now apparent. Here we saw that the real area
of contact depends only on the load, no matter what the overall
size of the geometric contact. Since friction must occur at
41
the real areas of contact, we can now see that this law has a
sound physical basis. The proof of this proposition, as out-
lined in Chapter 2, came some 260 years after the original
empirical statement.
(ii) That the friction force is directly proportional to the normal load.
The ratio of the friction force to the normal load is called the
' coefficient of friction ', µ. Again, it is only in recent years
that the scientific explanation of this law has become at all
understood. A material does not on its own have an intrinsic
coefficient of friction; ,ve must always consider the value ofµ
between two solids. Also the environment within ·which the
bodies slide upon each other can markedly affect the value ofµ.
Thus clean dry steel sliding on steel in air gives µ = O· 5,
whereas the same materials in ultra-high vacuum give values
of µ some ten times greater; note that this means F = 5 TV!
Graphite sliding on graphite in the ordinary atmosphere gives
µ = O· l but with very dry air as the environment we findµ= 0· 5,
an increase of five times in the absence of water vapour.
(iii) That kinetic friction is independent of the velocity of sliding.
Although this ' law ' is reasonably obeyed by most common
solids at modest sliding speeds, its validity as a scientific
statement is lower than that of the other two laws. Indeed
our current understanding of the effects of speed on kinetic
friction is much less developed than our understanding of the
basis of the first two laws.

3.2. Possible causes of friction


The most obvious single fact about friction is that it is an energy
dissipation mechanism; we always do work against friction. Let us
therefore consider the possible effects that can occur between the
interacting asperities, which we know to exist on virtually all solid
surfaces, which lead to an energy dissipation.
We shall assume that the bulk materials of the sliding solids are in
parallel motion so that no work is done on the bulk of the two bodies.
Our candidates for energy dissipation are then only three:
(a) Plastic deformation which we know represents a permanent
change, and therefore dissipation of energy. (b) Less obviously,
elastic deformation where ideally the work clone during deformation
is recovered during unloading. This perfect behaviour is of course
never actually achieved and there is always some energy dissipation
during an elastic cycle. This is called the ' elastic hysteresis ' loss
and, although it is fairly small for most metals, say 0· 5 per cent for
steel, its value with synthetic rubbers can be as high as 15 per cent.
(c) \Vhenever the deformation leads to fracture and the creation of new
42
surface area energy must be dissipated. Again the energy dissipated
in creating fracture is generally small, certainly as compared with the
energy dissipated in the plastic deformation of metals.
It follows that we shall probably find that the deformation
mechanisms, and particularly that of plastic deformation, are the
most profitable areas in which to seek an explanation of the friction
of metals.

3.3. The simple adhesion theory


We have already seen that all surfaces are rough at the microscopic
level so that when they are placed in contact they will touch at the
tips of mating asperities. As the normal load is increased the
asperities will deform, initially elastically and finally plastically, such
that the real area of contact increases. Furthermore, this real area
of contact will be directly proportional to the load and if the tips of
the asperities are spherical in shape the real area for a given asperity
(fig. 3.2) is given by
(3.1)

where II is the indentation hardness of the material and is defined as


load divided by the surface area of the contact. Here we are speaking
of ' the material ' as if both surfaces are made of the same material.
This point is cleared up shortly, but whether the materials are the
same or not, there is obviously only one value of A 1 between them.

Body 1

W = IW1 F= IF1
Fig. 3.2. The local asperity contacts between surfaces.

\Ve shall now assume that, due to surface preparation or because


of their disruption during the deformation process, the asperities'
contacts are clean, and that there are no surface films so that there is
metal-to-metal contact. If we now wish to slide one asperity relative
to its mate we must shear through the metallic junction which has
been created. Since such a junction must have a critical shear
strength, we can sec that for each junction
F 1 = A 1 s, (3.2)
where F 1 is the friction force for the junction ands is the shear strength
per unit area which is a property of the material.
43
Combining equations (3.1) and (3.2) for this junction gives:
Fl s
w1 H.
For the whole of the contact there will be many such junctions, say n,
so that we see that the total effect is given by:
F=nF1
W=nW1
and
F F1 s (3 3
µ = w= W = H' - )
1
This shows that µ is related to the two constants of the materials in
contact, namely the ratio of the shear strength to the hardness. Both
of these properties are stated in terms of stress, i.e. load per unit area,
so that we see that µ is dimensionless as expected. Both these
properties relate to the plastic properties of the material. In Chapter 2
we saw that, for instance, the hardness is about three times the yield
stress Y of the material. From plasticity theory we know that the
shear strength s is about one half of the yield stress Y. It therefore
follows that

Hard Hard Soft

~ Soft
Xr< Hard Soft

Fig. 3.3. Various combinations of hard and soft asperities.

So far it looks as if we have been rather careless in not stating


whether the materials have the same or different mechanical proper-
ties, and, if they are different, which values of s and H we are using.
In fact it transpires that this does not matter since the ratio s / H is
substantially the same for both hard and soft materials. However
the physical effects would be different even though the result is the
same forµ whatever the combination of materials, see fig. 3.3. W'ith
hard on soft material, shearing must occur in the soft material whereas
with hard on hard or soft on soft shearing may occur at the interface.
Although we shall find some objections to this theory and require
to extend it to meet these objections, it does offer a useful model to
explain many observed effects. The theory suggests that µ would
44
have the same value for all material combinations and therefore does
not appear to allow us to reduce the value of fl· Fortunately, this is
not true as can be seen from the following physical argument. In
fig. 3.4 \Ye consider the effects on Ai, F 1 of three combinations of
material subjected to the same load W1 . The cases of hard on hard
and soft on soft have already been considered and produce the same
value of fl· Now consider the third case where two hard materials
are separated by a thin film of soft material. The area of contact
will now be somewhere between the values for the previous two
conditions and will depend on the film thickness. When t is large
A 1 will approach the hard-soft combination, whereas when t is small
it will approach the hard-hard combination. Since W1 is the same
for all the junctions the coefficient of friction is determined by the
friction force F 1. In Case (a) F1 =A 1s, with A1 large ands small, so

(a) Hard iw,


~ t 1--F,
~
l•Af"1
Soft
Fig. 3.4. The effect of surface films on asperity contacts.

Fig. 3.5. A micrograph of a typical bearing alloy.


45
that as in fig. 3.3, µ,:::::1/6. In Case (b) Fi=Ais, with Ai small and
s large, so that again as in fig. 3.3, µ,::::: 1/6. In Case (c) Fi =A 1s, with
Ai intermediate ands small, thus Fi is less than in Case (a) or (b) and
produces a lower value ofµ,.
This is an important result since it shows that low shear strength
surface films produce a reduction in µ,. In effect what happens is
that load-carrying capacity is obtained by the underlying hard
material, which keeps the areas of contact small, whilst transverse
shearing is confined to the soft surface film. This principle is the
basis of many bearing systems, such as the use of thin soft-metal
films on hard steel backing material in many bearings, or the combina-
tion of soft materials with a hard metal matrix in other bearing
alloys, fig. 3.5.

Fig. 3.6. The general contact between surface asperities.

3.4. Extensions to the simple theory


We may now consider the limitations of the foregoing theory. It
will be recalled that the theory assumed adhesion at the metallic
contacts and produced a value ofµ, of approximately 0· 17, whereas
in experiments in high vacuum where such clean contacts may be
assumed values ofµ, much greater than this are obtained. The theory
also predicts the same value of µ, for all materials and, although the
values ofµ, for most materials are fairly similar, they are not identical.
Finally, if we examine fig. 3.3 in detail we can see that the transverse
displacement of asperities for the hard-soft combination is not a case
of simple shear but must involve considerable additional displace-
ment of material. We may extend this latter reservation by observing
that asperities may interact in rather more complicated ways than
those postulated so far. Thus in fig. 3.6 we consider asperities which
make contact along their flanks with considerable deformation as well
as simple shearing. Such deformation effects may be accommodated
by the inclusion of an extra force called the ' ploughing term ' in the
expression for F.
The above theory may be improved by adopting a more realistic
approach to the mode of deformation. In the preceding theory we
considered the deformation due to the normal load to define the real
46
area of contact A, and then superposed a tangential force and assumed
that A still had the same value. This is not the case, and plasticity
theory shows that the application of the tangential force will actually
increase the area Ai at each junction. In fact one finds that
2 2
2 Wi
Ai = (
H ) Fi )
+ ex ( H (3.4)

where ex is a constant. Clearly when Fis zero equation (3.4) is the


same as equation (3.1), but as Fi increases Ai increases, so that we
introduce the concept of 'junction growth ' as Fi is slowly increased
until slip occurs.
In the previous theory we also assumed that the surfaces were
clean, in the chemical sense, whereas most surfaces have some form
of contamination. Thus many metals are covered with oxide films
in air, whilst on many surfaces thin lubricant films are present. If
we consider the strength of junctions in terms of their resistance to
shear we may introduce a coefficient c which relates the actual shear
strength sr to the basic shear strength of the softer material,
Sr= CS,

where c lies between O and 1. When c = 0 we have no junction shear


strength whereas when c = 1 we have the condition described in the
earlier theory.
Combining these two effects yields an expression for µ,,
C
(3.5)

\;Ve now see that when c approaches 1 (i.e. complete adhesion of the
junctions) µ, reaches very high values, whereas when c is small the
values are more typical of those observed.
Finally we may extend these arguments by considering a more
general asperity interaction of the type shown in fig. 3.6. If we define
the angle of the initial contact of such asperities as 0 we can carry out
detailed analysis which ultimately yields values of µ, for each 0 and
each value of c defining the interfacial shear strength. The curve for
0 = 0 is of course the solution given by equation (3.5). Typical results
are shown in fig. 3.7, while fig. 3.8 shows schematically the progress of
the interaction between the asperities when c is small and when c is
large. It is interesting to note that when high adhesion occurs,
i.e. c is large, the junctions finally break by being pulled in a tensile
fashion as one surface moves relative to the other. Evidence of such
behaviour is found in experiments carried out in ultra-high vacuum
on clean surfaces where adhesive effects would be expected to be
strongest. This theory also predicts that when 0 = 0, as in the
conditions considered earlier, one obtains the solution represented by
equation (3 .5).
47
8=20°

µ.

c---+- 10
Fig. 3.7. The variation ofµ, with c for different asperity slopes.

~ interface - (a)

) --(
>==< >
--+- -

(
------l>--
------ -+-

~-c
(b)

·c_ ~"''"'""
~ ____ __.._
~ -► --+

Fig. 3.8. The deformation of a single asperity during sliding (a) when c is
small and (b) when c is large.

48
Although the developments to the basic theory have only been given
in outline, it should be clear how they have produced a more useful
approach and how they have overcome the objections to the simple
theory. In particular we are now able to appreciate the great effect
of surface films on frictional behaviour. Thus we note how much
oxide films on metal surfaces, which inhibit the adhesion at the asperity
junctions arc responsible for the low values of µ, usually observed.
Indeed in mechanisms for space vehicles which have to operate in a
non-oxidizing environment many tribological difficulties arise due
to the absence of such oxide films, as c tends to approach unity.

3.5. Friction between elastic bodies


The preceding explanations of friction have invoked the plastic
behaviour of materials and we now consider the results ·with materials
which are not likely to be plastically deformed. Consider the inter-
action of a rubber surface with a steel surface. The surface inter-
actions in such a case will be elastic, and even metallic surfaces after
they have been running together for a long period, i.e. the ' run in '
condition, may be behaving in an elastic rather than a plastic fashion.
In Chapter 2 we concluded that for many surfaces the real area of
contact is directly proportional to the load, even when the asperities
are deforming elastically. Such surfaces, whose asperity distributions
are of a Gaussian type, thus produce an equation of the same type as
equation (3 .1 ). Thus,
A=KW (3.5)
where K is a constant depending upon the elastic properties of the
materials and the characteristics of the surface geometry. The
tangential force will now be concerned with the shearing of surface
films at the contact whose strength may be defined by cs. Therefore
F=Acs,
and µ,=csK. (3.6)
This again produces a value forµ, whose value depends on constants
of the system.
These arguments may again be extended to take into account the
energy losses in the deformation which always occurs in surface
asperity interactions. In the case of elastic materials this energy
loss arises from the hysteresis effects which arc negligible for metals
but are very significant for rubber and some other materials. Indeed
with synthetic rubber where the hysteresis losses are as high as 15
per cent, this term becomes the predominant frictional effect. Such
rubbers have high friction and are therefore used for the treads of
vehicle tyres, since high friction between the tyre and the road is
required for effective traction and braking.
49
3.6. Rolling friction
Wheels of all types and rolling elements such as belt drives (where
the pulley effectively rolls along the belt) and a wide range of rolling-
contact bearings are so well established that we must consider the
nature of friction in such devices. \Ve can attempt to classify the
various types of rolling contact as
(a) Free rolling
This is the condition when a rolling element rolls along a plane,
and the resistance then arises from the basic rolling friction
between the element and the plane.
(b) Rolling subjected to tractions
If a rolling element is subjected to a braking or driving torque
these torques give rise to frictional effects at the contact.
(c) Rolling in conforming grooves
When a ball rolls around the inner race of a ball bearing the
geometrical conformity between the ball and the groove gives
rise to a frictional resistance.
(d) Rolling around cun:es
\Vhen a rolling element travels along a curved path, frictional
effects at the contact are inevitably introduced.
\Yhenever rolling occurs the free rolling friction must occur,
whereas (b), (c) and (d) occur separately or in combination depending
on the particular situation. Thus the real wheel of a car involves
(a) and (b) whereas in a radial ball bearing (a), (b) and (c) are involved,
and in a ball thrust bearing (a), (b), (c) and (d) occur.

3.6.1. Free rolling


Consider a cylinder of length L subjected to a load W rolling along
a plane (fig. 3.9). The size of the contact zone and the pressure
distribution arc (Chapter 2, section 2.6):
= 18WR(l-v2)1112
b rrLE
2 1 2
_ 2W ( x ) 1 (3.7)
P- rrbL l - b2
\Ve assume that no slip occurs within the contact zone but we note
that the material of the wheel in the front half of the contact zone
is being compressed elastically, whereas it recovers from such com-
pression in the rearward part of the contact zone. This process of
compression followed by unloading continues as rolling proceeds.
If the materials were perfectly elastic the work done in forward
50
compression would be exactly equal to the work received in rearward
expansion, so that no nett energy would be dissipated. \Ve now recall
that in elastic stress cycles there is always a nett energy loss due to
hysteresis. There will be a loss of energy during rolling due to this
cause, and this constitutes the free rolling resistance. \Ve may
define this resistance by the following argument.

L ~surep

~.x
~ 2b i - ~zb__::;:--dx

Fig. 3.9.
fl
The contact of a cylinder and a plane and the resulting pressure
distribution.
Consider only the forward part of the contact zone. The contact
pressure acting over a small element width dx and length L gives rise
to a force pLdx. The resisting moment due to this force about the
centre line of the contact is pLdx x. Thus the total resisting moment
Ai due to the forward compression is
/)

M= f p . L . x . dx.
0

Substituting for p and b from equation (3.7) gives


~f=2Wb (3.8)
lY. 1T ,
3
Now consider the cylinder to roll a distance S, when it will turn
through an angle 0. The total work done in forward compression is
M0. Elastic recovery in the rearward part of the contact zone means
that most of this work is recovered, but let elastic hysteresis result in a
nett loss E per cent. The work done against hysteresis is thus

1~0 M0.
Suppose the friction force moving the cylinder the distance S is F
and note that S = 0R.
51
Work done by F =Work done against hysteresis
E E S
F S=-M0=-M-
. 100 100 R.
Thus
Elli/
F= lOOR' (3.9)

Substituting for M from equation (3.8) in equation (3.9) gives


F= E. 2Wb_ (3.10)
3001TR

This result is interesting in that it shows that the free rolling resistance
of a wheel will be less if W is small and R is large. This clearly
explains the advantages of a spoked wheel in reducing friction.

3.6.2. Rolling subjected to tractions


Consider the driving wheel of a locomotive or of a car (fig. 3.10).
At the contact between the wheel and its track we must produce a
force T on the wheel in the direction of motion to drive the vehicle.
Furthermore this force must be less thanµ W otherwise the wheel will
slip and forward rolling motion will cease. We immediately see that
when T= 0 we have no slip between the wheel and the track, whilst

-v

l•2b+j
t
L
Non -slip area
t
Slip
area .j 2 .BI+

Fig. 3.10. The microslip in the contact zone of a cylinder rolling along a plane.
52
when T = µ, W slip occurs throughout the contact area. The interes-
ting question is what happens when O < T < µ, W. The answer may
at first sight seem somewhat strange but it will eventually be seen to be
entirely reasonable. What actually happens is shown in the details
of the contact patch shown in fig. 3 .10. A part of the contact patch
is subjected to slipping whilst in the remainder at the front of the
contact zone no relative motion occurs. As T increases the size of
the slip area increases and the whole area is slipping when T = µ, W.
Theoretically it can in fact be shown that the degree of slipping is
related to T by the formula,

.I_ = [
µ,W l
-(@)2]1/2
b '
(3.11)

i.e. when f3=b, T/µ,W=O, and when /3=0, T/µ,W= 1.


To explain this effect we must recognize three things: the degree
of slip is microscopically small, the materials in contact may deform
elastically, and the motion being rolling means that particles on the
rim of the wheel pass through the contact zone and out the other side
as motion proceeds. Thus if we imagine particles approaching the
contact zone they do not slip relative to the track because of a build-up
of elastic deformation, but in due course these built-up deformations
are released and microslip between the particles and the track ensues.
The particles then pass out of the contact zone and all the strains arc
finally released.
One interesting effect of these microslip mechanisms is that the
wheel does not roll exactly the distance one might expect from purely

8R

t rm
I
1 T l·O
µW

Fig. 3 .11. The variation of the creep with the tangential traction for a rolling
cylinder.
53
geometric considerations, i.e. S #- RB, the condition which was used
in the previous section. Thus a driving wheel rolls slightly further
than RB whilst a braking wheel rolls slightly less. These so-called
' Creep effects ' have been measured and shown to agree with this
type of theoretical treatment, fig. 3 .11, which yields

(3.12)

where 8 is the fractional creep which occurs due to the traction T


for a wheel of radius R producing a contact width b clue to the load W.

3.6.3. Rolling in conforming grooves


Imagine a ball rolling along a groove as shown in fig. 3.12. A basic
kinematic condition for rolling motion is of course that there must be
an instantaneous axis in the contact zone where no relative motion
occurs between the ball and the track. Thus the ball in fig. 3.12
might be considered to be instantaneously rotating about the axis AA.
This implies that slip must occur at contact points remote from this
axis. Such slip is of course of fairly small magnitude and the actual
behaviour is complicated by surface elastic straining taking up some
of the requirement to slip. This results in complex patterns of
sticking and microslip areas of the type shown in fig. 3.12.

w w

I
' I Contact

~ffl
' I
' / - width - -
----~--r'•~-___.,. __ -·-·-A- - - A

j 2bL
1+ +120
Non slip
2b a~a
_t_ 1

Slip_,,f
areas

Fig. 3.12. The microslip in the contact zone for a ball rolling along a grooved
track.
54
3.6.4. Rolling around a curl'e
A rolling ball has to be forced to roll around a curved path by the
application of a torque 1112 , fig. 3.13. This torque then creates
microslip patterns of the type shown in fig. 3 .13, where once again
the elastic straining of the surface can accommodate some of the slip
requirement.

'"', Direction of rolling


/

'-.

Plan view of ball rolling around a curved path

Fig. 3.13. The microslip in the contact zone for a ball rolling around a
curved path.

3.6.5. Vehicle cornering


As an example of the preceding arguments imagine driving a front
wheel drive car around a corner. \Vhatever the details of the road
wheel friction we can represent the effects by a friction force F acting
at some point X. This system is entirely equivalent to the three
components Fx, Fy, and sWz as shown in fig. 3.14. These forces and
moments are identifiable as:
Fx The tractive force driving the vehicle in the direction of
motion;
Fy The tractive force transverse to the direction of motion giving
rise to the required change of direction of the vehicle;
1\12 The torque ·which demonstrates the reluctance of the wheel
to changing its direction of motion, i.e. the self-aligning
torque which we feel at the steering wheel as a desire to
straighten the car.
These three effects are similar in kind, but rather more complex
in treatment, to those discussed earlier. For example the effects due
to the force Fx was discussed in section 3.6.2. The relationship
between the cornering force F 11 and the self-aligning torque l\12 are
best understood from the Gough plot, fig. 3.15. Along AB we see
increased cornering force associated with increasing l11z, i.e. ' feel at
the wheel '. From B to C the cornering force increases very much
E 55
w

I y I F

Contact
zone ~ x
~

MF~
Fig. 3.14.
0· bF·
The friction forces in the contact between a tyre and the road.

tI D
LL'."
Q)
;:
-2
0, C
'=
~
Q)
C

0
u

A Self-aligning torque M2 ---..-

Fig. 3.15. The Gough plot relating cornering force and self-aligning torque.

56
but we hardly notice any change at the steering wheel due to Mz.
From C to D is the most remarkable situation where we find reversed
feel at the wheel as Fy increases. Regions B to C and C to D are
strictly for the professional driver. As we move from A to B the
slip means that the centreline of the wheels is a few degrees out of line
from the direction of motion, but from C to D one has much greater
misalignment as the car drifts around the corner. The ultimate value
of F y is defined by µ W, the point at which pure sliding occurs and
control is lost.

3. 7. The experimental measurement of friction


The simplest method of determining the friction between solids
is to use the principle of the inclined plane, fig. 3.16. The normal
force is now W cos 0 and the frictional force is W sin 0, giving
µ,=tan 0. Thus by slowly inclining the plane until sliding occurs
we obtain a value for µ,. Although basically simple this device is
clearly limited to relatively small loads and is totally unable to study
the effects of sliding velocity.
A more sophisticated arrangement is shown in fig. 3.17 which
overcomes the disadvantages of the inclined plane. A stationary

µ=ton 8
Fig. 3.16. The value ofµ defined by an inclined plane.

Loading lever Load

Fig. 3.17. A simple crossed cylinder friction and wear machine.


57
specimen is loaded against a rotating cylinder. The system is loaded
by dead weights usually through a lever system, although in some
cases hydraulic or magnetic forces may be used. The stationary
solid is in the form of a sphere, a pin or a cylinder with its axis at
right angles to the rotating cylinder. The stationary specimen is
mounted in a frame ·which is supported such that it may have small
transverse movement. This frame is connected to a force transducer
which records a measure of the friction force. The value of the
friction force may be continuously recorded at any desired load,
speed and surface condition, e.g. with some form of lubrication.
An alternative arrangement loads a pin against the top surface of a
rotating disc, as in fig. 3.18. Such machines can be arranged to
operate under various environmental conditions such as temperature,
humidity, pressure, and in a variety of gaseous or liquid environments.

Load

Pin

Friction
force

Fig. 3.18. The pin and disc friction and wear machine.

Where friction has to be studied in conditions of ultra-high vacuum,


although the basic arrangements remain the same, several additional
refinements have to be incorporated. Such an apparatus is shown in
fig. 3.19, and the following points should be noted. The machine is a
pin and disc arrangement but it will be noted that the rotary drive
outside the vacuum chamber drives the disc via a magnetic drive.
The normal load is applied externally by a lever system which passes
through a flexible bellows moubted on gimbal bearings. By reversing
the direction of normal load using the loading motor it is also possible
to determine the adhesion between the pin and the disc. The friction
force is obtained by measuring the transverse force on the loading
58
lever. Such an arrangement allows one to study friction and adhesion
effects under carefully controlled clean conditions.
l\!Iany other arrangements for the measurement of friction are
clearly possible but the foregoing illustrates the most usual arrange-
ments.
For the study of rolling friction the most commonly employed
system is the disc machine in which two rolling discs arc loaded
together (fig. 3.20). One disc is driven at constant speed \vhile the
other disc is subjected to a resisting torque M so that such effects as
the creep due to microslip may be studied.
For the study of free rolling resistance, difficulties are experienced
because of the very small value of such resistance, particularly with

Flexible metal
Disc bellows

Vertical force
...- transducer

~-~--=L"'-'----L..J-----~----__.,__.-To friction
'----~~----~~------r force
transducer

Base
Vacuum pumps

Fig. 3.19. A pin and disc machine for use in a U.H.V. chamber.

Driven
disc

Fig. 3.20. The principle of the disc machine.


59
metals. One ingenious method for such studies is to cause a ball to
roll in an oscillatory fashion along a plane surface by incorporating it
into a compound pendulum as shown in fig. 3.21. The rolling ball
acts like the knife edge in a normal pendulum and it is worth noting
that all knife edges are in effect very small radius pivots on which the
levers roll. By studying the decay of the ensuing vibrations of the
system one may evaluate the very small rolling friction. The effect
of load on rolling friction is studied by varying the weights used in a
series of experiments.

Test surface

w w

Fig. 3.21. A pendulum rolling friction machine.

3.8. Project suggestions


Many forms of friction tester are used and some have been briefly
described in Section 3.7. The inclined plane is probably the most
elementary form, but pulling loaded bodies along by a spring balance
also offers a very simple machine. Try pulling weighted sections of
rubber tyres along a variety of road surfaces using a spring balance to
measure the friction force.
A lathe offers a useful basis for a fairly complex friction tester and
one such instrument has been described by Halling ( Wear, 4, 22,
1961). The basic arrangement is shown in fig. 3.22 (sec also fig. 4.15),
the load being applied via a lever and the friction force measured by
the deflection of a cantilever spring indicated by a dial gauge. Other
methods can be used such as a friction band loaded against a rotating
shaft as shown in fig. 3.23. A method suggested by Boothroyd
(Bulletin of Mechanical Engineering Education, 9, 219, 1970) is shown
60
Fixed cylinder

Cylinder mounted in lathe

Fig. 3.22. The principle of the crossed cylinder machine.

Shim steel
\

Steel shaft held in variable


speed motor or in lathe or
in drill chuck

Frictionµ.= "#loge(;~)

Spring balances

Fig. 3.23. A kinetic friction tester.

61
in fig. 3.24. A slider is pushed along a flat surface by means of a
straight edge. The new position of the slider can be shown to be
defined by the friction between it and the straight edge.
In all friction experiments such effects as surface finish, surface
coatings and lubricants may easily be studied.

F µ. 13 W sin (8 +1/t)
µ. ---
12- N - fl-13 W cos (8+1/r)

= cot (8+1/rl

I /
\ I
\-:i;7
I . Flat surface ( 3)
. I
I• I
I
\ !
I I
I

8
~ Sliders

Straight
edge
(2)
Initial position


Fig. 3.24. The lloothroyd friction tester.

62
CHAPTER 4
wear and the properties of materials

4.1. Introduction
AT the most elementary level wear is manifested by a loss of surface
material from one or both surfaces when they are subjected to relative
motion. Sometimes the wear is clearly visible to the naked eye; in
other cases it requires the most elaborate measurement techniques
for its detection. The relative motion producing the wear may be
macroscopically observable as in many machine elements, whereas in
other cases the motion may be a small microscopic vibration, when we
speak of fretting. An example of fretting is the wear and subsequent
corrosion we can see between a bolt or rivet head and the surface in
systems subject to severe vibration.
At the outset we might also note some other characteristics of all
wear phenomena. Wear takes place in real time and we cannot affect
the current rate of wear of materials unless we alter the load, speed,
lubrication or environmental conditions. Nonetheless as time
elapses the rate of wear may either increase or decrease due to the
complexity of the mechanics involved. Although wear clearly occurs
at the surface contact which is also the seat of the friction mechanism,
the two are not simply related one to the other. Cases have already
been shown in table 1.1 where low friction does not necessarily mean
low wear and vice versa. Indeed even with a given combination of
materials one sometimes meets situations where as time proceeds the
wear rate increases and the friction decreases.
Two other general observations about wear can be made. In
Chapter 1 it was pointed out that wear may often prove beneficial as
in ' running in ' or when employing the principle of ' planned
obsolescence '. In a more obvious way we use wear in the workshop
in filing, grinding, and in a whole range of abrasive processes.
Secondly, we must be careful about the units in which we define and
measure wear. Suppose we define it as the mass or volume of material
removed per unit distance of sliding. Now consider the wear of a
shaft rotating in a bearing. If the shaft has diameter d and the bearing
length is L, the change or in radius r per revolution due to a volume
wear V per unit distance of sliding is obtained from
LTrdor = V Trd,
63
so that
V
8r=]_;

The engineering designer is interested in 8r, which is the increasing


slackness due to wear in each revolution, and clearly he must think of
V /L rather than just Vin defining this value. Thus, although Vis a
valuable method of defining wear in identical geometrical situations
such as may be used in standard wear test apparatus, we must always
recognize that the designer is usually more interested in the dimen-
sional changes due to wear.
The most general definition of wear is usually taken as ' the
progressive loss of substance from the surface of a body brought about
by mechanical action '. Although it may be subject to some criticism
this definition of the phenomena will suffice for our purposes.

4.2. Types of wear


Before proceeding to any detailed study of the mechanisms which
can result in wear, we shall introduce the simplest classification of
wear into mild wear and severe wear. In Chapter 3 we have noted that
surface films such as oxides can reduce the degree of interaction
between the bulk materials in contact. Mild wear is therefore
generally associated with low loads where metallic interactions are
somewhat inhibited and the wear debris consists of fine particles and
is usually in the form of oxides. This does not imply that metallic
contacts have never occurred at all, since the resulting metallic debris
would tend to become oxidized at the high local temperatures at the
interface. Nevertheless the nature of the surface-asperity interaction
is relatively gentle, resulting in characteristically mild wear and a
smoothing of the surfaces.
At higher loads a much coarser wear process occurs. The wear
debris is of a much larger particle-size, the worn surfaces are much
rougher and the increase in volume wear rate changes by several
orders of magnitude. This is the so-called severe wear regime. A
startling fact about these two types of wear behaviour is the very
rapid transition from one mode to the other as the load is increased
(fig. 4.1 ). In this figure it should be noted that the wear rate suddenly
changes by more than one hundred times. With some materials at
even higher loads the increasing temperatures cause metallurgical
changes in the materials such as to increase their inherent hardness.
These effects can then lead to a second transition from the severe
wear back to the mild wear regime.
64
10-6

10-7

!':! 10-S
l?
L..
0

~
,o-9 ~ /
Transition

10-,o

Mild

10 10 2 103 10 4 10 5
Lood
Fig. 4.1. The transition phenomena in wear.

4.3. Mechanisms of wear


We now consider the detailed mechanisms by which material may
be removed from the surface. The most common mechanisms are:
(a) Adhesive wear,
(b) Abrasive wear,
(c) Surface fatigue,
(d) Corrosive wear.
In some situations more than one of these mechanisms may be
operative at the same time, and this is one of the reasons for the
complexity of wear studies.

4.3.1. Adhesive wear


In Chapter 3 we saw that the contact between surfaces occurs at the
tips of the asperities which then deform under load. The nature of
the adhesion between such asperity interactions is modified by surface
films so that the metallic adhesion characterized in the simple friction
theory is somewhat modified. But as translation occurs these surface
films are to some extent disrupted and adhesion will occur at a certain
proportion of these contacts, as can be appreciated in view of the
distribution of asperity heights discussed in Chapter 2.
From these considerations and our knowledge of the nature of the
contact of rough surfaces we can now predict a ' wear equation '.
If one assumes that the wear particles are geometrically similar, the
wear volume would be expected to be proportional to the real areas
65
of contact at which adhesion occurs, and also to the distance of sliding.
Since the real area of contact for the plastic interaction of asperities
is given by
A= H
w (see equation (3.1))
the wear volume V is given by

Vcx:AxLcx;xL
where L 1s the distance of sliding. Thus the adhesive wear law
becomes
V=KWL
H'
or in words the ' laws ' of adhesive wear are:
(a) The volume of wear is proportional to the distance of sliding.
This relationship has been justified by experience for a wide
range of conditions.
(b) The 'l'olume of wear is proportional to the applied load. This has
also been shown to be true in many tests for limited ranges of
load, although as wear mechanisms change with increasing
load some abrupt transitions have been observed; see fig. 4.1.
(c) The z•olume of wear is in'l•ersely proportional to the hardness of
the softer material. This has also been shown to be valid,
particularly for pure metals.
Recalling the physical nature of the wear process arising from
adhesion we can give a physical meaning to the constant of propor-
tionalitv K, often called the adhesive wear coefficient or the Archard
constari't. Surface films together with the height distribution of the
asperities sec to it that adhesive effects are only significant at a certain
proportion of the asperity contacts. Also the mere fact of adhesion
does not of itself detach ,vear particles, since shearing clue to sliding
could still occur at the junction interface. Wear particles are
probably formed where the junctions are stronger than the underlying
material, so that some material from the surfaces is torn away and
eventually released by the continued sliding. So we can see that K
may be interpreted as a probability factor, that is, the factor which
indicates the probability of wear particles being created by the
adhesive effects between the populations of asperities on the two
rubbing surfaces.

4.3.2. Abrasive wear


This type of wear arises from the cutting action of hard surfaces
rubbing on softer materials, as for example, when hard surface
66
asperities act rather like cutting tools and remove material from
softer materials. Another example arises when loose debris of
any kind is trapped between sliding surfaces. Such debris may be
extraneous, such as sand particles, or may be the actual wear particles
created by the primary wear process.
One method of reducing the first type of abrasive wear is to ensure
high quality of surface finish of the mating surfaces, particuarly the
hard surface. \Vith modern production methods this type of wear
is no longer a serious problem. The second type of abrasive wear is
more difficult to eliminate. Suitable sealing and filtration can reduce
the wear due to extraneous material but the abrasive action of wear
debris is a greater problem, and one which shows the importance of
correct design of the contact geometry. It is often desirable to
provide grooves or other such recesses on the surfaces of bearings
which allow the debris to ' escape ' from the contact geometry.

Fig. 4.2. \Vear due to a single conical asperity.

A simple but useful equation representing abrasive (cutting) wear


is obtained by considering the action of a single asperity. For
simplicity, suppose the asperity has the conical geometry of fig. 4.2.
In traversing a distance L this asperity displaces a volume of material
Lr 2 cot 0. If the yield contact pressure is Po the normal load carried
by this asperity is ½rrr 2p 0 , the factor 1 arising because only the front
part of the asperity makes contact with the soft material, and if there
are n such asperity contacts between the surface, the total load TV is
given by
W=Jnrrr2Po·
Likewise the total volume of material removed by the n asperities is
V = nLr 2 cot 8.
2
Eliminating nr from these equations gives,
V= 2 co~ W L.
rr Po
67
From Chapter 2 we recall that Po= ½H, where H is the hardness of
the material, so that we may write

This equation bears some resemblance to the adhesive law discussed


earlier, though one difference is noticeable, for Ka is now seen to be
related to the characteristic geometry of the hard asperities as defined
by the angle 0. Also, when abrasive action is the major wear mechan-
ism, Ka has a much higher value than the constant K in adhesive
wear mechanisms.
Although the abrasive law has been derived using a simple conical
asperity model it is also appropriate to the cutting action of loose
wear debris. In such cases the value of Ka is rather smaller. Many
investigators have found this law to be substantially true in wear
experiments.

4.3.3. Fatigue wear


It is well known that if materials are loaded and unloaded cyclically
they exhibit fatigue failure. This type of failure can occur after a
large number of loading cycles, even though the load is less than that
which we would normally expect to produce failure in a single load
application. It is usual to express such behaviour by a logarithmic
graph of stress S against the number N of cycles to failure (the S/N
curve, as it is usually called) such as fig. 4.3. Here we see that the
lower the applied cyclical stress level the longer the life of the material
before failure occurs.

"'c,,
0
..J
I

Log. N - looding cycles to failure


Fig. 4.3. A typical S-N fatigue curve.
68
If we consider the interaction of asperities during the sliding of one
surface over another we can see the possibility of fatigue mechanisms
leading to failure of the asperities. This will result in material
being broken off the asperities to produce wear debris. A simple
experiment to illustrate this is to run one's finger many times along
the teeth of a comb. The teeth (asperities) are continuously being
loaded and unloaded due to repeated traversals by one's finger and
after many cycles they ultimately break due to fatigue.
Fatigue mechanisms of failure are even more important when ,,·e
consider rolling contact. In Chapter 2 it was seen that during rolling
material is compressed at the front of the contact zone and is then
released as the element rolls forward. This is a perfect example of a
fatigue loading cycle so that in a rolling bearing the number of fatigue
cycles is simply related to the number of revolutions of the bearing.
In Chapter 2 we also saw that the maximum stress at the contact of a
cylinder or a sphere occurs a small distance below the surface.
Fatigue failure, which will be initiated where the stress is highest,
therefore often occurs below the surface. Cracks are produced by
fatigue, and these lead to the removal of relatively large pieces of metal
giving the characteristic ' pitting ' failures associated with rolling
contact, fig. 4.4.

Fig. 4.4. A typical surface fatigue failure.

69
Rolling-contact bearings show life characteristics significantly
different from those which apply to rubbing contacts. In the latter,
any wear process is initiated at the start of sliding and the wear is
progressive. The life of such a contact is therefore defined by the
allowable wear which can occur before the the component ceases to
function adequately. In rolling contacts the fatigue failure only
occurs after many cycles of operation, but when it does occur it
produces pitting which brings the bearing to the end of its useful life.
Indeed it is standard practice to define the life of rolling-contact
bearings in terms of their fatigue resistance. If we consider that a
bearing will fail by fatigue after N revolutions of the bearing under
an applied load W it has been found that:
For roller bearings W 2 N = Constant.
For ball bearings W 3N = Constant.
From Chapter 2 we already know that the stresses produced by
spheres are greater than those produced by cylinders so that this result
is not surprising. It is also apparent from the above that the lower
the applied load ( and hence stress) the longer is the life, which is
just the sort of behaviour \Ve ,votild expect from the S /N curve shown
in fig. 4.3.
There is one final point about wear due to fatigue as opposed to
other wear mechanisms, particularly adhesive wear. To produce
fatigue failure all that is required is that the surface material should be
loaded whereas the other mechanisms require not only loading but
also actual physical contact. If the surfaces are separated by a
lubricant film, adhesive and abrasive wear are virtually eliminated.
But the applied load is still transmitted to the solid surfaces through a
lubricant film, and can still cause stresses in the surface, so that
fatigue-type failures arc not excluded. This is particularly true in the
almost perfect fatigue environment of rolling contacts. It is found
that the propensity to fatigue-type failure decreases as the oil film
thickness increases. This is not surprising since as the oil film
becomes thicker it tends to smooth out the local surface stresses.

4.3.4. Corrosi•ve wear


Any clean metal surface reacts with its environment to form
contaminant films, and the rate of formation of such films is initially
very rapid but decreases as the ' corrosive ' film thickens. In many
instances, such as oxide (rust) films on steel, these surface films adhere
only loosely to the surface. Rubbing therefore removes the films
leaving exposed ' clean ' metal which immediately reacts with its
environment to provide new surface films, which are again removed
during rubbing. So material is continuously being removed from
the surface, and wear is taking place. The chemistry of such reactions
70
is beyond the scope of this book, but fortunately is not needed for the
understanding of the basic mechanism.
A further effect of corrosive environments is to enhance the abrasive
action of wear debris. Most metal oxides are harder than the metal
itself so that if metal debris is created, this becomes oxidized and
gives a rate of abrasion greater than that which would otherwise occur.
A good example of this occurs with relatively soft aluminium, where
the oxidized wear debris is a very hard abrasive. Indeed aluminium
oxide is often used as the cutting agent in grinding wheels and the
like.
Corrosive effects are not entirely deleterious. In the chapter on
friction it has already been shown that the presence of oxide films in
preventing metal to metal contact greatly reduces the coefficient of
friction. In other applications surface films are deliberately produced
to avoid metallic contact. The so-called E.P. (extreme pressure)
additives to lubricating oil produce surface films such as chlorides and
sulphides and provide protective surface layers. In a sense these
could more properly be called extreme temperature films rather than
extreme pressure films. Their main characteristic is their chemical
stability at the high temperatures of the high pressure contacts m
such situations as hypoid gears as used in motor car back axles.

4.3.5. Fretting
This is not really a separate mechanism of wear but it is treated
separately because it arises in rather special circumstances. It shows
how one particular wear process may be a complicated combination of
several mechanisms of wear, and also demonstrates the deleterious
effects of any wear debris which may become trapped in the contact
system.
Fretting effects are associated with the contact of surfaces in which
the sliding motion is an oscillation of relatively small amplitude, often
only a few micrometres. Since vibrations occur in virtually all
machines we find fretting occurring between surfaces in contact such
as bolted components, splines and components located by friction
such as flanges shrunk onto shafts. In such joints the vibration gives
rise to small amplitude oscillatory displacements between the surfaces
in contact.
Such contacts can exhibit a form of adhesive wear. The debris
is trapped between the surfaces and can cause a second stage of
abrasive wear. In certain cases this effect leads to the interesting
combination of a reduction in friction, but an increase in wear. The
increase in wear arises from the second stage abrasive effects while
the reduction in friction occurs because the oscillatory motion tends
to produce rough spherical wear debris particles which then act rather
like a series of small balls separating the surfaces.
F 71
Fretting often occurs in a corrosive environment such as air where
the wear processes tend to be accelerated by the rubbing off of loosely
adhering corrosion films and the increased abrasive action of the
harder oxidized wear debris. In such cases the process is known as
fretting corrosion and with steel in air is characterized by a fine reddish-
brown powder known as ' cocoa '. Fretting effects often occur in
machines during transit due to the vibrational shocks incurred. For
this reason great care is usually taken to unload all critical contacts
in the machine before its transit. The repeated ' hammering '
suffered by the stationary ball bearings of a machine during a long
journey leads to a form of fretting called ' false brinelling ', and
consequent unsatisfactory running ·when the machine is subsequently
operated.

4.4. Factors affecting wear behaviour


4.4.1. Hardness
We have seen that wear tends to be less the harder the material.
Furthermore as the load increases it is common for the wear to increase
rapidly as the nominal contact pressure approaches H /3, where H is
the indentation hardness (fig. 4.5). Such a transition from mild to
severe wear is associated with a plastic yielding of the material
immediately below the asperities, a condition which was discussed in
Chapter 2. It must also be noted that the hardness of a material
depends on the temperature, although some materials have ' hot
hardness ' characteristics which enable them to retain their hardness
to high temperatures. Since all sliding results in a high local tem-
perature rise in the regions of contact, the hot hardness characteristics
of materials are particularly significant (fig. 4.6). In general, we
find that high hardness (in particular at the operational contact
temperature) is associated with reduced wear. In any event, one

...
0

~
H/3
I
Contact pressure
Fig. 4.5. The effect of contact pressure on wear rate.
72
12% chrome steel

I Ni-Cr-Mo
(/)
(/)
steel
Q)
C:
~
0
I

400 800 1200


Temperature /°C
Fig. 4.6. Typical hot hardness characteristics for metals.

should always try to keep the applied normal load below that value
which gives rise to nominal contact pressures approaching H/3, where
H is the hardness of the materials at the operating contact temperature,
which equals the ambient temperature plus the temperature rise due
to frictional heating.

4.4.2. Mutual solubility


It has been found that the mutual solubility, or compatibility, of
materials in contact is correlated with the wear rate. In general,
wear is less where the mutual solubility is low and is greatest where
this is high (table 4.1 ). For this reason the greatest wear occurs
between identical metals in contact, where the mutual solubility is
100 per cent, and such combinations should therefore be avoided.

4.4.3. Crystal structure


The crystal structure of nearly all metals may be one of three
configurations: body-centred cubic (bee), face-centred cubic (fee) or
hexagonal close packed (hep). These arrangements are shown in
fig. 4.7. Of these three, the hexagonal close packed has the most
limited deformation characteristics, for it can only deform by slip
along the basal plane. Since wear is associated with plastic deforma-
tion it is not surprising to find that metals such as cobalt which have
a hexagonal close packed structure give good wear resistance. Un-
fortunately, such crystal structures are not always stable under all
operating conditions. Thus, although cobalt has good wear resistance
at low temperatures its wear increases one hundred-fold when the
73
Table 4.1. Comparison of coefficients and compatibility ratings
of metal pairs.

Metal pair Compatibility \Vear coefficient, K


(solubility)

Copper-lead Incompatible 0·10 X 10- 4


Nickel-lead Incompatible 0·21
Iron-silver Incompatible 0·68
Iron-lead Incompatible 0·69
Aluminium-lead Incompatible 1·4
Silver-lead Limited compatibility 2·5
Zinc-lead Incompatible 2·6
Silver-silver Compatible 3·4
Aluminium-copper Compatible 4·8
Aluminium-silver Compatible 5·2
Aluminium-iron Compatible 6·0
Silver-zinc Compatible 8·4
Zinc-zinc Compatible 12
Zinz-copper Compatible 18
Iron-copper Limited compatibility 19
Silver-copper Compatible 19
Lead-lead Compatible 24
Aluminium-aluminium Compatible 30
Iron-magnesium Limited compatibility 38
Iron-nickel Compatible 59
Iron-iron Compatible 77
Copper-nickel Compatible 81
Copper-copper Compatible 130
Nickel-nickel Compatible 290

Note that K varies by a factor of 3000 for the metal pairs in the above table.

contact temperature rises above 417°C. This sudden change is


directly related to the change in the crystal structure of cobalt, which
is hexagonal close packed below 4 l 7°C and face-centred cubic above
this temperature.
In normal environments the oxide films on metal surfaces offer
considerable protection against wear by restricting the extent of
metal-to-metal contact. Other films such as the so-called E.P.
additives fulfil the same function. There are a large number of such
films suitable for this purpose and these are often termed ' boundary
lubricants '. This implies a kind of lubrication which is different to
film lubrication where the surfaces are completely separated by a
pressurized film of lubricant. The purpose of boundary lubrication
is essentially to produce a chemical reaction between the metal surface
and the lubricant in order to produce a protective film. Fatty acids
are often used for this purpose since they react with metals to produce
' soaps ' with excellent lubrication properties. Unfortunately when
74
the local contact temperatures are high these soaps undergo thermal
degradation so that other chemical reactions are used. It is for this
reason, as we have already seen, that the E.P. additives are so valuable
since they produce films of greater thermal stability. Sulphide films
operate reasonably efficiently up to 650°C.
Another class of surface film which is widely employed uses the
so-called ' lamcllar solids ' such as graphite and molybdenum
disulphide. These materials have a plate-like structure, rather like
a pack of playing cards. Although the exact mechanisms of their
behaviour arc still not completely understood, there is no disagreement
as to their effectiveness.

b.c.c. f.c.c.

h.c.p.

Fig. 4.7. The three common types of crystal structure.

In some applications these materials, particularly graphite, may be


used as a solid block rather than as a surface film. A well known
example of such an application is in the thrust collar which \Vas used
for many years in car clutches. Recently these materials have also
been extensively used by bonding them to metal surfaces using organic
resin binders.

4.5. Tribological properties of plastics


In recent years there has been a significant growth in the use of
plastics to replace metals in many bearing applications. In general
the friction and wear of plastics can be explained by the adhesion
75
theories already discussed. The friction coefficients of plastics are
not particularly low, but their main advantage is that they wear at
reasonably low and predictable rates. One notable exception to this
behaviour is PTFE (polytetrafluoroethylene) whose friction coefficient
may be very low, about 0·05. This very low friction value seems to
be associated with the very low adhesion of this material, which is why
it is used in non-stick kitchenware. It is also reasonably hard due
to the mechanical interlocking of its molecules. So it is extensively
used in bearings, where the loads and speeds are modest, since it is
almost self-lubricating and is highly reliable.
A useful design parameter for all types of plastics used in bearings
is the' pv 'factor (which is the product of the nominal contact pressure
and the sliding speed).

-----v

Area A
Fig. 4.8. The change in dimension due to wear.

Consider a block of this material sliding over a metal surface, fig. 4.8.
The rate of energy dissipation against friction isµ, Wv. It is reasonable
to assume that the volume wear rate of the block, V = - d V /dt, is
proportional to this rate of energy dissipation, so
Vocµ,Wv.
The block will therefore wear to a depth d such that the volume
wear V is given by
V=Ad.
Hence its rate
V=Ad
where d is the rate of increase of d with time. Thus
. Ji µ,Wv
d= A oc -r ocpv.
This shows that the rate of change of dimension of the block is
proportional to the product pv. This is the rate of change of bearing
clearance with time in any practical application and as such is a more
useful parameter than the actual volume of material removed.
76
For any material the allowable value of the pv product may be
defined and fig. 4.9 shows a typical result of PTFE if the wear rate is
to be a dimensional change of 25 µmin 100 hours.
Plastic bearings and particularly PTFE bearings must be operated
within their approved pv ratings. These ratings are associated with
wear and are particularly subject to thermal effects due to the de-
composition of the surface. Thus at higher ambient temperatures the
pv factor for such materials is considerably reduced. Such materials
when used in practical bearing designs are often associated with a
metallic matrix which provides additional strength and improves the
thermal conductivity thus allowing the easier escape of the heat
generated in rubbing. When plastics slide, problems can arise from
the generation of electrostatic charges. In such situations, designers
must incorporate earth paths to minimize the build-up of charges.

N
I
107
E
z
0
a..
'- 106
...:,
Q)

en
en
Q)

a: 105

10- 1 10
Speed / m s- 1

Fig. 4.9. A typical' pv' curve.

4.6. The measurement of wear


In most engineering machinery the rate of wear is relatively small,
typically changes in dimensions of micrometres per year. Also wear
takes place in' real time ' and laboratory tests have to devise conditions
where the wear processes are considerably accelerated so that results
are produced in days rather than years. In so far as such tests are
artificially accelerated their results should be viewed with some caution.

4.6.1. High pressure contact tests


These tests produce accelerated results by applying loads over very
small areas of contact. Various geometrical arrangements are used
77
as shown in fig. 4.10. In each of these tests specimen A is the material
being worn away, the degree of wear being measured by either a
change in dimensions or a loss of mass from the material. Many such
machines also measure the frictional force at the contact .

. ·A

Pin on ring Pin on disc Crossed cylinders

Fig. 4.10. Three common geometries used in wear measurement.

Such contact geometries can be studied either in the ordinary


atmosphere or in a totally enclosed chamber where the atmosphere
may be controlled, as to such properties as the gaseous environment,
pressure, temperature and humidity. A very common form of such
apparatus is designed to carry out these tests at various reduced
pressures. In a high vacuum the formation of oxide films is inhibited,
so we obtain useful information on the friction and wear of the
materials themselves. UHV apparatus is also useful for measuring
the adhesion between surfaces in contact, and we have already seen
that such information is very valuable to our understanding of friction.
More complex geometrical arrangements are shown in fig. 4.11.
Fig. 4.11 a shows the four ball arrangement in which a rotating ball
rubs against three stationary balls to produce wear scars whose size is

{a) Load W (bl

I
Retaining cup
Fig. 4.11. The four ball wear machine.
78
an indication of the volume wear. This arrangement is extremely
popular for industrial testing since the test specimens, the balls, are
readily available at low cost from ball bearing manufacturers. Fig.
4.11 b shows the so-called 'disc machine' which is very useful for the
study of wear under combinations of rolling and sliding. When the
peripheral speed of both discs is the same one has pure rolling whilst
a difference in speeds implies some additional sliding. This rather
complex roll/slide process often occurs in machinery, perhaps the
best example being the contact between meshing gear teeth depicted
in fig. 4.12. At the initial contact we have rolling and sliding between
the teeth which becomes pure rolling at the pitch point B followed by
rolling with sliding in the opposite sense during the arc of disengage-
ment.

A Initial contact-rolling and sliding


B Pitch point contact-pure rolling
C Final contact-rolling and sliding

Fig. 4.12. The progress of contact between gear teeth.

4.6.2. Radioactii·e tracer techniques


Many interesting wear results have been obtained by ' tagging '
one of the materials in contact with a radioactive isotope. The
transfer of this isotope onto the non-tagged mate specimen gives an
indication of the wear. Fig. 4.13 shows a typical autoradiograph from
such a test, the bright areas represent the active wear debris distribu-
tion over the wear track. This method also enables the quantity of
wear debris to be measured with the help of standard radioactive
counting instruments.
79
This has also been successfully employed to study wear in actual
operational machinery. Thus in an ordinary reciprocating internal
combustion engine if one ' tags ' the piston rings, the ensuing wear
debris may be monitored continuously in the engine sump oil. This
enables one to measure very small quantities of wear under operational
conditions, which is perhaps the most attractive feature.

Fig. 4.13. An autoradiograph of a wear track. The light areas represent


radioactive material transferred to the non-active specimen.

4.6.3. Surface monitoring


As any wear process proceeds it alters the properties of the machined
surfaces in contact. By measuring changes in such physical proper-
ties as surface hardness and the changes in the surface geometry one
may interpret the extent of the wear. There has recently been con-
siderable research on the surface properties discussed in Chapter 2,
so that now the changes in such properties may be used to assess wear
80
Initial
surface

After
sliding

Fig. 4.14. The change in surface profiles due to wear.

behaviour. Fig. 4.14 shows two profiles of the same surface, one
before use and the other after wear.

4.7. Project suggestions


The wear of load-bearing surfaces is of vital importance in tribology
and occurs under very wide ranges of conditions of load, speed,
temperature, etc. As a guiding principle the wear tester should
reproduce as closely as possible the practical situation, although some
form of accelerated procedure will usually be necessary to obtain
results in a reasonable time.
Wear testers in which a pin (thin rod) of one material is rubbed
against a rotating disc of another have proved to be very versatile.
V cry high contact pressures can be obtained by using a hemispherically
ended pin or small ball. An old gramophone turn-table might serve
as a basis for a tester for light loads.
The wear characteristics of metals and a wide range of materials
may be obtained using a simple wear machine which is made as an
attachment to a standard centre lathe. The test specimens are
conveniently available either as small cylinders or as balls, and these
are pressed against a cylindrical bar, which is mounted between
centres in the lathe.
To achieve this it will be necessary to construct a system which can
be located on the saddle of the lathe, which will be capable of applying
a load normal to the specimen and which should also have a facility
for measuring the friction force. The normal load may most
conveniently be applied by a lever system while the friction force is
most readily recorded by noting the deflection of a simple spring,
such as a cantilever, with a dial gauge. The essential ingredients of
the system are indicated in fig. 4.15. Halling (Wear, 4, 22, 1961)
describes a possible arrangement of this system.
The wear of the test specimen may be measured by either weighting
the specimen or by measuring the size of the scar produced in the
81
contact between the specimen and the rotating cylinder. Simple
geometry will enable one to convert the measurement of the scar into
a volume wear value.
Having once constructed a piece of equipment such as that discussed,
one is able to use it for a wide range of valuable experimental studies.
Such factors as the effect of load, speed, various material combina-
tions, surface finish and different types of lubricants may all be
considered in relation to their effect on the friction and wear charac-
teristics.
Flexible support

Dial
gauge .____.--"9
Cantilever ---- .,,,- .-~

fr1ct1on
indicating()=
r c,'"
0
oe\ ~o 0\e
Weights

Fig. 4.15. A simple crossed cylinder wear machine for use in a lathe.

It is very interesting with this type of equipment to consider the


wear of the specimen when the rotating bar has been thoroughly
cleaned, preferably by a machining process, immediately before the
test. Such machining can most easily be done by rubbing emery
paper against the bar in the area adjacent to that where the contact
between the specimen and the bar occurs.
Abrasive wear is most easily studied by use of the lathe wear tester.
Abrasive cloth wrapped around the cylinder mounted in the lathe
wears away pins loaded against this surface. The effects on wear of
abrasive type and size may then be readily studied.

82
CHAPTERS
properties and testing of lubricants
5.1. Introduction
ALTHOUGH oil is a lubricant, it is not true that lubricants must be oil.
In fact almost any fluid may act as a lubricant, and its most important
single property is its viscosity. The term fluid is also to be interpreted
in its widest sense as including all liquids, gases and vapours. Thus
in many applications it is possible to use that very cheap fluid, air,
as the lubricant in high speed bearings.
It is interesting to ask why oils should have become so extensively
used as lubricants during the past two hundred years. Their
effectiveness resides in their viscous properties, their non-corrosive
properties and their price and availability. After air and water, oil is
the cheapest and most plentiful fluid available on this planet.
Although air is increasingly used as a lubricant it is really only
effective for certain load and speed conditions, and whilst water is
also used as a lubricant its corrosive effect on metals makes it unattrac-
tive as a general lubricant.

5 .2. Viscosity
This is most readily recognized as that physical property which
defines the resistance to flow so that we often associate high viscosity
with such adjectives as ' thick '. The scientific definition of viscosity
is due to Newton, who in the seventeenth century introduced the
idea of layers of fluid sliding over one another, as shown in fig. 5.1.
This sliding induces interfacial shear stresses between the layers and
N e,vton postulated that these shear stresses were proportional to the
rate at which the shear strain was occurring. Thus in fig. 5.1 if we
assume that each layer has a velocity which increases linearly by an
amount ou across a vertical height oy,we may write
OU
TOC-
oy
or in the more conventional form, the tangential stress T is given by
du
T=Y)-
dy
where the constant of proportionality is the dynamic viscosity YJ and
du/ dy is the velocity gradient across the film thickness. The units
83
of viscosity are clearly those of a stress multiplied by a distance and
divided by a velocity. Thus the SI unit is N s m 2 , but the most
commonly used unit is a c.g.s. unit the poise (P) dyn s cm- 2 • For
many lubricants the poise is too large a unit so that the centipoise
(cP) which is one hundredth of a poise is more commonly employed.
Since one poise is 0· 1 N s m- 2 , 1 cP = 10- 3 N s m- 2 • Water has a
dynamic viscosity of 1 cP, air a value 0·02 cP, and lubricating oils
have values which may be as high as 400 cP or as low as 2 cP.

t
t
8y-------~----.-' +-u + 8u
,-;'--'-------,---'- u
y

t
-...-u
Fig. 5.1. Newtonian flow.

From the foregoing we appear to have assumed that the dynamic


viscosity is a constant of the fluid and this is perhaps reinforced when
its other name of coefficient of viscosity is used. In fact Y/ varies
markedly with temperature for all fluids, varies with pressure for a
great many fluids, and in some cases varies with the velocity gradient
(rate of shearing) du/dy. Where YJ is independent of the rate of
shearing we refer to the fluid as a Newtonian fluid, whereas if YJ
varies with du/dy the fluid is non-Newtonian in character. Our
major preoccupation will be with the behaviour of simple Newtonian
fluids such as mineral oils rather than the more complex non-
Newtonian fluids such as greases and oils with certain additives.
In many situations it is convenient to use the kinematic viscosity
rather than the dynamic viscosity. The kinematic viscosity v is the
dynamic viscosity YJ divided by the density p of the fluid:
V=YJ/p.
The S.I. unit of kinematic viscosity is 1 m 2 s- 1 but the c.g.s. unit
the stokes, or its submultiple the centistokes ( cS), is most often used;
1 stokes is 10- 4 m 2 s- 1 .

5.3. Effect of temperature on viscosity


It is a matter of common experience that the viscosity of most oils
decreases with temperature and this is an important effect in tribology.
84
The changes in temperature arise from two causes. First, lubricants
are used in a wide range of ambient temperatures due to both climatic
variations and local thermal environments, such as the temperature
associated with refrigeration machinery or internal combustion engines.
Second, during the shearing of lubricant films the energy dissipated
manifests itself as a heat release in bearings, etc. \Ve would therefore
anticipate that the operational viscosity of the oil will be considerably
less than that of the oil in the can. Designers of machines and their
operators must therefore know the way in ,vhich the viscosity of an
oil varies with temperature.

Z' VI =O
'iii L
0
u
VI
·;; u
.S!
0 H
.,
E
C:
:.::

Temp./°C

Fig. 5.2. The specification of the viscosity index=~=~ x 100.

For lubricating oils the viscosity index (V.I.) is now commonly


used to indicate the approximate effect of temperature on kinematic
viscosity. This index is based on the behaviour of two standard oils.
Gulf Coast oils were thought to have the greatest variation of viscosity
with temperature and were ascribed a viscosity index of zero, whilst
Pennsylvanian oils were considered to have the least variation and
were defined by an index of 100. All other oils are defined by an
index between these two values, as defined by the formula shown in
fig. 5.2. The particular reference oils are chosen so that they and
the oil under study have the same viscosity at a temperature of l00°C.
Since the introduction of the viscosity index classification many
new oils and additives have been produced so that oils are now
available with much better viscosity /temperature characteristics than
the Pennsylvanian oils, i.e. they have an index greater than 100.
A typical modern engine oil will be found to have a V.I. of 160.
85
Finally mention should be made of some fluids (a term which
includes gases and vapours) whose viscosity actually increases with
temperature. Thus in gases the increased velocity of the molecules
at higher temperatures causes an increase in the momentum transfer
which is the primary cause of viscosity of gases. The effect of
temperature on the viscosity of air at atmospheric pressure is shown
in fig. 5.3.

~
·.;;
0

-~ I ·5 X 10- 2
>

30 60 90 120·
Temp.l°C

Fig. 5.3. The effect of temperature on the viscosity of air.

5.4. Effect of pressure on viscosity


Just as temperature rise reduces the viscosity of a mineral oil, so
an increase in pressure produces a rise in its viscosity. The simplest
expression which represents the effect of pressure is,

YJ = YJo exp (exp)

when Y/ is the current dynamic viscosity at a pressure p whilst Y/o is


the viscosity at atmospheric pressure. The constant ex is referred to
as the pressure exponent of viscosity. The exponential form of this
equation shows that the viscosity rises fairly slowly at low pressures,
but that the rate increases rapidly at the higher pressures. Thus for
many bearing applications the changes in viscosity due to the pressures
produced in the lubricant are negligible. This pressure effect
becomes much more important in so-called Hertzian contacts ( see
Chapter 3) such as between balls or rollers and the races of rolling-
contact bearings, or between gear teeth in contact. In such situations
the pressures can rise to values in the region 109 Pa= 10 4 atm. where
it would be found that the dynamic viscosity of a mineral oil is raised
86
by about one hundred times its value at atmospheric pressure. In a
sense the oil which is a simple fluid at atmospheric pressure begins to
behave almost like a plastic solid at these very high pressures. It is
also worth noting that at such high pressures the oil behaves in a
non-Nevvtonian fashion when the shear stresses are no longer directly
proportional to the shear strain rates.

5.5. The measurement of viscosity


A very wide range of instruments have been designed for the
measurement of viscosity. l.viost of these instruments called visco-
meters are of three main types:

(a) Instruments in which the viscosity is determined by measuring


the rate of flow of the fluid through a capillary tube or through
an orifice.
(b) Instruments in which the viscosity is determined by measuring
the motion of a solid object through the fluid.
(c) Instruments in which the viscosity is determined from the
shearing of the fluid between surfaces in relative motion.

In general instruments of types (a) and (b) are only suitable for
measuring the viscosity of Newtonian fluids, whilst type (c) may be
used for a wide range of fluid characteristics. Since the effects of
temperature and pressure have already been established as of para-
mount importance they must be controlled in all measurements of
viscosity. \Vhilst the effects of temperature on viscosity may be
studied in almost all types of viscometer the effects of pressure may
only be studied in a very few instruments.

5.5.1. Capillary viscometers


These provide a measure of viscosity Y/ by measuring the volume
q flowing through a capillary of diameter d and length l in a time t.
Here Y/ is obtained by applying Poiseuille's formula, where

_ 7Td4t/j.p
Y)- 128lq

/j.pbeing the pressure drop along the capillary. In practice the above
equation has to be slightly modified due to the end effects where the
fluid is suddenly accelerated at entry to and exit from the capillary.
Provided l / d is large, say greater than 500, these end effects are fairly
small corrections.
G 87
This type of viscometer may be used to make ' absolute ' measure-
ments of the viscosity using the Poiseuille formula together with any
' end effect' correction. More often, however, these viscometers
are used as comparative instruments. The instrument is calibrated
using fluids of known viscosity to establish its constants whence it
may then be used to determine the viscosity of any other fluid. A
commonly used viscometer of this type is shown in fig. 5.4. In use
the fluid level is adjusted to mark A and is then drawn up to a point
just above B. The time taken for the fluid level to fall from B to C
is noted and used to determine its viscosity.

-8

-c
Capillary
A- tube

Fig. 5.4. A typical capillary viscometer.

5.5.2. Efflux viscometers


In these instruments the viscosity is determined by measuring the
time for a given volume of liquid to flow through a calibrated short
tube orifice in the base of the instrument. The Redwood, Saybolt
and Engler viscometers are all of this type and are of the form shown
in fig. 5.5. These instruments are purely comparative and define the
viscosity in such units as Saybolt seconds. Fortunately, there are
tables which enable one to convert these arbitrary units into kinematic
viscosity units such as table 5 .1. In these instruments the viscosity
may be determined at any desired constant temperature because of
the surrounding water bath.
88
E E

A - oil container B - thermometer


D - orifice F - stirrer
Fig. S.S. A typical effiux viscometer.

Table 5.1. Viscosity conversion table.

Kinematic Engler Redwood Savbolt


viscosity degrees No. 1 uni~ersal
centistokes seconds seconds

2 1-14 31 32
6 1·48 41 46
10 1·84 52 59
14 2·22 65 74
18 2·65 78 90
25 3·46 105 120
35 4·71 144 164
45 5·99 184 209
60 7·92 245 279
80 10·60 326 371
100 13-2 406 463
150 19·9 620 700
200 26·8 820 940
300 40·0 1230 1410
500 66·0 2040 2320
700 93·0 2820 3250
1000 133 4100 4750
2000 260 8100 9200
3000 400 12 300 14000
5000 660 20 000 23 000

89
5.5.3. Falling-body viscometer
In the simplest type of falling-body viscometers one finds the
viscosity from the terminal velocity of a steel sphere falling under
gravity (fig. 5.6). When the sphere has reached its constant terminal
velocity v it is in equilibrium under the action of gravity, buoyancy
and drag forces, since the buoyancy and drag forces must equal its
weight.
Now weight force is J4 7TY 3 psg,
4
the buoyancy force is 7TY 3p1g
3
and the drag force is 6mirv (Stokes's formula),
where r is the radius of the sphere and p 8 and p1 are the densities of
the steel and the liquid respectively.
From weight= drag+ buoyancy we obtain
2 r 2(ps - p1)g
1)=9 V •

It is essential that the radius of the sphere should be small compared


with that of the vessel containing the fluid. If not, it is necessary to
correct for the proximity of the walls of the containing vessel. This
type of viscometer may be designed to measure viscosity at any
desired temperature and pressure.
Several commercial viscometers based on the same general prin-
ciples as that described above are available but will not be described
further.

Guide tube

Gloss tube

Sphere----+-~
Timing
marks
Test oil---l--

Fig. 5.6. A falling sphere viscometer.


90
5.5.4. Rotational viscometers
In these viscometers there are two elements, one fixed and one
rotating, the space between being filled by fluid. By measuring the
transmitted torque across the fluid film and the speed with which it
is sheared one may readily obtain an instantaneous measure of
viscosity. The torque may be converted to shear stress and the
speed difference to the rate of shearing.
Two basic geometries of this type of viscometer are in common use.
The rotating cylinder form is shown in fig. 5.7 where the inner
rotating cylinder is concentric with the stationary outer cylinder.
The space between the cylinders contains the test fluid and the torque
transmitted across the film is recorded. By varying the speed of the
inner cylinder and/or changing the clearance between the cylinders
the rate of shearing may be varied. Such viscometers are therefore
useful for measuring the viscosity for a range of values of shear
strain rate. This is particularly important with non-Newtonian
fluids where the shear stress and shear strain rate are not in a constant
ratio at all values of strain rate, i.e. the viscosity varies with strain rate.

Clearance
for test oil

Fig. 5.7. A typical rotational Yiscometer.

The other form of rotational viscometer is shown in fig. 5.8. This


is the so-called cone and plate arrangement. Again one element
rotates relative to the other and they are separated by the test fluid.
The conical shape of one element is necessary to ensure a linear
variation of film thickness in the radial direction so that the shear
strain rate is the same throughout the film. The shear strain rate
is u / h where u is the linear relative velocity and h the film thickness
at that point. Since u is a linear function of the radius, h must follow
the same law if u/h is to be constant.
91
Fig. 5.8. A cone and plate viscometer.

5.5.5. Disc machine viscometers


For the measurement of viscosity at very high pressures most of
the common viscometers described earlier are either no use or at the
very least inconvenient, since they require external pressure systems.
The disc machine offers a very simple approach to this particular
problem since the pressures are inherent in the design of the system.
Consider two rotating discs subjected to a load Pas shown in fig. 5.9.
The fluid in the gap between these discs is sheared due to the motion
of the discs and is subjected to very high pressures. That this is so
will be appreciated by recognizing that the contact area of the discs
is only a few square millimetres so that a load P of 100 N leads to

Fig. 5.9. The principle of the disc machine high pressure viscometer.
92
pressures of the fluid in the contact region of the order of 108 Pa (N m- 2 ).
From a measurement of the torque transmitted across the fluid film
and the rate of shearing one may thus readily obtain the viscosity at
very high pressures. Such information is essential where lubricants
are used between gear teeth or between balls and rollers in rolling
bearings. Indeed it will be readily appreciated that the disc machine
itself is useful simply because it represents these types of high-
pressure contact.

5.6. Other properties of lubricants


Although viscosity is the most important property of a lubricant,
other properties must not be neglected. Other important properties
are: density, specific heat, thermal conductivity, acidity or alkalinity,
foaming characteristics, flashpoint (the lowest temperature at which
vapour is produced sufficient for first ignition), pour point (the
temperature at which oils cease to flow freely) and oxidation stability
(since oxidation of lubricants leads to a shortening of their life and
may result in corrosive effects). Test procedures are defined for each
of these properties but detailed discussion is beyond the scope of this
book.
Most modern lubricants have chemical compounds added to
improve the basic characteristics of the base mineral oils. Some
such additives are:
(a) Viscosity improvers
These arc usually high-molecular-weight polymers designed to
reduce the rate of change of viscosity with temperature.
(b) Pour point depressants
When mineral oils are cooled, waxy crystals are precipitated
which combine to form a semi-solid mass. The pour point
depressants, usually complex polymers, form a skin around
such crystals and inhibit their coalescing.
(c) Oxidation inhibitors
These reduce the rate of formation of oxidation products.
\Vhere such products are acidic the oxidation inhibitor also
acts as a corrosion inhibitor.
(d) Detergents or dispersants
These are mainly used in engine oils to hold insoluble material
in suspension, thereby inhibiting the formation of sludge.
A variety of organometallic salts or polymers are used for this
purpose.
(e) Corrosion inhibitors
Alcohols, ethers, organic acids or soaps are used to produce
protective films on the metal surfaces to inhibit corrosion.
93
(f) E.P. additi'l·es
Extreme pressure additives are used to form protective,
thermally stable surface films in situations where metallic
contact might otherwise occur, e.g. when starting up an engine
and before hydrodynamic films of lubricant have been generated.
Compounds of chlorine, phosphorus and sulphur are most
commonly used.
(g) Anti-foam additi'l·es
These are usually silicone compounds designed to mm1m1ze
foaming where this is likely to present difficulties.
(h) Solid lubricants
These are materials like graphite and molybdenum disulphide
which are lubricants in their own right. They are often used
as dispersions in mineral oil and offer protection even when
the base oil is no longer present.
(i) Emulsifiers
In such operations as metal cutting an emulsion of oil and
,vater is often used as a cutting fluid. Emulsifiers such as
petroleum sulphonates and metallic soaps stabilize such
emulsions.

5.7. Greases
Greases usually consist of a thickening agent mixed with a standard
mineral oil, often with other special additives. For certain special
greases the mineral oil may be replaced by other fluids such as silicone
fluids where a high temperature grease is required. Metallic soaps
such as calcium, lithium and sodium soaps are commonly used as
thickeners, whilst clays and silicates are also used for this purpose.
The grease has the characteristic appearance under the microscope
of a tangled web of fibres of the thickener suspended in the base oil
as seen in fig. 5 .10. The inherent stiffness of greases enables them
to be used in a wide variety of practical situations. In ball bearings
they both fulfil the role of lubricant and act as a very effective seal
against the ingress of dirt and moisture. Their stiffness also allows
them to be used in situations where a liquid lubricant would readily
drain away.
As with other lubricants several tests are carried out to define the
characteristics of greases. Perhaps the most important aspect of
greases is that they tend to be altered during operation rather more
than simpler lubricants. In high pressure contacts such as ball
bearings the fibrous structure of the grease is broken down and this
leads to a reduction in lubricating efficiency.
94
Calcium base.

Sodium base.

Fig. 5.10. Some typical grease structures.


95
5.8. Project suggestions
Many forms of viscometer of the types described in the foregoing
chapter may readily be constructed. The falling ball viscometer
offers a useful project activity since such effects as the size of the ball
relative to the diameter of the cylinder, the material of the ball, and
the temperature of the lubricant are all easy to ascertain.
Efflux-type viscometers should be capable of easy construction and
are particularly useful for the study of temperature effects. Cone
and plate and disc machines offer the greatest degree of complexity
in both design and construction, but even these are within the range
of a well equipped workshop and laboratory.

96
CHAPTER6
externally pressurized bearings

6.1. Introduction
Where a fluid film separates two solid surfaces it must be under
pressure. Indeed, in the absence of such a pressure any load applied
to the solids would simply squeeze the fluid out of the gap. Consider
a block of area A subjected to a load W and separated from a plane
by a film of fluid held at a constant pressure p, as in fig. 6.1. The
film can only be sustained provided that
W=pA.
If p is not constant at all points, so that each element SA carries a
load SW,
SW=p SA.
For the total effect
W= Jp dA (6· 1)
where the integral is taken over the total area. For simplicity we
shall restrict our examples to two-dimensional cases where the above
equation would produce the results shown in fig. 6.2.

,,-, Pressure p

Area A

Fig. 6.1. The relationship between load, area and pressure.

~
I ! ! : I i :
'
tt H
! : I
rt t +~ l ~ :!+i
I
I ◄--· L--~:
!
i-.------·-· L----·--+-i

Fig. 6.2. Examples of load and pressure relationships.


97
We now consider some elementary fluid flow arguments. Suppose
a fluid flows along a passage in which the pressure difference is dp
over a short element of length dx (fig. 6.3). The flow is proportional
to dp/dx; the greater the pressure gradient the greater the flow.
One would also expect a greater flow the larger the cross-sectional
area of the passage, but that the flow would be inversely proportional
to the dynamic viscosity.

-q
-,~i,------,:h
p +3p
3x
ii'
t
p
',

Fig. 6.3. The flow of an element of fluid due to pressure.

Fluid flow theory justifies these arguments and yields the following
two results for the cases which are of most interest.
For flow along a uniform pipe of diameter d (fig. 6.4) one obtains
the Poiseuille formula:
q= _ d (dp) = ~
77 4

12817 dx 12817 l
(P2-P1). 6 .2 ) (
For flow between substantially parallel surfaces (fig. 6.4) one obtains
3 3
q= _h dz (dp) = h dz (p 2 -p 1 ) ( 6 _3 )
1217 dx 1217 l '
the negative sign simply indicating that the flow must be in the
direction of decreasing pressure, i.e. if dp /dx is positive the flow

Pz
.e- - - - - - ~P1
q-
( t
d

t I --q

Pz P1
-.e
82
q-
t
h

Fig. 6.4. The flow down a circular tube and between parallel plates.
98
would be in the negative x direction. The similarity of these two
equations and their consistency with our physical arguments is
evident. Equation (6.3) is important in fluid film lubrication and we
shall return to it when we deal with hydrodynamic lubrication.
This type of fluid flow, accompanying a uniform pressure gradient,
is often referred to as Poiseuille flow.
Applying equation (6.3) to the simple example shown in fig. 6.5
we find that the pressure variation is of the form shown since the flow
of fluid in each section must be the same.

. dp 1217q h
1 e -=
. · dx
---=-
h 3 dz h3
(6.4)

where h is a constant.
We immediately see that the pressure drop along the large section
is very small in relation to the other section, and for such large sections
we can therefore assume a constant pressure.

t
q- -q

~ Pressure distribution

Fig. 6.5. The pressures due to flow from a large to a narrow gap.

6.2. The simple pad bearing


Suppose we supply fluid at a pressure Ps to the gap between two
parallel plates subjected to a load W per unit width. We shall for
simplicity assume that the plates are very wide so that there is no flow
in the z direction and the pressure is the same for all values of z at
any x coordinate position. At the edges of the plates the pressure
is taken to be zero, so that from the supply point the pressure falls
linearly to zero as shown in fig. 6.6, i.e. dp/dx = constant from equation
99
(6.4). Consider the load per unit width in the z direction. We
see that
W=2 (Ps;O) l/2x 1 =Psl/2.
If we now increase the load we obviously have to increase Ps to main-
tain equilibrium.
There are two possible solutions to our dilemma. \Ve may use
not a constant pressure but a constant flow supply, or we may use a
constant pressure supply together with a ' compensating ' element.
Although the latter method is of most practical value, we shall first
briefly consider the former solution.

jW pe, ,oit width

~~~~~~~~1!~~~~
-=-=-=-=~-=-=-=-=-=-=-=-=-=-=-
J_ h

Pressure
/ distribution

I-+-----
Fig. 6.6. The pressure distribution in a simple hydrostatic pad.

6.2.1. Constant flow supply


Suppose a pump supplies fluid at a constant flow rate q. Neglecting
flow in the z direction, the flow from the bearing in the x direction
is from the centre to the two edges. Applying equation (6.3) for
parallel-walled channels to one half of the bearing gives (fig. 6.6)
1 . _h 1(Ps.-0)
71q-12r]
3
712 .
X

Thus
S= 3qY]l
P ha
and
W =Psl = 3qYJz2 (6.5)
2 2h 3
100
In this arrangement since q is constant we see that any change in W
is accommodated by a change in the film thickness h, and the supply
pressure Ps now varies to ensure a constant flow rate q. The
stiffness of this bearing is defined as the rate of change of W with
film thickness h.
dW 9(/YJl 2
Thus stiffness - dh
- =+---- 2h 4 •
(6.6)

At this point we may also consider a major improvement in the


bearing design which will significantly increase its load capacity.
Recalling the simple example in which we saw that the pressure drop
due to flow is very small for large passages we can usefully compare
our present arrangement with the recessed design shown in fig. 6. 7.
Since the dimensions of the film thickness in the recess is relatively
large with respect to the film thickness at the lands the pressure in
the recess is substantially constant. The load capacity of such a
bearing is, fig. 6. 7,
W =Psb X 1 + (Ps/2)/ X 1.
Considering the constant flow supply through the lands gives equation
(6.5) as before so that the load capacity is now

W=
3
f;;t
(b+l/2)
and the stiffness is
dW 9q,YJl(l + 2b)
s= - dh =+ 2h 4 • (6.7)

W per unit width

I ~
~~~
1
J. .- -

-I ½1~~-b------4 ½1--
Pressure
distribution

t t
Fig. 6.7. The modification of the pressure distribution due to a recess in a
hydrostatic pad.
101
This is an increased load capacity as can be seen by comparing
figs. 6.6 and 6.7 and this type of geometry is that most commonly
employed in externally pressurized bearing designs.
We appear to have satisfied the design requirements for this type
of beating, so why should we consider any alternative? The answer
is that the use of a constant flow type of pump implies that every
bearing would require its own separate pump, a very expensive
arrangement. The advantage of the other alternative, a constant
pressure pump, is that the same pump may supply a whole series of
bearings provided that its flow capacity can meet their total require-
ments.

W Bearing width win

l the z direction

~
/. ~,%·,.,. ,. ,.-~1/~~~~.~
----------------·

t
le
t

Fig. 6.8. The capillary compensated hydrostatic pad.

6.2.2. Constant pressure supply


We have seen that a direct connection of a constant pressure supply
to the bearing is no use since we would have to adjust the supply
pressure to accommodate changes in load. But if we connect such
a supply to our bearing through a compensating element we can
achieve our object. The simplest compensating element is a length
of capillary tubing whose flow and pressure characteristics are governed
by equation (6.2). Consider the system shown in fig. 6.8, where a
constant supply pressure Ps drops to a pressure Pr at the bearing
supply point. From the flow through a tube we note, from equation
(6.2), that

q= 1;~: ( Ps -;/!!) . (6.8)

102
The linear pressure drop along each half of the bearing ( through
which flows half of the fluid) is governed by:
<j _ h 3w Pr
(6.9)
2- 1271 l/2'
where w is the width of the bearing m the z direction. Thus
eliminating q one obtains
_ Ps
Pr-1 +ha/k
where
31rd4l
k= 128/cw ·
The load carried is then given by
W=Prlw
2
Pslw
=~--- 3
2(1 +h /k)
and the stiffness is
dW 3 (
s= - dh = 2k 1 +h 3 /k
h )2 p slw. (6.10)

This equation relates the load to the film thickness for a given
geometry of bearing. The flow through such a bearing is then found
from equations (6.9) and (6.10),
. k(pslw-2W) (6.ll)
q= 371[2 •

Thus although the supply pressure is constant, changes in load cause


a change in the film thickness and in the flow rate. By examining
equation (6-8) we see how the compensating capillary element fulfils
our requirements. For a constant Ps any increase in load necessitates
an increase in Pr which is accommodated by a reduction in flow rate.
Equation (6.8) is a kind of hydraulic Ohm's law, where the pressure
drop (voltage) is linearly related to the flow rate (current), the constant
of proportionality being the ' hydraulic resistance '.
A more useful geometry in practice is to have a large central recess
together with narrow lands (as shown in fig. 6.9, where the com-
pensating element is represented as a resistance). The load capacity
for such a bearing then becomes
W =Ps(b + l/2)w (6.12)
1 +h 3 /k .
II 103
The stiffness of the bearing is then
dW 3( h
s=- dh =+k l+ha/k Ps(b+l/Z)w.
)2 (6.13)

In a similar fashion the flow is calculated as


. _ k[Ps(b + l/Z)w- W]
(6.14)
q- 3YJl(b+l/2) .
Equations (6.12) and (6.14) reduce to equations (6.10) and (6.11)
·when b is zero, as we ,vould expect.

Bearing width w in
the z direction

Fig. 6.9. The compensated hydrostatic pad with a recess.

For more complicated cases in which oil-flow in the z direction


occurs, or with more complex geometries, the basic behaviour is still
similar to that indicated by these equations.
Any form of hydraulic resistance which produces a relationship
between the oil flow and the pressure drop is acceptable as the
compensating clement. One popular alternative to the capillary
compensator is to use an orifice in the supply line. The formula
for the flow through such a constriction is

• _ 1T C d2 ( Z(Ps -Pr))
q- - cl (6.15)
4 p
where p is the fluid density, d is the orifice diameter and Cc1 is the
discharge coefficient of the orifice. When such a compensator is used
the preceding analysis must be reworked starting from equation (6.15)
instead of equation (6.8), and this leads to rather more complicated
expressions for the load capacity, stiffness and oil flow.
104
6.3. The characteristics of capillary-compensated bearings
An examination of the physical meaning of equations (6.12),
(6.13) and (6.14) is rewarding since they indicate the many interesting
features of this type of bearing.
(a) For a given geometry of bearing we note from equation (6.12)
that the load capacity depends on the operational film thickness
and the form of the relationship is shown in fig. 6.13 a. We
see that the load capacity rapidly increases with reducing film
thickness.
(a)
w

(bl
s
I

~
h
(cl
q

h
Fig. 6.10. The load, stiffness and flow rate in hydrostatic pads related to the
film thickness.

(b) For the same bearing we find that equation (6.13) for the
bearing stiffness has the form shown in fig. 6.10 b. Here we
notice that there is an optimum film thickness which provides
the maximum stiffness for the bearing, although the load
capacity is smaller than it would be for thinner films. This is
important since one of the main attractions of this type of
bearing is its high stiffness and the ability of the designer to
optimize this characteristic.
105
(c) Substituting for Win the flow equation shows that the oil flow
increases with increasing film thickness in the manner shown
in fig. 6.10 c.
(d) We note in passing that the load capacity, the stiffness and the
flow rate are all increased by increased supply pressure.
(e) Also, the load capacity depends very little on YJ, whereas the
oil flow depends on 1 / YJ· So this type of bearing, being less
dependent on viscosity than hydrodynamic bearings, may
operate using a wide range of fluids.
(f) For a bearing of the same overall dimensions changing the
geometry to make b large and l smaller increases both the load
capacity and the stiffness and slightly reduces the oil flow.

w
(a)

t
(b)

I
--------~
J _ _ _ _ _ _ _ ..J

I
/ ..... - .....
I
I ' \
- ___J . ..!.... - 0 - · I-, I_,
I I / I
! \....... _ ..,/ !
I

I
Drain
channel Supply
t
Fig. 6.11. Single and multi-pad hydrostatic journal bearings.
106
6.4. Hydrostatic journal bearings
So far we have considered the hydrostatic bearing in the form of
plane pads, but the same system may be applied to journal bearings.
Fig. 6.11 shows a single-pad and a multi-pad arrangement for such a
bearing. Each pad is supplied separately from a constant pressure
supply via appropriate compensating elements. The oil then flov,·:;:
over the lands into the drainage channels and the calculations of the
preceding section apply. In fig. 6.11 a the load is carried by the
single pad, and this must therefore be aligned in the direction of the
load. In fig. 6.11 b the sequence of pads allows such a bearing to
accommodate a load in almost any direction and is useful when the
direction of the load may change during operation.

t
R
____ Velocity
distribution

1
h' not to scale

Fig. 6.12. The velocity distribution in the oil film.

\Ve have referred to hydrostatic bearings of various types but as


yet we have not introduced the question of relative motion. In all
bearings one part must move relative to the other, and we would
expect such motion to meet some frictional resistance. Friction
arises from shear stresses in the lubricant film caused by the relative
motion. Consider the film between the lands and the shaft in the
single pad journal bearing in which the shaft rotates at an angular
velocity w; the velocity distribution throughout the oil in such a pad
is as shown in fig. 6.12. For a Newtonian fluid the shear stress T
is given by:
du
'T = 11 cir
where 17 is the dynamic viscosity and du/dr is the constant gradient
of the radial velocity distributions shown in fig. 6.12. Since the
recess depth is probably one hundred times the film thickness at the
107
lands, du/dr may be large across the lands but negligible across the
recess. So the friction is mainly due to the shear stresses at the
lands, and these are
wR
T= Y) h (6.16)

The frictional force F is then the integral of these shear stresses taken
over the operational area. Thus for a unit length of shaft,
F=TlX 1
wR
=YJhl (6.17)

The frictional torque is FR for this simple case. For the plane pad
type of bearing a relative sliding velocity U gives rise to a frictional
force YJUl/h for each unit width of the pad in the z direction. For
such a system the coefficient of friction would then be F divided by W.
We can see the major advantage of these bearings. The friction is
directly proportional to the sliding velocity, being zero when the
bodies are at rest, and still relatively small at fairly high velocities.

Drain
channel
Pressure
di sir i bution
Fig. 6.13. The multi-pad and the multi-recess hydrostatic journal bearing.

In a single pad bearing, one other effect arises because of relative


motion. The velocity of the fluid across the film thickness varies
linearly from zero at the land surface to U = wR at the shaft surface,
so it has a mean velocity of U /2. Since for unit axial shaft length
the cross-sectional area of the film is h x 1, the flow of fluid in the
direction of motion is Uh/2. In the static cases considered earlier
108
we assumed that the rtow across each land was <j_ /2 from the recess
to the outside. The effect of motion is therefore to modify this
symmetrical flow condition so that <j_ /2 + Uh/2 flows out of the leading
land and <j_ /2 - Uh /2 flows out of the trailing land.
The nett flow from the bearing is still <j_, the sum of the above two
outflows. Nevertheless, the difference in flow across the two lands
clearly affects the simple arguments already considered in the static
case. No further analysis will be pursued, but it will be noted that

(a)
!Applied load

Supply
pressure

/Tapered
f i I rn

Fig. 6.14. Apparatus for the study of hydrostatic pads.

109
if Uh /2 = q/2 no flow occurs out of the trailing land and if Uh /2 > q/2
the fluid tends to flow into rather than out of the trailing land.
We should mention one final point which arises from motion.
Under certain circumstances pressures may be generated in the fluid
films by hydrodynamic effects. Such pressures will be additional to
those due to the external pressurizing system so that the load capacity
for bearings in motion may be greater than that predicted above.
Examination of the pressure distribution around a multi-pad
bearing in fig. 6.13 a shows that in the drainage channels the pressure
is zero, which means ambient atmospheric pressure, so if one
eliminates the drainage channels one increases the load capacity.
This leads on to the so-called multi-recess design of externally
pressurized journal bearings shown in fig. 6.13 b. The fluid from
one recess now flows across the land into the adjacent recess and in
this way increases the load capacity.

6.5. Project suggestions


The apparatus shown in fig. 6.14 a offers a wide range of studies
for static hydrostatic pads. The central cylinder is maintained in a
vertical plane with minimum friction by a simple form of air bearing.
Compressed air is led to a series of small holes drilled around the
periphery of the containing tube at two different height levels. The
hydrostatic pad at the base of this cylinder is then supplied through
an appropriate capillary restrictor with fluid from a pressurized
supply. This type of bearing may then be studied to relate film
thickness, measured by a capacitance method, to the applied load.
The effects of supply pressure and restrictor length may then be
investigated for various geometries of pad defined by the ratio R 1 /R 2 .
Since the cylinder is supported vertically the effects of taper on the
hydrostatic pad may also be studied as shown in fig. 6.14 b.
The effects of motion and the pressure distribution may be studied
by replacing the hydrodynamic tilted pad by a hydrostatic pad,
restrictor and oil supply in the apparatus shown in fig. 7.21 and
described in section 7.5.

110
CHAPTER 7
hydrodynamic lubrication

7 .1. Introduction
WE have already established that if two loaded solids are to be
separated by a continuous fluid film such a film must be pressurized.
The preceding chapter gave examples of externally pressurized
bearings, hydrostatic bearings as they are often called. Now we shall
consider situations where the pressure is produced by virtue of the
viscosity of the fluid and the relative motion of the two solids. Such
bearings are often called self-acting bearings, and the lubrication
method is referred to as hydrodynamic lubrication.

.
t St t nary
ed pad
Direction of
pressure build-up~- _____ _
_-:.._-_-_-_-____ -- ---
_-_-_ - - - - - - - - - -_-__
-----u
Fig. 7.1. The convergent oil film.

The moving surfaces must converge geometrically-that is the two


surfaces are non-parallel, as shown in fig. 7.1. The viscous fluid is
being squeezed into a converging space as it is dragged along by the
motion of the solids, and pressure is therefore generated in the film.
This pressure enables the film to withstand the applied loading.
Even without mathematical analysis, we can sense that the load
capacity (that is, the pressure) will be greater at higher speeds and
with higher values of viscosity.
We must now ask how we can create the geometrical conditions.
The most obvious solution is to use tapered pads for either linear or
journal bearings as shown in fig. 7.2. The degree of taper of the oil
films is relatively modest, the main requirement being that some
tapering exists. Typically we therefore think of a change in film
thickness of say 40 µm per cm length, with film thicknesses of about
400 µm. In all the diagrams of films and tapers, the scales have had
to be greatly distorted and magnified to facilitate explanation.
111
----u

Fig. 7.2. Convergent oil films in slider and partial journal bearings.

Using this distorted diagram, let us now consider a plain journal


bearing. If the shaft is to fit into the bearing it must be smaller in
diameter, but again only by something in the region of 100 µm or
less. With the centre of the shaft coinciding with the centre of the
bearing, the clearance (the film thickness) is uniform around the
bearing. Such a situation, figure 7.3 a, creates no hydrodynamic
pressure and therefore has no load-carrying capacity, and the film
would tend to collapse due to the applied load. In fact what actually
happens is that the shaft runs in the position shown in fig. 7.3 b; the
film thickness varies around the circumference as shown by the
developed diagram, fig. 7.3 c. In such a situation the film converges
towards the point of closest approach so that the conditions for the
generation of hydrodynamic pressure and thereby load capacity arc
created. Thus in practice with journal bearings the shaft rotates
about a centre which is eccentric to the centre of the bearing, and both
the degree of convergence and the thickness of the hydrodynamic
film are such as to created a pressure sufficient to support the load.
The mathematical relationships for such bearings will be discussed
later.
Bearings of this type are among those most extensively employed,
largely because of the simplicity of their construction. Thus a shaft
rotating inside a hole and supplied with lubricant would appear to
make a perfectly satisfactory bearing for many applications in cars,
gear boxes, pumps, etc. Unfortunately, all such bearings arc
subjected to many starting and stopping cycles during their lifetime.
During these periods the sliding speed of the shaft relative to the
housing may be very small so that some metallic contact occurs as
112
the surface asperities break through the very thin fluid films formed
at low speeds. For this reason, rather than because of any fluid film
considerations, the actual construction of the bearings is rather more
complicated. Soft metal constituents have to be used to prevent
unduly high friction and wear during the periods before full fluid
films are generated. Once the films are created the two surfaces are
completely separated and the friction only depends on the shearing
characteristics of the fluid.

Fig. 7.3. The oil film geometry in a full journal bearing.

7.2. Basic theoretical considerations


7.2.1. Assumptions
We shall first simplify our problem by making the following
assumptions.
(a) We assume that the fluid is incompressible. This is reasonably
true for most liquids and particularly oils, but is clearly not
true for fluids such as air, gases and vapours, although as we
shall see later these can still be satisfactory lubricants.
(b) We assume that the fluid is Newtonian, with the shear stress
proportional to the shear strain rate. Some fluids, and indeed
almost all fluids at very high pressures, depart from this
Newtonian behaviour, but fortunately for mineral oils acting
as lubricants within bearings this assumption is reasonably valid.
113
(c) It will be assumed that the fluid properties, particularly the
viscosity, are independent of pressure and temperature. This
is not generally true, as we saw in Chapter 5, but for many
cases we can obtain reasonable results by assuming the fluid
properties have a characteristic constant value at the particular
operating values of temperature and pressure.
(d) We neglect any effects due to the inertia of the fluid and assume
the fluid flow to be laminar, i.e. turbulence in the fluid is
neglected, although at very high speeds this is no longer an
acceptable assumption.
(e) The solid bodies constituting the bearing surfaces are assumed
to be rigid. Again at very high pressures such as those existing
between the balls and their races in ball bearings this assumption
is no longer valid.
(f) Although we shall consider variations in pressure along the
bearing it will be assumed that at any point the pressure is
constant across the thickness of the fluid film.
(g) We assume that the bearings are infinitely wide in the z direction
so that we may neglect any fluid flow in this direction; sec
fig. 7.4.
In view of the above assumptions it may at first appear that any
results based on such a model will be of little value. In practice,
they give fairly accurate predictions of observed behaviour over a
wide range of bearing designs.

W per unit width of z

Fig. 7.4. The geometry of the inclined pad slider bearing.

7.2.2. Reynolds' equation


It is reasonable to expect that there will be relationships between
the build-up of pressure, the sliding speed, the operational viscosity
and the geometry of the hydrodynamic film. We shall now derive
114
the basic relationship by considering a simple taper geometry of the
type shown in fig. 7.5 a. Since the fluid that is in contact with the
moving surface moves with the velocity U of the surface, we shall be
subjecting the film to a continuous shearing so that the velocity 11 8 of
the film at any value of y is given by

Us= U(Jz y)_ i (7.1)

(a)
~ h
T X y 0

~L U t

(cl
=O dp.
- 1s- ve
dx
------ Pressure
I distribution

x=O x=L

(d)

up distribution up is - ve up is 0
I

~
(el
1/ .,/-'

==
Velocity
distribution
hm ~
U = Us+ Up
u u u

Fig. 7.5. The Yelocity and pressure distribution in an inclined pad bearing.

115
The form of Us is shown in fig. 7.5 b; Us is + U at y = 0 and zero at
y = h. This is known as Couette flow.
With converging walls the relationship between the fluid flow and
the pressure build-up (Poiseuille flow) is

U p =1- ( - dp)
- y(h-y). (7.2)
271 dx
We must know how the values of dp/dx vary with x, and shall assume
that this variation is as shown in fig. 7.5 c. The form of up is then
as in fig. 7.5 d and Up is zero at y = 0 and y = h. Furthermore, at the
entry section dp /dx will be positive, near the centre it will be zero,
and after this point it will be negative. This gives rise to the patterns
of up shown in fig. 7.5 d.
Since the pressure gradient and the shear flow are the only two
causes of fluid flow, the resultant velocity of the fluid follows the
patterns shown in fig. 7.5 e and is given by
_
ll-Us+Up-
_ GT(h-y)
-- + -1 ( - - dp) y(h-y). (7.3)
h 271 dx
As we are neglecting any side flow, the area of an element of film is
1 x dy per unit width of the film. The quantity flowing per unit
time is thus 11 <ly for each such elemental area and the total flow q is
given by:
+h
q= J 11dy.
0

Substituting for u from equation (7.3) and integrating yields

. (2U) + 1271
q= h
h
3
(
- dp)
dx • (7.4)

At some point the pressure is a maximum and dp/dx=O. Let the


value of h at this point be hm whence

qm=tZm
1 u
Z. (7.5)

But the flow through the film must be the same at all values of h so
that q111 = q and thus

. (2u) =h (2u) + 1271


q=hm 1z - dp)
dx .
3
( (7.6)

Rearranging terms we obtain the relationship we have been seeking:


dpdx = 12 71 ( U)
2
h-hm.
h 3
(7.7)

116
This equation was first derived by Osborne Reynolds and is known
as Reynolds' equation. It shows why geometric convergence is
essential. If h did not vary with x, then h would equal hrn and dp /dx
would be zero at all values of x so that no hydrodynamic pressure
would ever be produced. In the case considered ,vc can also sec that
h at any value of x is defined by the equation,

Il= Il j - -hi - ho
L--X.

Equation 7.7. is Reynolds' equation in its simplest form but the


physical arguments used for its derivation apply to a very wide range
of situations. Thus if we have two surfaces with velocities U1 and
U2 , as in fig. 7.6, Reynolds' equation may be stated as

dp = 1211 ( U1 + U 2 ) h-hm. (7 .S)


dx 2 Jz3
In using this rather more general form of Reynolds' equation we must
always impose the restriction that the film thickness at any.'\: coordinate
must not change with time. A couple of examples will illustrate the
use of this form of Reynolds' equation and will also confirm our
earlier derivation.

Fig. 7.6. The scale of roughness and film thickness in hydrodynamic


lubrication.

Consider the two convergent surfaces shown in fig. 7.7 a. Evidently


the value of h at any given x value does vary with time. By resolving
this system into its two component parts we can make h invariant
with time for one of the components whilst the second component is
simply a translation of the whole system with uniform velocity UtJ•
This second component does not generate any pressure so we apply
equation (7 .8) only to the first component with U1 = U a - llti and
U2 =0. It will also be seen that the first component is identical to
our original simplified problem with U = U a - UIJ, so that for the
case considered Reynolds' equation becomes

dp. = 121 ( Ua- UtJ) h-hm. (7.9)


dx 1 2 Jz 3
117
As a second example consider two discs rotating about fixed centres
with a lubricant film separating the two discs, fig. 7.7 b. For this
case we can see that h does not vary with time at any given value of x
so that equation (7.8) applies directly with U1 = U a and U2 = [Jt;

dp = l Z ( U a+ U
dx Y/ 2
b) h -h hm
3 •
(7.10)

It is interesting to note that in fig. 7.7 a there will only be a pressure


generated when Ua=/= Ub, whereas in fig. 7.7 b a pressure is generated
even when U a= Ub, i.e. for conditions of pure rolling. A special
case of fig. 7.7 b occurs when a cylinder rolls along a plane with some
degree of slip, as might occur in a driving or braked car wheel rolling
along a wet road. This situation is shown in fig. 7.7 c which has been
resolved into the velocity components shown. Again we see that
equation (7.8) applies to the first of these components whilst the
second component simply represents a uniform motion of the whole
system.

~ _____._LJb

~
(a)
~
• hf - +
X ~-+-Ua X -----.- Ua - Ub ---.- Ub
h invariant

(b)

+ +
U0/
/

h invariant
(cl

~ + +-+-Ub

Fig. 7.7. Some examples of the resolution of the velocity components for the
application of Reynolds' equation.
118
7.2.3. Practical examples
We are now in a position to identify a number of practically
important situations in which pressures may be developed by hydro-
dynamic action. The first simple example is a fixed tapered pad
(fig. 7.2) which can be used in a simple linear bearing, in a thrust
bearing, or in a journal bearing. Such pads, instead of being fixed,
may be mounted on pivots, springs or elastic supports, as shown in
fig. 7.8. These flexible mountings are used to enable the pads to
take up that inclination which produces the best solution to the
particular operational conditions of load, speed and viscosity. This
is a classical example of the principle that generally ' Nature knows
best ', so that if systems are flexible they tend to take up the optimum
operating position.

~ ~ ~
u u u
Pivoted pad Elastic support Spring mounted

Fig. 7.8. Typical examples of non-rigid pad mountings.

With journal bearings we have already seen that if the shaft runs
about a centre which is eccentric to the bearing centre the necessary
convergent film is obtained. Thus in fig. 7.9 a a partial journal
bearing is essentially the same as a curved pad bearing whilst the full
journal bearing is shown in fig. 7.9 b together with the associated
hydrodynamic pressure distributions.
w w

Pressure Pressure
distribution distribution
(a) (b)
Fig. 7.9. The pressure distribution in a partial and a full hydrodynamic
journal bearing.
119
For balls and cylinders rolling around the races of rolling-contact
bearings we are dealing with hydrodynamic films generated as shown
in fig. 7.7 c. A rather interesting example of the same kind of hydro-
dynamic film generation occurs between mating gear teeth, which as
shown in fig. 7.10) may be considered to be instantaneously rolling
and sliding one upon the other. Thus at any instant they are entirely
equivalent to two cylinders rolling and sliding one upon the other,
that is, to two cylinders rolling together with different peripheral
speeds U1 and U 2 . Such a system is then able to generate hydro-
dynamic pressure, and can operate with a fluid film under load,
according to the arguments discussed in fig. 7.7 busing equation (7.10).

Fig. 7.10. The hydrodynamic film between gear teeth and the equivalent
system.

Other practical examples of hydrodynamic films are the films of


water which might be formed between tyres and roads according to
the argument of fig. 7.7 c and equation (7.10). This undesirable
phenomenon leads to lack of vehicle control and is called ' aqua-
planing '. The analysis of such a situation must take into account
the flexibility of the tyre since it is clear that this must modify our
assumption (e). We shall return to this problem later when we shall
find that even such fairly rigid materials as steel when used in gears
and ball bearings cannot be accepted as entirely rigid.
Many examples of the formation of hydrodynamic films occur in
metal-working operations. Thus in rolling forming operations the
conditions of fig. 7.7 c are likely to lead to the formation of hydro-
dynamic films as in fig. 7.11 a. In extrusion processes the formation
of such films between the \Norkpiece and the dies is also possible
(fig. 7.11 b).
We can see that the formation of hydrodynamic films is a very
widespread effect. The simple derivation of Reynolds' equation has
indicated the way in which such films are created, since it establishes
the relation between the pressure build-up and the viscosity, the
speeds and the film geometries. Let us therefore go on to show
how, once Reynolds' equation has been set out for a particular
120
(a) (b)

E
-- a::
0
Material --+-

Fig. 7 .11. Examples of hydrodynamic films in metal working processes such


as rolling and extrusion.

situation, we may obtain other information about the situation such


as the load which may be carried and the friction which results from
shearing of the fluid film.

7.2.4. Load capacity


The load to be carried is simply the integral of the pressure over
the area of the bearing. As yet, we have not even derived the
pressure distribution, although for the simple case considered in
fig. 7.12 we know that

dp = 12 ( U) h - h111
dx YJ 2 1z3
and

where
n=---
hi l
ho
provided x is measured to the left from the ongm O; sec fig. 7 .12.
This choice of origin simplifies the ultimate solution. We obtain by
integration

P= _ 6ri~[- L/n + hm L/n O


+A]. (7.11)
1z 0 ~ 1 +nx/L 2h 0 (1 +nx/L)-
121
In a practical case we would know Y/, U, h 0 , n and L but we are still
not able to show the way in which p varies with x because there are
two unknowns, the integration constant A, and the value hm at which
dp/dx=O because the pressure is a maximum.

_ _ ____.._u
-+--- L ------
x....__ __..,0

~--
.,",,I.,..,. ']\ Pressure
/I .
,/r t l
1--------distribution

\
x=L x~O

Fig. 7.12. The pressure distribution along an inclined pad bearing.

Fortunately, we still have two pieces of information as yet unused.


The supply pressure at the inlet and the outlet of the bearing will be
that of the atmosphere so that p = 0 when x = L and p = 0 when
x = 0. Substitution of these values in equation (7.11) then gives hm
and A as

hm=2h 0 ( 21 +n)
+n

and (7.12)

A= L
n(2+n)
Substituting these back into equation (7 .11) gives the form of the
pressure distribution for this case as

6riUL [ nx/L(l -x/L) ] (7.13)


P= hT (2+n)(l +nx/L) 2 •

122
The form of this pressure distribution is shown in fig. 7.12 and we
then define the load W per unit width of the bearing by
x=L
W= J p dx
x=O

which on substitution for p from equation (7.13) and integrating


yields
(7.14)

where
K=loge(l+n)_ 2 (7.15)
n2 n(2+n)
Before we consider the physical meaning of these results let us
recapitulate the steps ,,·hich must always be taken in dealing with
any hydrodynamic lubrication problem.
Step 1. State the correct form of Reynolds' equation for the case
being considered.
Step 2. Define the ,vay in which the film geometry changes with x
for the particular contact and substitute for h in the
Reynolds' equation.
Step 3. Integrate the Reynolds' equation to obtain the pressure
distribution as a function of x. Since we shall always have
two unknowns such as h111 and the constant of integration,
we introduce two appropriate boundary conditions.
Step 4. Decide which are the most appropriate boundary conditions
and use these to find the unknowns h111 and A, and so
completely define the pressure distribution in terms of x
and the constants of the system.
Step 5. Obtain the load appropriate to the derived pressure distri-
bution by integrating the pressure over the area over which
it acts.
\Ve now return to the important physical aspects of the results
obtained for our simple slider bearing, since these will have a general
validity in hydrodynamic bearings of many types.
The effect of the degree of inclination of the slider is totally defined
by the value of K, since this itself depends on n = lzi(h 0 - 1. The
values of K for various values of n are given below:

n 0·1 0·5 1·0 1·2 1·5


K 0·02 0·022 0·0264 0·0267, 0·0261

123
Changes of more than ten times in the inclination factor n cause
very little change in K, and therefore the load which is carried for
given operating conditions. This confirms the earlier statement that
the main requirement is that a convergent film shall exist rather than
that the degree of convergence has to be very large.
Secondly we note that by rearranging equation (7.14) the minimum
film thickness h 0 is defined by

(7.17)

Thus for a given inclination factor n and operational variables Y/, U


and W the minimum film thickness increases with length L of the
convergent film. \Ve should also note the appearance of the group
YJU/W which appears in all hydrodynamic situations. This is a
dimensionless group as is (h 0 /L) and we shall say more about this in
due course. Equation (7 .17) is the equation which we should use to
design a bearing of this type. We are interested in carrying a certain
load W at some sliding speed U, and to ensure that the minimum
film thickness h 0 is sufficient we then choose values of L and K (i.e. n)
which, together with an oil of appropriate viscosity Y/, satisfies equation
(7.17). In choosing h 0 we must remember that metal surfaces are
not perfectly smooth but arc covered with asperities. We therefore
design for a value of h 0 which is such that these surface asperities are
completely separated, as in fig. 7.13, where we note that h 0 must
always be greater than the sum of the roughness on the two surfaces.

Fig. 7.13. The scale of the roughness to the mm1mum film thickness 111
hydrodynamic lubrication.

7 .2.5. Friction in hydrodynamic bearings


So far we have seen how we can relate load, speed, viscosity,
geometry and film thickness in hydrodynamic bearings but ,ve have
not considered friction. Lubrication with continuous hydrodynamic
films will cause a marked reduction in the coefficient of friction to
values of about 0·001 but it does not reduce it to zero. Friction will
of course lead to a heat release and therefore to a rise in temperature
of the lubricant which in turn heats the bearing. The temperature
rise will reduce the viscosity of the lubricant and so the designer must
124
evaluate the friction so that he may decide the actual viscosity needed
for the lubricant in the bearing.
\Ve shall again consider the simple slider bearing since this produces
results typical of the behaviour of all hydrodynamic films. \Ve use
the Newtonian fluid equation
T= Y/ du/dy, (7.18)
where T is the shear stress at the solid surface when 17 is the viscosity,
and du/dy is the rate of shearing of the fluid at the surface. We
note that u varies with y in the manner shown in fig. 7 .14, and that
u is defined by
U=U ,( r) +-217
1-"-
h
1 ( --
dp) y(h-y),
dx
see equation (7.3). Thus
du=_ C _ _!_ (dp) (lz- 2y).
dy 1z 217 dx
du
At y = 0 we therefore define - as
dy

(~~ t~o =- f - 2\ (ix) h.


We already know dp/d,v from equation (7.7) using lzm as defined by
(7 .19)

equation (7.12). Furthermore we know that

h=h 0 (l+ nx)


L
and

u u u
----u
du) < U
(dy du)
(dy U (du) > U
y=O hm y=O = hm dy y=O hm

~-------- L -----------~
Fig. 7.14. The Yelocity distribution in an inclined pad bearing.
125
Putting these equations together therefore defines (du/dy)v,~o at all
values of x along the lower surface. The frictional force at the lower
surface is then easily obtained, since it is simply given by

F= Jo dx= oJ17 (du)


-r
dy v~o
dx. (7.20)

We may now substitute all the appropriate values from the preceding
equations into equation (7 .20), carry out the integration and obtain
(7.21)
where
4 loge (1 +n) __6_)
(
Kr= n 2+n
[
6loge(1+n)_ 12 ]1;2
n2 n(2+n)
The coefficient of friction µ, = F / W for this bearing then becomes

µ,= : = ( t)1;2 Kr. (7.22)

The coefficient of friction for the bearing depends on the dimension-


less group (1/U/W) and the constant Kr which is wholly dependent
on the inclination factor n. Comparing equations (7.17) and (7.22)
we note that for a given geometrical factor n, the minimum film
thickness and the coefficient of friction both increase with increases
in the dimensionless group ( 17 U / W). This is a feature which is
common to all hydrodynamic films and once again illustrates the
importance of this particular dimensionless group.

7.2.6. Hydrodynamic squeeze films


\Ve have examined the behaviour of fluid films when the surfaces
which they separate are moving tangentially to each other. It is
also of interest to consider what happens when the two surfaces have
normal velocities V1 and V2 as shown in fig. 7.15 a. Physical insight
suggests that when V1 - V2 is positive, so that the film of lubricant
is getting thinner, some additional pressure will be developed in the
film. This is to say that as surfaces approach each other they will
tend to increase the load-carrying capacity of the fluid film.
Consider the case shown in fig. 7.15 a, where we separate the
motion into one component decreasing the film thickness provided
V1 - V 2 is positive, and one component resulting in a motion of the
whole system, which does not affect the film thickness. Therefore
126
the basic problem is that shown in fig. 7.15 b, and it is obvious that
as the film thickness decreases a pressure will be developed and
fluid will flow to the left and to the right from the centre of the pad,
the coordinate position Xm,
Recalling equation (6.3) in Chapter 6 for the flow between parallel
plates and adopting the value of dz as unity to simplify the argument,
we may write

(7.23)

(a)

t
2

(b) '/////////##///////##/&

Direction of flow -
t h _____._ Direction of flow

o--~t____ x
(cl

Fig. 7.15. The hydrodynamic pressure arising from the squeeze film effect.

The fluid flow will be to the left when x < Xrn, where dp /dx will be
positive; and to the right when x > Xrn, where dp /dx will be negative.
Thus at x = Xrn we are not surprised to find dp / dx = 0, so that Xrn is
the point of maximum pressure.
By considering a small reduction in film thickness dh in a time dt it
is easily seen that the outward flow rate at a given value of xis given by

(Jn=(x-xrn) ~;=(x-xm)(V1 -V2 ). (7.24)

127
This means that the flow rate increases from zero at Xrn to a
maximum value at each end of the pad. Combining equations (7.23)
and (7.24) produces the expression for the pressure development:
dp_12 (V: V)X-Xm (7.25)
dx - Y/ z- i ~ -
The term V2 - V1 occurs since dp /dx must be positive when x < Xm.
Integration of this equation using the appropriate boundary
conditions of p = 0 at the edges of the pad then produces the pressure
distribution of the type shown in fig. 7.15 c. The summation of this
pressure over the area over which it acts then gives the additional
load capacity due to squeeze film effects.
Examination of equation (7.25) shows a remarkable similarity of
form to our original Reynolds' equation (7.8) and ,ve may combine
these two equations to give
dp =l 2
dx Y/
(U +U
2
1 )h-hm 12 (V-V)X-·Xm
2
Jz3 + r1 Ji3 . z i
(7.2 6)

In equation (7.26) the first term defines the pressure development


due to the tangential velocity components whilst the second term
defines the effect of any normal velocity components. If we apply

vbk tvb
~ --ub

+
0 X 0 0
Ua~va --ua It Va
I t
-ub
~
No pressure No pressure
effect effect
0 0
--ub t Vb
+ +
~di::~ L»',w'~~

0-----
-ua-Ub
0------'-
tVa-Vb

~=1277(Ua-Ub)h-hm +1277(Vb-v) X-Xm


dx 2 h3 ° h3
Fig. 7.16. The method of formulating Reynolds' equation for a complex
velocity system.
128
this equation to a problem of the type shown in fig. 7.16 we may
resolve the original problem into its various component parts, which
leads to the substitution of U a - Ub for U1 + U2 and Vb - Va for
V 2 - V1 in equation (7.26).
Having set up the appropriate equation for the pressure distribution
we may use it to obtain values of such quantities as the minimum film
thickness, load capacity and friction, by the methods discussed in the
earlier part of this chapter.

7.3. Gas-lubricated bearings


It has already been mentioned that any fluid is suitable as a lubri-
cant, and gases and vapours are fluids which may often be used for
this purpose. Indeed since air is free there are obvious advantages
in using this fluid in preference to relatively expensive oils in those
situations where it is appropriate. Air bearings were suggested as
long ago as 1854 and were actually demonstrated in 1897, but it is
only during recent years that they have been extensively employed.
The modern dentist's high-speed drill is one application of ·which
most of us have personal experience, but these bearings also do useful
service in machine tool spindles, gyroscopes, turbomachinery and
many other applications.
The principles of operation are identical to those already discussed,
and pressure is generated by drawing air into a convergent clearance.
Since air is considerably less viscous than oil the pressures generated
are less. This means that the load capacity of air bearings is less
than that of their oil counterparts, but because of this low viscosity
the friction is also decreased. Thus the coefficient of friction of such
bearings is of the same order as that of liquid-lubricated bearings.
Strictly speaking, these bearings should be called aerodynamic
rather than hydrodynamic and this term is often used. \Ve might
summarize the use of such bearings by the following observations.
Ad,vantages
1. Very low friction.
2. The lubricant is' free'.
3. The lubricant does not cause contamination.
4. The lubricant 1s operational from very low to very high tem-
peratures.
Disadvantages
1. The load capacity 1s many times less than with liquid bearings
of the same size.
2. They arc susceptible to instability.
3. They require a much higher accuracy of machining since the
lubricant films are generally much thinner with gases than with
liquids.
129
Although the general design principles are similar to those already
discussed, there is one additional complication. Air is very com-
pressible so that the assumption that the fluid is incompressible has
to be relaxed. This leads to a rather more complicated form of
Reynolds' equation as follows

dp =l 2
dx YJ
(!J +U2) lph-pmhml·
1
2 ph 3
(7.27)

In this equation p is the pressure at any particular point in the


lubricant film. At constant temperature a gas obeys Boyle's law,
and under working conditions we can suppose that this law holds so
that the relation between the pressure p and the density of the gas p is
P=pC, (7.28)
where C is a characteristic constant that depends on the temperature
and the relative molecular mass (molecular weight) of the gas. Using
this relation converts equation (7.27) into

(7.29)

In this equation p is the density of the gas at the point considered


and p 111 is the density at the point where h = h111 , i.e. where dp /dx = 0.
If Pm= p, that is, if the density does not change with pressure, this
equation is the same as equation (7.8) as we would expect. Particular
bearing problems are then solved using equation (7.29) instead of
equation (7.8), but following the same sequence of steps.

7.4. Elastohydrodynamic lubrication (E.H.L.)


7.4.1. Introduction
In the preceding discussion on hydrodynamic lubrication we have
assumed the viscosity to be constant. This assumption is reasonably
acceptable for the types of situation which have been so far considered.
We now wish to consider some other applications where the intrinsic
contact pressures are much higher than those occurring in say a
journal bearing. Such high pressures may be generated in situations
where the contact geometry is basically described by two ' non-
conforming ' circular arcs. Typical examples are wheels on rails,
ball bearings and roller bearings, and gear teeth contacts. The main
features of such metallic contacts are that the two materials in contact
are both hard and that the contact zones are relatively small ( see
Chapter 2). With such contact areas often measured in fractions of a
square millimetre, even modest loads give rise to very high contact
pressures which may approach 3 x 109 Pa (N m- 2 ).
130
For the analysis of this family of situations we shall consider the
behaviour of a pair of discs pressed into contact in the presence of a
lubricant. The peripheral speeds of the discs may be the same as
in pure rolling contact, or they may be different, and so give a com-
bination of rolling with some degree of sliding. vVith such a system
the hydrodynamic film will have a thickness and a pressure distribution
such that the integral Jp dA of the pressure over the contact area
must be in equilibrium with the applied load.

7.4.2. Film thickness


From our previous discussion which led to the formulation of the
simple Reynolds' equation (see fig. 7.7 b) the equation governing
hydrodynamic lubrication is
dp = 12 ( U1 + U2 ) (h - hm) (7 .30)
dx YJ 2 Jz3 ·
Recalling the basic steps in the solution of lubrication problems on
page 123 we now need to find how h will vary with x. It ·will suffice
for our purposes to show that it is possible to solve the problem using
the available information. Clearly h varies with x according to the
geometry defined by the two circular arcs of the disc. Solving
equation (7.30) using appropriate boundary conditions gives the
pressure distribution in the contact region, the minimum film thick-
ness and the load as indicated by steps 2 to 5 on page 123.
The answer we obtain is relatively simple and is

(7 .31)

where h 0 is the minimum film thickness and U is the average value


of U1 and U2 , and where the pressure distribution is as shown in
fig. 7.17 a. The value R in this equation represents the size of the
contacting discs and is given by
1 1 1
-=-+- (7.32)
R R1 R2
the subscripts referring to the discs 1 and 2 respectively.
If we now compare this prediction with many experimental results
we find an unfortunate lack of agreement. Experience shows that
with steel discs the film thickness at a given load and speed using
mineral oils is some ten or more times greater than that suggested by
equation (7.31). In other words our theory is totally inadequate and
we must therefore examine our assumptions to identify the cause of
the discrepancy. The error has already been hinted at and it arises
131
because in these very high pressure contacts the viscosity increases
markedly due to the effect of the high pressure (see Chapter 5).
What then will be the effect on our solution if we allow the viscosity
to increase with pressure? If we examine the physical meaning of
equation (7 .30) we see that it tells us that the build-up of pressure
as x increases (i.e. the value of dp/dx) is a function of the viscosity
which is itself now increasing with the pressure. Thus the pressure
build-up must occur at an ever-increasing rate, in the form shown
in fig. 7.17 b. This leads to a much better load-carrying capacity,
which is nearer to our experimental observations.

2
/
0-'l.-
r (al

/
:::----.,._ U2
h Ye c,_,--:: lsoviscous pressure
0

1/U1

Pressure viscosity effect

Fig. 7.17. The hydrodynamic pressure distribution between loaded discs for
the isoviscous and the pressure viscosity cases.

Unfortunately we again find a problem arising from this solution.


The very rapid pressure build-up actually predicts a rise to infinite
pressures in the oil film. Since this pressure must also act on the
solid discs we must ask, what are the physical consequences of such
tendencies? No solids can accept infinite pressures and they must
therefore deform and in so doing they lose their circular arc geometry
in the manner shown in fig. 7.18. In such situations the geometry
of the film thickness used in equation (7.30) should be different from
the circular-arc geometry that was originally assumed.
We can now appreciate the mathematical complexities of E.H.L.
solutions, although the physics! arguments are essentially simple.
We need to solve the basic lubrication equation, (7.30), incorporating
the facts that the viscosity increases with pressure and that the film
132
geometry is the elastically deformed shape produced by the operational
pressures. Such solutions generally require the use of digital
computers, but it is nonetheless interesting to note the form of the
results. It should also be recognized that viscosity does not increase
with pressure for all fluids and that the susceptibility of materials to
deformation also varies. \,Ve can therefore identify four possible
situations that can arise in practice.
//Pressure

~ Basic geometry
low loads

-,
/ , /Pressure
I I
/ I
~ I I
\
,,, Initial deformation
due to high

~
loco I pressure

/i
/ \/Pressure
,.--✓ \
/ \

~ Nature of film shape


due to deformation
at even higher loads

Fig. 7.18. The nature of the deformation of the contacting discs clue to the
hydrodynamic pressures generated.

1. Rigid-isoviscous
These are situations where the elastic deformations may be
neglected and the effects of pressure on the viscosity of the fluid may
be negligible. This is the situation discussed at the beginning of
this section, so that the minimum film thickness h 0 is given by

t~ = (it). 4 ·9 (7.33)

2. Elastic-isoviscous
Herc we incorporate the effects of elastic deformation but again
assume that the viscosity is constant. For such situations we find
that
Ji 0 = 2 .35 (
R rv
0 0 4
·o ( ~ ) ·
RE'
riU)
(7.34)

133
In this solution we note two groups, one as it were defining the
hydrodynamic effects, and the other the elastic deformation effects.
The latter group contains the elastic constants of the materials since

(7.35)

the subscripts 1 and 2 referring to the values of the elastic constants


of the two materials in ' contact '.
A typical example where this solution would apply is with rubber
tyres and water lubricant, i.e. the behaviour of car tyres on very wet
roads leading to aquaplaning. In this case E for the road is virtually
infinite relative to E for the rubber, so the equation becomes 1 IE'
=(l -v 2 )/E, the values of v and E being those of the rubber.
3. Rigid-viscous
Here we neglect the effects of elastic deformation but allow the
viscosity to increase with pressure. The dependence may be
expressed as
YJ = YJo exp rx(p - Po), (7.36)
where rx is a constant and Y/ = YJo when p = p 0 , the atmospheric pressure.
The resulting equation for the minimum film thickness is then
lz 0 _ • (YJU) 2 l 3 (cxW)2/3 (7.37)
R -166 w R .
In this case the two dimensionless groups may be identified as
representing the hydrodynamic effect and the pressure/viscosity
effect respectively. Such solutions are reasonably correct for hard
steel materials with mineral oil lubricants.
4. Elastic-viscous
This incorporates both the elastic deformation and the pressure/
viscosity effects and is the most complex situation. The film thick-
ness is now given by all three dimensionless groups representing the
hydrodynamic, viscosity /pressure and elastic deformation respectively.
h0 = 2 .6 (YJU) 0 ·7 (rxW)o-54
R ~V, R
(~)o·o3
RE' .
(7 .38)

7.4.3. Surface tractions


The preceding section has indicated the formulae for calculation
of the film thickness in the various cases considered. As far as film
thickness is concerned, the most important feature is that a hydro-
dynamic film must exist which will completely separate the surfaces.
This means that it will be significantly thicker than the size of the
asperities which are found on all practical surfaces.
134
We now turn to the problem of calculating the surface shear stresses
which in turn give us the value of the frictional forces in this type of
lubrication. The analysis of these frictional forces is beyond the
scope of this book, but some observations are apposite in identifying
the main features of surface frictional drag in elastohydrodynamic
lubrication.
In conventional hydrodynamic lubrication the friction forces are
very small and typically give coefficients of friction of the order of
0·001. In E.H.L. lubrication with conventional lubricating oils
between steel surfaces we find that the coefficient of friction may be
some two hundred times greater than this value. This is illustrated
by the typical results shown in fig. 7.19 where F/Wis plotted against
the slide/roll ratio in E.H.L. experiments. In pure rolling the F/W
values are about some ten times greater than the value of conventional
lubrication but they then rise towards a maximum value after which
they tend to fall slowly to a steady value. Increasing the speed U
at the same slide/roll ratio tends to decrease the value of Ff W whilst
increases of load have the opposite effect.

01

--------+- Slide/ roll = u, - uz


u, + uz
Fig. 7.19. The F/J¥ value as a function of the slide/roll ratio between rotating
discs.

The complex shape of the curve in fig. 7 .19 is explicable in terms


of three interacting physical effects. The high pressures cause an
increase in viscosity so that the friction is increased and this effect is
enhanced by increases in load. The sliding and increase of speed
tend to cause more shearing of the lubricant which results in higher
lubricant temperature and therefore a reduction in viscosity and the
ensuing friction. The interactive effects of pressure and temperature
lead to a non-Newtonian type of behaviour which is characteristic
of visco-elastic effects. The very high values of friction are believed
to indicate that fluids such as mineral oils begin to be so compressed
K 135
Fig. 7.20. The nature of the breakdown of grease structure in heavily loaded
contacts.
136
that they behave almost like a viscous solid such as pitch, but the oil
appears to return to its more or less normal fluid state after passing
through the contact zone, i.e. after the release of the high pressures.
Finally, it is of some interest that with more complicated lubricants
such as greases the effect of the high pressures in E.H.L. contacts is
actually to break down the grease structure. The long fibrous strncture
of grease shown in fig. 7.20 is reduced to a series of very small solid
particles suspended in the carrier oil.

(a)

Disc

W/2 w 0 W/2

Oil
' --- -- - - -

(bl

Pressure
distribution

Oil

Fig. 7.21. A simple hydrodynamic pad bearing test apparatus,

7.5. Project suggestions


The hydrodynamic behaviour of bearings is most easily studied by
an apparatus of the form shown in fig. 7.21. A pad is made which
conforms to the periphery of a disc which is driven by a variable speed
137
motor. The pad is supported by two coil springs as shown so that
it has freedom to adjust its inclination to the disc, but is at the same
time reasonably constrained. The oil supply is provided by the disc
which dips into an oil reservoir and thereby carries oil to the contact.
The load is most easily applied by deadweights as shown. The film
thickness may be measured by the capacitance between the pad and
the disc provided these are carefully designed to avoid any alternative
electrical conducting path, so the pad should be supported by springs
attached to insulated supports such as a wooden board. With this
apparatus the effects of load, speed and viscosity on the film thickness
and friction are readily determined. The friction is recorded by
measurements of the deflection of a transverse support spring or by
an appropriate displacement transducer.
Further refinements to such a test rig are to devise methods of
measuring the inclination of the pad with respect to the disc. The
pressure distribution over the pad may easily be obtained by providing
a series of pressure tappings connected to simple manometers as in
fig. 7.21 b.

138
CHAPTERS
the selection of tribological solutions

8.1. Introduction
THE foregoing chapters have identified the scientific principles on
which most tribological solutions depend. Although some examples
of these principles have been discussed they have so far only been
included as illustrations of specific methods of solution to tribological
problems. As yet no attempt has been made to demonstrate the
rules which are used in deciding on, say, a rolling-contact bearing
rather than a hydrodynamic bearing, in a small electric motor. In
most practical situations one method of solution will have advantages
over the others. Also not surprisingly the ' rules of the game '
follow logically from the knowledge of the scientific principles
involved in each method of solution. Since industrial situations offer
the widest scope for the application of tribological knowledge we shall
concentrate on problems taken from this area.

Normal load

~
~
- Friction
wear

Environment

Fig. 8.1. The basic tribological system.

The ' black box ' concept of the tribological solution is of two solid
bodies subjected to a normal load, having relative motion to each other
and the whole operating within a defined environment as in figure 8.1.
The effects from such a system which concern the tribologist are
simply friction and wear.
139
8.2. Environment
The environment may be atmospheric, vacuum, chemical vapours,
water, etc., e.g. the bearings supporting a ship's propeller operate in
sea water whilst some of the bearings of a space vehicle must operate
in the very high vacuum of outer space. The majority of industrial
contacts operate in a normal atmospheric environment, although even
here the ambient temperature may vary from very low values in
certain refrigeration machinery to very high values in such industrial
plants as steelworks.
The complete range of environmental problems cannot be dealt
with in this book, but it will be obvious that in what follows the
designer must always consider the particular environmental problems
when choosing a tribological solution. Thus mineral oils cannot be
used in hydrodynamic bearings operating at very low temperatures
since the oil would solidify, whilst the use of dry plastic bearing
materials at very high temperatures is impossible due to the thermal
degradation of such materials. Over the years engineers have shown
considerable enterprise in overcoming the problems of the environ-
ment as the following examples indicate.
In the liquid-metal-cooled nuclear reactor certain bearings in the
pumping circuit have to be capable of prolonged operation within
the liquid metal environment. This is an essential safety feature
since any atmospheric contamination of a material such as liquid
sodium would be clearly disastrous, so that such bearings have to be
' sealed ' into the system. In these situations the use of conventional
bearings appears impossible until we remember that we already have
a liquid available, liquid sodium! The design which is employed
therefore uses pressurized liquid sodium as the lubricant in a logical
design extension of the conventional hydrostatic bearing.
In bearings which are to be used in ultra-high vacuum applications
such as certain X-ray equipment and in space vehicles, the metallic
parts of bearing components would rapidly weld together. This
problem was discussed in Chapter 3 where it was shown that in the
absence of protective surface films, such as oxides in the atmosphere,
the surfaces weld together to give very high values of friction. The
use of liquid lubricants in such situations is unacceptable since in
such a vacuum the lubricant vapour pressure is still high enough for
the lubricant to evaporate. A well tried solution to this problem is
to use ball bearings and to coat the balls with a solid lubricant. The
' lubricant ' which is often used is a soft metal such as lead, coated
to a thickness of a few micrometres. Bearings treated in this way
can operate satisfactorily for several years in such applications as
communication satellites.
An alternative solution to this problem uses rather more conven-
tional solid lubricants such as molybdenum disulphide. With such
140
lubricants the cage of the ball bearing is made from the solid lubricant
and on each revolution of the balls in the bearing a little lubricant is
transferred to the balls as they slide against the cage. This small
quantity of lubricant is sufficient to lubricate the contacts between the
ball and the races of the bearing.

8.3. Load
The single resultant force acting between bodies in contact is
conveniently considered in terms of two components at right angles
to one another. The normal component of this force is the load
which is being applied across the contact, although it may not be the
externally applied load. Thus if an applied load W is carried on an
inclined slider, the normal load at the contact is W cos ex; see fig. 8.2.
Likewise the contact of a dry rubbing journal bearing would be as
shown in fig. 8.3, where an applied load W gives rise to a normal load
of value W cos ex at the contact. (In these diagrams the clearances
between the shaft and the bearing housing are grossly exaggerated
for ease of understanding.) If, and only if, the two bodies in fig. 8.2
and 8.3 had a constant relative velocity would the system be in
equilibrium, when the components W sin ex would be the friction force
at the contact and we should have tan ex=µ. In the case of the dry
journal bearing the driving torque would then be WR sin ex.

Fig. 8.2. The inclined plane.

In what follows we shall use the term load to define the normal
load carried across the tribological contact. Thus each particular
practical application will have to be analysed to determine the relation-
ship between the load as defined .here and the actual applied load
carried by the particular device. In most cases these two values of
load will be very nearly the same and in many cases are actually
identical.
Load, defined as a force, is a vector quantity, with direction as well
as magnitude. The magnitude of the load may be constant, as for
example in the main bearings of a large electric alternator. It may
vary in magnitude as happens in the main bearings of a car engine
141
throughout the inlet, compression, power and exhaust strokes of the
cylinders. The actual nature of the variation in magnitude of course
depends on the particular application considered. Fortunately the
changes in magnitude of the load do not normally cause much of a
problem since, although an increase in load produces an increase in
the frictional force, the coefficient of friction is not materially affected.
The main difficulties arising from varying magnitude of the load are
unwanted vibration and/or instability and, if the magnitude becomes
very large, breakdown of the bearing. On the other hand, a rapidly
increasing load may even have beneficial effects. For example, in a
lubricated bearing the higher the pressure of the lubricant film the
greater the load-carrying capacity. Consequently a rapid increase
in load will tend to produce higher pressures in the lubricant film by a
squeeze film effect (Chapter 7, section 7.2.6).

Fig. 8.3. The simple journal bearing.

The direction of the applied load may be constant, again as in large


electric alternators, or it may vary; a simple example being the
connecting rod bearing of a car engine. In this case the applied
load will be in the general direction of the connecting rod whose
inclination to the vertical varies as the crankshaft rotates, fig. 8.4.
Again changes in the direction of the load may sometimes prove
beneficial as in the following simple example. Consider a simple
dry journal bearing of the type already shown in fig. 8.3. Assume
142
that an applied load of constant magnitude is such that its direction
varies between the limiting positions A and B in fig. 8.5. As the
load direction changes the rubbing contact on the bearing surface,
which gives rise to wear, will also change over the range P to Q.
The wear of such a bearing will be somewhat less than would occur
for the same load acting in a constant direction, where all the wear
would occur at the same contact patch. Thus a varying direction of
load is seen to give rise to a longer service life of a dry rubbing bearing.

~Piston

' '\ '


\ I'
I I
I I
I I
\
Cronk / \
..... __ __,,,✓/ ''

Fig. 8.4. The bearings in a reciprocating engine.

w
/

r2'_-;:i;- Load at A contact at P


: w/ /
I / I Load at B contact at Q
// I
I I
,L____ _J

Fig. 8.5. The effect of a varying direction of load,


L 143
8.4. Speed
Although misunderstanding is rarely possible in the context,
strictly speaking speed is a scalar representing the magnitude of rate
of motion (such as a speedometer reads), while velocity is the term
for the corresponding vector quantity with both magnitude and
direction. Speed, like load, is a major operational variable in all
tribological devices. The magnitude of the sliding velocity tends to
be rather critical. At very low speeds most solids demonstrate a
rather jerky motion due to stick/slip effects, whilst with hydrodynamic
designs such speeds are insufficient to create continuous fluid films.
At high speeds one finds problems due to such effects as frictional
heating, inertia force effects, instabilities and the onset of turbulence
in fluid film bearings.

Fig. 8.6. The contact of a pair of gears.

The direction of the sliding velocity may be continuously in one


direction as in a simple journal bearing, or it may be continually
reversing as in all reciprocating motions. The contact betvveen the
piston rings and the cylinder in our car engine is one such obvious
example, while the component parts of the car's suspension system
are clearly subjected to a rather more random form of oscillatory
motion. \Vhen the amplitude of oscillation is fairly small we can
often solve our tribological problem by using some form of elastomer
which allows the relative motion without utilizing the more traditional
tribological solutions of fig. 1.5.
There is a further speed property to be considered which occurs in
those situations where any form of rolling motion occurs. It has
been shown in Chapter 3 that in many rolling contacts the general
rolling motion is accompanied by some degree of sliding. Thus the
wheels of a car may be rolling with only negligible slip, rolling with
144
measurable sliding or, in the case of a skid, simply sliding. All these
conditions may be defined by the ratio roll/slide, which increases as
the degree of rolling increases. This ratio may be constant in both
magnitude and direction, as in a car wheel running under steady
conditions. Alternatively, the roll /slide ratio may vary in both
magnitude and direction. This is best typified by the contact of a
pair of gear teeth, fig. 8.6. Initially the teeth roll with sliding in one
direction, the degree of sliding reducing to zero at the pitch point
contact P and thereafter increasing again, but in the opposite sense
as the teeth move out of contact.

8.5. Tribological limits of load and speed


8.5.1. General
From the preceding discussion of load and speed we may summarize
the possible operational parameters of load and speed as follows.
The applied load may be constant or varying in both magnitude and
direction, which leads to the following four possible loading conditions:

l\Jagnitude Dircct£on
Constant Constant

Varying Varying

Applying the same format to the speed conditions we find there arc
eight possible speed conditions:

l\Jagnitude Direction Roll/Slide


Constant Constant Constant

Varying
~ Varying
X Varying

If we recognize that each tribological contact has a specific load and


speed condition we can suppose that there are no less than thirty-t\rn
different combinations of load and speed. Furthermore, in these
arguments we have not included any specific definition of the \Yay
in which variations of load and speed may occur. A few common
examples of load and speed conditions in tribological contacts are
illustrated in fig. 8.7.
Fortunately a very large number of tribological situations are
substantially of the constant load and speed type, and it is therefore
this situation \Ve shall now examine in greater detail. Furthermore
for the sake of simplicity we shall in general assume that the tribological
device is operating under normal atmospheric conditions.
145
Journal bearing Connecting rod
bearing
Load and speed constant Load, speed and Lood varies in both
in both magnitude and roll /slide ratio all magnitude and
direction vary during contact direction

Com Car wheel

Load and speed Load, speed and roll/slide


both vary in ratio ore constant in
magnitude steady state

Fig. 8.7. Some practical tribological contacts.

8.5 .2. Strength considerations


In what follows we shall interpret the word ' failure ' to mean those
conditions when the system fails to operate in the manner for which
it was designed. This does not necessarily imply that the system fails
to function completely. Thus a ball bearing may fail, as evidenced
by a deterioration in smooth running, although it is still able to rotate
and support load.
Almost all engineering designs are based on the supposition that
the bulk materials will behave elastically although local plastic deforma-
may occur at the surface asperities during contact. We can calculate
146
the maximum load which may be carried across solid body contacts
if the bulk material 'is to be restricted to elastic behaviour. This
load will be independent of the speed and might be termed the static
strength of the system. It is represented in fig. 8.8 a by the horizontal
line defining this maximum load whose value is calculated from our
knowledge of the mechanics of solids. Its actual value will depend
on the elastic limit of the materials used and the geometry of the
system. Thus with rolling clement bearings we use the Hertzian-type
analysis mentioned in Chapter 2 to define this load; in this case this
being the load which ensures that the balls do not plastically deform
the races.

(a) (b)
logW logW
~
Static strength

Inertia
force
limit

log U log U

(cl (d)
log W

logU

Fig. 8.8. Some typical tribological limitations.

The relationship between speed and failure is rather more com-


plicated. However, we recognize that many systems involve rotating
machinery, of which bearings of many types are the most obvious
examples. The effect of rotating any solid is to set up inertia forces
arising from the necessary centripetal, and possibly Coriolis, accelera-
tions. The forces are resisted by the strength of the solid materials,
147
but at high speeds they can produce large enough stresses in the solids
to lead to failure of the component. Examples are the bursting of
cages in ball bearings and the actual distintegration of large diameter
solid shafts at high speeds. The inertia forces and therefore the
resulting stresses are proportional to the square of the rotational
speed. The strength limitation arising from such speed effects is
therefore defined by a maximum speed for the bearing, etc. (fig. 8.8 a).

8.5.3. Wear considerations


Any engineering component must have a reasonable life before it
fails clue to the cumulative effects of wear. In Chapter 4 we saw
that in both the major types of wear considered, adhesive and abrasive
wear, the volume of wear V was directly proportional to the applied
load W and the distance travelled L, and inversely proportional
to the hardness H. The constant of proportionality K is much
greater for abrasive wear than for adhesive wear. In both cases
we may write
V =KWL
H . (8.1)

In practice we are most interested in the rate at which wear occurs


rather than in its absolute extent. Thus by dividing equation (8.1)
by the time interval we obtain a relation between rate of wear V, iv
and the velocity U, of the form

V = ~ WU. (8.2)

This is most easily illustrated graphically by a log W against log U plot,


since from equation (8.2) we see that

log W= -log U +log (Kl'H) . (8.3)

Thus a specified wear rate for given materials will always be defined
by a straight line of slope - 1 when log Wis plotted against log U as
shown in fig. 8.8 b.

8.5.4. Fatigue considerations


In many systems as motion proceeds clements of material arc
being loaded and unloaded in a prescribed pattern. An example
occurs in a rolling contact bearing where the rolling elements are
continuously loading and unloading the material of the races. This
type of situation is just that in which we would expect metal fatigue
to occur. The essential feature of fatigue behaviour is that the
higher the stress imposed (i.e. the higher the load) the less will be
the number of cycles of loading and unloading before failure occurs.
148
With rolling bearings the higher the load applied the smaller the
number of revolutions before failure occurs, for the number of
revolutions defines the number of loading cycles on each ball in the
bearing.
For a ball bearing it has been shown that the total number of
revolutions leading to fatigue failure is inversely proportional to the
cube of the load, so we can write
L 1.fe 111
. 1
revs. rx H'3 •
For a bearing rotating at frequency U this life may be defined by a
'time to fatigue' t since the life (in revolutions) is equal to Ut, and
hence Utrxl/W 3 • No,v if we require the bearing to have a given
fatigue life as we vary both the load and speed, the required relation-
ship between the load and speed is that UW 3 must be constant.
Taking logarithms this relationship becomes
log U + 3 log W = C = Constant
or log W = - ¼log U + f.
This clearly results in a fatigue limit defined by a straight line of
slope - ½, as in fig. 8.8 c. For roller bearings the relationship is
different since these bearings are found to have a life in revolutions
which is inversely proportional to the square of the load.

8.5.5. Hydrodynamic films


In Chapter 7 we saw that in certain circumstances self-acting
hydrodynamic films may be generated between moving surfaces.
Furthermore we noted that such films are characteristically dependent
in thickness on the non-dimensional group ( 11 U / W) where 1/ is the
viscosity of the lubricant, e.g. equation (7 .17). It was also stressed
in Chapter 7 that this oil film thickness must always be sufficient to
provide complete separation of the surfaces, remembering that all
solid surfaces are covered with microscopic asperities.
The guarantee of the existence of a minimum film of lubricant in
hydrodynamic bearings thus requires that some function of the
group ( 1/ U / W) is never less than a certain critical value. For the
simple slider bearing we note for instance from equation (7 .17) that
( 1/ U / W) 1 12 is the particular function involved, whilst in other cases
the index would have a different value; for two discs the index is 1 as
shown by equation (7.33). In general the physical meaning of this
argument is that as speed is increased the load ,vhich may be supported
is also increased without any change in the thickness of the lubricant
film. On the log/log plot this leads to the type of limit shown in
fig. 8.8 d.
149
The initial part of this limit is essentially a straight line as we would
expect from the foregoing arguments. As the speed increases so also
does the friction of such bearings as shown by equation (7.21).
This friction is dissipated as heat producing a rise in temperature of
the lubricant, which in turn produces a fall in its viscosity. The nett
effect of this sequence of events is to reduce the load capacity of the
bearing at any particular speed and this is represented by the
curvature in fig. 8.8 d.

8.5.6. Other considerations


The foregoing has shown how one may use the scientific principles
of tribology to define limits of performance. There are, of course,
many other limits of performance which could equally well be included
but a few examples must suffice in a text of this type. For the
highly loaded contact conditions discussed at the end of Chapter 7
one may define the limits which ensure adequate hydrodynamic films.
For very slow speed sliding one may identify the limits below which
frictional vibrations may occur. Try sliding your hand slowly over
the table and you can feel this intermittent type of motion. For
certain systems one may wish to define a limit which ensures that
certain maximum temperatures are not exceeded. Limits may be
calculated which define the conditions if one is to avoid certain
instabilities in behaviour such as cavitation or turbulence of the
lubricant.

8.6. The use of tribological limits


8.6.1. General
We have seen how we may decide the limits of load and speed;
the designer must choose those limits which he knows will decide
the limits of performance of a particular tribological device and
thereby identify the safe operating regime of load and speed. Let
us examine this process for three different types of journal bearing:
the dry rubbing journal bearing based on such tribological materials
as PTFE; the simple radial ball bearing and the hydrodynamic
oil-lubricated journal bearing.
(a) The dry rubbing bearing
The performance of such a bearing is limited by ( 1) the
maximum static strength limit, (2) the pv factor for the material
and (3) the maximum speed resulting in thermal degradation
of the material. These three limits are shown in fig. 8.9 a,
and the bearing will be entirely satisfactory at any load and
speed combination lying within the regime defined by these
three limits.
150
(b) The radial ball bearing
This bearing is defined by the limits of static strength, fatigue
strength and a maximum speed above which inertia forces lead
to failure (fig. 8.9 b).
(c) The hydrodynamic journal bearing
This bearing is limited by the hydrodynamic film limit and a
maximum speed limit arising from failure due to inertia forces,
in this case centrifugal forces (fig. 8.9 c.)

log W

Safe Safe
operating operating
regime regime

log U log U log U

Ory rubbing bearing Bal I bearing Hydrodynamic bearing

Fig. 8.9. The definition of the safe operating regimes for three different types
of bearing.

8.6.2. Tlze selection of bearing type


\iV e may now put together our scientific knowledge of tribological
limits in such a way as to facilitate the correct choice of bearing type
for any particular application. \iV e shall only consider the problem
of choosing between the three types of journal bearing already
discussed, although it is only a matter of increased complexity to
include all other types of journal bearing such as hydrostatic journal
bearings and self-lubricated bearings.
In selecting a journal bearing for a particular application we simply
require to support a shaft of a given diameter and we shall therefore
use shaft diameter as our basic specification. For a shaft of a given
diameter, that is a particular size of bearing, we may readily compute
the appropriate tribological limits for each type of bearing. The
results of such computations are shown for shaft diameters of ¼inch
and 20 inches in fig. 8.10. It will be noted that these computations
cover a very wide range of both load and speed in the log/log plot of
fig. 8.10, and cover virtually the whole range of loads and speeds used
in engineering practice. Computation of the tribological limits for
151
20
_ in} . shaft
m d1a.
0 51

z - - Hydrodynamic bearing
C C
- · - Dry rubbing bearing
- - - - Ball bearing

10

10 2
3
---~--
'\.
"'--
. ............ ,

·"'· -----
10
1
0·25 in } .
_ m dta shaft
0 0064
10

10
R.P.M.

Fig. 8.10.~ A chart for the selection of bearing type.

bearings based on intermediate sizes have been excluded in fig. 8.10,


although they are available in Design Data Item 65007 published by
the Institution of Mechanical Engineers.
We are now in a position to use fig. 8.10 to help us in the selection
of bearings for particular applications. Let us suppose that we wish
to select the best bearing for a ¼ inch diameter shaft running at
1000 r.p.m. From fig. 8.10 we see that the load which may be
safely applied to a ball bearing at this speed is much greater than either
the hydrodynamic or the dry bearings. We should therefore use a
ball bearing for this application and in engineering practice this is
the bearing normally used in small electric motors, etc. At the same
speed for a 20 inch diameter shaft the greatest load capacity is
provided by the hydrodynamic bearing, some 10 6 N in this case.
The ball bearing of this size has a much lower load capacity while
the dry rubbing bearing is not even a practical proposition for this
size of shaft. This explains why hydrodynamic bearings are used
in very large steam turbine alternator sets for electricity generation.
152
Fig. 8.10 also indicates how the maximum speed of ball and hydro-
dynamic bearings, which is defined by inertia forces, reduces as shaft
size increases. Indeed for hydrodynamic bearings this limit is
defined by centrifugal force effects alone and is found to be inversely
proportional to the shaft diameter as we would expect.

8.7. Conclusions
Our discussion has shown how the scientific principles of tribology
may be used to help us in the choice of the best tribological solution
for a particular application. In practice, however, life tends to be
rather more complicated than these simple ideas might suggest.
For instance, having decided to use a ball bearing for a particular
situation we must now examine the wide range of different types of
ball bearing which are available. A glance at any ball bearing
manufacturer's catalogue will indicate the extent of this range and
further engineering expertise is required if the correct choice is to
be made. Such knowledge is beyond the scope of this book, but the
foregoing should have illustrated the value of tribological knowledge
in arriving at the optimum solution.

log U
Fig. 8.11. The matching of bearing characteristics to the operational loads
and speeds.

A further point to be borne in mind is that we have restricted our


arguments to constant load and speed situations. In many industrial
applications these conditions do not apply, so that the effects of
varying load and speed must also be considered when choosing our
bearing type. As an example consider the problem of choosing the
bearing type for a large gearbox to be used in a marine installation.
In this situation the load on the bearings will increase as the speed
is increased in the fashion shown in fig. 8.11. It is therefore obvious
153
108
10000
(\J

E
'-

@))
9
Ll-

(\J ~
:,
E I 1000 U>
'-
z I
Elastomer
U>
~
a.
~ c,,
:, C
U>
U>

Q)
Q)
d:: Ll
10 6 E
:,
100 E
"><'
0
:z

10 5 -
10
0 10 20 30
Maximum angular deflection of oscil lotion degrees

Fig. 8.12. The characteristics of elastomer type bearings.

that the tribological limits defining the performance of hydrodynamic


bearings are a very good match for this particular application.
Finally one should always be asking whether a conventional type
of bearing is really necessary. Thus if the motion to be dealt with
is a simple oscillatory circular motion one may obtain a satisfactory
solution by using a rubber bushing bonded to the two surfaces. The
motion is then accommodated by the flexure of the rubber and one
therefore has virtually a zero-wear bearing. Bushes of this type are
commercially available and fig. 8.12 gives a typical example of the
relationship between their load capacity and the maximum angle of
oscillation. Naturally such devices are not satisfactory for very high
speed oscillations.

154
INDEX

abrasiYe wear 66, 147 design, application to 151


additives 8 design data 151
adhesion 46, 76 dimensionless groups 123, 133
adhesion theory 43 disc machine 59, 92
Amontons's Laws 4, 41 distributions, ordinates 27
aquaplaning 133 peaks 28
Archard constant 66 valleys 28
area of contact 33, 41 slopes 28
arthritis 13 gaussian 29, 35
asperities 33, 43 standard deviation 27
asperity interaction 43 driving traction 50
autoradiograph 80 dry rubbing bearing 150

ball bearings 6, 150 eccentricity 111


bearing area curve 28 economics 11
bearing alloy 45 elastic hysteresis 42, 51
Beauchamp Tower 6 elastic work 51
boundary lubrication 47, 74 elastohydrodynamic lubrication 129
braking traction 50 elastomer 9, 153
electron microscope 19
electron micro-probe 25
centre line 27 energy dissipation 42
centre line average 27 environment 113
coefficient of friction 4 evolution 2
cobalt 8 externally pressurized bearing (hydro-
compatibility 74 static bearings) 10, 96
conformity 54
connecting rod bearing 141 fatigue failure 68
contact, ellipse of 54 fatigue limit 146
contact pressure 32 film thickness 104, 123, 130
contact region 50 films, soft 46
contact zone 54 four ball machine 78
contaminant films 8, 47 free rolling 50
cornering force 5 5 fretting 71
Couette flow 115 friction 41
Coulomb 4 adhesion theory 43
creep 54 kinetic 41
crossed cylinder machine 57, 78 laws of 4, 41
crystal structure 18, 73 measurement of 57
cylindrical contact 32 coefficient of 36, 44
force 42
deformation, elastic 30, 33, 42
plastic 34, 42 gas lubrication 128
of surfaces 33 Gaussian distribution 28, 35
155
gear contact 79, 119 microslip 53
Gough plot 55 mineral oils, additives 93
graphite 8, 75 extreme pressure 94
greases 94, 135 flashpoint 93
foaming of 94
oxidation of 93
hardness 32 pour point 93
effect on wear 72 molybdenum disulphide 8, 135
heat source, stationary 37
moving 37
Hertz, equations 32 Newton 6
Hertzian, pressure distribution 31 Newtonian behaviour 84
historical survey 3 nominal pressure 35
hydrodynamic limit 146 non-Newtonian behaviour 84, 134
hydrodynamic lubrication 10, 110 normal load 138
hvdrostatic lubrication 10
hydrostatic bearings, (externally
pressurized) 96
characteristics 104 oxide films 17, 139
compensation of 101
constant flow 99
constant pressure 101 pad bearings 99, 113
effect of sliding 107 partial journal bearings 111
flow in 102 Petroff 6 ·
load capacity of 104 pin and disc machine 58, 78
stiffness of 104 planned obsolescence 12, 63
plastic behaviour, of solids 33, 42
of fluids 136
instantaneous axis 54 plastics, effects of temperature 77
friction of 7 5
inertia effects 143
inertia force limit 146 pv factor 76
wear of 75
interferometry 19
Poiseuille flow 87, 97, 115
Poisson's ratio 30
polytetrafluoroethylene 8
journal bearing 105, 112, 143 profilometer 21
junction growth 47 pressure on viscosity 86
pressure distribution, between solids,
31
lamella solids 8, 75 hydrodynamic 114
Leonardo da Vinci 4 elastohydrodynamics 132
load effect on wear 64 hydrostatic 99
load-carrying capacity 120 pv factor 76
locomotive wheels 52
lubricants, properties of 93
lubrication, boundary 8 radioactive tracer 79
hydrostatic 96 real area of contact 33
hydrodynamic 110 of rough surfaces 33
elastohydrodynamic 129 Reynolds, Osborne 6
Reynolds' equation 116
gas lubrication 129
macro-roughness 18 sliding motion 116
magnetic fields 10 squeeze motion 127
micro-roughness 18 road surface 56
microscopy 19 rolling contact 9, 33, SO, 69, 143
156
rolling friction 50 turbulence 143
rolling friction machine 60 tyre-road contacts 55
root mean square value 27
roughness 18
rough surface contact 33 vehicle cornering 55
running-in 12, 63 visco-elastic effects 134
viscometer, capillary 87
efflux 88
self-aligning torque 55 falling body 90
slip areas 52 rotational 91
soft films 45 viscosity 83
solubility 74 dynamic 83
spherical contact 32 index 85
kinematic 84
squeeze motion 125, 141
static strength limit 146 measurement of 87
statistical properties of surfaces 27 units of 84
stick-slip 143
Stokes 6
waviness 18
stress distribution due to contact 31
stylus 24 wear 7, 63
surface, nature of 16 abrasive 65, 66
adhesive 65, 147
ordinate distribution of 27
peak distribution of 28 corrosive 65, 70
slope distribution of 28 definition of 64
valley distribution of 28 fatigue 65, 68
fretting 71
surface asperities 33, 43
surface films 4 7 measurement of 77
mild 64
surface profile 21, 81
severe 64
wear limit 146
wheel 3, 52
tangential traction 52
Talysurf 22
taper section 21 X-ray diffraction 25
temperature of surfaces 36
temperature effects on hardness 73
on viscosity 84 yield stress 18, 44
tribology 1 · Young's modulus 30

157
sen

THE WYKEHAM SCIENCE SERIES


1 tElementary Science of ;l,Jetals J. \V. MARTIN and R. A. HULL
2 Neutron Physics G. E. BACON and G. R. NOAKES
3 ·\·Essentials of 1\/leteorology D. 1-1. McINTOSH, A. S. TnoM and V. T. SAUNDERS
4 Nuclear Fusion H. R. HULME and A. McB. CoLLIEU
5 Water Waves N. F. BARBER and G. GHEY
6 Gravity and the Earth A. H. CooK and V. T. SAUNDERS
7 Relativity and High Energy Physics W. G. V. ROSSER and R. K. McCULLOCH
8 The Z\llethod of Science R. HARRE and D. G. F. EASTWOOD
9 tintroduction to Polymer Science L. R. G. TRELOAR and \V. F. ARCHENHOLD
10 t The Stars; their structure and e~,olution R. J. TAYLER and A. S. EVEREST
11 Superconductivity A. \V. B. TAYLOR and G. R. NoAKES
12 Neutrinos G. M. LEWIS and G. A. WHEATLEY
13 Crystals and X-rays H. S. LIPSON and R. M. LEE
14 tBiological Effects of Radiatio11 J. E. CoGGLE and G. R. NOAKES
15 Units and Standards for Electromag11etis111 P. VIGOUREUX and R. A. R. TRICKER
16 The Inert Gases: 1'\llodel Systems for S'cie11ce B. L. SMITII and J.P. \VEB13
17 Thin Films K. D. LEAVER, B. N. CHAPMAN and H. T. RICHARDS
18 Elementary Experiments with Lasers G. \VmGHT and G. FoxCROFT
19 tProduction, Pollution, Protection \V. B. YAPP and lVI. I. SMITH
20 Solid State Electronic Devices D. V. l\loRGAN, :\1. J. HowEs and J. SUTCLIFFE
21 Strong Materials J. W. MARTIN and R. A. HULL
22 tElementary Quantum 1\/lechanics Sm NEVILL MOTT and M. BERRY
23 The Origin of the Chemical Elements R. J. TAYLER and A. S. EVEREST
24 The Physical Properties of Glass D. G. HOLLOWAY and D. A. TAWNEY
25 Amphibians J. F. D. FRAZER and 0. II. FRAZER
26 The Senses of Animals E. T. BURTT and A. PRINGLE
27 tTemperature Regulation S. A. RICHARDS and P. S. FIELDEN
28 tChemical Engineering in Practice G. NoNHEl3EL and M. BERRY
29 tAn Introduction to Electrochemical Science J. O'l\il. BocKRIS, N. BONCIOCAT,
F. GUTMANN and M. BERRY
30 Vertebrate Hard Tissues L. B. HALSTEAD and R. HILL
31 tThe Astronomical Telescope B. V. BARLOW and A. S. EVEREST
32 Computers in Biology J. A. NELDER and R. D. KIME
33 Electron 1'\llicroscopy and Analysis P. J. Goommw and L. E. CARTWRIGHT
34 Introduction to 1'\llodern Microscopy I I. N. SOUTHWORTH and R. A. HuLL
35 Real Solids and Radiation A. E. 1-Iuc1rns, D. POOLEY and B. WooLNOUGH
36 The Aerospace Environment T. BEER and lVI. D. KUCHERAWY
37 The Liquid Phase D. H. TREVENA and R. J. CooKE
38 ·J·From Single Cells to Plants E. THOi\lAS, l\il. R. DAVEY and J. I. \V1LLIAMS
39 The Control of Technology D. ELLIOTT and R. ELLIOTT
40 Cosmic Rays J. G. WILSON and G. E. PERRY
41 Global Geology l\!I. A. KHAN and B. iVIATTIIEWS

THE WYKEHAl\1 ENGINEERING Al'JD


TECHNOLOGY SERIES
1 Frequency Conversion J. THOMSON, \V. E. Tum, and M. J. BEESLEY
2 Electrical Measuring Instruments E. HANDSCOM13E
3 Industrial Radiology Techniques R. HALMSHAW
4 Understanding and 1\/leasuring Vibrations R.H. WALLACE
5 I11troduction to Tribology J. HALLING and \V. E. \V. SMITH
All orders and requests for inspection copies should be sent to the appropriate agents.
A list of agents and their territories is given on the verso of the title page of this book.

t (Paper and Cloth Rditions available.)

You might also like