Вы находитесь на странице: 1из 464

Rolling Contacts

922$$$$pg1

02-10-00 12:18:34

This page intentionally left blank

Rolling Contacts
by
T A Stolarski
and
S Tobe

Professional Engineering Publishing Limited


London and Bury St Edmunds, UK

First published 2000


This publication is copyright under the Berne Convention and the International
Copyright Convention. All rights reserved. Apart from any fair dealing for the
purpose of private study, research, criticism, or review, as permitted under the
Copyright Designs and Patents Act 1988, no part may be reproduced, stored
in a retrieval system, or transmitted in any form or by any means, electronic,
electrical, chemical, mechanical, photocopying, recording or otherwise, without the prior permission of the copyright owners. Unlicensed multiple copying
of this publication is illegal. Inquiries should be addressed to: The Publishing
Editor, Professional Engineering Publishing Limited, Northgate Avenue, Bury
St Edmunds, Suffolk IP32 6BW, UK.
T A Stolarski and S Tobe

ISBN 1 86058 296 6


ISSN 1470-9147

A CIP catalogue record for this book is available from the British Library.

The publishers are not responsible for any statement made in this publication.
Data, discussion, and conclusions developed by the authors are for information only and are not intended for use without independent substantiating
investigation on the part of the potential users. Opinions expressed are those
of the authors and are not necessarily those of the Institution of Mechanical
Engineers or its publishers.
Printed by J W Arrowsmith Ltd, UK.

Related Titles of Interest


Title

Author

ISBN

History of Tribology

Duncan Dowson

1 86058 070 X

Lubrication and Lubricant


Selection

A R Lansdown

1 86058 029 7

Lubricants in Operation
(English translation edited
by A R Lansdown)

U J Moller and
U Boor

0 85298 830 3

A Tribology Casebook

J D Summers-Smith

1 86058 041 6

An Introductory Guide to
Industrial Tribology

J D Summers-Smith

0 85298 896 6

Guide to Wear Problems and


Testing for Industry

M J Neale and M Gee

1 86058 287 7

IMechE Engineers Data Book


Second Edition

Clifford Matthews

1 86058 246 6

A Guide to Presenting Technical


Information Effective Graphic
Communication

Clifford Matthews

1 86058 249 4

Journal of
Engineering Tribology

IMechE Proceedings,
Part J

ISSN 13506501

Professional Engineering Publishing (publishers to the Institution of Mechanical


Engineers) offers a valuable and extensive range of engineering books, conference and seminar transactions, and prestigious learned journals, as well as a
successful list of magazines. For the full range of titles published by Professional
Engineering Publishing contact:
Sales Department
Professional Engineering Publishing Limited
Northgate Avenue, Bury St Edmunds
Suffolk, UK
IP32 6BW
Tel: C44 (0) 1284 724384
Fax: C44 (0) 1284 718692
E-mail: sales@imeche.org.uk
www: pepublishing.com

This page intentionally left blank

Contents
Series Editors Foreword

xiii

Preface

xv

Notation

xvii

Chapter 1 Introduction to Rolling Contacts


1.1 Historical perspective
1.2 Practical use of rolling contact
1.3 Need to lubricate the rolling contact
1.3.1 Lubrication in the contact area
1.3.2 Reasons for lubrication
1.4 References
Chapter 2 Elements of Surface Contact of Solids
2.1 Introduction
2.2 Distribution of stresses within the contact zone
2.3 Deformations resulting from contact loading
2.4 Contact between bodies of revolution
2.4.1 Stress distribution within the contact
2.4.2 Contact with combined normal and tangential
loads
2.4.3 Three-dimensional contact
2.5 Contact of real surfaces
2.6 Criterion for deformation mode
2.6.1 Surface plastic deformations
2.7 Thermal effects during rolling
2.7.1 Moving source of heat
2.8 Contact of bodies with interposing film
2.8.1 Background to the analysis
2.8.2 Case of contacting cylinders
2.8.3 Contacting spheres
2.9 Crack formation in contacting elastic bodies
2.9.1 Description of the contact

1
1
2
5
5
7
9
11
11
12
18
19
21
23
24
30
34
35
38
39
40
41
42
45
45
46

viii

Rolling Contacts

2.10 Contacts deviating from the Hertz theory


2.10.1 Friction at the contact interface
2.10.2 Adhesion at the contact interface
2.11 References

50
50
51
54

Chapter 3 Fundamentals of Rolling Motion


3.1 General features of rolling contact
3.2 Source of friction in rolling contact
3.3 Rolling friction force
3.4 Free rolling
3.4.1 Cylinder on a plane
3.4.2 Sphere on a plane
3.5 Material damping during rolling
3.6 Slip at the surface of contact
3.7 Internal friction
3.8 References

55
55
58
60
62
62
64
65
68
72
74

Chapter 4 Dynamic Characteristics of Rolling Motion


4.1 Introduction
4.2 Analytical evaluation of friction torque
4.2.1 Friction during rolling
4.2.2 Friction torque in the rolling contact
4.2.3 Total friction torque
4.2.4 Variable components of friction torque
4.3 Elastic and damping characteristics of the rolling
contact
4.3.1 Static stiffness of the rolling contact
4.4 Dimensional accuracy and contact stiffness
4.4.1 Radial stiffness as a function of inaccuracies
4.4.2 Effect of variable dimensions and variable
stiffness
4.4.3 Effect of waviness of raceways
4.5 Ball motion in a rolling contact bearing
4.5.1 Inertia forces and moments acting on the ball
4.5.2 Relative motions of the rolling elements
4.5.3 Friction at the contact interface
4.6 References

75
75
76
76
79
88
91
106
106
115
116

Chapter 5 Rolling Contact Bearings


5.1 Phenomenology of friction during rolling
5.2 Friction torque
5.2.1 Friction coefficient

145
145
150
152

121
122
125
125
129
135
143

Contents

5.3 Contact stresses and deformations


5.3.1 Contact between elastic bodies
5.3.2 Elastic deformations in bearings
5.3.3 Permanent deformations
5.4 Load distribution within bearings
5.4.1 Radial bearings
5.4.2 Thrust bearings
5.5 Kinematics of bearing elements
5.5.1 Rotational speed of the elements and the cage
5.5.2 Contact cycles due to rolling
5.6 Inertia forces
5.6.1 Centrifugal forces
5.6.2 Crankpin bearings
5.6.3 Forces of gyration
5.7 Load-carrying capacity
5.7.1 Dynamic capacity
5.7.2 Static capacity
5.7.3 Equivalent bearing loads
5.8 Lubrication of bearings
5.8.1 Elastohydrodynamic lubrication
5.9 References

ix

153
153
157
160
163
163
167
167
167
171
171
171
173
174
175
175
181
184
195
196
198

Chapter 6 Rolling Contacts in Land Locomotion


6.1 Railwheel systems
6.1.1 Traction at the railwheel interface
6.1.2 Braking process
6.1.3 Traction enhancing techniques
6.1.4 Consequences of wheel and rail wear
6.1.5 Ribbed tyre
6.2 Tyreroad interactions
6.2.1 Relationship between friction and traction
6.2.2 Characteristics of the traction
6.2.3 Analysis of dry road traction
6.2.4 Traction under wet conditions
6.2.5 Analysis of wet road traction
6.2.6 Practical approach to traction modelling
6.3 References

201
201
202
205
205
207
207
210
211
214
218
222
226
234
236

Chapter 7 Machine Elements in Rolling Contact


7.1 Contact of meshing gears
7.1.1 Peculiarities of contact between gear teeth
7.1.2 Geometry of contact between gear teeth

239
239
239
241

Rolling Contacts

7.2 Friction in meshing gears


7.2.1 Tooth losses
7.3 Outline of elastohydrodynamic theory
7.3.1 Estimates of film thickness
7.4 Application of elastohydrodynamic theory to gears
7.4.1 Film thickness between gear teeth
7.4.2 Operating temperature
7.4.3 Oil viscosity in relation to surface condition
7.5 Boundary contact in gear lubrication
7.5.1 Running-in process
7.6 Scuffing in meshing gears
7.6.1 Flash temperature as a criterion for
scuffing
7.6.2 Phenomenon of scuffing
7.6.3 Probability of scuffing
7.7 Tooth face pitting
7.7.1 Fatigue fracture
7.7.2 Impact fracture
7.7.3 Tooth loading
7.8 Camfollower system
7.8.1 Reciprocating engine cam
7.8.2 Analysis of the follower motion
7.8.3 Tangent cam with a roller follower
7.8.4 Camshaft torque
7.8.5 Convex cam with a roller follower
7.8.6 General case of a convex cam with a roller
follower
7.8.7 Convex cam with a flat follower
7.8.8 Stresses within the camtappet contact
7.8.9 Lubrication of the camtappet contact
7.8.10 Design considerations
7.9 References
Chapter 8 Non-metallic Rolling Contacts
8.1 General considerations
8.1.1 Approaches to polymer fatigue
8.1.2 Loading conditions in rolling contact
8.2 Phenomenology of polymer fatigue
8.2.1 Physical states of stressed polymers
8.2.2 Response to applied stress
8.2.3 Phenomenological description of fatigue

243
246
248
251
252
253
257
259
261
262
263
263
266
268
274
279
280
281
285
286
289
291
297
300
302
306
309
312
313
314
317
317
319
319
320
321
323
324

Contents

8.3 Behaviour of polymers in rolling contact


8.3.1 Characteristics of rolling contact conditions
8.3.2 Thermodynamic equilibrium in rolling contact
8.3.3 Mechanics of polymer rolling contact
8.3.4 Fatigue considerations
8.4 Model rolling polymer contact
8.4.1 Experimental setting
8.4.2 Kinematics of the model contact
8.4.3 Performance of some polymers in rolling contact
8.5 Technical ceramics in rolling contact
8.5.1 Ceramic materials
8.5.2 Ceramic bearings
8.5.3 Manufacture of silicon nitride balls
8.5.4 Dimensional quality of ceramic bearing
components
8.5.5 Material quality
8.5.6 Surface quality
8.5.7 Failure modes of ceramics in rolling contact
8.6 References

xi

329
330
331
337
340
344
344
345
348
350
350
352
356
357
358
359
361
362

Chapter 9 Coated Surfaces in Rolling Contact


9.1 Introduction
9.2 Coating processes
9.2.1 Thermal spray
9.2.2 Electroplating
9.2.3 Physical vapour deposition
9.2.4 Chemical vapour deposition
9.3 Application of coatings to rolling contact elements
9.3.1 Rollers for steel forming
9.3.2 Rollers for papermaking and printing
9.3.3 Fracture of coatings during rolling
9.4 References

365
365
365
365
372
377
379
383
383
385
388
389

Chapter 10 Rolling in Metal Forming


10.1 Introduction
10.2 Forces acting in the contact region
10.2.1 Forces acting in the roll gap
10.2.2 Neutral point and no-slip angle
10.2.3 Expressions for the no-slip angle
10.2.4 Maximum value of the no-slip angle
10.2.5 Rolling when bar motion is impeded

391
391
392
392
395
397
398
399

xii

Rolling Contacts

10.3 Forward slip during rolling


10.3.1 Introduction
10.3.2 Expressions for forward slip
10.4 Friction between the rolls and the material
10.4.1 Role of friction in rolling
10.4.2 Friction in hot rolling
10.4.3 Friction in cold rolling
10.4.4 Evaluation of friction measuring methods
10.4.5 Coned compression tests
10.4.6 Friction coefficient variation along the arc of
contact
10.5 Theories of metal rolling
10.5.1 Introduction
10.5.2 Equation describing the friction hill
10.5.3 Theory of von Karman
10.5.4 Simplification of von Karmans equation
10.5.5 Modification of von Karmans equation
10.5.6 Effect of front and back tension on the pressure
distribution
10.5.7 Shortfalls of rolling theories
10.5.8 Theory of rolling by Orowan
10.5.9 Variations of the Orowan equation
10.6 Discussion of metal rolling theories
10.7 References
Index

400
400
400
404
404
405
407
408
409
412
416
416
417
421
423
424
426
428
429
437
438
439
441

This page intentionally left blank

Series Editors Foreword


Almost every aspect of industry and many fields of science and medicine
require a practical understanding of tribology. The Tribology in Practice
Series aims to be accessible and industry-relevant bringing the latest
developments in tribological research, as well as established ideas and
techniques, to bear on real engineering problems.
The Series Editors are pleased to introduce Rolling Contacts by T. A.
Stolarski (Brunel University, UK) and S. Tobe (Ashikaga Institute of
Technology, Japan) as the second title in this Series. This book is a
rigorously researched monograph on rolling contacts covering all
aspects of their use. It covers the fundamentals of rolling contacts of
all types with particular attention paid to important engineering
applications including rolling bearings, gears, railwheel and road
tyre interactions, camtappet systems, as well as the roll-forming of
materials. We are sure that this comprehensive treatment of the subject
will be a valuable addition to the available literature and will prove
useful to researchers, practising designers, and students.
As a Series, Tribology in Practice is particularly concerned with
design, failure investigation, and the application of tribological understanding to the products of various industries and to medicine. We are
also including reviews of areas of research in the subject, which will be
useful both for research workers and engineers in industry.
The link with the IMechE through its publishing operation, Professional Engineering Publishing, is appropriate as tribology has its
natural home within this branch of engineering. The scope of the series
is as wide as the subject and applications of tribology. Wherever there
is wear, rubbing, friction, or the need for lubrication, then there is scope
for the introduction of practical, interpretative material. The Series
Editors and the publishers would welcome suggestions and proposals
for future titles in the Series.
M J Neale
Neale Consulting Engineers, UK
T A Polak
Neale Consulting Engineers, UK
C M Taylor
University of Leeds, UK

Preface
Rolling friction is a very old problem in engineering and undoubtedly
one of the most important from a practical point of view. According to
estimates, the losses in the United Kingdom resulting from friction and
wear related problems amount to 500 million annually. The availability of reliable, low-friction rolling contacts has become an important
factor in the development of micro-machines and miniaturization in
general. Many devices in high-precision engineering applications, such
as magnetic storage and recording systems, miniature motors, laser
scanners, machine tools for micro- and nano-level machining, and scanning microscope techniques, require bearings with extreme accuracy of
motion. Undoubtedly, the modern high-tech world depends upon and
demands tribological systems of the highest quality. Despite this, many
aspects of rolling friction are still not entirely understood, and research
into mechanisms and processes governing the operation of rolling contacts at an atomic level is just starting to emerge.
This book presents a general introduction to the fundamentals of
rolling friction with the emphasis on important engineering applications
of rolling contacts. Usually, a rolling contact is taken to be synonymous
with a rolling contact bearing. This, however, is not necessarily true as
there are a number of technologically important applications, such as
gears, roadtyre and camtappet systems, and roll-forming of materials,
where rolling contact configuration is at the heart of the matter. Analytical treatment of the topics discussed, wherever feasible, was considered to be of prime importance and, in the majority of cases, this
was achieved. It is very much hoped that the procedures and techniques
of analysis presented in this book will be found useful through
improved understanding, selection, and design of rolling contacts for
mechanical devices and systems. It is also hoped that the book will be
seen as a comprehensive monograph on rolling contacts in all aspects
of their utilization. Therefore, it should prove useful to practising
designers, researchers, and postgraduate students. Students on engineering degree courses in universities should also benefit from this book, as
it will give them an introduction to rolling contacts that are commonly
used in engineering.

xvi

Rolling Contacts

Many years of research collaboration between the authors provided


the inspiration to write this book. This was a natural progression, as
the subject matter of the book is firmly rooted in their research interests.
The material presented is grouped according to leading themes: sources
of rolling friction, mechanics of contact between solid bodies, dynamics
of rolling motion, land locomotion, rolling contact bearings, gears, the
camfollower system, non-metallic rolling contacts, coated surfaces in
rolling contact, and rolling in the metal forming process. Chapter 1
presents a general introduction to the area of rolling contacts, with
some information about the history of development of rolling contacts.
Chapter 2 deals with elements of surface contact of solids. It is by no
means a comprehensive treatment of the subject as there are specialist
monographs that focus on contact mechanics. Nevertheless, the information contained in this chapter is considered to be adequate for proper
understanding of the problems involved in contact between solids.
Chapter 3 is an attempt to explain the origin and sources of friction
during rolling motion. The dynamic characteristics of rolling motion, a
topic frequently neglected in design practice, is discussed in Chapter 4.
Rolling contact bearings are introduced in Chapter 5. As this is a topic
dealt with in almost all books on tribology, it is treated with the
assumption that a fairly good understanding of problems pertaining to
rolling contact bearings already exists in the engineering community.
Nevertheless, some topics such as inertia forces, although important in
practice, are felt to be less publicized and therefore are presented more
thoroughly. Chapter 6 is devoted to land locomotion where two applications of rolling contact are of prime importance, i.e. the railwheel
system and tyreroad interaction. Both topics are treated with sufficient
depth to allow understanding of the problems involved. Gears and
camfollower systems are commonly encountered machine elements
where rolling motion plays an important role, and these are introduced
in Chapter 7. Non-metallic rolling contacts, both polymeric and ceramic, are discussed in Chapter 8. In order to meet ever-increasing
demands for better performance, surfaces in rolling contact are coated.
Chapter 9 presents the various coating techniques available and investigates how coating can improve the performance of a rolling contact.
Finally, Chapter 10 deals with rolling in metal forming an important
area of engineering.
The authors would like to thank Ms Sheril Leich, Commissioning
Editor, Professional Engineering Publishing, for facilitating the project
and for having confidence in them.

Notation
a
A
c
E
E
Fe
Fr
G
H
l
M
n
N
p
pm
po
R
R
t
T
Tf
W
V
Y

width of contact
area of contact
radial clearance
modulus of elasticity
equivalent modulus of elasticity
axial (thrust) load
radial load
shear modulus
identation hardness
contact length
moment (torque)
speed of rotation
normal load
contact pressure
mean contact pressure
maximum contact pressure
radius of curvature
equivalent radius of curvature
traction
tangential load
flash temperature
power
velocity
tensile yield strength

thermal conductivity
asperity radius
surface energy
normal approach
strain
coefficient of friction
coefficient of rolling friction
Poissons ratio
density

xviii

max
(z)

Rolling Contacts

normal stress
tangential stress
maximum tangential stress
distribution of peak heights
plasticity index

Chapter 1
Introduction to Rolling Contacts

1.1 Historical perspective


Technological progress usually leads to increasing demands on all elds
of design and manufacture, and consequently on the design and application of contacts in relative motion supporting load. The solution of
the diversied contact problems involved occasionally requires a
detailed knowledge of friction and wear mechanisms and theories on
the part of the designer. This knowledge is a part of an ever-growing
area called tribology. The interdisciplinary nature of tribology, with
knowledge drawn from different disciplines such as mechanical engineering, materials science, chemistry, and physics, leads to a general tendency for the chemist to describe in detail, for instance, lubricant
additives, and for the mechanical engineer to discuss, for example, sliding journal bearings, with no overall guide to the subject. Also, it is
difcult to nd, in books dealing with tribology problems in general, a
focused and advanced treatment of certain practically important topics
in a comprehensive and thorough way. This is certainly true with rolling
contacts the subject of this book.
It is quite probable that primitive man at a very early stage in the
development of civilization discovered that it is far easier to move a
heavy object over the ground by placing it on logs and rolling rather
than sliding it. Though refuted by some scholars on the grounds of lack
of positive archaeological evidence, it is difcult to see how the wheel
with its axle and sliding bearings could have been developed other than
from logs. It could thus be argued that experience of the rolling contact
bearing predates that of the sliding bearing.

Rolling Contacts

The ancient Egyptians and Greeks are believed to have made effective
use of the principle of the rolling contact, and a reference exists to a
rolling bearing devised by the Greek Diades in 330 B.C. for a battering
ram which incorporated the essential principles of a rolling bearing as
made at the present time. Fragments of what appears to resemble a
ball thrust bearing were found in Lake Nemi, Italy, in 1928 (1). It was
speculated that it was used to support a rotatable statue and may have
been made about 12 A.D. Leonardo da Vinci (2) studied, among other
things, the differences between sliding and rolling, but this aspect of his
work was not generally known until a publication that appeared in the
late nineteenth century.
On the British scene, an iron ball thrust bearing with many design
characteristics of a modern bearing made its appearance about 1780 for
use in a post mill in the Norwich area. A book published by Varlo in
1772 (3) describes a ball bearing he designed and tted to his postchaise. British Patent 1580 was assigned to John Garnett of Gloucester
in 1787 for interesting arrangements of various types of rolling element
to form a bearing. In 1794, British Patent 2006 was granted to Philip
Vaughan of Carmarthen for a radial ball bearing, the rst of its kind
on record.
Important engineering developments of the rolling bearing continued
in the early and middle part of the eighteenth century, but the main
impetus that led to the foundation of the rolling bearing manufacturing
industry came from the invention of the bicycle in Scotland in about
1840.
In 1881, Heinrich Hertz (4) published in Germany his study on deformation of curved elastic bodies in contact, providing the growing industry with a mathematical theory that is used to the present time.
Other than papers that have since been presented to learned bodies
and other institutions, there appears to be very little literature in English
that describes the development of the rolling bearing industry and provides information on the complex technology of design and production
developed in the closing years of the nineteenth century and the rst
half of the twentieth century.

1.2 Practical use of rolling contact


Though much study seems to have been devoted to rolling resistance,
the basis of rolling itself seems to have been given comparatively little
place in the literature. It may be helpful to understanding rolling motion
if an ideal concept is visualized, a perfect cylinder on a perfect plane,

Introduction to Rolling Contacts

Fig. 1.1

both made of the same rigid, inelastic, frictionless material. If the cylinder is made to rotate about its own long axis, theoretically it might
continue to rotate indenitely. If pushed along the plane it might slide.
If the coefcient of friction of the surface is then assumed to be raised
by some means and the hypothetical experiment repeated, the cylinder
should roll as soon as the value of static friction between roller and
plane exceeds the value of the force previously applied to the cylinder,
causing it to rotate or slide. The static friction may be regarded as a
force acting in the opposite direction to the applied force, thus creating
a couple. Since the same friction prevents it from sliding or rotating, it
must roll. This is illustrated in Fig. 1.1. If the applied force is greater
than the static friction the roller may rotate or slide as in the case of
the locomotive wheel. The signicance of the concept will not be altered
if a sphere is substituted for the simpler case of the cylinder. If the case
of a cylinder and plane both made of an elastic material such as steel
is then considered, when the system is at rest, the metal in the contact
area, as shown by Hertz, is deformed elastically under load (Fig. 1.2).

Fig. 1.2

Rolling Contacts

Application of a force tending to push the cylinder along the plane will
cause it to start to roll owing to static friction. The displaced metal then
forms waves preceding and following the rolling cylinder. Differences
in the rate of recovery of the deformed area, because of elastic hysteresis, lead to an imbalance that produces a couple acting on the roller,
causing it to continue to roll. This mechanism would theoretically apply
even if the cylinder and plane were frictionless but elastic. In a rolling
bearing, for example, the cylinder or sphere is not freely rolling as in
the hypothetical case described but is in contact with two surfaces, one
of which normally is stationary and the other, in motion, may be
regarded as providing the force causing the rolling element to move.
This, of course, is an oversimplication. However, what is often referred
to as rolling friction is really rolling resistance, which will be considered
later.
The widespread use of the rolling contact principle in industry is, in
large part, due to the lower power losses expected in rolling contacts
compared with sliding contacts. However, in the practical situation, a
signicant amount of sliding motion occurs and it is important to consider this. The situation in the area of contact between rolling element
and track is then examined rst.
Hertz was concerned with the deformation of curved elastic solids in
dry contact and calculation of the stresses thereby created. Figure 1.2(a)
shows the pattern of deformation and stress he proposed for a stationary cylinder resting on a at surface, while Fig. 1.2(b) shows the change
in the stress pattern created by rolling the cylinder under load. Some
modication of the deformation pattern occurs when the surfaces are
separated by a lm of lubricant. Since steel is highly elastic, and provided the elastic deformation limit is not exceeded, recovery is almost
instantaneous, but the rate varies to the extent that the elasticity is
imperfect; other factors are also involved. The hysteresis effect produced by differences between deformation rate and recovery time
accounts in large measure for resistance to rolling, while the repeated
stress and relaxation cycles themselves in a rolling contact have the
major inuence on its fatigue life.
A perfect cylinder stationary on a perfect plane has contact on a
mathematical line if no load is applied. On the application of load, the
projected area of contact is a rectangle increasing in size according to
the load. Similarly, a perfect sphere would have point contact under noload conditions, but the projected deformation area would be a circle of
size increasing with load. If a force acting parallel to the plane is applied
to a cylinder or ball under loaded conditions, it will roll along the plane,

Introduction to Rolling Contacts

but the rolling cannot be perfect and a degree of slip must occur, producing some resistance to motion owing to increased friction.
Osborne Reynolds (5) studied the nature of rolling even before he
made his classic study of lubricated sliding. Using the simple system
consisting of a cylindrical roller on a plane, Reynolds proposed a theory
of rolling resistance due to microslip. In a more recent publication,
Tabor (6) showed that the apparent slip observed by Reynolds was due
to unequal stretching of the surfaces. Tabor also argued that resistance
to pure rolling is due largely to the elastic hysteresis described earlier.
Heathcote (7) demonstrated that any departure from mathematical
straight line contact towards curvature introduces an element of slip
arising from the variations in circumferential speed. In practice, this
situation always applies even to a cylinder on a at track, since neither
is perfect in a mathematical sense.
Any force operating to displace a ball or roller from its true rolling
path will also cause slip. Surface nish inevitably affects the amount of
friction when slip occurs and must therefore be included in the factors
creating rolling resistance.
Thus, it is possible to summarize the main components of resistance
to rolling of the element on the track as: elastic hysteresis which will
be affected by the properties of the materials; temperature; load and
frequency of the stress relaxation cycle, that is, rolling speed; the shape
and surface nish of the contacting surfaces; the effect of any deviation
from the rolling path.

1.3 Need to lubricate the rolling contact


1.3.1 Lubrication in the contact area
Taking the hypothetical case introduced earlier of an ideal system consisting of an inelastic true cylinder of perfect surface nish on a true
plane of the same material, there would be mathematical line contact,
no deformation, and rolling should take place without slip or wear.
There must be sufcient static friction or interaction of surface force
elds between the roller and plane to promote rolling. This might of
course be inconsistent with perfect surface nish and the ideal system
represents only a concept. In practice, all metal surfaces exhibit some
surface roughness and, with exaggeration, the plane and roller may be
conceived as resembling a crude rack and pinion (Fig. 1.3).
If a lubricant is introduced at the contact and the roller is rotated on
its own long axis, hydrodynamic lubrication may be established. For
hydrodynamic conditions to be established, laminar ow must take

Rolling Contacts

Fig. 1.3

place within the lubricant lm. Since the rolling element may be considered as being rotated by tangential forces from the two tracks transmitted through the lubricant lm, slip must then occur in such a case,
as a hydrodynamic lm cannot transmit forces parallel to its motion
without slip.
Archbutt and Deeley (8) suggested that a hydrodynamic wedge is
created between roller and track. Purday (9) made a mathematical study
of hydrodynamic conditions which suggested they could exist between
a roller and plane. Osterle (10), working with roller bearings and using
Purdays analysis, conrmed the existence of hydrodynamic conditions
between roller and track. Smith (11), studying wear problems in roller
bearings on the main shafts of aircraft gas turbines, demonstrated that
rollers and cage could travel at different speeds. He found, by using a
roller with magnetic inserts, that slip between the rotating inner race
and the roller exceeded that between the roller and the xed outer race.
Sudden increase in radial load could cause breakdown of the hydrodynamic lm with resultant wear if slip continued. It is possible that the
somewhat unexpected experimental results reported by Fogg and
Webber (12), working with cageless roller bearings at high speeds, may
also be explained by postulating some form of hydrodynamic conditions both between rollers and tracks and between the rollers themselves. They found that the cageless bearings, which were made by
removing the cage from a standard bearing and increasing the number
of rollers, operated with lower torque and no increase in friction, compared with the standard caged bearing. Lubrication was by oil mist in
both cases.
The conclusions of Palmgren and Snare (13) about bearing behaviour
in conditions of high speed and no load require some qualication in
what must be taken as the normal case. They suggested that the regime
changed from hydrodynamic to boundary lubrication when speed was
reduced and load increased. However, since Grubin (14) published
his work on the elastohydrodynamic theory of lubrication, the entire
concept of lubricant behaviour in rolling contacts has changed.

Introduction to Rolling Contacts

1.3.2 Reasons for lubrication


Before considering the signicance of the elastohydrodynamic theory,
it may be useful to review the reasons for lubricating rolling contacts
at all. If, in the hypothetical simple case, rolling takes place in the dry
state without deformation or sliding, and no wear occurs, lubrication is
unnecessary. Even in a practical bearing, Goodman (15) claimed to
have shown that introduction of lubricant into a well-designed rolling
bearing increased friction. However, the potential advantage of lubricated rolling contacts over sliding contacts in respect of friction is illustrated by the following typical ranges of friction coefcient.
Fully lubricated sliding bearing
Lubricated rolling bearing

Static
0.10.3
0.0020.005

Kinetic
0.0010.005
0.00100.0018

Rosenfeld (16) stated that lubrication may greatly increase frictional


resistance, but ball and roller bearings cannot operate for any considerable time without a lubricant. Rare exceptions may exist when it is
expedient to operate a rolling bearing in a dry condition and the
resulting wear is tolerable, but, in general, without lubrication the bearing life would be unacceptably short.
The main reasons for lubricating rolling contact bearings can be summarized as follows:
(a) to prevent metalmetal contact between races and rolling elements
at points of sliding;
(b) to eliminate any harmful effects of surface irregularities that cannot
be completely removed even by the most careful polishing;
(c) to support the sliding contact between the cage and the rolling
elements andor race shoulder;
(d) to carry away the heat developed in the bearing;
(e) to protect the highly polished surfaces from corrosion.
Thus, the main functions of a lubricant can be classied as lubrication,
heat transfer, and protection.
It has already been noted that some sliding takes place between the
rolling element and the running track in commercial rolling bearings
and, if diametral clearance and relaxation of load encourage it, a hydrodynamic regime is established. However, in normal loaded conditions
the preponderant function of the lubricant in this area according to the
elastohydrodynamic theory is the generation of a lm between rolling
element and track as a result of deformation. The shape of this is

Rolling Contacts

Fig. 1.4

shown schematically in Fig. 1.4 with the theoretical pressure distribution. In practice, as the pressure in this region is very high, possibly
1.53.0 GPa, the viscosity of this lm is greatly increased. Therefore, to
the extent that such a regime prevails in the contact area, it would
seem likely that this lm, momentarily of such high viscosity as to be
comparable with dry static friction, transmits the elements of the couple
which produces rolling. At the entry to the region the lubricant is under
shear and may be regarded as behaving hydrodynamically, but within
the contact zone negligible slip occurs.
Clearly, the pressureviscosity characteristics of lubricants are of
importance in the behaviour of such elastohydrodynamic lms. Usually,
a close agreement between measured thickness of an oil lm and predictions of the theory is found. Where anomalies are found, notably in the
case of a silicone uid and a solution of polymethylmethacrylate in
oil, the possible effect of non-Newtonian behaviour is thought to be
responsible.
Where full hydrodynamic or elastohydrodynamic conditions can be
maintained, the full fatigue life expectancy of the rolling contact may
be achieved. If boundary conditions prevail even for part of the time,
fatigue life may become unpredictable.

1.4 References
(1) Cellini, B. (1949) Autobiography The Life of Benevenuto Cellini,
(Phaidon Press, London).

Introduction to Rolling Contacts

(2) Da Vinci, L. Codie Atlanticus (British Museum Library, Milan).


(3) Varlo, C. (1772) Reections upon friction with a plan of the new
machine for taking it off in wheel carriages, windlasses of ships
etc., London.
(4) Hertz, H. (1896) The Contact of Elastic Bodies (Macmillan,
London).
(5) Reynolds, O. (1875) On rolling friction. Phil. Trans. R. Soc., 166.
(6) Tabor, D. (1954) The mechanism of rolling friction. Phil. Mag.,
45.
(7) Heathcote, H. L. (192021) The ball bearing in the making, under
test and on service. Proc. Instn Auto. Engrs, 15.
(8) Archbutt, L. and Deeley, R. M. (1927) Lubrication and Lubricants
(Grifn, London).
(9) Purday, H. F. P. (1949) Streamline Flow (Constable, London).
(10) Osterle, J. F. (1959) On the hydrodynamic lubrication of roller
bearings. Wear, 2.
(11) Smith, C. F. (1962) Some aspects of the performance of high-speed
lightly loaded cylindrical roller bearings. Proc. Instn Mech. Engrs,
176.
(12) Fogg, A. R and Webber, J. S. (1955) The inuence of some design
factors on the characteristics of ball bearings and roller bearings
at high speeds. Proc. Instn Mech. Engrs, 169.
(13) Palmgren, A. and Snare, B. (1957) Inuence of load and motion
on the lubrication and wear of rolling bearings. In Proceedings of
IMechE Conference on Lubrication and Wear, London.
(14) Grubin, A. N. and Vinogradova, I. E. (1994) Book No. 30 (Central
Scientic Research Institute for Technology and Mechanical
Engineering, Moscow).
(15) Goodman, J. (1912) Roller and ball bearings. Proc. Inst. Civ.
Engrs, 189.
(16) Rosenfeld, L. (1942) Friction of ball and roller bearings. Instn
Auto. Engrs paper 19426.

This page intentionally left blank

Chapter 2
Elements of Surface Contact
of Solids

2.1 Introduction
This chapter deals with the stresses and deformation resulting from the
contact of the surfaces of two solid bodies. Usually, conforming and
non-conforming contacts can be distinguished. A contact is dened as
conforming when the surfaces of the two bodies t exactly together
without deformation. An example of a conforming contact is a journal
bearing and a thrust bearing. A non-conforming contact is formed by
bodies that have different proles. Depending on the overall contact
geometry, they will form a point contact or a line contact. A ball bearing represents the case of point contact because the ball makes point
contact with both raceways. On the other hand, in a roller bearing, the
roller makes line contact with raceways. In general, line contact is
created when the proles of the bodies are conforming in one direction
and non-conforming in the perpendicular direction. The area of nonconforming contact is usually small compared with the dimensions of
the bodies in contact. The stresses are highly concentrated in the region
close to the contact zone and are not signicantly inuenced by the
shape of the bodies at a distance from the contact area.
In engineering applications, the points of surface contact are quite
often executing complex motions and are required to transmit both
forces and moments. For instance, the point of contact between a pair
of gear teeth moves in space, while, at the same time, the two surfaces
move relative to each other at that point and the motion combines both

12

Rolling Contacts

rolling and sliding. In this chapter a frame of reference will be dened


in which the motions and forces that arise in any particular circumstances can be generalized. In this way, the problems of contact can be
formulated and studied independently of any application context. This
approach also facilitates the application of the results of such studies
to the wide variety of engineering problems.

2.2 Distribution of stresses within the contact zone


It should be stated right at the beginning that the stresses acting within
the outermost layers of material (typically up to a depth of 1 mm or so)
will mainly be considered here. The effects at several centimetres below
the surface are of only secondary importance. For that reason, it is
permissible to treat the surfaces as though they represent the surface of
bodies of innite depths; i.e. they may be considered as semi-innite
bodies. This approach facilitates the analysis of details of the surface
contact of solids rather than considering their overall geometrical shape
and thus leads to considerable mathematical simplication.
Stresses produced by a single normal line load P per unit length in
the plane xz and applied at a point O and dened by the coordinates
, O on the surface (zG0) of a semi-innite body and having the same
value for all values of y are shown in Fig. 2.1(a). The elastic stress eld
in the plane xz is easily obtained. Considering a unit length in the y
direction, the radial stress r is given by
2P
r G
(2.1)
cos
r
as the tangential stress and the shearing stress r are equal to zero.
The case depicted in Fig. 2.1(a) represents a state of simple radial
compressive stress. The stress increases with decreasing radius r and
decreasing angle . The use of the two-dimensional Mohrs circle of
stress, shown in Fig. 2.1(b), for these stresses gives the resulting
Cartesian stresses with respect to O
r
2P
zx2
x G (1Acos 2 )G r sin2 G
2
(x2Cz2)2

2P
r
z3
(1Ccos 2 )G r cos2 G
2
(x2Cz2)2
2P
r
z2x
xz G sin(2 )G r sin cos G
2
(x2Cz2)2

zG

(2.2)

Elements of Surface Contact of Solids

Fig. 2.1

13

14

Rolling Contacts

Equations (2.2) can be written with respect to the origin O as

X G

2P

Z G

2P

Z(XA )2

Z3

[(XA )2CZ 2]2

[(XA )2CZ 2]2

XZ G

Z 2(XA )
2P
[(XA )2CZ 2]2

(2.3)
In a similar way, the stresses due to a single tangential line T acting at
O (Fig. 2.2) can be obtained
2T
cos
r G
r
G r G0
(2.4)
and
Z 2(XA )
2T
[(XA )2CZ 2]2

(XA )
2T
G

(XA ) CZ ]
Z(XA )
2T
G

[(XA ) CZ ]

X G

2 2
2

XZ

2 2

(2.5)

Fig. 2.2

Elements of Surface Contact of Solids

15

Taking into account the fact that TGP, where is the appropriate
friction coefcient, and adding the stress components due to P and T
at any point (x, y), the stress distribution arising in a simple frictional
contact can be obtained. Examination of equations (2.1) and (2.4)
reveals that at O (rG0) the stresses are innite which in practice is
unacceptable. This is due to the assumption that the load acts at a single
point, i.e. over zero contact area. In reality, there is always some nite
area of contact associated and this changes the formulation of the initial
problem.
Figure 2.3 shows a uniformly distributed load producing a contact
pressure p over the region Oa on the surface (zG0) of a semi-innite
body. Taking the length along the y direction to be equal to unity, the
external load itself is given by
a

p dxGpa

PG

In the case of a very small load p d at some point dened by the


coordinates ( , O), it is possible to nd the stress at any point (X, Z)
that is caused by this load using equations (2.3). In this case, P will be
replaced by p d . The total stress at a point X, Z that is due to the
distributed load P is then obtained by the summation of the effects of
all the p d loads acting at different values of from O to a, or, in

Fig. 2.3

16

Rolling Contacts

mathematical terms

X G
Z G

[(XA ) CZ ] d
2

2 2

[(XA ) CZ ] d

2p

XZ G

Z(XA )2

2p

Z3
2

2 2

2p

Z 2(XA )

[(XA ) CZ ] d
2

2 2

(2.6)
If a tangential load TG P acts over the region Oa (Fig. 2.4), then at
every point it follows that t dxGp dx, and thus

TG

p dxGP

t dxG

Using equations (2.5) for each elemental tangential load t d acting on


element d (O, ), the stresses at any point (X, Z) that are due to the
total distributed load T can be obtained

X G

2t

2t
Z G

XZ G

Z 2(XA )

[(XA ) CZ ] d
2

2 2

(XA )3

[(XA ) CZ ] d

2t

2 2

Z(XA )2
d
[(XA )2CZ 2]2

(2.7)

Fig. 2.4

Elements of Surface Contact of Solids

17

The total stresses for a sliding contact subjected to a normal load P


uniformly distributed over the contact region Oa are given as the sum
of the stresses dened by equations (2.6) and (2.7). It is obvious that
the solutions for the normal and tangential point load may be used to
obtain the resultant stress distribution for any type of load over the
contact region. In the above solutions, the basic assumption is that of
elastic behaviour of the bodies in contact. However, in reality the
possibility of plastic effects must be taken into account. The simplest
criterion dening the onset of plastic deformation assumes that this
occurs when the maximum shear stress attains the critical level, k, for
the material, where kGY2 (Y denotes the tensile yield stress).
Under plane strain conditions, the maximum shear stresses always
occur in the xz plane. The maximum shear stress in this plane is simply
the radius of the Mohrs circle of stress, which is shown in Fig. 2.5.
Thus

max G

P
r
G cos
2
r

By drawing a circle of diameter b in the way shown in Fig. 2.5(a), it


can be found that rGb cos and

max G

P
b

which means that the stress remains constant at all points on the circle.
It is therefore benecial to plot the stress distribution as isochromatics

Fig. 2.5

18

Rolling Contacts

or lines of constant max . It is then possible to determine the location


at which max will reach its limiting value of k, i.e. the location of the
onset of plastic deformation. These diagrams are useful since they also
indicate the pattern of isochromatics obtained in photoelastic stress
studies. In the case of a point normal load and a uniformly distributed
normal load, analysis of max produces the pattern of isochromatics
shown in Fig. 2.5(b) and (c). It is clear that in both cases the material
will rst reach a yield condition at the surface where increasing load
gives max Gk, which denotes the yield strength of the material.

2.3 Deformations resulting from contact loading


The logical step from assessing the contact stresses is to examine the
displacements in a solid using the known relations between stress and
strain. Thus, for a single normal load P acting at O [Fig. 2.1(a)], the
horizontal and vertical displacements u and w respectively are
u

1
2P
Ger G ( rA )G
cos
r
E
rE

u w
2P
1
C
cos
Ge G ( A r )G
r r
E
rE
r

u w w
1
C A G r G r G0
r r
G

The solution of the above equations requires information on the boundary conditions. For this, it can be assumed that points on the z axis,
i.e. at G0, have no lateral displacements and that at a point on the z
axis at a distance b from the origin there is no vertical displacement.
The displacements occurring at the boundary zG0 are of interest. Thus,
by putting GJ2 in the solution of the above equations it can be
shown that the horizontal displacement is given by
(u)z G0 G

(1A)P
2E

(2.8)

This indicates that at all points on the boundary of the solid there is a
constant displacement directed toward the origin. Also, it is possible to
nd the vertical displacement of a point on the boundary zG0 at a
distance x from the origin
(w)z G0 G

2P
b (1C)P
log A
E
x
E

(2.9)

Elements of Surface Contact of Solids

19

At the point of load application (xG0) the displacement in the vertical


direction tends to innity. This is the result of the assumption of a point
load which, in reality, is not valid as the load is usually distributed over
a small nite area. If the load is distributed over the region Oa, as
shown in Fig. 2.3, giving rise to a constant pressure p, the vertical displacement at any point with coordinates X, O produced by an element
of load p d at a distance from point O is known from equation (2.9).
It can be estimated by substituting p d for P and (XA ) for x so that
the total displacement at point (X, O) is given by
(w)z G0 G

2
E

p log

b
XA

d A

(1C)
E

p d

(2.10)

All the above solutions are valid for two-dimensional problems only.
The three-dimensional problems are far more complex and their
detailed treatment can be found in standard books on the theory of
elasticity.

2.4 Contact between bodies of revolution


The contact between bodies whose geometry is dened by circular arcs
is essential for rolling contacts. Hertz (1) was the rst to solve this type
of problem for elastic bodies and, for that reason, this contact is known
as Hertzian contact.
Figure 2.6 illustrates the contact of two identical cylinders under conditions of plane strain. Utilizing the symmetry of the contact, it can be

Fig. 2.6

20

Rolling Contacts

argued that the zone of contact is created by compression of the cylinders to generate a straight line, i.e. to produce a plane contact zone
[Fig. 2.6(a)]. Although this is not strictly true for a cylinder in contact
with a plane, the error is small and can be neglected. Thus, a plane
contact zone may be assumed.
When two identical elastic cylinders are in contact under a normal
load P per unit axial length, the resulting plane contact zone has a
width of 2a [Fig. 2.6(b)]. The normal deformation at the centre of the
contact zone is greater than at the extremities, and the contact pressure
distribution p is given by
pG

2P
a

x2
1A 2
a

(2.11)

Using a simple physical argument, it can be shown that


Stress T(Pa)
Considering the deformation, it is justied to say that increasing the
load increases a and, thereby, increases the strain. Thus
Strain T(aR)
where R is the radius of the cylinder. From the above two relations

a
P
TE
a
R
or
a2T

PR
E

The analytical solution for this case gives


2

a G

4PR(1A 2 )
E

(2.12)

The solution dened by equations (2.11) and (2.12) is approximately


true for other identical cylinders, i.e. plane contact geometries. Provided
that the angle subtended by the contact width at the centre of the cylinder is less than 30, the results may be used for other contact geometries

Elements of Surface Contact of Solids

21

Fig. 2.7

shown in Fig. 2.7 by using the equivalent modulus of elasticity E and


the equivalent radius of curvature R. Thus
1
E

1A 21 1A 22
C
E1
E2

1
1
C
R1 R2

and
1
R

and nally
a2 G

4PR
E

In the case of contact between a cylinder and a plane, the radius of the
plane is taken as innity. Therefore, R becomes the radius of the cylinder only and for concave curvatures the radius is taken as negative. It
is important to note that when E S, the solids become rigid, resulting
in a single point contact where a 0.

2.4.1 Stress distribution within the contact


It was argued earlier in this chapter that the onset of plastic deformation may be associated with the maximum shear stress reaching a
critical value k. Therefore, it is of practical importance to examine the
distribution of the maximum shear stress for a contact loaded by the
pressure prole given by equation (2.11) and acting over the region
from Aa to Ca. Using equations (2.6) for elemental loads p d and
integrating for the actual distribution of P will result in the Cartesian

22

Rolling Contacts

stress distribution within the body in contact. Thus


0

X GA

Z(XC )2
d AA
[(XC )2CZ 2]2

Z GA

[(XC ) CZ ]
Z3

XZ GA

2 2

Z 2(XC )
2

2 2

d AA

[(XC ) CZ ]
B

Z(XA )2
d
[(XA )2CZ 2]2

[(XA ) CZ ] d
B

Z3
2

2 2

d AA

Z 2(XA )

[(XA ) CZ ] d
B

2 2

(2.13)
where
AG

4P
2a

BG

2
a2

1A

The maximum shear stress for plane strain conditions is given by the
radius of Mohrs stress circle, i.e.

max G

XA Z 2
C 2XZ
2

(2.14)

where X, Z, and XZ are dened by equations (2.13). Therefore, equation (2.14) denes the values of max at all points. Equation (2.14) can
be used to draw the isochromatics, from which it can be seen that the
greatest value of max occurs below the surface at a distance of 0.67a.
Besides, as the load is increased, max at this point also increases,
attaining the value k when the maximum pressure at the centre of the
contact zone po is 3.1k. This is because the surface elements are subjected to compressive stresses in all three orthogonal directions, allowing po to reach a value greater than 2k without producing yield. This is
an important result since it means that contact pressures in excess of
the yield value for the material do not result in plastic deformation.
Thus, higher loads than might have been expected can be supported
elastically within Hertzian contacts. In addition, even if yielding has
taken place below the surface, very little plastic deformation takes place
on the surface itself because the plastic zone is constrained by elastic
material on all sides.

Elements of Surface Contact of Solids

23

With the further increase in load, the plastic zone also increases in
size and ultimately spreads to the surface of the body. Plastic ow may
then occur quite readily and the cylinder will indent the surface of the
body. This happens when the mean contact pressure pm is about 6k, i.e.
more than twice the contact pressure at which initial yield occurred.
The mean pressure under these conditions is essentially the indentation
hardness value of the material, H, which is why for metals the following
is applicable
H6k3Y
where Y is the material uniaxial tensile yield strength.
Another case of contact loading that is important in practice concerns
the combined action of a normal load, P, and a tangential load, P.
Furthermore, it is obvious that at all points within the contact zone the
tangential traction is given by tGp. Combining the stress distribution
due to the normal and tangential loads and calculating the values of
max leads to the isochromatics. When the pattern of isochromatics is
plotted it will be seen that the location of the greatest value of the
maximum shear stress is now much nearer the surface. Thus, plastic
deformation can take place more readily than in the previous case.
In practical terms it means that macroscopic plastic deformation is
facilitated by the presence of friction traction.

2.4.2 Contact with combined normal and tangential loads


Contact conditions under which bodies are subjected to tangential loads
less than P so that macroscopic sliding does not occur are quite frequently encountered in engineering applications. This takes place in
situations where friction is used as the mechanism for preventing slip
between mating components, for instance, nuts and bolts, interference
ts, and friction devices such as clutches. This mechanism will be
explained by considering a cylinder pressed against a plane and loaded
by a tangential load less than P. Under such loading conditions there
is a central area within the contact zone in which no slip occurs, while
at the two extremities a small degree of slip takes place as depicted in
Fig. 2.8. The coexistence of a zone of sticking and zones of microslip is
possible because of the deformable nature of the materials in contact
and the deformation pattern being such as to allow slip at the extremities of the contact zone. As the value of T increases, the areas of microslip increase until, when TGP, they meet at the centre of the contact
and microslip occurs over the whole contact zone. It is possible for
to have a constant value wherever slip takes place, that is, within the

24

Rolling Contacts

Fig. 2.8

slip regions tGp while within the stick region tFp. Since T is the
integral of t over the contact zone, this can satisfy the requirement of
the problem with always having a constant value.
The case of TGP is the next contact situation to be considered.
Increasing the normal load induces equal compression strains x in both
bodies so that no slip occurs owing to this effect. With the tangential
load, on the other hand, slip must occur throughout the contact zone
since the load must be acting in opposite directions on the two bodies
in contact.
It must be concluded that, even when no macroscopic motion takes
place, some degree of microslip exists when TFP and this gives rise
to a phenomenon known as fretting. For more complicated contact
geometries these arguments are still qualitatively valid and microslip
will occur at the extremities of the contact zone.

2.4.3 Three-dimensional contact


Many engineering applications of rolling contacts involve more complicated three-dimensional problems. In general, the patterns of behaviour
are similar to two-dimensional contacts, but some of the previously
introduced expressions must be modied.
If two identical spheres are brought into contact under a normal
load N (Fig. 2.9), the area of contact will be a plane circle of radius
a and the pressure distribution will be of hemispherical form and is
given by
pG

3N
2a2

x2 z2
1A 2A 2
a a

(2.15)

Elements of Surface Contact of Solids

25

Fig. 2.9

The magnitude of a is given by

1 2E

aG

3NR

(2.16)

The contact of two dissimilar spheres does not result in a plane circular
contact area, and the results given by equations (2.15) and (2.16) still
hold with substantial accuracy. The contact area radius is dened by

1 4E

aG

3NR

(2.17)

where R is related to the radii of contacting spheres R1 and R2 by


1
R

1
C
R1 R2

In the case of contact between a sphere and a plane, R is equal to the


radius of the sphere as the radius of the plane is taken to be innity.
Consider the contact of two bodies 1 and 2 whose geometry is dened
by the principal radii of curvature of each body in two orthogonal

26

Rolling Contacts

Fig. 2.10

planes as shown in Fig. 2.10. The area of contact is now elliptical in


shape and the contact pressure distribution is given by
pG

3N
2ab

x2 z2
1A 2A 2
a b

(2.18)

The size of the contact ellipse is dened by the semi-major and semiminor axes a and b as follows
aGka

bGkb

1
1
3

(2.19)

(2.20)

3N
4E(ACB)
3N
4E(ACB)

where ka and kb are constants depending on the values of the principal


curvatures of the contacting bodies and on the angle between the
normal planes that contain these curvatures. If the principal radii of
curvature of body 1 are denoted by R11 and R12 , the principal radii of
curvature of body 2 are denoted by R21 and R22 , and then constants A
and B are found from
BAAG12 1(C 2CD 2C2CD cos 2 )
ACBG

1
1
1
1 1
C C C
2 R11 R12 R21 R22

where CG1R11A1R12 and DG1R21A1R22 .

(2.21)
(2.22)

Elements of Surface Contact of Solids

27

In the above expressions a concave curvature is taken as negative.


Coefcients ka and kb in equations (2.19) and (2.20) are numbers
depending on the ratio (BAA)(ACB) and they can be found by introducing an auxiliary angle dened as
cos G

BAA
ACB

With the help of equations (2.21) and (2.22), the value of can be easily
obtained. In order to determine the values of ka and kb corresponding to
a certain value of , quite complicated numerical calculations involving
elliptical integrals are required. Figure 2.11 shows typical results of such
calculations.
The assumption of a plane area of contact is no longer valid for
complicated geometries. While the pressure distribution and the size of
the contact as determined from the Hertz theory are generally correct,
sometimes there is a need to know the actual shapes of such contacts.
For materials having the same elastic properties it is sufcient to assume
that the deformed surface, which has some common radius R c , is about
mid-way between the two original surfaces as shown in Fig. 2.12. The
value of the common radius of curvature is given by
RcG

2R1R2

(2.23)

R1AR2

Fig. 2.11

28

Rolling Contacts

Fig. 2.12

Obviously, for two identical spheres in contact the above equation gives
the expected result of a plane contact area. The radius is taken to be
negative where concave curvatures occur.
It is necessary for the analysis of contact to dene the normal
approach of a sphere owing to the application of normal load and the
consequent deformation. Figure 2.13 depicts the contact of a sphere
and a plane. It can be seen that the separation u of the surfaces at a
distance r from the centre of the contact zone is given by
uGRA1(R2Ar 2)GRAR

1A

r2
r2
GRARC
A
R2
2R

Fig. 2.13

Elements of Surface Contact of Solids

29

If r is small compared with R, then


uG

r2

(2.24)

2R

The normal approach is dened as the distance over which points on


the two bodies remote from the deformation zone move together on
application of a normal load. The reason for that is the attening and
general displacement of the surface within the deformation region. If a
is the radius of the contact zone and w is the displacement of the sphere
at the boundary of this zone, then the normal approach will be given
by
a2

GuCwG

Cw

2R

(2.25)

At the centre of the contact zone, is given by the degree of deformation and it is therefore justied to assume that the normal approach
will be proportional to the attening of the sphere. Thus

a2
R

With the help of equation (2.17)

1
3

aT

NR
E

so that

1
3

N2

E2R

The exact solution gives

9N 2
16E2R

or nally
NG43 E 1(R )

(2.26)

Combining equations (2.17) and (2.26), the area of contact, A, is given


by
AGa2 G R

(2.27)

30

Rolling Contacts

Equation (2.27) indicates that the surface outside the contact region is
displaced in such a way that the actual area of contact is only one-half
of the geometrical area, which is equal to 2 R.

2.5 Contact of real surfaces


All engineering surfaces are rough, and surface nishing processes available nowadays can substantially reduce the level of roughness but cannot eliminate it altogether. It is convenient to consider a simplied
contact between a single rough surface with a perfectly smooth plane,
as the result from such an approach is then reasonably indicative of the
effects to be expected from the contact of real surfaces. Moreover, the
problem will be simplied further by introducing a theoretical model
for the rough surface in which the asperities are represented by spherical
segments so that their elastic deformation characteristics may be
described by the Hertz theory. Also, it is assumed that there is no interaction between individual asperities; i.e. the displacement due to a load
on one asperity does not inuence the heights of the neighbouring
asperities.
Figure 2.14 shows, schematically, a surface of unit nominal area composed of an array of identical spherical asperities all of the same height
z with respect to some reference plane XX As the smooth plane moves
towards the rough surface as a result of the application of load, the
normal approach will be given by (zAd), where d is the current separation between the smooth surface and the reference plane. It is apparent
that each asperity is deformed equally and carries the same load Ni so
that for asperities per unit area the total load will be equal to Ni .
For each asperity, the load Ni and the area of contact Ai are known
from the Hertz theory [see equations (2.26) and (2.27)]. Thus, if is the

Fig. 2.14

Elements of Surface Contact of Solids

31

asperity radius, then


Ni G43 E 0.5 (zAd)1.5
and
Ai G (zAd)
and the total load is given by
4

NG E
3

0.5


Ai

1.5

Thus, the load is related to the total real area of contact AG Ai by


NG

4E
3 0.5
1.5

A1.5

(2.28)

Equation (2.28) indicates that the real area of contact is related to the
two-thirds power of the load when the deformation is elastic.
In the case of a load causing plastic deformation of asperities characterized by a constant ow pressure H, which is closely related to the
hardness, it is assumed that the displaced material moves vertically
down and does not spread horizontally. In this way, the area of contact
A will be equal to the geometrical area 2 . The load on an individual
asperity Ni is then given by
Ni GHAi G2H (zAd)
Therefore
NG Ni G HAi GHAG2HA

(2.29)

which means that the real area of contact is linearly related to the load.
All engineering surfaces have asperity peak heights distributed in a
probabilistic way. Therefore the surface model introduced earlier must
be modied accordingly and the analysis of contact between real
surfaces has to include a probability statement as to the number of
asperities in contact. Assuming that the separation between the smooth
surface and the reference rough plane is d, then there will be contact at
any asperity whose height was originally greater than d as shown in

32

Rolling Contacts

Fig. 2.15

Fig. 2.15. If (z) is the probability density of the asperity peak height
distribution, then the probability that a particular asperity has a height
between z and zCdz above the reference plane will be (z) dz. Thus,
the probability of contact for any asperity of height z is given by

(z) dz

prob(zHd)G

Considering a unit nominal area of surfaces containing asperities, it


is not difcult to show that the number of contacts created, n, is equal
to
nG

(z) dz

(2.30)

The total area of contact is given by


AG

(zAd) (z) dz

(2.31)

and the expected load is dened by the following expression


4
NG 0.5E
3

(zAd)1.5 (z) dz

(2.32)

It is common practice to express the above equations in terms of standardized variables, i.e. hGd and sGz , where is the standard

Elements of Surface Contact of Solids

33

deviation of the peak height distribution of the surface. Therefore


nG Fo (h)
AG F1(h)
NG43 0.5 1.5EF3/2 (h)
where F32(h) is a statistical function which, in general terms, is given
by

Fm(h)G

(sAh)m *(s) ds

In the above expression, *(s) is the probability density function standardized by scaling it to give a unit standard deviation.
In cases where the asperity obeys the plastic deformation mode, equations (2.31) and (2.32) are modied to assume the following forms
AG2

(zAd) (z) dz

(2.33)

NG2 H

(zAd) (z) dz

(2.34)

It is apparent that the load is linearly related to the real area of contact
by NGHA and this result is independent of the height distribution
(z).
The analysis presented above was based on a theoretical model of a
rough surface. An alternative approach to the problem is to use actual
surface roughness prolograms and obtain from them the surface bearing area curve. In the absence of asperity interaction, the bearing area
curve provides a direct method for determining the area of contact at
any given normal approach. Thus, if the bearing area curve is denoted
by (z) and the current separation between the smooth surface and the
reference plane is d, then for a unit nominal surface area the real area
of contact is given by

AG

(z) dz

(2.35)

34

Rolling Contacts

In the case of plastic surface deformation, the total load on the contact
is
NGH

(z) dz

(2.36)

It is apparent from the analyses presented above that the relationship


between the real area of contact and the load will be dependent on both
the mode of deformation and the distribution of the surface prole.

2.6 Criterion for deformation mode


In most real contact situations the higher asperities are deformed plastically while the lower asperities are still elastic. Therefore, a mixed
elasticplastic system exists in most real contacts. Usually the greater
the load, and hence the increase in the normal approach, the greater is
the number of plastic asperity contacts. Thus, the normal approach can
be regarded as an indicator of the extent of plastic deformation within
the contact area. Taking into account equations (2.26) and (2.28), it is
possible to dene the mean pressure, pm, for the case of an elastic
asperity contact as
pm G

4E 0.5
3 0.5

or

0.5 G

3 0.5pm
4E

(2.37)

The transition from a purely elastic contact to a completely plastic contact takes place over a range of loading for a contact between two
spheres. Plastic deformation is initially located under the surface when
the maximum contact pressure is 3.1k or the mean pressure is approximately equal to Y. The extent of plastic deformation becomes macroscopic when the mean contact pressure is about 3Y; i.e. it is equal to
the hardness of the material. Thus, from equation (2.37) it can be seen
that the transition from elastic to fully plastic behaviour occurs in a
range of values of 0.5, and the initial deviation from elastic behaviour
occurs when pm GH3, where

0.5 G0.78

0.5H
E

Elements of Surface Contact of Solids

35

The transition from elastic to fully plastic behaviour is gradual, and


therefore a transition point can be assumed at

0.5

H 0.5

It is useful to normalize the above expression by dividing both its sides


by 0.5

( *)0.5 G

0.5

0.5

where is the standard deviation of asperity heights.


The left-hand side parameter decreases as the surface roughness,
dened by , increases. It is common to name the inverse of ( *)0.5 as
the plasticity index

E
H

0.5

(2.38)

The plasticity index is indicative of the onset of plastic deformations


within the contact zone. It is large when the contact is basically plastic,
and small, less than unity, when the contact is essentially elastic.

2.6.1 Surface plastic deformations


In this section, a heavy, rigid roller moving steadily over the surface of
an elasticplastic half-space is considered. Resulting deformations of
the roller are assumed to be in plane strain. The elastic properties of
the half-space are homogeneous and isotropic, and the plastic deformation is assumed to be in-plane shear on planes normal to the direction of rolling. The shear is taken to be of constant magnitude within
a surface layer of thickness h, extending backwards through the rolled
zone from the leading point of contact between the roller and the
surface.
In a Cartesian coordinate system (x1 , x2 , x3) the half-space is dened
as x20 and is occupied by an elasticplastic body, isotropic and homogeneously elastic and subject to conditions of plane strain normal to
the x3 axis. The surface of the half-space is loaded by a rigid roller
assumed to move steadily in the positive direction of the x1 axis, leaving
behind it a plastically deformed layer of material with thickness h.
Owing to the assumed steadiness of the movement, the plastic deformation, when checked behind the roller, will be constant at a given
depth under the surface.

36

Rolling Contacts

In front of the surface layer there will be a region in which plastic


deformation takes place. It is assumed, tentatively, that this region is a
narrow slip band emerging from the leading point of contact perpendicularly to the surface. Removal of the material in the slip band would
leave the unloaded body with displacements
v1 Gv 1s (x2)

(2.39)

where v2 G0 for ASFx1Fb and Ahx2 0, and elsewhere v1 Gv2 G


0. Here, b is the coordinate in the x1 direction of the slip band. In
particular, v1s (h)G0, and so the body is left stress free.
To calculate the deformation of the surface in the actual body that
is due to plastic straining, let g (x1At, x2 ) be the x direction displacement of the point (x1 , x2 , x3 ) of the body when deforming a purely
elastic material to a unit-concentrated line load acting in the x direction
at (x1 , x2 )G(t, 0). The Greek subscripts range over the values 1 and 2.
Also, let
G (x1At, x2 )G

2
l l l
r

(2.40)

be the in-plane stress of the body thus loaded. Additionally


rG1[(x1At)2Cx22]
and
1
l G (x1At, x2 )G(cos , Asin )
r
denote the distance between the points (x1 , x2 ) and (t, 0).
In order to restore the body to its state prior to removal of the slip
band, it is necessary to subject the anks at x1 Gb and x1 Gb+ to tractions T (x2 ) and T + (x2 ) respectively, with
T (x2)CT + (x2 )G0

(2.41)

In this process, the anks will experience displacements v (x2 ) and


v + (x2 ) respectively.
The tractions are determined by the conditions
v 1 (x2 )Cv 1s (x2)Gv +1 (x2 )
v 2 (x2 )Gv +2 (x2 )
This loading will give rise to elds of displacements v (x1 , x2 ) and
residual stresses (x1 , x2 ) satisfying the following conditions:
, G0 for x2 0 and 2 G0 for x2 G0.

Elements of Surface Contact of Solids

37

Utilizing the reciprocity relations

S AC

G (x1At, x2 )v , (x1 , x2) dS

S AC s

g , (x1At, x2 ) (x1 , x2 ) dS

In the above expression, S is the region x2 0, taken up by the body,


and C s denotes the slip band. Application of the divergence theorem is
admissible for the region S AC s. Thus

G 11 (bAt, x2 )v 1s (x2 ) dx2

v (t)G

(2.42)

Here, v (t)Gv (t, 0) are the surface displacements. However, the derivative v (t) rather than v itself is required. Therefore

G 11,1 (bAt, x2 )v 1s (x2 ) dx2

v (t)GA

G 12,2 (bAt, x2 )v 1s (x2 ) dx2

G 12 (bAt, x2 ) (x2 ) dx2

where

(x2 )Gdv1s (x2 ) dx2

(2.43)

is the magnitude of slip. Finally, as l1 dx2 Gr d


v1 (t)G

v2 (t)G

cos sin d

sin2 d

where

Garctan

Ax2
;
bAt

o Garctan

h
bAt

for tFb

38

Rolling Contacts

and
v1 (t)G
v2 (t)G

cos sin d

sin2 d

where

GAarctan

Ax2
tAb

o GAarctan

h
tAb

for tHb

If, for simplicity, it is assumed that is constant, then


v1 (t)G
v2 (t)G

h2
hC(bAt)2

(2.44)

h
h(bAt)
arctan
A 2

bAt h C(bAt)2

(2.45)

It should be noted that for H0 the surface has a wedge at tGb with
a jump in inclination of magnitude . It can be assumed that the wedge
is situated at the leading point of contact. There are, however, no shear
stresses available to drive the plastic deformation in front of the contact
zone, and the wedge would be attened under the roll.

2.7 Thermal effects during rolling


In order to create a qualitative physical picture of the problem, it is
advisable to consider the situation shown in Fig. 2.16(a) where two

Fig. 2.16

Elements of Surface Contact of Solids

39

contacting discs are assumed to be rolling with a very small amount of


relative slip. It is clear that all particles on the surface of both discs pass
through the contact zone where heat is being generated and afterwards
undergo considerable periods of rest. Temperature rise for them will
therefore be rather modest owing to the small magnitude of frictional
work and low relative sliding velocity, and also because the generated
heat will be readily dissipated during the rest periods. If one of the discs
is stationary [Fig. 2.16(b)], then conditions of pure sliding will be
created. In this case, surface particles on disc 2 will be subjected to
relatively high temperatures when passing through the contact zone and
have a considerable rest period outside the contact zone where cooling
takes place. Surface particles on disc 1, however, never leave the contact
zone and are subjected to a continuous build-up of temperature towards
a steady state determined by the thermal properties of the whole system.
Thus, the contact zone which is xed in space can be regarded as a
stationary heat source with respect to disc 1, while the contact zone
with respect to disc 2 can be regarded as a moving heat source. It is
obvious that for rolling contacts the relevant and important case is that
of a moving heat source.

2.7.1 Moving source of heat


The problem of a moving heat source traversing the surface of a semiinnite body at a relatively high speed V can be somewhat simplied
by neglecting the effects of the transverse ow of heat. Thus, the problem can be treated as one of linear heat ow. If the heat is supplied at
a constant rate of q per unit area, then the mean temperature rise of a
point on the surface of the body is given by

1 c
2qt

where t is the time during which heat is supplied, is the thermal


conductivity of the body, is the density, and c is the specic heat of
the body. If heat is supplied through a circular area of radius a, then,
by assuming that qGQa2 and by considering the effective value of t
for all points within this area, a mean surface temperature can be
obtained. The time to traverse the contact area for any point dened
by (x, y) is given by
tG

2x
V

2
V

1(a2Ay 2 )

40

Rolling Contacts

and therefore the mean effective time is


tG

1(a Ay ) dyG
2a (a Ay )
4V
1

Thus

mG

0.318Q
2Q1(a)
G
2
1
1
2a ( cV ) a (acV )

(2.46)

It is convenient to introduce a normalized surface temperature *


dened as

*G

cV

so that for a moving source of heat, utilizing equation (2.46), the result
is

m
*G

0.3181(2acV )
G0.438 0.5
0.5
(2 )

(2.47)

where

ac
V
2

2.8 Contact of bodies with interposing lm


The performance of rolling contact bearings coated with thin solid
lubricant lms, particularly those deposited by such processes as ion
plating and sputtering, is of practical importance and interest, as in
many cases the size of the contact zone is considerably inuenced by
elastic deformation of the lm material as well as by that of the
contacting bodies.
In this section, expressions are provided from which the dimensions
of all-elastic cylindrical and spherical contacts may be determined when
the ratio of the contact dimensions to the lm thickness is large and
where the ratio of the modulus of elasticity, , of the lm material to
that of the contacting bodies is less than unity.

Elements of Surface Contact of Solids

41

2.8.1 Background to the analysis


It is known that a wide variety of contacts possess an approximately
elliptic pressure distribution of the form
qGqc

r2
1A 2
a

(2.48)

where a is the contact half-width, q is the pressure at a distance r from


the contact centre, and qc is the pressure at the centre. The above
approximation is not justied for values of greater than unity.
If the load on the contact is P per unit length, it is permissible to
write
a

PG2

q drG 2 aq
1

(2.49)

and, making the reasonable assumption that the same pressure distribution in an axisymmetric form will apply in the case of spherical
contact
a

PG2

rq drG 3 a q
2

(2.50)

in which P is the total contact load and a is now the contact radius.
In the special case of Hertzian contact between equal cylinders or
spheres of the same material
(i) cylinder
a2o G

4PR(1A 2 )
E

(2.51)

(ii) sphere
a3o G

3PR(1A 2 )
2E

(2.52)

where ao is the Hertzian half-width or radius for direct contact between


surfaces of radius R with modulus of elasticity E and Poissons ratio .

42

Rolling Contacts

In the case of Hertzian contact, equation (2.48) becomes


qGqh

1A

r2
a2o

(2.53)

where qh is the pressure at the centre of the Hertzian contact. Fig. 2.17
shows the model of the contact under consideration. The contact consists of two equal cylinders or spheres (radius R and elastic constants
E2 and 2 ) pressed together with load P or P into a lm of thickness
h and elastic constants E1 and 1 . As the contact is symmetrical, all the
derivations are signicantly simplied. However, the results of analysis
may be translated into any conformal or counterformal cylindrical or
spherical contact provided that RZaZh.

2.8.2 Case of contacting cylinders


With reference to Fig. 2.17, the relative displacement normal to the
contact surface between points at the contact edge (rGa) and the
contact centre is given approximately by
w(0)Aw(a)G

a2

(2.54)

2RAh

where h is the central compression displacement of one-half of the


lm. Assuming the lm to be sufciently thin for the normal component
of the pressure across it to be virtually constant, it is permissible to
write
hG

hqc

(2.55)

2E1

Fig. 2.17

Elements of Surface Contact of Solids

43

as it is assumed that there is no friction or bonding between the cylinders and lm, i.e. that free compression conditions exist. At the other
extreme, it is possible to assume that the lm is bonded to both cylinders
and, since the lateral strain is negligible at the centre, that the lm in
this region is subjected to bulk compression. Thus
hG

(1A21)(1C1 )hqc

(2.56)

2(1A1)E1

It is convenient to write equations (2.55) and (2.56) as


hG

khqc

(2.57)

2E1

where k is a factor describing the nature of the bonding.


Considering now an innite line load of qr per unit length, acting
on the surface of a semi-innite plane, the relative perpendicular displacement between two points whose distances from the line load are
A and B can readily be shown to be
w(A)Aw(B)G

2(1A22 )
B
ln
qr
E2
A

(2.58)

where E2 and 2 are the elastic constants of the plane material. It is


now possible to use equation (2.58) to dene the relative displacement
w(0)Aw(a). Figure 2.18 shows the contact with two line load elements
at a distance r from the centre-line. The total relative displacement for

Fig. 2.18

44

Rolling Contacts

the whole contact is

1A 22
E2

w(0)Aw(a)G2qc

aAr
r

ln

1A

r2
dr
a2

r 1
a

ln

aCr

1A

G2aqc

1A 22
E2

r2

dr

a2

ln 1Aa 11A a d a
1

r2

r2

a 1
1

A2

ln

1A

d
a a
r2

which can readily be evaluated by trigonometric substitution and integration by parts to give
w(0)Aw(a)Gaqc

1A 22

(2.59)

E2

By combining equations (2.54), (2.57), and (2.59)


aqc

1A 22 a2 khqc
G A
E2
2R 2E1

(2.60)

from which, after substituting for qc from equation (2.49), the following
is obtained
1A 22
kh
aC
a G4PR
E2
2 (1A 22)

(2.61)

which reduces to the Hertz formula given by equation (2.51) when hG


0. By substituting from equation (2.51) and rearranging
h
ao

G2

1A 22
k

a Aa
a

(2.62)

so that, provided that ao is known for Hertzian contact between cylinders of material 2, it is possible to dene a for given values of h.

Elements of Surface Contact of Solids

45

2.8.3 Contacting spheres


In order to obtain w(0)Aw(a) for the case of contacting spheres, it is
convenient to refer to some results of Timoshenko and Goodier (2),
which may be presented in the form
w(r)Gqc

1A 22
(2a2Ar 2 )
4aE2

(2.63)

where w(r) is the normal displacement at a point located at a distance


r from the centre of an elliptical contact. Hence
w(0)Aw(a)Gaqc

1A 22
4E2

(2.64)

and, in a way analogous to that used in the case of contacting cylinders,


the following is obtained
1A 22
h
G
ao
2k

a Aa
a

(2.65)

It is important for both cases considered to discuss the signicance of


factor k. It is clear that the extreme values of k are unity for no lateral
constraint and (1A21)(1C1)(1A1) for both surfaces rigidly bonded.
By taking 1 G0.25, these limits are 1 and 56 ; therefore it is justied to
conclude that k is not expected to have a very pronounced effect for
most materials. In friction situations, one surface of the lm is bonded
while the other is constrained to some extent by friction forces. Thus,
it is reasonable to expect the lower value of k to be more applicable in
this case.

2.9 Crack formation in contacting elastic bodies


The problem of the development of a crack in contacting elastic bodies
has some common elements with the contact problem of the theory of
elasticity (3). This section analyses the problem of two elastic halfplanes pressed together, without any friction, by distributed loads of
constant intensity q that are orthogonal to their common boundary, i.e.
the x axis, and are applied far away from that axis.

46

Rolling Contacts

2.9.1 Description of the contact


The model of the contact is shown in Fig. 2.19. It can be seen that two
equal but opposite forces P and P tend to break the contact of these
half-planes. Force P is oriented in the positive direction of the y axis
and is the resultant of a uniformly distributed load p which is applied
to the part of the boundary dened by cFxF+c and yG0 of the upper
half-plane. Force P has an analogous signicance; the load p, the
resultant of which is P, acts on the lower half-plane.
It is obvious that
2pcGPGPG2pc
Let it be assumed that, as a result of the action of forces P and P, the
contact of the elastic half-planes is broken along that part of the
interface which is dened by aFxF+a, where a gap, 2a, appears, the
length of which has to be determined.
In order to nd the length of the gap, a plane problem of the theory
of elasticity will be considered. In agreement with the earlier considerations, the following boundary conditions at the boundary of the upper
half-plane are applicable

x 0,
y G0,
y Gp,

xy 0,
xy G0
xy G0

y Aq
for y S
for AaFxFc and for cFxFa
for AcFxFc

Fig. 2.19

Elements of Surface Contact of Solids

47

and, moreover, xy G0 and vG0 at all points of the part of the boundary of both half-planes at which they are in contact, i.e. for xHa, yG
0. Here, as is usual, x , y , and xy are the components of the stress
tensor and u and v are the components in the x and y directions of the
displacement of an arbitrary point on the elastic half-plane.
By using well-known methods of solution of the plane problem of
the theory of elasticity, the functions u and v of the variables x and y
can be determined from the functions x , y , and xy which satisfy both
equilibrium equations and the boundary conditions. It is found that,
on the segment of the interface where there is no contact, i.e. xFa and
yG0, the displacement of the points of the boundary of the upper elastic half-plane in the vertical direction is given by the following
expression
vGv(x, 0)G

c
C2 2
p arcsin Aq 1(a2Ax2 )
2 ( C)

1(a2Ax2)C1(a2Ac2)
p
C c ln
1(a2Ax2)A1(a2Ac2)

Ax ln

c 1(a2Ax2)Cx1(a2Ac2)
c 1(a2Ax2)Ax 1(a2Ac2)

(2.66)

where and are Lames constants. The expression for the distributed
load maintaining the contact between the two half-planes in the domain
where there is contact has the following form

p arcsin aAq 1(x Aa )

yG

a2Acx
a2Ccx
C arcsin
Aarcsin

a(xAc)
a(xc)
p

(2.67)

which is valid only for xHa.


Let l denote the root in the interval (c, S) of the trigonometric
equation
c
2
p arcsin AqG0

(2.68)

in which the quality a is unknown. Under the condition pHq, equation


(2.68) will always have exactly one such root.

48

Rolling Contacts

It is not difcult to show that the half-length a of the gap formed as


a result of forces P and P which tend to separate the half-planes is
exactly equal to the root l of equation (2.68). Let it be assumed that
aFl. In this case, the gap has the shape shown in Fig. 2.20. The tangent
to the contour of the gap is vertical in its extreme points xGJa and
yG0. In the immediate proximity of the ends of the gap, the stress y
at the points of the x axis becomes positive. This means that the elastic
half-planes must exert a tensile traction on each other in order to maintain contact.
In the present case, as in the classic statement of the contact problem
of the theory of elasticity, a tensile traction between contacting bodies
is not admitted. Thus, the case aFl is impossible.
Now, let it be assumed that aHl. Thus, everywhere in the contact
area, i.e. for xHa and yG0, the stress y is negative and, consequently,
the elastic half-planes are pressed against each other. In this case, however, the function vGv(x, 0), i.e. the displacement of the boundary of
the upper elastic half-plane, is positive near the ends of the gap and
negative in its central part. In consequence of this, the contour of
the gap acquires the shape of a curve that intersects itself as shown in

Fig. 2.20

Elements of Surface Contact of Solids

49

Fig. 2.21

Fig. 2.21. This means that the elastic half-planes penetrated one another
which, of course, is impossible, and the case aHl has to be eliminated.
It can then be concluded that the solution of the plane problem of
the theory of elasticity leads, for a l, either to a physically impossible
stress state or to an impossible displacement eld. Thus, the only possible case left is lGa. In this case, the following is applicable
1(l 2Ax2)C1(l 2Ac2)
C2 p
v(x, 0)G
c ln
1(l 2Ax2)A1(l 2Ac2)
2 ( C)

Ax ln

c1(l 2Ax2)Cx1(l 2Ac2)


c1(l 2Ax2)Ax1(l 2Ac2)

(2.69)

and

y (x, 0)G

p
a2Acx
a2Ccx
arcsin
Aarcsin

a(xAc)
a(xCc)

(2.70)

Now the ends of the gap smoothly converge to each other. The tangent
to the contour of the gap in the points xGJl and yG0 is directed

50

Rolling Contacts

along the x axis. The stress y and the distributed pressure of one elastic
half-plane upon the other at the given points is equal to zero and
increases gradually, tending to the negative value Aq as the distance
from the gap increases. The length of the gap increases with growth of
the cleaving forces P and P. The equations of the theory of elasticity
and the boundary conditions are satised.

2.10 Contacts deviating from the Hertz theory


The theory of elastic contact put forward by Hertz has been proved to
predict the area of contact and stresses remarkably well. It has been
used by both the engineer and physicist to deal with problems arising
from the contact between elastic bodies. In general, deviations from the
Hertz theory are small; nevertheless in special circumstances, physical
conditions that were omitted by Hertz can be important. These special
cases are known as non-Hertzian contacts.
It is useful to list the relevant restrictions in the Hertz theory:
the bodies in contact are homogeneous and isotropic;
their surfaces are smooth and continuous;
their proles are represented by a second-order surface;
the stresses and displacements may be found from the small strain
theory of elasticity applied to a linear elastic half-space;
there is no friction between surfaces in contact;
the surface tractions are the result of the contact forces only, i.e.
adhesive forces are ignored.
In this section, modications to the theory that are necessary if some
of the above restrictions are to be relaxed will be considered.

2.10.1 Friction at the contact interface


No engineering surface is ever frictionless, so that an obvious practical
extension of the Hertz theory is to remove that restriction and to consider the inuence of friction at the contact interface. In the normal
contact of spheres, friction introduces a rst-order correction only when
the two bodies have dissimilar elastic constants. Under the action of
their mutual contact pressure, the two surfaces undergo radially inward
tangential displacements whose magnitudes, in the absence of friction,
would be proportional to the respective values of the parameter
(1A2)G, where G is the shear modulus and is Poissons ratio. With
dissimilar materials, the displacements will be different for the two
surfaces and the resulting slip will be resisted by interfacial friction,

Elements of Surface Contact of Solids

51

which will act radially outwards on the more compliant surface and
inwards on the more rigid one.
Theoretically, interfacial slip could be prevented by a sufciently high
coefcient of friction at the interface. This will constitute the fully
adhesive or no-slip contact conditions. At the other extreme, if the coefcient of friction is very small, slip will take place over the whole contact area and the tangential traction will be radial and equal to Jp
everywhere ( p denotes normal contact pressure). This is the so-called
complete slip contact condition. In reality, the contact comprises a
central circular region of adhesion surrounded by an annulus of slip.
This case is usually referred to as the partial slip contact condition.
The contact behaviour is controlled by two non-dimensional parameters, namely

[(1A21)G1]A[(1A22 )G2]
[(1C1)G1]C[(1C2)G2]

and coefcient of friction .


For practical values of the elasticity parameter (F0.4) and the coefcient of friction (F0.5) the inuence of frictional traction upon the
compliance and load-bearing capacity of most contacting bodies is
small, so that the use of the Hertz theory is justiable. An exception
arises in the case of brittle solids which fracture as a result of the radial
tensile stress surrounding the contact area. This component of stress in
the absence of friction is given by

rr G

(1A2)P
2r 2

(2.71)

and is valid for ra.


In the above expression, P is the load on the contact and r denotes
the distance from the centre of the contact. It is particularly sensitive
to frictional tractions at the interface.

2.10.2 Adhesion at the contact interface


Owing to intermolecular forces, clean, smooth surfaces when pressed
into intimate contact may in some circumstances strongly adhere
together. The theory proposed by Hertz does not involve such forces
and, as a matter of fact, measurable adhesion between non-conforming
elastic bodies is rarely observed in practice. A notable exception to this
common experience is the contact of an optically smooth rubber sphere

52

Rolling Contacts

of very low elastic modulus with a smooth glass surface. The combination of smooth surfaces and a compliant solid allows for intimate
contact between the rubber and the glass surfaces to be realized
throughout the whole of the apparent contact area. Under a compressive load, the sphere makes contact with the glass over a circular area
appreciably greater than that predicted by Hertz theory and a measurable tensile force is required to separate the two surfaces. As expected,
a thin layer of uid with a wetting agent between the surfaces destroys
the adhesion and the deformation of the sphere reverts to that predicted
by Hertz theory.
As it is not known what law governs the force acting agent against
separation, so it is convenient to express the adhesive action in terms
of a surface energy , dened as the work required to separate a unit
area of the adhered surfaces. Then, if two elastic spheres of radii R1
and R2 are brought into contact under a force P1 , according to Hertz
they make contact on a circle of radius a1 given by
a31 G

RP1
M

(2.72)

where
RG

R1R2
R1CR2

MG

4E
3

and
1
E

1A 21 1A 22
C
E1
E2

Reducing the load acting on the contact to Po , while maintaining the


surfaces in contact by adhesion over the same contact radius a1 , allows
the resulting distribution of surface traction to be found by subtracting
the stress under a at cylindrical punch given by Hertz theory, namely
p(r)G

3P1
2a21

1A

1
r 2 P1APo
A
2
2
a1
2a1 1(1Ar 2a21)

(2.73)

Elements of Surface Contact of Solids

53

Fig. 2.22

This traction is of tensile nature at the edge of the contact and compressive in the centre, as shown in Fig. 2.22 (line B). It is possible to
estimate the elastic strain energy UE associated with the state of stress
within the contact zone. Thus
1
UE G 15
M 2/3R 1/3 (P15/3C15P 2o P11/3C5Po P12/3)

(2.74)

where P1 is related to a1 through equation (2.72). The surface energy


associated with a contact area of radius a1 is
Us G a G
2
1

M
P1R

2/3

(2.75)

According to Grifth, the equilibrium value of a1 , for any given value


of Po , occurs when the variation in elastic strain energy with a1 just

54

Rolling Contacts

matches the variation in surface energy, i.e. when d (EECVs)da1 G0.


This condition results in
Ma31
GP1 GPoC3 RC1[6 RPoC(3 R)2]
R

(2.76)

If it is assumed that Po GP and a1 Ga, an expression is obtained for the


equilibrium value of a under the action of load P in the presence of an
adhesive force. Thus
a3 G

R
1[PC3 RC36 RPC(3 R)2]
M

(2.77)

When the surface energy is zero, equation (2.77) reduces to the simple
Hertz equation. It is also apparent that an equilibrium contact area can
be maintained even though the force is tensile up to a maximum value
given by
PG 32 R

(2.78)

This is also the force required to separate the contacting spheres.

2.11 References
(1) Hertz, H. (1896) The contact of elastic bodies. Miscellaneous Papers
(Macmillan, London).
(2) Timoshenko, S. and Goodier, N. N. (1951) Theory of Elasticity
(McGraw-Hill, New York).
(3) Barenblatt, G. I. (1959) On the equilibrium cracks due to brittle
fracture (in Russian). Prikl. Mat. i Mekh., 23.

Chapter 3
Fundamentals of Rolling Motion

Why is it that spherical and cylindrical forms are easier to move? Firstly
because they have a very slight contact with the ground and secondly because
there is no friction, for the angle is well away from the ground.
Why is it that it is easier to convey heavy weights on rollers than on carts
although the latter have large wheels and the former a small circumference?
Is it because a weight placed upon rollers encounters no friction, whereas when
placed upon a cart it has the axle at which it encounters friction?

This quotation from Mechanica, Section 8.11, by Aristotle adequately


describes a popular understanding of the resistance to motion in rolling
contacts. It also emphasizes the fact that there is a different way in
which one surface can be moved relative to another and that is by
rolling. It is undoubtedly much easier to roll surfaces than to slide them,
an observation discussed above with characteristic clarity by Aristotle.
With hard materials, the coefficient of rolling friction may be as little
as 0.001.

3.1 General features of rolling contact


It is a common experience of each of us that rolling contacts operate at
a far lower resistance to motion than those involving sliding. However,
in order to secure this low resistance to motion, the rolling component
must be placed on a relatively hard and smooth substrate so that the
stresses in both interacting surfaces are entirely elastic. For a given load
and geometry, increasing the rigidity of the rolling components will
result in further reduction in the resistance to motion. It is, however,

56

Rolling Contacts

not free and there is a cost to pay in the form of increasing the surface
and near-surface stresses at the point of contact.
During the rolling motion, for instance, of a cylinder over a nominally flat surface under a load N, deformations of both the cylinder
and the surface occur. As a consequence, the macroscopic contact
between them takes place on a finite area ab (Fig. 3.1). Because of the
roughness of both interacting surfaces, the real contact takes place at
discrete locations within the nominal contact area which is analogous
to a sliding contact. However, owing to deformations of interacting
surfaces a special characteristic feature of the rolling contact is microslip. The first to point out the occurrence of microslip during rolling
was Reynolds (1). Assuming that deformations at the contact between
the cylinder and a flat surface are elastic, Reynolds pointed out that
within the instantaneous contact area ab the cylinder is under compression and the flat surface under tension. Thus, as rolling progresses,
points a and b, and for the matter other points located within the contact zone on the substrate surface, will have a tendency to get closer to
each other, whereas corresponding points located on the cylinder surface will tend to move apart. This phenomenon is responsible for microslip taking place at discrete points within the contact zone where the
energy of elastic deformation is greater than the cohesive energy of
junctions formed as a result of intimate contact between the cylinder
and the substrate. Thus, aa1 and bb1 are the regions where slip occurs
and a1b1 is the region where there is no relative sliding between the

Fig. 3.1

Fundamentals of Rolling Motion

57

contacting bodies. Therefore, the resistance to rolling originates from


elastic hysteresis losses within the material of the contacting bodies.
Another particular feature of rolling motion is manifested by a continuously growing distance between surfaces at the rear of the contact
and their steady approach in front of the contact. Thus, for the rolling
to continue, surface forces and eventual adhesive junctions must be
overcome at the rear of the contact. Some of the work expended on
that can be retrieved by the action of surface forces in front of the
contact. When a lubricant is applied to the substrate, the creation and
destruction of discrete direct contacts between the interacting bodies
must be preceded by action against lubricant viscosity.
During rolling there is a continuous change in the nominal area of
contact. Contact pressures within the nominal area of contact, as well
as deformations caused by them, are uneven. Moreover, the relative
motion of surface elements takes place along curvilinear trajectories.
Therefore, it is justified to state that the rolling friction is a more complicated physical phenomenon than the sliding friction. For that reason
it is more difficult to present rolling friction in analytical terms.
Two familiar and practically important examples of rolling contacts
are those of pneumatic tyres and steel-rimmed railway wheels. The
loads on each are quite similar but the properties of the materials
involved are very different. The rolling resistance of a tyre is very much
greater than that of a steel wheel on a steel rail because of the very
much lower value of the equivalent elastic modulus of a rubber tyre on
a concrete road as compared with that of a steel wheel on a steel rail.
On the other hand, the magnitude of the contact stresses, which to a
large extent control the potential damage to both road and rail surfaces,
is very much lower for a pneumatic tyre than for a steel-rimmed wheel.
The smooth operation of a very large number of machines depends,
at least to some extent, on the utilization in their design of rolling contacts. A typical practical example of this is a rolling contact bearing
where a number of balls or rollers are placed between the two moving
elements so that the sliding motion is transformed into a rolling motion.
This consequently leads to a substantial reduction in the resistance to
motion.
Although friction in a rolling contact is generally much less than in
a sliding contact, it is, nevertheless, finite and its origins have been the
subject of much investigation. As a result of the difference in the elastic
deformation of the contacting surfaces, there is always likely to be an
element of microslip within the contact region, and this makes a contribution to the overall resistance to motion. However, in steel components rolling on one another, the magnitude of this element of slip and

58

Rolling Contacts

the friction associated with it are quite small and insufficient to account
for the whole of the observed resistance to motion in rolling contacts.
A second characteristic mechanism for rolling contacts is known as
Heathcote slip. It contributes to the rolling resistance of ball bearings.
If the geometry of the rolling components is conformal, exemplified by
that of a sphere rolling in a groove, then within the contact zone there
will be two lines along which there is zero surface slip. However, on
either side of them there are regions in which the relative sliding
between the sphere and the track is in opposite directions and therefore
the rolling resistance is quite substantial. In addition, the small but finite
hysteresis loss can also be an important factor. Finally, in real bearings
the contact is usually lubricated and energy loss within the viscous
lubricant film will also play a part.

3.2 Source of friction in rolling contact


What is the ultimate source of rolling friction and why is it generally
so small? Until recently, most engineers considered that rolling friction
arises from minute slip between the ball and the surface over which it
rolls. However, recent studies, both experimental and theoretical, have
shown that, although such slip does occur, it generally contributes only
a small part to the rolling resistance. This is supported by the observation that lubricants, which greatly reduce sliding friction, have very
little effect on rolling friction (2). A clue as to the source of rolling
friction can be obtained by considering what happens if a hard steel
ball rolls over a softer metal such as lead or copper [Fig. 3.2(a)]. As it
rolls along, the ball displaces metal plastically around and ahead of it
and produces a permanent groove in the metal surface. It is easy to
demonstrate that the force required to displace the metal is almost
exactly equal to the observed rolling friction. Thus, the rolling friction
is essentially a measure of the plastic work of deformation or grooving.
It is now easy to see why lubricants have little effect on rolling friction.
What happens, then, if rolling takes place over an elastic solid such
as rubber? No permanent groove is formed. Does this mean that the
rolling resistance is zero? The answer is no. As the ball rolls forward, it
deforms the rubber ahead of it and in doing so does work on it. The
rubber recovers elastically and does work on the rear portion of the
ball pushing it forward [Fig. 3.2(b)]. If the recovered energy in the rear
portion were equal to that expended on the front portion, the net work
required to roll the ball would be zero. However, no material is ideally
elastic. In the course of deforming and relaxing rubber, some energy is

Fundamentals of Rolling Motion

59

Fig. 3.2

lost by elastic hysteresis. Hysteresis losses originate from the rubbing


of rubber molecules over one another during the deformation process.
Indeed, elastic hysteresis is sometimes described as internal friction.
With a very bouncy rubber, the hysteresis losses are very small; only a
few percent of the deformation energy may be lost in this way. On the
other hand, with a very soggy rubber as much as 95 percent of the
deformation energy may be lost. This lost energy appears as heat within
the rubber.
Ball bearings are made of hard steel and are so designed that the
contact stresses are not high enough to produce plastic flow of the balls
or races. Only elastic deformation occurs. For bearing steels, the hysteresis losses are extremely small (less than 0.5 percent) so that the rolling
resistance is very small. In practice, the balls must be surrounded by a

60

Rolling Contacts

cage to separate them and prevent them rubbing on one another. The
cage friction is often far greater than the rolling friction. Lubricants are
used to reduce the sliding friction between balls and cage. As was said
earlier, they play little part in rolling friction itself.
With rubber-like materials, the rolling resistance can be very much
larger. For example, with a ball loaded so heavily that it is half-buried
in the rubber and using a soggy rubber of high hysteresis loss, a rolling
resistance equivalent to a coefficient of 0.3 can be achieved.
An interesting question in the context of rolling is about the role of
adhesion at the regions of contact. The first point to observe is that the
rolling process imposes much gentler deformation on the surface than
the sliding process. Consequently, break-up of surface films is likely to
be less marked so that strong adhesion is less likely but cannot be ruled
out completely. Secondly, even if adhesion does occur, the junctions are
peeled apart. This is a much easier process than shearing which actually
takes place in sliding contact. The overcoming of interfacial adhesion
consumes only a small part of the total energy expended during rolling.
Thus, even if a lubricant reduces the adhesion, this has very little effect
on the total rolling friction. This is the basic reason why lubricants play
a small part in rolling friction.

3.3 Rolling friction force


A fundamental principle describing rolling of a cylinder over a flat surface was formulated by Coulomb (3). It says that the rolling friction
force is directly proportional to the applied load and inversely
proportional to the radius of the cylinder
TGf

(3.1)

where r denotes the radius of the cylinder and f is the coefficient of


rolling friction, expressed in the same units as the radius.
The rolling friction force acts in the opposite direction to the relative
motion and its point of application is not precisely defined. The rolling
friction coefficient depends on the type of material and surface finish
but is independent of rolling velocity. According to Reynolds (1), the
rolling friction coefficient is represented by tangent ac (Fig. 3.1).
Figure 3.3 shows the forces acting on a cylinder rolling over a nominally flat substrate. Rolling motion is due to the application of force
F. During steady state motion
FGT,

NGZ,

NfGTr

(3.2)

Fundamentals of Rolling Motion

61

Fig. 3.3

In Fig. 3.3, represents the shift of reaction force Z and is called the
arm of rolling friction.
The Coulomb law does not have a general character as, depending
on the mechanical properties of the materials of the cylinder and the
substrate, the relationship between friction force and load may be
expressed by a different equation and, sometimes, the effect of rolling
velocity may be included in it. Because of the practical importance of
rolling contacts, a considerable amount of research effort has gone into
studies of rolling resistance. As a result, new equations taking into
account conditions under which rolling occurs have been derived. In
Fig. 3.4 a schematic representation of forces acting on a driving wheel
of a vehicle is shown. Shift a of a normal reaction depends on the
deformation of the tyre, although the magnitude of a is influenced by
the normal load on the wheel and the tractive force. The coefficient of
resistance to rolling for a rigid wheel is given as the ratio of the force
resisting rolling Pk and the normal reaction Zk

rG

Pk

Pk

Zk Gk

(3.3)

Shift of the reaction force Zk is caused, in the case of a rigid wheel, by


deformation of the substrate. The coefficient of rolling friction is given
by
a

rG

(3.4)

62

Rolling Contacts

Fig. 3.4

In the case of a deformable wheel


a Mk rkArd
rG A
rd Zk rk rd

(3.5)

where rd denotes the dynamic radius of the wheel, i.e. the distance
from its centre to the substrate over which it is rolling, and rk is the
rolling radius of the wheel, i.e. the radius of a fictitious rigid wheel
which, rolling without slip, has identical angular velocity and linear
velocity to a real wheel.

3.4 Free rolling


Free rolling is the most basic form of rolling motion. During free rolling, material ahead of the contact region is compressed, while material
to the rear of the contact zone has this stress state relieved. For these
conditions, the Hertz theory applies. Two cases will be analysed here:
a cylinder on a plane and a sphere on a plane.

3.4.1 Cylinder on a plane


The distribution of pressure p in the contact zone owing to an applied
load N is equal to
2N

pG
al

1A

x2
a2

where l denotes the length of the cylinder.

(3.6)

Fundamentals of Rolling Motion

63

The above equation represents a semicircular pressure distribution


with a maximum pressure at the centre of the contact zone of value
2N(al ), as shown in Fig. 3.5. The value of the contact semi-width a
is given by

4Nr 1A 21 1A 22
C
l
E1
E2

aG

(3.7)

where is Poissons ratio, E is Youngs modulus, r is the radius of


the cylinder, and suffixes 1 and 2 denote the cylinder and the plane
respectively.
During the forward compression of the material in contacting bodies,
a resisting moment M about the centre of the contact region is
developed
a

MG

plx dxG 3

2Na

(3.8)

Under ideal conditions the cylinder would be subjected to an equal and


opposite moment arising from the pressure distribution in the rear part
of the contact zone, but the rearward recovery does not replace the

Fig. 3.5

64

Rolling Contacts

whole of the elastic work of forward compression, which is given by


Mx

2Nax

(3.9)

3r

This imbalance may be explained by an energy dissipation due to the


elastic hysteresis loss occurring in the complex straining of the material
which must occur during the rolling process. If this loss is represented
by a coefficient , then
FxGG

2Nax

(3.10)

3r

where F is the force required to overcome the resistance to motion. The


ratio FN is usually called the coefficient of rolling friction
F 2a
rG G
N 3r

(3.11)

It must be noted that r depends on the geometry of the roller, the


applied load, and the elastic constants of the two bodies in contact.
Moreover, the hysteresis loss factor used in the expression for r is
not the same as that which would be measured in a simple tensile test.
An approximate estimate of the hysteresis loss coefficient suggests that
it should be almost 3 times the loss factor measured in a simple tension
test.

3.4.2 Sphere on a plane


The pressure distribution in the contact zone is given by
3N
pG
2a2

r2
1A 2
a

(3.12)

where r denotes the radius of the sphere.


The radius of the contact can be found from

aG

Nr

1A 21 1A 22
C
E1
E2

1/3

(3.13)

The contact configuration of the case under consideration is shown in


Fig. 3.6. The moment resisting the motion is
3N

MG 3
4a

(a Ax )x dxG 16
2

3Na

(3.14)

Fundamentals of Rolling Motion

65

Fig. 3.6

The work done in forward compression in rolling a distance x is


Mx

3Nax

16r

(3.15)

Introduction of a loss factor similar to the one discussed earlier gives


FxGG

3Nax
16r

(3.16)

Therefore, the coefficient of rolling friction is


F 3a
rG G
N 16r

(3.17)

The hysteresis loss coefficient for this case is only about twice that
which will be obtained in a simple tension test.

3.5 Material damping during rolling


The rolling of a sphere on a nominally flat plate is a cyclic process. The
stresses resulting from the contact of ball and plate are carried along
with the moving contact. Elements of volume in the material are cyclically stressed as the stress field passes through them. This cyclic stressing

66

Rolling Contacts

is accompanied with a hysteresis loss that appears to be the primary


source of energy dissipation during rolling.
Figure 3.7 shows the ball and plate in the vicinity of the contact area
subdivided into blocks of thickness x, each bounded by the sample
surface and by planes perpendicular to the direction of rolling. The
strain energy stored in each of these blocks is a function of the contact
pressure and the elastic constants of the material. The stored strain
energy in the blocks in the ball V1 and in the blocks in the plate V2 is
plotted as a function of the distance of each block from the centre of
the contact area. The exact shape of these curves is not known, but the
essential fact is that the strain energy is a maximum for the block at
the centre of the contact area. As the ball rolls a distance x, the energy
dissipated is x times the friction force F and the strain energy in the
blocks is changed according to these curves. Effectively, this energy
change between blocks is equivalent to one block in the ball and one
block in the plane being subjected to the cycle of zero strain energy to
maximum strain energy to zero strain energy. By definition, the specific
damping capacity is the ratio of the energy dissipated during a cycle to
the maximum strain energy stored during the cycle. Thus, the specific

Fig. 3.7

Fundamentals of Rolling Motion

67

damping capacity for the rolling ball is


Fx
G
V1 maxCV2 max

(3.18)

Since F is determined experimentally, all that is needed for calculation


of the specific damping capacity is the value of Vmax for ball and plate.
Choosing a coordinate system with the origin at the centre of the contact area, the x axis extending in the direction of rolling, and the z axis
extending into the sample, the complete stress distribution in the yz
plane for a sphere contacting a plane is given by
1A2 a2

xGqo

Cqo

yGAqo
Aqo

zGqo

y2


1A

z
1u

1u
z
(1A)u
a
2C 2
Aqo (1C)
tan1
1u
1u
a Cu
a

1A2 a2
y2

z 3 a2u
1u u2Ca2z2


1A

z
1u

Aqo

1u
z (1A)u
a
Aqo (1C)
tan1
A2
2
1u a Cu
1
a
u

z 3 a2u
1u u2Ca2z2

yzGqo

yz2

(3.19)

a21u

u2Ca2z2 a2Cu

(3.20)

(3.21)

(3.22)

where u is defined by the equation


y2

z2
C
G1
a2Cu u

(3.23)

and
3 P
qoG
2 a2

(3.24)

68

Rolling Contacts

In the above expression, P is the force pressing the sphere against the
plane, a is the radius of the circular contact area, and is Poissons
ratio. The strain energy in an element of volume
S

dy dz

xG0

is Vmax and is given by


x

VmaxG

[12 ( 2xC 2yC 2z )

A (x yCy zCz x )C(1C) 2yz] dy dz

(3.25)

where E is Youngs modulus. Assuming that G0.3 and expressing relevant terms in equation (3.25) by parameters given by expressions introduced earlier, the resulting double integral can be evaluated numerically
to give
VmaxG0.1315

P2
a2E

(3.26)

Equation (3.18) for the specific damping capacity thus becomes

GF 0.1315

P2 1
a2

1
C
E1 E2

(3.27)

where E1 and E2 are the Youngs modulus for the ball and plate,
respectively. Equation (3.27) is only applicable for values of the load
that do not cause any plastic deformation. When plastic deformation
does occur, the geometry is no longer a sphere contacting a flat plate
and equations giving stresses within the contact zone are no longer
valid.

3.6 Slip at the surface of contact


If a perfectly rigid ball is rolling on a perfectly rigid flat substrate, a single
contact point occurs at any moment in time, and the rolling resistance
disappears. However, in practice, all solids have a finite elasticity, and the
bodies will deform elastically or plastically if the local contact pressure
reaches the yield stress of any of the solids. Thus, the actual contact area
will be finite. It follows from the theory of elasticity that a tangential stress
may develop at the interface between the ball and the substrate. If the
local tangential stress becomes high enough, local slip will occur

Fundamentals of Rolling Motion

69

between the surface of the ball and the substrate. During this local slip,
external energy will be converted into heat motion; i.e. the slip will
contribute to the rolling resistance. The slip velocities are generally
small, usually less than 1 percent of the rolling velocity, but nevertheless
produce in many cases a major part of the total resistance to rolling.
The basic theory of slip contribution to rolling resistance is due to
Carter (4). His main interest was in the action of a locomotive driving
wheel, but he considered the simpler problem of two cylinders of radius
R and length l of like materials pressed together and rolling on one
another. One of the cylinders is subjected to a torque and the other to
an equal countertorque (see Fig. 3.8). Thus, any state of stress or strain
in one member, resulting from tangential tractive forces only, is
matched by an equal and opposite state in the other, and the distribution of pressure between the members is unaffected by the traction,
since the radial displacements of the surfaces in contact are complementary. Assuming that both bodies can be considered as half-spaces and
that the theory of elasticity is applicable, the normal stress distribution
in the contact area AaFxFa is given by the Hertz theory (5)
4Po

PG

x2

1A

a2

(3.28)

where PoGL2al is the average pressure in the contact area. Carter


used the Coulomb friction law but assumed that the static and kinematic friction coefficients are equal. Thus, in the contact area, if the
tangential stress is smaller than P(x), no slip occurs, but if the tangential stress reaches P(x), local slip occurs.

Fig. 3.8

70

Rolling Contacts

(a)

P

leading
a
edge

-a
slip zone

stick zone

(b)
1.0

F/F1= R/

0.5

0
0

2
v/v

Fig. 3.9

The solid line in Fig. 3.9(a) shows P(x), which is equal to the maximum tangential stress possible in the contact area. The thicker line A
shows the actual tangential stress distribution (x). No slip has occurred
for cFxFa where FP(x). In the other part of the contact region,
AaFxFc, slip occurs, and GP(x). It should be noted that, when the
friction force F increases, so does the size of the slip region, until F
reaches F1GL, the maximum friction force, at which point slip occurs
in the whole contact area. In general, the friction force is given by
2

2a

FGF1 1A

aAc

(3.29)

so that FF1 as c a (complete slip) and F 0 as c Aa (no slip).


Thus, for a free-rolling cylinder (no applied driving torque), there is no
rolling resistance due to local slip. When a torque is applied (with constant rolling velocity), the friction force F0, and the rolling speeds of
the two cylinders differ. If R 1 ( is the angular velocity) and R 2 are

Fundamentals of Rolling Motion

71

the rolling speeds of the two cylinders, then the ratio vv, where vG
R( 1C 2)2 is the average velocity and vGR( 1A 2 ) is the difference in the rotational velocities, increases in proportion to the fraction
(aCc)2a of the contact area where slip occurs
v 4(1A)Po (aCc)
G
v
Ga

(3.30)

where G is the shear modulus. Using equations (3.29) and (3.30), it is


possible to express the friction force as a function of vv
FGF1[1A(1A vv)2]

(3.31)

where

G
8(1A)Po

The solid line in Fig. 3.9(b) shows FF1 as a function of vv. It is


obvious that FGF1 when
1 8(1A)Po
G G
v
G

If only elastic deformations take place in the contact area, the pressure
Po must be below the plastic yield stress c of the solids in contact.
Since Gc 100 for steel, and since typically 8(1A)1, it follows
that at the onset of complete slip vvF0.01. In the wheelrail contact
area, the pressure Po is usually close to the plastic yield stress and in a
typical case vv0.005 at the onset of complete slip. Thus, if, for
example, v20 ms, then v0.1 ms.
It is quite clear that the static friction coefficient is generally higher
than the kinematic friction coefficient. Furthermore, the steady sliding
friction coefficient may be velocity dependent and usually decreases
with increasing sliding velocity. Taking into account these two effects,
the relation between F and v is of a qualitative nature, as indicated by
the dashed line in Fig. 3.9(b). Another important effect may be the
dependence of the static friction force on the time of stationary contact.
If a train moves with the velocity v and if the contact area between the
wheel and the rail has the diameter 2a, then the maximum time of
stationary contact (for no slip) is *G2av104 s. In the example,
2a0.01 m and v100 ms. This is a very short contact time and it is
clear that the static friction coefficient will, in general, not have reached
its maximum value before the contact is broken. The actual dependence

72

Rolling Contacts

of the static friction force on the time of stationary contact is known


for the wheelrail system and is probably not very well defined as it is
sensitive to the nature of the contamination layer, which varies with
the spatial and temporal location. It is well known, however, that the
maximum traction of a locomotive decreases with increasing rolling
speed, as is expected and in line with what has been said above.
The curve denoted by the dashed line in Fig. 3.9(b) forms the basis
for designing anti-skid braking systems. Since the friction is maximum
at point A in Fig. 3.9(b) an optimally designed braking system should
keep the motion of the wheel close to this point during emergency braking. However, even with a feedback system, it is difficult to stay continuously close to point A, since the dashed-line curve is a function of
time, i.e. the coefficient of friction varies with changing road surface
conditions. Most anti-skid systems for automobiles use a pulsed system
where the wheelroad surface slip oscillates around the maximum
denoted by A, continuously adjusting itself to the road conditions. In
fact, the most important effect is not the increase in the sliding friction
compared with gross slip (locked wheels), but rather the fact that rolling
wheels allow steering or directional control which is otherwise lost.

3.7 Internal friction


During rolling, different regions of the contact interface are first
stressed, and then the stress is released as rolling continues. Each time
a volume element in either body in contact is stressed, elastic energy is
stored by it. Most of the elastic energy is later released as the stress is
removed, but the cyclic deformations cannot occur entirely adiabatically and some energy will be converted into heat by the internal
friction. This transfer of energy from the rotational and translational
motion to heat will contribute to the rolling resistance.
Numerous physical processes contribute to the internal friction of
a solid and are usually described by a complex shear modulus
GGG1CiG2 . For example, in metals, various modes of switching of
segments of pinned dislocations can take place. In chain polymers the
viscoelastic properties may be due to switching of chain segments
between two configurations. In both cases, the stress, on its own, does
not force the switch. Instead, its result is to bias the activation energies
of the switch in the favourable and unfavourable directions, and the
jump over the barriers occurs due to thermal excitation. For a weak
applied external stress that varies very slowly with time, the deformations occur adiabatically and the deformation vanishes as the

Fundamentals of Rolling Motion

73

external load is removed. However, at finite frequencies this is not the


case and external energy is converted into heat in the solid. Any specific
mechanism of internal friction will, in general, give rise to a characteristic peak in the frequency-dependent dissipative response G2() of a
solid at a frequency 1, where is the relaxation time characterizing
the average time a molecule (or a segment of a molecule) spends in a
particular state before jumping to the next state. Most solids have many
such peaks at different resonance frequencies. It needs to be noted that
depends strongly (in fact, exponentially) on temperature, and the
viscoelastic properties (at a fixed frequency) of a solid depend on the
inverse temperature T much as they do on the frequency, often exhibiting several peaks as a function of T.
The magnitude of internal friction varies considerably from material
to material. Thus, for a steel ball rolling on a steel substrate, the contribution from the internal friction is usually rather small. Using the
theory of viscoelasticity, it is possible to estimate that

Po Im

G()GP G CG
G2

2
1

2
2

where Po is the average pressure in the contact area; G() is evaluated


at the frequency vR, where v is the rolling velocity and R is the
radius of the ball. For steel on steel, Po cannot exceed the plastic yield
stress c 0.01G1 so that PoG1F0.01. Furthermore, the ratio G2G1 for
steel is typically smaller than 0.001 for the frequencies of interest, so
that F105. However, for a steel ball rolling on a polymer or a wheel
rolling on a road, the internal friction may contribute a substantial part
of the rolling resistance.
It was stressed earlier that the internal friction of solids depends on
the frequency of the external oscillating stress, and, in general, it will
exhibit several peaks (resonances) at well-defined resonance frequencies.
This fact is utilized in many practical applications. For example, in
order to minimize the friction losses during steady driving of an automobile, the rubber used for tyres is designed in such a way that during
rolling at normal speeds the deformations at the contact interface are
mainly elastic, while during braking the tyres behave viscoelastically,
producing high internal friction. This can be achieved since the typical
deformation frequencies during rolling are different to those during
sliding; in the latter case the relevant deformations result from surface
asperities on the road sliding relative to the tyres. This generates pulsating forces acting on the tyres at characteristic frequencies S vr,

74

Rolling Contacts

Fig. 3.10

where r 0.11 mm is the typical extent of surface asperities on the road.


With v 10100 ms this gives S 104106 s1. During rolling, the
characteristic frequency is instead the rolling frequency R vR, where
R 0.11 m is the radius of the wheel, which gives R 10103 s1.
Figure 3.10 shows schematically the complex shear modulus G()G
G1()CiG2() of a viscoelastic amorphous polymer. The shaded areas
denote the frequency intervals typical for rolling and sliding friction. It
should be noted that in the sliding regime G2HG1 , while in the rolling
region G2FG1 . It is apparent that the friction coefficient will be much
larger during sliding than during rolling.

3.8 References
(1) Reynolds, O. (1876) On rolling friction. Phil. Trans. R. Soc., 166.
(2) Tabor, D. (1955) The mechanism of rolling friction Part II: The
elastic range. Proc. R. Soc., A229.
(3) Coulomb, C. A. (1809) Theorie des machines simples-en ayant en
regard au frottement de leures parties et a la roideur des cordages
(in French), Paris.
(4) Carter, F. W. (1926) Proc. R. Soc., A112.
(5) Hertz, H. (1896) The contact of elastic bodies. Miscellaneous
Papers (Macmillan, London).

Chapter 4
Dynamic Characteristics
of Rolling Motion

4.1 Introduction
The dynamic characteristics of motion in rolling contact and their variation during rotation are of importance for the precision running of
rotors supported by a rolling contact bearing. It is quite a difcult problem to study as the rolling contact bearing represents a complex mechanical system.
A number of studies wholly or partly devoted to the problem of
stiffness in rolling contacts have been undertaken. Novikov (1) derived
expressions in a linearized form for the axial and radial stiffness of a
radial thrust bearing, taking into account the non-linearity of elastic
characteristics. Kharlamov (2) analysed the case of static equilibrium
of a radial thrust bearing when the axial load appreciably exceeds the
radial load. Simple expressions for the radial, axial, and transverse stiffness of the bearing were obtained for this case by linearizing the equations. The formulae obtained for radial and axial stiffness were
compared with those obtained by Novikov.
In the theoretical analysis carried out by Szucki (3) a procedure is
given for computation of the elasticity of a ball bearing. The load is
represented in the form of radial force, axial force, and the torque in
the plane of these forces. The assumption was made that, in a mounted
bearing, clearance and preload are both absent.
In all the above studies the dependence of stiffness on angular
rotation of the cage owing to changes in the position of rolling elements

76

Rolling Contacts

and errors of shape of the elements has not been considered. The
importance of this problem was signalled in the theoretical work of
Tamura and Shimizu (4) and Neubert (5), but the problem itself has
not been solved.
In this chapter, an advanced analysis of rolling motion dynamics will
be undertaken. Also, the different types of resistance to rolling motion
in bearings and their contribution to the total resistance will be
considered.

4.2 Analytical evaluation of friction torque


4.2.1 Friction during rolling
The resistance to relative motion in rolling contact bearings is due to
many factors, the major one being rolling friction. This was long
assumed to be the only resistance to motion in rolling contacts. As a
result of studies into the nature of rolling motion it was established that
the contribution of rolling friction to the overall resistance to motion is
small, though its effect on wear and tear and operating temperature
conditions is important. These factors are especially important for
miniature instrument rolling contact bearings operating in very accurate
mechanisms of servo-systems, magnetic recorder mechanisms, and
other precision parts of instruments.
The rst research into rolling friction was carried out by Leonardo
da Vinci in 1508. The concept of the coefcient of external friction arose
as a result of these experiments. Leonardo da Vinci concluded that for
identical surface roughness this coefcient is a constant for different
bodies and equals 0.25.
In early research the rolling friction was treated as a mechanical process; i.e. the interaction of rough surfaces of absolutely rigid solid bodies was considered. The concept of rolling friction as a mechanical
process was introduced by Delagir and was further developed in studies
by Parin, Euler, Lesley, and others (6, 7).
Coulomb was the rst to introduce the hypothesis of the dynamic
nature of external friction. The experimentally obtained relationship of
the friction force F due to rolling along a plane is given by the classic
equation
FGf

N
GN
r

(4.1)

where f is the coefcient of rolling friction having the dimension of


length, N is the normal load, r is the radius of the cylinder, and is a

Dynamic Characteristics of Rolling Motion

77

coefcient similar to the coefcient of sliding friction. This relationship,


known as the AmontonCoulomb law, relating frictional force and normal load, appeared to mark the end of a long period in the development
of the science of rolling friction (7).
In 1876, Osborne Reynolds (8) was the rst to propose the hypothesis
of relative slipping, according to which the frictional force due to rolling
of a perfectly elastic body along a perfectly elastic substrate appears to
result from relative slipping of contact surfaces on account of their
deformation. This is also discussed by Solski and Ziemba (9). It can be
seen from Fig. 4.1 that, in the region AC, surface layers of the cylinder
are compressed (in a direction along the area of contact), and in the
plane they are elongated. In the region CB of the contact area, these
deformations take place in the opposite direction, as a result of which
microslip occurs.
Akhmatov, Tabor, and Tomlinson (10, 11, 12) have shown that slipping in the contact zone is not the only source of friction losses in
rolling. Tomlinson and Tabor showed experimentally that, even during

Fig. 4.1

78

Rolling Contacts

the rolling of two cylinders or two balls of identical size made of identical materials, loss of energy occurs. According to Tomlinson, when
two bodies approach one another, atomic and molecular interactions
appear and energy is lost in overcoming them. This accounts for the
additional energy loss.
Taking into account the forces of molecular interactions during rolling of a cylinder over a plane, Tomlinson determined the frictional
coefcient as follows
KG

3
4

ef
2 1(Nr )

(4.2)

where K is the coefcient of rolling friction, N is the load on the cylinder, r is the radius of the cylinder, e is a crystal lattice constant, f is the
coefcient of sliding friction, and is a function of the elastic constants
of the materials of the bodies in contact.
Tomlinson (12) carried out experimental verication of the coefcientof rolling friction by introducing an additional coefcient related
to the coefcient of rolling friction K (mm) by the expression KG l,
where l is the moment arm applied to the rolling body. Coefcient
was determined by the damping of the oscillations of a pendulum.
The value of thus obtained was twice the theoretical value for amplitudes of pendulum oscillation ranging from 4 to 0.61, which is
explained by the fact that the relative displacement and molecular sizes
are commensurable.
Akhmatov (10) developed formulae for determination of the coefcient of rolling friction based on perfect contact between nominally
at surfaces. The works of Ishlinskii (13, 14), Drutowski (15), Palmgren
and Snare (16), Tabor (11), and many others have great practical
importance. They studied the origins of rolling friction during contact
of real solids with imperfect elastic properties. Ishlinskii considered the
resistance to motion in the case of the rolling of a perfectly rigid cylinder
over a viscoelastic surface and along a plastically deforming plane. The
contact stress distribution is shown in Fig. 4.2. The presence of frictional force is reected by the asymmetrical contact stress distribution.
Ishlinskii developed an expression for determination of the resistance
force F (at low rolling velocities) in the case of the rolling of a perfectly
rigid body along a plastically deforming plane
FG

v
N
Cr

(4.3)

Dynamic Characteristics of Rolling Motion

79

Fig. 4.2

where N is the load on the cylinder, v is the rolling velocity, is the


coefcient of internal friction of the substrate material, r is the radius
of the cylinder, and C is the hardness of the substrate material.
The coefcient of rolling friction is thus given by
KG

v
C

(4.4)

The corresponding corrections, especially in the mechanical and molecular concepts of external friction, led to further development in this
area. The large amount of research, basically of an applied nature,
devoted to the mechanical concept has improved the understanding of
the laws of elastoplastic deformation and failure.

4.2.2 Friction torque in the rolling contact


In 1919, Palmgren and Snare (16) developed a theory of resistance to
motion in roller bearings, assuming differential slipping as the main
cause of rolling friction. According to Palmgren, the contact region is
a point or a line in the case of an unloaded bearing, but in the case of
a loaded bearing surface, contacts appear owing to elastic deformation
of contacting bodies. When rolling motion commences there is,
although small, relative local displacement and slipping. The energy
spent on this slipping is considered as a loss. The hysteresis loss is added
to the losses arising from elastic slipping. These losses depend on the
load and the properties of the materials of the contacting bodies.

80

Rolling Contacts

Poritsky (17) extended the theory of friction in rolling contact by


analysing the losses resulting from sliding friction caused by the gyroscopic effect. Kalker (18) analysed roller bearing friction under axial
load as the friction due to differential slipping but he did not consider
the loss arising from the motion of the rollers. Johnson (19) considered
the movement of the rollers as a starting point to develop a theory of
friction. However, he did not take into account all the factors acting on
the roller, in particular the losses due to elastic hysteresis. The current
tendency is to generalize the early theories into a unifying theory of
total friction torque.
When considering the total friction torque it is necessary to break
down the torque into a number of its constituent components. It can
be represented by the following equation (20)
MG(MdsCMgyrCMhysCMdeCMcCMeCMT )K

(4.5)

where Mds is the friction torque due to differential slipping of rolling


elements on a raceway, Mgyr is the friction torque arising from gyroscopic spin or deviation of the axis of rotation of a rolling element,
Mhys is the friction torque arising from elastic hysteresis losses, Mde is
the friction torque due to the deviation of rolling elements from the
required geometric shape and surface roughness, Mc is the sliding friction torque along the edge of the raceways, Me is the friction torque
due to shearing of the lubricant lm, MT is the friction torque arising
from temperature effects, and K is a correction coefcient taking into
account any other factors contributing to the friction torque such as
vibration.
Friction torque due to differential sliding
Consider the friction torque caused by differential sliding, Mds , for the
case where the ball rolls along a groove with a radius of curvature R in
a plane perpendicular to the direction of rolling. As shown in Fig. 4.3,
pure rolling will occur along two lines located on an ellipse of contact
at a distance 2ac . In other parts of the contact ellipse there will be
sliding because of the unequal distance of contact points from the axis
of rotation.
Friction torque due to differential sliding of rollers can be expressed
in terms of the work done, A, by the rolling contact bearing in unit
time
AG(Fi liCFo lo )z

(4.6)

where Fi and Fo denote sliding friction caused by differential sliding


during rolling on a ball along the surface of the inner and outer ring

Dynamic Characteristics of Rolling Motion

81

Fig. 4.3

respectively, li and lo are the distances travelled in unit time by a point


located within the contact zone of the ball with the inner and outer ring
respectively, and z represents the number of balls. The above equation
is valid for the case where the inner ring rotates and the outer ring is
stationary.
The distances li and lo can be estimated from the following
expressions
li G2Rl i nb

(4.7)

lo G2Rlo (nAnb )

(4.8)

where n is the rotational speed of the inner ring of the bearing, nb is the
rotational speed of the ball, and Rli and Rlo represent the radii of curvature of the groove on the inner and outer ring respectively.
The friction torque resulting from differential sliding of the ball is
Mds G

A
2n

(4.9)

82

Rolling Contacts

or after expansion
Mds G

PDo z
4db

1A

d 2b
D

2
o

cos2 (ro BoAri Bi )

(4.10)

where is the coefcient of sliding friction, Bi and Bo are the coefcients


determined by the geometry of contact deformations of the inner and
outer race respectively, and ri and ro are the respective radii of contact
areas on the inner and outer ring.
Friction torque due to gyroscopic spin
Torque caused by gyroscopic spin, Mgyr , appears in the presence of
contact angle on the balls. In order to estimate this torque, it is
necessary to determine the moment of inertia of the ball, I
IG

60

d 2b

(4.11)

where is the density of the material of the ball.


The friction torque due to gyroscopic spin then equals
Mgyr GI b z sin

(4.12)

where b G(nb )30 is the angular velocity of the ball and the
rotational velocity of the ball is given by
nb Gnshaft

D 2o d 2b cos2
2Do db

(4.13)

where nshaft is the rotational velocity of the shaft on which the bearing
is mounted.
Torque Mgyr increases with an increase in contact angle and attains
a maximum when G2. In the case of a radial bearing, G0, so
that Mgyr G0. In order to avoid gyroscopic spin of balls, it is necessary
to satisfy the inequality
MgyrFNdb z

(4.14)

where N is the axial load. The effect of gyroscopic spin is especially


noticeable in radial thrust bearings.
Friction torque resulting from hysteresis losses
Energy losses due to elastic hysteresis in the material of contacting bodies can be determined by assuming that during the loading cycle a
specic quantity of energy is expended. Additional loading inuencing
hysteresis losses is created during high-speed rolling on account of centrifugal forces. Therefore, in all the expressions presented below, torque
Mhys must be treated as a function of rotational speed.

Dynamic Characteristics of Rolling Motion

83

Fig. 4.4

When a ball rolls along a raceway, Hertzs contact ellipse is formed


between them (Fig. 4.4). Energy developed during the rolling motion of
a body about one point of the raceway is proportional to the load p
and to the magnitude of deformation . The contact time t of the rolling
body with the raceway is directly proportional to the length of the contact area in the direction of motion 2b and inversely proportional to
the circumferential velocity r (Fig. 4.5). Thus
t

2b

(4.15)

Fig. 4.5

84

Rolling Contacts

Energy losses during contact of the rolling body with one point on the
raceway are given by
Mhys tp

(4.16)

According to the Hertz theory, where there is point contact between


two bodies the magnitude of deformation is expressed as
3
GK 1
( p3 r)

(4.17)

where r is the sum of the reciprocals of the principal radii of curvature


of the bodies in contact, given by
(r1Cr2)I

(4.18)

(r1Cr2)II
and K is a coefcient depending on the auxiliary function F(r)

F(r)G

(r1Ar2)IC(r1Ar2)II

(4.19)

Coefcient F(r) is usually determined from appropriate tables. It is


possible to consider approximately
Mhys G p43

(4.20)

The above expression is valid for a point contact when p23 and
bp13. A more exact equation for loss caused by elastic hysteresis is
given in the following form
Mhys G1.25B104Do db2/3 pi4/3

(4.21)

where pi is the load on the ith ball. In practice hysteresis losses are
independent of the shape of the raceway.
Friction torque due to geometric errors
The main cause of changes in friction torque are errors in the shape of
rolling elements. Torque due to error of shape can be expressed as
Mde G

dx
N
dk

(4.22)

where k is the angle of rotation of the moving ring relative to the


stationary ring, x is the linear displacement of the moving ring in the
direction of the load N acting on it and arising because of the deviation
in the geometric shape and surface roughness of the contacting surfaces.

Dynamic Characteristics of Rolling Motion

85

In a real radial bearing, the centre O of the circle of the moving inner
ring does not coincide with the theoretical centre O1 (Fig. 4.6). Errors
z1 and z2 arise owing to geometric deviations from the nominal
dimensions. While determining the centre O1 , it is assumed that the
bearing is loaded purely by a radial load R applied at point O1 . In the
case analysed, the load is supported by two balls which roll without
slip. Thus
Mde GR
Mde G

dx

(4.23)

dk
R

sin( z)

dk sin AAz 2D cos AC dk Cz 2D cos

di

dz1

AG

di

dz2

2
A 1
z

Fig. 4.6

(4.24)

86

Rolling Contacts

where di is the diameter of the circle of the inner ring, z1 and z2 are
the geometric errors, and Do is the diameter of the outer ring.
Friction torque arising from raceway effects
It it is assumed that the inner ring rotates about a vertical axis and
touches, under the action of its own weight, the ball at one point only,
then the friction torque due to the contact of rolling elements with the
raceway is given by an expression of the form
McI G

Do

1A

d 2b
D

2
o

cos2 sin Ctan1

db sin
2RI

G
c

(4.25)

where Gc is the raceway mass, RI is the radius of the inner ring, and db
is the diameter of the ball.
The bearing housing is in contact with the inner or outer ring and
hence at the time of operation a friction force appears as a result of
contact of the housing with the guiding edges. Consequently, the friction force is reduced or increased depending on the type of t between
the housing and guide rings. The expressions for the torque are as
follows:
housing tted on to the outer ring
M ocII G1.38B104Gc n2Do

DoAdb cos

Do

(4.26)

housing tted on to the inner ring


M icII G1.38B104Gc n2Di

DoAdb cos
Do

(4.27)

where Do and Di denote the diameter of the outer and inner ring respectively, and is the eccentricity of the housing with respect to the bearing
axis.
The total torque is the sum
Mc GMcICMcII

(4.28)

The housing is in its most suitable position when it rests on the outer
ring as there is a reduction in the deformation of the housing owing to
the centrifugal force.
Friction torque due to shearing of the lubricant lm
The presence of a lubricant between contacting elements at the rolling
contact leads to additional energy dissipation on account of the viscosity of the lubricant which, in turn, depends on the lubricant physical

Dynamic Characteristics of Rolling Motion

87

characteristic, the contact pressure, the relative velocity of lubricant


ow, and the temperature within the contact zone. The lubricant lm
formed in the contact zone prevents direct contact of rolling elements,
with a subsequent reduction in wear and tear of the interacting surfaces.
The cross-section of the contacting bodies is shown in Fig. 4.7.
The energy lost on overcoming friction in the lubricant lm per unit
length of a cylindrical roller bearing is
W oc G

6.5v2
1(ho a)

(4.29)

where vGxt is the longitudinal velocity of the point of contact during rolling of the cylinder, is the viscosity of the lubricant, ho is the
minimum thickness of the lubricating lm, and aG12 (1RxJ1R1 ).
Here, Rx is the radius of curvature of the surface of the cylinder within
the loaded contact zone, and R1 is the radius of curvature of the surface
of the raceway within the loaded contact zone.
The frictional loss for a ball has the following form
W bo G

6v2 1a
4m2
ln
(2b3a)1b
ho

(4.30)

where 2m is the maximum thickness of the lubricating lm, and


bG12 (1RyJ1R2 ), Ry and R2 are the radii of curvature perpendicular
to the direction of rolling.

Fig. 4.7

88

Rolling Contacts

The energy expended on overcoming friction is sensitive to changes


in the values of m and ho and is logarithmically dependent on the thickness of the lubricating lm. The resistance moment due to shearing of
the lubricating lm can be approximately expressed in the following
way
Me G2zW bo

(4.31)

where z denotes the number of rolling elements in the bearing.


Friction torque caused by elevated temperature
Temperature does not uniformly affect friction torque. With an increase
in temperature to around 100120 C, the friction torque actually
decreases, which is explained by the decrease in the viscosity of the
lubricant. An increase in temperature beyond 120 C could cause an
appreciable increase in the constant component of the friction torque
as a result of changes in dimensions of individual elements of a rolling
contact bearing.

4.2.3 Total friction torque


So far the discussion has been conned to the existing state of the theory
of friction torque in rolling contacts. However, the methods presented
above are not very useful for practical applications. Quite often, it is
sufcient to determine friction torque magnitude approximately,
especially in the case of a bearing of secondary importance for the
reliable operation of a whole system. In most practical applications, it
is sufciently accurate to assume that the coefcient of rolling friction,
, is 0.0010 for a self-aligning ball bearing, 0.0011 for a cylindrical roller
bearing, 0.0013 for a thrust ball bearing, 0.0015 for a deep groove ball
bearing, 0.0018 for a tapered and bevelled roller bearing, and 0.0045
for a needle roller bearing. A simplied form of friction torque is
MGF

d
2

(4.32)

where F is the load on the contact and d is the diameter of a rolling


element.
The friction torque of an assembly of rolling contact bearings and
also the torque due to individual bearings are determined in different
ways by various researchers. According to Palmgren and Snare (16),
the torque of resistance to rolling is expressed in the form
MGM1(F )CMo (n)

(4.33)

Dynamic Characteristics of Rolling Motion

89

where
M1(F)Gf1 Fdb

C
F

(4.34)

Mo (n)Gfo d 2b ( n)2/3

(4.35)

where fo and f1 are factors depending on bearing design, load, and lubrication method, c is an exponent depending on the type of bearing, is
the kinematic viscosity, Cs is the static carrying capacity of the bearing,
db is the diameter of a rolling element, and n is the rotational velocity.
The torque required to overcome the resistance to rolling motion can
be approximately calculated using the following formula
MG

Q d
2

(4.36)

where Q is the radial load on the bearing and d is the diameter of the
rolling body.
The expression for the estimation of friction torque in a ball bearing
is
MG1.4Q

d C1
Do

(4.37)

where Do is the diameter of the ball race of the outer ring and db is the
diameter of the ball.
When a radial load acts
MGMoC1.25

Do
db

Pr

(4.38)

For a bearing supporting an axial load only


MGMoC1.5

Do
Pa
db

(4.39)

When both radial and axial loads act simultaneously


MGMoC(1.5PaC1.25Pr )

Do
db

(4.40)

All these expressions give an approximate value of friction torque since


they do not take into account the inuence of many factors present
during the operation of a rolling contact bearing.

90

Rolling Contacts

The formula for the determination of the friction torque in a lubricated bearing is as follows
MG z

Do

LF

D1n
30db

[ACBCCDC16.5ao ]

(4.41)

and
AG

4.5db ao E
a2oCho db
9db ao E

BG

4a2oCho db

CG

15F 3/2
9a20 1F
A
21(2ho ) 2ho 1(2ho )

1(h d )Ctan
ao

DGtan1

EG
FG

2ao
1(ho db)

db hoCa2o
2ho
db
2

where is the viscosity of the lubricant, n is the rotational velocity of


the inner ring, z is the number of rolling elements, D1 is the diameter
of the outer ring, LF is a coefcient depending on the ball race radius
of curvature, ao is the length of the hydrodynamic lubricating lm, and
ho is the minimum thickness of the lubricant lm.
The above equation is not very handy for practical situations and an
alternative to it has the form
MG0.452

(1A )(1C )2 2

D o z 0.625B103

K1r

db
Do

KG

1
1C sin

6.5d 0.65
b

(4.42)

Dynamic Characteristics of Rolling Motion

91

where is the contact angle, Do is the pitch circle diameter of the balls,
is the allowable contact stress (usually 8502800 MPa), and r is the
radius of the ball race.
When a bearing operates under vibration conditions and if its main
axis is horizontal, then the friction torque can be computed using the
formula
MGv Q

ds

(4.43)

where Q is the radial load on the bearing, ds is the diameter of the shaft,
and v is given by

v Gk

2.8(1C ) Do
ds g

V m V mv

where Vmv is the mean relative collision velocity of the inner ring and
balls with the outer ring of the bearing, Vm is the mean frequency of
impacts of the inner ring and balls against the outer ring, is the rolling
friction coefcient, d is the ball diameter, g is the gravity constant, and
is the coefcient of speed restoration during impact.
In the case of vertical vibration of a bearing with a vertical axis of
rotation, the friction torque equation has the form
MGfA Q

ds

(4.44)

where
2k(1C )

d Ccos f d g sin V

fA G

Dri

mv

Vm

Dri is the diameter of the ball race on the inner ring, fn is the coefcient
of non-uniformity of load distribution between balls (usually 0.9), and
is the angle at which the load acts on the ball.
All these formulae give only an approximate value of friction torque
or contain parameters that can only be determined experimentally.

4.2.4 Variable components of friction torque


When a ball rolls along the ball race, the associated friction torque is
then given by the formula
MGW

(4.45)

92

Rolling Contacts

where W is the power expended during the rolling of the ball and is
the angular velocity of the ball.
Taking into account the contribution of the oil lm to rolling losses,
the power expended during rolling of the ball along the ball race can
be expressed in the following form
WG

17.3v2 1
(2 C3 )1

(4.46)

Here the speed v is practically constant, the viscosity can be considered independent of contact pressure, and and are parameters
depending on the radii of curvature of the bodies in contact
1
1
G J
;
r Rin(out)

1 1
G C
r rz

For balls and ball races of ideal geometric form, the radii r, Rin(out) , and
rz are constant. In actual bearings they are all variables.
The geometric errors in ball races can be expressed in terms of a
Fourier series. In the case of the outer ball race, the function for errors
has the form
m

1( )G Ak cos(kCk )

(4.47)

k G1

and for the inner ball race


m

2 ( )G Bk cos(k C k )

(4.48)

k G1

where m is the number of harmonics, Ak , and Bk are the amplitudes of


the harmonic components, and k and k are the corresponding phase
angles.
The amplitude of the harmonic components is expressed in the form
Ak G1(a2kCb2k)

(4.49)

where
ak G

bk G

1 cos k dxG

1 sin k dxG

1 N
oi cos kix
i G1

1 N
oi sin kix
i G1

Dynamic Characteristics of Rolling Motion

93

Here is the interval for computing errors in a ball race. Usually,


1. Phase angles are expressed in the following way

k Gtan1

bk
ak

Analyses carried out with the help of the expressions introduced above
show that the magnitude of the variable component of friction torque
is negligibly small compared with that obtained experimentally.
The second approach to determination of the variable friction torque
is as follows. Assume that the ball rolls without slip between two surfaces as shown in Fig. 4.8. The lower surface differs slightly from an
ideal plane surface and is stationary. The upper surface is an ideal plane
surface and moves with velocity v in such a way that at any moment in
time no point on it changes its position along the height [Fig. 4.8(a)].
For the displacement of the upper plane it is necessary to overcome a
certain rolling friction force Qc . Here, the ball will be either in the valley
or on the peak. During rolling of the ball over the peak [Fig. 4.8(b)], it
is necessary to do work proportional to the magnitude of elastic deformation and the height of the peak. The force Qc will also change. If the

Fig. 4.8

94

Rolling Contacts

height of the peak depends on its position along the length, then the
force Qc will be a function of time. Furthermore, the model adopted
also helps to explain the effect of lubricating oils of different viscosities
on Qc (t). It is well known that the lubricant forms a lm able to separate the contacting surfaces, and an increase in lubricant viscosity helps
to achieve complete separation. Also, with increasing viscosity the
amount of work done increases, since the deformation takes place
within a larger nominal contact area. Force Qc changes proportionally
to the work done. Thus, with an increase in the viscosity of the lubricant
the amplitudes of the variable component of friction torque increase.
Besides that, when the viscosity of the lubricant increases, the microroughness effects regarded as high-frequency components of the force
are suppressed. These assumptions correspond well with the results of
experiments which show that, with increase in viscosity, the highfrequency components of the variable friction torque are damped but
the amplitudes of the lower components are increased.
In the model outlined in Fig. 4.8 it can be seen that Qc (t) is possibly
related to the force developed by the deformation of the projection and
acting in the normal direction. The magnitude of the force when the
ball approaches the uneven surface depends on the height of the bump
and is described by the Hertz theory of elastic contacts. Thus, it is
possible to write Qc GF(P).
The friction torque developed when the ball rolls along the race can
be expressed by

MG 11.24

o 5
rrC0.5db
ao bo
ho
rr db

e n o
(n o )3/2

(4.50)

where
y 4
1 y 2
3
A
C
ao
n o ao
2(n o )3/2

BG

In the above expressions, o is the dynamic viscosity of the lubricant,


ho is the thickness of the lubricating lm, ao and bo are the major and
minor axes of the ellipse of Hertzian contact respectively, rr and db are
the radius of the ball race and the diameter of the ball respectively, n is
the pressureviscosity coefcient of the lubricant, o is the maximum
compressive stress at the centre of the contact zone, and is the angular
velocity of rolling of the ball.
Equation (4.50) was obtained on the basis of the solution of a hydrodynamically lubricated contact. Since the major axis of the contact

Dynamic Characteristics of Rolling Motion

95

ellipse is larger than the minor axis by one order of magnitude, it is


assumed that the lubricant ows only in the direction of rolling
(plane contact hydrodynamic problem). The resistance to differential
slipping when the ball rolls along the ball race has also been taken
into account. In addition, the hydrodynamic ow of lubricant, the
deformation of surfaces, and the exponential dependence of viscosity
on pressure are all included. Equation (4.50) is most suitable for heavily
loaded contacts.
An important element in the calculation of friction torque is the
thickness of the lubricating lm, ho . The following expression can be
used to determine it
hoin(out) G

3.17[ o (UaCUb )]0.75n0.6


K o0.15 ( )0.45

(4.51)

where
Ko G

3
31
3

8
4

3
1
[P 2E( )2]

G A
db

Rin G

1
J
Rin Rin(out)

0.5DinCrin

Rout G

cos in
0.5DinArin
cos in
A(rinA0.5db )
C0.5db
cos out
cos out

Ua Ub ( inA o )Din 2
In the above expressions, Ua and Ub are the relative speeds of the surfaces with respect to the contact point, is a coefcient that is a function of an elliptic integral, P is the normal load on the contact point, E
is the elastic modulus, in is the angular velocity of the inner ring, and
o is the angular velocity of the centre of the ball. Subscripts in and
out represent the inner and outer rings.

96

Rolling Contacts

Equation (4.50) can be rewritten to facilitate its use at different loads.


Thus

MG

6.11 o s65 rinC0.5db 2 1.6 3/2tP 1/3


P e
B
kt 3/2
rin db

(4.52)

where

a
ym

BG

tG

2 1 ym 2
1
A
C0.407 3/2 1/2
1/3
3 tP
ao
t P

n
s2v

sG1.08B103

kG

1/3

3.48 0.15 o0.75 (UaCUb )0.75n0.6


E 0.05 ( )0.45

ym ao is the relative position of the trajectory of pure rolling, ao is


determined from the Hertz contact theory, and v is a coefcient that is
a function of an elliptical integral.
Close estimation of the sum of two terms

a
ym
o

2 1 ym
A
3 tP 1/3 ao

shows that it does not exceed 5 percent of the term 0.407(t 3/2P 1/2).
Hence, it is safe to neglect it in further analyses.
In Fig. 4.9 the dependence of friction torque on lubricant viscosity is
shown for two sliding velocities. From this graph it can be seen that
the friction torque, even at low sliding velocity, sharply rises and can
reach a very high value. This is contrary to experimental observations
and is due to the close dependence of viscosity on pressure.
A formula to determine shear stresses in an oil lm that takes into
account the dependence of the viscosity of the lubricant on pressure by
the exponential law and on temperature by the polytropic law has the
form

21(427 Ts s)
Z
ho1 1[k2 C2A1(a1C1)]

(4.53)

Dynamic Characteristics of Rolling Motion

97

Fig. 4.9

where
ZGln

kG

1(a1C1)(k 2 C1C1{k 2 [k 2 C2A1(a1C1)]})


1[k 2 C(1Ca1)]

(UaAUb )1 s
21(427 Ts )

In the above expressions, Ts is the temperature of the lubricating lm,


s is the viscosity of the lubricant at the given pressure and temperature
Ts , is the coefcient of heat conductivity of the lubricant, and and
a1 are experimentally determined coefcients.
The friction torque due to rolling is given by

dF

MK G

(4.54)

The above equation is difcult to use because of its complexity. However, at low sliding speeds and small loads the shear stress can be

98

Rolling Contacts

computed using the classical formula for the isothermal domain

G o

(4.55)

ho

The following simplifying assumptions are introduced:


(1) The contact radius of curvature m (Fig. 4.10) is determined,
approximately, by the formula

m G

ri db

(4.56)

2riCdb

(2) The instantaneous rolling axis of the ball relative to the ball race
passes through the trace of pure rolling (i.e. rolling without slipping)

Fig. 4.10

Dynamic Characteristics of Rolling Motion

99

in the plane passing through the axis of the ring and the centre of
the ball.
(3) Since the stress in the contact zone is relatively small, it is possible
to consider that viscosity does not depend on pressure.
(4) The area of contact in the outer direction along the axis is assumed
to be plane.
According to Fig. 4.10, the distance from any point on the prole of
the contact area to the axis of pure rolling is
z

1
2m

y2

(4.57)

The sliding speed at the contact during rolling is


VG

1
( y 2mAy2 )
2m

(4.58)

The elemental resistance to sliding is equal to


dTG o

v
ho

dx dyG

o ( y2mAy2 )
dx dy
2m ho

(4.59)

The elemental friction torque in rolling with slipping is


dMGdT(zmAz)G

o
( y2mAy2) dx dy
ho 42m

The friction torque is


ao

MG4

bo

dMG

o a5o bo 4 2 2 1
ymA ymC
ho 2m
3
5

(4.60)

where ym Gymao , i.e. the position of the trajectory of pure rolling.


Detailed calculations show that y4mA23 y2m does not exceed 5 percent
of the expression in brackets. Taking this into account, the expression
for the torque, after transformation, is
MG

0.8 o s65v rinC0.5db


k

rin db

2.1

(4.61)

Equation (4.61) can be written in the following form


MGCP 2.1

(4.62)

where C is a constant for the given type and identical geometric size of
bearing.

100

Rolling Contacts

The power expended in rotation of the inner ring of the bearing is


NGNinCNout

(4.63)

where Nin and Nout are the power spent during rotation of the ball along
the raceway of the inner and outer rings respectively.
In order to overcome the resistance to inner ring motion it is necessary to apply a torque given by
MGMinC(MinCMout)

(4.64)

where b and are angular velocities of the ball and the inner ring of
the bearing respectively.
On the basis of equation (4.61), it is possible to write for one ball

MG CinC(CinCCout )

b 2.1
Pi

(4.65)

and for a bearing with s balls

MG CinC(CinCCout )

P
s

2.1
i

(4.66)

i G1

Here Cin and Cout are constants for the inner and outer rings
respectively.
Load acting on a single ball
In the case of point contact, the load acting on the ith ball bearing is
Pi GKi3/2

(4.67)

where K is a coefcient depending on the curvature of the contact


elements of the bearing at the contact points and on the elastic characteristics of the materials from which these elements are made, i is the
elastic deformation of the contacting bodies.
The deformation in the direction of the ith ball depends not only on
the force due to preliminary tension and displacement of the rotor but
also on the geometric errors in the rings. Taking the above into account,
the deformation in the direction of the ith ball in the case of a radial
bearing can be expressed in the form

i G oCx cos i cos oCy sin i cos oCz sin o


Cin ( i ) cos oAout(i ) cos 0

Dynamic Characteristics of Rolling Motion

101

where x, y, and z are displacements of the rotor centre in directions X,


Y, and Z.
The static displacement of the rotor centre can be expressed by the
equations
xGF 1out(t)AF 1in (t)
yGF 2out(t)AF 2in (t)
zGF 3out(t)AF sin (t)
(4.68)
For a radial bearing, the radial load in the direction of the ball owing
to axial preload has the following expression
Pio GK{ oC[F 1out(t)AF 1in (t)] cos i cos o
C[F 2out(t)]AF 2in (t)] sin i cos oC[F 3out(t)AF 3in (t)] sin o
Cin ( i ) cos oAout (i ) cos o }3/2
This formula is valid if the inequality oH x is satised, where x is the
variable component of deformation. The load acting in the direction of
the ith ball from the radial force is found using the formula
Pi p G4.37

P
s

(cos i )2/3

(4.69)

where P is the radial load and i is the angle between the direction of
the radial load and the position of the ith ball. In this case, i coincides
with i . Finally, the expression for the total load on the ith ball, taking
into account the preload and radial load, takes the form
Pi GPioCPi p

(4.70)

Friction torque resulting from the exponential dependence of viscosity


on pressure
The friction torque when a ball rolls along the ball race with an angular
velocity is given by
MG

The power W used by friction while rolling in the lubricating lm can


be determined in the following way. Let an element of the surface of a

102

Rolling Contacts

Fig. 4.11

ball move with velocity V to the left (Fig. 4.11). Obviously, h at point
xGx1 is a function of time t. On the basis of elementary considerations
h
t

GV tan

(4.71)

where tan Ghx, and hence


h
t

GV

(4.72)

An element of the ball surface touching the lm does an element of


work on a unit distance p dx dyh. The total work done is
h

Ap dx dy t dt

FG

Assuming the expression for the thickness distribution of a lm hG


hoC x2C y2, the power required is
WG2 V

px dx dy

(4.73)

where
12Vn o1
1
p
pGA ln 1A
n
(2 C3 )1h3o

(4.74)

Dynamic Characteristics of Rolling Motion

103

Here p in non-dimensional coordinates is expressed in the following


way
pG

(1C C 2 )2
2

In the expression for pressure distribution [equation (4.74)] the following exponential dependence of viscosity on pressure is assumed:
G o e np .
For further analysis, the following notations are introduced
RG

2Vho3/2
n1

LG

12V o n1
(2 C3 )1h3o
(4.75)

After transformation
WGR

x ln(1ALp) dx dy

(4.76)

where
pG

x
(1Cx Cy2 )2
2

The integration limits are determined by the following values

G0FxFS
Gya
From equations (4.74) and (4.75) it follows that the solution is meaningful in the interval
0FLpF1
The function ln(1ALp) is expanded into a Taylor series

(Lp)2 (Lp)3
(Lp)n
C
C C
ln(1ALp)GA LpC
2
3
n
(Lp)k
k
k G1
S

(4.77)

104

Rolling Contacts

Taking into account equations (4.75) and (4.76), where kG1, 2, 3, . . . ,


the power loss is expressed by the expression

W(L)G

k G1

Lk

x kC1
(1Cx2Cy2 )2k

dx dy

(4.78)

It is not difcult to show that equation (4.78) converges uniformly relative to k.


By putting 1Cy2 Gb2, the following is obtained
2Lk

W(L)GR

k G1

x kC1

dy

(b2Cx2)2k

dx

(4.79)

Equation (4.79) can be rearranged into the following form


S

Lk

k G1

W(L)G2R

TkC1,2k dy

(4.80)

where

x kC1

TkC1,2k G

[b2( y)Cx2]2k

dx

Assuming that
a

TkC1,2k GAk ( y);

Ck G

A ( y) dy
k

the result is
Ck k
L
k G1 k
S

W(L)G2R

(4.81)

In equation (4.81), Ck can be expressed as


a

Ck GTkC1,2k

1[(1Cy )
1

2 3kA2

dy

The limits of integration are determined by the boundaries where the


pressure is zero. Therefore, the limits of integration are taken as the
mean value of pressure isobars in the direction of the y axis. A few
isobars are taken for this purpose with different values of pressure.

Dynamic Characteristics of Rolling Motion

105

Fig. 4.12

According to Fig. 4.12, aGa( p), and therefore the following


expression applies
x

(4.82)

ApG0
(1Cx2Cy2)2
The value of a is found in the following way
a( p)G

1
x2Ax1

x2

x1

11

A(1Cx ) dx

(4.83)

where x1 and x2 are points of intersection of isobars with the x axis.


For nal computation of power losses, the value of a is taken as the
value corresponding to pG0.0001. It is interesting that, in spite of the
fact that the values of a are different for different p, the integrals
a

1[(1Cy )
1

2 3kA2

dy

(4.84)

differ very little.


Thus, the nal expression for determination of the power losses due
to the rolling of a ball along the ball race, assuming an exponential

106

Rolling Contacts

dependence of viscosity on pressure, takes the form


WG2R(2.4998LC0.03276L2C0.002619L3
C0.0005112L4C0.00009736L5C0.00002076L6 )

(4.85)

where R and L are determined from equation (4.75). It is possible to


use two terms of the series to compute the power loss, and the error
will not be more than 0.5 percent.
The friction torque due to rolling of one ball is expressed in the following form
MG

2R
(2.4998LC0.03276L2)

(4.86)

where is the angular velocity of the ball.


The friction torque of an assembly of balls when the inner ring rotates
is given by
MG

(WoutiCWini )

(4.87)

i G1

where s is the number of balls, and Wini and Wouti are the power losses
due to rolling of the ball along the inner and outer rings respectively.

4.3 Elastic and damping characteristics of the


rolling contact
4.3.1 Static stiffness of the rolling contact
Stiffness of radial bearings
The analysis of radial stiffness is based on the assumption that the
bearing remains in static equilibrium throughout rotation. According
to the Hertz theory, the total elastic force at the point of contact of the
ith ball with the inner and outer rings is expressed in the following way
Pi GKi3/2

(4.88)

where i is the deformation of both rings in the direction of the ith ball
and K is a proportionality coefcient.
Equation (4.88) is also applicable to roller bearings, in which case the
exponent is equal to 109. Thus, the total elastic force at the contact
points of rolling elements with the inner and outer rings can be

Dynamic Characteristics of Rolling Motion

107

expressed as
Pi GK ni

(4.89)

and the projection of this force on to the line of action of the applied
radial load on the bearing is given by
Pi GK ni cos i

(4.90)

where n is the exponent mentioned above, i GCi is the angle


between lines of action of the radial load (direction of displacement of
the moving ring) and the radius passing through the centre of the ith
rolling element, represents the turning angle of the cage, G2m is
the angular distance between the rolling bodies, and m is the number
of rolling elements in the bearing (iG0, 1, 2, . . . , mA1).
It is assumed rstly that the bearing elements are ideal in shape,
dimensions, and material properties. Then, the deformation in the direction of the ith rolling element is expressed by

i Gc cos(0.5 max sin i )C 1C 2 cos( max sin i )

(4.91)

where c is the radial preload or clearance, and 1 and 2 are deformations caused by mutual displacement of rings at the contact points
of the rolling element with the inner and outer rings respectively.
With reference to Fig. 4.13, the magnitude of angle max is expressed
by the formula

max G2 sin1

(4.92)

Fig. 4.13

108

Rolling Contacts

where d is the diameter of the raceway of the bearing inner ring and z
is the mutual displacement of the rings.
Usually, max is small and may be taken as being equal to zero. Then
equation (4.90) is transformed into
Pi GK(cCx cos iCy sin i )n cos i

(4.93)

where x is the displacement of the moving ring in the direction of the


radial load, and y is the displacement of the moving ring in the direction
perpendicular to the radial load.
The reaction force of the complete bearing is equal to the sum of
elastic forces of all rolling elements taking part in the transfer of load.
Thus
mA1

PG Pi GK [cCx cos(Ci )Cy sin(Ci )]n cos(Ci )


i

(4.94)

i G0

In equation (4.94) only the terms with positive values in the square
brackets are summed, since the rolling element does not participate in
load transfer when it enters the unloaded zone.
For the determination of y the condition of zero elastic force of the
bearing along the y-axis will be used. Thus
Py G0

(4.95)

since
mA1

Py G Pyi GK ni sin i
i

(4.96)

i G0

where Pyi is the projection on to the y axis of the elastic force of the ith
rolling element. Equation (4.95) can be expressed in the following form
mA1

K [cCx cos(Ci )Cy sin(Ci )]n sin(Ci )G0

(4.97)

i G0

Expanding equation (4.97) in a McLaurin series in powers of y and


limiting it to the rst two terms, the following is obtained
mA1

0GK [cCx cos(Ci )]n sin(Ci )


i G0

mA1

CyKn [cCx cos(Ci )]nA1 sin2(Ci )


i G0

(4.98)

Dynamic Characteristics of Rolling Motion

109

Hence, in a rst approximation an expression for y is obtained

i G0 [cCx cos(Ci )] sin(Ci )


yG
mA1
n i G0 [cCx cos(Ci )]nA1 sin2(Ci )
mA1

(4.99)

The elastic properties of bearings can be conveniently studied by


employing their stiffness, determined as the rst derivative of the restoring force with respect to the displacement
kG

(4.100)

Substituting equation (4.94) in equation (4.100) and taking into account


equation (4.99), the following is obtained
mA1

kGK

i G0

cCx cos iA

B cos iA

A
Bn

sin i

CBnAAF(nA1)
(Bn)2

nA1

sin i cos i

(4.101)

where
mA1

AG (cCx cos i )n sin i


i G0

mA1

BG (cCx cos i )nA1 sin2 i


i G0

mA1

CG (cCx cos i )nA1 sin i cos i


i G0

mA1

FG (cCx cos i )nA2 sin2 i cos i


i G0

(4.102)
Thus, the stiffness of each typical size of a bearing depends on preload
or clearance c, the mutual displacement of the rings under the action of
load x, and the angle of cage turn . However, equation (4.101) does
not give a clear picture of how k depends on these values and is not
useful for direct application. This equation can be used to obtain graphs
of functions k(x) and k( ), employing a computer for further approximation by simple analytical expressions.
In considering the stiffness of the bearings as a function of k(x),
depending on c as a parameter, it is useful to take G 2. For this

110

Rolling Contacts

position of the cage (radial load is exactly between two adjacent rolling
elements) yG0, and equation (4.101) is appreciably simplied. The
value yG0 also corresponds to the value G0, but this position is
unstable in a bearing with initial clearance and can be impracticable.
The derived expression for stiffness ought to be multiplied further by
coefcient K, which takes into account the design and material of the
bearing elements.
Radial and torsional stiffness of radial thrust bearings
In this section an analysis of the radial stiffness of a radial ball thrust
bearing will be carried out. It is assumed that the contract angle of all
balls is identical; i.e. it does not depend on radial displacement of the
moving ring and the rotation of the cage remains constant. The centrifugal forces that appear during the rotation of a cage with balls affect
the value of the contact angle. As a result, the angle between the balls
and the outer ring somewhat decreases, and that between the balls and
the inner ring increases. However, this phenomenon can be realized in
practice only when the inner ring rotates at high speed (tens and
hundreds of thousands of revolutions per minute), and is observed only
in ball bearings.
Let the outer ring of the bearing remain stationary while xing the
load, and let the inner ring and rolling elements be displaced to the
right as shown in Fig. 4.14. The radial clearance of the free bearing co
has an effect only on the value of contact angle . When the ball is
displaced from the initial position (the location of the centre of the ball
is at O1 ) to a position with its centre at O3 , deformation takes place
only in the region of displacement of the centre O2 , O3 and is equal to
the axial preload ca . This load corresponds to radial preload c and gives
the total given contact deformation of the ball resulting from preload

gG

c
cos

(4.103)

The total contact deformation is equal to

1 G cC xiC yi

(4.104)

where xiC yi is the deformation of the ith ball caused by the radial
displacement of the inner ring. Thus

iG

c
cos
sin
Cx
Cy
cos
cos
cos

(4.105)

Dynamic Characteristics of Rolling Motion

111

Fig. 4.14

Substituting equation (4.105) in equation (4.89) and adding the projections Pi on to the radial plane and on to the x axis, the following is
obtained
mA1

PGK

i G0

cos
c

Cx

cos i
sin i n
Cy
cos cos i
cos
cos

(4.106)

Using the above equation, it is possible to determine the radial stiffness


of a radial thrust bearing while neglecting the transverse displacements
of the moving ring
mA1

kGKn cos1An [cCx cos(Ci )]nA1 cos2(Ci )

(4.107)

i G0

Now, let a torque act in the plane xOz (see Fig. 4.15) on a radial thrust
ball bearing with a preload. As a result there appears a reaction torque
of rings through an angle . Let it be assumed that the twist appears

112

Rolling Contacts

Fig. 4.15

only in the plane xOz, and angular free transverse displacement is


absent. Then the contact angle of each ball is changed by a value i ,
depending on and i ( i G cos i )

i G C cos(Ci )

(4.108)

The angle is very small when compared with angle and hence it can
be neglected. In such a case it is assumed that the contact angle of all
balls remains unchanged and equal to . Then the deformation of the
ith ball caused by the twisting of rings will be expressed by

i G

ro sin cos(Ci )
sin

(4.109)

where rG(dkCdb )2, i.e. the radius of the circle passing through the
centres of the balls, dk is the diameter of the raceway of the inner ring,
and db is the ball diameter.
The total contact deformation of the ith ball is determined by the
expression

iG

c
ro sin cos(Ci )
C
cos
sin

(4.110)

Dynamic Characteristics of Rolling Motion

113

After summing i elastic moments from each rolling element, an


expression for elastic torque about the y axis is obtained
mA1

MGKro cosn [cCsin ro cot cos(Ci )]n cos(Ci )

(4.111)

i G0

Considering the magnitude of , it may be presumed that sin . Then


the contact stiffness of a radially supported bearing is determined from
angle as the rst derivative of the elastic moment
k GKnro sin1 cos1An
mA1

B [cCro cot cos(Ci )]nA1 cos2(Ci )

(4.112)

i G0

Consequently, the regularity of the variation in stiffness of a radial


bearing is true for radial and torsional stiffness of a radial thrust bearing. Although equations (4.107) and (4.112) have been derived for ball
bearings, they are also valid for roller bearings if nG109 is assumed.
Radial stiffness of thrust bearings
For a thrust bearing (Fig. 4.16), deformation of the ith rolling element
caused by the displacement of the inner ring in the axial direction with

Fig. 4.16

114

Rolling Contacts

axial preload and mutual twisting of rings is given by

i GcaCzCro sin cos i cos

(4.113)

where ca is the axial preload and z is the displacement of the moving


ring in the axial direction.
Considering the low value of angle and equation (4.89), the axial
load transferred by the whole bearing is expressed by
mA1

Pa GK [caCzCro cos i ]n

(4.114)

i G0

and axial stiffness is given by


ka G

P
z

mA1

GKn [caCzC ro cos(Ci )]nA1

(4.115)

i G0

In the above equation, iF0 means that the ith rolling element does
not take part in the load transfer and hence the terms in the square
brackets with a negative sign are not summed up. If the value of ca is
negative (bearing with an initial radial clearance), then for zCro ca
the axial rigidity of the bearing is equal to zero.
For a radial thrust bearing the deformation of the ith rolling element
is
x cos i

i GcaCzCro sin cos iC

(4.116)

cos i

and the axial load transferred by the bearing is expressed in the form
mA1

Pa GK

i G0

x cos i n
caCzCro cos iC
sin
cos

(4.117)

In this case it is assumed that, for small values of , sin . Differentiating equation (4.117) with respect to z, the following is obtained
mA1

ka GK sin

i G0

x
caCzC ro C
cos(Ci )
cos

nA1

(4.118)

In the case where caF0 the axial rigidity of a radial thrust bearing is
equal to zero when

x
zC ro C
ca
cos

Dynamic Characteristics of Rolling Motion

115

From equations (4.115) and (4.118) it can be found that, when there is
no radial preload or angle of mutual twisting of the rings, axial stiffness
does not depend on the turn angle of the cage if the bearing is mounted
ideally and accurately on a perfect rigid rotor without radial preload.
If the shaft allows a constant deection both in magnitude and direction, or if the moving ring is skewed or x0, the axial rigidity of the
bearing varies periodically as it rotates. Hence, in actual conditions,
axial stiffness of the bearing ka , like radial or torsional stiffness,
depends on the turn angle of the cage .
Another characteristic case will be considered now. Let the moving
ring of the bearing have a twist with respect to the axis of the shaft.
Then the deformation of the ith rolling element is expressed by
x cos i

i GcaCzC ro C

cos

cos Cr cos q
i

(4.119)

and the axial stiffness of the bearing is


mA1

ka GKn sin

i G0

c CzCr Ccos cos(Ci )


x

Cro o cos q(Ci )

nA1

(4.120)

where o is the angle of twist of the moving ring with respect to the axis
of rotation and q is a coefcient depending on the relation of rolling
element diameters to the circle of their centres. Since, in general, q and
are not repeated values, it follows that ka is not a periodic function
if 0 and o 0 simultaneously.

4.4 Dimensional accuracy and contact stiffness


Although components of precision rolling contact bearings can be
manufactured with high accuracy, certain errors in the geometric shape
and dimensions and deviations in the properties of materials are
unavoidable. These errors affect only the periodic component of bearing stiffness. It is known that even in ideal bearings the amplitude of
the periodic component of stiffness is considerable in the case of initial
clearance. Hence the effect of errors is considered only for the case of
preloaded bearings when the change in stiffness caused by imperfections
of geometric shape and the characteristics and sizes of rolling elements
outdo the changes in stiffness in an ideal bearing.

116

Rolling Contacts

4.4.1 Radial stiffness as a function of inaccuracies


During the rotation of a bearing, the rolling elements roll along the
raceways of the outer and inner rings and, hence, their contact points
continuously change. In the case of an ideal bearing, this condition has
no real signicance. In an actual bearing, the rolling elements are either
climbing over bumps or entering troughs while rolling along the raceway and the mutual position of the unevenness of the outer and inner
rings changes continuously.
At the initial moment let the ball or roller be in contact with point 1
(Fig. 4.17) of the stationary outer ring and with point 2 of the inner
ring rotating in an anticlockwise direction. After a time t, point 2 moves
away from stationary point 1 by an angle in t, where in is the angular
velocity of the inner ring. The ball is displaced during this time in the
same direction by an angle t, and tF in t since the angular velocity
of the cage, , is less than in . Then the rings make contact with the
ball at points 1 and 2. This means that the point of contact of the ball
with the outer ring moves along the prole of the ring in an anticlockwise direction through an angular distance t, and the point of contact
of the ball with the inner ring moves in a clockwise direction through
a distance in tA tG( inA )t, or q t, if
qG

in
A1

(4.121)

Fig. 4.17

Dynamic Characteristics of Rolling Motion

117

Thus, during rotation of the cage through an angle tG, the contact
point between the stationary ring and the ball moves through an angular distance and the contact point of the inner ring with the ball
moves through a distance q. For the ith rolling element these angular
distances can be written respectively in the form Ci and qCi , where
G(2)m is the angular distance between the rolling elements, m
denotes the number of rolling elements in the bearing, and iG
0, 1, 2, . . . , mA1. It is possible to consider the raceway as a circle of variable radius. Since it is a closed curve, the function describing it is periodic
and can be expanded in a Fourier series. Taking the above into account,
the prole of the raceway of the outer ring can be represented by
RoC Rl sin[l(Ci )C l ]

(4.122)

and the prole of the raceway of the inner ring is given by


roC rp sin[ p(qCi )C p ]

(4.123)

where Ro and ro are the constant components of the radii of the raceways of the outer and inner rings respectively, l and p are the orders of
the harmonic of unevenness in the raceways of the outer and inner rings
respectively, Rl and rp are the amplitudes of these harmonics, and l
and p are the phase angles.
Consider the limit of summation of the series given by equations
(4.122) and (4.123). Obviously, it is meaningful to take the upper limit
as innity since, for sufciently large values of l and p, unevenness of the
raceway is covered by the contact area and stops affecting the periodic
component of stiffness. Besides, the upper limit of quantities l and p,
which will be denoted by s, cannot be determined rigorously and
uniquely since it depends on the value of the preload, the mutual displacement of the rings, the mechanical properties of the materials of
the components, and other factors. The lower limit in equation (4.123)
is equal to unity since, without exception, all lower harmonics of errors
in raceways of the rotating ring affect the change in contact deformation during rotation under the condition xGconst. It is impossible to
say the same about the stationary outer ring. The rst harmonic of
unevenness is caused by eccentricity which does not affect the change
in contact deformation since the bearing is automatically centred in the
inner ring. If, on account of design or some other factor, the centre of
the outer ring still does not coincide with the centre of rotation of the
inner ring, this misalignment is taken as the value of x in equation
(4.124) given below. Thus, the lower limit of summation in equation

118

Rolling Contacts

(4.122) or the minimum value of l should be taken to be equal to 2, but


the minimum value of p in equation (4.123) should be taken to be equal
to unity.
Let the value of the radial displacement x of the inner ring with
respect to the outer ring remain constant during rotation of the bearing.
Then the total contact deformation at places where the rings rub against
the rolling elements is determined by the equation

i G giC(x1Cx0 ) cos(Ci )C( y1Cy0 ) sin(Ci )


s

C Rl sin[l(Ci )C l ]
l G2

C rp sin[ p(qCi )C p ]
p G1

(4.124)

where gi GgCDi and Di is the deviation of the actual diameter of


the ith rolling element from the nominal diameter, x1 is the mutual
displacement of rings in the x direction under the action of radial load,
x0 is the mutual displacement of rings in direction x for zero radial
load, y1 denotes the mutual displacement of rings in direction y under
the action of radial load in an ideal bearing, and y0 represents the
mutual displacement of rings in direction y for zero radial load. Also,
x0 GxAx1 .
Then substitute the expression for total contact deformation [equation (4.124)] in the equation dening the applied radial load on the
bearing
Pi GK ni cos i
where n is the power index (for ball bearings nG32 and for roller
bearings nG109) and i is the angle between the line of action of the
radial load (direction of displacement of the moving ring) and the radius
passing through the centre of the ith rolling element.
Summing with respect to i, an equation for the radial elastic force for
the entire bearing is obtained
mA1

PG Ki giC(x1Cx0) cos(Ci )Cy0 sin(Ci )


l G0

C Rl sin[l(Ci )C l ]
l G2
s

C rp sin[ p(qCi )C p ] cos(Ci )


p G1

(4.125)

Dynamic Characteristics of Rolling Motion

119

where Ki is the coefcient of proportionality. Its value, except for a


number of geometric parameters, depends on the mechanical properties
of the material of the rolling elements and the rings, i.e. the modulus
of elasticity and Poissons ratio. It can therefore be said that the value
of K for the contact of each rolling element with the rings depends on
the mechanical properties of the material of a particular rolling element
and can be used to describe the differences in the mechanical properties
of the materials. In equation (4.125), there is no y1 , because here a
bearing with a preload is considered for which the values of y1 are small
and only slightly affect the stiffness.
Values of x0 and y0 are determined from the condition that, in the
absence of radial load, the projections of the elastic force of the complete bearing on to axes x and y are equal to zero, i.e. PG0 and
Py G0.
In this case, x1 G0 and y1 G0, and the condition written above takes
the form
mA1

Py G Ki giCx0 cos(Ci )Cy0 sin(Ci )


l G0

C Rl sin[l(Ci )C l ]
l G2
s

C rp sin[ p(qCi )C p ] sin(Ci )


p G1

(4.126)

and
mA1

PG Ki giCx0 cos(Ci )Cy0 sin(Ci )


i G0

C Rl sin[l(Ci )C l ]
l G2
s

C rp sin[ p(qCi )C p ] cos(Ci )


p G1

(4.127)

By expanding equations (4.126) and (4.127) in a McLaurin series in


powers of x0 and y0 and limiting the terms containing x0 and y0 to the
zero and rst order, the following is obtained
ACBny0CCnx0 G0
BCCny0CEnx0 G0
(4.128)

120

Rolling Contacts

where
mA1

sin

AG Ki giCC
i G0

mA1

sin

BG Ki giCC
i G0

mA1

mA1

CG Ki giCC
i G0

mA1

EG Ki giCC
i G0

nA1

sin i cos i
n

DG Ki giCC
i G0

cos
2

nA1

cos2 i

G Rl sin[l(Ci )C l ]
l

l G2
s

G rp sin[ p(qCi )C p ]
p

p G1

Solving the system of equations given above with respect to x0 and y0


results in
xG
yG

AEACD
n(C 2ABE )
BDAAC
n(C 2ABE)
(4.129)

After substituting these expressions in equation (4.125) and differentiating with respect to x1 , the formula for radial stiffness of a radial
bearing incorporating the errors in the components is obtained
mA1

kGn Ki giCx1 cos(Ci )C


i G0

BDAAC
2

n(C ABE )

AEACD
cos(Ci )
n(C 2ABE )

sin(Ci )CC
l

nA1

cos2 (Ci )
(4.130)

Dynamic Characteristics of Rolling Motion

121

4.4.2 Effect of variable dimensions and variable stiffness


In order to demonstrate the effect of variable dimensions and variable
elasticity of rolling elements on the radial stiffness of the bearing it is
assumed that: (i) there are no other errors, (ii) the bearing has an even
number of rolling elements, and (iii) there are two identical defective
rolling elements placed opposite each other. Under such conditions
y1 Gy0 Gx0 G0
and equation (4.130) for the radial stiffness is considerably simplied
mA1

kGn Ki [giCx cos(Ci )]nA1 cos2(Ci )

(4.131)

i G0

The maximum allowable value of variable dimensions of the rolling


elements depends on the class of accuracy and the dimensions of the
bearing. Coefcient Ki takes into account the difference in the mechanical properties of the rolling elements. Any difference in this coefcient,
such as that between the rolling elements included in one assembly, is
due to the prevailing manufacturing technology. The assembly of bearings does not guarantee that rolling elements with identical mechanical
properties will be tted into a particular bearing. Nor does it exclude
the possibility of a difference in mechanical properties arising during
heat treatment. The deviation in the value of Ki in equation (4.131) is
equal to the deviation in the contact rigidity of the ball, since
ki GnKi nA1
i

(4.132)

In Fig. 4.18 the function k( ) is graphically shown according to equation (4.131) for mG6 and nG32. The graphs correspond to the following values of the parameters involved:
(1) Ki G1 and gi G20 m for all values of i (the case of an ideal bearing).
(2) Ki G1 for all values of i, go Gg3 G19.7 m, and g1 Gg2 Gg4 Gg5 G
20 m (the stiffness of all balls is identical, but the deformation
by preload of two balls is lower than the deformation of the remaining balls by 1.5 percent because of negative deviation in their
diameters).
(3) Ko GK3 G0.985, K1 GK2 GK4 GK5 G1, and gi G20 m for all values
of i (the stiffness of two balls is lower than the rigidity of the
remaining balls by 1.5 percent; the diameter of all balls is identical).

122

Rolling Contacts

Fig. 4.18

A comparison of the graphs shows that classication of the rolling contact bearings, especially precision bearings with preload, on the basis
solely of dimensional and geometric parameters is not enough. It is
necessary to verify the rigidity of the rolling elements too. Besides, it is
not rational to manufacture precision bearings with a small number
of rolling elements since the change in their rigidity during rotation is
considerable even in the case of absolute dimensional uniformity of the
rolling elements.

4.4.3 Effect of waviness of raceways


Assuming that the unevenness of the raceway of the outer and inner
rings of the bearing comprises only harmonics whose order is equal to
m, or less or more than that by a factor of 2, and that phase angles of
all harmonics are equal, i.e. x0 Gy0 G0, it is possible to simplify
equation (4.130) appreciably
mA1

kGn Ki giCx cos(Ci )C Rl sin[l(Ci )C l ]


i G0

l G2

C rp sin[ p(qCi )C p ]
p G1

nA1

cos2(Ci )

(4.133)

Dynamic Characteristics of Rolling Motion

123

The graphs shown in Fig. 4.19 were plotted using equation (4.133) for
mG6, xG0, Ki GKG1, gi GgG5, qG1.64, and l G p G0. All the
parameters correspond to three cases:
(a) the inner ring is ideal, and the raceway of the outer ring has three
harmonics of unevenness whose amplitudes are R3 G1, R6 G0.5,
and R12 G0.3;
(b) the outer ring is ideal but the inner ring has the same type of
unevenness;
(c) both rings simultaneously have the unevenness described above.
As a result of superposing the periodic component owing to the unevenness of the raceways of the inner ring on the component arising from
the unevenness of the outer ring, and in the general case also owing to
variable dimensions, variable elasticity, and change in the position of
rolling elements during rotation (x0), the function k( ) becomes
almost periodic because the rotational speed of the moving ring and the
cage are, in general, not repeated.
The repetition of the rotational speeds is characterized by the quantity q, which is determined by the relation
qG

in
A1

(4.134)

However
nin

nc

2Do
DoAdr cos

(4.135)

where nin and nc are the rotational speeds of the inner ring and cage
respectively. Also
nin

nc

in

(4.136)

Do is the diameter of the circle passing through the centres of the rotating elements, and dr is the diameter of the rolling element.
By substituting equation (4.135) in equation (4.134), the following is
obtained
qG

2Do
A1
DoAdr cos

(4.137)

124

Rolling Contacts

Fig. 4.19

Dynamic Characteristics of Rolling Motion

125

Thus, the quantity q depends not only on the dimensional and


geometrical parameters of the bearing, i.e. Do , dr , and , but also
on the value of the preload. With changes in the preload the effective
values of diameters Do and dr and contact angle also change. Consequently, the quantity q not only cannot be computed with sufcient
accuracy but also does not remain a constant during the operation of
the bearing.

4.5 Ball motion in a rolling contact bearing


In modern high-speed ball bearings the pressure areas, which result
from elastic deformations at the ballring contacts, are appreciably
curved and interfacial slip can occur at most points within the pressure
areas. These slippages give rise to friction forces acting on the balls
which are held in equilibrium by reactions from the rings and the inertia
effects of the motion of the balls.
In this section, a method is derived for determining the motion of the
ball and sliding friction in a high-speed, angular contact ball bearing
under thrust load in terms of the inertia effects on the ball and the
friction resistances resulting from interfacial slip at the contact areas.
Possible elastic compliance at the interface, hysteresis, and dynamic perturbations of ball motion are neglected.

4.5.1 Inertia forces and moments acting on the ball


The instantaneous position of an element of mass in the ball of a highspeed, angular contact bearing is shown in Fig. 4.20. The following
notation is used (Fig. 4.20): x, y, zGxed, right-handed coordinate
system with x being the axis of the bearing about which the mass
centre of the ball revolves; x, y, zGright-handed coordinate system
with the origin at the mass centre of the ball and revolving about x
at the radius e (x is parallel to x); U, V, WGright-handed coordinate system with the origin at the origin of x, y, z and moving with
x, y, z; U is directed along the axis of rotation of the ball about
its own centre; W is in the plane of U and z; U, r, Gcylindrical
coordinates rotating with the ball; Gangle between W and z axes;
Gangle between the trace of uz on xy and the x axis; Gangle
between the z and z axes.
The ball motion is dened by G B and G e . The position of the
mass element dm in the ball at location U, r, is related to the xed

126

Rolling Contacts

Fig. 4.20

coordinates x, y, z through
UGU

(4.138)

VGr sin

(4.139)

WGr cos

(4.140)

xGU cos cos AV sin AW sin cos

(4.141)

yGU cos sin CV cos AW sin sin

(4.142)

zGU sin CW cos

(4.143)

xGx

(4.144)

Dynamic Characteristics of Rolling Motion

127

yGe sin Cy cos Cz sin

(4.145)

zGe cos Ay sin Cz cos

(4.146)

From equations (4.138) to (4.146)


xGU cos cos Ar(sin sin Csin cos cos )

(4.147)

yGe sin CU(cos sin cos Csin sin )


Cr(cos sin cos Ccos cos sin )
Ar sin sin cos cos

(4.148)

zGe cos CU(cos sin sin Csin cos )


Cr(cos sin sin Ccos cos cos )
Cr sin sin cos sin

(4.149)

Treating , , B , and e as constants, two differentiations of equations (4.147) to (4.149) with respect to time yield expressions for the
instantaneous accelerations of a mass particle of the ball as
x Gr 2B (sin sin Csin cos cos )

(4.150)

y G2 B e r(cos cos sin Acos sin cos


Asin sin sin sin )C 2e [e sin CU(sin sin
Acos sin cos )Cr(cos sin Acos cos sin
Csin sin cos cos )]C 2B r(cos sin cos
Acos cos sin Csin sin cos cos )

(4.151)

z G2 B e r(cos cos cos Ccos sin sin


Asin sin sin cos )C 2e [e cos CU(cos sin sin
Asin cos )Cr(cos sin sin Acos cos cos
Asin sin cos sin )]C 2Br(cos sin sin
Acos cos cos Asin sin cos sin )

(4.152)

For the purposes of evaluating the forces and moments acting on the
ball and referred to x, y, z, the value of can be arbitrarily chosen.

128

Rolling Contacts

Setting G0 and with material density


r

Fx G

Fy G
Fz G

(r2AU 2)

(r2AU 2)

(r2AU 2)

xr d dr dU

(4.153)

yr d dr dU

(4.154)

zr d dr dU

(4.155)

The moments about the x, y, z axes are


r

Mx G

(r2AU 2)

{y [U sin Cr cos cos ]

Cz [U cos sin
Cr(cos sin Asin sin cos )]}r d dr dU
r

My G

(r2AU 2)

(4.156)

{x [U sin Cr cos cos ]

Az [U cos cos
Ar(sin sin Csin cos cos )]}r d dr dU
r

Mz G

(r2AU 2)

(4.157)

{x [U cos sin Cr (cos sin

Asin sin cos )]Cy [U cos cos Ar(sin sin


Csin cos cos )]}r d dr dU

(4.158)

The integrations give


Fx G0

(4.159)

Fy G0
Fz Gme

(4.160)
2
e

(4.161)

Mx G0

(4.162)

My GIp B e sin

(4.163)

Mz GIp B e cos sin

(4.164)

where m is the mass of the ball, and Ip is the polar moment of inertia
of the ball.

Dynamic Characteristics of Rolling Motion

129

4.5.2 Relative motions of the rolling elements


The contact conguration between the ball and the outer race is shown
in Fig. 4.21. The pressure surface is an ellipse of semi-axes ao and bo .
The radius of curvature of the deformed pressure surface in the plane
of the paper is Ro . Because the ratio bo ao is small, the radius of curvature of the pressure surface in the rolling direction can be taken as
innity with little error.
Assume that the ball centre is xed in the plane of the paper. Let the
outer race rotate with angular velocity o . The components of the ball
rotational angular velocity that lie in the plane of the paper are x and
z .
There is some radius, as ro , that sufces for determination of the
relative translational velocities of ball and outer race. This radius is not

Fig. 4.21

130

Rolling Contacts

necessarily restricted to points that lie on the deformed pressure surface


if gross slip between ball and race occurs.
According to the Hertz theory
Ro G

2fo d

(4.165)

2foC1

where fo is the ratio of outer race curvature radius to ball diameter, and
d is the ball diameter.
Through o cos o , a point (xo , yo ) on the outer race has the linear
velocity V1o
V1o G o cos o
B

cos C1(R Ax )A1(R Aa )C1(r Aa )


e

2
o

2
o

2
o

2
o

2
o

(4.166)

V1o Ge o
A[1(R 2oAx2o)A1(R 2oAa2o)C1(r 2Aa2o)] o cos o

(4.167)

Through x cos o and z sin o , a point (xo , yo ) on the ball has the
linear velocity V2o
V2o G( x cos oC z sin o )
B[1(R 2oAx2o)A1(R 2oAa2o)C1(r 2Aa2o)]

(4.168)

and the velocity with which the outer race slips on the ball in the Y
direction is
Vyo GV1oAV2o

(4.169)

Thus
Vyo GAe oC( x cos oC z sin oA o cos o )
B[1(R 2oAx2o)A1(R 2oAa2o)C1(r 2Aa2o)]

(4.170)

Through y all points within the pressure area have a velocity of slip
of race on ball, Vxo , in the direction of the major axis. Therefore,
Vxo G y [1(R 2oAx2o)A1(R 2oAa2o)C1(r 2Aa2o)]

(4.171)

Dynamic Characteristics of Rolling Motion

131

Through the components of velocity that lie along the line dened by
o there is a spin of the race, so , with respect to the ball

so G o sin oC x sin oA z cos o

(4.172)

A similar condition exists at the inner race contact as shown in Fig.


4.22. The pressure surface is an ellipse of semi-axes ai and bi . The radius
of curvature of the deformed surface in the plane of the paper is denoted
by Ri . As with the outer race contact, the radius of curvature of the
pressure surface in the rolling direction is taken as innity.
In Fig. 4.22, ri is the effective rolling radius which determines the
relative translational velocities of ball and inner race. As with ro , ri is
not restricted to points within the inner race pressure area if gross slip
occurs.
The radius of curvature Ri is
Ri G

2fi d

(4.173)

2fiC1

Fig. 4.22

132

Rolling Contacts

Through i cos i a point (xi , yi ) on the inner race has the linear velocity V1i . Thus
V1i Ge iC[1(R2iAx2i)A1(R2iAa2i)C1(r 2Aa2i)] i cos i

(4.174)

Through x cos i and z sin i , a point (xi , yi ) on the ball has the
linear velocity V2i
V2i G( x cos iC z sin i )
B[1(R2iAx2i)A1(R2iAa2i)C1(r 2Aa2i)]

(4.175)

The velocity with which the inner race slips on the ball in the Y direction
is Vyi
Vyi GV1iAV2i

(4.176)

Vyi GAe iC( x cos iA z sin iC i cos i )


B[1(R 2iAx2i)A1(R 2iAa2i)C1(r 2Aa2i)]

(4.177)

Through y , all points within the area have a velocity of slip of race
on ball in the direction of the major axis
Vx G y [1(R 2iAx2i)A1(R 2iAa2i)C1(r 2Aa2i)]

(4.178)

Through the components of velocity that lie along the line dened by
i there is a spin si of the race with respect to the ball

si G i sin iA x sin iC z cos i

(4.179)

It is convenient to write the expressions for the linear velocities of slip


and the angular velocities of spin in terms of the parameters , , and
B.
From Fig. 4.20

x G B cos cos

(4.180)

y G B cos sin

(4.181)

z G B sin

(4.182)

Dynamic Characteristics of Rolling Motion

133

Then, from equations (4.170) to (4.172) and (4.177) to (4.179)


Vyo GAe oC[1(R 2oAx2o)A1(R 2oAa2o)C1(r 2Aa2o)]

cos cos cos oC

B
sin sin oAcos o o
o

Vxo G[1(R 2oAx2o)A1(R 2oAa2o)C1(r 2Aa2o)] o

(4.183)

cos sin
o

(4.184)

B
cos cos sin oA sin cos oAsin o o
o
o
2
2
Vyi Ge iC[1(R i Axi )A1(R 2iAa2i)C1(r 2Aa2i)]

so G

(4.185)

cos cos cos C sin sin Ccos


i

(4.186)

Vxi G[1(R 2iAx2i)A1(R 2iAa2i)C1(r 2Aa2i)] i

cos sin
i

(4.187)

si G

B
B
cos cos sin iC sin cos iCsin i i
i
i

(4.188)

At the effective rolling radii ro and ri on the ball, the translational velocity of the ball and race is the same.
From Fig. 4.21

cos Cr

cos o Gro ( x cos oC z sin o )

(4.189)

(eCro cos o ) o Gro (cos cos cos oCsin sin o ) B

(4.190)

B
eCro cos o
G
o ro (cos cos cos oCsin sin o )

(4.191)

From Fig. 4.22

cos Ar cos Gr(

cos iC z sin i )

(4.192)

A(eAri cos i ) i Gri (cos cos cos iCsin sin i )

(4.193)

B
A(eAri cos i )
G
i ri (cos cos cos iCsin sin i )

(4.194)

134

Rolling Contacts

For the outer race to be stationary, the retainer must be given the absolute angular velocity c , such that

c G o

(4.195)

Then the inner race rotates with the absolute angular velocity

i G iC c

(4.196)

From equations (4.191) and (4.194)

o G

ro (eAri cos i ) cos cos cos oCsin sin o


i
ri (eCro cos o cos cos cos iCsin sin i

(4.197)

and from equations (4.194) to (4.197)

iG

i
1CA1

(4.198)

and

o G c G

A i
1CA2

(4.199)

where
A1 G

ro (eAri cos i ) cos cos cos oCsin sin o


ri (eCro cos o) cos cos cos iCsin sin i

A2 G

ri (eCro cos o ) cos cos cos iCsin sin i


ro (eAri cos i ) cos cos cos oCsin sin o

Also

BG

A i

(4.200)

B1CC1

where
B1 G
C1 G

ri (cos cos cos iCsin sin i )


eAri cos i
ro (cos cos cos oCsin sin o )
eCro cos o

Dynamic Characteristics of Rolling Motion

135

Likewise, for the inner race to be stationary the retainer must be given
the absolute angular velocity c such that

c G i

(4.201)

Then the outer race rotates with the absolute angular velocity o

o G oC c

(4.202)

From equations (4.191), (4.197), (4.201), and (4.202)

o
1CA3

oG

i G c G

(4.203)
A o
1CA4

(4.204)

where
A3 G
A4 G

ri (eCro cos o) cos cos cos iCsin sin i


ro (eAri cos i ) cos cos cos oCsin sin o
ro (eAri cos i ) cos cos cos oCsin sin o
ri (eCro cos o) cos cos cos iCsin sin i

and

BG

o
B2CC2

(4.205)

where
B2 G
C2 G

ro (cos cos cos oCsin sin o)


eCro cos o
ri (cos cos cos iCsin sin i )
eAri cos i

4.5.3 Friction at the contact interface


Figure 4.23 is an enlarged view of the pressure area at either race contact as viewed from outside the ball. Owing to the spin velocity s and
the linear slip velocities Vy and Vx , an element of area dA at coordinates
(x, y) has a resultant velocity of slip V of the race on the ball acting at
an angle with respect to the Y direction.

136

Rolling Contacts

Fig. 4.23

The friction force with which the race acts on the ball is dF, taken
over the area dA, and is in the same direction as V. The normal pressure
is distributed over the elliptical pressure area in accordance with
Sxy G

3P
2ab

x2 y2
1A 2A 2
a b

(4.206)

With the coefcient of sliding friction , dF is


dFG

3P
2ab

x2 y2
1A 2A 2 dx dy
a b

(4.207)

Let
xGaq

(4.208)

yGbt

(4.209)

thus
dFG

3P
1(1Aq2At 2) dq dt
2

(4.210)

Dynamic Characteristics of Rolling Motion

137

The component of dF parallel to the Y direction is cos dF and, for the


whole ellipse
Fy G

3P

(1Aq2)

(1Aq )

1(1Aq2At 2 ) cos dq dt

(4.211)

The total force in the X direction is


Fx G

3P

(1Aq2)

1(1Aq2At 2) sin dq dt

(4.212)

(1Aq )

The moment of dF about the normal at the centre of the pressure ellipse
is
dQs G cos( A ) dF

(4.213)

With
kGba

(4.214)

dQs Ga1(q2Ck2t 2) cos( A ) dq dt

(4.215)

Integration over the whole ellipse produces


3P
Qs G
2

(1Aq2)

(1Aq )

1(1Aq2At 2 ) 1(q2Ck2t 2 ) cos( A ) dq dt


(4.216)

where

Gtan1

kt
q

(4.217)

The moment of dF about the y axis is


dQy G[1(R 2Ax2)A1(R 2Aa2)C1(r 2Aa2)] sin dF

(4.218)

Integration over the whole ellipse gives


3PR
Qy G
2

(1Aq2)

1 (1Aq2)

1(1Aq2At 2 ) sin [AABCC] dq dt


(4.219)

138

Rolling Contacts

where

1
1
1

AG

1Aq2

BG

1A

CG

The moment about an axis through the centre of the ball, perpendicular
to the line dening the contact angle and lying in the plane x, z, is
dQR G[1(R 2Ax2)A1(R 2Aa2)C1(r 2Aa2)] cos dF

(4.220)

For the whole ellipse


3PR
QR G
2

(1Aq2)

(1Aq2)

1(1Aq2At 2 ) cos [AABCC] dq dt


(4.221)

where A, B and C are dened above.


The value of is found from Fig. 4.23 as
tan G

s sin AVx
s cos CVy

(4.222)

and
tan G

ktAVx (a s )
qCVy (a s )

(4.223)

Figure 4.24 shows the moments acting on a ball. Figure 4.25 shows the
forces acting on a ball.
From Fig. 4.24
AQRo sin oCQso cos oCMzCQRi sin iAQsi cos i G0

(4.224)

and
AQRo cos oAQso sin oCQRi cos iCQsi sin i G0

(4.225)

MyAQyoAQyi G0

(4.226)

Dynamic Characteristics of Rolling Motion

139

Fig. 4.24

From Fig. 4.25


APo cos oAFxo sin oCFzCFxi sin iCPi cos i G0

(4.227)

Po sin oAFxo cos oCFxi cos iAPi sin i G0

(4.228)

FyoCFyi G0

(4.229)

In addition
T
GPo sin oAFxo cos o GPi sin iAFxi cos i
n

(4.230)

There is also a constraint imposed by the physical dimensions of the


bearing and the elastic deformations at the ball and race contacts.

140

Rolling Contacts

Fig. 4.25

Figure 4.26 shows the initial and nal relative positions of the ball
centre and the race curvature centres. In Fig. 4.26, is the initial contact angle of the bearing. The race centres are originally at (O) and (3)
while the ball is at (1).
When equilibrium is attained under high-speed conditions, the inner
race curvature centre has moved axially from (3) to (4), by an amount
H . The ball has moved from (1) to (2). These movements result from
the elastic deformations o and i occurring at the outer and inner race
contacts

oG

iG

Ko Po2/3
d 1/3

Ki Pi2/3
d 1/3

(4.231)

(4.232)

Dynamic Characteristics of Rolling Motion

141

Fig. 4.26

Then

Ko Po2/3
Ki Pi2/3
( foA0.5)dC 1/3 cos oC ( fiA0.5)dC 1/3 cos i
d
d
G( foAfiA1)d cos

(4.233)

The pitch radius of the bearing is also a function of the dynamic effects
so that

Ko Po2/3
eGeoC ( foA0.5)dC 1/3 cos oA( foA0.5)d cos
d

(4.234)

A nal condition is that the input and output torques on the races of
the bearing are equal and opposite.
The torque on the outer race is
r o Gn
Q

QRo (eCro cos o )


CQso sin o
ro

(4.235)

142

Rolling Contacts

and on the inner race


r i Gn
Q

QRi (eAri cos i )


ri

AQsi sin i

(4.236)

Then
QRi (eAri cos i )
QRo (eCro cos o )
CQso sin oC
AQsi sin i G0
ro
ri
(4.237)
Determination of the ball motion requires the evaluation of eight
unknowns. These are ro , ri , , , Po , Pi , o , and i . The eight necessary simultaneous equations are as follows.
From equations (4.224) and (4.225)
0GAQRo (sin oCcos o )CQso (cos oAsin o )
CQRi (sin iCcos i )AQsi (cos iAsin i )CMz

(4.238)

From equation (4.226)


MyAQyoAQyi G0

(4.239)

From equation (4.227)


APo cos oAFxo sin oCFzCFxi sin iCPi cos i G0

(4.240)

From equation (4.229)


FyoCFyi G0

(4.241)

From equation (4.230)


T
APo sin oCFxo cos o G0
n

(4.242)

Again, from equation (4.230)


T
APi sin iCFxi cos i G0
n

(4.243)

From equation (4.233)

Ko Po2/3
0G ( foA0.5)dC 1/3 cos o
d

Ki Pi2/3
C ( fiA0.5)dC 1/3 cos iA( foAfiA1)d cos
d

(4.244)

Dynamic Characteristics of Rolling Motion

143

Finally, from equation (4.237)


QRo (eCro cos o )
QRi (eAri cos i )
CQso sin oC
AQsi sin i G0
ro
ri
(4.245)
The solution of the eight simultaneous equations, equations (4.238) to
(4.245), cannot be attained analytically since closed-form solution of the
double integrations involved in some of them is not possible. However,
iterative methods used in conjunction with a computer enable a numerical solution for any particular case, by means of which the ball motion
is completely dened.
Throughout the analysis it has been assumed that Coulomb friction
law is applicable. Actually, the coefcient of friction is a complex function of a number of variables. Among these are: the unit contact pressure and sliding velocities at different points within the contact area,
the nature of the contacting surfaces, the temperature, and the type of
lubricant. The functional relationship between all factors is not known.
However, measurements of friction torque in actual ball bearings have
shown good agreement with calculated results using a coefcient of
friction of 0.060.07.

4.6 References
(1) Novikov, L. Z. (1964) On the elastic characteristics of radial-thrust
ball bearings (in Russian). Izv. AN USSR, OTN, Mekhanika and
Mashinostroenie, (3).
(2) Kharlamov, S. A. (1962) Stiffness of radial-thrust ball bearings
with axial loading (in Russian). Izv. AN USSR, OTN Mekhanika
and Mashinostroenie, (5).
(3) Szucki, T. (1955) Analysis of stiffness of radial ball bearing (in
Polish). Z. Nauk. Warsaw Technical University. Mechanika, (27).
(4) Tamura, H. and Shimizu, H. (1967) Vibration of rotor based on
ball bearings. Bull. JSME, 10(41).
(5) Neubert, G. (1968) Der Einuss der Lager den Lauffehler einer
Walzgelagerten Werkzeugmashinen-Hauptspindel (in German).
Maschinenbautechnik, 17(6).
(6) Kostetskii, B. I. (1970) Friction lubrication and wear in machines
(in Russian) (Tekhnika, Kiev).
(7) Kostetskii, B. I. and Edigoryan, F. C. (1964) Classication of the
basic types of wear and elements of the theory of wear due to

144

(8)
(9)
(10)

(11)
(12)
(13)
(14)

(15)
(16)

(17)
(18)
(19)

(20)

Rolling Contacts

rolling friction (in Russian). Trudy Kievskogo Grazhdanskogo Vozdushnogo Flota, 4.


Reynolds, O. (1876) On rolling friction. Phil. Trans. R. Soc., 166.
Solski, P. and Ziemba, S. (1965) Problems of dry friction (in Polish) (Polish Scientic Publishers (PWN), Warsaw).
Akhmatov, S. A. (1951) Inuence of prole and physicochemical
properties of rubbing surfaces on the type of dependence of
friction resistance on lubricant. In Proceedings of 1st All-Union
Conference on Friction and Wear, Moscow (USSR Acad. Sci.).
Tabor, D. (1955) The mechanism of rolling friction. Proc. R. Soc.,
A229.
Tomlinson, G. A. (1929) Molecular theory of friction. Phil. Mag.,
(7).
Ishlinskii, A. Yu. (1938) Prikladnaya Matematika i Mekhanika (in
Russian), 2.
Ishlinskii, A. Yu. (1949) Friction and wear in machines (in
Russian). In Proceedings of All-Union Conference, Moscow, Vol.
2 (USSR Acad. Sci.).
Drutowski, R. C. (1959) Energy losses of balls rolling on plates.
Trans. ASME, J. Basic Engng, June.
Palmgren, A. and Snare, B. (1957) Inuence of load and motion
on the lubrication of and wear of rolling bearings. In Proceedings
of IMechE Conference on Lubrication and Wear, London.
Poritsky, K. E. (1950) Stresses and deections of cylindrical bodies
in contact. Trans. ASME, J. Appl. Mech., 17.
Kalker, J. J. (1973) Simplied theory of rolling contact. Delft progress report, Series C, Vol. 1.
Johnson, K. L. (1959) The inuence of elastic deformation upon
the motion of a ball rolling between two surfaces. Proc. Instn
Mech. Engrs, 173.
Sprishevskii, A. I. (1969) Rolling contact bearings (in Russian)
(Moscow, Mashinostroenie).

Chapter 5
Rolling Contact Bearings

Rolling motion has been most extensively utilized in rolling contact


bearings. Ball and roller bearings are typical representatives of that.
They are considered to be machine elements that have found a use in
almost all kinds of machinery and devices with rotating parts. Their
properties have frequently contributed to technical and economic progress in different branches of engineering. Standardization of rolling
contact bearings has made possible the assignment of their design and
manufacture to specialists. Thus, the machine designer is not required
to be expert on fundamental bearing problems.
However, in order to utilize the properties of rolling contact bearings
to the best advantage, engineers not only must be well acquainted with
applied bearing engineering as used in their particular elds but also
must have a general knowledge of the bearing elements themselves.

5.1 Phenomenology of friction during rolling


Rolling contact bearings are commonly considered as offering lower
resistance to motion than sliding bearings. Although rolling friction,
or rather rolling resistance, is of small magnitude, it is a complicated
phenomenon. There is no complete theory for predicting the magnitude
of the rolling resistance under all possible conditions of bearing operation. Thus, only an outline of the principles involved will be given here,
together with limited information concerning the frictional resistance in
rolling contact bearings.

146

Rolling Contacts

Fig. 5.1

During rolling motion of a body along a substrate, certain forces


resisting the motion develop within the contact zone. This can be illustrated by considering two bodies made of nearly perfectly elastic material and in contact with each other (Fig. 5.1). The load on the contact
causes deformation of the contacting bodies so that a denite area of
contact surface is produced. If one of the bodies begins to roll over the
other one, the material in both bodies will be gradually compressed
in the forward parts of the contact area and gradually relieved of the
compressive stress in the rear parts. It is known that the relation
between load and deformation is different for increasing load and
decreasing load, although the limit of elasticity of the material is not
exceeded. When the load is increasing, a given deformation corresponds
to a higher stress than when the load is decreasing. Thus, an elastic
hysteresis exists, which could be compared to the magnetic hysteresis.
Rolling action, therefore, causes the surface pressure to be somewhat
higher in the forward than in the rear part of the contact surface, which
affects the resistance to the rolling motion. The difference in pressure
for stresses within the limit of elasticity depends, to a certain degree,
on the speed with which loading and unloading occurs. Therefore, the
resistance to rolling must be inuenced by the rolling speed. Temperature will also affect the rolling friction, as elastic properties are a
function of temperature.
It is known that elastic hysteresis is dependent not only upon the
properties of the material, the rolling speed, and the temperature but
also on the radii of curvature of the contacting surfaces and on the
specic load acting on the contact. These last two factors determine the
magnitude of deformation, and thus the magnitude of the hysteresis
losses. When the loads are heavy and produce appreciable plastic
deformation, the hysteresis losses increase considerably.
When rolling occurs under load perpendicular to the contact, a forward displacement of the load resultant in the contact surface develops.

Rolling Contact Bearings

147

If an additional tangential force is also acting in the contact surface,


this load displacement will considerably increase. Any tangential force
resulting from the sliding friction at the interface develops elastic deformations both in a tangential and in a perpendicular direction. Perpendicular direction deformation causes the material to be pushed up into
something resembling a bulge in front of the contact surface and in the
direction of the tangential force, and also to be stretched into a cavity
behind the contact surface (Fig. 5.2). The resultant effect may be different for different directions of the tangential force in relation to the
direction of rolling and for different elastic properties of the material.
When the rolling of the one body takes place in the same direction as
the tangential force on the other body (the substrate), the rolling resistance will increase, particularly if the substrate is deformed more or is
of more plastic material than the rolling body. However, the resistance
decreases if rolling takes place in the opposite direction to the tangential
force.
In theory, there is one more kind of resistance to rolling in addition
to the two described above. This is the resistance produced by damping
of the elastic vibrations that occur under uneven contact pressure
between two bodies in rolling contact. If the substrate is uneven or
the rolling body unbalanced, vibrations develop that have their energy
absorbed by the internal friction in the material, and thus contribute to
the increase in the resistance to rolling.
Apart from resistances due to the elastic properties of the material,
resistance due to slippage between the surfaces, resulting from their
shape (the original shape and the shape caused by deformation under
load), should also be considered.
Two perfect cylinders should roll on each other without slip when
their axes of rotation are exactly parallel and their limited length and

Fig. 5.2

148

Rolling Contacts

Fig. 5.3

the length of the contact zone in the direction of rolling are neglected.
However, if one or the other of the two surfaces has a curved generatrix,
the actual contact surface between the two bodies will have a curved
shape (Fig. 5.3). The various points in the contact area will thus be
located at different distances from the centre of rotation of the roller.
As the circumferential speed of each point relative to the centre of
rotation is the product of the radius and the angular velocity of the
whole body, the various points will move with different circumferential
speed. This means that only a few points can be considered to have
pure rolling motion on the substrate, while all the other points will slide
with different velocities; some in a forward direction and others in a
reverse direction. Those points which roll without slip form one or two
lines, the neutral lines A (Fig. 5.3) being parallel to the direction of
rolling. In the case depicted by Fig. 5.3, the points within area I on the
roller slide backwards in relation to the direction of rolling, and those
in area II slide in a forward direction. The location of the neutral lines
is determined by the requirement of equilibrium, that the geometrical
sum of the sliding friction forces in area I, the same forces in area II,
and the outside tangential forces on the roller should be zero. This
means that the distance between the neutral lines and their location
relative to the centre of the contact surface is not always the same but
varies with the distribution of the perpendicular forces and with the
magnitude and direction of the outside tangential forces acting on the
rolling element.
Figure 5.4 shows the case where only one neutral line, A, develops.
The forces in the contact surface form a moment that tends to turn the

Rolling Contact Bearings

149

Fig. 5.4

roller out of position. This is called pivoting. Only when the rolling
element is spherical can pivoting take place continuously. For any other
shape of the rolling element, the pivoting moment must be eliminated by
the action of a counteracting outside moment in order for rolling to continue. The location of the neutral line depends, in this case, on the outside
tangential forces which tend to displace it to one side or the other.
Energy losses resulting from this type of sliding constitute a large
proportion of the rolling resistance and are equal to the total of the
product of sliding force and sliding distance, summed over all points of
the contact surface. These losses are therefore inuenced not only by
the perpendicular load and the material properties but also by the
properties of the lubricating lm at the existing temperature, and nally
by the deviation of the radii of the different points in the contact surface
from the radius at the neutral line. This deviation depends on the original shape and the deformation of the two bodies and on the magnitude
of the tangential forces, as the latter inuence the location of the neutral
lines, and the locations of these in turn inuence the sliding distances
of different points located on the contact surface.
Microsliding within the contact surface also develops for other
reasons than those discussed above. As a result of deformation caused
by the perpendicular force, those parts of the contacting surfaces that
are located within the area of contact expand or contract. The line acb
on the periphery of the rolling element shown in Fig. 5.1 is shortened
to the length aeb, while the section adb of the substrate is increased
in length. A consequence of the differential deformation of contacting
surfaces brings about microsliding which contributes to the overall resistance to rolling. Conditions favourable to microsliding may develop as a
result of temperature at the interface. As a result of different types of
energy losses at various points of the contact surface and the different
shape and heat-dissipating ability of the two bodies in contact, the

150

Rolling Contacts

temperature increase may be higher in one part of the surface than in


another; a net result of that is an unequal heat expansion leading to
microsliding. A form of microsliding develops owing to the inuence of
outside tangential forces. If the rolling element is in contact with the
substrate and a tangential force acts in a certain direction, elastic deformation in the tangential direction takes place. During rolling, microsliding occurs in the direction of the tangential force at the points located
close to the periphery of the contact surface. The rolling element maintains a continuous motion in that direction. This special condition,
which develops despite the fact that the points closer to the centre of
the contact surface do not slide, has its roots in the known circumstance
that the perpendicular load is not evenly distributed over the contact
surface, but varies so that it is highest at the centre and decreases to
zero at the periphery. As the coefcient of sliding friction is almost
constant over the entire area, the tangential force acting on a surface
element rst attains the sliding friction force at or in the vicinity of the
periphery and causes a microsliding there, while the friction force near
the centre is still sufcient to prevent microsliding.
A practical consequence of the phenomenon described above is that
a ball never can remain rolling between surfaces that are inclined to
each other, regardless of the magnitude of the angle of inclination.
While rolling, the ball is always seeking surfaces that are exactly parallel. Macrosliding, which is the simultaneous sliding in a certain direction
of all points within the contact surface, occurring coincidentally with
rolling, should not be considered to be a part of the proper resistance
to rolling. Such macrosliding between rolling surfaces develops in the
case of non-parallel cylinders, or of cones with non-coinciding apexes.
Finally, resistances that are caused by the bodies coming in contact
with the lubricant and the atmosphere must be added to the overall
rolling resistance. At the contact surface only a very small amount of
lubricant participates in the rolling process as the excess lubricant is
squeezed out. Within and immediately adjacent to the contact region,
pressure and suction forces develop in the lubricant that oppose the
movement of the surfaces towards or away from each other. This
increases the hysteresis action during rolling. The resistance due to the
motion of bodies through the surrounding atmosphere is less important
but, nevertheless, must often be considered under practical conditions.

5.2 Friction torque


Friction in ball and roller bearings results from several types of movement within the contact surface. In the contact surfaces between rolling

Rolling Contact Bearings

151

elements and raceways, rolling takes place. Because of the curved shape
of the contact surfaces and other deviations from the requirement for
pure rolling, microsliding develops at the interface. Moreover, certain
losses due to sliding friction must always occur at the contacts between
the rolling elements and the cage, between the rolling elements and
the guide anges in roller bearings, between the bearing parts and the
lubricant, and within the lubricant itself. Even the atmosphere may
be considered to be a source of losses as the bearing elements move
in it.
Apart from the material of the bearing and the nish of the contact
surfaces, the following factors affect the total friction torque:
(1)
(2)
(3)
(4)
(5)
(6)

The bearing design.


The bearing dimensions.
The load and its distribution over the rolling elements.
The speed of rotation.
The quantity of lubricant.
The lubricant characteristic and properties at the operating temperature, namely:
(a) the viscosity,
(b) the lubricity,
(c) the strength of the surface lm.

An equation to calculate the friction torque, encompassing all important parameters, would be rather complex. Besides, the inuence of some
individual factors on the friction torque is not known in an analytical
form. Studies have usually been limited to examining the inuence of
different loads and speeds for a specic bearing size under constant
conditions of lubrication for various types of bearing. It is, however,
possible to put forward a simple formula for friction torque based on
test results
MF GM0CMGCMD

(5.1)

where M0 is the moment with zero loading. This is primarily due to


friction with the lubricating lm and, therefore, decreases with decreasing viscosity and increases with the speed of rotation and the amount
of lubricant.
In the loaded bearing, friction develops at the contact surfaces
between rolling bodies and raceways. This friction is composed of a
resistance to sliding motion, MG , and a damping resistance, MD . During rolling, microsliding takes place within a pressure region where a
boundary lm of lubricant is formed. Therefore, resistance to sliding

152

Rolling Contacts

motion, MG , does not follow the laws of hydrodynamic lubrication but


is nearly independent of viscosity, speed of rotation, and amount of
lubricant in the system.
For a loaded ball bearing that has a constant contact angle, the
moment due to microsliding usually varies according to the formula
MG Gf (const F 4/3Cconst F 5/3)

(5.2)

where f is the coefcient of sliding friction in the pressure region and


F is the load acting on the bearing. The constant depends on the bearing
design and dimensions as well as on the load combination.
The moment due to damping resistance, MD , is caused by the elastic
hysteresis of the material. Experiments have shown that the damping
losses increase with F 43 for a bearing with point contact and a constant
contact angle, while in the case of a deep groove ball bearing it is of
the same order of magnitude as the loss due to microsliding.
An approximate formula applicable to all types of bearing can be
written as
MF GM0CfF F c

(5.3)

Coefcient fF depends on the size of the bearing and its construction.


It also depends on the bearing material and the lubricant characteristic.
Load F is the bearing load or, more precisely, the imaginary radial or
thrust load that would give the same friction moment as the actual
combined load acting on the bearing. The magnitude of the exponent c
is different for different bearing types. Generally, it is between 1.0 and
1.2 for roller bearings with good roller guiding and between 1.2 and 1.6
for various types of ball bearing and different load conditions; it is
essentially dependent on the sliding loss caused by the curvature in the
contact surfaces.
During the start-up period of a ball or roller bearing, some dry friction must be overcome in the region where microsliding takes place.
Therefore, the starting torque is slightly higher than the torque during
steady state rotation.

5.2.1 Friction coefcient


When carrying out approximate calculations, it is useful to use a constant value of the friction coefcient. In reality, the friction coefcient
varies considerably, but a value can be chosen that can be applied to
normal operating conditions and adequate lubrication without introducing any serious error. The friction coefcient values given below

Rolling Contact Bearings

153

represent the bearing load that will give a 1B109 revolutions life for the
respective bearings (1).
Self-aligning ball bearings

f G0.0010
Cylindrical roller bearings with ange guided short rollers

f G0.0011
Thrust ball bearings

f G0.0013
Single row deep groove ball bearings

f G0.0015
Tapered and spherical roller bearings with ange guided rollers

f G0.0018
Needle bearings

f G0.0045
All the above friction coefcients are referred to the bearing bore.
Higher values apply for new bearings, particularly roller bearings that
do not have run in guide anges. This is also the case when starting
bearings and when using an excessive amount of lubricant or a lubricant
with high viscosity.

5.3 Contact stresses and deformations


5.3.1 Contact between elastic bodies
During a loaded contact between two bodies with curved surfaces, an
area of contact is created, the form of which depends on the curvatures
of the surfaces at the point of contact, and the extent of which is a
function of the force with which the bodies are pressed together. At
zero load, the contact generally is a mathematical point and is commonly called point contact. However, when the two surfaces have the
same curvature in any common plane, the contact is a mathematical
line and is called line contact. This is the case, for example, in the
contact between parallel cylinders. When one body is pressed against
the other, an area of contact is formed; in the case of point contact
it is an ellipse or a circle, whereas in the case of line contact it is a
trapezoid or rectangle. The area of contact increases in size with

154

Rolling Contacts

increase in the load. Also, the bodies approach each other as a result
of the attening or deection that takes place in and adjacent to the
area of contact.
The magnitude of deformation that develops is clearly dependent on
the stresses in the material which are felt to some depth below the surface of the body. The stresses of a compressive nature are highest at the
surface itself, but the shear stress, which is more serious because of the
potential to cause surface fatigue or cracks, seems to develop inside the
body at a certain small depth under the surface.
The theory developed by Hertz (2) and its later modications and
amplications are used to calculate stresses and deformations. The
results are in accurate agreement with experiment regarding the deformations and the compressive stresses at the surface. It is thus possible
to calculate the form and size of the contact area, the normal pressure
on the contact area, its distribution within the area, and the compressive
deformation of the body. However, as far as the static and dynamic
bearing capacity is concerned, only a certain amount of guidance can
be obtained from the present theory. Therefore, calculations of bearing
capacity and endurance, of importance for practical application, ought
to be supported by very comprehensive experimental data.
It is mainly because of the friction between bodies in rolling contact
that the shape and size of the contact area is of interest. The nearest
approximation to pure rolling is obtained when the two surfaces in
contact are parallel cylinders, or cones with a common apex. For any
other form of the contact surface, more or less pronounced sliding
motions develop.
The theory of contact between elastic solids developed by Hertz
assumes that the proportionality limit of the material is not exceeded
and that the surfaces of the bodies are perfectly smooth, and thus no
friction forces are allowed to exist in the area of contact. Half the major
axis, half the minor axis, and the normal stress in the case of point
contact can be calculated as a function of an auxiliary quantity F(r) (3).
This is dependent on the principal radii of curvature which are located
in perpendicular planes, and also on the angle between the planes of
principal curvature of the two bodies. When these planes coincide
F(r) G

(r1Ar2)IC(r1Ar2)II
r

(5.4)

where r1 and r2 are the reciprocals of the principal radii of curvature of


the respective bodies at the point of contact, and rG(r1Cr2)I

Rolling Contact Bearings

155

C(r1Cr2)II is the sum of the reciprocals of all the principal radii of


curvature. The radius of curvature is considered to be positive for a
convex surface and negative for a concave surface.
The half-axes of the contact ellipse are

1
1

aG a E

(5.5)

(5.6)

bG b E

where a is the major half-axis, b is the minor half-axis, W is the normal


force acting on the contact, a and b are coefcients dependent on F(r) ,
and E is a material constant.
Constant E can be calculated from the equation

E G

1
3

115.5B109(EICEII)
EI EII

(5.7)

where
EI G

EI
1A1v2I

EII G

EII
1A1v2II

In the above expressions, EI and EII are the moduli of elasticity (Nm2)
and 1vI and 1vII are the Poissons ratios for bodies I and II
respectively.
The elastic compression, or the approach of distant points in the two
bodies, is obtained from the equation

G 2E 31(W 2 r)

(5.8)

where is a coefcient whose magnitude is a function of F(r) . The


mean normal pressure in the contact area is

mG

W
ab

(5.9)

and the maximum pressure at the centre of the area of contact is

max G1.5 m G

1.5W
ab

(5.10)

156

Rolling Contacts

From equations (5.2) and (5.3) it follows that

max G

1.5
a b 2E

1
3

W( r)2

(5.11)

Assuming that the permissible load on the contact surface depends on


a certain limiting value of max , and that a constant relationship
between the radii of curvature in the contact exists (as in the case of
bearings of the same type but of different size), i.e. if r is proportional
to 1Dw , where Dw is the diameter of the rolling element, this limiting
value for max is obtained when WD 2w reaches a certain value. The
quantity WD 2w is known as the specic load and denoted by .
For line contact between cylinders, the half-width of the contact area
is
bG0.0105B103E3/2

lw r

(5.12)

where lw is the length of contact (m).


Deformation in line contact is rather difcult to express in a simple
formula. When a cylinder of nite length is pressed between two plane
surfaces of innite size, and all three bodies are of the same material,
then the approach between the axis of the cylinder and a distant point
in either of the surfaces is approximately given by

G0.0003B103 2.7
E

W 0.9
l 0.8
w

(5.13)

The above expression may also be used with a sufcient degree of accuracy in the case of ordinary roller bearings. It is quite clear that the
diameter of the cylinder has no inuence on the magnitude of the
compression.
The mean normal pressure in the contact area is

mG

W
2blw

(5.14)

and the maximum normal pressure along the centre-line of the contact
is
4
max G m

(5.15)

Rolling Contact Bearings

157

By inserting equation (5.9), the following formula for the maximum


normal pressure is obtained

max

60B107

3/2
E

W
lw

(5.16)

When setting r 1Dw is permissible, as in the case of point contact,


the specic load for line contact is

W
lw Dw

5.3.2 Elastic deformations in bearings


When load is applied to ball and roller bearings, elastic deformations
will be produced. They will develop at and adjacent to the point of
contact and will have a signicant inuence on the degree of axial and
radial rigidity with which the shaft is supported. Under a load, the
bearing yields and the rings are displaced in relation to each other by a
certain small amount. This elastic action can be calculated approximately in both magnitude and direction.
Direction of elastic deformation
When a bearing is under simultaneous action of a radial load Fr and a
thrust load Fa , then the ratio of forces is
Fa
Fr

Gtan

(5.17)

where denotes the angle that the resultant load forms with the radial
plane of the bearing (Fig. 5.5). Assuming that the outer ring of the
bearing is stationary and that the load acts on the inner ring, the centre
of the inner ring is displaced in a certain direction. This direction, however, does not always coincide with the direction of the resultant load.
When the contact angle of the bearing is constant and greater than
0, and the bearing has no internal clearance, the ratio of axial to radial
displacements a r can be calculated. Negative values of a r indicate that the axial displacement takes place in such a direction that the
bearing centre moves away from the plane in which the rolling elements
are located.
For a thrust load only, Fr G0 and r G0, so the following apply:

G90,

tan tan G0,

r a G0

158

Rolling Contacts

Fig. 5.5

When tan tan G0.823 radial displacement takes place only in the
case of point contact, and when tan tan G0.785 radial displacement
occurs in the case of line contact since, under these conditions, a r G
0 and therefore a G0. It can be concluded that in all single-row angular
contact bearings a pure radial displacement occurs when
Fa 1.25 tan Fr . A single-row bearing can support a pure radial load,
i.e. G0, only when the contact angle G0. In such a case, the displacement is also purely radial, i.e. a G0.
Magnitude of elastic deformation
When the maximum load W on one rolling element in the bearing is
known, the compression in both contact areas of the most heavily
loaded rolling element can be calculated provided that the material constants and the principal radii of curvature of the surfaces at the points
of contact are given. The curvatures of the surfaces are similar for different bearing sizes and approximately proportional to the diameter
Dw of the rolling elements for any ordinary standardized bearing type.
Therefore, it is possible to derive approximate formulae of loading that
give a value for the displacement from a zero location of the bearing
centre in relation to the surrounding housing.

Rolling Contact Bearings

159

In the case of steel bearings under load conditions that give purely
radial deection ( a G0), the following applies

rG

0.0032B103

cos

W2
Dw

(5.18)

In the case of deep groove and angular contact ball bearings

rG

0.002B103
cos

1
3

W2
Dw

(5.19)

In the case of roller bearings with point contact at one raceway and
line contact at the other

rG

4
0.001B103 1
W3
1lw
cos

(5.20)

In the case of roller bearings with line contact at both raceways

rG

0.0006B103 W 0.9
cos

(5.21)

l 0.8
w

In addition to elastic deformations, displacement resulting from possible bearing clearance should also be taken into account. Under practical conditions, particularly in the case of radial bearings, deformations
in the bearing housing, which make the displacement of the shaft
greater than the calculated value, should also be included.
The axial deection under pure axial load ( r G0) is approximately
as follows.
Self-aligning ball bearings

aG

0.0032B103
sin

1
3

W2
Dw

(5.22)

Angular contact ball bearings

aG

0.002B103
sin

1
3

W2
Dw

(5.23)

Roller bearings with point contact at one raceway and line contact at
the other

aG

4
0.001B103 1
W3
1lw
sin

(5.24)

160

Rolling Contacts

Roller bearings with line contact at both raceways

aG

0.0006B103 W 0.9
sin
l 0.8
w

(5.25)

Thrust ball bearings

aG

0.0024B103
sin

1
3

W2
Dw

(5.26)

The case of deep groove ball bearings with G0 is special, because


rather large axial deections develop under pure thrust load. Even for
a very light thrust load an axial displacement of the magnitude 0.1 mm
is possible. Any internal bearing clearance increases this displacement
even further. Under heavier thrust loads, the displacement increases
4
Fa . The absolute value of the axial displacealmost in proportion to 1
ment mainly depends on the radius of curvature of the ball groove, the
internal clearance, and the ts with which the bearing is mounted, and
therefore no simple expression for calculation of the axial displacement
can be derived.
Occasionally, it is important to maintain as near constant deection
as is possible, even though the load may vary in both magnitude and
direction. A practical way to achieve this is by avoiding internal clearance and clearance due to shaft and housing ts and by using a preload.
The change in deection caused by any given change in load is therefore
reduced. In roller bearings this reduction is very small because the
deection is almost proportional to the load, and thus no advantage is
usually gained by preloading. In the case of ball bearings, the gain due
to preloading may be substantial.

5.3.3 Permanent deformations


When a sufciently high load is applied on the area of contact between
two elastic bodies so that the elastic limit is exceeded at some point of
the material, then a permanent deformation develops in the form of a
local indentation at this point. A ball pressed against a at surface with
an appropriate force creates a bowl-shaped indentation in the at piece;
the ball is also attened. In practical applications, permanent deformation occurs, even under relatively light loads, because the surfaces of
the bodies are not perfect. Highest surface asperities are rst subjected
to pressure and are attened by even a very small load because they
represent an insignicant area of support. It has to be noted that this
localized attening affects the functioning of the bearing insignicantly.

Rolling Contact Bearings

161

Fig. 5.6

With increasing load, however, the permanent deformation gradually


becomes more noticeable. Figure 5.6 is an illustration of that. At rst,
the total compression is almost entirely elastic and increases in proportion to W 23. Gradually, the inuence of permanent deformation
becomes more important and a deviation from this proportionality
takes place. This change is by no means abrupt. The increasing rate
of permanent indentation is shown in Fig. 5.7, where the permanent
deformation is a function of the load. It should be noted from this
gure that surface imperfections cause permanent deformation even at
very light loads.
The Hertz theory cannot be used to estimate the magnitude of permanent deformation since it applies only to elastic contacts. However, relatively simple formulae can be derived, which give the magnitude as well
as the variation of the permanent deformation within a limited range
of the vicinity of the elastic limit.
The point contact case can be represented by

p G1.25B108

W2
Dw

(r1ICr1II )(r2ICr2II )

(5.27)

where p is the total permanent deformation at a contact between a


rolling element and its support, W is the load at the area of contact,
Dw is the diameter of the rolling element, and r1I , r2I and r1II , r2II are

162

Rolling Contacts

Fig. 5.7

the reciprocals of the principal radii of curvature at the point of contact


between body I and II. In equation (5.27), as in the equation given
below, the radius of curvature is positive for a convex surface and negative for a concave surface.
Permanent deformation for line contact is not uniform along the
length of the roller if the roller generatrix has the same curvature along
its entire length as the support. In such a case, the permanent deformation is considerably larger at the ends of the contact zone than it is
in the middle, especially if the support has a greater length than the
roller has in its axial direction. For this case, the following equation for
the maximum total deection can be used

pG

2B1011 W
1(rICrII )
1Dw lw

(5.28)

where lw is the effective length of the roller, and rI and rII are the reciprocals of the radii of curvature of body I and II respectively in the radial
plane of the rolling element.
The distribution of permanent deformation between the rolling
element and the bearing ring is such that one-third takes place in the
rolling element and two-thirds in the ring. Before a fracture develops in
either of the two bodies in contact, the permanent deformation may
have assumed a considerable proportion. Therefore, in practice it is

Rolling Contact Bearings

163

necessary to take the risk of fracture into account only in exceptional


cases.

5.4 Load distribution within bearings


The transmission of external forces acting on the bearing occurs
through the rolling elements from one bearing ring to the other. When
two rolling elements are loaded simultaneously, a case of practical
importance, the bearing is of statically indeterminate design. This means
that the distribution of the forces among the individual rolling elements
depends on the elastic deformation of the various bearing parts.
The analysis of load distribution is important for static cases only,
i.e. when the loaded bearing does not rotate. Endurance calculations
for rotating bearings require not only the inuence of the load distribution but also the number of stress cycles to which each point in the
bearing material is subjected.

5.4.1 Radial bearings


According to Stribeck (4), the maximum load per ball for a radially
loaded radial contact bearing is
Wmax G4.37

Fr
n

(5.29)

where Fr is the radial load on the bearing. Equation (5.29) is quite


accurate for any conventional number of balls n in a bearing, provided
the internal clearance is zero.
Stribeck found that these conditions are not often satised in practice, so that the calculated value for the maximum ball load has to be
modied according to the expression
Wmax G5

Fr
n

(5.30)

A typical distribution of load over the balls in a ball bearing is shown


in Fig. 5.8. The ordinate in the diagram gives the ball load as (nWmax)
Fr , so that it is directly comparable with the constants in equation (5.29)
and (5.30). The abscissa gives the angle between the annular location
of the ball of interest and the direction of the load (Fig. 5.9). The dotted
curve in Fig. 5.8 gives the calculated load distribution for the ideal case
where there is no internal clearance and no out-of-roundness deformation of the rings.

164

Rolling Contacts

Fig. 5.8

Fig. 5.9

Elastic deformations in the areas of contact in a roller bearing


with line contact between rollers and raceways are governed by different laws to those applicable to point contact. In theory, the maximum
roller load under a radial force should be about 8 percent lower

Rolling Contact Bearings

165

than in the case of ball bearings. However, owing to the increased


clearance required by the roller bearing, a greater deviation from the
theoretical value is usually obtained. Nevertheless, practical experience
supports the use of equation (5.30) for the ordinary types of roller
bearing.
In the general case of a bearing with an arbitrary but constant contact
angle , loaded by an external force F at an arbitrary angle in relation
to the radial plane (Fig. 5.10), the analysis is more involved and can be
carried out only with the help of special tables. When the thrust load
Fa G0, the inner ring moves away from the outer ring so that the bearing load becomes concentrated on a single rolling element. If, however,
Fa increases above a certain critical value, the inner ring moves towards
the outer ring, and more rolling elements are loaded. When Fa G
Fr tan , the inner ring does not move away from the outer ring, but
only one rolling element is loaded by the force WGFr cos , while the
other rolling elements remain unloaded. When Fa G1.25Fr tan , the
rolling elements located half-way around the bearing circumference are

Fig. 5.10

166

Rolling Contacts

loaded, and the load on the most heavily loaded rolling element is
Wmax G4.37

Fr

(5.31)

n cos

If Fa is further increased to about 1.7Fr tan , all rolling elements are


loaded, i.e.
Wmax G3.93

Fr
n cos

G2.36

Fa
n sin

and Wmin G0. If is then allowed to approach 90, the bearing load
gradually becomes more evenly distributed among the rolling elements
and nally
WG

Fa
n sin

In the case of double-row bearings with a constant contact angle, an


internal thrust force develops between the two rows of rolling elements
when there is a radial load on the bearing. Thus, even under pure radial
load, the normal load distribution is obtained within the bearing, i.e.
both rows of rollers are equally loaded around half the bearing circumference. Therefore
Wmax G4.37

Fr
2n cos

where n denotes the number of rolling elements in one row.


Now, if additional thrust load is applied, the length of the loaded
zone is increased in one row and reduced in the other. When Fa G
1.7Fr tan , one row is loaded around its entire circumference and the
other row is completely unloaded. The maximum ball load in one row
is then
WI max G7.86

Fr
2n cos

G4.72

Fa
2n sin

and in the other row, WII G0. In the case of a heavy thrust load, the
loaded row acts in the same way as a single-row bearing.
There are other more complicated cases that have not been discussed
here. For example, any combination of two angular contact bearings,
mounted opposed with separated centres and loaded with radial and
axial forces as well as moments, can be analysed with the help of special

Rolling Contact Bearings

167

tables, provided the combination is symmetrical and the contact angle


and distance between bearing centres remains constant.

5.4.2 Thrust bearings


For a central force acting on a thrust bearing, the load on each of the
rolling elements is
WG

Fa
n sin

(5.32)

where Fa is the central thrust load and n is the number of rolling


elements in the bearing.
A thrust ball bearing able to support external load only in one direction with G90 may be subjected to an eccentric thrust load (Fig.
5.11) making the load distribution non-uniform. If the eccentricity of
the bearing load Fa is e and the pitch radius of the bearing is rm , one
ball is unloaded when eG0.6rm . The load on the diametrically opposite
ball is Wmax G2.36Fa n. If e is increased to 0.8rm , only half the balls
are loaded and Wmax G3.6Fa n. When eGrm , only one ball is loaded
and its load is WGFa .

Fig. 5.11

5.5 Kinematics of bearing elements


5.5.1 Rotational speed of the elements and the cage
Firstly, the case where rings of the bearing rotate with different speeds
is considered. In the analysis to follow, clockwise rotation is considered

168

Rolling Contacts

Fig. 5.12

as positive and counterclockwise as negative (Fig. 5.12). Assuming pure


rolling conditions, the absolute speed of point A on the circumference
of a rolling element, is equal to the circumferential speed vi at the instant
A contacts the inner ring raceway. At the same time, point B on the
rolling element contacts the outer ring raceway and has the absolute
speed ve . The absolute speed of each point on the rolling element is a
combination of the speed of the rolling element around its axis and the
speed of the set of rolling elements around the bearing axis. The absolute velocities of points between A and B on the diameter of the rolling
element are, at the instant shown in Fig. 5.12, parallel and vary linearly
between the velocities of A and B. As the centre of the rolling element
moves around the bearing axis only, its speed, which is the circumferential speed of the pitch circle, is given by
viCve
vm G
(5.33)
2
Because
Di ni
vi G
60
de ne
ve G
60

Rolling Contact Bearings

169

where ni and ne are the speeds of rotation of the inner ring and outer
ring respectively, Di is the diameter of the inner ring, and de is the
diameter of the outer ring.
According to Fig. 5.12
Di GdmADw cos
de GdmCDw cos
where dm is the bearing pitch diameter, Dw is the diameter of a rolling
element, and is the contact angle.
Finally, the circumferential speed of the rolling element centres and
the cage can be expressed as
vm G

120

ni d m 1A

Dw
dm

cos C

120

Dw

ne d m 1C

dm

cos

(5.34)

Consequently, the speed of rotation of the set of rolling elements and


the cage is

nm G0.5ni 1A

Dw
dm

Dw

cos C0.5ne 1C

dm

cos

(5.35)

The difference between the absolute speed of rotation of the set of rolling elements and the speed of rotation of the inner ring constitutes the
speed of rotation of the set of rolling elements relative to the inner ring.
It is given by the following expression

nmi GnmAni G0.5(neAni ) 1C

Dw
dm

cos

(5.36)

In a similar way, the speed of rotation of the outer ring relative to the
set of rolling elements can be found from

Dw

nem GneAnm G0.5(neAni ) 1A

dm

cos

(5.37)

The speed of rotation of the rolling element around its own axis is given
by
nw G

Di
nmi
Dw

nw G

de
nem
Dw

or

170

Rolling Contacts

Using one of the above expressions, the following equation can be


obtained
nw G0.5

dm
Dw

(neAni ) 1A

Dw
dm

Dw

cos 1C

dm

cos

(5.38)

It is rather rare that both rings of a bearing rotate simultaneously.


Normally, one ring is stationary. If the outer ring is stationary (ne G0)
and the inner ring rotates with the speed ni Gn, equations (5.35), (5.36),
and (5.38) respectively reduce to

nm Gnem G0.5n 1A

Dw

dm

1Ad

nmi G0.5n 1C

nw G0.5n

Dw

dm

Dw

cos

dm

cos
Dw

cos

1Cd

Dw

cos

In the case of a stationary inner ring (ni G0) and the outer ring rotating
with ne Gn, the following is obtained

Dw

nm Gnmi G0.5n 1C

nem G0.5n 1A

nw G0.5n

Dw
dm

dm

cos

cos

dm
Dw
1A
cos
Dw
dm

1Cd

Dw
m

cos

The case of a thrust bearing with G90, that is, cos G0, when ni G
n and the housing ring is stationary (ne G0) is described by the following
equations
nm Gnem G0.5n
nmi G0.5n
nw G0.5n

dm
Dw

Rolling Contact Bearings

171

5.5.2 Contact cycles due to rolling


Both the load-carrying capacity and the life of the bearing are of practical importance, and thus it is necessary to know the number of rolling
elements, fi and fe respectively, that pass a given point on the inner ring
or a given point on the outer ring, while one ring or the other makes
one revolution. Each time the set of rolling elements has moved one
full revolution in relation to one of the rings, the z rolling elements in
a row have passed every point on the ring. The number of rolling
elements passing per unit time is therefore znmi for the inner ring and
znem for the outer ring. If one ring rotates with a speed of n revolutions
per minute while the other ring is stationary, then
fi Gz
fe Gz

nmi
n

nem
n

Dw

Dw

G0.5z 1C

dm

G0.5z 1A

dm

cos

(5.39)

(5.40)

cos

In this analysis the speed of rotation of the set of rolling elements must
always be taken as positive. Besides, the number of rolling elements
passing a given point must not be confused with the number of stress
cycles at a given point, because the point in question is subjected to
load only when it passes or is passed by a rolling element that is in the
loaded zone.

5.6 Inertia forces


5.6.1 Centrifugal forces
Motion of a rolling element in a rotating bearing is associated with the
creation of a centrifugal force acting on the rings. The magnitude of
the centrifugal force is given by
TG

G
rm 2m
g

(5.41)

where G is the weight of the rolling element, g is the gravity constant, rm


is the pitch radius of the bearing, and m denotes the angular velocity of
a set of rolling elements.
In the case of a radial bearing, the force between the rolling element
and the outer ring is increased by the given amount, while the force
between the inner ring and the rolling element remains at the magnitude
determined by the bearing load. The centrifugal force is usually relatively small and therefore does not have the effect of moving the most

172

Rolling Contacts

highly stressed point in the bearing from the inner ring to the outer ring
in bearings in which the inner ring is the weakest element. Thus, in most
types of radial bearing, the centrifugal force does not contribute to the
reduction in the load-carrying capacity of the bearing. In bearings in
which the outer ring is the weakest element, the overall loading is
slightly increased by the centrifugal force. This increase, however, is
usually so insignicant that it can be safely neglected. Only for very
high speeds can the centrifugal force reach the magnitude of 10 percent
of the outside load on the bearing.
The centrifugal force in angular contact ball bearings causes a certain
angle to exist between the directions of load at the two contact points
of the ball with the rings. Usually, this does not have any effect of
practical importance on the magnitude of the forces, but the areas of
contact are displaced so that the microsliding within these areas
increases. This phenomenon is most pronounced in thrust ball bearings
where the balls at high speed must lean on the outer shoulders of the
ball grooves (Fig. 5.13). The areas of contact, therefore, assume a position that deviates more from the theoretical cone of rolling than in the
case when the centrifugal force is absent (Fig. 5.13). The balls also roll
on a larger pitch diameter and may exert a pressure on the cage, thus
facilitating the wear. A practical way to alleviate this effect is to make
pockets oblong in a radial direction and keep a suitable thrust load on
the bearing.
When the bearing centre makes a circular motion, as in the case of
an epicyclic gear, the rolling elements and cage are subjected to forces

Fig. 5.13

Rolling Contact Bearings

173

acting towards the centre of the circle. If the bearing has a cage, which
is usually the case, it is centred on one of the rings and presses against
this ring under pure sliding motion with a force equal to the centrifugal
force of the cage. The forces created during acceleration owing to the
eccentric motion and acting on the rolling elements are primarily evidenced as normal forces against the rings and as tangential friction
forces on the rolling contact areas. The load on the rolling elements,
which varies from one rolling element to another, is sufcient to develop
a tangential friction force of the required magnitude in the contact areas
for pressing the rolling elements against the side surface of the cage
pocket under the condition of pure rolling.
It is quite often advantageous to make bearings subjected to a crank
motion cageless. The centrifugal forces then cause the rolling elements
to press against each other within those parts of the circumference that
are not located in the loaded zone. However, quite difcult lubrication
problems are often encountered in such applications.

5.6.2 Crankpin bearings


The kinematics of a crankpin bearing is quite complex, as it moves in
several different ways. The inner ring rotates with the same speed of
rotation as the crankshaft. Also, its centre moves with the crankpin,
while the outer ring makes an oscillatory motion owing to the limited
length of the connecting rod. As a result of this, the motions of the
rolling elements are complicated, producing inertia forces within the
bearing that may be of far greater importance for the proper functioning of the bearing than the external load. Especially important is
the oscillatory motion of the outer raceway, which causes the speed of
the rolling elements around the bearing centre to vary considerably.
When the length of the connecting rod is 3 times the crack radius, the
relative speed of rotation between the inner and outer raceway varies
between 0.67 and 1.33 times the speed of rotation of the crankshaft.
For example, at a crankshaft speed of 3000 rmin the bearing speed
changes from 2000 to 4000 rmin and back during each revolution, that
is, 50 times per second. For the rolling elements to maintain only rolling
motion at their points of contact with the raceways during these rapid
changes in speed, high tangential forces must exist in the area of contact; the heavier the rolling elements, the higher the forces must be.
However, tangential forces can develop at the points of contact with
the raceways only where the rolling elements are loaded, and these in
turn must exert sufcient pressure on the unloaded elements so that the
whole set takes part in the motion.

174

Rolling Contacts

Practice shows that cylindrical rollers are better for high-speed crankpin bearings than for ball bearings. This is due to the more favourable
ratio of contact area to weight of the rolling element. Also, it is known
that rollers of small diameter are preferred to rollers of large diameter.

5.6.3 Forces of gyration


Any rotating body whose axis of rotation changes direction is subjected
to gyratory forces. It is obvious that in ball and roller bearings the
rolling elements change the direction of their axes of rotation if the
contact angle H0.
A ball subjected to a gyratory moment has an axis of rotation that
coincides with the tangent of the pitch circle. Thus
Mg GJ o m sin

(5.42)

where J is the mass moment of inertia of the ball, o is the angular


velocity of the ball around its own axis, and m is the angular velocity
of the ball centre around the bearing axis.
According to equation (5.42), the gyratory moment increases with
the contact angle . It is highest in pure thrust bearings where G90.
When the shaft ring rotates at n rmin, the housing ring is stationary,
the bearing pitch diameter is d m , and the ball diameter is Dw , then the
angular velocity of the ball around its own axis of rotation is

oG

no G

30

dm
n
60 Dw

(5.43)

and the angular velocity of the ball around the bearing axis is

mG

30

nm G

60

(5.44)

The spinning of the ball in the direction of the gyratory moment is


resisted by the friction moment caused by the ball load. Where sliding
and rolling occur simultaneously, the coefcient of sliding friction is
0.070.08 for low rolling speeds. In the case of high speeds, however, it
may assume considerably lower values owing to the formation of a
lubricant lm at the area of contact. An approximate formula for the
friction moment in high-speed bearings is
Mf G20

Dw Fa
z

(5.45)

Rolling Contact Bearings

175

where Fa is the thrust load on the bearing and z denotes the number of
balls in the bearing.
The condition for sliding not occurring is that MfHMg , which means
that
FaH5.75B1011zd mD 3w n2

(5.46)

Practical observations show that sliding between surfaces owing to spinning of the ball in the direction of the gyratory moment does not have
an adverse effect on the bearing provided the specic load on the ball
is less than 80 kNm2 and the lubrication is good.
All radial bearings with a contact angle H0 generate gyratory forces. However, they are small in these bearings since sin is usually
small and sliding can develop only when the ball load is low, so the
bearing surfaces cannot be seriously damaged. In roller bearings with
H0, the gyratory moments have a potential to cause a slight change
in the load distribution in areas of contact, but this is of no practical
importance.

5.7 Load-carrying capacity


5.7.1 Dynamic capacity
A rolling contact bearing can operate for a limited period of time only.
If a ball or roller bearing is exposed to moisture or dirty surroundings,
it may become unserviceable because of corrosion and wear after a
period in service that cannot be predicted. On the other hand, if it is
properly protected and well lubricated, all potential causes of damage
are eliminated except one, i.e. the fatigue of the material owing to cyclic
stresses produced by rotation. The manifestation of this fatigue process
is aking, which starts as a crack and develops into a spalled area on
one or other of the load-bearing surfaces. Thus, ultimately, fatigue is
unavoidable but the number of revolutions the bearing may make
before aking starts is a function of the bearing load.
Concept of service life
The concept of service life can be dened as a period of service that is
limited by fatigue phenomena. The service life is measured in the number of revolutions of the bearing or in the number of hours of operation
at a certain speed of rotation. Individual bearings that are identical and
operate under identical conditions may, however, have different service

176

Rolling Contacts

Fig. 5.14

lives. The scatter of service life values is illustrated in Fig. 5.14. With a
sufciently large number of bearings tested, a minimum of 30, it is
found that the longest life rarely exceeds 4 times the average life Lm ,
and that about 90 percent of the bearings have a longer life than onefth of the average.
On the basis of the relationship between life and bearing load, as well
as between bearing load and material stress, it is possible to calculate
that the material stress that 90 percent of the bearings can endure for a
given length of time is only 16.5 percent smaller than that stress which
can be endured for the same length of time by those bearings which
have a life in excess of the average. The difference in material strength
between individual bearings is therefore not as signicant as might be
thought when considering only the variation in service life.
The scatter of results makes it necessary to formulate a precise denition of the concept of service life, which is taken to be the period of
useful operation which can be anticipated with reasonable probability.
Thus, the estimated service life means the number of bearing revolutions, or the number of working hours at a certain rotational speed,

Rolling Contact Bearings

177

that will be reached or exceeded by 90 percent of all bearings. This


estimated service life is denoted by L in Fig. 5.14 and serves as a basis
for bearing selections. The average service life Lm is 5 times as long
as L.
Analyses concerning the load capacity, equivalent load, and the like
are carried out on the assumption that the most highly stressed individual point in the bearing material determines the capacity. The rst
fatigue cracks, however, do not always develop exactly at this point.
Signs of fatigue could be observed rst at certain other points, and their
location depends on variations in strength inherent in the material. For
that reason, the scatter of service life values is always observed. It is of
practical importance to be able to estimate the probability of fatigue,
permanent deformation, or fracture developing in various parts of a
bearing.
Link between load and service life
In an imaginary test, if one group of bearings of a certain type and size
is tested under the constant load F1 and another group under the load
F2, but otherwise under identical operating conditions, the service life
L1 and L2 respectively will be obtained for these two groups. It is found
that a lower load always results in a longer service life. On the basis
of a sufciently large number of test results, the following has been
established

F1
G
F2

L2

L1

or
3
3
F11
L1 GF2 1
L2 Gconst

(5.47)

Assuming that F is a constant pure radial load acting on a radial bearing whose inner ring rotates relative to the line of action of the load,
and this bearing has a life of LN million revolutions, the constant in
equation (5.47) is that load with which the bearing, under the given
operating conditions, can attain one million revolutions. This load is
called the dynamic specic load capacity or specic capacity C. In the
case of thrust bearings, the specic capacity is expressed in terms of a
pure thrust load. Because of their design, single-row angular contact
bearings ( H0) must always be subjected to a certain thrust load
which is obtained by opposing the bearing to another bearing.

178

Rolling Contacts

Another well-known form of equation (5.47) is

LN G

(5.48)

in which C is the specic capacity of the bearing, that is, the loadcarrying capacity for a service life of one million revolutions, P is the
bearing load, and LN is the service life, in millions of revolutions, under
the bearing load P. If the specic capacity of a bearing is known, then
the permissible load for any service life can be calculated from equation
(5.48) or the service life can be found for any known load.
Estimation of specic dynamic capacity
A number of factors inuence the specic capacity of a bearing and,
thus, its permissible load for a given life. Among the most important
factors are:
the
the
the
the

properties of the material,


degree of oscillation between rolling elements and raceways,
dimensions of the rolling elements,
number of rolling elements.

As theoretical estimation of the specic capacity is impractical, the only


way to determine it is through time-consuming tests during which the
inuence of the different factors can be established. It has been found
that the load-carrying capacity varies with the ball diameter Dw
approximately as
D 2w
1C0.02Dw
It is known from the Hertz theory that the load-carrying capacity
should be proportional to D2w . Practical experience shows that small
balls have a relatively higher load-carrying capacity than large balls.
This fact is accounted for by the correction factor 1(1C0.03Dw ). In
roller bearings with short rollers of diameter Dw and length lw , the loadcarrying capacity varies with Dw lw . For rollers with greater lengths
(lwH1.4Dw ) it is practically impossible to secure a uniform load distribution over the entire roller length owing to unavoidable skewing of
the rollers, misalignment, and other conditions. Therefore, the loadcarrying capacity does not increase in proportion to the roller length.
When the bearing has rolling elements made of a hard material, the
fatigue phenomenon that determines the life usually develops on the
raceway of one ring or the other. It is therefore obvious that the rolling

Rolling Contact Bearings

179

elements are not the weakest parts of the bearing, but their number
inuences the stress in the rings, partly because of the magnitude of
load per roller and partly owing to the number of stress cycles per
bearing revolution at the weakest point of the raceway. The magnitude
of the force with which each rolling element acts on the weakest point
of the raceway is inversely proportional to the number of rolling
elements. Thus, for a constant number of stress cycles at the point of
interest, the carrying capacity is proportional to the number of rolling
elements. On the other hand, the number of stress cycles for one revolution of a bearing ring is proportional to the number of rolling elements
z. Therefore, the same functional relationship between capacity and
number of stress cycles exists as that between capacity and service life
given by equation (5.48). In other words, for a constant ball or roller
load the capacity is inversely proportional to the cube root of the
number of rolling elements. Finally, the bearing capacity varies with
z
Gz2/3
3
1
z
provided all other factors remain unchanged.
The force of the rolling element acting against the raceway varies
with 1cos , so that the capacity is proportional to cos . Finally, if
the bearing has i rows of rolling elements, the capacity is i times higher
than that of a single-row bearing, provided the load is distributed evenly
among the several rows of rolling elements. This is, however, true only
for double-row self-aligning bearings. For all other bearings, with two
or more rows of rolling elements, it is necessary to allow for some
reduction in bearing capacity owing to non-uniform load distribution.
In summary, the specic load capacity of a ball bearing is
CGfc

iD 2w z2/3 cos
1G0.02Dw

(5.49)

and that of a roller bearing is


CGfc iDw lw z2/3 cos

(5.50)

where z is the number of rolling elements per row.


A special explanation is necessary for coefcient fc . It depends,
among other things, on the properties of the material, the degree of
oscillation, and the reduction in the capacity on account of uneven load
distribution within multiple row bearings and bearings with long rollers.
The value of this coefcient can only be determined by comprehensive

180

Rolling Contacts

laboratory tests. It usually has one value for all sizes of a given bearing
type. The values of coefcient fc are available in proprietary catalogues
published by all leading manufacturers of rolling contact bearings.
As far as the load-carrying capacity is concerned, there is no fundamental difference between radial bearings and thrust bearings. The term
thrust bearing is used for those bearings whose contact angle is large,
usually 4590, and whose rings take the form of washers. Thrust bearings are suitable for supporting loads that are primarily or exclusively
thrust loads, and their specic capacity is therefore given in terms of
pure thrust-carrying capacity. Thus, the loads on the individual rolling
elements vary with 1sin and the capacity with sin . Other factors
have the same effect as in radial bearings. The expression for calculating
the specic capacity for single-row thrust ball bearings with 90 is
CGfc

D 2w z2/3
1C0.02Dw

(5.51)

and for single-row thrust roller bearings


CGfc Dw lw z2/3 sin

(5.52)

Effect of the speed of rotation on the load-carrying capacity


With the help of equation (5.48) it is possible to nd the required bearing size if the constant bearing load P and the desired life are both
known. This can be carried out for service life expressed in number of
revolutions or in number of hours at a certain constant speed of
rotation. If Lh is the service life in hours and n is the speed of rotation,
then
LN G60B106nLh

(5.53)

The required specic capacity can be calculated from the equation


describing service life. The bearing manufacturers provide tables showing the so-called load rating, i.e. the load-carrying capacity at different
speeds of rotation based on a certain constant service life expressed in
hours of operation, usually 500 h. Thus, for service life, equation (5.53)
yields
LN G0.03n
According to equation (5.48), the load rating Cn at the speed of rotation
n is
C
Cn G 3
Gfn C
10.03n

(5.54)

Rolling Contact Bearings

181

Coefcient fn is a function of rotational speed and can be easily found


in catalogues published by bearing manufacturers. The specic capacity
is equal to the load rating at 33.3 rmin, as is apparent from equation
(5.54) when Cn GC.
Service life factors
When all forces acting on a bearing are given and the equivalent load
P is estimated on the basis of these forces, then the required specic
capacity C can be calculated from the equation
CGfN P

(5.55)

where fN is a factor that has a given relationship to the service life in


million revolutions. From equation (5.48) the following is obtained
3
fN G1
LN

(5.56)

This service life factor can be used in cases where the service life cannot
be easily expressed in hours as well as in the absolute number of revolutions. Examples of such instances are: automobiles, railroad boxes,
and other vehicles where the service life is given in number of miles of
operation.
In the majority of applications, rolling contact bearings work with
nearly constant speeds of rotation. Therefore, it is convenient to estimate the service life in number of operating hours. If the speed
of rotation is constant, it is practical to start from the load rating Cn
applicable at this speed. Therefore
Cn Gfh P

(5.57)

where fh is the service life factor for the life given in hours. According
to equation (5.47)
3
3
Cn 1
500GP 1
Lh

(5.58)

and therefore

fh G

Lh
500

(5.59)

5.7.2 Static capacity


Quite often a bearing is under external load while it is not rotating. The
service life equation (5.48) cannot be used in such a case because it gives
PGS for LN G0. Understandably, there is a limit to the load that the
bearing can carry under static conditions. However, this limit is not

182

Rolling Contacts

determined by the fatigue strength of the material but by the permanent


deformations that are created in the load-supporting surfaces.
It was mentioned earlier that permanent deformations could be produced even under relatively light loads and increase in size with increasing load. Consequently, there is no load limit below which the
deformations are entirely elastic or at which plastic deformation suddenly begins to develop. It is practically impossible to avoid some permanent deformations, and the only question is what size of deformation
can be safely tolerated if the bearing is to operate properly. Again,
application experience demonstrates that permanent deformations have
a negligible inuence on the operation of the bearing provided that their
combined magnitude at any one contact point is less than 0.0001 times
the diameter of the rolling element. When the load is very high, permanent dents are formed in the raceways and consequently the bearing
begins to run noisily and vibrate. Slight vibration and noise caused by
the bearing when running at high speed have an important effect on
certain applications. However, the static capacity implies the load to
which a bearing can be subjected while stationary without deformations
being produced, which would be noticeable when the bearing subsequently rotates under a lower load and normal requirements for
smooth running. The pure radial load or pure thrust load that corresponds to this static load-carrying capacity of the bearing is called the
specic static capacity and is denoted by Co .
The above denition of static capacity implies that the bearing can
safely be allowed to rotate under a load that is higher than the specic
static capacity. If the maximum load is of limited duration and acts
only while the bearing rotates, the permanent deformations that take
place will be evenly distributed around the entire periphery of the raceways and cause no practical harm until the deformations become relatively large. The second implication of the denition of the specic static
capacity is that the load may exceed the stated limit, in certain cases
quite considerably, even while the bearing is stationary, provided that
in subsequent running the need for smooth operation is not too important and the main requirement is that the bearing can still rotate.
Magnitude of static capacity
The maximum ball or roller load, Wmax, is used to determine the specic
static capacity, Co . The magnitude of Wmax depends on the degree of
oscillation in the weaker of the two ball or roller contacts as well as on
the material characteristics. By setting a limit to the permissible permanent deformation of rolling element and bearing ring at a contact as

Rolling Contact Bearings

183

0.0001 times the diameter of the rolling element, the allowable static
specic load po can be calculated for a given bearing from equations
(5.27) and (5.28) respectively by substituting p Dw G0.0001. Therefore, for ball bearings
po G

2.8B107

(5.60)

Dw 1[(r1ICr1II )(r2ICr2II )]

and for roller bearings where the line contact at the inner ring is the
location of interest
po G

17B107

(5.61)

1[Dw(rICrII )]

In the above equations, the allowable static specic load, po , is given in


Nm2.
Bearings of standard design have the relationship between diameter
of rolling element and the radii of curvature of the ring surface xed
for different sized bearings of a given type. Thus, it is convenient to
have average values for po for some bearing types prescribed in order
to facilitate approximate analyses. Table 5.1 gives values of po .
Since for ball bearings
po G

Wmax
D 2w

and
po G

Wmax
Dw lw

for roller bearings, the specic static capacity, Co , can be determined


from the relations between bearing load and maximum ball or roller
load which were discussed earlier. For radial ball bearings, the following
Table 5.1
Bearing type

po (B107 Nm2)

Ordinary ball bearings


Self-aligning ball bearings
Thrust ball bearings
Deep groove ball bearings
Roller bearings

1.5
1.7
5.0
6.2
11.0

184

Rolling Contacts

is obtained
CoG15 Wmaxiz cos G 15 po iz cos D2w

(5.62)

and for radial roller bearings


Co G15 Wmax iz cos G15 po iz cos Dw lw

(5.63)

The axial specic static capacity for thrust ball bearings is


Co GWmax iz sin Gpo iz sin D 2w

(5.64)

and for thrust roller bearings


Co GWmax iz sin Gpo iz sin Dw lw

(5.65)

Table 5.1 contains values of po that are applicable to bearings so


designed and made that the load distribution within the bearings can
be regarded as conforming to that calculated. In bearings where this is
not the case, for example, in double-row radial contact ball bearings,
roller bearings with long rollers, and the like, the static capacity is
slightly reduced.

5.7.3 Equivalent bearing loads


Equation (5.48) has to be used in all calculations of the service life of a
bearing under given operating conditions. In this equation, LN is the
service life in millions of revolutions, C is the specic bearing capacity,
and P is the bearing load acting under operating conditions applicable
to that specic capacity. It is assumed, therefore, that for a radial contact bearing the load P is constant and purely radial, and that the inner
ring rotates in relation to the direction of the load P, while the outer
ring is stationary in relation to the line of action of P.
However, there are many practical applications in which the load is
not constant and does not have a purely radial direction. Also, in these
applications, the outer ring or both rings rotate in relation to the direction of load. In all these cases, the equation for the service life [equation
(5.48)] can only be used if it is possible to calculate that constant pure
radial load which, if only the inner ring were to rotate in relation to the
direction of load, would have the same effect on the bearing as the
actual load. In other words, it would give the same service life as that
which the bearing will attain under the actual operating conditions.
This imaginary load is called the equivalent load.
The thrust bearing case requires calculation of the equivalent thrust
load, because the specic capacity for this type of bearing refers to the
capacity under constant centric pure thrust load. Calculation of the

Rolling Contact Bearings

185

capacity of a bearing that does not rotate under load also requires the
use of an equivalent load, although this is related not to the fatigue
process but to slight permanent deformations that limit the service
potential of the bearing.
It is known that the service life of a bearing, expressed in number of
revolutions, is practically independent of the speed of rotation. Therefore, it is possible in the case of variable speeds to use the average speed,
or to transform the product of time and number of revolutions per unit
time during the different periods of operation into the total number of
revolutions.
Case of rotating bearings
According to equation (5.48), the service life is inversely proportional
to the third power of the load. Thus, a high load can exert a considerable inuence on the life, even though it acts only during a relatively
small part of the total service life. It is reasonable to assume that, if a
load acts for a certain fraction of that number of revolutions during
which the bearing can operate under this load before failing because of
fatigue, the same fraction of the ability of the bearing to endure load
under rotation is used up. Subsequently, if a load of another magnitude
is applied, the bearing thus loaded can function for that number of
revolutions for which this load alone would permit the bearing to operate before failing because of fatigue, minus the used up fraction of this
last mentioned number of revolutions.
Therefore, the relationship between the service life L1 and the radial
load F1 is

L1 G

If a1 denotes the fraction of the service life during which the load F1
acts, then a1L1 is the number of millions of revolutions during which
the load F1 acts. This number is designated N1. Thus
a1L1 GN1 Ga1

F
C

or
F 31N1 Ga1 C 3

186

Rolling Contacts

Using similar designations for other loads gives


F 32 N1 Ga2 C 3
F 3n Nn Gan C 3
Adding the above equations
3
3
(F N)G (a)C

According to the assumed principle for the consumption of the loadcarrying ability, it is apparent that (a)G1. Therefore
3
3
(F N)GC

or
LN G

C 3LN
3
(F N)

By denoting
Fm G

1
3

3
(F N )

LN

(5.66)

it leads to
3

LN G

Equation (5.66) shows that Fm represents the constant cubic mean


load the mean effective load which gives the same service life as the
actual variable load acting on the bearing.
Quite often it is possible to divide the bearing load history into a few
periods with constant load within each period as shown in Fig. 5.15.
By doing so it is easy to calculate Fm from equation (5.66). If a more
accurate estimate of the load is required or the load history diagram
has a complicated form, then it is necessary to divide the load diagram
into an innite number of small parts, i.e. replace (F 3N) by the
integral

LN

F 3 dN

Rolling Contact Bearings

187

Fig. 5.15

As a result, the following is obtained


Fm G

1
3

L
0 F 3 dN
N

LN

(5.67)

If the load varies periodically, the cubic mean load for all periods is the
same as the cubic mean load for one period. It is therefore possible in
equation (5.67) to let LN represent the number of revolutions in one
period.
Figure 5.16 illustrates loading conditions in which F varies from a
minimum Fmin to maximum Fmax . In such a case, the following approximate expression can be used
Fm G13 FminC23 Fmax

(5.68)

All the equations for the cubic mean load apply to cases where the
direction of the load is constant, that is, Gconst. The cubic mean load
can, however, be introduced in the service life calculations only after
having been converted to pure radial load, acting on the bearing as a

188

Rolling Contacts

Fig. 5.16

rotating inner ring load in the case of radial bearings, or pure centric
thrust load in the case of thrust bearings.
It was concluded earlier that the maximum load on the ball or roller,
as well as the length of the loaded zone and the load distribution within
that zone, changes when the direction of the resultant load changes.
For example, in a bearing whose weakest point is the inner ring ball or
roller contact track and whose inner ring rotates relative to the direction
of load, the pressure is exerted at a given point on the inner ring ball
or roller contact path every time this point passes a rolling element
located within the loaded zone. The magnitude of this pressure is dependent on the angular displacement of the rolling element when it passes
the point of interest, this displacement being measured from the action
line of the bearing load (see Fig. 5.9). The relation between the magnitude of the force and its angular position under pure radial load is
shown in Fig. 5.8. The potential fatigue effect that these cyclic loads
have on the point under consideration is proportional to the number of
times the point is loaded and to the cubic mean value of the different
forces that act during each respective instant of load.
Suppose that the bearing load acts at an angle to the radial plane
of the bearing (see Fig. 5.10). As a result of that, the loaded zone is
longer and the load variation, from rolling element to rolling element
within this zone, is not the same as that previously considered. Consequently, the fatigue effects are also different. In the meantime, other
factors are introduced, such as the tangential stresses due to the tendency of various rolling elements in the bearing to have different peripheral speeds, and the increased probability of the occurrence of fatigue

Rolling Contact Bearings

189

owing to a greater number of points being stressed, and these all have
the potential to inuence the fatigue strength. Therefore, the load-carrying capacity of the bearing varies with the direction of load. This is not
only because of change in the maximum ball or roller load but also for
other reasons. The analysis is quite complicated and will be illustrated
by an example of double-row and single-row angular contact bearings,
both with a contact angle such that tan G0.25 and in both of which
the inner ring is the weakest element.
A possible starting point for the analysis is that capacity which a
bearing would have with various directions of load for a given life,
provided that the maximum ball or roller load alone determined the
capacity. Therefore, this bearing load is chosen as a unit load that
theoretically could be applied to the bearing to give the same maximum
ball or roller load with all directions of load. Figure 5.17 shows the
conditions in the double-row bearing, where the curve i applies to the
rotating inner ring load. Usually, the permissible load (i.e. the loadcarrying capacity of the bearing) is lower than the unit load mentioned
above. Thus, in the case of a pure radial load, G0, the load carrying
capacity is only 0.875 of the theoretically calculated load, because the
actual load distribution of the rolling elements is less favourable than
the theoretical distribution owing to clearance within the bearing and
between the outer ring and housing.

Fig. 5.17

190

Rolling Contacts

The increased number of stress cycles per revolution at the weakest


point of the bearing must be taken into account when a thrust load is
applied, that is, H0. As the angle increases, the capacity gradually
becomes a smaller fraction of that capacity which was determined by
variations in the maximum ball or roller load only. When tan G
1.7 tan , that is, when G23, one row of rolling elements is loaded
around the entire circumference and the other row is entirely unloaded.
In order to obtain a certain life under a pure thrust load, G90, only
half of the maximum ball or roller load can be allowed as that which
theoretically would be permissible in the case of pure radial load.
An analogous condition for a single-row bearing with the same contact angle is shown in Fig. 5.18. As argued earlier, can never be
smaller than . Only when tan G1.25 tan will the bearing begin to
operate normally and in the same way as each row in the double-row
bearing when G0. Thus, when G17.5, the carrying capacity is 87.5
percent of the theoretical. When tan G1.7 tan , that is, when G23,
the row of balls is loaded around the circumference and then functions
in exactly the same way as the row in the double-row bearing that
carries the load. Therefore, the curves in Figs 5.17 and 5.18 coincide at
H 23.
When the inner ring rotates relative to the direction of the bearing
load, a given point in the ball or roller contact path with the inner
ring passes the different rolling elements. Within the unloaded zone, no

Fig. 5.18

Rolling Contact Bearings

191

pressure is exerted on the point in question. Within the loaded zone, a


lightly loaded rolling element passes rst, and thereafter other rolling
elements that are more loaded pass successively until a maximum pressure is obtained when the point passes a rolling element that is located
almost on the line of action of the bearing load. After this, less loaded
rolling elements pass again.
In a running bearing with the inner ring stationary in relation to the
direction of load, the case of a xed inner ring load, the point of the
inner ring located on the line of action of the bearing load is always
subjected to the maximum ball or roller load every time a rolling
element passes this point. Thus, the load-carrying capacity of the bearing for a given service life is reduced. In the case of a pure radial load,
it is usually about 70 percent of the capacity for a rotating inner ring
load. As the bearing load gradually changes to a more axial direction,
the load-carrying capacity changes according to the curves denoted by
e in Figs 5.17 and 5.18. When there is a pure thrust load, it is unimportant which ring rotates so that curves i and e coincide at G90.
In order to examine how the load-carrying capacity changes with the
direction of load, it is instructive to consider as a unit the practical
load-carrying capacity F1 of a single-row bearing with G0. Knowing
the distribution of load over the rolling elements, it is possible to calculate theoretically the load-carrying capacity that the bearing should
have for different directions of load, if the maximum ball or roller load
Wmax were a limiting load. In the single-row bearing with G0 and z
denoting the number of rolling elements, the maximum load is
Wmax G

5F1
z

In a double-row bearing under radial load F2 , the theoretical maximum


load is
Wmax G

4.37F2
2z cos

and therefore
F2 G2.29F1 cos
if Wmax is to be of the same magnitude in both cases.

192

Rolling Contacts

In theory, the double-row bearing should have 2.29 cos times as


high a capacity as the single-row bearing has in practical application.
In the case of a pure thrust load F2 for the double-row bearing
Wmax G

F2
z sin

because only one row of rolling elements is loaded.


Therefore, in this case the following is obtained
5F1
F2
G
z sin
z
or
F2 G5F1 sin
In general, the following is applicable
PGXFrCYFa

(5.69)

where P is the equivalent load, Fr is the radial component of the actual


load, Fa is the axial component of the actual load, X is the bearing
rotation factor, and Y is the bearing thrust factor. The rotation factor
X is an expression for the effect on the bearing capacity of the conditions of rotation. The thrust factor Y is a conversion value for thrust
loads.
It has been found that equation (5.69) is applicable to most bearing
types provided the load components Fr and Fa are constant during the
entire bearing life. The values for X and Y are different for different
bearing types and are available in catalogues published by rolling
contact bearing manufacturers.
Self-aligning ball bearings have the feature that the weakest point is
located on the contact path of the outer ring. The factor Y for these
bearings is smaller than that applicable to other double-row bearings.
The number of stress cycles per revolution of the inner ring, produced
by a thrust load, is the same at the most heavily loaded point on the
outer ring contact path as under a radial load. When the inner ring
load is xed, the rings have comparable strength and, therefore, the
factor X does not change its value in this case.
The relationship between the load-carrying capacity and the direction
of load for single-row bearings with a constant contact angle is exactly
the same as in double-row bearings provided that tan H 1.7 tan .

Rolling Contact Bearings

193

The specic capacity and equivalent load P is only half of that for
double-row bearings. Thus
PG0.5FrC0.4 cot Fa
and for a xed inner ring load
PG0.7FrC0.4 cot Fa
It can be seen that the factors X and Y have only half the magnitude
for a single-row bearing that they have for the corresponding double
row. However, equation (5.69) is not entirely applicable to single-row
bearings. In the case of a rotating inner ring load it must be remembered
when analysing a single-row angular contact ball bearing that the thrust
load should not be less than Fa G1.25 tan Fr. For an even smaller relative thrust load, the load is concentrated on a reduced number of rolling
elements, and when Fa G1.25 tan Fr , only a single rolling element is
loaded. Thus, the load-carrying capacity is reduced to a very small value
for this direction of load.
It was shown earlier that deep groove ball bearings do not have a
constant contact angle. It is known, however, that, with the conventional radii of curvature of the ball grooves, which is approximately 4
percent larger than the ball radius, it is possible to use the factor YG
1.52, the lower value applying to relatively heavy loads. Equation
(5.69) agrees well with test results under pure or predominantly radial
loads. The pure thrust capacity of deep groove ball bearings can be
raised by an increase in the radial internal bearing clearance to a maximum value of 0.005Dw. In this way, the thrust capacity is increased by
a maximum of 25 percent, which corresponds to a doubling of the
service life.
There are many operating conditions under which the bearing loads
are variable in magnitude as well as direction. If a radial load component continuously acts in the same axial plane in relation to the inner
ring or outer ring but the load resultant is variable and changes direction in the plane of interest, then the following expression for the equivalent load is obtained as a result of insertion of the expression for P
given by equation (5.69) instead of F resulting from equation (5.67)
L
0 (XFrCYFa )3 dN

PG

LN

(5.70)

194

Rolling Contacts

When the load changes in such a way that equation (5.68) can be
applied, then the following approximation results
PG13 (XFrCYFa )minC23 (XFrCYFa )max

(5.71)

If the load is not only variable but also has a radial component that
changes direction in an irregular way, or if the load is periodic with the
same frequency as the speed of rotation of the bearing, then a more
sophisticated and advanced analysis is required.
It is a well-known fact that pure thrust bearings, G90, never carry
combinations of radial and thrust loads. In such a case, both bearing
rings are usually equally strained and the equivalent thrust load may
be calculated according to equations (5.67) and (5.68) if the load on the
bearing is a variable centric axial load. With an eccentric thrust load,
the equivalent load is increased in proportion to the maximum ball or
roller load owing to the unchanged number of stress cycles per
revolution.
Nowadays, a special type of bearing is available which can be considered as a thrust bearing for which the specic capacity is given as
pure thrust capacity even though the bearing can also support radial
loads. One such type is the spherical roller thrust bearing which, as a
matter of fact, may also be considered as a radial bearing with a very
steep contact angle. The ratio of load-carrying capacity in a pure radial
direction to that in a pure axial direction, as well as the ratio of the
equivalent radial load P to the equivalent thrust load PA, is
P

GY

PA
Substituting P obtained from equation (5.69), the equivalent thrust load
is given by
X
PA GFaC Fr
Y
or by denoting
X
Y

GYa

the result is
PA GFaCYa Fr

(5.72)

Rolling Contact Bearings

195

Case of non-rotating bearings


When a load of such magnitude as to damage a bearing is applied to a
stationary bearing, then it is not subjected to fatigue comparable with
that prevailing if the bearing rotates. The possibility of permanent
deformation development in the material, which would affect the serviceability of the bearing, is quite real. In a case like that, the absolute
maximum load must be taken into account, and the equivalent load
must be calculated on the basis that it would cause the same deformation or give the same risk of fracture as the actual load combination.
Thus, when calculating this static equivalent load Po for radial bearings,
the following expression is recommended
Po GXo FroCYFao

(5.73)

where Fro is the radial component of the maximum static load, Fao is
the axial component of the maximum static load, and Xo and Yo are
factors available from the bearing manufacturer catalogue.
The eccentricity of the load in pure thrust bearings is the only factor
that affects the equivalent thrust load. Thus, the static equivalent thrust
load bears the same relation to the actual eccentric load as the maximum ball or roller load under eccentric load bears to the maximum ball
or roller load under centric load. For a thrust bearing that can carry a
radial load, the ratio of the static radial capacity to the static thrust
capacity is 0.2 cot . As the static equivalent radial load of the bearing
is
Po G0.5FroC0.2 cot Fao
the static equivalent thrust load is therefore
PAo GFaoC2.5 tan Fro

(5.74)

5.8 Lubrication of bearings


As stated in Chapter 1, there are three basic functions performed by
the lubricant, i.e. to lubricate, to remove heat, and to protect against
the inuence of environmental elements. It was also said that some
sliding contact occurs between the rolling element and the running track
in commercial bearings and, in some cases where diametral clearance
and relaxation of load encourage it, a hydrodynamic lubrication regime
is established. However, in normal loaded conditions the preponderant
function of the lubricant in this area, according to the elastohydrodynamic theory, is the generation of a lm between rolling elements
and track as a result of deformation.

196

Rolling Contacts

5.8.1 Elastohydrodynamic lubrication


The shape of the lm of lubricant as predicted by Grubin and Vinogradova (5) and conrmed by Crook (6) is shown schematically in Fig. 5.19
with the theoretical pressure distribution. In practice, as the pressure in
this region is very high, possibly 1.53.0 GPa, the viscosity of this lm
is greatly increased. Therefore, to the extent that such a regime prevails
in the contact area, it would seem likely that this lm, momentarily of
such high viscosity as to be comparable to dry static friction, transmits
the elements of the couple which produces rolling. At the entry to the
region the lubricant is under shear and may be regarded as behaving
hydrodynamically as suggested by Archbutt and Deeley (7), but within
the zone, presumably, negligible slip occurs.
Clearly, the pressureviscosity characteristics of lubricants are of
importance in the behaviour of such elastohydrodynamic lms. Experimental studies into elastohydrodynamic lubrication conrmed a close
agreement between lm thickness measurements and values predicted
by the theory. Where anomalies were found, notably in the cases of a

Fig. 5.19

Rolling Contact Bearings

197

silicone uid and a solution of polymethyl methacrylate in oil, the possible effect of non-Newtonian behaviour was suspected.
Tallian et al. (8) made a study of conditions in rolling contacts where
load is shared between an elastohydrodynamic lm and surface asperities. They called this partial elastohydrodynamic lubrication and considered that it could exist for a given system in speed conditions below
those at which a full elastohydrodynamic lm would be established.
The critical nature of the conditions that determine whether a fully
developed elastohydrodynamic lm can be established in relation to the
surface nish can be deduced from the following dimensions proposed
by Dowson (9).
The typical length of the Hertzian zone of contact (representing the
region of effective pressure generation for elastohydrodynamic contact)
is approximately 0.25 mm. The typical lm thickness in the zone is
approximately 0.00025 mm. The typical transit time for lubricant
through the zone (according to rolling speed) is approximately 0.0001
0.00001 s. Finally, the typical contact pressure on the lubricant is
approximately 2 GPa.
These quantities must be related to a typical surface nish for rolling
contact bearing elements of about 12.5B1062B104 mm. It is now
commonly accepted that, when the lm thickness exceeds the combined
peak-to-valley heights of the rolling element surface and the raceway
by about 34 factors, then a complete uid lm lubrication is established. Where full hydrodynamic or elastohydrodynamic conditions can
be maintained, the full fatigue life expectancy of the bearing, insofar
as it is determined by lubrication, may be achieved. If boundary conditions prevail even for part of the time, the fatigue life may become
unpredictable.
When considering the lubrication in the rolling elementraceway contact, some reference must be made to the behaviour of a ball in conditions where it is forced to depart from relatively true rolling, as in an
angular contact bearing. A study by Hirano and Tanoue (10) using a
magnetized ball, which enabled tracking of its path, revealed that the
motion of the ball is three-dimensional, including slipping and spinning.
It was also found that slip of the ball occurs on the unloaded side, where
its contact with the raceways is looser. Slip decreases with increase in
speed and load. It increases with radial clearance. The instantaneous
rolling axis relative to the ball changes regularly and this is closely
related to the slip. The cause of the change in the rolling axis is spin
during slip. The angular displacement of the spin is proportional to the
mean slip. The spin increases with the asymmetry of the contact of the

198

Rolling Contacts

ball with the cage or with the grooves of the races. Further increase in
asymmetry causes negative slip. Under pure thrust load, no slip is
observed although the contact is still maintained.
Friction resulting from the spin and slip described may be high in
some cases and not only must it be included in the sum of lubricating
requirements for the bearing but it may make severe demands necessitating special qualities in the lubricant. In general, lubrication of the
rolling element sliding in the cage pocket presents no problems as the
load at the contact point is negligible. Normal hydrodynamic lubrication would be expected to be maintained. Sliding of the roller end
on the lip, particularly in taper roller bearings, can result in boundary
conditions being produced under high loads, and heavy wear may ensue
if special lubricants are not used.
If the cage is centred on the inner or outer ring, lubrication is
demanded between the bore or circumference of the cage, as the case
may be, and the race. Again, the load may be negligible except where
out-of-balance forces exist. Lubrication of this, in effect, plain bearing,
is normally hydrodynamic, but in the case of the cage centred on the
inner ring a speed limitation must be observed if the lubricant is a
grease. If the limit is exceeded, the centrifugal force created by slight
imbalance of the cage may cause the lubricant lm to break and wear
occurs, aggravating the tendency of the cage to precession.
Finally, lubrication of the rubbing seal in sealed bearings must not
be overlooked. Although strictly not a part of the bearing proper, like
the inner centred cage this should be regarded as a plain bearing hydrodynamically lubricated and liable to wear if lubrication fails.

5.9 References
(1) Harris, T. A. (1996) Rolling Bearing Analysis (John Wiley, New
York).
(2) Hertz, H. (1882) Uber die Beruhrung fester elastischer Korper (in
German). J. Reine und Angewandte Mathematik, 92, 156171.
(3) Pinegin, S. W. (1965) Contact Strength in Machines (in Russian)
(Mashinostroenie, Moscow).
(4) Stribeck, H. R. (1901) Ball bearings for various loads (in German).
Z. VDI, 45.
(5) Grubin, A. N. and Vinogradova, I. E. (1949) Book No. 30 (in
Russian) (Central Scientic Research Institute for Technology and
Mechanical Engineering, Moscow).

Rolling Contact Bearings

199

(6) Crook, A. W. (1961) Elastohydrodynamic lubrication of rollers.


Nature, 190.
(7) Archbutt, L. and Deeley, R. M. (1927) Lubrication and Lubricants
(Grifn, London).
(8) Tallian, T. E., McCool, J. I., and Sibley, L. B. (1965) Partial elastohydrodynamic lubrication in rolling contact. In Proceedings of
IMechE Symposium on Elastohydrodynamic Lubrication, London.
(9) Dowson, D. (1965) Elastohydrodynamic lubrication: An introduction and a review of theoretical studies. In Proceedings of IMechE
Symposium on Elastohydrodynamic Lubrication, London.
(10) Hirano, F. and Tanoue, H. (1961) Motion of a ball in a ball bearing. Wear, 4.

This page intentionally left blank

Chapter 6
Rolling Contacts in Land Locomotion

The powerweight ratio of railway locomotives has been increasing consistently over the past two decades mainly owing to enlarged power
plants, the extensive use of lightweight materials, and improved construction methods. The only limiting factor to a continued increase in
this ratio seems to be the traction developed between the driving wheels
and rails (1). This statement is supported by the pronounced tendency
for wheel slip to occur, especially in starting from standstill and under
wet conditions.
Another major area of land locomotion and transportation involves
the rolling performance of pneumatic tyres on road surfaces. The contact interface between tyre and road is determined by complex interaction events during free rolling, braking, driving, cornering, skidding, or
any combination of these modes, and its properties reect the result of
these interactions (2). Nevertheless, in practice it is useful to separate
the individual contributions of tyre and road in order to understand
the fundamental events that subsequently determine the frictional
coupling in the contact area.

6.1 Railwheel systems


In order to gain a clear understanding of the fundamental mechanism
governing the traction between wheel and rail, a memory effect also
known as secondary conditioning has to be recalled. This phenomenon
arises from the fact that, when certain substances, especially oils, are
spread upon a rail, secondary effects take place which are manifested

202

Rolling Contacts

by the creation of a minute quantity of the substance so closely associated with the surface as virtually to form a part of the main material.
Therefore, it is possible for two apparently identical steels to possess
widely different coefcients of surface friction despite all the efforts to
clean the track. This memory capability, which is generally detrimental
for surfaces previously contaminated with oil, can have benecial effects
when the spark discharge method is employed to improve traction.

6.1.1 Traction at the railwheel interface


Basically, there are only three methods for improving railwheel traction, namely:
employing additives on the rail surface (chemical),
scoring, abrading, and sanding of the railhead (mechanical),
plasma arc or spark discharge between an electrode and the rail
(electrical).
It appears that both the chemical and mechanical methods have either
failed to provide satisfactory improvements in traction or they tend to
introduce other unwanted effects. The list of undesirable effects given
below is by no means comprehensive but illustrates well the problems
involved:
(1) The use of a colloidal dispersion of silica in water produces a considerable improvement in the traction of dry or wet rails having
medium or low values of secondary conditioning. However, this
treatment shows little improvement on rails covered by an oil.
(2) Sodium hydroxide solutions are able to reduce the effects of oil
contamination by attacking traces of oil which give low secondary
conditioning, thus increasing adhesion. However, excess or longterm use of these compounds produces a sludge which may have an
adverse lubricating effect.
(3) Sanding of the rail surface provides a practical way of securing an
instantaneous increase in railwheel traction. However, the sand
must be absolutely dry, and there are practical difculties with
appropriate storage and particle size control. Besides, sanding is not
permitted near switch gears and switching points because of the
danger of clogging. Also, an increase in surface damage occurs
owing to pitting.
(4) Rails previously treated with silicone uids tend to give high traction values even when subsequently treated with oil. On the other
hand, initially clean rails give low traction when covered with oil.
However, when the silicone-treated rails are covered with water, the

Rolling Contacts in Land Locomotion

203

formation of innumerable droplets brings about a large reduction


in traction.
The method of spark discharge, consisting of ionizing the air gap
between an electrode placed ahead of each driving wheel and the rail,
effectively removes the contaminants that normally appear on the rail,
and thereby produces a marked improvement in traction under the most
adverse conditions.
Table 6.1 shows values of railwheel traction values under dry, wet,
or greasy conditions (3). It is apparent that a large variation in the
values of the traction coefcient exist, from a minimum value of 0.07
on damp rails to a maximum of 0.35 under dry, clean conditions. Lower
traction values usually point to the presence of an oil contamination on
the rail surface. Also, the slippery conditions are usually found on
curves, near points, near stations, and at road crossings. Oil contamination from axles and lubricating pads ows on to the wheel rims and
nds it way into the contact path.
The traction coefcient values listed in Table 6.1 are approximate
and most appropriate for low speeds. When speed is considered as a
variable, then it can be seen that the traction systematically decreases
with increasing locomotive speed. A number of empirical relationships
have been put forward to approximate the adhesion versus speed.
Usually, they take one of two forms
k1
fT Gk1C
VCk3

(6.1)

fT Gk4Ak5V n

(6.2)

Table 6.1 Examples of railwheel traction


coefcients
Condition of rail surface

Traction coefcient

Dry rail (clean)


Dry rail (with sand)
Wet rail (clean)
Wet rail (with sand)
Greasy rail
Moisture on rail
Sleet on rail
Sleet on rail (with sand)
Light snow on rail
Light snow on rail (with sand)
Wet leaves on rail

0.250.30
0.250.33
0.180.20
0.220.25
0.150.18
0.090.15
0.15
0.20
0.10
0.15
0.07

204

Rolling Contacts

where fT denotes the traction coefcient, ki represents positive constants


(iG15), and n has the value 1 or 2. From the results of many investigations, the following approximate and simplied equations can be used
for dry rail surfaces
(a) 0FVF60 kmh
fT G0.25
(b) 60FVF225 kmh
fT G

30
VC75

In the case of a wet rail surface, the following is recommended:


f T G0.6 fT .
In general the traction coefcient decreases with increasing brake
block pressure. This can be expressed by the following relationship
fT p0.38

(6.3)

where p is the pressure exerted by the brake block.


The inverse dependence of fT on p is in accordance with the simple
theory of adhesion for metals and is valid for pressures of up to about
2 MPa. It can be shown that at greater contact pressures the traction
coefcient, which is directly related to the adhesion at the interface,
begins to increase with increasing p and, at the same time, becomes
independent of speed. The reason for this is probably the combination
of high pressure and speed causing overheating of the brake block,
softening it, and eventually bringing about more intimate contact with
the rail. In order to secure an acceptable wear resistance, the brake
blocks should have a hardness of 220240 HB (Brinell hardness) and
the wheel rims a hardness of 240300 HB.
There are a number of other variables affecting traction between rail
and wheel, namely wheel load, wheel size, and braking or driving mode.
The wheel load seems to have an insignicant effect on the traction
coefcient fT. This is largely due to the fact that, for very high loading,
the case of railwheel contact, the real area of contact approaches the
apparent area size. The mean contact pressure is close to the yield limit
in compression for steel, and, according to the simple theory of the
adhesion component of friction, the traction coefcient remains
invariant
fT Gsp*

(6.4)

Rolling Contacts in Land Locomotion

205

where s represents the shear strength of the weaker material in contact


and p* denotes plastic ow or yield pressure. Increased wheel load will
also increase the apparent area of contact, but this affects both the
numerator and the denominator in equation (6.4) so that there is no
overall change in fT. For clean surfaces, the value of sp* is about 0.8
for steelsteel contact. In most cases, however, the usual presence of
oxides, rain, oil, or other contaminations reduces the effective shear
strength of the interface, s, so that the adhesion values in Table 6.1 are
obtained. With increasing wheel diameter, there is a small reduction in
mean pressure and a corresponding increase in apparent contact area.
Also, it appears that the traction coefcient is little different whether
braking or driving conditions prevail at the wheelrail interface.

6.1.2 Braking process


It is instructive to consider the effect of braking a railway wheel from
an initial travel speed of 100 kmh. The time taken for wheellock to
develop is approximately 1 s or less from the instant of brake application. During this period, both tractive effort and wheel angular velocity decrease progressively, whereas the velocity of slip between wheel
and rail rises in a non-linear way to attain a nal locked-wheel skidding
value of 100 kmh. The contact area between wheel and rail consists,
during that time, of a region of adhesion and a slip zone. This is schematically shown in Fig. 6.1. Both theory and experiment indicate a
longitudinal shear stress distribution within the area of contact, which
takes the form of the upper curve in Fig. 6.1. It should be noted that
the shear stress is conned to relatively low values within the adhesion
zone and reaches its maximum value within the slip region. Thus, the
well-known requirement that a certain relative slip velocity between surfaces is essential for a maximum friction is satised by the above observation. The distribution of longitudinal slip velocity in the contact zone
tends to have a non-linear increase towards the rear of contact according to the lower curve in Fig. 6.1. The slip region is followed by a kind
of overshoot as the band velocity attempts to return to its undeformed
value just outside the contact area. Another interesting fact is that,
although the apparent area of contact could be quite substantial, the
real area of contact is much smaller.
6.1.3 Traction enhancing techniques
One of the techniques used with considerable success in practical situations is the spark discharge method of volatilizing railwheel contaminations. It has shown a considerable improvement in traction values

206

Rolling Contacts

Fig. 6.1

after its application. In general, the technique consists of mounting a


single electrode of mild steel ahead of the test wheel and at a height 15
17 mm above the railhead. A similar electrode is placed to the right of
the test wheel surface. The test wheel is rigidly attached to an axle, on
the other end of which a similar wheel is free to roll. A hydraulically
operated disc brake applies a braking torque to the experimental axle
until the test wheel is on the point of slipping. This torque is proportional to the coefcient of traction between the wheel and rail and
can be easily measured using a dynamometric arm. A high voltage circuit triggers and sustains a spark discharge simultaneously across
the gap between the electrode and rail and across the gap between the
electrode and wheel at prearranged time intervals.

Rolling Contacts in Land Locomotion

207

6.1.4 Consequences of wheel and rail wear


It is inevitable that, as a result of wear occurring at the interface formed
by the contact between wheel and rail, the performance of the rolling
contact is affected. For new wheel and rail proles, the contact area is
little more than 300 mm2, and the contact pressure approaches the yield
limit in compression of the railhead material. Plastic ow of metal
occurs both at the surface between wheel and rail and more readily at
a small depth within the rail prole itself. This accounts for the smooth
and shiny appearance of rails in service compared with the rusted
appearance of those that have not been used for some time. As wear
progresses in both the steel tyre of the wheel and the railhead, the contact area widens and becomes elliptical in a transverse direction. Figure
6.2 shows the relative proles and contact conditions for three combinations of rail and wheel in both cornering and straight running. It can
be seen that wear of the rail top causes a progressively increasing slope
of its contact surface to the horizontal. Wear of both tyre and rail
increases the slope of the contact interface from about 1 :20 for the new
condition to 1 :6, while the centre of contact remains in about the same
position. If the rail but not the wheel is reground to the original prole,
this slope remains at a value of about 1:6, but the centre of contact
moves inwards with respect to the vehicle or in the direction of the
wheel ange. Meanwhile, the cornering condition produces only one
contact spot for this combination compared with the normal two
contact positions when the rail and tyre wear progressively together.
6.1.5 Ribbed tyre
The wheel used in land locomotion may consist either of a exible
(pneumatic tyre) or rigid structure (locomotive wheel). The locomotion
problem of a wheel on deformable terrain, such as soft soil or sand,
nds particular application in farm tractors and off-road vehicles. Here,
the properties of the soil are of primary importance in determining the
tractive capability or otation of the vehicle wheels. Soil, in general,
has both plastic and frictional properties, and the shearing stress in
soil depends on the coefcient of cohesion c and angle of friction .
Thus
GcCp tan

(6.5)

where p is the mean loading pressure acting at the wheelsoil interface.


Plastic masses such as saturated clays or certain types of melting snow
can be considered to have a zero angle of friction , and thus Gc. The
more common granular type of soil exhibits no cohesion or internal

208

Rolling Contacts

Fig. 6.2

bonding (cG0), so that Gp tan . One of the complications in


attempting to apply equation (6.5) to an actual soil is that the relative
proportion of c to p tan is very sensitive to water content. Furthermore, soils generally lack homogeneity which is characteristic of other
deformable materials.
Figure 6.3 shows, schematically, the traction of a ribbed tyre in soft
terrain. Both theory and practical experience indicate that the spuds or
ribs of elastic tyre treads clog with soil once they are in contact with
the ground. Therefore, they play an insignicant and secondary role in
developing traction. This is in contrast to the effect of tread pattern on
automobile tyres, as will be demonstrated later on in this chapter. The
main effect of the ribs is to increase the effective wheel diameter from
D to DC2t, where D is the undeected smooth tyre diameter and t

Rolling Contacts in Land Locomotion

209

Fig. 6.3

represents the rib depth as shown in Fig. 6.3. Let TD be the driving
torque applied to the wheel and FD the average value of tractive force
developed at distance h below the wheel centre. Then, taking moments
about the centre O
TD GFD hCWa

(6.6)

where a is the distance forward of the wheel centre at which the load
reaction vector W is effective at the tyresoil interface. Dividing both
sides of the above equation by h gives
TD
GFDCFR
h

(6.7)

where FR GW(ah) is the rolling resistance of the wheel. It should be


noted that the ratio ah for locomotion in soft terrain is considerably
larger in magnitude than that corresponding to the more usual tyre
road case. This is because the load is effectively supported on soft soil
only from the point of initial contact to the maximum penetration position, i.e. along the arc bc in Fig. 6.3. If both sides of equation (6.7) are
divided by W, the coefcients of friction rather than frictional forces
are obtained on the right-hand side. Although the FD and FR terms in
equation (6.7) appear to have the same sign, each is taken in a positive
sense according to the directions indicated in Fig. 6.3. Thus, they are
actually numerically subtractive. It is convenient, then, to divide each
term in equation (6.7) by the contact area A, in order to obtain
TD h
A

G Ap fR Gp (tan AfR )

(6.8)

From equation (6.5), for granular soils, cG0. The shear strength of
the soil at the shearing boundary is equal to FDA, p GWA, and the

210

Rolling Contacts

coefcient of rolling resistance fR GFR W. Equation (6.8) shows clearly


that the application of driving torque TD to the wheel is a function of
soil frictional properties and normal loading. It is important to assess
the bearing capacity or otation characteristics of soft soils and terrain,
and this is expressed quantitatively by p in equation (6.8).

6.2 Tyreroad interactions


The most important aspect of tyreroad interactions seems to be traction at the interface of their mutual rolling contact. Traction is essentially a friction process where tangential forces are transmitted and
controlled by varying the relative motion of rolling and sliding between
the contacting bodies. Land transportation relies, to a large extent,
upon tyreroad traction to achieve the expected high standard of
mobility, efciency of performance, and safety of road vehicles. The
force of traction or the tractive force is dened as the resultant tangential force of friction in the plane of contact across which a normal force
acts reciprocally between the tyre and the road. The traction forces
transmitted by tyres to the vehicle are primarily responsible for the
entire class of vehicle manoeuvres, i.e. acceleration, braking, cruising,
cornering, steering, handling, and even parking. In order to achieve
high performance and manoeuvrability of vehicles, the tyre and road
should have the potential for generating large frictional forces in
relation to the normal force being transmitted. Also, they should have
the ability to control smoothly the magnitude and direction of traction
forces as dictated by the dynamics of the intended manoeuvre.
It is well known that pneumatic tyres perform remarkably satisfactorily on dry roads on account of some of their unique properties. The
requirement of a large frictional force can be met easily because the
friction of tyre tread rubber on dry surfaces can be, under suitable
conditions, quite high. The smooth control of traction forces can be
achieved mainly through the elastic compliance of the tyre. However,
much of the benets coming from the excellent frictional properties of
rubber tyres may be lost when the tyreroad interface is wet or contaminated in any way. The frictional force resulting from shearing liquid
lms or contaminant layers is usually too small to generate adequate
traction. Also, on wet roads, the tyre compliance may turn out to be a
liability since it can promote elastohydrodynamic lubrication. In
addition, the low elastic modulus of the tread rubber can induce microelastohydrodynamic or mixed lubrication depending upon the texture
of the road surface. The nal remaining problem pertaining to tyre

Rolling Contacts in Land Locomotion

211

road interaction is direct tyreroad contact across a boundary lubricated interface. The possibility of activating various kinds of lubrication
mechanism on a wet road poses a serious problem for tyre and road
designers.

6.2.1 Relationship between friction and traction


Three distinct aspects are involved in the study of the process of traction
between tyre and road.
(a) the basic problem of generating friction through direct contact (dry
or boundary lubricated) at the contact interface of the tyre and the
road;
(b) the mechanics of slipping of a deformable tyre under the combined
action of rolling and sliding;
(c) the mechanics of elastohydrodynamic lubrication on a wet road.
In order to study the rst aspect, there is a need to know how large the
forces of static or sliding friction are, how they respond to the relative
motion between contacting surfaces, the normal force, and other inuential variables. The second aspect of traction is concerned with the
mechanics of slipping, and thus the extent of sliding and rolling of the
tyre is varied in order to control the force of traction. Owing to the
elasticity of the tyre structure, it follows that the contact area under a
normal force is nite and that distributions of normal force, relative
motion, and other important variables that affect friction are not uniform within this area. These two aspects of traction are sufcient for
analysing the traction process of tyres on dry roads. It is reasonable to
assume that the same type of analysis also applies to traction on wet
roads provided that water present at the interface gives rise to boundary
lubrication only. The effect of boundary lubrication may be taken into
account by appropriate modication of the basic laws of friction.
The third aspect of traction becomes relevant on wet roads where the
presence of water leads to some form of hydrodynamic lubrication of
the contact interface. The region of direct tyreroad contact and the
normal force on that region are both inuenced by lubrication. How
this affects the friction and traction between the tyre and the road will
be discussed later in this section.
The determination of available friction requires knowledge of the
basic laws or constitutive relations of friction of an elastomer on hard
solids. The frictional force is expressed in terms of a non-dimensional
quantity, i.e. the coefcient of friction , which is the ratio of the tangential force and normal force transmitted through the contact area.

212

Rolling Contacts

From a fundamental point of view, is not strictly a basic property of


the interface but an outcome of the friction process depending upon the
adhesion of the surfaces. On a macroscopic scale, for to be viewed as
a basic frictional property it is necessary for properties and the variables
inuencing friction to be prescribed uniformly over the entire contact
area. In particular, the most important variable, i.e. the relative motion,
should be uniform for all points at the contact interface.
Taking into account the non-uniform distribution of governing variables within the contact of tyre with road, it is assumed that the basic
laws of friction are only applicable locally. The local coefcient of friction at some point within the contact interface depends upon the
relative motion of surface elements of tyre and road at that point. The
local coefcient is dened as the ratio of the local tangential or shear
stress and normal pressure. The resultant of the local friction forces of
different magnitude and orientation is the tractive force, which, when
divided by the normal force, is expressed as the non-dimensional coefcient of traction f. The second aspect of the traction process is concerned with the utilization of friction (represented by ) by the slipping
tyre to control the traction coefcient f. It should be noted that, unlike
, coefcient f is not generally considered to be a basic tribological
property.
Under some optimum set of variables, the coefcient of friction
attains the maximum value that can possibly be generated with a given
combination of tyre, road, and interface. The maximum available friction sets an upper limit for the force of traction that can be generated
by the tyreroad system. Typical values of the two key coefcients
and f, representing the friction and the traction potential under different
conditions, are shown in Fig. 6.4. It is interesting to note that the
value of low-modulus rubber on smooth, clean, and dry surfaces can
be very large under favourable conditions (low normal pressure and
low sliding speed). The values of for dry friction of tread rubber
compounds on road samples taken from typical roads are usually much
lower. However, these values are of the same order of magnitude as
the traction coefcients of tyres measured on dry roads. The traction
coefcients of tyres on roads determined under different surface conditions also exhibit wide variations depending upon the variables affecting the traction process. On dry and reasonably clean roads, tractive
forces of tyres are not very sensitive to small variations in road properties, speed, and load.
The highest anticipated traction coefcients of tyres on dry roads are
in the range 0.81.2; under conditions favourable to traction it is not

Rolling Contacts in Land Locomotion

213

Fig. 6.4

unlikely to come across values of 1.5 on dry roads. In the case of racing
tyres, it is possible to generate a traction coefcient on dry roads that
exceeds a value of 2. Under laboratory conditions, the maximum friction coefcients for the various rubber compounds used in tyre treads
range from 0.8 to 4 depending upon the type of counterface in sliding
contact. These results show that the friction potential indicated by basic
studies is not fully utilized for generating tyre-on-road traction. Some
design and operational constraints are responsible for this but there is
still scope for achieving further improvements.
Figure 6.5 shows typical variations in with sliding speed (4). The
data for the low speed part of the curve are based on laboratory tests
with a tyre tread compound on a road surface of dry bituminous concrete. The part of the curve at high speeds is plotted from the results of
tyre tests on roads made from materials similar to that used during
laboratory tests. In the range of low sliding speeds, friction is almost an
isothermal process, but for higher speeds, the temperature rises rapidly
because of frictional heating. The non-stationary nature of heat transfer

214

Rolling Contacts

Fig. 6.5

strongly inuences the transient value of . The problem of severe wear


limits the duration of locked wheel sliding.
It is possible to control the tyreroad traction by varying the extent
of sliding motion of the rolling wheel. In order to dene the kinematics
of sliding and rolling more precisely, it is convenient to suppose that
the wheel and the road are two rigid bodies linked together by an elastic
body, i.e. the tyre. The translation of the wheel, together with rotation
about its axis, implies that there is a relative motion between the wheel
and the road in the plane of contact. The relative motion consists of
translation in the contact plane and rotation about a normal to the
plane. The two components of the relative motion are known as the
slip and the spin. The elastic tyre has to accommodate this relative
motion between the two rigid bodies if it is to roll without sliding.
The ratio of the components of relative velocity between the wheel
and the road, V r, referred to the interface, and the average rolling velocity of the wheel, V 0, is dened by a kinematic parameter termed the
slip, . It may be generalized to include both the slip and the spin
motions. The slip parameter may be regarded as an overall measure of
relative motion of the wheel with respect to the road, but it does not
contain any explicit information concerning the relative motion of sliding at the tyreroad interface. The kinematic description of rolling with
slip is a relation involving three vector quantities, namely: the linear
velocity of the rigid wheel, the angular velocity of the wheel about the
axle, and the radius vector normal to the contact interface.

6.2.2 Characteristics of the traction


It has been stated earlier that traction is strongly inuenced by slip and
it can be controlled by varying slip on the condition that it does not

Rolling Contacts in Land Locomotion

215

lead to sliding of the entire contact area. The functional relation connecting the two non-dimensional quantities f and slip is known as the
traction characteristics of the system consisting of the tyre, the road
surface, and the interface. The traction characteristics are inuenced
jointly by the tyre and the road through an interaction of two kinds of
properties, that is, the elastic properties of the tyre and the tribological
properties of friction and lubrication of the tyreroad interface.
Usually, the practical interest is in the effect of the resultant forces of
traction on vehicle dynamics, and the usual approach is to treat traction
as a black box system with and f taken as input and output quantities.
Bearing in mind that the road surface and its contamination change
continuously, the question of generality of the f relation arises.
Both the slip and traction, i.e. the force and moment, can be
expressed as vector quantities with components in the longitudinal, lateral (slip angle), and spin (turnslip). The components of V r, expressed
as three elements of slip ( x , y , ), are the longitudinal, lateral slip
(translation), and spin (rotation). The traction components are the two
important force components in longitudinal and lateral directions Fx
and Fy , the self-aligning torque Mz , the overturning couple Mx , and
the wheel torque My .
A set of f functions for a given tyreroad interface system may be
determined, corresponding to a set of time-invariant parameters such
as the normal force, slip, and other vectors. They are known as the
stationary traction characteristics. A family of such characteristics for
various permutations and combinations of parameter values is useful
for predicting the stationary type behaviour. In vehicle dynamic practice, non-stationary traction characteristics are also important because
variables and parameters such as the slip, load, speed, and road inputs
vary continuously with time and any attempt at analysis is an extremely
difcult task. Even the simpler stationary problem, outlined later on in
this section, poses a number of difculties. The traction characteristics
in the stationary state can be expressed in the form
[Fx , Fy , Mx , My , Mz ]G [ x , y , ]

(6.9)

Keeping x , y , , V 0, and Fy constant and assuming constant properties of the tyre, the road surface, and the interface, only a short rolling
distance equal to a few contact lengths is sufcient for generating a
stationary response. It is convenient to study traction when only a single
component of slip is considered at a time. This type of relation may be
described as the partial traction characteristics. One of them is the lateral traction characteristic for stationary rolling. The variation in lateral

216

Rolling Contacts

traction and the self-aligning torque with lateral slip, with the other slip
components equal to zero and the parameters constant, has the form
[Fy , Mz ]G1( y )

(6.10)

where function 1 denes the partial stationary characteristic of lateral


traction.
A typical stationary traction curve ( f as a function of ) may be
described, qualitatively, without performing a detailed analysis. As
shown in Fig. 6.6, the curve exhibits three broad regions. A region of
small slip where the curve is almost linear, a region of nite slip where
traction increases more slowly with increasing slip, and nally the
region of large slip where substantial sliding takes place. In the last
region, the traction curve can display a wide variety of shapes
depending almost entirely on the process of friction. On dry roads and
on wet roads where no hydrodynamic lubrication takes place, the initial
part of the f curve for very small has the same slope for all road
surfaces. The reason for this lack of sensitivity of the initial slope to the
tribological conditions of a dry and boundary lubricated contact
interface is explained later in this section. In Fig. 6.6 the parameter M

Fig. 6.6

Rolling Contacts in Land Locomotion

217

is dened as the ratio of the thickness of the undisturbed water layer T


and D a measure of the combined drainage of the road and the tyre
surfaces; the parameter Ghm , where hm is the minimum lm thickness within the contact area, calculated on the basis of smooth proles,
and is the root mean square value of peak heights of the surface
roughness of the road surface.
Traction on a dry road
From the practical point of view, the traction of a tyre on a dry road
is of considerable importance as the driving takes place mainly on
reasonably dry roads. The friction of representative tyres on a dry road
is taken as the norm to design vehicles and to specify its performance.
The performance estimates are based usually on the assumption that
the coefcient of friction (kinematic) of a tyre on a dry road is roughly
of the order of unity.
The traction curve starts from vanishing slip corresponding to the
conditions of free rolling where a small force of traction is needed to
overcome the rolling resistance of the tyre. When a wheel rolls with
small slip, the tangential forces are, on account of a nite static o ,
generated primarily by static friction. The slip is accommodated almost
entirely by elastic deformation of the tyre. The sliding involved in this
process is minimal and conned to a small region towards the trailing
edge of the contact. The frictional properties such as the coefcients o
and are not involved so long as the friction coefcients required to
sustain the elastic deformation do not exceed o. Therefore, traction is
insensitive to surface conditions. The action of static frictional force is
passive because elastic deformation of the tyre is predetermined. This
initial part of the f curve can be described as the elastic regime.
Because very little sliding is involved, the initial slope of the traction
curve for a dry and a boundary lubricated road does not depend on
values of o and .
As slip increases, sliding friction becomes increasingly important and
the traction depends essentially on the outcome of the detailed interaction
between the friction and elastic forces acting on the tyre. This part of the
traction curve could be named the triboelastic regime. Traction in this
regime is a basic property neither of the tyre nor of the road, taken separately. For large magnitudes of slip, sliding takes place on a large scale,
almost over the whole contact area. Then, traction is dictated mainly by
the process of friction. Traction, under conditions of pure sliding, is of
little use because the ability to direct traction forces and consequently the
directional control of vehicle motion are completely lost. Besides, gross

218

Rolling Contacts

sliding at high speeds is clearly undesirable from the point of view of


preventing damage by overheating and excessive wear. Practical interest
is in modelling traction characteristics in the range of small to moderate
slip. Occasionally, operation at higher slip is inevitable when, for
instance, driving takes place on a low-friction icy road.

6.2.3 Analysis of dry road traction


Triboelastic interactions of the tyre with the road govern the mechanics
of how traction is generated and controlled by varying the slip of the
wheel. It is concerned with the equilibrium of surface elements of the
tyre as they pass through the contact interface under the action of the
frictional and elastic forces. The kinematics of slipping inuences the
local sliding velocities and hence the frictional forces generated by the
elements of the contact surface. The resultant of the local frictional
forces within the contact is the tractive force generated by the tyreroad
interface. When rolling with slip, the basic tribological variables, i.e. the
speed of sliding and the local tangential stresses, vary from point to
point within the contact.
A proper analysis requires the independent specication of basic frictional and elastic properties, the formulation of the kinematics of rolling with slip, and solution of the rolling contact problem. Besides, heat
dissipation caused by rolling and sliding of the tyre results in an increase
in tyre temperature and consequently affects friction. Calculation of
the temperature of the tyre surface under various operating conditions
requires a thermal model of the contact between the tyre and the road.
It is usual practice to obtain a rough estimate of the surface temperature
of the tyre based on rolling and sliding speeds. This estimate utilizes
simple semi-empirical equations without the use of a more elaborate
thermal model.
The basic relations of friction of elastomers on a dry and hard road
surface are well established and can be found in published textbooks
dealing with the friction of elastomeric materials. The elastic properties
of interest here are those of the tyre because the road is relatively rigid.
The specication of the elastic properties of the tyre required for a
specic application is a difcult task.
Modelling of the response of an elastic tyre
Modelling of elastic deformation of a tyre is one of the most difcult
tasks in traction research. The problems under consideration are mainly
in the area of applied mechanics rather than pure tribology. Three
major difculties arise because the tyre is an air-inated structure

Rolling Contacts in Land Locomotion

219

which, when loaded, undergoes large displacements and deformation.


Unlike Hertzian contacts, the standard analytical techniques cannot be
applied to determine the shape and size of the contact area and the
normal pressure distribution under a given normal load. Although nite
element techniques may be used, nevertheless, it was found that their
use is rather limited and it is common practice to determine these allimportant parameters by experimentally testing an existing tyre.
A more serious problem of traction analysis is that the shape, the
size, and the normal pressure distribution are affected signicantly by
tangential forces. Thus, the contact problem becomes strongly non-linear and extremely difcult to solve. However, if such effects are ignored
in order to make the problem easier to analyse, the results, though
approximate, are likely to be acceptable.
Another task is to calculate the tangential deformations of the tyre
on account of surface forces of friction. This is the most important part
of any analysis of the triboelastic interaction. Assuming that the tyre
has a linear elastic response in the tangential plane, the deformation
due to the distributed contact force can be expressed by an integral of
a Green-type function. However, the well-known inuence functions of
massive, semi-innite bodies do not apply to tyres because of obvious
structural differences. Since the analytical solutions for such complex
structures are not known, it is necessary to determine the inuence functions, with the help of nite elements, of empirical techniques.
Assuming that the inuence functions are somehow determined and
provided that the tyre response is linear, the elastic displacement components u, v, w in the X, Y, Z coordinate system may be expressed by
an integral of the inuence functions. The XY plane is taken as the
plane of contact (road surface) and Z is the normal to the plane from
the centre of the wheel that intersects the contact plane at the origin of
the coordinate system. The following inuence, or Green, functions on
the boundary surface are needed to specify the elastic response of the
tyre to surface forces of traction, i.e. Gxx , Gyy , Gxy (GGyx ) for tangential
displacements due to the tangential traction stresses and Gzz for the
normal displacement due to the normal pressure on the contact.
In order to describe interactions between the normal and the tangential stresses and displacements, additional functions are required. It is
interesting to note that the elastic properties of the shell-like structure
of the tyre are quite different from those of solid bodies. The boundary
elastic response of a solid body is such that the inuence of normal
pressure on tangential displacement is more signicant than that of tangential stresses on normal displacement. In the case of the shell-like

220

Rolling Contacts

structure of tyres, these dependencies are reversed. Radial tyres are


especially sensitive in this respect as the lateral tangential components
of surface tractions have a strong inuence on the normal displacements
to the tread surface. As a result, for a given normal load, the shape and
size of the contact area and the distribution of normal pressure p are
signicantly inuenced by lateral components of tangential stresses. The
inuence of longitudinal components is generally much smaller. The
effect of the lateral traction stresses on the contact may be taken into
account by introducing the inuence function Gzy .
Tangential displacement

[ (, )G

u(x, y)G

xx

(x, y; , )C y (, )Gxy (x, y; , )] d d

(6.11)

[ (, )G

v(x, y)G

yx

(x, y; , )C y (, )Gyy (x, y; , )] d d

(6.12)
Normal displacement

[ p(, )G (x, y; , )C (, )G

w(x, y)G

zy

(x, y; , )] d d

(6.13)
The size and shape of the contact area C are determined by the compatibility of the deformed surface of the tyre with the road surface. Both
the geometry of C and the distribution of normal pressure p are
obtained for prescribed normal displacements within C. Taking into
account the difculties of solving normal contact problems, the modication of C on account of y has to be determined either by empirical
means or by the use of the nite element method. Under a purely
normal force, the contact geometry and normal pressure, denoted by
CGC 0 and pGp0 respectively, have to be modied using certain empirical functions and . Then, CGC 0 (w) and pGp0 (w). In addition,
the following two conditions apply to points (x, y) on the tyre surface:
pG0 outside C and pH0 inside C.
Owing to the complexity of the elastic structural response of a pneumatic tyre, a number of simplied models have been proposed. Most of

Rolling Contacts in Land Locomotion

221

them are well suited to represent, in an approximate way, the elastic


deformation of a tyre for a single mode of slip.
Rolling motion combined with slip
The kinematics of rolling with slip is an important element of the analysis of the mechanics of interaction between pneumatic tyre and road
surface. Consider a tyre rolling with a linear velocity of the wheel V 0.
The slip components x and y and the spin resulting from relative
motion between the wheel and the road surface, considered as rigid
bodies, may or may not lead to relative motion of points within the
contact zone. Whether the local sliding at certain points within the contact zone takes place depends upon the tangential displacements of the
tyre. The kinematics of sliding is expressed by the relations dening the
x and y components of the local sliding velocity, i.e. Vsx and Vsy

V C
x
t
0

Vsx (x, y)G xA yA

Vsy (x, y)G yC xA

V C
y
t

(6.14)

(6.15)

For stationary contact, the non-stationary terms disappear. Local sliding occurs when
V 2sGV 2sxCV 2syH0

(6.16)

If Vs G0, then there is adhesion or static friction at that point.


At any point given by (x, y) inside C, the tangential displacements of
the tyre are caused by local stresses produced by frictional forces which,
in turn, depend upon the local sliding velocity Vs and the normal pressure p at that location. The contact described by C can be split into two
regions, namely the region C A of static friction (due to adhesion) and
region C B representing kinematic friction (sliding).
For (x, y) in C A, i.e. if Vs G0
1( 2xC 2y) o ( p)p

(6.17)

and for (x, y) in C B, i.e. if Vs 0


1( 2xC 2y )G (Vs , p, T)p

(6.18)

222

Rolling Contacts

where (Vs , p, T)H o ( p). Also

x Gp

(6.19)

(6.20)

Vsx
s

y Gp

Vsy
s

The two non-linear functions o( p) and (Vs , p, T ) are the basic


relations of dry friction. If a thermal model of the tyre is available, the
calculations of T can be coupled to the present model to take into
account the inuence of on the surface temperature. In most applications it is sufcient to calculate the temperature T approximately
using an elementary expression with heat input calculated from V 0 and
, and with a reasonable estimate for .
The contact problem, dominated by interactions between friction and
elastic deformations, is completely formulated when elastic properties
of the tyre given by the surface inuence functions Gxx , Gyy , Gxy , Gzy ,
and Gn are specied, and the traction-free contact area C 0, the pressure
p0, the functions and for the traction inuence, and the friction
parameters are all known. The set of elastic response, kinematic, and
friction law equations together form a mixed boundary problem where
C A and C B are also unknown. The longitudinal and lateral components
of the traction force, Fx and Fy , and the moments can simply be
expressed as integrals over C, once the distributions of the tangential
stresses have been determined.
The solution of the equations requires a study of its own. An iterative
approach is needed to determine C A and C B. In spite of the difculties,
which are mainly of a technical nature, the possibility of being able to
predict tyre characteristics from the basic properties of the wheelroad
system is of considerable practical interest.

6.2.4 Traction under wet conditions


Undoubtedly, the friction on a wet road surface is a problem of considerable concern for safety of driving. The elastohydrodynamic lubrication (EHL) taking place at the interface is the main cause of the
degradation of traction on wet roads. The main reason for that is the
intrusion of water into the contact interface which is facilitated by the
elastic deformation of the tyre. At some speed, a point is reached where
the frictional interface created by the contact between the tyre and the
road is separated by a lm of water which, while transmitting the

Rolling Contacts in Land Locomotion

223

normal force, ceases to transmit any sizeable tangential force. The viscous friction due to shearing in the water layer is negligible in relation
to dry friction, and traction is neither sufcient nor controllable. Under
wet conditions, three distinct mechanisms of lm formation may be
identied. The conditions under which one of these mechanisms is
dominant depend upon the speed of rolling and the slip of the tyre.
Also, the thickness of the water layer present on the road is of signicance. In addition, surface features of tyre and road surfaces can play
an important role in draining the water out of the contact zone. The
geometrical features of the road surface are usually characterized by the
average size of surface irregularities, i.e. a macroscale texture with a
linear dimension of 110 mm and a microscale texture with features of
the order of 0.1 mm. The depth of the tread prole of a tyre is of the
same order of magnitude as the size of the macroscale texture of the
road. The three EHL mechanisms are: the inertial or thick layer mechanism, the viscous or thin layer mechanism, and micro-EHL which
occurs on tips of surface asperities.
Inertial EHL most often takes place on roads ooded after heavy
rain when thick layers of water may be present at various locations. If
the combined drainage due to the road macrotexture and tread patterns
is insufcient, the impact of water with the tyre causes large hydrodynamic pressure to build up ahead of the contact area. Clearly, the
effect of uid inertial forces acting on the tyre depends on the thickness
of the water layer and the speed of rolling. Under the high pressure of
water, the surface of the tyre is deformed inwards, which permits the
water lm to penetrate further into the contact zone. There is a critical
speed at which the entire contact is water-borne. This inertial EHL
where inertial forces are dominant is known as dynamic hydroplaning.
The speed V D at which there is onset of dynamic hydroplaning is a
function of the parameter M GTD, where D is a measure of the
combined drainage of surfaces of the road and the tyre, and T is the
thickness of the undisturbed water layer. The structural factors of the
tyre are the membrane stiffness due to the ination pressure pi and the
exural rigidity EI of the tread. The relation describing the propensity
for hydroplaning has the following form: VV D Gfunction of
[ M , V 2( piCEIl 4 )]1, where is the density of the uid and l is
the contact length.
There are, basically, two ways to suppress dynamic hydroplaning.
The rst consists of reducing the cause, i.e. the force of impact of water.
The strategy of a designer is to offer a part of the useful contact to ease
the ow of uid so that the remaining part is starved of lubricating

224

Rolling Contacts

uid. This is usually done by providing grooves and macroroughness


on the running surface of the tyre and the road. The second approach
is to reduce the effect, i.e. the inward deformation of the tyre. This is
accomplished by increasing the radial stiffness (the ination pressure)
of the tyre.
Under ooded conditions, the combined drainage of the road and
the tyre helps to reduce the hydrodynamic pressure due to the impact
of water against the tyre surface. The importance of effective drainage
is manifested at high speeds of rolling, especially when it takes place on
a road with little or no macrotexture and with tyres whose tread grooves
have been worn out. The use of open macrotexture and porous surfacing materials contribute to the effective drainage of water from the
tyreroad surface contact zone. If the road macrotexture is inadequate,
primary drainage can only take place through the grooves on the tyre
surface. An inherently stiffer tread construction helps to reduce the
deformation due to water pressure and also gives the tyre designer
greater freedom in the geometrical layout of tread patterns for increasing the rate and dispersion of the ow of uids. The danger of dynamic
hydroplaning is very much reduced under normal driving conditions
when tyres have an adequate groove depth.
There is no signicant benet to be gained from the tread patterns
on the tyre surface if the macrotexture of the road is sufciently large
to secure adequate drainage. On roads covered by a thin layer of water,
full lm lubrication can still occur owing to viscous effects. This can
take place at relatively low rolling speeds of the tyre and is termed
viscous aquaplaning or viscoplaning. Viscoplaning can occur, despite
sufcient drainage, on smooth roads at vehicle speeds of only 40
60 kmh. Two different scales can be distinguished for the viscous EHL
effect. Firstly when a thin but essentially continuous lm is formed on
a signicant part of the contact, and secondly when very thin lms are
created locally at tips of asperities. The effectiveness of repelling water
from the contact zone and preventing the build-up of EHL lm by the
rst scale effect depends on the ratio m Ghm , where hm is the minimum lm thickness calculated for smooth surfaces and is the root
mean square value of peak heights of the surface roughness (ignoring
the waviness) of the road surface. In order to prevent EHL lm buildup, mF1. The parameters affecting viscous EHL and the structure of
the equation describing isoviscous EHL of low-modulus materials have
been thoroughly reviewed by specialists in the area of lubrication. However, utilization of the results for the case of tyreroad contact has not
been fully implemented.

Rolling Contacts in Land Locomotion

225

The second scale effect, taking place at the tips of asperities, may be
viewed as a problem of point contact EHL applicable to soft materials.
The way to prevent the formation of very thin lms at the contact
between individual asperities is to increase the intensity of local pressure
so that the lm breaks down. It appears that somewhat higher intensity
of local pressure within the contact owing to tread patterns is not
enough to destroy thin lms. Thin micro-EHL lms are formed readily
on smooth tips of road surface asperities in contact with the low-modulus rubber tread of the tyre. The high pressure required to prevent the
formation of thin EHL lms can only be achieved on road surfaces
having a sufciently sharp microstructure.
The microstructure of the road surface is usually described in qualitative terms; i.e. at one extreme the road surface may have asperities
with smooth, polished and rounded tips while at the other extreme surfaces are described as harsh or sharp. A sharp microstructure is more
effective in penetrating thin lms and hence provides a signicant
reduction in the risk of viscoplaning. A rounded microstructure may
actually help build up thin lms. A change in microstructure from sharp
to rounded may occur on roads where intense road trafc produces
polishing of the tips of asperities. The choice of suitable road surfacing
material as well as the geometry of surface asperities (average slope, tip
curvature) are all important in controlling viscoplaning. The transition
from micro-EHL to boundary lubrication may be dened by the ratio
of the thickness of the boundary layer and the thickness of the microEHL lm.

hb
h

(6.21)

The signicant contribution of tread patterns to wet road friction is


generally recognized, but the precise function of the complex geometrical features is only vaguely understood. The distribution of pressure in
the contact area depends on the structural behaviour of the tyre under
normal load as well as on the geometry of tread patterns. Ideally, the
pressure distribution should be such that more water is expelled much
faster out of the contact through the grooves. Also, it is thought that
the grooves serve as low-pressure reservoirs into which water may accumulate temporarily during the passage of a tread element through the
contact zone. Several secondary features such as sipes or knife cuts are
also provided, presumably for squeezing out the thin lms. The narrow
slits can collect water by a capillary action and act as tiny reservoirs.
Experimental studies of friction on wet road surfaces do not report any

226

Rolling Contacts

signicant effect of these features except on extremely smooth surfaces.


There is no conclusive evidence on the benets resulting from such elaborate and intricate designs.
Undoubtedly, the deep grooves in the tread are clearly benecial for
the drainage of water. However, the inuence of lowering the tread
stiffness on traction under the action of inertial and viscous EHL
should also be considered. The situation where the pressure in the central region is lower than at the boundary is regarded as unfavourable
because water tends to be channelled towards the centre. This may arise
when the tyre is underinated or overloaded. There is also the
accompanying deformation which tends to close the grooves and
thereby increases the risk of dynamic hydroplaning at lower vehicle
speeds.
It is obvious that such a heuristic approach may not always lead to
correct solutions of this complex problem. Nevertheless, the tyre
designer must take care to verify that the tread patterns provided for
drainage of the contact area do not help to enhance the build-up of
hydrodynamic pressure. Only through a detailed analysis of surface
deformation of the tyre, considered together with tread pattern and
sipes, is it possible to assess whether the design is able to achieve the
desired effect. It is very important to make sure that the tread surface
does not act as a seal trapping pockets of water within the contact area.

6.2.5 Analysis of wet road traction


Undoubtedly, the traction phenomenon on a wet road is much more
complex than that on a dry road. The ingress of the lm of water into
the tyreroad interface inuences the entire traction characteristic,
including the initial slope of the traction curve in the elastic region. The
extent to which the EHL lm intrudes and how large the associated
lm pressure can become is determined by the competing action of
hydrodynamic and elastic forces. As stated earlier, dynamic hydroplaning and viscoplaning are the two extreme conditions initiated by distinct
EHL mechanisms operating during wet road traction. The rst type is
dominated by inertial forces and the second by viscous forces. It was
emphasized earlier that the EHL due to viscous effects can occur at two
levels. At the macrolevel, the formation of a continuous lm is determined by the overall geometry and the elastic deformation of the tyre
as a whole. At the microlevel, when the lm is dispersed, the microEHL process occurs because a tread element has to go through the uid
lm and squeeze the lm away before it can establish a direct contact
with the surface of the road.

Rolling Contacts in Land Locomotion

227

There has been signicant progress in understanding dynamic hydroplaning (5). A detailed analysis of the uid ow between a at road and
a rigid surface having a shape that corresponds to the deformed shape
taken by a hydroplaning tyre has been carried out. The importance of
inertial forces in the phenomenon of dynamic hydroplaning of tyres has
been stressed, while the role of viscous forces has been considered to be
only marginal. The micro-EHL problem has also been studied (6).
It has to be stated quite clearly that the conditions of contact between
tyres and roads for ordinary driving are not such that tyres can encounter a thick layer of water at high speeds leading to dynamic hydroplaning. Full lm dynamic hydroplaning occurs very rarely. Likewise, total
viscoplaning due to the persistence of micro-EHL lms on roads devoid
of any texture is an extremely rare event. From the practical point of
view, the regime of lubrication most representative of average driving
conditions on wet roads is the partial EHL regime. This conjecture is
based on the experience that, while the traction forces on wet roads are
signicantly smaller than those on dry roads, the forces are still quite
substantial. This suggests that direct tyreroad contact takes place on
a signicant part of the contact interface.
Partial EHL of contact interface
The interaction between the uid lm and the elastic tyre forces is critically important for modelling EHL effects at the contact interface. The
usual approach to the analysis of lubrication is to start from the
assumption of a full lm and then to calculate the minimum lm thickness by solving the combined elastic and hydrodynamic equations. The
two equations may be coupled together as integral or integraldifferential equations which can be solved numerically. The more common techniques use either direct or inverse iteration and select a trial solution
based on a physically justied simplication of the problem. Regardless
of the method of solution, an implicit assumption made in the formulation of the problem is that the solids are separated totally by a continuous lm of lubricant. The probability of solid asperities making
contact is considered to be low if 2.5, where is the dimensionless
ratio of the minimum lm thickness, calculated assuming smooth surfaces, and the standard deviation of heights of asperities on interacting
surfaces.
In the outline of partial EHL presented here, the conguration is
adopted a priori to ensure that direct solidsolid contact is established
on a part of the interface bordering on the trailing edge of the contact
area. Also, it is assumed that the pressure build-up in the inlet zone is

228

Rolling Contacts

Fig. 6.7

caused by inertial effects. The situation leading to partial EHL at the


interface of the rolling tyre is depicted in Fig. 6.7. For simplicity, the
region of interaction is rectangular and identical to the region of dry
contact C. This assumption is probably justied if the water layer on
the surface of the road is thin.
It can be seen from Fig. 6.7 that the component of the uid ow in
the direction normal to boundary b, which separates zones II and III,
is zero. The boundary is a plane curve b(x, y)G0, which cuts across the
width of the contact and terminates at the side edges in the shoulder
region of the tyre. The tangential component of the ow will decrease
owing to the boundary layer along b. The interfacial area C is thus
divided into two regions: region C f is the area of unbroken water lm
and region C b belongs to the third zone where direct tyreroad contact
is established under the conditions of boundary lubrication. It is obvious that CGC fCC b. In an actual analysis, the curve b could be represented by a parabola y 2 G (xAx), with parameter assumed to be
a constant, and a variable parameter x to locate the point of transition
from zone II to zone III along the axis of symmetry, which is the x axis.
In region C f, the uid lm thickness h(x, y) at any point is linked
with the lm pressure p f through the Reynolds equation of lubrication
(7). The boundary conditions on b are dened by the requirement that
there is no uid ow normal to this boundary. In the following equations, superscript f is omitted for convenience and p denotes the lm
pressure p f
1 p 3 1
Qx G
h C h(2V oAVsx )
(6.22)
12 x
2

Rolling Contacts in Land Locomotion

Qy G

229

1 p

1
h3C hVsy
12 y
2

(6.23)

The scalar product of the ow and the local normal vectors vanishes at
b; the expressions for Vsx and Vsy were introduced earlier in this section.
Assuming isothermal conditions in the case of wet road traction, the
behaviour of water may be taken as that of an isoviscous Newtonian
uid having dynamic viscosity . Under conditions of stationary rolling
with slip, the Reynolds equation for region C f expressed in a Eulerian
frame has the following form

h
C
h
x x y y
3

G6 (2V oAVsx )

h
x

C6 Vsy

h
y

C6 hV o

2u

2v
C
x2 y2

(6.24)

In stationary rolling and slipping, the local surface velocity of the tyre
varies both in the x and y directions depending on the elastic strains,
which can be expressed in terms of the displacement components u and
v occurring in these two directions. The variation in the tangential component of surface velocity resulting from elastic strain differentials of
rolling bodies represents a stretch effect. The stretch effect can signicantly affect the lm pressure and thickness in the case of a slipping
tyre. Under certain conditions, this effect could be utilized either to
reduce or to increase the lm pressure and thickness of the intruding
uid.
The mechanics of uid ow in inlet zone I is determined by the overall ow pattern around the tyre. If the water layer is thicker than the
depth of the macrotexture on the road, inertial forces become signicant
and the theoretical framework has to start from the NavierStokes
equations. In simplifying the complex problem, it is now propounded
that the effect of the inlet zone is to boost the pressure pi at the inlet
edge to some signicant fraction of the stagnation pressure. If, however,
the water layer is very thin, inlet pressure is boosted mainly by viscous
forces which depend on the shape of the deformed surface of the tyre.
The pressure at the inlet edge may be expressed as some fraction of
the stagnation pressure, although in the latter case this fraction may be
quite small. The average pressure of the uid will be reduced as a result
of drainage through the channels on the surfaces of the road (macrotexture) and the tyre grooves. This reduction may be taken into account
by decreasing further the value of factor . The effective pressure at the

230

Rolling Contacts

inlet edge leads to the boundary condition


pi G1V o

(6.25)

where F0.5. At all the other free boundaries, pG0.


The nite pressure developed at the inlet edge will force the uid to
ow through the relatively narrow interfacial region bounded by the
solid surfaces. This may be considered as a pressure-driven ow and it
is governed mainly by viscous forces. In addition, there is the usual
shear ow due to the velocity eld imposed by the solid boundaries
relative to the uid lm.
The expressions relating the local elastic displacements of points on
the surface of the tyre and the interfacial pressure and shear traction
stresses are given by equations (6.11) to (6.13). These can be
expressed explicitly once the inuence functions of the tyre have been
determined. The most characteristic feature of the EHL problem is
the elastic response expressed by the normal surface displacement due
to the lm pressure in C f and the pressure exerted directly by the road
in C b.
Taking into account the highly compliant structure of the tyre and
the road surface covered with a thin layer of water, the normal displacements of the tyre are expected to be much larger than the thickness of
the uid lm. This implies that the pressure distribution in the contact
region of the partial EHL interface may not be much different from
that in the corresponding region of the dry interface. Using superscripts
for the normal pressure and displacements (d for dry contact, f for the
water lm, and g for the boundary region), the EHL problem may be
formulated in terms of the difference between the normal displacements
in the dry and the partial EHL problems.
The difference between normal displacements at any point (x, y)
resulting from the difference in the normal pressures between the dry
and partial EHL contacts consists of two main parts, i.e. the lm thickness h(x, y) in C f and a small change in the rigid body displacement r o
of points remote from the contact. The change in the normal pressure
at any point from the dry contact situation to the partial EHL situation
is denoted by p(, ) which is
in C f
p(, )Gp f (, )Ap d(, )
in C

(6.26)

p(, )Gp g(, )Ap d(, )

(6.27)

Rolling Contacts in Land Locomotion

231

From equation (6.13) and ignoring the effect of shear traction


in C f
w f (x, y)Awd(x, y)Gh(x, y)Cr o

p(, )G (x, y; , ) d d

p(, )G (x, y; , ) d d

(6.28)

in C g
wg(x, y)Awd(x, y)Gr o

p(, )G (x, y; , ) d d

p(, )G (x, y; , ) d d

(6.29)

The condition that the total normal load remains unchanged is

p(, ) d dC p(, ) d d G0
C

(6.30)

While the basic structure of the analysis is outlined here, it remains to


investigate the detailed steps necessary to implement even an approximate numerical solution to this problem. The use of variational formulation may be required to solve the numerical problem.
A simplied approach to the partial EHL problem characteristic for
a tyre on a wet road is based on the assumption that the ow is onedimensional (longitudinal ow, the lateral ow is neglected). The analysis uses the choked lm conguration as shown in Fig. 6.8, where the
owrate Qx G0. As an illustration of the approach, the pressure developed by the uid for a slider with a parallel lm prole will be considered. Since the lm thickness is constant until the point of choking,
it is observed that, if the inlet pressure is zero, the pressure prole
increases linearly within the lm. The solution can be obtained without
making explicit use of the Reynolds equation as the condition that ow

232

Rolling Contacts

Fig. 6.8

vanishes, together with the inlet boundary condition, is sufcient to


determine the pressure prole. Since there is also no pressure ow, the
inlet pressure acts as hydrostatic pressure, and this will lead to a gross
overestimation of the load supported by the lm. This deciency can
be corrected if the decrease in the mean pressure owing to side leakage
over the front part of the interface region can be estimated either
empirically or from the approximate solution of two-dimensional ow.
The decrease in the quasi-hydrostatic pressure along the lm may be
taken into account by letting G (x).
The result of the parallel lm problem may be extended to a more
general lm prole h(x). As an illustration, a normalized contact region
(1, 1) to represent the initial dry contact between a free rolling tyre
and the road under a constant normal load Fz will be considered. The
lm of water initially enters the inlet region and the pressure is boosted
at the inlet edge, that is, xG1, just before entering the contact. The
hydrodynamic pressure at the inlet edge is
p(1)Gpi G (1)1V o
(6.31)
It is assumed that the lm intrudes into the interface up to the point
xGx. Assuming that Qx G0 results in
dp 12 V o
G
(6.32)
dx
h2
Also
p(x)G12 V o

dx
2
1 h

(6.33)

Rolling Contacts in Land Locomotion

233

Adding the quasi-hydrostatic component of p f produces the total uid


pressure pf in 1xx in the form
p f (x)G12 V o

dx
C (x) 1V o
2
1 h

(6.34)

The normal displacement w(x) at any point on the surface of the tyre
can be expressed using the integrated one-dimensional version of the
normal inuence function Gn given by equation (6.13). Thus

w(x)G

[ p( )Gn (x; )] d

(6.35)

The formulation of the integral equation for the change in pressure


p(x) follows the same lines as that in the two-dimensional ow case
discussed above. However, the numerical solution in the present case is
likely to be easier to obtain. The tyre inuence functions are based on
a model of the tyre as a circular ring on an elastic foundation in both
the radial and tangential directions. This simple model is able to represent and retain the most essential characteristics of the elastic
response of a radial tyre. The simplicity of the model and the onedimensional approach make it easier to explore the inuence of the
tangential deformations of the tyre rolling with longitudinal slip.
When steady state conditions prevail and only rolling with longitudinal slip is considered, the one-dimensional Reynolds equation is
h
G6 (2V AV
dx dx

dp

sx

dh
dx

C6 hV o

d 2u
dx2

(6.36)

where is the dynamic viscosity of water, V o is the linear velocity of


the wheel, u is the longitudinal elastic displacement of the tyre, and Vsx
is the longitudinal sliding velocity which was introduced earlier.
The formulation of the partial EHL problem discussed above clearly
indicates the nature of inuence of the wet road conditions on the traction characteristics. The inuence function of normal deformation has
a role of some signicance for wet road traction of the slipping tyre. It
is not only limited to reduction in the maximum force of traction; the
entire traction curve can be affected. The inuence of the decrease in
the region of direct contact on the initial slopes of the traction curves,
at vanishingly small values of slip and spin, can have signicant effects
on vehicle handling and stability. The measures required to alleviate
these undesirable effects are the responsibility of the tyre designer and

234

Rolling Contacts

tribologist, as this task is clearly beyond the scope of the vehicle engineer. However, the ability to estimate or predict the changes in initial
slopes of the traction curves is one of the most basic requirements for
counteracting and controlling the vehicle behaviour under these
conditions.

6.2.6 Practical approach to traction modelling


The specication, prediction, and control of the dynamic performance
of vehicles require some sort of model able to predict the traction
characteristics. In view of physical and mathematical complexities, only
a few researchers have attempted to solve the triboelastic problem using
the actual and independently determined properties of friction and elastic response of the tyre. Also, practising tyre and vehicle engineers have
developed highly simplied models to deal with a wide range of traction
problems. The main aim has been to use simple representations based
on models or formulae, which, although relatively simple, exhibit qualitative features similar to results obtained from experimental work on
traction. If the qualitative behaviour is similar, the quantitative behaviour may be matched empirically by using techniques of parameter
identication. In this way, simple expressions to describe the dry traction characteristics of stationary rolling tyres when a single mode of
slip is prescribed at a time have been constructed. They give a good t,
especially in the elastic regime. The empirical traction formulae are useful for the preliminary analysis of problems of vehicle dynamics. Two
approaches have evolved for modelling the traction behaviour of the
tyreroad interface system. The rst one is a semi-empirical approach
where the two basic properties of friction and elasticity are specied
independently, although in a rudimentary form. Using these properties,
the simplied mathematical problem of the triboelastic interaction of
the slipping tyre is then solved. The parameters representing the fundamental properties are the best ts determined from suitable traction
experiments in the laboratory or on the road. The experiments are performed with representative tyres and roads and with a dry interface.
Usually, the elastic properties of the tyre are represented by a variety
of simple mechanical elements such as a set of isolated springs or a
beam on an elastic foundation. The frictional properties of the interface
are modelled according to the AmontonCoulomb law, sometimes
including modications inspired by experimentally observed trends of
the traction characteristics.
Taking into account the fact that the basic properties are independently specied but not determined in a rigorous way from fundamental

Rolling Contacts in Land Locomotion

235

experiments, the results of analysis can serve only as qualitative guidelines. Although quantitative estimates may be made to match empirical
results by providing a sufciently large number of free parameters, it is
doubtful whether such ne tuning of the model has any physical signicance and meaning.
Although the semi-empirical models are not entirely satisfactory,
their main strength is that the formulae derived to represent the traction
behaviour possess a basic structure that is qualitatively sound. It also
implies some limited degree of generality for predicting the main trends
of traction for small but nite deviation from the nominal operating
conditions and parameters. The models are known to perform reasonably well when describing traction on dry roads and to a limited extent
on wet roads under boundary lubrication conditions. The models do
not strictly apply to traction under the more general conditions of wet
roads because fundamental principles of lubrication are not included in
such models.
A more pragmatic approach to the problem is represented by the
second type of model. Empirical models are derived directly by curvetting of data obtained from indoor and outdoor traction measurements. One of the main goals is to nd easy-to-use and time-efcient
representation of the empirical relations. Among the various possible
methods of function representations, those making use of special functions that match the required shape of the empirical curves are generally
both accurate and economical. The value of such models can be further
enhanced if the salient features of traction curves can be expressed
directly and simply by the parameters of the special function.
The main motivation for developing empirical traction expressions is
twofold. Firstly, the vehicle dynamicist requires only the overall input
output relations of traction in order to connect the forces and moments
acting on the wheel to the wheel slip, without going into details of
interface tribology. Secondly, the modest traction requirements under
average driving conditions correspond to operation of tyres mainly in
the elastic regime, i.e. small slip, of the traction curve where traction is
relatively insensitive to variations in the surface condition of the road.
The empirical traction models make use of the special frictional
behaviour of dry, high-quality road surfaces and, therefore, there are
severe limitations to the range of road conditions and operational variables for which such models can provide useful predictions. Under more
general surface conditions, the strong inuence of friction is bound to
cause large variations in traction in the triboelastic and frictional
regime. Ideally, the traction versus slip curves should be determined

236

Rolling Contacts

from tests over the whole range of slip. Another problem with the accuracy of empirical models is the difculty of specifying and controlling
test procedures and subsequent evaluation and interpretation of test
results. This is particularly true for the results pertaining to both the
elastic (low slip) and the frictional (high slip) regimes. In the elastic
regime, the difculties arise partly on account of deviations and uctuations in geometrical or physical properties which are inherent to the
tyre and the road surfaces. Also, inaccuracies related to test equipment
and procedures contribute to the difculties mentioned above.
The parameters of the traction model that are most difcult to determine are those representing the fundamental frictional property of the
contact interface. The obvious option for determining the frictional parameters is to measure them under conditions of uniform sliding, i.e.
when the wheel is locked. Unfortunately, the friction coefcient measured under locked-wheel sliding conditions is characterized by a large
scatter as the value is sensitive to the duration of the test. The reason
for this is the transient rise in the surface temperature which depends
on the frictional heat dissipation. For the same reason, measurements
carried out at different test speeds give different values of sliding friction
coefcient. Locked-wheel testing at representative road speeds can be
sustained only for a short period of time to avoid risk of damaging the
tyre as a result of overheating and excessive wear. All the above difculties of measurement and interpretation point to the complexity of
the system and the futility of representing friction by a single constant
coming from the Coulomb model.
It is quite clear that the model parameters related to traction
measurements in both regimes cannot be determined with a sufciently
high degree of accuracy. The data required to identify parameters are
probably most accurate only for moderate slip values, i.e. in the triboelastic regime on condition that the test road surface is sufciently isotropic. As it happens, traction in the triboelastic regime is governed by
the combined properties of friction and elasticity, and therefore it is
difcult to identify the separate parameters associated with the two
basic properties. However, such a separation is required if the traction
model is to apply for the same tyre rolling on different road surfaces.

6.3 References
(1) Bekker, M. G. (1962) Theory of Land Locomotion (University of
Michigan Press, Ann Arbor, Michigan).

Rolling Contacts in Land Locomotion

237

(2) Hills, D. A., Nowell, D., and Sackeld, A. (1993) Mechanics of


Elastic Contacts (ButterworthHeinemann, Oxford).
(3) Moore, D. F. (1975) Principles and Application of Tribology
(Pergamon Press).
(4) Grosch, K. A. (1963) The relation between the friction and viscoelastic properties of rubber. Proc. R. Soc., A274.
(5) Browne, A. L. (1975) Mathematical analysis for pneumatic tyre
hydroplaning. ASTM special technical publication, p. 583.
(6) Moore, D. F. (1967) A theory of viscous hydroplaning. Int. J. Mech.
Sci., 9.
(7) Reynolds, O. (1886) On the theory of lubrication and its application
to Mr Beauchamp Tower experiments including an experimental
determination of the viscosity of olive oil. Phil. Trans. R. Soc.,
Lond., 177.

This page intentionally left blank

Chapter 7
Machine Elements in Rolling Contact

7.1 Contact of meshing gears


7.1.1 Peculiarities of contact between gear teeth
If two parallel, curved surfaces, such as the proles of meshing spur
gear teeth made of a rigid material, were pressed together, they would
make contact along a line, which implies that the area of contact would
be zero, and the contact pressure innite. However, there are no absolutely rigid materials, so deformation of an elastic nature occurs, and a
nite, although small, area of contact supports the load.
The case of two cylinders of radii R1 and R2 was solved by Hertz (1).
If the case of two steel cylinders is considered, for which G0.286, then
the maximum compressive stress is given by
pmax G0.416

PE

R CR
1

(7.1)

where P is the compressive load per unit length of the cylinders and E
is the equivalent Youngs modulus. If the radius of relative curvature
R of the cylinders is dened as 1R1C1R2 , then
pmax G0.416

PE
R

(7.2)

It should be noted that the stress given by equation (7.2) is one of the
three compressive stresses and as such is unlikely to be an important
factor in the failure of the material. The maximum shear stress occurs
at a small depth inside the material and has a value of 0.3pmax. At the

240

Rolling Contacts

surface, the maximum shear stress is 0.25pmax. When sliding is introduced, however, a tangential stress eld due to friction is added to the
normal load. As the friction increases, the region of maximum shear
stress, located at half the contact area radius beneath the surface, moves
upwards while, simultaneously, a second region of high yield stress
develops on the surface behind the region of contact. The shear stress
at the surface is sufcient to cause plastic ow when the coefcient of
friction reaches a value of about 0.27. These stresses are much more
likely to be responsible for the failure of the gear teeth. The important
point for the designer at this stage is that each of these stresses is proportional to pmax, and therefore for any given material they are
proportional to 1(PR).
For a number of reasons, however, this result cannot be directly
applied to gear teeth. The analysis assumes two surfaces with constant
radii of curvature and an elastic homogeneous isotropic material free
of residual stresses. Firstly, a gear tooth prole has a continuously varying radius of curvature, and the importance of this departure from the
assumption may be emphasized by considering the case of an involute
tooth where the prole starts at the base circle. The radius of curvature,
R1, is at all times the length of the generating tangent, so at this point
it is, from a mathematical point of view, zero. However, it remains zero
for no nite length of the involute curve, growing rapidly with the
height of the tooth and having an unknown value within the base circle.
If contact were to occur at this point, the stress would not be innite,
as an innitely small distortion would cause the load to be shared by
the adjoining part of the involute prole, so that there would be a nite
area of contact. It is quite clear that Hertz analysis is rather inapplicable
at this point; all that can be said is that the stresses are likely to be
extremely high. In the regions where contact between well-designed gear
teeth does occur, the rate of change in R1 is much less rapid, and it is
not unreasonable to take a mean value at any instant for the short
length which is of interest here.
The assumption that the material is elastic will certainly break down
if the resulting shear stress exceeds the shear yield strength of the material. The consequences are quite beyond the ability to predict them mathematically. An analysis might be possible if one load application at one
instant were all to be dealt with; but the microscopic plastic ow that
would then occur would completely upset the calculations for contact
at the next point on the tooth prole, and so on. The situation when
the original contact reoccurred would be quite different; and there is a
need to deal with millions of load cycles as the gears revolve. All that

Machine Elements in Rolling Contact

241

can be said is that the repeated plastic ow is likely to lead to fatigue


failure, but that it will not necessarily be so, since the material may
perhaps build up a favourable system of residual stress, and will probably work harden to some extent. If such a process does go on, then
there is no longer a homogeneous isotropic material free of residual
stress.
Gears transmitting more than a nominal power must be lubricated.
The introduction of a lubricating lm between the contacting teeth
might be expected to alter the situation drastically, but it does not,
simply because the lubricant assumes the form of an extremely thin lm
of almost constant thickness.
In view of all these qualifying remarks, it is hardly to be expected
that the gear teeth can be designed on the basis of the maximum shear
stress being equal to 0.3pmax and having to equate to the shear strength
of the material in fatigue. Nevertheless, the Hertz analysis is of vital
qualitative value in indicating the parameter PR, which, for any given
material, can be taken as a criterion of the maximum stress, the actual
value to be allowed being determined experimentally.

7.1.2 Geometry of contact between gear teeth


In order to apply fundamental equations given by the Hertz theory to
the case of contact between a pair of gear teeth, it is necessary to dene
the reduced radius of curvature and the contact load for a specic gear.
Figure 7.1 shows, schematically, two involute gears in mesh. The
surface velocity at the point of contact C is expressed as
UG

U1CU2 1R1 sin C 2R2 sin


G
2
2

(7.3)

where R1 and R2 are the pitch circle radii of the driving and driven gear
respectively; is the angle of pressure dened as the acute angle
between the contact normal and the common tangent to the pitch
circles, and 1 and 2 are the angular velocities of the driving and
driven gear respectively.
Since
R1 2
G
R2 1
the contact surface velocity is therefore
UG 1R1 sin G 2R2 sin

(7.4)

242

Rolling Contacts

Fig. 7.1

Assuming that the total load is carried by one tooth only, then, referring
to Fig. 7.1, the contact load in terms of the torque exerted is given by
WG

T2
T2
G
h2 R2 cos

(7.5)

where W is the total load on the tooth; h2 is the distance from the centre
of the driven gear to the interception of the locus of the contact with
its base circle (=R2 cos ), and T2 is the torque exerted on the driven
gear.
The torque exerted on the driving and driven gear, expressed in terms
of the transmitted power, can be calculated from
T1 G

T2 G

G9.55

M
n1

M
M
G9.55
2
n2

(7.6)

(7.7)

where n1 and n2 are the rotational speeds of the driving and driven gear
respectively (rs) and M is the transmitted power (kW).

Machine Elements in Rolling Contact

243

Fig. 7.2

The line from C1 to C2 in Fig. 7.1 is the locus of the contact and it
can be seen that the distance S between the actual contact point of the
gear teeth and the pitch point P is continuously changing with the contact position during the meshing cycle of the gears. Thus, it is possible
to model any specic contact position on the tooth ank of an involute
gear by two rotating circular discs of radii R1 sin CS and R2 sin AS
as shown in Fig. 7.1. This idea is applied in a testing apparatus generally
known as a two-disc machine. Figure 7.2 shows, schematically, the
principle of the two-disc machine.

7.2 Friction in meshing gears


Attempts to measure the instantaneous coefcient of friction as a function of the meshing position have not been very successful, and at present it is necessary to rely mainly on disc tests for information on the
relations between the coefcient of friction and other factors. Tests utilizing the two-disc machine are designed to reproduce the state of affairs
at the point of contact between gear teeth, in a particular position of
the mesh but under conditions permitting continuous observation. Two
discs are run together in tangential contact with each other while a
normal load calculated to reproduce the stress conditions under consideration is applied. The circumferential velocities 1r1 and 2 r2 correspond with the rolling velocities of the teeth under the conditions
simulated (i.e. velocities of sweep of the point of contact), the sliding

244

Rolling Contacts

velocity Vs being given by


Vs G 1r1A 2 r2

(7.8)

where the disc with the higher rolling velocity is identied as disc 1.
The discs may be connected by gears or chain, the former reproducing the relative motion of the surfaces in actual gears; chain connection
produces a condition that is abnormal in spur and helical gears, though
it occurs on the inlet side of the teeth in worm gears.
The practical conditions represented by a disc test depend not only
on the surface velocities but also on the relative (reduced) radius of
curvature. Denoting this by r, it is dened by
1 1 1
G C
r r1 r2

(7.9)

and thus
rG

r1r2

(7.10)

r1Cr2

It should be remembered that r1 and r2 are the radii of equivalent circles


as shown in Fig. 7.1 and are, therefore, given by
r1 GR1 sin CS
r2 GR2 sin AS
(7.11)
The relative radius of curvature may be thought of as a measure of
the degree to which the opposing contours conform to each other; the
expression radius of conformity for this quantity is more informative,
but relative radius of curvature is in general usage. In involute teeth the
radii of curvature of course vary over the proles. The relative radius
of curvature also varies. At the pitch point it has a value given by
1
r

C
sin R R

(7.12)

where is the pressure angle and R1 and R2 are the pitch circle radii.
Qualitatively, there is a considerable measure of agreement between
the results of various tests that have been carried out by different investigators with gear connected discs on the relations between the coefcient of friction, , and other variables. The discs that have been used
in experiments have usually been given a nish similar to that typical

Machine Elements in Rolling Contact

245

of high-quality gears, and the results thus obtained are consistent with
the prevalence of quasi-hydrodynamic, or perhaps fully hydrodynamic,
lubrication at the zone of conjunction.
A comparative study leads to the conclusion that the principal correlatives of are the surface velocities, the relative radius of curvature,
and the representative or effective viscosity of the oil, more specically
its viscosity at the surface temperature at the inlet to the lubricating
lm. To these may be added disc or tooth loadings, though its effect in
this respect, like that of temperature, is secondary.
In order to describe the inuence of surface speeds on the coefcient
of friction, it has been found expedient to accord primary importance
to the entraining velocity, Ve , dened by
Ve G 1r1C 2 r2

(7.13)

i.e. the sum of the velocities of sweep. The two velocities of sweep over
mating proles on involute teeth vary continuously along the path of
contact. The sum of the two also varies, increasing in reduction gears
from the point of rst engagement to that of disengagement and similarly falling in speed-increasing gears. In a typical case the variation is
of the order of 2040 percent. At the pitch line the value is 2V sin ,
where V is the pitch line velocity and is the angle of pressure.
The most noticeable of the relations between the coefcient of disc
or tooth friction and other variables is an inverse dependence on Ve ,
attributable to its signicance as a measure of the inuence of surface
velocity on the entrainment of oil into the load area, and to its effect
on hydrodynamic relief of the contact load component, i.e. that portion
of the load borne by areas in direct contact.
The sliding velocity Vs has also been cited as another velocity factor
governing the friction coefcient . At a given viscosity, the traction
needed to shear the oil lm will increase with the rate of shear, and the
local rate of power loss will vary with the product of the sliding speed
and the shear stress exerted locally on the rubbing surfaces. The result
of an increase in sliding speeds is thus an increase in the temperature
of the metal and hence a reduction in the viscosity of the oil beyond
the lm inlet, and possibly an increase in the bulk temperature, resulting
in a reduction in the representative (inlet) viscosity itself. Recent work
suggests that the net effect of Vs is comparatively limited and that the
predominant velocity factor governing is Ve . There is also an inverse
relation between and the relative radius of curvature r. It seems that
this effect is again of the primary importance.

246

Rolling Contacts

The inverse relation between and the viscosity of the oil, , is well
established. The relevant viscosity, at the inlet to the zone of conjunction, is taken to be that of the bulk temperature of the discs or gear
teeth. The magnitude of the variation in with is the subject of some
disagreement, but to a large extent the reason for this appears to be
that in most of the earlier tests the oil viscosity was taken at the supply
temperature and the bulk temperature of the metal is unknown.
The variation in with tooth loading at constant temperature can be
predicted theoretically, but its magnitude is relatively slight and some
researchers have been unable to detect it. Tooth loading will, however,
be reected in the temperature at the metal surface, and hence in the
effective viscosity of the oil at the inlet to the zone of contact, and will
thereby indirectly inuence to some extent.
Hydrodynamic action in gears, discussed more fully later, is responsible for their successful operation, and it is interesting at this point to
touch upon the reasons for the inverse relation between coefcient of
friction and viscosity, an effect which is the opposite to that rst
expected from the theory of hydrodynamic lubrication as applied to
bearings. Since viscosity is the coefcient of proportionality between
the shear stress on the uid and its rate of shear, if the latter is xed it
follows, of course, that the resistance to motion set up by the uid, and
hence the power it absorbs, will vary with viscosity. However, where
lubrication in the zone of conjunction is only partially hydrodynamic,
some of the load is still borne by direct contact. Under such conditions,
the higher the inlet viscosity, the greater is the hydrodynamic relief of
the contact load component, and accordingly the smaller is the coefcient of friction. Under fully hydrodynamic conditions, an increase in
viscosity is accompanied in gears with a general increase in lm thickness; where an oil lm becomes thicker but the relative sliding velocity
between its boundary surfaces remains the same, the rate of shear is
thereby reduced. The effect of the increase in lm thickness with
increase in the viscosity of the oil would appear to predominate over
the corresponding increase in resistance to shear. Thus, on both scores
the coefcient of friction tends to vary inversely with viscosity.

7.2.1 Tooth losses


For any particular gearset, (the overall mean value of friction coefcient ) is essentially a measure of the transmission efciency of the
teeth. It can be derived by measuring tooth losses in the gearset concerned or in similar units.

Machine Elements in Rolling Contact

247

In estimating tooth losses, or comparing different gearsets or


operating conditions in this respect, empirical relations based on disc
test data are helpful as a guide to the values of to be adopted. A
number of empirical formulae have been put forward relating to
other variables. To some extent they conict with each other, and a
good deal of scope remains for further work in this eld. It would
appear, however, that over the usual range of values VsVe the following relation gives a reasonable approximation of the mean coefcient
of friction between the teeth of high-quality spur and helical gears

308

(7.14)

0.25 0.5

r V 0.5
e

where v is the viscosity (cSt) at the bulk temperature of the gears, r is the
relative radius of curvature at the pitch line (m), and Ve is the entraining
velocity at the pitch line (ms). For spur gears, another form of the
expression is
436

G
0.25

sin

(7.15)

dD

0.5

dCD

where is the pressure angle, d and D are the pitch circle diameters of
the gears (m), and V is the pitch line velocity (ms).
For helical gears the comparable expression can be taken to be
436

G
0.25

sin n

dD

cos0.5

(7.16)

0.5

dCD

where n is the normal pressure angle and is the helix angle. The
value of under any specic conditions of speed and geometry may,
of course, be considerably increased because of inaccuracy, including
misalignment, or wear of the gears.
Tooth losses expressed as efciency vary considerably with the size,
proportions, and speed of the gears. In the largest and fastest sets, for
example the reduction gears of marine turbine main engines, the tooth
loss with modern units may be as low as 0.2 percent per train. At the
other extreme, with spur and helical gears, the tooth loss in a small set
with rather coarse teeth may be as high as 1.5 percent. A typical value
in spur and helical gears of middling size is 0.7 percent per train. In

248

Rolling Contacts

other types of gear, notably worm gears, tooth losses may be very much
higher.

7.3 Outline of elastohydrodynamic theory


Under ideal conditions of hydrodynamic or uid lm lubrication the
working surfaces are completely separated from each other by a continuous lm of lubricant. A typical example of hydrodynamic lubrication in action is provided by plain journal or thrust bearings.
Normally, in such bearings the thickness of the lm at the loaded area
is of a distinctly higher order of magnitude than that of the height of
the asperities on the opposing surfaces.
For many years it was commonly considered that in heavily loaded
contacts, such as those in rolling bearings or between gear teeth, hydrodynamic lubrication was out of the question. This view rested on the
conclusion drawn from classical hydrodynamic theory that the lm
thickness between gear teeth that could be attributed to viscous
entrainment of oil was of a much lower order than that needed to separate the surfaces even under the most favourable conditions envisaged
in that theory. Contact between gear teeth was therefore classed as a
case of boundary lubrication, dened as a condition in which the load
on the engaging surfaces is wholly, or nearly wholly, borne by direct
contact. The view was commonly accepted that an attenuated lm of
oil remaining under such conditions owed its existence primarily to
physical or physicochemical effects at the metal surface. The lm was
sometimes even regarded as monomolecular.
Nevertheless, considerable practical evidence remained that in many
gears the thickness of the uid lm is impossibly large for boundary
conditions as thus dened, even where it does not separate the two
surfaces completely. In particular, the long life with negligible wear
often achieved with high-speed precision gears strongly suggested the
existence between the teeth of lubricating lms of substantial thickness
by comparison with molecular dimensions.
Classical hydrodynamic theory, as pioneered by Reynolds (2), rested
on various assumptions, notably that the lubricant is of constant viscosity in the contact, and that bearing surfaces are perfectly smooth
and rigid. From 1940 onwards a number of investigators attempted to
modify the classical theory for application to heavily loaded conjunctions. A general agreement was reached that the thickness of the predicted lm increased somewhat, by a factor of 2 or 3, when account
was taken of the increase in the viscosity of the lubricant under pressure.

Machine Elements in Rolling Contact

249

This theoretical renement thus effected some improvement, but was


still far from sufcient to account for the apparent existence of full, or
nearly full, uid lms in practice. Attention turned also to the effect of
elastic deformation of the contacting surfaces. Prediction of the
operating lm prole on this basis involves determination of the local
deformation of the contours of the two rubbing surfaces, providing
compatibility between the elastic effects in the metal and the hydrodynamic effects in the uid.
A theoretical solution that has proved to be essentially sound was
published in 1949 by Grubin and Vinogradova (3) and by Petrusevich
(4). According to their analyses, the surfaces confronting each other at
the loaded area are almost parallel along the greater part of the contact
zone, and the corresponding stress distribution is approximately Hertzian. At the outlet there is a pronounced constriction, which is often
associated with a sharp and highly localized pressure peak. Figure 7.3
shows (a) the lm prole and (b) the pressure distribution indicated in
references (3) and (4). The analyses also predicted that the thickness of
the lm would be largely unaffected by load, given the effective oil
viscosity.
In 1958, Crook (5) reported the direct measurement of the oil lm
between the peripheries of the discs of a disc machine. He found that
under heavy load the lm thickness was of the same order as that predicted by Grubin and Vinogradova and was, in fact, highly insensitive
to changes in load.
The joint action of the elasticity of the surfaces and the increase in
viscosity of the lubricant under pressure affords a striking example of
synergy, i.e. the situation in which the whole is greater than the sum of
its parts.
The effect of the compressibility of the lubricant has also been taken
into account by Dowson and Higginson (6). According to this study,
the load capacity, or the lm thickness at a given load, is not signicantly affected by oil compressibility, but the shapes of the contact and
of the pressure diagram are more rounded. The predicted shape of the
contact zone is almost exactly as found experimentally by Crook.
Because viscosity varies greatly with pressure and temperature, it has
been a problem to decide on the value of the viscosity to be used in
lubrication calculations. It is acceptable to take as the representative
viscosity that existing at the conditions of pressure (i.e. atmospheric
pressure) and temperature of the surfaces before they enter the zone of
conjunction. The reason for this is twofold: in the rst place the oil
lms on the contacting surfaces are only comparatively slowly displaced

250

Rolling Contacts

Fig. 7.3

by new oil coming from the bath or other source of supply; in the
second place, because there is negligible side leakage from the zone
of conjunction in the direction perpendicular to that of the peripheral
velocities, the thickness of the lm there is decided by the volume of oil

Machine Elements in Rolling Contact

251

present and not directly by its viscosity, or variations in viscosity, whatever the case may be, as a result of the local increases in pressure and
temperature.

7.3.1 Estimates of lm thickness


Grubin and Vinogradova (3) included the effects of elasticity and of
variation in viscosity under pressure, although they introduced certain
approximate assumptions. Their formula, arranged in non-dimensional
groups, can be expressed as
0.091

w
h
G1.17
r
Er

( E)0.73

oVe
Er

0.73

(7.17)

where h is the thickness of the nearly parallel part of the lm, r is the
relative radius of curvature of the contacting surfaces, given by 1r=
1r1C1r2 (r1 and r2 being the radii of curvature of two contacting
surfaces), w is the load per unit length of contact, E is the reduced
Youngs modulus for the materials of the contacting bodies, o is the
viscosity of the lubricant at atmospheric pressure and at the bulk temperature of the surfaces before contact, and is the pressure coefcient
of viscosity of the lubricant.
The reduced Youngs modulus can be found from
1

1 1A 21 1A 22
C
2 E1
E2

where is Poissons ratio.


The pressure coefcient of viscosity is given by the formula

o
dp
p

where p is the viscosity at pressure p. Under the elastohydrodynamic


conditions for which the GrubinVinogradova formula applies, p can
be assumed to vary exponentially with p according to

p G o e p
Also, Ve Gv1Cv2, where v1 and v2 are the velocities of the moving
surfaces.
DowsonHigginson analysis of elastohydrodynamic lubrication
produced a formula for minimum lm thickness very similar to the

252

Rolling Contacts

GrubinVinogradova formula. Its form is as follows

h min
w
G0.96
r
Er

0.13

( E)

0.6

oVe
Er

0.7

(7.18)

The formula may be more conveniently written as


IG0.96

G 0.6M 0.7
J 0.13

(7.19)

where I, J, G, and M represent the four non-dimensional groups


respectively.
The results from the DowsonHigginson formula and the Grubin
Vinogradova formula are said never to differ by more than 20 percent.
The formula can be further simplied, and quite drastically so, without
serious error. Experience shows that, in practice, h min is virtually independent of E since the power to which it is raised is low and the range of
variation for engineering materials such as steel and bronze is relatively
limited. Furthermore, the value of does not vary much for conventional lubricants, so that G 0.6 will remain fairly constant. Also, h min is
only slightly dependent on the load parameter J. Dowson and Higginson therefore concluded that, to a reasonable degree of accuracy, G 0.6
and J 0.13 could be replaced by constants, so that the minimum lm
thickness h min could be simply calculated as follows
h min G3.51( oVe r)

(7.20)

where h min is in micrometres, o in poises, Ve in centimetres per second,


and r in centimetres.
In the application of elastohydrodynamic theory to gears, the question arises as to the effects at the ends of the zone of contact. To the
extent that elastic deformation under load might decrease somewhat at
and near these positions, the oil lm there might be slightly reduced in
thickness.

7.4 Application of elastohydrodynamic theory to


gears
The above theories assume constant temperature conditions, but a useful guide to the interpretation of practical experience can be given in
terms of elastohydrodynamic theory, together with the ash temperature theory.

Machine Elements in Rolling Contact

253

7.4.1 Film thickness between gear teeth


Even in the many cases where complete uid lm lubrication is unattainable and some degree of actual contact has to be accepted, hydrodynamic relief of the contact load component remains important to
the preservation of the gears. As indicated earlier, the lm thickness is
governed primarily by the viscosity of the oil at the temperature of inlet
to the zone of conjunction, by the entraining velocity, and by the
relative radius of curvature.
The solution to the elastohydrodynamic problem achieved by Grubin
and Vinogradova and Petrusevich and subsequently rened by others
refers strictly to isothermal conditions. Such conditions characterize
pure rolling, in which frictional losses and non-uniform lm heating are
quite slight, and low sliding speeds, when such losses and heating are
still slight. When there is a considerable amount of sliding, the rate of
generation of heat in the zone of conjunction will be much increased,
but it does not appear that the lm thickness will be directly affected,
since the viscosity that is in this respect governing remains that at the
inlet. It has been suggested that a forward ow of heat takes place from
the conjunction, with the effect that the viscosity of the incoming oil is
reduced. If this were so, the effect would be to reduce the lm thickness
to some extent. The reduction would be least at or near the pitch line
and would rise to maxima at or near the points of rst engagement and
of disengagement. However, the grounds on which such a ow of heat
has been inferred are not very convincing, and there is uncertainty as
to its rate. According to ash temperature theory, which is discussed
later, forward ow of heat from the conjunction would be negligible at
any normal velocity of sweep. Moreover, the lm thickness between
rollers in slidingrolling contact has been found to be substantially as
predicted by isothermal elastohydrodynamic theory over a wide range
of conditions of speed, slidesweep ratio, and load.
At the pitch line in spur gears, the entraining velocity Ve (ms) has a
value of 0.08V sin , where V is the pitch line velocity (ms). For spur
gears the radii of curvature r1 and r2 at the pitch line are
r1 G
r2 G

d sin
2
D sin
2
(7.21)

where d and D are the pitch circle diameters and is the pressure angle.

254

Rolling Contacts

The relative radius of curavature r at the pitch line is thus [see also
equation (7.12)]
rG

sin
2

dCD
dD

(7.22)

Assuming that the ratio of wheel to pinion diameters DdG results in


rG

d sin
2

C1

(7.23)

Comparing gears with the same centre distance but differing in ratio ,
i.e. reduction ratio, with a given speed and viscosity, the lm thickness
at the pitch point will be less the higher the ratio. Beyond about 2 or
3, however, the effect of will be limited.
For comparative purposes, the lm thickness in spur gears at the
pitch line can thus be taken to be
h min G0.7 sin

o dV

C1

(7.24)

where h min is in micrometres, o is in poises, d is in metres, and V is in


metres per second. The viscosity o is that at atmospheric pressure and
at the surface temperature of the teeth. For mineral oils, and for synthetic oils with specic gravities about equal to those of mineral oil, the
formula becomes
h min G0.065 sin

vo dV

C1

(7.25)

where vo is the viscosity at atmospheric pressure and at the surface


temperature of the teeth (cSt). It is important to note the effect of pressure angle on lm thickness.
In helical gears the relative radius of curvature appropriate to the
calculation of Hertzian stress is that in the pitch ellipse of the virtual
spur gear and has the value
rG

d sin n
2
2 cos C1

(7.26)

where n is the normal pressure angle and is the helix angle. At the
pitch line, the corresponding value of the entraining velocity (ms) can

Machine Elements in Rolling Contact

255

be taken to be
Ve G0.08V sin n cos
On this basis, the foregoing relation for spur gears becomes
h min G0.065

sin n
0.5

cos

vo dV

C1

(7.27)

where h min is measured in the normal direction. The factor cos0.5 has
the value 0.931 for G30 and 0.983 for G15. An approximate
expression for h min in helical gears at the pitch line is thus
h min G0.07 sin n

vo dV

C1

(7.28)

As before, h min is in micrometres, vo is in centistokes, d is in metres,


and V is in metres per second. In the transverse direction, the thickness
of the lm is, of course, h min sec , while
sin t Gsin n sec
where t is the transverse pressure angle.
Assuming that isothermal elastohydrodynamic theory is valid for all
positions of contact between the teeth, it is to be expected that the lm
thickness will vary over the active proles of the teeth with the factor
0.43
V 0.7
, or approximately with 1(Ve r).
e r
The relative radius of curvature of involute proles in mesh varies
parabolically over the common tangent of the base circles. Figure 7.4
shows this effect. The interference points I1 and I2 are the points of
tangency of the common tangent and the base circles, centres C1 and
C2. At I1 and I2 the relative radius of curvature r is zero. At the point
mid-way between I1 and I2 the relative radius r has its maximum value.
The effect of r on the thickness of the oil lm between the teeth is
indicated by the lower curve, which on some scale represents r0.43. This
curve has a comparatively at top and falls off sharply near the interference points.
The entraining velocity Ve , the sum of the velocities of sweep, is constant over the line I1I2 in gears of 1 :1 ratio, but in all other cases it
varies from I1 to I2. The variation is linear, increasing from I1 to I2 in
the simple relation that the ratio of Ve at I2 to Ve at I1 is equal to the
gear ratio. Thus, in speed-reducing gears Ve increases continuously over
the path of contact, and in speed-increasing gears it falls continuously.
Figure 7.4 shows the variation in Ve with a gear ratio of 2 :1. If it is

256

Rolling Contacts

Fig. 7.4

taken that the thickness of the oil lm will vary with V 0.7
e , the effect
will be as indicated in the diagram.
In a particular pair of meshing gears the magnitude of the variation
in lm thickness due to variation in r and Ve along the path of contact
will depend on the position and extent of the section of the common
tangent between I1 and I2 occupied by the path of contact. Figure 7.5
shows the combined effect of the variation over I1I2 of r and Ve on the
foregoing basis, in terms of the ratio of the lm thickness at any point
to that at the pitch point. As the gear ratio increases, the pitch point,
denoted by P in Fig. 7.4, moves towards I1. Except in gears of very low
ratio, the effect insofar as Ve and r are concerned is that in general the
lm thickness will increase continuously over the active prole in speedreducing gears and will decrease continuously in speed-increasing gears.
In general, the variation is greater the higher the gear ratio. Figure 7.5

Machine Elements in Rolling Contact

257

Fig. 7.5

represents either a spur gear or a transverse section through a helical


gear. In the latter case the base circles to be considered are those on
which the involute helicoids are generated, and not the base circles of
the virtual spur gears.
The length of the path of contact in proportion to I1I2 will depend,
at a given pressure angle, on the diametral pitch of the teeth. Figure 7.6
shows the approximate variation in the lm thickness over the path of
contact of standard full depth teeth with a 20 pressure angle, in terms
of the percentage of the pitch point value at or near the point of rst
engagement and at or near that of disengagement. The lm thickness is
shown increasing from I1 to I2 , i.e. the condition of reduction gears.
Increase in the pinion addendum, and decrease in that of the wheel,
moves the point of rst engagement nearer to the pitch point, and
reduces the extent to which the lm thickness at the point of rst
engagement falls short of that at the pitch line.

7.4.2 Operating temperature


The signicant operating temperature with respect to lm thickness,
i.e. that at which oil enters the zone of contact, is identied with the
temperature of the metal surface prior to meshing. The viscosity
temperature dependence of lubricating oils is well known, but it is

258

Rolling Contacts

Fig. 7.6

relevant here to note its magnitude. Over the range of conditions encountered in gear lubrication, the viscosity of a mineral gear oil is
halved by an increase in temperature of approximately 15 C, the corresponding effect on lm thickness being a reduction of about one-third.
The temperature of the teeth is thus of primary importance: the lower
the temperature, the thicker is the lm of oil between the teeth. The
viscosity grade of the oil is also important, but it should be noted that
the use of a thicker oil, while reducing tooth friction, increases bearing
friction and churning losses, and thus normally leads to a general
increase in operating temperatures which partially offsets the viscosity
increase at the surface of the teeth.
Comparing geometrically similar gears of the same size, operating
temperatures generally increase with speed and at any given speed they
increase with the viscosity grade of the oil. With bath lubricated gears
it is a reasonable assumption that the bulk temperature of the teeth will
not differ much from the temperature of the oil charge. With spray
lubricated gears, however, there is an element of uncertainty as to the
temperature of the teeth. It is to be expected that it will vary in relation
to the oil inlet temperature according to the amount of oil directed

Machine Elements in Rolling Contact

259

on to the teeth and the effectiveness of application of the sprays. The


temperature of the oil rebounding from the gears would give an indication of the surface temperature, but it is difcult to measure this
reliably and in any case some doubt would remain as to the extent to
which the temperature of the gears would exceed that of the oil. Some
differential is inevitable since the period of contact of the bulk of the
oil with the metal surfaces is very short indeed.
Measurements of the operating temperature of the teeth in spray
lubricated industrial gears led to the conclusion that with good design
the margin over the temperature of the oil leaving the teeth is commonly
in the range 510 C and does not exceed 20 C in the extreme cases.
The tooth operating temperatures depend on the tooth losses in relation
to the efciency of cooling. Both of these factors may vary considerably.
The tooth losses tend to increase with tooth loading and sliding speeds,
but in practice designed loads are reduced and the teeth are made ner
in pitch as peripheral speeds increase. Generally speaking, the operating
temperature of the teeth does not show so much dependence on peripheral speed where lubrication is by spray.
As a rule, the oil inlet temperature should not be more than about
10 C below the estimated tooth temperature, so that the oil will not be
of too high viscosity for effective spraying over the gears. In large units
the inlet temperature is often about 32 C.
Typically, the temperature peak in the metal bounding the conjunction (the ash temperature) may, where sliding velocities are greatest
(i.e. at the tips of the teeth), be perhaps 6595 C over the bulk temperature of the teeth outside the zone of contact. A temperature rise of 20
percent of the peak in the constriction has been suggested for the oil at
entry, but the extent to which the oil temperature at entry is increased
by forward ow of heat from the conjunction, if at all, is highly
conjectural.

7.4.3 Oil viscosity in relation to surface condition


The selection of the viscosity grade for gears is based on the general
necessity for the viscosity grade of the oil, i.e. its viscosity at a reference
temperature, to vary inversely with speed. If the viscosity grade typical
of gears at a given speed is adopted for comparative purposes and the
corresponding typical temperature of operation is taken into account,
the characteristic value of lm thickness in gears of given form shows
the trend in relation to speed shown in Fig. 7.7. This has been drawn
on the basis of pitch line conditions for a hypothetical spur gearset of
standard design assuming a pinion diameter of 210 mm and a ratio of

260

Rolling Contacts

Fig. 7.7

diameters of 2.6. It is interesting, though perhaps not surprising, to note


that over a wide range of speed the average lm thickness is uniform.
At low speeds it increases, and the reason for this could be that surface
nish is not so good in slow running gears, with the effect that experience has led to the general use of oils of higher viscosity than would be
expected by comparison with faster running gears. The values indicated
in the diagram are those that would be obtained between geometrically
true surfaces.
In selecting the oil for a given gearset, distinction must be made
between conditions during running-in and those when a smooth surface
has been achieved. It would be reasonable to assume that for freedom
from wear the minimum lm thickness indicated by elastohydrodynamic theory should exceed the sum of the peak-to-valley heights of
the asperities on the two surfaces, i.e. about 34 times the sum of the
two centre-line average (c.l.a.) values. It appears, however, that freedom
from wear is obtained with lm thicknesses less than this, as though
some depth of oil, in the valleys between asperities, acts as a foundation
for the hydrodynamically induced lm. Experience suggests that, as
indicated by the theory, a lm thickness of about twice the sum of
the c.l.a. values will ensure freedom from wear. This condition may be
achieved at the outset with some precision quality gears operating at
high speed, but in general the initial roughness of the surfaces will
exceed the value at which contact will occur, and the establishment of
a predominantly hydrodynamic regime will depend on running-in.
A practical question concerning lm thickness is that of where in
relation to the asperities it is presumed that the datum exists. Measurements of lm thickness, using a voltage discharge method, indicate

Machine Elements in Rolling Contact

261

variation in lm thickness between pitch line and tips in gears of 1 :1


ratio, shown to be of the order of 5 :1. However, the base here is the
average position of zero voltage, which will occur where contact takes
place between the most prominent asperities. If the lm thickness is
assessed from the estimated position of the c.l.a. datum, the variation
between pitch line and tips of the hydrodynamically induced lm
appears to be about 2 :1 or less. Variation in lm thickness of this order
exceeds that attributable to variation in the relative radius of curvature
and has been cited as evidence of forward ow of heat from the constriction. If, however, the lm thickness is dened as the effective clearance between the opposing asperities, it is arguable that the value of
this at a particular position may, on average, vary with the rate at
which opposing asperities are traversing that position; that is to say,
the chances of an encounter or near-encounter between asperities over
average size may increase with the local amount of sliding. Finely divided solid matter such as metallic debris of extremely small particle size
in entrainment in the oil may similarly bridge the oil lm with greater
frequency the greater the sliding velocity. It is therefore possible that the
elastohydrodynamic lm thickness, assessed from, say, the c.l.a. datum,
could be much the same at the tips of the teeth as at the pitch line, notwithstanding that the effective lm thickness or clearance diminishes
from the pitch line to the tips. If this is so, it is to be expected that the
variation in lm thickness will diminish with running-in, but the scales
will always be weighted against the lm thickness at the tips of the teeth.

7.5 Boundary contact in gear lubrication


Full uid-lm lubrication implies that the surfaces are completely separated from each other, and partial uid-lm lubrication that the normal
load is partly borne by areas of oil under pressure and for the remainder
by limited areas of direct contact between opposing asperities separated
only by adsorbed lms of lubricant. The amount of such boundary
contact can be referred to in terms of the proportion of the load so
borne or in terms of the frequency of opposition or engagement of
asperities. It is of note that, as sliding speeds vary, this amount will also
vary with the amount of sliding, which is why sliding velocities are of
more consequence under conditions of boundary lubrication than in
uid-lm lubrication.
For little or no contact between the surfaces, the calculated nominal
lm thickness must exceed some critical value in relation to the roughness of the asperities. In practice there are very large variations in

262

Rolling Contacts

surface roughness. Such variations are caused both by machining and


wear, including running-in, and there is little doubt that many anomalies in gear performance can be traced to this source.
Specications for spur and helical gears do not usually include stipulations as to surface nish. Requirements as to surface accuracy imply
particular conditions of surface nish, and in this respect there is less variation among gears of precision quality than among less accurate gears.
Given that the oil lm thickness indicated by elastohydrodynamic theory
will have an order of magnitude of 0.751.25B103 mm, it is clear that in
the large majority of gears much contact between asperities will be
inevitable when the gears are rst run. In gears of ordinary commercial
quality, such contact will initially occur on a gross scale, while the load
between the teeth is inevitably carried largely by opposing asperities.

7.5.1 Running-in process


The process of running-in, particularly with comparatively ductile
steels, consists primarily of the progressive smoothing of the surfaces
through a combination of ploughing, abrasive wear, and plastic ow.
With the more ductile steels the process is very rapid at the outset, with
comparatively high rates of wear accompanied with rapid smoothing of
the surface. As the surfaces become smoother, the rate of wear diminishes, falling swiftly to a low value in a matter of days. Subsequently,
slow wear persists over a period of weeks or months, usually tailing off
and in many cases falling to a negligibly low rate. This means that in
general a state of near-hydrodynamic or marginal hydrodynamic lubrication will then have been reached. However, nothing much better than
that marginal state can be achieved, even after prolonged running.
Generally speaking, the harder the surface material of the gear teeth,
the greater is their resistance to running-in. Ready running-in, leading
to a highly polished surface condition, is characteristic of the steels of
low to medium hardness widely used in industrial gears. Harder tooth
surfaces may similarly undergo a progressive improvement in smoothness, though the process tends to take longer the harder the surface
material and may occupy several years. Thus, with these harder gears
the risks typical of a running-in period tend to cover a much longer
period. Such teeth may well be rather sensitive towards the choice of
lubricant. With hard teeth many anomalies have been noticed, and there
is little doubt that to a considerable extent these are due to the inuence
of the material of the mating gear. If one gear is considerably harder
than the other, the course of events tends to be governed by the surface
condition of the harder of the two. A hard pinion with a surface produced by prole grinding may, for example, have a nish superior to

Machine Elements in Rolling Contact

263

that of the mating wheel where the latter goes into service as hobbed,
and the effect will be a smoother nish, or it will tend to plough and
roughen the wheel. On the other hand, if the disparity in hardness is
large, the pinion itself may not run in very well, or may hardly run in
at all, with the result that real damage is done to the tooth faces of the
wheel. As a rule, however, where both the gears have a surface hardness
only slightly below the limit of machinability with normal cutting tools,
running-in generally follows the same course as with the softer materials, though more slowly. Failure of the gears to run in well is associated
mainly with the use of case carburizing or nitriding to obtain a fully
hardened surface, particularly where the mating gear is not so hard.
It is sometimes said that a certain degree of microroughness of the
working surfaces of the teeth is advantageous, presumably because trapping of oil between asperities is supposed to help to maintain the oil
lm. There are indications that at increasing speeds the shear elasticity
of the oil comes into play, i.e. that its behaviour begins to be to some
extent that of an entrapped elastic solid. However, there is no evidence
that there is an optimum degree of surface roughness within the range
attainable in practice. The belief that the surface can have too high a
polish is possibly due to the observation that breakdown of lubrication
may be preceded by burnishing to a mirror-like nish. It may also be
due to the fact that the harder teeth tend to receive the best initial nish,
and that these have a prolonged and hazardous running-in period.
During the early stages of running-in, gears are particularly vulnerable to scufng. As running-in continues, the danger of scufng
becomes less acute, but it persists until the process has reached an
advanced stage say for 48 or 72 h running under load with comparatively soft gears, but for months or even years with hard gears. This is
especially so if the gears are subjected to overload torque during these
stages. Particularly with the harder tooth surfaces, it is often advantageous, and may be essential, to use an oil embodying an effective
extreme-pressure (EP) additive during the running-in period. When the
running-in process is completed it is usually not essential to retain an
EP oil, but with hardened gears it may be 2 or 3 years before this stage
is reached.

7.6 Scufng in meshing gears


7.6.1 Flash temperature as a criterion for scufng
Scufng is the severe surface damage attributed to local welding of
microasperities on the mating surfaces followed by rupture of the
welded microjunctions. It is apparently due to gross breakdown of

264

Rolling Contacts

lubrication over a distinctly dened area, which is likely to be badly


damaged. Usually, scufng is progressive and its development fairly
slow, although it can develop at a catastrophic rate. It normally begins,
or is more severe, when sliding speeds are highest, i.e. at the root and
tip of the teeth, and will never occur at the pitch line, where rolling
dominates and there is little or no sliding.
For scufng to occur, contact between the surfaces is essential. However, this is far from being the only requirement. Although it has been
noted that a considerable amount of actual metalmetal contact takes
place in practically all gears during the early stages of operation, nevertheless scufng is normally avoided.
It has been established that whether or not contact gives rise to
scufng can be related to the local increase in temperature in the adjacent metal as the zone of contact passes through the particular position
considered. A reasonable estimate of this increase, or ash temperature,
is given by the following equation derived by Blok (7). The transient
contact temperature Tc is dened by
Tc GTbCTf

(7.29)

where Tb is the bulk temperature of the teeth and Tf is the ash temperature superimposed on Tb. For spur gears, Tf is given by
Tf G1.11

w
1U1A1U2
b 1z

(7.30)

where is the instantaneous coefcient of friction for the contact area,


bG(s)0.5 denotes the thermal contact coefcient for the tooth face
material, in which is thermal conductivity, is density, s is the specic
heat per unit mass, w is the tooth normal load per unit length of contact
in the meshing position considered (the actual tooth load should be
used, allowing for tooth load sharing and the dynamic load increment),
z is the Hertzian contact band width, and U1 and U2 denote tangential
speeds of teeth 1 and 2 perpendicular to the line of action for the meshing position considered. The formula for Tf is valid for any consistent
system of units.
The meshing position considered has commonly been the lowest
point of single-tooth contact. It is advisable, however, to calculate Tf
for a number of positions on the prole to nd the highest value. Only
in a minority of gears is the lowest point of single-tooth contact likely
to be the least favourable position. Unless there is some tip relief, either
as manufactured or after running-in, it may be preferable for general

Machine Elements in Rolling Contact

265

comparative purposes to take the value Tf at the position of maximum


sliding velocity, i.e. the position of initial contact at the dedendum of
the driving tooth and the addendum of the driven tooth. Practice has
varied as to the load assumed. In the position of transition to singletooth contact, the whole tangential load, together with the dynamic
load increment, is borne by the tooth; in positions where two teeth are
in contact, some proportion of the tangential load upwards of one-half
is assumed, depending on the accuracy and conditions of operation of
the gears, but owing to dynamic loading this may exceed unity. Where
more than two teeth can engage at once it is advisable to assume for
this purpose that only two of them are load bearing at any instant.
For helical gears, Tf is given by
Tf G

w
1U1A1U2
1(sin n csc t) b 1z
1.11

(7.31)

Since a number of teeth will be in contact at the same time, there is some
doubt as to the value of w to be adopted. For comparative purposes it
is assumed that the load is equitably distributed at any moment over
the total length of line of contact.
The appropriate value of in a particular case is likely to be difcult
to decide. For the purpose of estimating the maximum ash temperature at which a thermally stable system can exist between the meshing
gear teeth, the appropriate value is that under conditions approaching
those of incipient scufng, as might be indicated in a rig test by the
beginning of a runaway increase in the bulk temperature of the gears.
For some time a value of 0.06 for was thought appropriate in this
context, and more recently a range of 0.070.09 has been put forward
as more representative.
It cannot be said with certainty that the scufng temperature is a
property of the oil, in the sense of it being a unique characteristic of
the grade concerned, with the effect that in any situation where this
value is attained the dynamic system in which the oil is taking part will
undergo thermal collapse. There is room for speculation that the volatility of the oil, or perhaps other intrinsic properties, imposes such a
limit, but there is not at present clear evidence of this. It is also possible
that the limiting ash temperature with a particular oil may vary
between gears differing in materials and surface nish. The value in a
practical case may be to some extent higher or lower than that referring
to laboratory gears used in scufng tests.
In estimating the ash temperature occurring in practice in a gear set,
the relevant value of will be that under the least favourable operating

266

Rolling Contacts

conditions, including the effects of dynamic load increment. If, for


example, the expected average value over the meshing cycle is 0.03, it
may be appropriate to assume a value of 0.070.09 to take account of
variation in the instantaneous value and dynamic loading effects.

7.6.2 Phenomenon of scufng


The term scufng is used for the more rapidly damaging forms of wear
due to oil-lm breakdown under high load, characterized by welding
and tearing and consequent roughening of the tooth surfaces. In many
cases where gears are subjected to severe loads, the danger of scufng
can be avoided by the use of an extreme pressure additive.
Scufng varies in degree, being referred to as slight, moderate, heavy,
and so on; the more severe cases show more obvious signs of metallic
tearing. Severe scufng, in which metal may be dragged radially over
the tips of the teeth, is sometimes called galling.
Scufng results from asperity contact under conditions conducive to
disruption of the oil lm. Such conditions are characterized by the generation of heat at high rates at the points of contact and necessarily
involve comparatively high sliding speeds, or a combination of high
sliding speeds and loads, where contact occurs. The sliding speeds
between gear teeth at a given surface temperature are seen to be rather
immaterial to the thickness of the oil lm at the point of contact, and
the lm thickness is also virtually independent of tooth loading. However, because of the inuence of sliding speeds and load on surface
temperatures, both of these quantities are subject to practical limitations for the avoidance of scufng. The ash temperature theory discussed earlier is also applicable to full uid-lm lubrication, since the
product w may be replaced by uid frictional force. The load w would
still be inuenced via the width z of the Hertzian area.
Although the frictional force is still affected by sliding speed and load
under elastohydrodynamic conditions, the inuence of these factors
increases comparatively greatly with the amount of direct contact, i.e.
with the relative magnitude of the contact load component. Generation
of heat thus increases with the frequency and severity of collisions
between asperities. Scufng will result from exceptionally violent
engagement of asperities where the oil lm is, or becomes, too thin in
relation to their height.
The inception of scufng implies a situation both thermally and
mechanically unstable, and its development tends to snowball progressively. More often than not it spreads very rapidly. When scufng
occurs, it very often starts at the tips and roots of the teeth, where the

Machine Elements in Rolling Contact

267

conditions of load and sliding speed are generally worst. Sometimes it


starts at the points where tip relief ends or, in small spur gears, at the
beginning or end of single-tooth contact. It tends to spread towards the
pitch line, sometimes leaving a band there which remains unscuffed but
is overstressed and consequently succumbs to progressive pitting. The
designer should avoid unduly severe combinations of load and sliding
in relation to pitch line speed. The pitch of the teeth and amount of
addendum are important in this respect; large teeth, though stronger,
have higher maximum sliding speeds, while increase in the addendum
also increases sliding speeds. At a given peripheral velocity the tendency
to scuff varies with the size of the teeth; gears with large numbers of
small-pitch teeth are less likely to scuff than comparable gears with a
small number of teeth of coarse pitch. This is because the smaller teeth
are weaker in regard to bending fatigue and therefore the loads to which
they are subjected are limited to lower values, and at the same time the
maximum sliding velocities, at the tips and roots, are lower the smaller
the teeth.
Generally speaking, slow-running gears have little tendency to scuff,
since the sliding speeds between the teeth are as a rule too slow; metal
metal contact tends to result in relatively slow abrasive wear rather than
scufng. As speeds increase, metallic contact is increasingly liable to
result in scufng, while running-in becomes more critical, and to counter this tendency increasingly high standards of nish and accuracy are
required. At the same time, the hydrodynamic effect of increase in speed
is to tend to increase the oil-lm thickness. Thus, while metallic contact
becomes more dangerous, the amount of such contact tends to diminish. In practice, the former effect appears to be predominant in mediumspeed gears and in some high-speed gears up to pitch line speeds of
about 25 ms. At higher speeds, although torque or load must be
reduced, the net effect with regard to power transmissible is a net
increase in the margin of safety against scufng.
The choice of gear materials and their microstructure and hardness,
as well as surface nish, are important to the scufng resistance of the
gears, some material combinations being more prone to scuff than
others. It should be borne in mind, however, that what is being considered is the result of an interaction between the metal surfaces and
the lubricant; in some cases the inuence of the foregoing variables on
scufng resistance may be modied to some extent by changes in the
lubricant. This applies particularly to changing from an inactive oil to
an oil with extreme-pressure additives or vice versa. With spur and
helical gears it is of note that a high content of nickel, or of nickel

268

Rolling Contacts

and chromium in combination, appears to reduce scufng resistance,


particularly where these materials are running together, while direct
hardening steels free from nickel show very good scufng resistance.
Case-hardened and nitrided teeth are also highly resistant to scufng.
However, there are numerous anomalies in the relation between hardness and resistance to scufng. The harder teeth are more resistant to
running-in, and the gears are thereby exposed to a greater and more
protracted danger of scufng; extreme-pressure oils, or oils of higher
viscosity grade, or both, may then be advisable.
Any factor causing local or general overloading while sliding speeds
are high enough may cause scufng. Overloading may stem from many
sources, including magnication of dynamic load increments as a result
of abrasive wear, or reduction in the load-carrying area through pitting.
Damage to the working surfaces of any kind will tend to increase local
stresses and is capable of leading ultimately to scufng. Discontinuities
such as large pits also tend to cause an extended interruption of the oil
lm. Local overloading may be caused by inadequate tip relief, misalignment, or thermal distortion.
Similarly, any factor causing a local or general reduction in the oil
lm thickness may lead to scufng. This included, notably, increase
in the operating temperature of gears. In gearsets lubricated by spray,
inadequate supply of oil to the working face, or inefcient distribution
of the oil, will result in local or general overheating and consequent
reduction in the oil viscosity. Oil starvation of the tooth face will also
be liable to lead to dry running. Conversely, an increase in the oil supply
is generally advisable where scufng threatens.
In addition, the scufng resistance of the oil is to be considered. The
mechanism of oil-lm breakdown may vary between different lubricants, and in some cases instability or dissociation at high temperatures
may be involved. In mineral-based lubricants, resistance to scufng
depends on the viscosity grade and on the additives that may be contained in the oil. The hydrodynamically induced lm thickness will
increase with the effective viscosity of the oil. It should be noted, however, that, under conditions of partial hydrodynamic lubrication, a
heavy oil showing a certain viscosity at a given operating temperature
will give better protection against scufng than a thinner oil at some
lower temperature at which its service viscosity is the same.

7.6.3 Probability of scufng


It is permissible to say that scufng is controlled by a number of factors
such as load, lubrication, sliding speed between contacting bodies, surface topography, and temperature within the contact region. Although

Machine Elements in Rolling Contact

269

a number of deterministic models and scufng hypotheses have been


put forward (8, 9), it seems that the only rational approach in predicting
the potential for the onset of scufng is in probabilistic terms.
Lubrication regimes and thermal effect parameters
As a rst step in the analysis it is necessary to nd out if the contact
conditions are those dened by Martin (10). This is done by examining
the values of the viscosity parameter gv and the elasticity parameter ge
(11).
gv G

w
Re

w
ERe

(7.32)

oV

0.5

ge G

w
oV

(7.33)

where is the pressureviscosity coefcient, w is the normal load per


unit length of contact, o is the absolute viscosity, and
1

1A 21 1A 21
C
E1
E2

1
1
1
G C
Re R1 R2
VG

V1CV2
2

Hydrodynamic lubrication is possible when gvF1.5 and geF0.6.


Operating conditions outside the limitations for gv and ge are usually
dened as elastohydrodynamic (EHD) lubrication. Practical EHD
lubrication is limited by the values of another pair of parameters:
speed parameter
gs G

E 3 oV
Re

0.25

(7.34)

and load parameter


gl G

wE

2Re

(7.35)

270

Rolling Contacts

For EHD lubrication 1.8FgsF100 and 1FglF100. Equations (7.32) to


(7.35) help to establish whether or not the contact is in the hydrodynamic or EHD lubrication regime.
Thickness of lubricant lm
For the hydrodynamic lubrication regime, the minimum lm thickness,
ho , in the case of smooth surfaces can be calculated (10) from
ho G4.9

oVRe
w

(7.36)

where 4.9 is a constant for a rigid solid and an isoviscous lubricant.


Moreover, equation (7.36) is valid for high velocities and low loads on
the contact.
In elastohydrodynamic contacts the oil-lm thickness is nearly
constant throughout the Hertzian zone. For line contacts of smooth
surfaces, lm thickness can be calculated (12) from
ho G2.65

0.54 ( oV )0.7R e0.43


w0.13E 0.03

(7.37)

Film thickness at EHD point contacts is given (13) by


ho G0.84( oV )0.74R e0.41

0.07

W
E

(7.38)

where W is the total load on the contact.


The thickness of the lm developed at the contact between smooth
surfaces has to be related to the topography of real surfaces. Usually,
the specic lm thickness, dened as the ratio of the minimum lm
thickness for smooth surfaces, ho , to the peak-to-valley roughness of
the contacting surfaces, is used

2ho
Rsk1CRsk2

(7.39)

where Rsk G1.11Ra is the root mean square height of the surface and
Ra is the centre-line average height of surface asperities.
The value of specic lm thickness is an important indicator of the
lubrication regime. If H3, the probability of metalmetal asperity contact is practically negligible and therefore no further analysis of the
contact is required. However, if F1, the operating conditions are
characteristic of boundary lubrication and the mode of imminent
asperity contact should be examined. The parameter that can be used

Machine Elements in Rolling Contact

271

to do this is the plasticity index. The range of values between 1 and


3 points to a mixed lubrication regime where the probability of asperity
contact is much smaller than in the boundary lubrication regime.
It is commonly accepted that most machined surfaces have asperity
height distributions that are closely Gaussian. This is particularly true
for the upper half of the surfaces, which includes the proportions of the
surfaces in contact. This observation makes the mathematical characterization of surfaces much more tenable. It is therefore possible to
determine a value for the standard deviation of the peaks, knowing the
value of the standard deviation of the surfaces. Thus, the peak height
r p1 GKm1 Rsk1 and
distribution may be expressed approximately as X
r
Xp2 GKm2 Rsk2 and the standard deviation of the peaks is given by p1 G
Kd1Rsk1 and p2 GKd2Rsk2 , where Km is the peak-to-surface mean proportionality factor determined for each type of surface manufacturing
method and Kd is the surface-to-peak standard deviation factor.
When F1, the mode of asperity contact has to be determined
through examination of the plasticity index. The plasticity index is given
by (14)

E
Pm

p
r

(7.40)

where Pm is the ow hardness of the softer material in contact, r is the


average radius of curvature of an asperity peak, and E is the effective
elastic modulus.
When F0.6 the contact between the asperities will be elastic under
all practical loads, and when H1 the contact will be partially plastic
even at the lightest loads. The deformation mode is mixed when
0.6 1, and an increase in load can change the contact of some of
the asperities from elastic to plastic.
Temperature criterion of scufng
Scufng is very unlikely when F0.6, but when H0.6 the probability
of scufng should be examined. This can be done by comparing the
surface contact temperature Ts and the characteristic lubricant temperature. Special attention must be given to the relation between the plasticity index value and the surface topography as affected by the
running-in process. It is commonly accepted that running-in improves
the surface nish in the sense that the highest peaks are removed and
the initial Gaussian height distribution is markedly skewed because of
the attening of the peaks. Taking into account the parameters involved
in the plasticity index equation, it can be said that the elastic nature of

272

Rolling Contacts

contact will be a dominant feature of a well run-in surface. Entirely


elastic contact requires F0.6. If, however, the plasticity index F1,
it means that 98 percent of the contact area is composed of elastic
asperity contacts. The remaining 2 percent will have yielded plastically.
The surface contact temperature, Ts depends on many factors such
as load, sliding velocity, the thermal and mechanical properties of contacting materials, and the coefcient of friction. An approximate estimate of the temperature rise due to frictional heating can be made using
a simple theory proposed by Bowden and Tabor (15)
Ts GToCT

(7.41)

where To is the bulk lubricant temperature entering the contact zone.


The equation for estimating the temperature rise, T, is
TG0.443

gV1(WPm)

(7.42)

J(k1Ck2)

where is the friction coefcient, V is the relative sliding velocity of


the contacting surfaces, W is the load on contact, J is the mechanical
equivalent of heat, k1 and k2 are the thermal conductivities of the materials involved, and g is the gravitational constant.
The fractional lm defect strategy is most appropriate for examining
the characteristic lubricant temperature Tcr . In physical terms, the fractional lm defect, , is a measure of the probability of two bare asperities coming into contact. It is given by the expression (16)

G1Aexp

30.9B105
V

exp RT

Tm
M

Ec

(7.43)

where V is the relative sliding velocity of the contacting surface, Tm is


the melting temperature of the lubricant, M is the molecular weight of
the lubricant, R is the universal gas constant, and Ec is the heat of
adsorption of the lubricant on the surface being lubricated.
Because of the double exponential dependence of the fractional lm
defect on the heat of adsorption, Ec , and the contact surface temperature Ts, a critical moment for the lubricating lm durability is when
just starts to increase. This is dened by the point on the versus Ts
curve where the change in curvature rst becomes a maximum. Mathematically, this is where the third derivative of the expression for is
zero. By calculating the third derivative and equating it to zero, an
expression for the characteristic lubricant temperature, Tcr , can be

Machine Elements in Rolling Contact

273

derived in the following form


Tcr G

3[(Ec RTcr)A2]A1{[5(Ec RTcr)2A12(Ec RTcr)C12]}


[(30.9B105)V ]1(TmV )(Ec RT 2cr) exp(Ec RTcr)
(7.44)

If the value of surface temperature, Ts , is greater than Tcr given by


equation (7.44), the probability of scufng is high, but when TsFTcr
scufng is not probable.
Probability of asperity contact
The fact that the thickness of the lubricant lm is not sufcient to provide a complete separation of the contacting surfaces and asperity contacts take place does not necessarily mean the onset of scufng. As long
as the asperity contact is elastic the probability of scufng is practically
negligible. Plastic asperity contact is only considered as able to initiate
scufng. Thus, the probability of asperity contact in general can be
estimated (17) from
P(h0)G

1
1(2)

exp

(hr Ah)2
2 2

d(h)

(7.45)

where h is the separation of the highest surface peaks and hr is the


distance between the mean lines of the surface peaks.
The probability of plastic contact is given by
P(h p )G

1
1(2)

exp

(hr Ah)2
2 2

d(h)

(7.46)

Plastic asperity deformation, p , is given by

p Gr

0.6Pm
E

(7.47)

The composite standard deviation of surface peaks, , is estimated


from

G1( 2p1C 2p2)


If hGhC p , then the probability of asperity plastic contact is
P(h0)G

1
1(2)

exp

(hr C pAh)2

2 2

d(h)

(7.48)

274

Rolling Contacts

The probability P(h0) of asperity plastic contact can be found from


the normalized contact parameter hr , where hr G(hr C p) is the
number of standard deviations from the mean value of separation hr .
Criterion for scufng
The probability of asperity contact given by equation (7.45) is related
to the probability of scufng, i.e.
P(scuffing)FP(asperity contact)
However, the probability of scufng is mainly inuenced by the plasticity index, , and the temperature in the contact zone. The necessary
condition for scufng to be seriously considered is when H0.6, which
means plastic asperity contact. The sufcient condition is fullled when
TsHTcr . When both conditions are simultaneously fullled, then
P(plastic asperity contact)FP (scuffing)FP(asperity contact)

7.7 Tooth face pitting


Elucidation of the origin of pitting has been quite an intractable problem. Areas undergoing pitting are found to have numerous minute pits
visible only under the microscope, besides those, upwards of several
micrometres across, visible to the naked eye. Although pitting generally
involves a progression from such small beginnings, the term progressive
is used only for the worst cases. Gears of the usual industrial quality
very often develop pitting that does not proceed beyond an early stage
in which the larger pits are no more than about 1 mm across. This socalled initial pitting is associated primarily with running-in, and commonly starts in the vicinity of the pitch line. It is attributable to local
overstressing caused by the presence of asperities and minor departures
from the geometrically correct proles of the teeth; as running-in continues, and the asperities are reduced, the surface stresses become more
uniform and the pitting is arrested. Sometimes, gears that have been in
use for a considerable period develop pitting that similarly comes to a
halt. For example, bearing wear may cause slight changes in alignment,
accompanied with local overstressing of the teeth which is in due course
reduced by corrective wear. Pitting beginning gradually and showing
no tendency to increase rapidly is likely to be of this temporary kind.
In a relatively few cases the pitting does not halt but continues to
increase, the pits gaining in size or total area or both. This is the kind
of pitting specicially known as progressive. It may not become manifest until the gears have been in service for some time, but if it occurs

Machine Elements in Rolling Contact

275

at all it is most likely to do so within about 18 months of commissioning


the gears. Unfortunately, in the initial stages of pitting it is generally
hard to tell whether it will develop into the arrested or the progressive
type. Progressive pitting is most often concentrated in the dedendum of
the teeth, though it may extend to the pitch line area, and in some cases
is concentrated there. The addendum is seldom much affected, until
damaged by engagement with the opposing dedendum. If such pitting
goes far enough, it will ultimately result in failure of the gears.
It is well established that the pitting of gear teeth is a surface fatigue
phenomenon, occurring where the material succumbs to repeated alternating shear and tensile stresses set up in rolling contact. Metals may
be fractured by such stresses at values, considering the loaded region as
a whole, well below their ultimate tensile stresses, which are determined
under static conditions. Fatigue failure comes about as a result of local
change in the material, or increase, brought about by stress alternation,
in local stress concentrations in the vicinity of stress raisers such as
inclusions or microcracks in the material or dislocations in the crystal
structure. It is accepted that in many cases the alternating stress causes
local gliding at the stress raisers, but with the effect that the material
becomes work-hardened and eventually gliding is prevented. If in such
a situation the applied stress exceeds a certain value, the metal will
crack, and the crevice thus produced will in turn act as a stress raiser
causing a stress concentration, enabling the process to continue as long
as the local stress is sufcient for its propagation.
Gear teeth commonly show surface cracks penetrating into the metal
in a direction inclined with respect to the tooth face, this direction
depending on that of sliding at each part of the tooth prole; metal
tends to be displaced by frictional traction in the direction in which the
opposing tooth slides over the surface, i.e. towards the root and tips of
the driver and towards the pitch line in the follower (Fig. 7.8). More
precisely, surface cracking tends to occur in lines parallel to the grain
distortion near the surface, commonly at an angle to the tooth face in
the range 520. It has been observed that the general tendency of
cracks to spread depends on the direction of sweep in relation to the
orientation of the crack. The explanation may be that if, as in Fig.
7.9(a), a crevice is inclined in the direction of sweep, it will tend to be
opened by the compressive stress immediately before the Hertzian band
passes over it, with the result that, when the mouth of the crevice is in
the loaded zone, oil will tend to be forced into it under pressure. Propagation of the crack will thus be furthered. On the other hand, it may
be that, if a crack is inclined in the opposite direction, as in Fig. 7.9(b),

276

Rolling Contacts

Fig. 7.8

although the crack will be squeezed and some oil pressure may thereby
be set up, the oil will tend to escape. Thus, although the crack may
certainly spread, the oil will contribute little or nothing to its propagation. Conditions favouring crack propagation by oil pressure arise
primarily where there is negative sliding, i.e. where sliding is opposite
in direction to sweep. This is so in the dedendum of the teeth, both for
the driver and the follower. In the addendum of the teeth there may be
occasional random cracks with the inclination favouring propagation
by oil pressure, but they will be exceptional.
When progressive pitting reaches a certain stage, the metal between
adjacent pits tends to be weakened and eventually breaks off. This
effect, known as spalling, takes very different forms according to the
scale and course of the pitting. It has been observed that in some cases
the cracking causing the pitting tends to travel downwards into the
metal to a certain depth, following a roughly uniform direction. In other

Machine Elements in Rolling Contact

277

Fig. 7.9

cases the crack might well start in one direction only, but after some
time a second crack may be initiated, so as to form a bifurcation. The
new limb may spread more rapidly, with the effect that the main line
of the crack changes course. There is a tendency for a crack to change
course to a direction parallel with the surface when it reaches the plane
of maximum shear stress. Where the cracks penetrate deeply into the
metal, the result tends to be formation of deep, coarse pits at a relatively
few localities. The collapse of intervening metal then produces gross
enlargement of the surface craters through separation of quite large
chips. If, however, the predominant tendency is for the cracks to travel
laterally in a subsurface stratum, the result is the development of large
numbers of pits of shallow and more or less uniform depth. As the
intervening metal is undermined and akes off, the surface is left rough
but with a contour following the original fairly closely. Commonly, the
process continues until the whole width of the tooth is affected. A certain additional period of operation is often required before further

278

Rolling Contacts

fatigue damage develops, and this may in turn take the form of fairly
uniform spalling. In the meantime, the appearance of the gears may
actually improve as the roughness remaining from the rst stage of
spalling is smoothed down by a renewed process of running-in. In practice, conditions intermediate between the two described are encountered, uniformly spalled areas being interspersed with some deeper pits.
Any factor increasing the stresses at the surface of the teeth will be
liable to promote pitting. Such an increase may result either from a
general increase in the load on the gears or from variation in stress
from point to point or between limited areas of the tooth face. In considering the loading of the gears, dynamic load increments must be
taken into account. Experience suggests that detrimental changes in
dynamic conditions, e.g. owing to vibration, are often the source of
an increase in loading. Differences in performance between similar or
identical gearsets may originate from differences in dynamic load
increment. The time to fatigue pitting in gears is inversely proportional
to load to some high power (probably 810). Hence, even a small
increase in dynamic load may very greatly increase the tendency to
pitting. Local variation in surface stress will vary from point to point
with the roughness of the surface, and will vary between distinct areas
with the accuracy of the gears. Usually, the better the surface nish and
prole accuracy the greater the resistance to pitting. In this respect it is
the initial state of the gears that matters. The attainment of a smooth
surface through running-in does not substitute for initial smoothness
and accuracy. Apart from the initiation of cracking during the relatively
rapid deformation during running-in, the running-in process does not,
in general, improve conformity of the surface with the geometrical
prototype and may in fact reduce such conformity. Localized variation
in loading, with a tendency to overstressing in places, will therefore tend
to persist. Other sources of local overloading and consequent pitting
include misalignment or thermal distortion, and also contamination of
the oil by hard particles, sources of which include the pitting itself at
earlier stages.
In gears prone to destructive pitting, an increase in the viscosity grade
of the oil used from the outset tends in some cases to defer the onset of
the pitting and retard it subsequently. A change to a thicker oil for
gears suffering destructive pitting similarly tends to retard its further
development. These tendencies are attributable to the increase in oillm thickness obtained at the loaded area, and a reduction in the distortion of the surface layers of the tooth material. However, experimental
results suggest that the time to pitting varies with viscosity to a low

Machine Elements in Rolling Contact

279

power of the order of 0.20.3. By comparison with that of load, the


inuence of viscosity is thus slight. Other factors governing the lm
thickness, notably the temperature at inlet to the loaded area, will be
capable of a similar effect. Improved cooling will have an effect similar
to that of using a higher viscosity grade.

7.7.1 Fatigue fracture


Gear teeth can break either through bending fatigue or through impact
or overload. Fatigue failure results from the progressive extension of a
crack, usually at or near the root of the tooth, until the remaining sound
material is so attenuated as to give way under load. As a rule, the
broken surface has a comparatively smooth, laminated texture, except
for the part that nally snaps, which generally has a granulated appearance. Within the laminated area there is often a point from which the
fault originated.
When a tooth breaks off, the shock as the mating tooth jumps the
gap in subsequent revolutions often carries away a number of other
teeth. There is usually little doubt, however, which tooth failed rst. If
other teeth are not carried away at once, the gears may not break down
immediately but may continue to run for some time; they are likely,
however, to suffer scufng or pitting, or accelerated development of
scufng or pitting, while running with a broken tooth.
There is often evidence of a stress riser such as an inclusion or cavity
at the initial point of fatigue failure. In other cases the failure must be
attributed to the progressive extension of surface or subsurface cracks
because of local overstressing, which have already resulted in heavy
pitting. It is to be borne in mind that in some elds, e.g. the automotive
eld, the normal practice in the interests of reduction in size and weight
is to require gears to have no more than a specic fatigue life beyond
which ultimate fatigue fracture is practically certain; as a rule, however,
this is not the case with industrial or marine gears, which are normally
designed for a practically indenite life.
An important point is that fatigue endurance data usually assume
dry conditions, by which is meant the absence of corrosive inuences.
If corrosion becomes involved, which usually means that moisture is
present, the corrosion feeds on the fatigue but at the same time promotes it, e.g. work-hardened material tends to suffer preferential corrosive attack, probably through the development of corrosion cells owing
to electrochemical differences between work-hardened and unaffected
metal, while crevices tend to retain moisture and become the seat of
corrosion cells through effects such as differential aeration. Under

280

Rolling Contacts

conditions of such corrosion fatigue there is no stress range below which


stress alternation can go on indenitely without progressive surface
fatigue, which can therefore cause damage at stresses well below the
limit under dry conditions.
Commonly, small amounts of red stain can be seen on the laminated
part of the broken surface of a tooth that has failed as a result of fatigue
accelerated by corrosion. Occasionally, red staining is detected at the
mouth of a crevice in which corrosion fatigue is going on. Unfortunately, the gears are liable to be in a dangerous state, if failure has not
already occurred, before corrosion fatigue is diagnosed as such. The
best preventive measure is to use an oil with good resistance to corrosion and as far as possible to keep the gears free from water, particularly salt water.

7.7.2 Impact fracture


Failure through sheer overload does not in itself have anything to do
with lubrication, particularly when a machine is suddenly stopped by
unintentional jamming of one of its components. However, it may be
the end-effect of progressive conditions in which lubrication was a
factor.
The surface of metal where impact or overload fracture has occurred
is generally much coarser than that characteristic of fatigue and often
includes areas where the material has been elongated and left with a
brous appearance. Overload may be due simply to the application of
excessive torque by the driving unit. It is, of course, unlikely that the
steady load greatly exceeds the capacity of the gears unless they have
been seriously weakened in some way. However, the dynamic load
increments cannot always be predicted with total accuracy and are often
liable to increase as a result of extraneous changes or deterioration of
the gearset. Dynamic effects causing overload include transmitted
shock, torsional vibration, lack of dynamic balance, misalignment of
the gears, distortion of the casing or mounting, excessive shaft deection, and bearing wear. Deterioration of the gears or bearing may also
be due to inadequacy of lubrication.
Except where wear has the effect of producing some benecial tip
relief, wear of the teeth usually increases the dynamic load increments
as the teeth pass through the meshing cycle. Any departure of the teeth
from the correct prole will have a similar effect, by causing either
impact and shock, or excessive deections, or both. A common result
of severe overloading is plastic yielding of the surface material; this
will in the same way increase the dynamic overloads and accelerate

Machine Elements in Rolling Contact

281

deterioration of the gears. Frequently, overloading leads to cracking


due to bending fatigue in the root llet, where there is a tendency to
stress concentration.
Fracture will occur where the metal is weakest in relation to the loads
on it. Commonly, this is at the root of the tooth, given that it is essentially a cantilever. Sometimes, however, progressive pitting at the pitch
line or in the dedendum may initiate fracture in these positions. Casehardened teeth in which the case is too thick are left with too weak a
core and are liable to break near the crest of the core, which may be as
low as the pitch line.
Fracture at the extreme ends of the teeth, rather than some distance
towards the middle of the tooth face, may be an indication of misalignment or damage to the ends of the teeth during or prior to assembly,
although a factor here is that the ends of the teeth are weakest since
there is no material beyond them to act as a buttress. In other cases
the pattern of fracture may be due to the haphazard position of stress
raisers such as inclusions in the material.

7.7.3 Tooth loading


The classic solution of the problem of stress distribution and deformation between elastic bodies in point or line contact was achieved by
Hertz for cases where the dimensions of the actual contact area are
small by comparison with the radii of curvature of the engaging bodies.
In relation to spur and helical gears it is assumed that the relevant case
is the comparatively simple one of line contact between cylinders with
parallel axes.
The contact zone between the cylinders takes the form of a narrow
rectangular band. Across this band the pressure distribution is semielliptic, with a maximum at the mid-point. It follows that the maximum
compressive stress qmax is given by
qmax G

4w

(7.49)

where w is the normal load per unit length of contact and z is the width
of the contact band. The maximum shear stress is equal to 0.295qmax ,
and this occurs at a depth of 0.393z.
The bandwidth has the value
zG4

wr

1A 21

E1

1A 22
E2

(7.50)

282

Rolling Contacts

where r is the relative radius of curvature (radius of conformity) as


dened below, 1 and 2 are the Poissons ratios for materials of cylinders 1 and 2 and E1 and E2 are the Youngs moduli for materials of
cylinders 1 and 2.
The relative radius of curvature r is dened by
1

1 1
G C
r r1 r2

where r 1 and r 2 are the radii of cylinders 1 and 2.


For steels in general, variation in Poissons ratio can be neglected.
Assuming a value of 0.3 for 1 and 2 , the value of z becomes
zG3.82

wr 1
1
C
E1 E2

(7.51)

A combination of materials having Youngs moduli E1 and E2 will produce the same stress conditions as if both have the harmonic mean
value of Youngs modulus E prescribed by

1
1 1 1
G
C
E 2 E1 E2

Hence
zG3.05

wE
r

(7.52)

and
qmax G0.418

wE
r

(7.53)

In the United Kingdom, a criterion of surface stress known as the Sc


value is dened by
ScGwr

(7.54)

where w is the normal load per unit length of contact (Nm) and r is
the relative radius of curvature (m). The Sc value as dened above has
the dimensions of stress but is not itself the actual maximum or mean
stress at the loaded area. From the above
qmax G0.4181(Sc E)

(7.55)

Machine Elements in Rolling Contact

283

Assuming for steelsteel gears a value of 211 GPa for E, this reduces
to
qmax G192B1031Sc

(7.56)

A somewhat more practical criterion of the stress conditions between


gear teeth is that known both in the United Kingdom and the United
States as the K factor. It is dened by
KG

Ft C1
d

(7.57)

where Ft is the tangential load per metre of face width, d is the pinion
pitch circle diameter, and is the reduction ratio.
In spur gears the relative radius of curvature r at the pitch line has
the value
rG

d sin
2

C1

(7.58)

where is the pressure angle.


Assuming the least favourable condition in which the whole of the
load is borne by one pair of teeth, the maximum tangential load per
unit face width, Ft, is
Ft Gw cos

(7.59)

From
qmax G192B1031(wr)
qmax G192B103

2
Ft C1
sin cos d

or
qmax G271.5B103

K
sin cos

(7.60)

With a pressure angle of 20


qmax G479B1031K

(7.61)

Thus, for spur gears with a pressure angle of 20


Sc G6.3K

(7.62)

284

Rolling Contacts

In helical gears the relative radius of curvature r at the pitch line has
the value
rG

sin n

2 C1 cos

(7.63)

where n is the normal pressure angle and is the helix angle.


The total load on the tooth normal to the tooth proles and in the
normal plane, say W, has the value
WG

FtL

(7.64)

cos cos n

where L is the face width.


In accurate helical gears the load is distributed over a total length of
contact, say ct, to which several pairs of meshing teeth contribute. This
is the product of the full length of contact between one pair of meshing
teeth, say c, and the number of axial pitches in the face width, say ln.
Thus
ct Gcln
cG

nct pbt

(7.65)

sin

where nct is the transverse contact ratio and pbt is the base pitch in the
transverse plane
ln G
ct G

L tan
pbt

(7.66)

Lnct

(7.67)

cos

Hence, the normal load per unit length of contact w has the value
wG

Ft
W
G
ct cos n nct

(7.68)

From
qmax G192B1031(wr)
qmax G192B103

2 cos2
Ft C1
nct sin n cos n d

(7.69)

Machine Elements in Rolling Contact

285

or
qmax G271.5B103

K cos2
nct sin n cos n

(7.70)

When the pressure angle is 20 and the helix angle is 30, assuming a
typical value of nct of 1.5
qmax G339B1031K

(7.71)

Sc G3.12K

(7.72)

or

7.8 Camfollower system


Generally speaking, a cam is understood to have rotary motion, either
continuous or oscillatory, while the follower may have either a rotary or
reciprocating motion. The common conguration of the camfollower
system is schematically shown in Fig. 7.10.

Fig. 7.10

286

Rolling Contacts

Cam motions have wide application in all classes of automatic


machinery, since by varying the cam prole it is possible to impact any
desired motion to the follower or driven element, within its range of
action. The discussion presented here will, however, be limited to those
types in which the follower moves in a plane perpendicular to the axis
of rotation of the cam and which are generally referred to as radial or
plate cams.

7.8.1 Reciprocating engine cam


The cam plays the same part in an internal combustion engine as the
eccentric of a steam engine. Cams operating the valves of internal combustion engines are of the type shown in Fig. 7.11. A tappet is usually
placed between the cam and the valve and left with a small clearance
to ensure that the valve closes properly when the tappet or push rod,
which is the follower immediate to the cam, is in its lowest position.
The cam prole will depend primarily upon the desired timing of the
various operations in relation to the crankshaft angles of the engine,
and upon the nature of the valve, i.e. whether inlet or exhaust. As an
illustration of the principles involved, the valve settings on a four-stroke
cycle engine, the indicator diagram of which is shown in Fig. 7.12(a),
will be considered. At point 1, just before the completion of the exhaust
stroke, the fuel and air inlet valve opens and remains open until point
2, just after the commencement of compression. Similarly, at point 4,
just prior to the completion of the expansion stroke, the exhaust valve
opens and remains open until point 5, just after the completion of the
exhaust stroke. The corresponding angles of rotation of the crank are
shown by 2 and 2 in Fig. 7.12(b). The object of the angle of overlap
between operations 1 and 5 is to improve scavenging effects on the
exhaust gases by the incoming charge, and it is usually small except for
high-speed engines. Typical average values for these two angles are
2 G196 and 2 G232. Since the operations occur in two revolutions
of the crank, it follows that the camshaft must rotate at half the speed
of the crankshaft, so that average camshaft angles are G98 and G
116. To take account of tappet clearance, these angles must be
increased slightly in designing the cam prole. If the camshaft is gear
driven it will rotate in the opposite direction to the crankshaft of the
engine.
The tangent or straight-sided cam was probably the rst to be used
in internal combustion engines and is shown in Fig. 7.11(a) and (b).
Suppose the follower commences to lift when the roller makes contact
at point A. Then, if clearance is neglected and the angle AOD represents

Machine Elements in Rolling Contact

Fig. 7.11

287

288

Rolling Contacts

Fig. 7.12

the period of opening of the valve, the follower must reach its position
of rest when the roller makes contact at D. Tangents AB and CD are
drawn to the circle representing the least radius of the cam. This circle,
of radius OA, will be referred to as the base circle, and AB and CD as
the anks of the cam. The prole of the nose BC of the cam is made
up of one or more circular arcs, forming a continuous curve in a manner
depending upon the total valve lift and the total angle of action during
which valve lift occurs.
Denoting the radius of the base circle by r1 and the total lift of the
follower by L and considering the exhaust valve cam [Fig. 7.11(b)], the
arc BC of radius OBGOCGr1CL subtends a small angle at the
centre of rotation O of the camshaft, and extends equal distances on
either side of the line of symmetry. The prole is completed by circular
arcs BB and CC of equal radii connecting BC with the straight
portions.
It will be shown in the next section that the velocity and acceleration
of the follower progressively increase against the resistance of the spring
control of the valve, while the contact point moves from A to B. When
point B is reached, retardation of the follower commences, contact
between the roller and the cam being maintained by the retarding force
of the spring, until at B the velocity is reduced to zero.

Machine Elements in Rolling Contact

289

From B to C, represented by the angular displacement , the valve


is temporarily at rest. This interval is termed a period of dwell and the
valve is fully open. When point C is reached, the spring force accelerates the valve and follower, and the downward velocity reaches a maximum value at C. Thereafter, deceleration occurs and the velocity of the
follower again falls to zero at D.
The longer the period of dwell, the greater are the accelerating and
decelerating forces to which the follower is subjected at points B and
C. These forces are reduced to a minimum when there is no dwell at
full lift, i.e. when the nose of the cam is circular, as shown for the inlet
valve in Fig. 7.11(a).
A more even distribution of acceleration and deceleration while the
follower is passing from A to B and from C to D is obtained by using
convex anks [Fig. 7.11(c)]. In this case, a at follower may be used.
The greater the convexity, the smaller are the accelerating forces at B
and C.
The other extreme arises with concave anks and a long period of
dwell. This gives rapid opening and rapid closing of the valve, and the
long period of maximum lift allows maximum piston speed. Stresses
and associated wear of the valve gear is, however, excessive, and a cam
of this nature results in noisy action as the valve comes abruptly into
contact with the seating. Strong springs are needed to prevent the valve
from bouncing. It ought to be noted that a at follower cannot be used
with cams having at or concave surfaces.

7.8.2 Analysis of the follower motion


Consider a radial cam with a roller follower in which the line of reciprocation of the follower is offset by a distance p from the centre of
rotation O of the cam, as shown in Fig. 7.13. Suppose OX and OY are
two mutually perpendicular axes of reference in the plane of the cam
and rotating with it. The centre of the roller is at P, and at any instant
(x, y) are the coordinates of P referred to the axes OX and OY respectively. Line OQ is xed in space and perpendicular to the path of P
shown by EP.
Assuming the cam rotates with uniform angular velocity , then
G

d
dt

(7.73)

where is the inclination of OX to OQ at the instant considered.


Furthermore, if h is the height of P above OQ at time t, then
pGx cos Ay sin

(7.74)

290

Rolling Contacts

Fig. 7.13

and
hGx sin Cy cos

(7.75)

Differentiating equation (7.74) with respect to t, p being constant, gives


cos

dx
dy
Asin G (x sin Cy cos )G h
dt
dt

(7.76)

Similarly, from equation (7.75)


VG

dh
dt

G (x cos Ay sin )Csin

dx
dt

Ccos

dy
dt

and so
VG pCsin

dx
dy
Ccos
dt
dt

(7.77)

where V is the velocity of the follower along EP.


Now, suppose that P moves on the curve yGF(x), where F denotes
a function of x, referred to the moving axes. Then
dy
dt

dy dx
dx dt

GF(x)

dx
dt

(7.78)

Machine Elements in Rolling Contact

291

where dydxGF(x) is the slope of the path of P at (x, y). Substituting


for dydx, equations (7.76) and (7.77) become
dx

VG pC

dt

[sin CF(x) cos ]

and

hG

dx
dt

[cos AF(x) sin ]

Eliminating dxdt, the velocity of the follower is given by


VG pC h

sin CF(x) cos

cos AF(x) sin

(7.79)

For the particular case where pG0


VG h

sin CF(x) cos

cos AF(x) sin

(7.80)

If F(x) is known and F(x) is determined, the acceleration of the follower can be found by differentiation of the expression for V. The application of this result is illustrated by the special cases detailed below.

7.8.3 Tangent cam with a roller follower


For simplicity, suppose the line of reciprocation passes through O, i.e.
pG0. Referring to Fig. 7.14(a) and (b), let r1 be the radius of the base
circle, r2 the radius of the roller, r3 the radius of the tip circle BB, s the
length of the straight ank AB, and d the distance of the centre O of
arc BB from O.
Roller is on the tangent
This case is depicted by Fig. 7.14(a). Taking the axis of x parallel to
AB, the point P will move along a line parallel to OX relative to the
moving axes. Thus, yGk, where kGr1Cr2 Gconst, and so
dy

GF(x)G0

dx
Hence the general equation for V reduces to
VG h tan

(7.81)

292

Rolling Contacts

Fig. 7.14

Machine Elements in Rolling Contact

293

Fig. 7.14 Continued

From the gure, kGh cos or hGk sec so that


VGk sec tan

(7.82)

Differentiating with respect to t


dV
G 2k sec (1C2 tan2 )
dt

(7.83)

Since s denotes the length of the straight ank AB, these results apply
from G0 to Gtan1 (sk). When point B is reached, i.e. the roller
makes contact with the straight ank at B as shown in Fig. 7.14(b),
then
xGs,

yGk,

hG1(x2Cy2 )G1(s2Ck2 )

and thus
VG h tan G

s
1(s2Ck2 )
k

294

Rolling Contacts

and

2s2
dV
G 2h(1C2 tan2 )G 21(s2Ck2) 1C 2
dt
k

(7.84)

where kGr1Cr2 .
Roller is on the circular arc
In this case, illustrated in Fig. 7.14(c), suppose OX1 and OY1 to be the
moving axes of reference, where OX1 passes through point O. Further,
let (x1 , y1) be the coordinates of the centre of the roller with reference
to these axes, and 1 be the inclination of OX1 to xed line OQ at the
instant considered.
The equation of the circular of P is then
(x1Ad)2Cy 21 G(r2Cr3)2
where x1 GOG, y1 GPG, and OOGd.
Upon differentiation with respect to x1, the following is obtained
2(x1Ad)C2y1

dy1

G0

dx1

or
x1Ad
dy1
GF(x1)G
dx1
y1
Substituting in the general equation for V when pG0 gives
VG h

y1 sin 1Ax1 cos 1Cd cos 1


y1 cos 1Cx1 sin 1Ad sin 1

However, since pG0


x1 cos 1Ay1 sin 1 G0
x1 sin 1Cy1 cos 1 Gh
The equation for V thus reduces to
VG h

d cos 1
hAd sin 1

Now
hAd sin 1 GOPAONGPN
d cos 1 GON

(7.85)

Machine Elements in Rolling Contact

295

and thus
VG h

ON
PN

G h tan

(7.86)

where is the inclination of OP to the line of reciprocation. For determination of , use can be made of
sin G

ON

d cos 1

OP

(7.87)

r2Cr3

and for the height of P above OQ


hGd sin 1C(r2Cr3) cos

(7.88)

Alternatively, the expression for V can be determined by direct differentiation of the equation for h and sin , thus
VG

dh
dt

G d cos 1A(r2Cr3) sin

d
dt

where
cos

d
dt

d
r2Cr3

sin 1 G

r Cr Acos
h

Eliminating 1 and d dt, the velocity equation reduces to


VG h tan

(7.89)

To determine the acceleration of the follower, the equation for V is


differentiated, and thus
dV
dt

G V tan C h sec2

(7.90)

dt

Eliminating V and d dt, this becomes


dV
dt

G 2h tan2 A 2h

d
sec3 sin 1
r2Cr3

or
dV
dt

G 2h

r Cr sec sin Atan


d

(7.91)

The two extreme positions B and B [Fig. 7.14(c)] are important, and it
is necessary to examine them in detail.

296

Rolling Contacts

Contact point at B
At this critical position, represented by Fig. 7.14(b), G , so that
VGh tan

(7.92)

and
dV
dt

G 2h

r Cr sec sin Atan


d

(7.93)

where tan Gsk and hG1(s2Ck2). The value of 1 in relation to


will depend upon the cam prole. Thus, for the symmetrical cam of
Fig. 7.14(b), if is the total angle of rotation and is the angle of
dwell, then

1A G 12 A 12 ( A )

(7.94)

and when G0

1A G 12 (A )

(7.95)

Comparing these results with those already obtained for the straight
ank, it can be seen that the velocity when the contact point is at B is
the same in each case. As the contact point passes on to the arc BB,
the positive acceleration of the follower suddenly changes to a retardation, the magnitude of which is given by equation (7.93).
Contact point at B
In this position, 1 G90 and G0, so that VG0 and
dV
dt

G 2h

d
r2Cr3

Writing hGdCr2Cr3 , this becomes

d
G 2d 1C
dt
r2Cr3

dV

(7.96)

For the system to function properly, it is quite clear that external force
is necessary to maintain contact between the follower and the cam. This
force is supplied by a compression spring in the case of a cam-operated
valve and tappet. Not only does the spring provide the necessary
retarding force during the latter part of the upward movement of the
valve, but in addition it produces a downward acceleration during the
early part of the return stroke. During the latter part of the return
stroke the motion of the follower is retarded by the cam prole.

Machine Elements in Rolling Contact

297

The necessary spring stiffness and initial compression will depend


upon the maximum value of the retardation during the upward movement of the valve, together with the valve lift, and for this reason the
straight-sided cam is unsuited to high speed. For a cam with convex
anks, the change in curvature in passing from the ank to the nose
arc is less marked and the instantaneous retardation at the point of
discontinuity is therefore less than with the straight-sided cam.
The effect of tappet clearance is shown in Fig. 7.15. Thus, if the
clearance is denoted by c, the valve will commence to lift when hG
r1Cr2CcGkCc, and if o is the corresponding value of , then
hGk sec o
and so
cos o G

(7.97)

kCc

so that the effective angle of cam action is reduced to ( A2 o ).

7.8.4 Camshaft torque


Referring again to the case of a cam with tangent anks and base circle
radius r1, operating a follower through a roller of radius r2 (Fig. 7.14),
suppose that the follower acts against a spring of stiffness S. Let x be
the initial compression of the spring, M the equivalent mass effect of
the valve, spring, and follower, assumed to be concentrated at the centre
of the roller, l the valve lift at angle , and F the force exerted by
the cam on the roller in a direction normal to the contact surfaces.
Gravitational and frictional effects are neglected.
Referring to Fig. 7.16, it can be seen that F cos is the component
of F in the direction of motion of the follower, and S(lCx) represents
the spring force opposing the motion of the follower. Thus
F cos AS(lCx)GM

dV
dt

or

FG S(lCx)CM

dV
dt

sec

(7.98)

Hence, the reaction torque on the camshaft is

Rh sin G S(lCx)CM

dV
dt

h sec sin

298

Rolling Contacts

Fig. 7.15

or

Reaction torqueG S(lCx)CM


where lGhA(r1Cr2).

dV
dt

h tan

(7.99)

Machine Elements in Rolling Contact

299

Fig. 7.16

Thus, when the contact point is on the straight ank, G and


hG(r1Cr2) sec
dV
dt

G 2h(1C2 tan2 )

Again, if dVdt Gf, where f is the maximum retardation to which the


follower is subjected, then, if contact is to be maintained
S(lCx)AMfH0
that is
SH

Mf
lCx

(7.100)

300

Rolling Contacts

7.8.5 Convex cam with a roller follower


Consider now the case of a cam with circular arcs for both ank and
nose proles. Let the centre of the ank arc be at O at a distance e
from the centre of rotation O of the camshaft, so that nose and ank
radii are r3 and r4 GeCr1 respectively.
When the contact point is on the nose arc, the equations representing
the motion of the follower are
hGd sin 1C(r2Cr3) cos

(7.101)

VG h tan

(7.102)

dV
d
G 2h
sec3 sin 1Atan2
dt
r2Cr3

(7.103)

where sin Gd cos 1(r2Cr3).


When the contact point is on the ank arc, as shown in Fig. 7.17, the
same expressions may be used by substituting r4 for r3, e for d, and
( A12 ) for 1, so that
hG(r2Cr4 ) cos Ae cos

(7.104)

VG h tan

(7.105)

dV
e
G 2h tan2 C
sec3 cos
dt
r2Cr4

(7.106)

where sin Ge sin (r2Cr4).


The proportions of the cam will depend upon the desired ratio q,
which is dened as the ratio of the duration of the acceleration period
during lift to the duration of the deceleration period during lift. Thus,
for a symmetrical cam with no dwell, the value of corresponding to
the contact point at the junction of the nose and ank arcs is given by
qG 1
2

and since 1A G12 (A ), then 1 and are determined when the total
angle of action is known.
Furthermore, suppose that the maximum lift L, the base circle radius
r1 , and the radius of the roller r2 are given. Then
LGdCr3Ar1

(7.107)

eGr4Ar1

(7.108)

Machine Elements in Rolling Contact

Fig. 7.17

301

302

Rolling Contacts

and at the end of the acceleration period [Fig. 7.17(b)]


sin G

d cos 1
r2Cr3

e sin
r2Cr4

(7.109)

Also, from the triangle OOO [Fig. 7.17(b)], as cos( A12 )Gcos 12
cos 12 G

(r4Ar3)2Ad 2Ae2
2de

(7.110)

These four equations determine the nose and ank radii r3 and r4 , and
the corresponding centre distance d and e respectively.

7.8.6 General case of a convex cam with a roller follower


For the tangent cam and the convex cam with a circular nose and ank
prole, described in the last section, the discontinuities in curvature
result in a sudden change from acceleration to deceleration of the
follower during lift.
Consider now the general case in which the nose and ank proles
form a continuous curve such that the equation of the path of P referred
to the moving axes OX1 and OY1 is given by y1 GF(x1) as illustrated in
Fig. 7.18.
At any instant, let be the inclination to the line of reciprocation of
the follower of the common normal at the point of contact of the roller
and the cam prole. Then, for point P on the curve y1 GF (x1), the
following applies
dy1
dx1

Gtan[A( 1A )]Gtan( 1A )

(7.111)

where 1 is the inclination of OX1 to xed line OQ at the instant


considered.
The velocity of the follower is given by
VG h

sin 1CF(x1) cos 1


cos 1AF(x1) sin 1

so that
VG h

sin 1Atan( 1A ) cos 1


tan 1Atan( 1A )
G h
cos 1Ctan( 1A ) sin 1
1Ctan 1 tan( 1A )

Machine Elements in Rolling Contact

303

Fig. 7.18

or
VG h tan
This result is perfectly general and applies to any cam prole.

(7.112)

304

Rolling Contacts

Parabolic prole
Suppose the prole is parabolic, and that A and D are the points of
tangency where the parabola touches the base circle of radius r1. The
cam is symmetrical and H is the value of h when 1 G12 .
The equation of the parabolic path of P may be written as
y21 G2(HAx1)

(7.113)

where is the radius of curvature of the parabola at its vertex.


Let r2 be the radius of the roller and be the total angle of action.
Then, writing kGr1Cr2, constants and H are given by the condition
y1 Gk sin 12 ,

dy1dx1 Gcot 12 ,

when x1 Gk cos 12

Thus
2y1
dy1
dx1

dy1
dx1
G

G2

y1

and nally

Gk cos 12

(7.114)

and
HG12 (sec2 12 C1)

(7.115)

Now, for any position (x1 , y1), the equations y1 Gh cos 1 and
dy1
dx1

Gtan ( 1A )G

h cos 1

are valid so that


tan ( 1A )G

h cos 1

(7.116)

Differentiating this expression, and recalling that d 1dtG and


dhdt GVG h tan , the following is obtained
d

A dt sec ( A ) cos G h sin( A ) sec


2

Machine Elements in Rolling Contact

305

However, from equation (7.106) the acceleration of the follower is given


by
dV
d
G 2h tan2 C h sec2
dt
dt
and eliminating d dt

sin( 1A ) sec
G 2h tan2 Csec2 1A
dt
h sec2( 1A ) cos2 1

dV

Substituting
cos2 1 G

2 2
cot ( 1A )
h2

then

h
G 2h tan2 Csec2 1A sec sin3( 1A )
dt

dV

(7.117)

Commencement of motion of the follower


The following equations are applicable for this stage of camfollower
system operation

1A G12 (A )
so that, if G0, 1 G12 A12 , G0, and hGk
k

dV

G 2k 1A cos3 G 2k sin2
dt

2
2

(7.118)

Maximum lift
Assuming that, 1 G12 , G0, and hGH, it follows that
H
1
dV

G 2H 1A
GA 2H tan2
dt

2
2

(7.119)

Consider now the case of a symmetrical cam with a roller follower, the
cam having a parabolic tip with circular anks. If the radius of the
ank arc is made equal to the radius of curvature of the parabola at
the junction of the two curves, then, by the crack and connecting rod
analogy, sudden change in the magnitude of the acceleration of the
follower as it leaves the circular ank will be avoided.

306

Rolling Contacts

As the follower passes over the circular ank, the acceleration will
increase, reaching a maximum positive value when the contact point is
at the end of the ank arc. Beyond this position the acceleration will
diminish, reaching its maximum negative value at full lift. At the junction of the ank arc and the parabolic tip there will be a sudden change
in the slope of the acceleration curve, resulting from the difference in
the rate of change in curvature of the prole.

7.8.7 Convex cam with a at follower


Considering now the case of a radial cam operating a at follower (Fig.
7.19), suppose that yGF(x) represents the prole of the cam referred
to rectangular axes OX and OY in the plane of the cam and rotating
with it. Point P is the point of contact of the prole with the surface of
the follower, and at any instant (x, y) are the coordinates of P. The
cam rotates with uniform angular velocity Gd dt, where is the
inclination of OX to xed line OQ at the instant considered, and the
contact surface of the follower remains parallel to OQ.
Referring to Fig. 7.19(a), let h be the height of P above OQ at time
t. Then
hGx sin Cy cos
so that
dh
dt

G (x cos Ay sin )Csin

dx
dy
Ccos
dt
dt

Now the slope of the prole at P referred to axes OX and OY is


dy
dx

Gtan(A )

that is
F(x)Gtan
Further
dx
dx
dy
GF(x) Gtan
dt
dt
dt
so that
dh
dx
G (x cos Ay sin )C (sin Acos tan )
dt
dt

Machine Elements in Rolling Contact

307

Fig. 7.19

The second term is zero and


x cos Ay sin Gp

(7.120)

so that
VG p

(7.121)

308

Rolling Contacts

where V is the velocity of the follower in a direction perpendicular to


OQ and p is the perpendicular distance from O to the normal at P at
the instant considered. The acceleration of the follower is thus given by
dV
dt

dp
dt

(7.122)

Equations (7.121) and (7.122) are valid for any convex cam with a at
follower.
Applying the above results to a convex cam with circular nose and
ank proles [Fig. 7.19(b)], let OOGd and OOGe, where O and O
are the centres of the nose and ank arcs respectively. The gure is
drawn for the position where P is at the junction of the two arcs.
When the contact point is on the circular ank arc, then
pGe sin

(7.123)

VG e sin

(7.124)

and
dV
G 2e cos
dt

(7.125)

These equations apply from G0 to the position shown in the diagram,


and during the corresponding period on the return stroke.
When the contact point is on the circular nose arc, then
pGd cos 1

(7.126)

VG d cos 1

(7.127)

and
dV
dt

G 2d sin 1

(7.128)

These equations apply from the position shown in the diagram to the
position (A 1), when the opposite extremity of the nose arc is in contact with the at follower. Maximum lift occurs when 1 G12 and the
two sets of equations are connected by the relation

1A G 12 (A )

(7.129)

where is the total angle of action of the cam.


In order to determine the principal dimensions of the cam prole,
suppose that and 1 relate to the position where P is at the junction

Machine Elements in Rolling Contact

309

of the two arcs as shown in Fig. 7.19(b). Also, let L be the maximum
lift of the follower, r1 the radius of the base circle, r3 the radius of the
nose arc, and r4 the radius of the ank arc. Thus
LGdCr3Ar1

(7.130)

r4 GcCr1

(7.131)

pGd cos 1 Ge sin 1

(7.132)

and from triangle OOO, since angle OOO is equal to A


1
2

1
d 2Ce2A(r4Ar3)2
cos A G
2
2de

(7.133)

The above results show that, as point P passes the junction of the two
arcs, the sudden change from acceleration to deceleration of the follower is accompanied with a sudden reversal in the sign of dpdt, i.e.
in the lateral velocity of sliding at P. At this critical point pGd cos 1 G
e sin has its maximum value. When P is on the ank arc, the maximum velocity of sliding occurs at the commencement of lift when G
0. Similarly, when P is on the nose arc, the maximum velocity of sliding
occurs at full lift when 1 G12 . The changes in the velocity of sliding
are similar to the changes in acceleration of the follower.

7.8.8 Stresses within the camtappet contact


The stresses within the contact between the cam and tappet are always
an important aspect of the system operation, inuencing, to a signicant
extent, its reliability and durability. Most tappets and cams can be
classied into one of the forms shown in Fig. 7.20. Also, the Hertz
theory can be used to evaluate the contact stresses.
In the equations presented below, the following symbols are used: W
is the load between the cam and tappet, b is the width of the cam, Rc
is the cam radius of curvature at the point under consideration, Rt is
the tappet radius of curvature, Rt1 is the tappet radius of curvature in
the plane of the cam, and Rt2 is the tappet radius of curvature at rightangles to the plane of the cam.
In the case of a at tappet face on the cam [Fig. 7.20(a)], the maximum Hertzian stress at the point under consideration is
pmax GK

W
Rc b

(7.134)

310

Rolling Contacts

Fig. 7.20

The centre-line of the tappet is often displaced slightly axially from the
centre-line of the cam to promote rotation of the tappet about its axis.
This improves scufng resistance but is considered slightly to reduce
the pitting resistance.
Since the theoretical line contact of the at tappet face on the cam is
often not achieved, on account of dimensional inaccuracies including
asymmetric deection of the cam on its shaft, edge loading occurs. In
order to avoid this, a large spherical radius is often used to the tappet
face. Automotive engines use a spherical radius of 7602540 mm.
Usually, to promote tappet rotation, the tappet centre-line is displaced

Machine Elements in Rolling Contact

311

slightly from the axial centre-line of the cam and the cam face is tapered. Alternatively, the longitudinal tappet axis is tilted by a suitable
amount to the camshaft axis. The theoretical point contact extends into
an elongated ellipse under load to give a better contact zone than with
the nominally at face.
The maximum contact stress is given by
pmax GXK

1 2
C
W
Rt Rc

(7.135)

where K can be obtained from the diagram shown in Fig. 7.21 after
evaluating (1C2Rc Rt) and XG8380 for a steel on steel material combination, XG7223 for steel on cast iron, and XG6396 for cast iron on
cast iron.
The maximum contact stress for curved and roller tappets with a at
transverse face is given by the following expression
pmax GY

1 W
C
Rc Rt b

(7.136)

Fig. 7.21

312

Rolling Contacts

where YG1875 for a steel on steel material combination, YG1676 for


steel on cast iron, and YG1526 for cast iron on cast iron.
Finally, the contact pressure in the case of a curved tappet with large
transverse curvature (crowning) can be obtained from
pmax GXK

1
1 2
C C
W
Rt1 Rt2 Rc

(7.137)

where the values of X for material combinations are as those listed


above and K can be obtained from Fig. 7.21.

7.8.9 Lubrication of the camtappet contact


The schematic geometry of the contact between a cam and at follower
is depicted in Fig. 7.22. The parameter relating the lubricating lm
thickness to the roughness of the contacting surfaces can be estimated
from the following expression
G1034.35

(bSn)0.74R0.26

(7.138)

Fig. 7.22

Machine Elements in Rolling Contact

313

where n is the camshaft rotational speed (rmin), SG(1011 o p) is a


lubricant parameter ( o is the lubricant viscosity at atmospheric pressure and p is the pressureviscosity coefcient), and bG2r1Al, where
l is the distance from the nose tip to the shaft axis (Fig. 7.22) and r1 is
the nose radius. The equivalent radius of the system is
1

1 1
G C
R r1 r
where r is the radius of the follower (in the cases of a at follower rG
S), G1( 21C 22) is the composite roughness of the system ( 1 and
2) denote the root mean square surface roughness (m) of surface 1
and 2 respectively. If the arithmetical average Ra is available, it is necessary to multiply it by 1.3 to convert to the root mean square measure
of surface roughness.
In general, the value of in cam systems is well below unity. In
this regime, elastohydrodynamic lubrication is not very effective and
boundary lubrication is the last frontier in the battle against scufng
and excessive wear of contacting surfaces.

7.8.10 Design considerations


Safe values of contact stress are dependent on a number of factors such
as: the materials and the material combination in use; heat treatment
of the materials, including the use of antiscuff treatments such as phosphating; surface nishes; the required fatigue life; the relative sliding
velocity between the cam and tappet; the oil viscosity at operating temperature; the inuence of any oil additives such as zinc
dithiophosphates.
Practical experience indicates that the maximum pitting tendency is
usually at around half-speed in engines. Also, practical experience
shows that the highest contact stresses can only be used with extremely
good surface nishes. In internal combustion engines, which work to
high contact stress levels, a cam surface nish of about 0.4 m is used,
and a tappet surface nish of about 0.15 m. The required surface nish
of the cam and tappet should be estimated in the light of the parameter given by equation (7.138).
Experience shows that a high working oil viscosity is desirable, with
a copious supply of lubricant for the best results. Studies on cams in
relation to scufng indicate that scufng is likely to occur when the
local metal temperature reaches about 200C. For this reason, the oil
should be supplied at the lowest practical temperature for the best

314

Rolling Contacts

results. In internal combustion engines, zinc dithiophosphates are added


to the oil as an antioxidant and antiscufng agent. However, there is
evidence that this additive worsens the pitting tendency and that this
effect rapidly increases when oil temperatures in excess of about 110 C
are reached, apparently with some corrosive inuence.

7.9 References
(1) Hertz, H. (1896) The contact of elastic bodies. Miscellaneous Papers (Macmillan, London).
(2) Reynolds, O. (1886) On the theory of lubrication and its application to Mr Beauchamp Tower experiments including an experimental determination of the viscosity of olive oil. Phil. Trans. R.
Soc., Lond., 177.
(3) Grubin, A. N. and Vinogradova, I. E. (1949) Book No. 30 (Central
Scientic Research Institute for Technology and Mechanical
Engineering, Moscow); Trans. D.S.I.R., (337).
(4) Petrusevich, A. (1951) Fundamental calculations from the contact
hydrodynamic theory of lubrication (in Russian). Izv. Akad. Nauk
SSSR, Otd. Tekh. Nauk, 2.
(5) Crook, A. W. (1961) Elastohydrodynamic lubrication of rollers.
Nature, 190.
(6) Dowson, D. and Higginson, G. R. (1963) The theory of roller bearing lubrication and deformation. In Proceedings of IMechE Convention on Lubrication and Wear, London.
(7) Blok, H. (1937) Surface temperature under extreme pressure conditions. In Proceedings of Second World Petroleum Congress,
Paris.
(8) Blok, H. (1970) The postulate about the constancy of scoring
temperature. In Interdisciplinary Approach to the Lubrication of
Concentrated Contacts, SP-237 (NASA).
(9) Snidle, R. W., Rossides, S. D., and Dyson, A. (1984) The failure of
ehd lubrication. Proc. R. Soc., A395.
(10) Martin, H. M. (1916) The lubrication of gear teeth. Engineering,
August.
(11) Johnson, K. L. (1970) Regimes of elastohydrodynamic lubrication.
J. Mech. Engng Sci., 12.
(12) Dowson, D. (1970) Elastohydrodynamic lubrication. In Interdisciplinary Approach to the Lubrication of Concentrated Contacts.
SP-237 (NASA).

Machine Elements in Rolling Contact

315

(13) Archard, J. F. and Kirk, M. I. (1961) Lubrication of point contacts. Proc. R. Soc., A261.
(14) Greenwood, J. A. and Williamson, J. B. P. (1966) Contact of
nominally at surfaces. Proc. R. Soc., A295.
(15) Bowden, F. P. and Tabor, D. (1954) Friction and Lubrication of
Solids, Part I (Oxford University Press, London).
(16) Stolarski, T. A. (1979) Adhesive wear of lubricated contacts.
Tribology Int., 12.
(17) Stolarski, T. A. (1989) Probability of scufng in lubricated
contacts. Proc. Instn Mech. Engrs, Part C, J. Mech. Engng Sci.,
203, 361369.

This page intentionally left blank

Chapter 8
Non-metallic Rolling Contacts

8.1 General considerations


The issue of fatigue in engineering materials has been recognized as an
important problem for the last 200 years. It came into prominence when
the steam engine revolutionized transportation and fatigue in rolling
stock axles became a serious problem. Thus, the stage was set for the
scientic and applied study of fatigue in ferrous materials, which has
been continued since then. A particularly strong incentive to research
on fatigue in metals has been given by the advent of steel hulls, crude
oil pipelines, and jet aircrafts.
Fatigue can be dened as the loss of strength or other important
properties by a material as a result of stressing uctuating in magnitude
and direction over a period of time. Alternatively, fatigue can be
described as the ultimate failure of a material or component at the
application of a varying load whose maximum amplitude is insufcient
to cause failure. Various atomic or molecular processes may take place
during the fatigue phenomenon. Some of them may be benecial and
some deleterious. The deterioration usually prevails over the strengthening and failure, generally as a result of stresses that are small in comparison with those required for failure resulting from a static stress to
occur. In the case of certain materials, there may be a clearly dened
so-called endurance limit. This is a limiting low stress below which failure does not occur within any practical time. Failure may also be
dened in terms of the loss of functionality owing to the inability to
meet certain design criteria, such as required strength, stiffness, or integrity of shape.

318

Rolling Contacts

Fatigue failure in polymers is controlled by a number of competing


factors, such as loading conditions, material structure and morphology,
composition, and time. It is therefore important that the engineer
responsible for material selection knows the fatigue characteristics of
polymers and understands the effects of loading and variables of materials. This is simply dictated by the nature of the fatigue process and the
much more complex structure of polymers compared with metals. Cyclic loading is a characteristic feature of rolling contacts. For example,
teeth in a polymer spur gear are subjected to loadingunloading every
revolution, although the torque applied is constant. The same takes
place in the case of a rolling contact bearing with inner and outer rings
made of polymeric material.
In material selection and component design for rolling contact applications, the empirical characterization of fatigue in polymers is useful
but not sufcient. Growing sophistication in these functions requires a
more fundamental understanding of the mechanisms underlying the
overall fatigue process. Although curves of stress versus number of load
cycles are helpful when selecting a material for a particular application,
they usually tell nothing about the initiation and propagation of crazes
or cracks, yielding, and drawing, or the storage, release, and dissipation
of energy within the rolling contact. The state of the art in the fatigue
of polymers is not as well advanced as is the case with ferrous materials.
The same is true with fatigue of polymers resulting from rolling contact
conguration.
Another class of materials of interest to engineers and designers is
that comprising technical ceramics. This is mainly due to demands on
the load-bearing rolling contacts in all kinds of machinery to operate
at high speeds, hostile environments, increased unit loads, and restricted
lubrication. The design and manufacture of such contacts is at the limit
of established technology. Ceramics as materials for rolling contacts
show some considerable practical advantages over traditional bearing
steels. The properties of ceramics, specically low density and high stiffness, are of most interest to gas turbines and machine tool manufacturers. High hardness, a low coefcient of thermal expansion, and a
high temperature capability are all properties most desired by rolling
contact designers. Research over the past three decades on structure,
quality control, and manufacturing technique of ceramics has produced
materials of high quality and structural integrity. At present, however,
there is rather little reliable information on the load-carrying capacity,
wear, surface fatigue, and failure modes of ceramic materials when used
in rolling contact applications.

Non-metallic Rolling Contacts

319

8.1.1 Approaches to polymer fatigue


A number of major problems might be identied as being responsible
for hindering the understanding of fatigue processes in polymers. The
traditional continuum-mechanical approach to fracture developed by
physicists and mechanical engineers excludes atomic and molecular processes taking place during polymer fatigue. The ultimate fracture is usually accompanied with large-scale, irreversible, and non-linear
deformations. On the other hand, polymer scientists are accustomed
to measure small-scale, reversible, and linear deformations in order to
characterize the effects of the molecular properties and composition of
a polymer. Besides, the exact nature of the competitive processes that
are active during fatigue and are responsible for the damaging effect of
cyclic loading is not sufciently known. Thus, in order to progress,
it seems necessary to accept the duality of continuum and molecular
descriptions as representing two different aspects of the same reality
and to investigate the limits of using a linear approximation to model
non-linear systems such as polymers.
8.1.2 Loading conditions in rolling contact
One of the promising applications of polymers is in rolling contact bearings. In practical circumstances, the contact between the rolling
elements and the rings consists more of sliding than of actual rolling.
The condition of no interfacial slip is seldom achieved because of material elasticity and geometric factors. Moreover, inherent to the state of
loading on rolling elements and inner and outer rings is uctuating
load, although the external load applied to a bearing treated as a system
is static.
The fatigue life of a rolling contact bearing is a function of a number
of factors which are interwoven in a highly complex manner. The effect
of increasing speed on the fatigue life, for instance, is mainly manifested
in the operating load due to the centrifugal force, with a corresponding
reduction in the loading zone. The contact angles also change, that at
the inner raceway increasing and that at the outer ring decreasing with
rising speed.
Another common example of a nominally rolling contact are gears.
The fatigue failure of toothed gearing may occur in any one or a combination of three basic models, i.e. (i) tooth bending, which, if continued
for a sufcient number of cycles, results in fracture at the root, (ii)
surface distress as a result of mesh stresses exceeding the compressive
fatigue limit in some localized areas of the face, and (iii) pitting resulting
from rollingsliding action in the contact region of two meshing teeth.

320

Rolling Contacts

8.2 Phenomenology of polymer fatigue


Polymeric materials subjected to strong mechanical and environmental
excitation show, like many other materials, gradual deterioration in
their performance including eventual failure. If the changes in properties are mostly due to chemical reactions, then the prevailing mode of
deterioration is corrosion or radioactive degradation. When, however,
the deterioration in material properties is caused by repeated cyclic or
random application of mechanical stresses, then the resulting damage
is of the fatigue type.
Polymers and their composites exhibit a much more complex behaviour when subjected to fatigue loading than ferrous materials. The net
effect of the several competing processes depends on a number of factors, which include the temperature, time, environment, and basic molecular properties of the polymer. The most important factors
determining the fatigue behaviour of a polymer are:
thermal effects during the loadingunloading cycle,
morphological changes within the polymer,
transition phenomena,
molecular characteristics (molecular weight, thermodynamic state),
chemical changes (degradation of bonds),
homogeneous deformations,
inhomogenous deformations.
It can be said that the deformation and ow in a polymeric material
depend to a large extent on the characteristics listed above. However,
the molecular structure, molecular weight, composition, and morphology are especially important in this respect. Polymers in a state of
stress are, almost invariably, in a non-equilibrium state, an equilibrium
state or steady state response being attained only under special test
conditions. In fact, the mechanical properties, fatigue included, desired
of polymeric materials in most engineering applications result in large
measure from their non-equilibrium response to the applied stress.
At elevated temperatures resulting, for instance, from frictional heating within a rolling contact zone, polymeric chains continuously
undergo random changes in conguration. The rate of such diffusional
processes, which depends strongly on temperature, is the dominant
factor that affects the response to stress.
Essentially, all available information on polymer fatigue is of an
empirical nature. General data on the fatigue of engineering polymers

Non-metallic Rolling Contacts

321

Table 8.1 Fatigue strength (MPa) for common engineering


polymers

>28 MPa at 104 cycles


Polyimides
Phenolics, moulded
Polysulfones
Acrylics
Polycarbonates
Nylons

<28 MPa at 104 cycles


38
34
28
49
31.5
49

Polystyrenes
Vinyls, rigid
Acetals
Polyethers
Polypropylene
Polyethylene

14
21
24.5
21
21
11

are arbitrarily summarized in Table 8.1 on the basis of whether the S


N curve passes above or below the point marked by 28 MPa and 104
load cycles.

8.2.1 Physical states of stressed polymers


At elevated temperature a polymer is either an elastic solid or an elastic
liquid, depending on whether or not it is crosslinked. When a constant
stress is applied to an elastic solid, the specimen rapidly attains a constant deformation; when the stress is removed, the specimen returns
rapidly to its initial shape. This is not exactly true since many elastomers, even at elevated temperature, rst deform rapidly under a constant stress but then creep very slowly, even when the network is
chemically stable. Under a constant stress, a non-crosslinked polymer
undergoes ow much like a low molecular weight liquid. However,
some energy is stored elastically, because the molecular chains have
been deformed, on average, from their most probable congurations.
Hence, when the stress is removed, some elastic recovery occurs and
therefore the material is called an elastic liquid because energy is both
stored and dissipated during ow.
The volume of an amorphous polymer decreases in a near-linear way
with decreasing temperature (Fig. 8.1). Thus, the free space available to
the molecular chains is reduced and, consequently, it is more difcult
for congurational changes to occur or, in other words, the mobility is
less.
With progressive decrease in temperature, the mechanical properties
become highly time dependent (viscoelastic) and change from elasticlike to leathery. In some narrow temperature range, long-range congurational changes take place only very slowly, and below a temperature called the glass temperature, Tg , long-range motions become frozen
and only short segments and side groups continue to execute thermally

322

Rolling Contacts

Fig. 8.1

activated motions. This is demonstrated by the thermal expansion coefcient, which is roughly 3 times greater above Tg than below. Below
Tg , the polymer is called a glass, although glassy polymers usually show
some form of ductility.
Semi-crystalline polymers have both a glass temperature and a crystalline melting point (Fig. 8.1). When such a polymer is slowly heated,
the volumetemperature relation indicates that melting occurs over a
temperature range. The temperature Tm shown in Fig. 8.1 is that at
which the largest and most perfect crystallites melt.
Above Tm , a crystalline polymer has properties similar to an amorphous polymer and at temperatures considerably above Tm it is either

Non-metallic Rolling Contacts

323

an elastic solid or an elastic liquid, depending on whether it is


crosslinked or not. In the range TgFTFTm , a semi-crystalline polymer
consists essentially of crystalline domains and amorphous material
intermixed together. Below Tg , both crystallites and amorphous glassy
material exist.

8.2.2 Response to applied stress


Under applied surface traction, an elastic solid undergoes, in general, a
change in both shape and volume. According to the classic theory of
elasticity, a measure of the resistance to change in shape at constant
volume for a homogeneous isotropic material is the shear modulus, G.
Similarly, the measure of resistance to volume change at constant shape
is the bulk modulus, K. The resistance to a change in length is the
tensile or Youngs modulus, E. Because such deformation usually
involves both shape and volume changes, the tensile modulus is a function of G and K. The three moduli are comprehensively dened in Fig.
8.2. The reciprocal of each modulus is a compliance.
Linear elastic behaviour in either shear, tension, or bulk can be represented schematically by a spring. The spring symbolizes that stress is

Fig. 8.2

324

Rolling Contacts

proportional to strain, the response to stress is instantaneous, no permanent deformation occurs, and the energy to deform the spring is
stored completely, i.e. there is no dissipation.
A Newtonian uid is usually represented by a dashpot for which the
stress is proportional to the shear or strain rate; no elastic recovery
occurs when the stress is removed and the energy affecting ow is
entirely dissipated.
Polymers are classied as viscoelastic materials. The energy required
to deform a viscoelastic material is partially stored and partially dissipated, and therefore it exhibits characteristics of both an elastic solid
and a liquid. Under sufciently small deformations, the behaviour is
linear, although only special tests can reveal whether or not the viscoelastic response is linear.
There are a number of mathematical ways to represent the linear
viscoelasticity. One of them is a mechanical model consisting of arrays
of linear springs and linear dashpots. The generalized Voight model
consists of a large number of Voight elements connected in series as
shown in Fig. 8.3. A Voight element is constructed from a spring and
a dashpot in parallel. The model often used to represent the response
of a polymeric material to a prescribed straintime history is the generalized Maxwell model shown in Fig. 8.4. This model consists of a
large number of Maxwell elements connected in parallel. Each Maxwell
element represents a spring and a dashpot connected in series.

8.2.3 Phenomenological description of fatigue


In fundamental terms, fatigue is due to irreversible processes that take
place when a cyclic load is applied to a polymeric material. The extent
of fatigue damage and its importance is dependent primarily on the
stress level at which irreversible damage occurs relative to the stress for
complete failure. Figure 8.5 shows two materials with different stress
strain curves. One material is almost perfectly elastic to fracture,
whereas the other undergoes viscoelastic ow at about 0.3 . The elastic
material will be insensitive to alternating stresses below , and the
stress versus number of load cycles curve will be almost at, pointing
to a very good fatigue resistance. The other material will start to deform
at relatively low stresses and fatigue damage will develop continuously.
The amount of damage will increase with increasing load and the material will have a poor resistance to fatigue.
When this simple principle is applied to composite materials, it is
clear that the response depends on the reinforcing bre arrangement
and volume fraction as well as on the matrix and the bre properties.

Non-metallic Rolling Contacts

325

Fig. 8.3

All these factors determine the way that the load is distributed between
the bre, matrix, and brematrix interface. For example, in an aligned
carbon bre material the load is carried almost entirely by the bres
and little irreversible damage occurs until bre fracture is initiated. Subsequent unloading and loading cycles result in a small redistribution of
the load in the region of the broken bres, and some fatigue damage
may develop. This will occur only at stresses close to ultimate fracture,
so that the fatigue resistance is considered as very good.
In a random glass bre material, microcracking by debonding occurs
in transversely oriented bre bundles at, usually, relatively low stresses,
and this marks the onset of clearly dened irreversible damage. Cracking may be preceded by resin ow. During fatigue cycling the transverse

326

Rolling Contacts

Fig. 8.4

Fig. 8.5

cracks propagate and this eventually leads to resin cracking, which can
occur at stresses well below those observed at monotonic loading.
Although the nature of the irreversible processes depends on the composite material and the loading conditions, the principle of progressive
fatigue damage can be understood by reference to the simple example
illustrated in Fig. 8.6. The sketch represents part of a bre bundle or
lamina oriented with the bre axis normal to the applied load. The
bundle or lamina is constrained by adjacent bundles, so that the growth

Non-metallic Rolling Contacts

327

Fig. 8.6

of a crack under monotonic loading occurs only under increasing strain


conditions. Suppose that the bundle is loaded by a cyclic load as shown
in Fig. 8.7. In the rst half-cycle, the deformation and fracture processes
that occur depend on the stress amplitude and presumably the initial
response is linear. As the stress increases, non-linear effects arise owing
to the viscoelastic properties of the resin and, at a later stage, to
debonding and resin cracking. Thus, at very low stress amplitudes,
when the response is fully elastic, fatigue damage will not develop.
When the stress amplitude is increased, viscoelastic ow occurs
preferentially between closely spaced bres because of the high strain
magnication in these regions. This ow will not be fully reversible
when the stress is reduced, and during cyclic loading additional stress
and strain concentrations develop, which lead to the initiation of

328

Rolling Contacts

Fig. 8.7

debonding at applied stresses below those observed in monotonic loading (Fig. 8.6). The debonding cracks grow during cyclic loading because
some ow occurs at the crack tip during the loading half of the cycle,
which is not fully reversed on unloading. As in uniaxial tensile tests,
the cracks nucleate and propagate in regions of closely spaced bres by
the growth and coalescence of individual bre debonds. When the bres
are widely spaced, the growth of the crack from one bre to the next
depends on the resistance to fatigue crack growth of the matrix.
It is obvious from this rather simplied presentation of the mechanism that the fatigue properties will depend on the temperature and the
cyclic loading frequency, since both these factors affect the amount of
matrix ow. An important additional effect is viscous dissipative heating during cyclic loading, leading to a rise in temperature which can be
in the region of 2550 C, depending on loading frequency. The magnitude of the rise depends on the specimen geometry and the efciency
of heat dissipation to the surroundings. Thus, carbon bre composite

Non-metallic Rolling Contacts

329

materials show lower temperature rises because of the relatively high


thermal conductivity of the bre. A similar mechanism can be used to
explain the fatigue behaviour of more complicated bre arrangements.
Progressive damage may involve intralaminar and interlaminar processes, and the rate of damage build-up depends on the effective stresses
causing different forms of microdamage.
A common feature of the fatigue failure of composite materials,
especially when the element is uniformly stressed, is the occurrence of
damage over a large volume of the material. The large-scale damage
usually produces three important effects: (i) the modulus of elasticity
decreases progressively during the fatigue life; (ii) the hysteresis or damage loop during cyclic loading becomes progressively more pronounced;
(iii) the residual strength of the material decreases progressively with
the number of cycles.
In the case of a rolling contact conguration, the failure almost
exclusively develops from local regions of high stress or strain associated with the nature of the contact. The micromechanism of failure in
these regions will be the same as that described above.

8.3 Behaviour of polymers in rolling contact


At rst sight, engineering polymers may seem to be rather unlikely
materials for rolling contact applications such as bearings, gears, etc.,
because of the generally accepted premise that polymers are weak and
soft. However, modern engineering polymers and their composites have
physical and mechanical properties that can be considered as very
attractive for rolling contacts. The main benets resulting from using
polymers in rolling contacts are as follows:
(a) corrosion resistance, as many polymers are considered to be chemically inert and capable of operating in environments hostile to
ferrous materials;
(b) the ability to operate without lubrication or to be lubricated with
process uids;
(c) many potentially useful polymers are available at a much lower
price than traditional engineering materials;
(d) ease of processing and cost of manufacture make polymers the most
competitive material for rolling contacts in underwater, marine,
chemical, and processing industries;
(e) injection moulding and extrusion eliminate the need, and therefore
lower the cost, of nishing and other surface treatments, since

330

Rolling Contacts

elements and components manufactured with the help of these techniques come nished to dimensions and ready for assembly. For
example, moulded toothed gears or components for rolling contact
bearings can be produced at a very low unit cost.
These obvious benets can be readily obtained provided that the application is characterized by light loads and low to moderate speeds. This
is not only because of the fatigue strength of polymers but also because
of the thermal softening and general decrease in mechanical strength at
elevated temperatures.
As mentioned earlier, three principal mechanisms may contribute to
the volume fatigue of polymeric materials. Although this view is now
commonly accepted, the case of surface fatigue of polymers, as manifested in rolling contact, probably requires a modied approach.
Despite the lack of sufcient understanding of the surface fatigue of
polymers, they have been successfully used as a material for gears and
rolling contact bearings.

8.3.1 Characteristics of rolling contact conditions


Pure or free rolling is the most basic form of rolling motion and is
probably most nearly approached in the case of a cylinder or ball rolling
without constraint in a straight line along a plane. Since all bodies are
made from deformable materials, some deformation must ensue when
they come into contact under load. The shape and size of the area of
contact will depend upon such factors as the individual geometry, the
load, and the deformation characteristics of the materials.
Most of the studies on the mechanism of rolling friction have been
concerned with perfectly elastic materials, and results for rolling contact
of polymers are relatively scarce. Basically, there are two scales of size
in the problem related to the contact of real engineering materials:
(a) the nominal contact dimensions and compressive deformations,
which could be calculated by the Hertz theory;
(b) the height and spatial distribution of surface asperities.
For the situation to be amenable to quantitative analysis, these two
scales of size should be very different. In other words, there should be
many asperities lying within the nominal contact area. When the two
bodies are pressed together, true contact occurs only at the tips of the
asperities, which are compressed as elastic or viscoelastic solids.
In the case of rolling contact of elastic materials, the process
consists of two main effects. The rst, which is usually predominant

Non-metallic Rolling Contacts

331

during the early stages of rolling, is primarily concerned with the


plastic displacement of material from the path of the rolling element,
and the resistance to rolling is essentially due to the work of plastic
deformation of the material. The second effect predominates after
repeated traversals of the same track; plastic displacement gradually
comes to an end and the deformation becomes primarily elastic. The
primary source of the rolling resistance is elastic hysteresis losses within
the contacting materials.
The situation is quite different in the case of rolling contact
between a hard and smooth cylinder and a viscoelastic material.
Tomlinson (1) explained his experimental observations in terms of
molecular adhesion between lightly loaded rolling surfaces. According
to him, the surface atoms are pulled away from their equilibrium
positions until the displacement exceeds a certain distance; they then
ick back to their old equilibrium positions and, in the process, energy
is dissipated. Although this theory was not universally accepted, it is
quite clear that the idea of the contribution of interfacial effects to
the friction in rolling contacts between viscoelastic materials should be
seriously considered.

8.3.2 Thermodynamic equilibrium in rolling contact


A system consisting of two bodies in contact over an area A is considered here and is shown schematically in Fig. 8.8. It can represent,
for instance, a microcontact created by a pair of asperities located on
the surfaces of bodies in rolling contact. Adhesive junctions formed by
the contacting asperities are ruptured due to tension exerted on them
by the rolling motion. The system can exchange work and heat with the
surroundings but no matter. A force P, representing the tensile load on
the contact and resulting from the rolling motion, can be applied either
by a xed load, as in Fig. 8.9(a), or by a more complex loading system
as shown in Fig. 8.9(b). It must be emphasized, however, that the area
of contact, A, is allowed to vary so that the geometry of the system can
change and linear elasticity must be excluded. Also, variation in A may
be accomplished independently of the load P, or of the elastic displacement , so that the state of the system depends, in general, on two
independent variables: P and A or . The area of contact can decrease
at xed P or xed , until separation of the two bodies occurs, corresponding to the rupture of a joint under xed load or xed grip conditions. The reduction in the contact area will be considered here as
crack propagation in mode I, i.e. the opening mode. The energy of the

332

Rolling Contacts

Fig. 8.8

system shown in Fig. 8.9 is given by


UGU(S, , A)
and is a function of and A besides the entropy S. It can be divided
into elastic energy Ue and interfacial energy Us . The interfacial energy
is solely a function of A and can be written as
Us G( 1C 2A 12 )AGwA

(8.1)

where 1 and 2 are the surface energies of the two bodies in rolling
contact; 12 is their interfacial energy and w is Dupres energy of
adhesion or the thermodynamic work of adhesion. The rst differential
of the energy is
U

dUG

,A

dSC

S,A

d C

S,

dA

(8.2)

Non-metallic Rolling Contacts

333

Fig. 8.9

or
dUGT dSCP d C(GAw) dA
where
U

G
S,A

Ue


Ue

Us

A C A

G
S,

GP
S,A

S,

S,

GGAw

(8.3)

334

Rolling Contacts

and G denotes the variation in elastic energy with A at constant and


is called the strain energy release rate.
The three equations of state
SGS(, A)
PGP( , A)
GGG( , A)
contain the same information as the fundamental equation UG
U (S, , A). For equilibrium under various constraints, the following
thermodynamic potentials can be used:
(i) at constant temperature, the Helmholtz free energy
FGUATS

(8.4)

dFGS dTCP d C(GAw) dA

(8.5)

(ii) at xed load, the enthalpy


HGUAP

(8.6)

dHGT dSA dPC(GAw) dA

(8.7)

(iii) at xed load and temperature, the Gibbs free energy


BGUATSAP

(8.8)

dBGS dTA dPC(GAw) dA

(8.9)

The problem can be simplied by assuming that thermal effects are


negligible, that is TG0 and SG0. Thus
dFGdUGd(UeCUs )GP d C(GAw) dA

(8.10)

dBGdHGd(UeCUsCUP )G dPC(GAw) dA

(8.11)

noting that UP GP is the potential energy of load P.


Appropriate relations can be written for GAw, leading to
Ue

Ue

UP

A G A C A

GG

Furthermore
G

GA
A

(8.12)

Non-metallic Rolling Contacts

335

and

P G A
A

Equilibrium at xed load conditions, that is, dPG0, corresponds to an


extremum of F or U, while equilibrium at xed grip conditions, that is,
d G0, is equivalent to an extremum of B or H. In either case, equilibrium is given by
GGw
This equilibrium relation links two of the three variables, namely , A,
and P, of the equation of state, so that the equilibrium curves (A),
A(P), and P( ) are a function of w. If GGw, the area of contact will
spontaneously change so as to lower the thermodynamic potentials. If
GFw, it is clear that A must increase and the crack recedes. Conversely,
if GHw the area of contact must decrease to have dFF0 or dBF0, and
the crack extends. Here, G dA is the mechanical energy released when
the crack extends by dA. The breaking of interfacial bonds requires an
amount of energy w dA and the excess (GAw) dA is converted into
kinetic energy, if there is no dissipative factor. Parameter G is often
called the crack driving force, but, strictly speaking, the crack driving
force is GAw, which is zero when GGw at equilibrium.
The equilibrium can be stable, unstable, or indifferent. A thermodynamic system under a given constraint is stable if the corresponding
thermodynamic potential is a minimum, i.e. if its second derivative is
positive. Stability at xed grip conditions (d G0) corresponds to
G

A H0

(8.13)

and stability at xed load conditions, dP G0, corresponds to


G

A H0

(8.14)

The stability problem can be studied with an experimental set-up having


a nite stiffness, km. Figure 8.9 is a schematic representation of such
a machine, and km is the stiffness characteristic of the coil spring. A
displacement can be obtained by turning the screw, which is thermodynamically equivalent to placing a load P on to the spring and then
clamping it without external work, and is divided into elastic displacement, m , of the spring and elastic displacement, , of the two solids

336

Rolling Contacts

in contact. The spring exerts a force PGkm m on the two bodies in


contact, and therefore
G C m G C

(8.15)

km

The stability of the system involving the two elastic bodies in contact
and the spring at xed crosshead displacement, , can be studied. The
energy of the system includes the elastic energy, Um, of the spring. Thus
UGUe(A, )CUm ( m )CUs (A)

(8.16)

and its rst differential is


Ue

Ue

A dAC

dUG

dUm
dUs
d C
d mC
dA
d m
dA
A

GG dACP d CP d mAw dA
GP dC(GAw) dA

(8.17)

Equilibrium at xed is still given by GGw, and the stability by


G

A H0

(8.18)

but this stability depends on the stiffness, km , of the apparatus. Intuitively, the spring can be seen as a reservoir that provides energy for
crack propagation at constant . It is of interest to calculate (GA)
as a function of (GA) by considering G[A, ( , A)] as a function of
G[A, (, A)]. Thus
G

A GA C A

(8.19)

By differentiating the expression for and rearranging the last equation, the following is obtained
G
G
P
G
A
A
A
A


The quantity
P

Gk
A

kmC

(8.20)

Non-metallic Rolling Contacts

337

is the stiffness of the two elastic bodies in contact, and is positive.


Therefore, (GA) can be zero while (GA) is still positive. It can
therefore be concluded that the stability range monotonically increases
with the stiffness from the xed load case (km G0) to the xed grip case
(km GS). When km 0, the xed load conditions are approached.
Estimation of the adherence or the normal load that the contact
formed by a pair of surface asperities can support is important for the
assessment of the rolling friction. Knowing the statistical properties of
a rough surface and the mechanism and the force required to cause the
junction failure, an estimate of the desired force may be made. Details
of analysis of this important problem can be found elsewhere (2).

8.3.3 Mechanics of polymer rolling contact


The stress in polymeric materials used for rolling contacts is inuenced
by the rate of strain and, therefore, the contact stress and deformation
will depend upon the speed of rolling. The simplest way to account for
time dependent characteristics of a polymeric material is to model it as
a linear viscoelastic material. This has been discussed in an earlier section in relation to the general response of polymers to applied stresses.
However, application of the linear theory of viscoelasticity to rolling
contact is not simple since the situation is not one in which the viscoelastic solution can be obtained directly from the elastic solution. It is
not difcult to appreciate the reason for that. During rolling, the material located in the front half of the contact is being compressed, while
that at the rear is being relaxed. With perfectly elastic material the
deformation is reversible, so that both the contact area and the stresses
are symmetrical about the centre-line. Viscoelastic material such as a
polymer, however, relaxes more slowly than it is compressed, so that
the two bodies in contact separate at a point closer to the centre-line
than the point where they rst make contact. This is illustrated in Fig.
8.10, where z1Fz2 and recovery of the surface continues after contact
has ceased. The geometry of rolling contact of a polymeric material is
different from that of the perfectly elastic case, and therefore the viscoelastic solution cannot be obtained directly from the elastic solution.
Moreover, the point at which separation occurs, that is, xGz2, cannot
be dened in advance. Usually, it has to be located where the contact
pressure drops to zero.
In what follows, a one-dimensional model of the contact between
polymeric material and a rigid cylindrical body of radius R (Fig. 8.10)
will be presented (3). The polymer will be modelled by a simple viscoelastic foundation of parallel compressive elements that do not interact

338

Rolling Contacts

Fig. 8.10

with each other. The rolling velocity is V and the cylinder makes rst
contact with the polymer substrate at xGz1. Since there is no interaction between the elements of the foundation, the surface does not
depress ahead of the roller. Assuming that z1[R, the compressive strain
in an element of the foundation at x is given by

Ax2 1
2R h

(8.21)

where z1 xz2 and is the maximum depth of penetration of the


roller.
In the case of a perfectly elastic foundation characterized by a modulus K, the contact would be symmetrical, that is z1 Gz2, and the stress
in each element would be K . The pressure distribution under the
roller and the total load would then be given by equations for elastic
contact proper. For a viscoelastic material the elastic modulus, K, is
replaced by a relaxation function (t). Thus, the stress in the viscoelastic element located at x is given by
t

(tAt)

p(x, t)G G

(t)
t

dt

(8.22)

At steady rolling tGVx, and, therefore, from the equation for


the strain

Vx
Rh

(8.23)

Non-metallic Rolling Contacts

339

Substituting this in equation (8.22) expressing stress in a viscoelastic


element and changing the variable from t to x gives
1
p(x)GA
Rh

x(xAx) dx

(8.24)

Further analysis requires specication of the relaxation function for the


polymer. One possibility is to use a simple delayed elasticity model for
which the relaxation function is given as
(t)GK(1C e t/T )

(8.25)

The material described by the relaxation function [equation (8.25)] has


a dynamic modulus K (1C ). Under static conditions the modulus is
K, and T denotes the relaxation time. Combining the relaxation function with equation (8.24) expressing stress in a viscoelastic element gives
p(x)G

x2
Ka2 1
x2
1A 2 A 1C 2 C (1C )[1Ae (1Cx/z1)/ ]
Rh 2
z1
z1

(8.26)
where GVTz1 represents the ratio of the relaxation time of the material to the time taken for an element to travel through the distance equal
to half the contact width. The contact pressure is equal to zero at xG
z1 and becomes zero again at xGz2.
The normal load is obtained from the expression
z2
Ka3
Fp ( , )
p(x) dxG
(8.27)
PG
Rh
z1

Because the pressure is distributed asymmetrically, the rolling motion


is resisted by a moment given by

z2

xp(x) dxG

MG

Ka4

z1

Rh

Fm ( , )

(8.28)

Finally, the coefcient of rolling friction can be expressed as

r G

M
PR

a
R

Fr ( , )

(8.29)

Examination of the above equations reveals that at slow rolling speeds,


when the contact time is long by comparison with the relaxation time
of the material, the pressure distribution and load are quite close to
those for a perfectly elastic material with modulus K. At the same time,
moment M is near zero.

340

Rolling Contacts

At high speeds, however, the pressure distribution and load are close
to the elastic contact results but with a dynamic foundation modulus
K (1C ). It is obvious that the relaxation effects play an important role
in the behaviour of a contact only when the contact time is approximately equal to the relaxation time of the material. Only then does
the contact become appreciably asymmetric and a maximum friction
moment arise.

8.3.4 Fatigue considerations


The characteristic feature of polymeric materials that largely denes
their fatigue behaviour is that they, unlike metals, are inhomogeneous
on a gross scale and anisotropic. They tend to accumulate damage in a
general rather than a localized way. Moreover, failure does not usually
occur by propagation of a single macroscopic crack. The mechanisms of
damage accumulation, including bre and matrix cracking, debonding,
transverse cracking, and delamination, occur sometimes independently
and sometimes interactively, and the predominance of one or other of
them may be strongly inuenced by both the material variables and the
testing procedure and conditions.
At the early stage of the life of a composite material subjected to
cyclic loading, one can expect damage of a very general type only. It
will usually be distributed throughout the stressed region and will not
have any immediate effect on the strength, although it could reduce the
stiffness. The slight reduction, if any, in strength in the early stage of
life might be compensated for by an opposite process leading to a slight
increase in strength. This increase in strength of a composite material
may arise from an improved bre alignment allowed for by small,
stress-induced viscoelastic deformations in the matrix. With the passage
of time, the amount of accumulated damage in a certain region may be
sufciently large as to reduce the load-bearing capacity of the composite
in that region to the level of the peak applied load in the fatigue cycle.
This inevitably leads to a gross failure. When the gross failure is a result
of a gradual process, it is referred to as degradation. On the other hand,
catastrophic failure is usually termed sudden death. Changes of this
type are not always related to the propagation of a single crack, and
this must be remembered when an attempt is made to interpret fatigue
data obtained during testing of composite materials by methods
specically developed for metallic materials.
The presence of a crack in a composite material does not necessarily
mean that it will propagate under cyclic loading, since crack propagation depends, to a signicant extent, on the nature of the composite

Non-metallic Rolling Contacts

341

material. In glass bre polymer laminates containing woven reinforcement, crack tip damage may remain localized by the complex geometry
of the bre array, and the crack may proceed through this damaged
zone in a fashion analogous to the propagation of a crack in plastically
deforming metals.
One noticeable feature of fatigue test results is their variability. This
variability stems not only from the statistical nature of the progressive
damage that leads to fracture of a composite, but more specically from
the highly variable quality that is usually found in many commercial
composite material. Another important problem is associated with the
denition of the criterion of failure of a composite material. A traditional denition of failure in the fatigue testing of metals states that
failure occurs when a complete separation of the broken halves of a
sample occurs. This can be quite meaningless in the case of a composite
material when the sample has lost its shape integrity and its ability to
sustain an applied stress as a result of extensive resin cracking.
The fatigue strength of many composite materials is strongly affected
by the loading frequency since it has a direct inuence on temperature
increases caused by hysteresis losses. Heat dissipation by conduction is
virtually impossible in many reinforced polymers. In polymeric materials, even quite small temperature increases may lead to signicant
changes in the mechanical properties, while excessive heating will certainly cause thermal degradation. The response of composite materials
to cyclic loading depends also on the size of the tested sample. This is
especially evident when the inhomogeneity resulting from the distribution of the reinforced agent is so great that it inuences the fatigue
response of a sample whose size is comparable with the scale of the
inhomogeneity. Therefore, in laminates with a woven structure, the
width of the testpiece should be sufcient to include several repeats of
the weaving pattern. In a hybrid laminate, it is important to make sure
that a test sample is an order of magnitude wider than the width of
individual bre tows in the composite.
The role played by holes and notches in the fatigue of a polymer
composite is still a matter of discussion. There are experimental data
demonstrating that sharp notches are more harmful than drilled holes,
although notches in general have little effect on the fracture strength
owing to the large number of debonding sites present in the material.
On the other hand, it is believed that holes are fully effective in initiating
fatigue damage, but they do not necessarily affect the nal failure.
The ability of polymer composites to arrest a crack that arises directly
from their inhomogeneity on a ne scale, i.e. the interface between bre

342

Rolling Contacts

and matrix, as well as on a large scale, makes it very often difcult to


apply fracture mechanics principles to fatigue life prediction and design.
However, in composites containing woven cloth or in mouldings containing randomly distributed chopped bres, notch-like defects tend to
propagate in a more normal fashion. In such a case, it is possible to use
fracture mechanics principles in order to predict the crack growth rate.
The damage progressively accumulated in the sample of a polymer composite during cyclic loading will inevitably affect the macroscopic mechanical properties of the material to an extent depending on the
composite geometry and the mode of loading.
The cumulative damage of polymer composites as a result of cyclic
loading has been the subject of many studies. All these studies have
aimed at obtaining a useful working design relationship similar to that
proposed by Miner and Palmgren. For resin cracking during cyclic
loading, the following non-linear damage law, independent of the stress
level, is often used
2

N N

G A

CB

(8.30)

where n is the number of cycles sustained by the composite at a stress


level that would normally produce failure after N cycles, and A and B
are constants (B is negative and A is equal to unity at failure).
The above model is not unreservedly accepted and an alternative one
has been put forward. This is, basically, a single-parameter, stress
independent damage model, where the damage factor, , is given by
G

e kxA1
e kA1

(8.31)

and kG1 and 0FxF1.


The model implies that the rate of damage accumulation, ddt,
depends, to a rst approximation, linearly on time, t, and t in turn is the
non-logarithmic cycle ratio nN. For different values of k this function
produces a family of exponential curves.
It is reasonable to expect an improved fatigue performance from
composites reinforced with high-strength but brittle bres, such as glass
and carbon. This is because the bres should carry the major part of
the load. As cyclic loading continues, however, even small movements
in the matrix, resulting from its viscoelastic nature, can lead to local
redistributions of stress which allow for some random bre damage to
occur by a normal brittle fracture process. Similar damage will also

Non-metallic Rolling Contacts

343

occur in the surroundings of any stress concentration. The damage


accumulation, being rather a function of time than a function of the
number of stress cycles, will build up to some critical level when the
overall strength of a composite is below the applied peak stress and will
lead, eventually, to failure.
When the applied stress is not uniquely uniaxial and aligned with the
reinforcing bres, then more severe loading conditions are imposed on
the matrix and the fatigue strength of a composite material is reduced.
A stress system of compression combined with shear, which is characteristic for rolling contact, is especially harmful to composites containing, for example, aramid bres. In this loading mode, the matrix and
interface would normally be expected to control the fatigue performance of a laminated composite, but not the bre. On the other hand,
the overall fatigue behaviour of hybrid composites ought to be governed
by the extent to which the particular bre mix controls the strain levels
in the composite.
It has been found that the matrix and interface are the weak links
within the polymer composite as far as fatigue strength is concerned.
The results suggest that resins of lower reactivity perform better when
exposed to low-stress cycling load. Any treatment or processing leading
to an improved resistance of the matrix material to crack propagation
will undoubtedly improve the fatigue strength. On the other hand, exposure of reinforced polymers to water and other surfactants often results
in some degree of plasticization of the matrix and weakening of the
interfacial bonds. Therefore, one can expect the fatigue strength of some
reinforced polymers to be environmentally sensitive when the behaviour
is not governed exclusively by the bres.
The fatigue strength of short bre composites is considerably lower
than that of polymers lled with high volume fractions of continuous,
aligned, and rigid bres. This is mainly because the weaker matrix has
to support a much greater proportion of the cyclic loading. Localized
matrix failures, which are easily initiated, can destroy the integrity of
the composite even though the bres remain intact. Since the interfacial
shear stresses probably reverse their direction at each load cycle, the
interface region is especially vulnerable to fatigue crack initiation and
eventual damage. An important cause of damage in reinforced polymers
is thermal degradation resulting from large hysteresis losses and subsequent heat dissipated within the material characterized by low thermal conductivity. A fatigue strength of reinforced nylon as low as half
the fatigue strength of unlled nylon could be found in practical
applications.

344

Rolling Contacts

8.4 Model rolling polymer contact


8.4.1 Experimental setting
The experimental conguration often used to test rolling contact fatigue
of polymers is schematically shown in Fig. 8.11. It consists of three steel
balls in contact with a polymer testpiece in the form of a hemispherically ended pin. Typical dimensions for one particular conguration are
provided in Fig. 8.12. The balls are contained in a polymer cup and are
free to roll. In this way, the contact conditions characteristic of a ball

Fig. 8.11

Fig. 8.12

Non-metallic Rolling Contacts

345

bearing could be imitated since the cup could be taken as representing


the outer race, the three balls as representing the rolling elements in a
ball bearing, and the upper piece as equivalent to the inner race. The
load on the assembly is applied through a piston and a lever arm by
means of dead weights. The upper testpiece is attached to a spindle via
a special collet. The spindle is driven by a shunt-type electric motor, the
speed of which can be continuously adjusted. The test time and
rotational velocity of the upper test piece are monitored. The apparatus
can be set to halt at a preset number of revolutions of the testpiece or,
by using a vibration sensor, when the level of vibration exceeds a preset
value owing to failure within the contact.

8.4.2 Kinematics of the model contact


The geometry of the test assembly in the form required for the
kinematic analysis is shown in Fig. 8.13. The angle denoted by is a
function of the dimensions characterizing the conguration and is given
by (4)
Garcsin

2R CR AR
RceARc
p

(8.32)

where Rc is the radius of the cup race, Rce is the external radius of the
cup, Ra is the radius of curvature of the upper testpiece, and Rp is the

Fig. 8.13

346

Rolling Contacts

radius of the lower ball. The angle denoted by is dened by the


following expression

Garctan

sin(90A )

[R (R CR ) sin ] cos(90A )
p

(8.33)

If the radius of curvature of the upper testpiece is equal to the radius


of the lower ball, then, by symmetry, angle will be equal to angle .
The third characteristic angle of the conguration is given by

G90A( C )

(8.34)

The angular velocity of the lower ball is

pG

a Ra
2(RaCRp )

(8.35)

and its linear velocity is


Vp G

a R2a sin
2(RaCRp )

(8.36)

The spin of the lower ball is equal to


Vs GVAVp

(8.37)

and

(8.38)

(8.39)

Ra
Vs G a Ra sin( ) 1A
2(RaCRp )
Therefore

sG

a Ra sin
Ra
1A
Rp cos
2(RaCRp )

The contact region between the lower ball and the upper testpiece is
characterized by a degree of slip resulting from the complex motion of
the lower ball. Depending on the location within the contact region, the

Non-metallic Rolling Contacts

347

slip is given by
(a) slip at point C is zero,
(b) slip at point C
VC A(VpCVs)
VC
(c) slip at point C
VC A(VpCVs)
VC
The magnitude of the slip within the contact area depends, among other
things, on the elastic properties of the contacting materials. The slip
contributes to such phenomena as heating, softening of contacting surfaces, and, eventually, to sliding wear which must be distinguished from
the wear resulting from surface fatigue produced by the rolling motion.
The number of load cycles experienced by the upper testpiece can be
estimated from the relationship
LGz

2R CR
RaC2Rp
a

(8.40)

where z denotes the number of lower balls in the assembly.


The normal load applied to the contact between the lower ball and
the upper testpiece is given by
NG

196M

(8.41)

3 cos 1

where M is the mass applied to the lever arm of ratio 1:20 and 1 is
the contact angle between the upper testpiece and the lower ball.
Knowing the normal load on the contact, it is possible to estimate
the contact stresses. The peak compressive stress is given by
po G

1
3

6NE 2
3R4

(8.42)

The principal surface plane stress is equal to


po(1C2)
x G y GA
2

(8.43)

348

Rolling Contacts

The contact region dimensions are


aGbG0.88

1
3

NR
E

(8.44)

The maximum shear stress at a depth of 0.638a is

max G13 po

(8.45)

The maximum tensile stress at radius a is

max G13 (1A2)po

(8.46)

8.4.3 Performance of some polymers in rolling contact


In order to illustrate the main features of polymer rolling contact, the
results of experimental testing of some polymers are presented in this
section (5). The performance of two polymers was assessed, namely
nylon 6.6 and acetal. Nylon 6.6 is a popular polymer and has been used
as a material for gears, cams, pipe ttings, and components of domestic
appliances. Acetal is a similarly widely used polymer, especially in the
automotive industry. However, acetal has a better dimensional stability
than nylon 6.6 mainly because of its low moisture absorption under
varying humidity conditions. The mechanical properties of both polymers were as specied by their suppliers. All tests were carried out under
the following conditions: a contact Hertzian stress of 34 MPa and
a rotational speed of the upper testpiece of 400 rmin. Tests were
performed under both dry and lubricated contact conditions.
Typical test results obtained under dry conditions are shown in Fig.
8.14, where the weight loss of material caused by surface fatigue is
plotted against the number of load cycles. A signicant loss of material
was observed for nylon 6.6. The damage was conned to surface layers
and was in the form of deep pits and cracks. The weight loss of the
acetal was considerably less. The surface of the contact path had a shiny
appearance and no apparent damage could be seen.
Under lubrication, the surface fatigue performance of the tested polymers markedly improved as shown in Fig. 8.15. In the case of acetal, no
weight losses were recorded after 216B103 cycles at a load of 34 MPa.
Therefore, the load on the contact was increased to 68 MPa in order to
create conditions under which a measurable loss of material could take
place. The surface of the contact path had the same shiny appearance

Non-metallic Rolling Contacts

Fig. 8.14

Fig. 8.15

349

350

Rolling Contacts

as in the case of dry contact. Nylon 6.6 also showed improved performance under lubrication. However, the results for nylon 6.6 in Fig. 8.15
are for a load on contact of 34 MPa. Clearly, compared with nylon 6.6,
acetal is a much better material for rolling contact applications.

8.5 Technical ceramics in rolling contact


8.5.1 Ceramic materials
Ceramic materials can generally be divided into traditional, structural
engineering, electricalelectronic, and composites. Traditional ceramics
include aluminosilicates, dolomite, and magnesite. They are used for
tableware, whiteware, refractories, etc. Engineering ceramics are much
more advanced materials with a low porosity and good mechanical
properties. Silicon nitride, silicon carbide, alumina, and zirconia are
typical examples of modern engineering ceramics. They are widely used
in engineering applications such as wear parts, cutting tools, engine
parts, and mechanical seals. Ferrite, titanates, and aluminium nitride
are the most popular ceramics used by the electricalelectronics industry
for such applications as insulators, magnets, packages, and superconductors. The main advantages and disadvantages of engineering
ceramics are summarized in Table 8.2.
Monolithic ceramics, also called rst-generation engineering ceramics, include silicon nitride, silicon carbide, zirconia oxide, aluminium oxide, boron carbide, titanium diboride, aluminium silicate,
and aluminium titanate. Second-generation engineering ceramics are
composites. A variety of reinforcements and matrixprocess techniques
are used. The main types are listed in Table 8.3. An example of a composite ceramic is alumina reinforced with SiC whiskers used for cutting
tips.

Table 8.2 Advantages and disadvantages of engineering ceramics


Advantages

Disadvantages

High hardness, strength


Low density
Structural integrity at high temperatures
Resistance to hot corrosion and
oxidation
Resistance to erosion and wear

Low toughness
Low reliability at present
Difcult to machine
Difcult to detect critical aws
High cost at present
Inadequate life at very high
temperatures

Non-metallic Rolling Contacts

351

Table 8.3 Ceramic composites


Reinforcement

Matrixprocess

Particles
Precipitates
Whiskers
Continuous bres

Glass ceramics
Oxidenitride powders
Chemical vapour inltration
Reactive processing

Ceramics used in engineering applications have a distinct stressstrain


behaviour which is different from that of ferrous materials. An illustration of this is shown in Fig. 8.16. Obviously, in the stressstrain
behaviour of ceramics there is no yielding, and failure occurs suddenly
without any warning signs. A key property of a ceramic material is its
toughness. Table 8.4 gives fracture toughness values for a sample of
engineering ceramics and, for comparison, for some metallic materials.
In order to rectify the inherent brittleness of ceramics, much research
effort has gone into the development of ceramic composites. The
current status of ceramic composites is summarized in Table 8.5.

Fig. 8.16

352

Rolling Contacts

Table 8.4 Comparative toughness


Material
Soft steels
Titanium alloys
Aluminium alloys
Hardened alloy steels
Glass
Silicon nitride
Zirconia (PSZ)
Ceramic composite

Kc (MPa m1/2)
100C
4090
2590
1650
1
57
812
20C

Table 8.5 Current status of ceramic composites

Toughness (MPa m1/2)


Composite

Matrix

Partially stabilized ZrO2


ZrO2 particleAl2O3
SiC breglass
SiC whiskerAl2O3
SiC breCVD SiC
SiC whiskermullite
SiC bre weaveAl2O3Lanxide

3
4
1
4
4
2
4

Composite
15
5
520
9
25
4
21

8.5.2 Ceramic bearings


The term ceramic bearing usually evokes the image of a bearing made
completely from ceramics running red hot without lubrication.
Although much of the early work was indeed focused on all-ceramic
bearings for high-temperature applications such as gas turbine engines,
recent development efforts have been concentrated on hybrid ceramic
bearings consisting of steel rings or races with ceramic balls. There is
still interest in all-ceramic bearings for high-temperature applications,
but further development of effective lubrication systems is required. On
the other hand, hybrid bearings in which only the rolling elements are
made of ceramic material are already in service in highly demanding
applications such as machine tool spindles, and their usage is becoming
more widespread. Any ceramic material considered for rolling contact
application must meet a number of basic requirements. Table 8.6 gives
a brief account of what is required.
There are a number of denite performance advantages offered by
hybrid ceramic bearings. They are outlined in Table 8.7.

Non-metallic Rolling Contacts

353

Table 8.6 Material requirements for ceramic bearing


components
Physical properties

Low density
Moderate Youngs modulus
Low thermal expansion

Mechnical properties

High compressive strength


High rupture modulus
Good toughness

Bearing properties

Good fatigue life


Spalling failure mode
Fine, homogeneous,
uniform, no porosity,
no inclusions

Microstructure

Table 8.7 Performance advantages of hybrid ceramic bearings


Higher speeds

Low density reduces centrifugal effects for aircraft


engines and machine tools (grinding)

Increased stiffness

Electrical conductivity
Greater exibility

High elastic modulus increases dynamic stiffness


for machine tools (milling)
Better kinematics reduces heat generation for
machine tools and computer accessories
Reduced vibration and run-out for reading,
tracking, and positioning equipment
Lower thermal expansion for machine tools
(roller bearings)
Improved insulation capabilities
More scope for designers to alter parameters

Reduced noise

Better ball geometry reduces operating noise

Lower operating
temperatures
Higher precision
Wider speed range

Comparison of traditional bearing steels with advanced engineering


ceramics shows that ceramics are offering a set of characteristics distinctly different from that of steels. Table 8.8 is a summary of the most
important quality aspects of the two types of material.
Silicon nitride has been found to have the best combination of physical and mechanical properties for use in rolling contact bearings (6). A
typical set of gures characterizing silicon nitride is shown in Table 8.9
and indicates that this material possesses an optimum combination of

354

Rolling Contacts

Table 8.8 Quality aspects


Bearing steels

Ceramics

Speciedoffered
Manufacture

Composition
Open

Claimed properties
Closed, proprietary

Chemical analysis
Properties modied by

Fast, standard
Heat treatment

Slow, difcult
Additives, densication

Mechanical testing

Standard and
reproducible

Non-standard, large
scatter

Physical characteristics

Conducting and
magnetic
Light, bright
appearance

Insulating and
non-magnetic
Various colours,
low reectivity

Optical characteristics

Table 8.9 Typical data for silicon nitride


Density
Elastic modulus
Compressive strength
Hardness (HV5)
Fracture toughness
Rolling contact fatigue
Failure mode

3.25 gcm3
310 GPa
H3500 MPa
16 GPa
58 MPa m1/2
Good
Spalling

properties for rolling contact applications. The current annual production of silicon nitride balls worldwide is estimated at over 1 000 000,
and this is expected to grow to several million over the next decade.
Angular contact ball bearings tted with silicon nitride balls are the
most common form of hybrid ceramic bearings. Other types of ceramic
balls may be used in special applications such as instrument and gyroscope bearings. Angular contact ball bearings are designed to operate
at high speeds under both radial and axial loads. These bearings have
a larger ball complement than deep groove ball bearings and are usually
tted with reinforced phenolic resin cages. Since an axial load can only
be accommodated from one direction, bearings are usually mounted in
pairs and preloaded to maintain the correct contact angle. Some types
of hybrid bearing involve only replacement of the steel balls with silicon
nitride balls; however, in other cases, the raceway geometry is altered
to optimize performance.

Non-metallic Rolling Contacts

355

Some of the performance advantages of hybrid bearings over their


all-steel equivalents are:
higher speeds; the low density of the silicon nitride balls reduces centrifugal loading and allows increases in operational speeds,
higher stiffness; silicon nitride has an elastic modulus 50 percent
higher than that of bearing steel, thus increasing the rigidity or
stiffness of the bearing;
reduced heat generation; the low friction characteristics of silicon
nitride balls combined with better kinematic behaviour reduce the
heat generated within the bearing and the temperature rise during
operation;
thermal stability; the thermal expansion of silicon nitride is one-third
that of steel so that the effects of temperature gradients across the
bearing are minimized; this reduces the risk of seizure;
greater exibility; the material properties of silicon nitride are such
that there is more scope for bearing designers to alter parameters
without counter effects.
In machine tool applications, these advantages give higher productivity,
better machining accuracy, and improved product quality. Bearings for
high-speed machine tool spindles are normally lubricated by an airoil
mist system. Hybrid bearings can be lubricated with grease without the
degradation of performance suffered by all-steel bearings. Elimination
of an airoil lubrication system can lead to signicant cost savings to
the machine tool manufacturer.
Hybrid ceramic bearings are also used for aircraft engines. Although
mainshaft bearings for gas turbines have a more complex form, their
principle of operation is the same as for angular contact bearings. Gas
turbine engines operate at high speeds and the centrifugal loads generated by the balls in the bearing are often the factor limiting the maximum speed of an engine. Hybrid bearings with low-density silicon
nitride balls offer the potential for higher-speed operation, which combined with less heat generation within the bearing will lead to increased
engine performance and efciency. Since aircraft engine mainshaft bearings are relatively large, there is also a signicant weight saving with
hybrid bearings.
Engineering applications of hybrid bearings and the nature of ceramic material require that the bearings be manufactured to a very high
standard of dimensional and running accuracy. Consequently, silicon
nitride balls for these bearings have to be made to the highest levels of
dimensional quality.

356

Rolling Contacts

8.5.3 Manufacture of silicon nitride balls


It is convenient to separate the stages in the manufacture of balls into
primary processing and nishing. The term primary processing is used
here to describe all operations from powder to the formation of
densied ball blanks, and nishing covers subsequent grinding, lapping, and polishing operations.
Primary processing
In common with most ceramic components, silicon nitride balls are
made from powder. Silicon nitride is a covalent compound and cannot
be densied by heat and pressure alone since the material dissociates
before becoming fully bonded. It is necessary to blend additives such
as yttrium oxide to the powder to promote densication by liquid phase
sintering. In the early stage of development of ceramic bearing components, the only method available for obtaining fully dense silicon nitride
was hot pressing which produced material in slab or disc form. Machining of ball blanks from hot-pressed slabs was laborious and expensive.
Apart from having a poor yield, such methods were clearly unsuited for
mass production. Consequently, net-shape manufacturing methods
have been developed for more efcient production of ball blanks.
The starting material is usually high-purity silicon nitride powder
with a mean particle size of less than 1 m. Mixingmilling of the powder with additives is an important stage that inuences nal product
quality. Some form of agglomeration, for example spray drying, is
necessary to make the powder ow for subsequent shaping into ball
preforms by die pressing or cold isostatic pressing. Ball preforms can
be soft machined in the pressed or green state to improve dimensional
quality. Silicon nitride requires temperatures of above 1700 C for
densication by the liquid-phase sintering mechanism. Desensication
by sintering alone (pressureless sintering) produces dense material but
with some residual porosity. Hot isostatic pressing (HIP) is now seen
as the preferred densication method for ball blanks since this process,
applied directly to glass encapsulated preforms, or as part of a sinter
plus HIP process, gives 100 percent dense material.
Ball nishing
Steel balls are manufactured from wire or rod which is cropped and
then cold pressed or stamped into a rough ball shape. Most of the
excess stock is removed by soft grinding (before heat treatment) and
hard grinding (after hardening) in ball grinding machines tted with
large silicon carbide grinding wheels. The balls are rough lapped

Non-metallic Rolling Contacts

357

between grooved cast iron plates in ball lapping machines using coarse
loose abrasives and then given a ne lapping or polishing operation
with a ne abrasive to obtain the required nal size and surface nish.
Owing to the high hardness of silicon nitride, it is not feasible to remove
stock material in ball grinding machines. Therefore, ball nishing operations are limited to rough and ne lapping using diamond abrasives.
Rough lapping times are longer than for steel balls because more material has to be removed, and these times can be extended further if the
dimensional quality of the blanks is poor. Since steel balls are made
from cropped wire or rod, there is some directionality present which
can affect the roughness of the nished balls. Modern silicon nitride
ball blanks are more isotropic and can be nished to better dimensional
quality than steel balls of the same size.

8.5.4 Dimensional quality of ceramic bearing components


There are various international and national standards for balls used in
rolling contact bearings, including ISO 3290. The dimensional quality
of balls is expressed in ball grades which are a combination of dimensional form, surface roughness, and sorting tolerances. The form
parameters used are:
ball diameter variation, which is the difference between the largest
and smallest actual single diameters of one ball;
deviation from the spherical form, which is dened as the greatest
radial distance in any radial plane between a sphere circumscribed
around the ball surface and any point on the ball surface.
Sorting tolerances are based on the nominal ball size, for example
10 mm, and the permitted variation in the actual mean diameter of balls
in a production lot from this value is expressed in gauge or subgauge
intervals. Balls for bearings generally have gauge intervals of 1 m and
subgauges of 0.2 m. Within a specied production batch or lot, the
main sorting parameter lot mean diameter variation is dened as the
difference between the mean diameter of the largest ball and that of the
smallest ball in the lot. Surface roughness tolerance values refer to the
arithmetical mean deviation, Ra, from the mean line of the prole,
evaluated according to the method specied in ISOR 468. The specication of ball grades in terms of these parameters is given in Table 8.10.
A further parameter that is used to describe the dimensional quality
of balls is waviness. There are no exact denitions of ball waviness
and it is broadly described as random or repetitive variations in surface texture that fall between form variations and surface roughness.

358

Rolling Contacts

Table 8.10 Maximum allowable tolerances ( l m) for


different ball grades

Ball grade

Ball
diameter
variation

Spherical
form
deviation

Lot mean
diameter
variation

Ra

3
5
10

0.08
0.13
0.25

0.08
0.13
0.25

0.13
0.25
0.5

0.012
0.02
0.025

Waviness of balls inuences the vibration and noise in bearings. Since


balls rotate more rapidly than the rings, ball waviness affects vibration
and noise in the high-frequency bands. In the absence of standardized
methods for measuring waviness, ball manufacturers apply their
internal standards or agree tolerances with customers.
High-precision angular contact ball bearings for machine tool
spindles typically contain balls of 614 mm diameter, and grade 3 or 5
dimensional quality is required. Hybrid bearings for aircraft engines
need larger balls, of 1740 mm diameter, nished to grade 5 or 10
quality.

8.5.5 Material quality


Stresses in bearings are high and are localized at, or near, the surface.
These Hertzian stresses can exceed 4 GPa. Although silicon nitride is the
ceramic material with the best combination of mechanical and physical
properties for use in bearings, in practice there are many types or grades
of material with different compositions, different properties, and different microstructures. Not all of these materials are suitable for the particularly demanding application of rolling contact bearings. Currently,
there are no national or international standards or specications for
silicon nitride; nor are there accepted standard methods for determining
some of the important properties. In view of this situation, ceramic ball
manufacturers have developed their own in-house specications and
quality assurance procedures to select and approve bearing grade
material.
A ne, uniform, and consistent microstructure is considered an essential prerequisite for good rolling contact fatigue resistance. The coarse
structure is usually associated with microstructural constituents of
porosity, metallic phases and ceramic second phases that can be present
in silicon nitride. A material with a coarse microstructure usually gives
poor fatigue performance.

Non-metallic Rolling Contacts

359

Silicon nitride materials with appropriate properties and microstructure usually have L10 fatigue lives that are considerably higher than
those of carbonchromium bearing steel and the M50 tool steel used
for aircraft bearings. Materials that contain porosity andor a coarse
distribution of metallic or ceramic phases fail after a low number of
fatigue cycles. In this context, a porosity level of 0.1 vol % would be
regarded as being high.
A further advantage of silicon nitride is that, if failure occurs, then
it does so by localized spalling in a similar manner to bearing steels.
Other types of ceramic material fail by catastrophic disintegration.

8.5.6 Surface quality


Surface roughness
In normal production, the surface roughness of silicon nitride balls, as
characterized by the Ra value, is in the range 0.0050.008 m. This level
of roughness is equivalent to that of the best-quality steel balls. However, by special lapping techniques, it is possible to produce very low
roughness values as shown in Table 8.11.
A very low level of surface roughness is not always necessary. A
rougher surface can have advantages in retaining lubricant, particularly
Table 8.11 Surface roughness (ISO) of 17.5 mm diameter
silicon nitride balls
S

Xr

Xmax

Xmin

Last X

R a ( m)
Rq ( m)
R y ( m)
R tm ( m)
R v ( m)
R p ( m)
Sm ( m)

0.000
0.000
0.012
0.004
0.010
0.003
3

0.002
0.003
0.038
0.024
0.030
0.010
36

0.003
0.004
0.058
0.031
0.046
0.017
40

0.002
0.003
0.023
0.018
0.018
0.007
31

0.003
0.004
0.058
0.028
0.043
0.014
40

q ( m)
q
R sk
R ku
S ( m)
R 3z ( m)
R pm ( m)
R 3y ( m)

35

39

30

38

0.0
2.0
16.1
15

0.0
0.0
32.8
16

0.0
3.4
5.4
14

0.0
3.4
32.8
16

0.0
1.0
10.8
1
0.001
0.001
0.002

0.013
0.007
0.017

0.015
0.009
0.021

0.012
0.005
0.013

0.014
0.009
0.021

360

Rolling Contacts

where solid lubricants are envisaged. Silicon nitride balls can be nished
to any level of roughness in the range RaG0.0020.05 m. The high
hardness of the material means that silicon nitride balls are less susceptible to dents, scratches, and indentations than steel balls and consequently have a more uniform and consistent surface roughness.
Surface appearance
The reectivity of silicon nitride is relatively high so that reected light
microscopy is the most convenient method for classifying the surface
appearance of balls. A random sample of balls from each production
lot is examined at a magnication of (B100) to classify the basic surface
in terms of small indentations and apparent surface porosity. The latter
feature includes both true porosity and apparent porosity arising from
microchipping during ball nishing. In addition, balls are assessed for
larger indentations, scratches, and metallic smears.
Surface defects
Surface integrity is an important requirement for bearing balls in view
of the high stress levels at the surface. Owing to the nature of the manufacturing processes for ceramic components, random or isolated material faults can occur at any stage from the initial blending of starting
materials to the nal nishing operations. Contamination by extraneous
materials can lead to metallic or ceramic inclusions in the densied
material. Cold pressing is also a potential source of defects when cracks
may be formed as a result of high ejection forces or stiction effects.
Pressing defects can remain as open cracks on the ball surface or may
be completely or partially healed during densication.
A number of non-destructive evaluation techniques are potentially
available for ceramic components. These include standard and microfocus X-ray radiography, acoustic microscopy, computed X-ray tomography, thermal wave microscopy, and high-frequency ultrasonic
methods. Currently, some of these techniques are not sufciently well
developed for the routine inspection of relatively large numbers of
components.
Although originally developed for the inspection of larger metallic
components, uorescent dye inspection has been applied successfully to
production quantities of silicon nitride balls. The combination of highsensitivity penetrants, carefully controlled processing, and specialized
examination techniques has proved to be effective in detecting isolated
material faults in nished balls. It is, however, necessary to supplement
uorescent dye inspection with visual inspection since not all types of
defect retain the penetrant. Shallow holes and healed pressing defects

Non-metallic Rolling Contacts

361

can be readily detected at low magnication using oblique white light.


Inclusions down to 50 m in extent can also be detected. In most cases,
ceramic inclusions are surrounded by material of altered composition
with different optical characteristics, which increases both the apparent
size and the contrast against the background. Metallic or metalloid
inclusions are also easily visible owing to their high reectivity.

8.5.7 Failure modes of ceramics in rolling contact


A comprehensive experimental study has been carried out to ascertain
the prevailing mode of failure of concentrated rolling ceramic contacts
subjected to cyclic loading conditions (7). Tests were carried out on
perfect and articially precracked silicon nitride balls under compressive cyclic stresses of up to 9.6 GPa. It was found that nominally perfect
silicon nitride balls fail in a non-catastrophic way. The prevailing mode
of failure is a delamination fatigue which is usually conned to a very
thin surface layer of material. Classication of crack initiation and subsequent propagation is schematically outlined in Fig. 8.17.
Shallow and deep initiations (types 1i and 2i) are dened according
to the initiation depth below the surface, i.e. below or above 10 m
respectively. As the classication is based on post-test examinations, it
is uncertain if surface or subsurface initiation actually occurs. There is

Fig. 8.17

362

Rolling Contacts

some evidence, however, of both surface and subsurface initiation.


Defect initiation is divided into two groups: no beach marks at defect
(type 3i) and beach marks at defect (type 4i). Beach marks are undulations present on fatigue fractured surfaces. Beach marks found within
the initiation region and classied as type 4i are commonly found on
sialon ball failures. Type 3i defect initiated failure is diagnosed by
premature cycles to failure and central spall at the centroid of the
delaminated surface. Precracks articially produced on the ball surface
usually lead to type 3i initiation failures.
Three propagation modes are identied from fractured ball surfaces.
Type 1p (fatigue typical beach marks) is dened by undulations measuring over 10 m peak to peak. Type 2p (fatigue shallow beach marks)
is characterized by undulations below 10 m peak to peak. Brittle crack
propagation (type 3p) is dened as a failed surface with no undulations
and a relatively smooth surface.
Another signicant nding is that silicon nitride balls with articially
produced cracks also fail in a non-catastrophic way. The pattern of
failure of precracked balls is strongly affected by the lubricating uid
and seems to be different from that observed for nominally perfect balls.
The dominating mode of failure of silicon nitride balls with an articially produced ring crack was found to be spalling fatigue.
It can be stated that currently available manufacturing techniques
can produce silicon nitride elements for rolling contact applications of
high quality and strength that are able to operate reliably for a long
time under severe loading conditions.

8.6 References
(1) Tomlinson, G. (1929) Molecular theory of friction. Phil. Mag., (7),
905.
(2) Stolarski, T. A. (1989) Fracture mechanics and the contact between
a pair of surface asperities during rolling. Int. J. Engng Sci., 27(2),
169.
(3) Johnson, K. L. (1985) Contact Mechanics (Cambridge University
Press, Cambridge).
(4) Hadeld, M. (1993) Rolling contact fatigue of ceramics. PhD thesis,
Brunel University.
(5) Stolarski, T. A. (1993) Rolling contact fatigue of polymers and
polymer composites. In Advances in Composite Tribology (Ed. K.
Fridrich), Vol. 8, Composite Materials Series (Elsevier).

Non-metallic Rolling Contacts

363

(6) Aramaki, H., Shoda, Y., Morishita, Y., and Sawamoto, Y. (1988)
The performance of ball bearings with silicon nitride ceramic balls
in high speed spindles for machine tools. J. Tribol., 110, 693.
(7) Hadeld, M., Stolarski, T. A., and Cundill, R. T. (1993) Failure
modes of ceramics in rolling contact. Proc. R. Soc., A443, 607.

This page intentionally left blank

Chapter 9
Coated Surfaces in Rolling Contact

9.1 Introduction
Interaction between elements in rolling contact takes place in various
situations. In some cases, rolling elements are in contact with very hard
materials or with high-temperature materials. Sometimes, the rolling
contact elements operate in a molten metal bath, such as molten zinc.
In all of these cases, the rolling contact elements suffer surface wear
due to, for example, the attack of the molten metal. Even when the
counter materials are very soft, such as paper, cloth, string, and so on,
the rolling contact elements are not free from surface wear.
The most effective way to prevent surface wear is to coat the surfaces
of rolling elements. When it is required to make a hard surface on a
steel roller, for example, it is possible to deposit a very hard coating,
such as WC (wolfram carbide)Co. If it is requested to make the surface
chemically stable and hard at high temperature, a Cr3C2 (NiCr) coating is the best choice. It is also possible to deposit a diamond coating,
and any ceramic coating in order to enhance surface performance. In
fact, coating processes are widely applied to rolling contact elements.
In this chapter, coating processes such as the thermal spray process,
electroplating, chemical vapour deposition (CVD), and physical vapour
deposition (PVD) are discussed, and their relevance to the performance
of rolling contacts is presented.

9.2 Coating processes


9.2.1 Thermal spray
Thermal spray processes are extensively applied to rolling contact
elements. In thermal spray processes, coating materials are in the form

366

Rolling Contacts

of a powder, wire, or rod. These materials are melted by burning fossil


fuels, mainly a mixture of acetylene and kerosene with oxygen, or by
electric energy. The high-temperature environment is created inside or
outside the spray gun or torch, and then the molten particles are
sprayed by the gun on to the surface of the substrate at a very high
speed. The most signicant features of a thermal spray process are:
(1) The highest deposition rate among all of the coating processes.
(2) The possibility of making coatings of any material except materials
that vaporize before melting.
(3) The possibility of making thick coatings.
The thermal spray processes are classied as shown in Fig. 9.1. The
speed range and temperature of molten particles appropriate for each
spray process are outlined in Fig. 9.2. It can be seen from this gure
that the highest particle speed can be attained in the high-velocity oxyfuel (HVOF) and detonation spray processes, while the highest temperature can be obtained in the plasma spray process. Among these
spraying processes, the arc and ame spray processes are mainly used
to deposit coatings for corrosion resistant, low melting temperature
metals such as zinc, aluminium, and their alloys. The HVOF and detonation spray processes are used to make hard coatings such as WCCo,

Fig. 9.1

Coated Surfaces in Rolling Contact

367

Fig. 9.2

Cr3C2 (NiCr). These materials, which are compounds of carbide and


metal or alloy, are called cermets. The cermet coatings are widely
applied for wear protection of rolling contact elements. The plasma
spray process is used to make coatings characterized by high melting
temperature. Ceramic coatings are deposited by the plasma spray
process.
High-velocity oxyfuel spray
In the HVOF spray process, acetylene or kerosene is used as a fuel and
is mixed with oxygen and then burned in a torch. In some spray systems, air is used instead of oxygen. In such a case, the process is called
the high-velocity airfuel (HVAF) spray process. In the HVOF and
HVAF spray systems, pressurized fuel and oxygenair are introduced
into the torch. An excess amount of oxygen or air is usually introduced
to the torch for an ideal stoichiometric ratio of fuel to oxygenair. In
some of the torches used, there is a combustion chamber and the diameter of the burning gas exit is reduced to produce a high-speed jet
stream. An example of the HVOF torch is shown in Fig. 9.3.
The particle speed depends not only on the spraying parameters but
also on the size and the specic gravity of the particle. As shown in

368

Rolling Contacts

Fig. 9.3

Fig. 9.2, the maximum particle speed can be more than twice the speed
of sound in air. However, the ame temperature of the HVAF is rather
low and certainly slightly lower than in the HVOF spray process shown
in Fig. 9.2. The high particle speed and low ame temperature are the
most typical features of the HVOF and HVAF spray processes. The
highest quality of cermet coatings can be obtained by the HVOF,
HVAF, and the detonation spray process which will be described later.
A recent tendency is to use pore-free stainless steel and titanium coatings deposited by the HVOF process for corrosion protection of steels.
The characteristics of thermal spray coatings deposited by various
processes are compared in Table 9.1, which demonstrates that coatings
with a very low porosity and low to medium oxide contents can be
deposited by the HVOF and HVAF spray processes. Also, the hardest
cermet coatings are deposited by this process.
Detonation gun spray
The detonation gun spray systems are only produced by one company
in the United States and a few companies in Russia. This system, however, is not widely used owing to restrictions on the sale of the spray
equipment.
Table 9.1 Porosity and oxide contents of some coatings
Spray process

Porosity

Oxide content

Atmospheric plasma spray


Vacuum plasma spray
HVOF (HVAF)
Detonation gun spray

High
Low
Very low
Very low

Mediumhigh
Very low
Lowmedium
Lowmedium

Coated Surfaces in Rolling Contact

369

Heat energy in both the detonation gun spray process and the HVOF
process comes from the combustion of fuel gases or kerosene. However,
there is one basic difference between them. In the HVOF process the
combustion is continuous, while in the detonation spray process the
combustion is intermittent. Usually, 510 detonations per second take
place in the spray system. The detonation gun is called a barrel, because
it resembles the barrel of a shot gun. A schematic illustration of the
detonation gun is shown in Fig. 9.4. In the detonation gun spray
machine, acetylene and oxygen are introduced into the barrel through
valves and ignited by a spark plug as shown in Fig. 9.4. Powder of the
material being sprayed is also introduced into the barrel and heated to
melting point by the detonation of the gases. The melted particles y
with very high velocity and impinge on a substrate surface to make a
coating. Owing to the high velocity of the melted powder, the lowest
porosity of the coating can be attained. The oxide content is said to be
the same as for the HVOF process. Antiwear materials such as WC
Co and Cr3C2 (NiCr) are the most suitable for the detonation spray
system. The hardness of such coatings depends on the coating process.
Figure 9.5 shows the hardness of WCCo coatings deposited by various
spray processes. The hardest coatings can be obtained by the detonation
spray and HVOF processes. Surprisingly, the hardness of a plasmasprayed WCCo coating is rather low. WCCo coatings deposited by
the detonation and HVOF spray processes at rather low temperature
have very low porosity. Owing to the low temperature, the resolution
of WCCo coatings into W2C or metallic W does not take place in the

Fig. 9.4

370

Rolling Contacts

Fig. 9.5

process. In the plasma spray process, described later, the temperature


of the plasma jet is very high, so that resolution of WC sometimes
occurs. For these reasons, the hardest WCCo coatings are deposited
by the detonation spray process.
Plasma spray
In the plasma spray process, electrical energy is used to produce high temperature. Spray material powders are heated, melted, and blown away
towards the substrate. Figure 9.6 illustrates, schematically, the crosssection of a plasma torch. The plasma torch consists of an anode, a
cathode, and a cooling system. When the arc is created between the anode
and cathode, the plasma gas, such as Ar, ArHe, and ArH2 , is introduced into the plasma torch. The gas is dissociated by the arc and a
plasma state is generated. As a mixed gas has higher enthalpy, it is more
suitable to melt powdered material. It is well known that plasma has a
very high temperature, up to 15 000 C in an ordinary plasma torch.
The plasma gas expands quickly owing to the high temperature, so the
gas leaves the torch with extremely high velocity. In some plasma torch
designs, spray material powders are introduced into the plasma jet when
it is outside the torch. Sometimes, the powder is injected into the plasma
jet near its exit from the torch as shown in Fig. 9.6. Occasionally,
water is used instead of a gas. Such a plasma spray system is called the

Coated Surfaces in Rolling Contact

371

Fig. 9.6

water-stabilized plasma spray process. Water is much cheaper than


either Ar or He, and thus low cost is a typical feature of the process.
In another conguration, the plama torch and a robot, which carries
the torch, are set in a chamber. The chamber is connected to a vacuum
pump so the coating is deposited at a low pressure. This is called the
low-pressure plasma spray (LPPS) or vacuum plasma spray (VPS) process. Before spraying, the chamber is evacuated to a high vacuum, and
then Ar gas is introduced into the chamber and spraying is carried out
under a pressure of 50100 torr.
The most typical feature of the plasma spray process is its high temperature. As mentioned earlier, the maximum temperature of a plasma
jet can reach 15 000 C. Thus, it is possible to melt any material and,
therefore, to make coatings of any material except materials that sublimate. As only inert gases are used for the plasma gas, oxidation of the
coating material is lower than in the arc or ame spray processes. The
bonding strength of the coating to the substrate in the plasma spray
process is also high. It is mainly due to the high temperature of the
melted particles. If, however, Ti is sprayed by the conventional plasma
spray process, most Ti changes to titanium nitride and titanium oxide.
In the low-pressure plasma spray process, it is possible to deposit coatings completely oxide free and with low porosity.

372

Rolling Contacts

The most suitable coating materials for plasma spraying are ceramic
oxides such as Al2O3 , ZrO2 , and TiO2 and high melting temperature
metals such as Mo. As the ceramic oxides are very stable at high temperatures, they are specially suitable for plasma spraying. Silicone
nitride (Si3N4) tends to sublimate, so it is very difcult to make coatings
of Si3N4 using the plasma spray process.
Mechanical properties of thermal spray coatings
The mechanical properties of thermal spray coatings are important
when they are used to modify the surface of a machine component.
Thermal spray coatings consist of a huge number of attened particles,
called lamellae, stacked together. Bonding between lamellae is usually
partial. The bonding region easily deforms or factures under a low
applied stress. As this behaviour controls the mechanical properties of
thermal spray coatings, the relationship between the applied stress and
the resulting strain is sometimes non-linear. Usually, the Youngs
modulus of a thermal spray coating is only 1412 of that for the bulk
material. The tensile strength is also very low compared with the bulk
material. The residual stresses in thermal spray coatings are very high,
because melted particles are attened, solidied, and cooled within microseconds. Thermal stress becomes the residual stress at room temperature. The random stresses, which have small stress elds and random
directions, are very high and can reach the level of fracture. However,
sometimes, the overall level of macroscopic residual stresses in thermal
spray coatings is low.
Finally, the thermal spray processes are evaluated from the viewpoint
of the environment. High jet stream velocities generate loud noise. The
HVOF and the detonation gun spray processes yield the loudest noise.
The noise sometimes exceeds 130 dB. The plasma spray torch emits loud
noise and bright light which includes ultraviolet rays. If a person stands
by the side of an operating plasma spray torch, the effect will be the same
as the person spending all day on the beach under intensive sunshine.
Smoke due to the formation of metal oxide, called fume, is generate in
any spray process. Blasting, which precedes any spraying operation,
generates dust. Therefore, a soundproof room and dust-absorbing or
extracting devices are required in any thermal spray installation.

9.2.2 Electroplating
Electroplating is a coating method in which metallic ions present in a
bath are precipitated on to the surface of the substrate by the electrochemical process to form a metallic coating. The principle of electroplating is shown schematically in Fig. 9.7. A substrate is connected to

Coated Surfaces in Rolling Contact

373

Fig. 9.7

the negative pole of the d.c. current, while the coating plate is connected
to the positive pole in a bath containing metallic ions.
The electroplating is carried out for three basic reasons. The rst is
for decoration. Gold, silver, platinum, rhodium, and their alloys are
electroplated for this purpose. The second reason is for corrosion protection of steels. Zinc, tin, and cadmium are mainly used for this purpose. The third reason is the enhancement of tribological performance.
Chromium and compound material coatings are used to reduce wear.
Tin and indium are used for sliding contacts. The electroplating coatings that are used for tribological purposes are listed in Table 9.2.
A coating bath containing CrO3 (250 gl) and H2SO4 (2.5 gl), is commonly used to deposit a Cr coating. Chromium electroplating is used
for both decoration of goods and industrial purposes. The coating lm
has good appearance and is a shiny blue. The thickness of the coatings
depends on the intended usage. When the Cr coating is used for decoration, its thickness is 0.250.50 m, but when the coating is used for
wear protection, the thickness of the coating is 10100 m.
The electroplated Cr coating has good tribological properties. The
Cr coating is chemically stable in the atmosphere. It has high hardness,
between HV (Vickers hardness) 800 and 1000, and also has a low

374

Rolling Contacts

Table 9.2 Electroplating coatings for tribological usage


Coating
material
Metal
Alloy

Composite

Specication

Structure

Hard Cr
PbSn
PbSnSb
CuSn
Matrix
Ni, Ni alloy
Co
Fe, Fe alloy
Co

Usage
Wear protection
Sliding contacts

Dispersed particles
Diamond, WC
TiC, TiN, Al2O3
PTFE (polytetrauoroethylene), MoS2

Wear protection
Low friction

friction coefcient. The Cr coating is dense, but it is possible to make


it porous or with a network of ne surface grooves. These pores and
grooves serve as reservoirs for a lubricant so that adhesive wear is effectively prevented. The Cr coating can be deposited on aluminium, copper, and copper alloys as well. Usually, a glossy surface is preferred and
can be obtained by choosing a rather low current efciency.
The hardness of any electroplated coating is higher than that of a
bulk material. This is considered to be due to the strain generation
within the coating during precipitation in the coating bath. A coating
that has higher strain also has higher hardness. The hardness of electroplated coatings depends on both bath temperature and current density.
Figure 9.8 shows the range of bath temperature and current density in
which a hardness of more than HV 1000 can be obtained. As shown in
this gure, the bath temperature and current density range from 30 C
and 10 Adm2 to 60 C and 140 Adm2.
The hardening mechanism of the Cr coating is similar to that of work
hardening. As described earlier, hardening is mainly due to the strain
generation during plating. Therefore, at elevated temperatures the hardness decreases. The decrease in hardness of the Cr coating is more pronounced when the overall coating hardness is higher. Usually, hardness
decrease occurs at 400 C. Similar phenomena, such as recovery and
recrystalization, also take place in the Cr coating when strain and hardness decrease.
The antiwear performance of electroplated Cr coatings has been
widely investigated. The Cr coatings were electroplated by changing the
bath temperature and current density. The antiwear performance was

Coated Surfaces in Rolling Contact

375

Fig. 9.8

tested on samples. It was found that hard Cr coatings have excellent


antiwear properties. The antiwear performance of the Cr coatings
changes slightly with the surface nish. As described before, the surface
can have a porous or plain nish. The Cr coating with a porous surface
nish shows an excellent wear performance, especially under lubricated
conditions. The network of grooves on the porous surface retains lubricant which, in turn, contributes to the wear reduction.
Electroplated alloy coatings are mainly used for corrosion resistance
and decoration. One application of alloy coatings for tribological usage
is for sliding bearings. Usually, very soft coatings are used for sliding
surfaces. In the case of direct contact between interacting surfaces, the
soft metal coating deforms easily and prevents damage to the
counterface.
One example of a composite coating is given in Fig. 9.9. This shows
a circular cutting tool, and the shining grains are diamond particles
dispersed in a base metal (nickel). By dispersing particles such as diamond in a plating bath, it is possible to deposit them with the nickel.
Composite coatings with base metal and low friction material additives,
for example PTFE, can also be deposited.
Chemical plating (electroless or autocatalytic deposition) is a process
in which electric current is not used but a chemical reaction is utilized to

376

Rolling Contacts

Fig. 9.9

deposit the coating. At present, this process is mainly used for corrosion
protection and as an undercoating of plastic materials. However, the
coatings deposited by this process have good tribological properties.
Typical characteristics of coatings deposited by chemical plating are as
follows:
(1) Deposition is possible on the surface of non-conductive materials
such as polymers and ceramics.
(2) The thickness of the coating is uniform even at the corner or edge.
(3) Very low porosity.
(4) High hardness.
(5) High strength of the interface.
The coating cost of chemical plating is higher than that of electric deposition, but this process is used for a number of advantages mentioned
above. Not all metals can be deposited by this process. Metals occupying a region in the middle of the periodic table, for example Fe, Co, Ni,
and Pd, are suitable as substrates for this process. Other metals and
most of the plastics require pretreatment.
As an example, the characteristics of the electroless Ni coating are
presented here. The Ni coating usually contains less than 10% P so, in
fact, the coating is an NiP alloy. The specic gravity of the Ni coating
is about 7.9 gcm3, and the hardness of the coating in the deposited
state is about HV 500. However, by applying a heat treatment, the hardness of the coating can be raised to more than HV 1000. Residual
stresses are very important for electroplated coatings. A residual tensile

Coated Surfaces in Rolling Contact

377

stress tends to be generated in electroplated coatings. On the other


hand, a residual compressive stress is generated in non-electrolyte Ni
coatings. The compressive stress is benecial in some cases.

9.2.3 Physical vapour deposition


All physical vapour deposition (PVD) processes need a vacuum
environment. PVD is also called vacuum plating or dry plating because
it does not use water or a water-based solution like electroplating. The
PVD comprises three basic processes, namely vacuum evaporation, ion
plating, and cathode sputtering.
The vacuum evaporation device is schematically shown in Fig. 9.10
The device consists of a vacuum chamber, a vacuum pump, and an
evaporation source. Less than 104 torr of chamber pressure is required
to deposit the coating. The heating method of the evaporation source
considerably affects coating quality. Three types of heating method are
used in the vacuum evaporation devices. One is to use the electric resistance of metallic materials such as NiCr, W, and Mo. The second heating method consists of utilizing an electron beam, and the third method
uses the high-frequency induction process. The process utilizing electron
beam heating produces contamination-free, high-quality coatings.
The coating material, such as Al, Cr, Cu, Ag, and also ceramic oxides
(Al2O3 and TiO2), is heated to a temperature sufcient to vaporize the
coating material. Particles of vaporized coating material y in the high
vacuum directly to the substrate surface on which they are stacked to
form a coating. The coating mechanism of this process is very simple,
and the process has the following features.

Fig. 9.10

378

Rolling Contacts

(1) The vapour of the coating material hits the substrate surface facing
the evaporation source and the coating is formed. The coating cannot be formed on the side or back faces of the substrate. It is rather
difcult to deposit a coating of uniform thickness.
(2) Coatings with excellent optical properties can be deposited.
(3) The thickness of the coating is less than 1 m and it is difcult to
deposit thicker coatings.
(4) The adhesive strength of the interface is rather low and the wear
performance of the coating is average.
Application of the PVD process is very wide. The electronic and optical
industries are the two major areas where PVD is used. In integrated
circuits, 80%Ni20%Cr alloy is commonly used for resistors, and aluminium for wiring. The process is also used for optical parts, such as
lenses, to prevent diffraction of light at the lens surface.
In the ion plating process, evaporated particles are attracted towards
the substrate by electric forces. The deposition mechanism of ion plating is illustrated schematically in Fig. 9.11. As shown in this gure,
negative high voltage (10005000 V) is applied to the substrate and an
evaporation source. The atmosphere of the ion plating is argon and the

Fig. 9.11

Coated Surfaces in Rolling Contact

379

pressure is 525B103 torr. Glow discharge of electricity occurs and


dark space is formed around the substrate. The metallic ions evaporated
from the evaporation source are accelerated in the dark space and
impinge on the substrate surface. Since the pressure is very low, evaporated particles collide with gas atoms several times before they reach the
substrate surface. Therefore, the particles also deposit themselves on
surfaces that are not facing the evaporation source. The use of ion
plating is very wide. The ion plating process is most widely utilized in
electronic applications. For example, a Cu coating is used for a printed
board or integrated circuit board, NiCr coatings are utilized for electric resistance elements, while Fe and Ni coatings are used in magnetic
tapes. In corrosion protection applications, Al, Zn, Cr, and Ti are used
as coating materials. To prevent wear, Cr, TiC, and TiN are used. The
ion plating process is also used for decorative applications.
The ion plating processes are classied according to the technique
used to evaporate coating materials. Thus, the following evaporative
PVD can be distinguished:
(a)
(b)
(c)
(d)

resistive,
inductive,
electron beam gun,
arc.

The third process of PVD is a sputtering process. In this process, unlike


the former two processes, the coating material is not melted but,
instead, the atoms of coating material are knocked out by ion bombardment. Thus, this process has the advantage over the evaporation and
ion plating processes that contamination from the crucible of the evaporation source is negligible. However, the deposition rate of the sputtering process is low compared with the ion plating and evaporation
processes. The deposition mechanism of sputtering is schematically
illustrated in Fig. 9.12. The comparison of the three coating processes
is summarized in Table 9.3.

9.2.4 Chemical vapour deposition


The chemical vapour deposition (CVD) process is a coating method in
which the coating material, mainly chloride, is in a vaporized form and
the coating material is generated and deposited from the chloride by a
chemical reaction. Heat, plasma, and light are used to promote the
chemical reaction. These terms are also used to classify CVD processes
into thermal CVD, plasma CVD, and photo-assisted CVD. A schematic
illustration of the CVD process is shown in Fig. 9.13.

380

Rolling Contacts

Fig. 9.12

Table 9.3 Characteristic features of PVD processes


PVD processes

Characteristic features

Evaporation

Simple and cheap; coating thickness up to 1 mm; high


temp. 10002000 C; vacuum 106 1 Pa; low strength
of interface; shadowing effect
No clear interface; mixing with substrate inevitable;
extremely well-bonded coating; limited to metals only;
thin diffused coatings
Very flexible process; any material can be coated;
deposition rate low; clean process

Ion-plating

Sputtering

The deposition of a coating material normally takes place in the following way. Hydrogen (H2), nitrogen (N2), or argon (Ar) gas is used
to carry the coating material which is usually in a liquid or solid state.
The carrier gas should be properly purified prior to use as some other
gases could react with the coating material. The coating material is first
introduced into a vaporising device. The coating material is mainly a
halogen composite. It should be pure and its evaporating temperature
low. Also, it should be able to precipitate a metal on a substrate surface
through a chemical reaction. The chemical reactions that take place to

Coated Surfaces in Rolling Contact

381

Fig. 9.13

deposit a metallic coating are as follows:


MA (vapour)CB (gas M (precipitate)CAB
MC (vapour)CD (gas) M (precipitate)CCCD
Thus, the metal M precipitates on the substrate surface and makes a
coating.
The thermal CVD process has a number of specic features. A coating material in the gaseous state is introduced into the chamber by a
carrier gas. In order to increase the rate of chemical reaction, the substrate is heated using electrical resistance, high-frequency induction,
infrared rays, and laser beam. The temperature is a very important
factor for the thermal CVD process as an increase in temperature
increases the deposition rate. However, too high a temperature results
in a deposition rate decrease. Atmospheric pressure has been widely
used as the chamber pressure, but low pressure ranging from 102 to
105 Pa is currently used to enhance the uniformity and deposition rate
of the coating.
The thermal CVD process is characterized by a number of typical
features. The rst one is of practical importance as any material can be
deposited by this process. Metals such as Al, Zr, and Ti, alloys such as
TaNb, MoW, carbonates (TiC, SiC, and WC), and nitrides (TiN,
AlN, and Si3N4) can all be deposited using thermal CVD. In addition,
borides, silicates, and even oxides can also be deposited with the help
of the CVD process.
The quality of coatings produced by CVD is very good. The coating
is characterized by high adhesive strength, high hardness, and uniformity and is defect free. CVD coatings are widely used for cutting tools.
A tool coated with TiN, for example, can easily be recognized by its

382

Rolling Contacts

shiny golden colour. One problem with the thermal CVD process is the
high temperature used to enhance the rate of chemical reactions. For
example, when TiC is deposited from SiCl4 , the plating temperature is
11001200 C. Occasionally, the substrate can be damaged by this high
temperature owing to grain growth and deformation caused by thermal
stresses.
Practical application of thermal CVD is very wide. The strength of
the interface is high, even at elevated temperatures, and therefore TiC,
TiN, SiC, and BN coatings are used to improve the wear resistance of
cutting tools. CVD coated ceramics are utilized in aero and rocket
engines as heat barriers and insulations. Semiconductor and insulation
coatings for electrical devices constitute another large area of application of the CVD process.
The plasma CVD process is a technique in which the reaction gases
are transformed to a plasma state. The plasma enhances the chemical
reaction and deposits a coating. Gas in a plasma state can be of two
types. One is a high-temperature plasma where the temperature reaches
several thousand degrees centigrade. In this case, the temperature of the
gas molecules and the temperature of their electrons are the same. This
state is called thermo-equilibrium plasma. The other type of plasma
state is low-temperature plasma where the gas temperature is only several hundred degrees centrigrade. Nevertheless, the temperature of electrons reaches 10 000 C. As the temperature of the gas molecules does
not coincide with the temperature of their electrons, this stage is called
non-equilibrium plasma. In the plasma CVD process, it is mainly lowtemperature plasma that is utilized. The gas temperature is around
300 C and the chamber pressure is 1100 Pa.
The plasma CVD process is indispensable in the manufacture of electronic devices such as semiconductors, insulators, and so on. The
plasma CVD process is also used to make coatings on optical lenses
and other high-precision optical components. Hard coatings, such as
diamond and TiC coatings, are deposited on mechanical parts in order
to enhance their tribological performance.
In the photo-assisted CVD process, optical rays, including a laser
beam, are radiated to activate the chemical reaction, leading to the
deposition of coating material. A YAG laser, CO2 laser, mercury lamps,
xenon lamps, and halogen lamps are used in the deposition process. In
photo-assisted CVD, the conformity between the wavelength of the ray
and the wavelength that is absorbed by the deposition material gas is
important. Therefore, a particular source of light is used for a given
material gas. Usually, the photolysis of coating materials is induced by

Coated Surfaces in Rolling Contact

383

ultraviolet rays. Deposition at temperatures lower than in plasma CVD


is possible in photo-assisted CVD. Metals, carbonates, and nitrides can
be deposited by photo-assisted CVD.

9.3 Application of coatings to rolling contact


elements
9.3.1 Rollers for steel forming
At steelmaking facilities, many rollers are used. Usually, coatings are
applied to the rollers in order to prolong the life of the rollers, lower
their maintenance costs, and enhance the surface quality of steel sheets
produced during rolling. The rollers to which coatings are applied can
be classied into three types, i.e. hearth rollers, process rollers, and sink
rollers.
Hearth rollers are used mainly in the continuous annealing lines. A
schematic illustration of the continuous annealing line is shown in Fig.
9.14. As shown in this gure, the hearth rollers are used to change the
direction of the steel sheet. As the temperature of the annealing furnace
reaches about 1200 C to anneal, for example, a stainless steel sheet,
oxidation and wear easily take place on the surface of the roller. The
most serious problem for the hearth rollers is the formation of an
accretion. The accretion is formed by stacking of oxides transferred
from the steel sheet. Once the accretion is formed, the surface roughness
of the roller increases and the surface quality of the sheet dramatically
decreases.

Fig. 9.14

384

Rolling Contacts

Process rollers are used in a steelmaking process including mills. The


mill rollers are exposed to extremely high contact stresses. In addition to
high contact stresses, the surface of the mill roller has to have specied
roughness as the hot billet is rather slippery. It would be difcult for
rollers with smooth surfaces to grip the billet and pull it in. The required
surface roughness, on the other hand, contributes to oxidation and
enhances wear which, in turn, are the main causes for the decrease in
surface quality of the steel sheet. It is very difcult to reconcile the need
for moderate surface roughness and durability of rollers.
Sink rollers are used to produce zinc-coated steel sheets for application in the automotive industry. More than twenty years ago, car
bodies were made of uncoated steel sheet and, therefore, were prone to
rapid corrosion. Car body steel sheets are quite thin, so corrosion can
easily penetrate them. Basically, there are two types of production process used to make zinc-coated steel sheets. One is electroplating and the
other is the hot dip process. In the hot dip process, the steel sheet is
continuously passed through a molten zinc bath. In order to maintain
the motion of the sheet, a roller is operating in the molten zinc bath. It
is well known that molten metals are very aggressive and attack solid
surfaces. Thus, the sink roller operates in a very corrosive environment.
A number of measures have been taken to solve the above-mentioned
problems. In some steelmaking facilities, ceramic rollers have been used
as hearth rollers. However, ceramic rollers, although able to resist corrosive wear, are very expensive to make and are prone to brittle fracture, especially under the high contact stress used. The most suitable
and effective way to solve the difculties associated with rollers in steelmaking facilities is to use coatings. The plasma transferred arc (PTA)
process, high-temperature isostatic pressing (HIP), CVD, PVD, and
electroplating have all been used to coat steelmaking rollers. However,
the use of thermal spray processes is now on the increase. The following
is an illustration of this trend. In a steel mill in Japan, 784 m2 of hearth
roller surfaces were thermally sprayed, as well as 456 m2 of process
roller surfaces and 35 m2 of sink roller surfaces (1). In this steel mill,
the tension bridle rollers, a kind of process roller with a diameter of
760 mm and a length of 1730 mm, were coated with electroplated hard
chromium. The service life of the electroplated roller was only
2.5 months. When the electroplated hard chromium was replaced by a
thermally sprayed WCCo coating, the service life was increased to
more than 6 years and the maintenance cost was reduced 10 times. In
another solution aiming to prolong the service life of process rollers, it
was proposed to use 80%Ni20%Al as the coating material (2). As a

Coated Surfaces in Rolling Contact

385

result, service life was considerably extended and the traction required
to drag the billet into the rollers was preserved. An increase in the
service life of the hearth rollers was achieved by applying a CoCrAlY
alloy coating reinforced by dispersed Ti2O3 CrB2 particles (1). By using
this particular coating, a very wear resistant surface was produced as
the surface roughness of the rollers did not change appreciably after 6
years service. Clearly, it is possible, using this coating system, to
increase signicantly the service life and to prevent the formation of
accretions.
As accretion formation on the surface of hearth rollers is a serious
problem for a continuous annealing line in a steel mill, various solutions
have been tried. One of them is to use an abradable coating that wears
easily during roller operation. When the accretion is formed, a very thin
layer of coating is delaminated together with the accretion. It was found
that a Ni-based cermet coating containing 5.5%BNNi14%Cr8%Fe
3.5%Al offers excellent protection against accretion formation (3). For
the same reason, a Cr3C2 coating was tested for use on hearth rollers
operating below 900 C (4). An abradable coating is also used inside
the casing of a turbocharger in internal combustion engines. The blades
of the turbocharger scrape the abradable coating so the gap between
the blades and the casing can be minimized and the pressure of the
compressed air can be increased, resulting in a higher power output.
A continuous galvanising line is shown schematically in Fig. 9.15.
After degreasing and acid cleaning, the steel sheet is introduced into the
molten zinc bath. In the bath, a sink roller and a support roller are
operating. The top rollers, although working outside the bath, suffer
considerable wear because the temperature of the zinc-coated sheet is
still high and the contact stress is considerable. Therefore, a WCCo
coating is usually applied to the top rollers.
The lifetime of a sink roller is usually 1 or 2 months. Thus, one of
the most important problems for a steel mill is to increase the service
life of the sink rollers. Thermally sprayed coatings are applied to sink
rollers in order to prolong their service life. The most commonly used
coating for the sink rollers is WCCo (5), but a borideWCCo composite and CoCr self-fusing alloy coating have also been tried recently.

9.3.2 Rollers for papermaking and printing


In the papermaking and printing industries, rollers are commonly used.
However, some differences between steel mill rollers and papermaking
and printing rollers exist. Although the contact stresses acting on papermaking and printing rollers are considerably lower compared with steel

386

Rolling Contacts

Fig. 9.15

mill roller contact stresses, the surface nish required is much higher.
The production of paper or print demands a mirror surface nish that
has to last for a considerable length of time. Therefore, coatings are the
only solution for this type of roller.
A number of specic properties are required from the surface of
printing and papermaking rollers. One of them is wear and corrosion
resistance. As mentioned above, a mirror nish of the coating surface
is required and it has to last as long as possible. The service life of a
roller is linked to maintenance costs. Moreover, repairability is also an
important consideration for the coating. When the coating is partially
worn out, the remainder has to be removed by blasting or a chemical
treatment and a new coating is deposited. This is economically justied
as the cost of a new roller is very high. The surface of a printing roller
is engraved by a laser beam, and it is therefore important that cracks
are not created during this process.
In order to satisfy all the above requirements, Cr2O3-based, Al2O3
TiO2 , WCCo, and stainless steel coatings have been applied to paper
and printing rollers. Some of the rollers are very long, as shown in Figs
9.16 and 9.17, and therefore the only coating system that can be used
is the thermal spray process. The reason for this is of a practical nature
as there is no sufciently long vacuum chamber for PCD and CVD or
long enough electroplating bath. Among various thermal spray

Coated Surfaces in Rolling Contact

Fig. 9.16

Fig. 9.17

387

388

Rolling Contacts

processes, the plasma spray and HVOF spray processes are the most
commonly used. Figure 9.16 shows a dryer roller that is used in the
papermaking process. Superheated vapour is introduced into the roller
to dry the paper. A WCCo coating is deposited by three HVOF spray
guns. Figure 9.17 shows a mirror nished metering roller utilized in a
gate roll coater.
Coatings are also applied to rolling elements of rolling contact bearings (6). Silver, lead and MoS2 are independently deposited by PVD
and CVD processes on rolling elements used in bearings without a
retainer. Tests carried out in a vacuum proved that such a bearing can
run for more than 108 revolutions under a contact stress of 800 MPa.

9.3.3 Fracture of coatings during rolling


The coating on a rolling contact element will be subjected to delamination wear after a certain period in service. This is mainly due to the
contact conguration and the contact stress magnitude.
Even if a coating material is very hard, it is possible for plastic deformation of the hard coating to be produced because the contact conditions create a state of stress resembling hydrostatic pressure. Contact
fatigue of a line contact created by thermally sprayed Mo, Al2O3 -based,
and Cr2O3 -based ceramic coatings was experimentally investigated (7).
The contact conguration used is shown in Fig. 9.18. As can be seen
from the gure, the steel counter-disc was in contact with the coated
disc at about two-thirds of the coating width. After rolling contact tests,
the surface prole of the coating in the axial direction was measured.

Fig. 9.18

Coated Surfaces in Rolling Contact

389

It was found that the prole of the coating was convex. This means
that the coating deformed even in the axial direction. The maximum
deformation was about 1 percent. From this, it can be surmised that
the deformation of the coating in the circumferential (rolling) direction
was much more extensive. It is known that in the case of thermally
sprayed coatings the bonding strength of each lamella interface is different, and thus slip can take place at weakly bonded lamella interfaces
(8) Also on account of deformation of the coating, a very high shear
stress is applied to the substratecoating interface. It is the shear stress
that is responsible for delamination wear. Therefore, deposition of a
coating of the same thickness as the depth of the maximum shear stress
in a rolling contact should be avoided. In the course of the study, the
residual stresses in the coating were measured after a certain testing
period (7). It was found that the residual stresses were low and compressive in nature but they increased linearly with time. Thus, it seems
possible to estimate the fatigue life of a coating from a measured
residual stress level.
Rolling contact fatigue performance of coatings was tested using a
well-known four-ball contact conguration (9). In this conguration,
three lower balls are contained in a steel cup and the fourth coated ball
or cone, attached to a spindle via a special collet, is in loaded contact
with them. A WCCo coating was deposited on to half the ball or cone
surface. The residual stresses within the ball or cone were measured
using a microbeam X-ray method before and after the rolling contact
fatigue tests. Very high compressive stresses were measured before the
test in the WCCo coating deposited by a HVOF thermal spray. It was
found, in the course of testing, that the residual compressive stresses
increased with testing time.
Available experimental results show that the residual stress of a coating depends on the material used and the coating process itself. Again,
it was found that the residual stress changes, approximately, in a linear
way with the rolling contact testing time. The measurement of residual
stress level before and after testing could be used to estimate the time
to failure of the contact.

9.4 References
(1) Sawa, M. and Oohori, J. (1995) Proceedings of International
Thermal Spray Conference, Kobe, Japan, (5), pp. 3742.
(2) Mayor, V. (1996) Proceedings of National Thermal Spray Conference, Ohio, USA, (10), pp. 6164.

390

Rolling Contacts

(3) Midorikawa, S., Sato, Y., Matumoto, S., and Ito, M. (1995) Proceedings of International Thermal Spray Conference, Kobe, Japan,
(5), pp. 4346.
(4) Swang, S. Y. and Seong, B. G. (1995) Proceedings of International
Thermal Spray Conference, Kobe, Japan, (5), pp. 5963.
(5) Kobayashi, Y., Tani, K., Harada, Y., Nakahira, A., Tomita, T., and
Takatani, Y. (1995) Proceedings of International Thermal Spray
Conference, Kobe, Japan, (5), pp. 211216.
(6) Nishimura, M. (1995) J. Surf. Finishing Soc. Jap., 46(7), 632635.
(7) Tobe, S., Kodama, S., Misawa, H., and Ishikawa, K. (1990) Proceedings of National Thermal Spray Conference, California, USA, (5),
pp. 171177.
(8) Makela, A., Vuoristo, P., Lahdensuo, M., Niemi, K., and Matyla, T.
(1994) Proceedings of National Thermal Spray Conference,
Massachusetts, USA, (5), pp. 759763.
(9) Ahmed, R., Hadeld, M., and Tobe, S. (1996) Proceedings of
National Thermal Spray Conference, Ohio, USA, (10), pp. 875
383.

Chapter 10
Rolling in Metal Forming

10.1 Introduction
Rolling, both hot and cold, is of enormous importance to the metalforming industry. It rst started with hot material rolling. As is usually
the case in an industry, knowledge of how to obtain the desired result
was gained without knowing why it was obtained. Practical knowledge,
in other words, was much in advance of theory. The process of deformation that a material undergoes in rolling has not lent itself readily
either to mathematical analysis or to experimental investigation. The
friction forces between the rolls and the material, which largely control
the rolling process, are not capable of direct measurement, and only
with very special apparatus in small laboratory mills can the pressure
distribution between the rolls and the material be measured. The pressure required to deform the material between the rolls is greater than
that needed for a similar reduction between at frictionless plates,
owing to the friction effects between the rolls and material, and moreover it varies with the thickness of the material. In hot rolling it also
depends on the speed of working, and in cold rolling the material work
hardens during the pass. Hence, mathematical theories designed to
enable the rolling pressure to be calculated have to be based on simplifying assumptions that often depart appreciably from the truth. In
addition, the essential data, for these assumptions, of the resistance to
deformation of the material under the particular rolling conditions and
of the coefcient of friction are unfortunately inexact.
Rolling parameters can be fairly easily measured experimentally, and
this has frequently been done. The power required by the rolls has also

392

Rolling Contacts

been measured, but such results apply only to the test model conditions,
and cannot be used without risk in a real situation. Therefore, the development of a complete theory of rolling that enables the rolling load,
power requirements, optimum conditions of front and back tension,
roll diameter, best strip lubricant, etc., to be predicted from the physical
characteristics of the material and the dimensions of the rolls has not
yet been accomplished, although much progress towards this goal has
been made.

10.2 Forces acting in the contact region


10.2.1 Forces acting in the roll gap
The friction between the surface of the rolls and the material rolled,
also called external friction, is the fundamental factor in the reduction
of material by rolling. It is the force that draws the material between
the rolls, and, if it did not operate, rolling, as distinct from drawing,
would be impossible. Friction greatly affects the magnitude and distribution of the pressure acting between the rolls and the material, and
consequently affects the power required for the reduction of the material. It also controls the amount of reduction, or draught, it is possible
to take. Generally speaking, the higher the coefcient of friction, the
greater is the possible draught. Similarly, the increase in width, or lateral spread, of the material as it passes between the rolls becomes
greater as the roughness of the rolls increases. Knowledge of the coefcient of friction in rolling, and particularly its variation, is undoubtedly of the greatest practical value, as it determines the extent to which
rolling can be carried in each pass, and consequently the design of the
roll drive.
Practical difculties in rolling may occur when the coefcient of
external friction is on the one hand very small and on the other hand
very great. Examples of the former case occur in cold rolling with
smooth rolls and a copious supply of an efcient strip lubricant, making
it difcult to enter the strip or sheet between the rolls. If, however, the
coefcient of friction has a high value, as occurs in cold rolling with
rough rolls and no lubricant, experience has shown that there is a
marked tendency for the rolled product to split and for the edges to
tear. In hot rolling, where the value of the coefcient of friction is generally high, a similar tendency exists and is sometimes revealed as a series
of hairline cracks in the surface of the rolled material which oxidize and
hence do not reweld in subsequent rolling. A further practical instance
of the effects of a high value of the coefcient of friction is the old cold

Rolling in Metal Forming

393

rollers trick of making a roll bite the material by applying parafn to


the roll and thereby considerably increasing the coefcient.
In view of the practical and theoretical importance of friction in the
rolling process indicated above, it is necessary to consider the manner
in which the friction and other forces in the region of contact between
the rolls and material interact. These forces are shown in Fig. 10.1,

Fig. 10.1

394

Rolling Contacts

which represents the rolling down of a parallel bar from an initial thickness h1 to a nal thickness h2 between two plain rolls each of radius R.
It will be assumed that the material does not spread laterally and that
the only forces acting on the bar are those exerted by the rolls. Then,
since the material does not appreciably change in volume during rolling
(except in the case of the rst few passes in rolling an ingot where
consolidation is apparent), the velocity V2 with which the material
leaves the rolls will be greater than the velocity V1 with which it enters
them. The periphery velocity of the rolls will be assumed constant. If
the roll grips the bar, this velocity cannot be less than V1, nor greater
than V2, and in general it lies between these two values.
If one makes the assumption common in theories of rolling that a
vertical plane section of the bar before rolling remains plane during
rolling, then it can be shown that there will be one point, and one point
only, in the surface of contact between the rolls and the material at
which the periphery of the rolls moves at the same velocity as the surface of the bar. This is termed the neutral point or no-slip point and is
shown at Y in Fig. 10.1. Between the neutral point Y and the plane of
entry X of the material between the rolls, the roll surface is moving
more rapidly than the surface of the bar, and the friction between the
rolls and bar tends to draw the bar between the rolls. Between the
neutral point and the point Z at which the bar leaves the rolls the
surface of the bar is moving more rapidly than the periphery of the roll,
and the friction between the bar and the rolls tends to oppose the delivery of the bar from the rolls.
In order to investigate the forces acting between the bar and the rolls,
consider a small arc of contact element subtending an angle d at the
roll centre, the radius to this element making an angle with the line
joining the roll centres (Fig. 10.1). Then the length of this element is
R d , and the forces acting on it are those due to the radial pressure pr
between the rolls and bar and the tangential friction force F; the direction of action, or sense, of F depends on whether the arc element lies
on the entry or exit side of the neutral point. Considering the unit width
of the bar, the radial force PR acting on the section is equal to pr R d .
For the purposes of the present analysis it will be assumed that the
radial pressure pr is uniform along the arc of contact. This assumption,
as will be shown later, is sufciently true only for cases of rolling in
which the draught is small in comparison with the thickness of the bar.
The tangential frictional force F is equal to PR, where is the coefcient of external friction which is assumed to be constant throughout
the arc of contact. Here, again, the assumption may not be valid as the

Rolling in Metal Forming

395

possible change in the surface condition of the bar from point to point
may cause to vary very appreciably along the arc of contact.
Considering the forces acting at point A (or, more strictly, on the arc
element having its centre at A, Fig. 10.1), which lies between the plane
of entry X of the bar and the neutral point Y, the resultant of the forces
PR and F is represented by K, while HE and HV are the horizontal and
the vertical components respectively of this resultant force K. It will be
noted from its direction that the horizontal component HE tends to
draw the bar into the rolls, and the value of this component at point of
entry X determines whether the bar will enter the rolls unaided. From
Fig. 10.1, the horizontal component of PR at X is PR sin , where is
the angle of contact and the horizontal component of the friction force,
F, is F cos . The horizontal component HE of the resultant K is therefore F cos APR sin . The entry of the bar between the rolls becomes
impossible when HE G0, i.e. when
F cos GPR sin
that is, when
F
PR

Gtan

However
F
PR

G Gtan f

where f is the friction angle, from which it follows that the bar cannot
be drawn into the rolls when the contact angle exceeds the friction
angle f.

10.2.2 Neutral point and no-slip angle


It will be seen from Fig. 10.1 that at point B, which lies on the contact
surfaces between the neutral point and the plane of exit of the bar from
the rolls, the frictional force F and also the horizontal component HE
of the resultant of PR and F are in an opposite direction to those at A.
All the forces acting on the bar can be replaced by an algebraic sum
of all the horizontal components and the resultant of the vertical components, and these two forces will be components of the nal resultant
of the whole system of forces acting on the bar. This resultant will be
vertical only if the algebraic sum of all the horizontal components is
zero. Ekelund (1) assumes this is the case when the roll speed is uniform,

396

Rolling Contacts

but this neglects the fact that the kinetic energy of the bar leaving the
rolls is always (unless the spread is sufciently large) greater than the
kinetic energy on entering the rolls. The force required to produce this
increase in kinetic energy is equal to the resultant horizontal force, say
H. If K is the resultant of all the forces exerted by the rolls on the bar
and P is its vertical component, then K will be inclined at an angle
to the vertical where tan GHP. Similar reasoning applies when the
rolls and bar are accelerated after entry. The resultant K is rarely, if
ever, absolutely vertical, but in practice the forces H required to accelerate the bar are so small in comparison with P that may be taken as
equal to zero, i.e. the resultant K may be assumed to be vertical without
any appreciable departure from the truth. Hence, in practical rolling
under the conditions specied it may be stated that the resultant force
between the rolls and the bar is vertical and that the sum of the horizontal components of K on the left of the neutral point must equal the
sum of the horizontal components on the right of the neutral point
(Fig. 10.1). The position of the neutral point is determined by these
conditions.
In the case where, after entry of the bar, the rolls are accelerated, the
neutral point moves to a position nearer the line joining the roll centres,
the exact position of the neutral point being determined by the condition that the resultant of the horizontal forces acting on the bar in
the direction of rolling equals the force required to accelerate the bar.
It must be clearly emphasized that, in the foregoing discussions and
in what follows, the assumption that plane vertical sections of the bar
remain plain during rolling that is, the rolls slip on the bar everywhere
except at the neutral point has been made, and also it has been postulated that the radial roll pressure pr is constant along the contact arc.
If the product of the coefcient of friction and the radial roll pressure
equals or exceeds the yield stress in shear of the material of the bar
under conditions obtaining in the roll gap, then, in the region of the
arc of contact in which this condition holds, the surface of the bar and
the roll surface will move together without slipping, and the neutral
point will become a neutral zone. In general, the radial roll pressure is
not constant but increases from the plane of entry to a maximum value
somewhere along the arc of contact, and then decreases to the plane of
exit. If this type of pressure distribution is accompanied with a neutral
zone, then the neutral point loses its original signicance and is best
dened as a neutral plane, the position of which coincides with the
plane of maximum radial roll pressure. It is still true, however, that the
friction drag between the bar and the rolls vanishes at the neutral plane.

Rolling in Metal Forming

397

10.2.3 Expressions for the no-slip angle


If the angle of contact is denoted by and the angle between the radius
to the neutral point Y and the line of roll centres is (Fig. 10.1), then
angle is dened as the no-slip angle and can be found from the condition of equilibrium of the horizontal components HE stated above for
slipping friction. This condition may be written as

HE G HE
1

(10.1)

If f is the friction angle, that is, Gtan f, and is the angle between
the radius to any point on the arc of contact and the line joining the
roll centres, then, from Fig. 10.1, equation (10.1) is equivalent to

K cos(2AfC0)G K cos(2AfA0)
Since
KG1(P 2RCF 2)G1(P 2RCF 2)

pr R d

G1(1C )p R d
2

PR

then

sin( fA ) d G

sin( fC ) d

After integration
sin G

cos( fA )Acos f

(10.2)

2 sin f

Equation (10.2) may be reduced to the approximate form

1
G A
2 2

(10.3)

where and are in radians.


Since equation (10.2) may also be written as
sin G

sin sin2( 2)
A
2

and
sin G

1[R(h1Ah2)]
G
R

2(h1 Ah2)
D

398

Rolling Contacts

The expression 1[R(h1Ah2)] for the projected arc of contact is a


common approximation of the accurate expression

(h1Ah2)2

R(h1Ah2)A

The error introduced by neglecting the term (h1Ah2)24 is less than


5 percent when (h1Ah2)R is less than 0.40.
Equation (10.2) may be expressed in the approximate form

h1Ah2
2D

1 h1Ah2

(10.4)

2D

where D is the diameter of the rolls and is again in radians.

10.2.4 Maximum value of the no-slip angle


The formula for the no-slip angle derived by Dahl (2) is in the form
sin G

sin Ccos A1 1
2

(10.5)

Using equation (10.5), it is possible to examine the conditions that make


G0. These are either G0 or sin G1Acos . From the latter
expression

Gtan fG

1Acos
sin

Gtan

and hence the no-slip angle is zero when G0 or G2f. It can also
be shown that is maximum when Gf, for then d(sin )dxG0.
Putting Gf in equation (10.5), the maximum value of the no-slip angle
is given by
1
f
sin max G tan
2
2

(10.6)

Equation (10.6) may also be expressed in the approximate form

max G

f
4

(10.7)

Since the bar will not enter the rolls unaided if the angle of contact
is greater than the friction angle f, it follows that the maximum value
of the no-slip angle is attained when the bar just enters the rolls
unaided.

Rolling in Metal Forming

399

It also follows from the above that if the bar is pushed into the roll
gap, rolling is possible with contact angles greater than f, the maximum
possible contact angle being reached when G2f, which corresponds
to a zero value of the no-slip angle . If rolling were attempted with a
greater angle of contact than this, the rolls would slide over the bar.
It should be noted that equations (10.2) to (10.7) have been derived
using the following assumptions: (i) the radial roll pressure pr is uniform
along the arc of contact, and (ii) the coefcient of external friction is
constant. As already pointed out, assumption (i) means that the formulae are strictly accurate only when the draught is small in relation to
the bar thickness.

10.2.5 Rolling when bar motion is impeded


If the angle of contact is less than the friction angle f, the bar is
drawn between the rolls unaided. Consider now the case in which the
motion of the bar is impeded so that it has to overcome a resistance.
Then, for equilibrium of the forces acting on the bar, the neutral point
must move towards the line of roll centres. The maximum value of the
resisting force that can be overcome by the issuing bar is reached when
the no-slip angle G0, for then the sum of the horizontal components
of the forces acting on the bar in the direction of rolling is maximum.
Since the resultant horizontal component at any point in the arc of
contact between the plane of entry and the neutral point is
2pr R cos d A2pr R sin d
the maximum value of the resisting force (RF) is
RFG2pr R

cos d A2pr R

sin d

G2pr R( sin Ccos A1)

(10.8)

In order to obtain an expression for the maximum reduction possible


with a pair of idle rolls, assume that the peripheral speed of the idle
rolls is equal to the speed of entry of the bar between them. Then the
no-slip angle ( 1) is equal to the angle of contact and the force opposing
the forward motion of the bar (RF1) is
RF G2 p R
1

1
r

G 1

cos d A2p R

G2p1r R1( sin Ccos A1)

1
r

G 1

sin d

(10.9)

400

Rolling Contacts

where R1, 1, and p1r all refer to the idle rolls. The forces given by
equations (10.8) and (10.9) must be equal for equilibrium, and hence if
R, R1 , pr, p1r , , and are known, 1, and consequently the draught
on the idle rolls, may be calculated.
The above reasoning neglects the energy required to overcome the
frictional losses at the roll necks of the idle rolls, and because of this
the draught obtained in practice will be less than that calculated.

10.3 Forward slip during rolling


10.3.1 Introduction
Forward slip is dened as the ratio (V2Av)v, where V2 is the velocity
of the material leaving the rolls and v is the peripheral velocity of the
rolls. This is the most obvious and easily measured manifestation of a
factor that is of such vital importance in rolling, i.e. external friction.
From it valuable information on the magnitude and variation of the
mean coefcient of external friction can be deduced by comparing
experimentally determined relationships between slip and other quantities with those predicted by theory on the basis of known assumptions.
It will become apparent later that the manner in which changes
from point to point along the arc of contact has a considerable effect
on the magnitude and nature of the pressure distribution between the
rolls and the material and consequently on the power consumption.
Direct experimental investigation of this frictional phenomenon
presents very considerable difculties, and therefore any measurement
capable of illuminating this particular subject is of value. Forward slip
comes within this category and for this reason alone is a matter of
importance.
10.3.2 Expressions for forward slip
An expression for forward slip in terms of the no-slip angle is readily
obtainable if it is assumed that initially plane vertical sections of the
bar remain plane during rolling, that the density of the bar is constant,
that lateral spread is negligible, and that the rolls are rigid. Referring
to Fig. 10.2, let V1 be the velocity of the bar entering the rolls and V2
its velocity when leaving them, and let V be the horizontal velocity of
the bar in the plane of the neutral point Y. On the above assumption
relating to plane sections of the bar, these velocities will be uniform at
any vertical plane throughout the thickness of the bar.
Let v be the peripheral velocity of the rolls, then the point on the
surface of the bar opposite the no-slip point Y is moving with a velocity

Rolling in Metal Forming

401

Fig. 10.2

v in a direction tangential to the roll, and hence V Gv cos . From the


condition of constant density
V1h1 GV2h2 Gvh cos

(10.10)

where h is the height of the bar at the neutral point and h1 and h2 are
the thicknesses of the bar at the entry to and the exit from the contact
respectively.
From Fig. 10.2
h Gh2C2R(1Acos )

(10.11)

and hence from equations (10.10) and (10.11)


V2 G

1
h2

v cos [h2CD(1Acos )]

(10.12)

402

Rolling Contacts

Thus, forward slip (FS) can now be expressed as


FSG(1Acos )

D cos

h2

A1

(10.13)

As an approximation this may be written as

1
D
FSG 2 A1
2
h2

(10.14)

where is in radians.
Since the no-slip angle reaches a maximum value when the angle
of contact equals the friction angle f, as shown earlier, it follows from
the above equation that forward slip will also be maximum for this
value of . It was shown previously that the maximum value of is
f4, and hence the maximum forward slip (MFS) is given by the
equation
MFSG

f2 D
A1
32 h2

(10.15)

In Fig. 10.3, two curves have been plotted to show the relationship
between the forward slip, expressed by the approximate equation
(10.14), and the angle of contact for two values of the coefcient of

Fig. 10.3

Rolling in Metal Forming

403

external friction G0.2 and G0.3. In the case presented here, the
thickness ratio h2D is 0.1. The friction angles corresponding to G0.2
and G0.3 are 1119 and 1642 respectively, and consequently the
maximum slip occurs in the two cases where the angle of contact has
these values. This fact is sometimes used to determine the coefcient of
friction between the rolls and the material, as it is possible to roll with
an angle of contact greater than the friction angle if the bar is assisted
in entering between the rolls by being pushed against them, or in the
case of strip, having its entering edge tapered.
In Fig. 10.4 the relationship between forward slip and percentage
reduction given by equations (10.2) and (10.14) is plotted for a thickness
ratio of 5.6 percent for various values of the coefcient of external
friction. These curves show that the percentage forward slip increases
with the coefcient of friction and is a function of the percentage
reduction. Forward slip reaches a maximum of about 2.1 percent when
G0.2 and 3.2 percent when G0.25 at reductions of about 25 and 35
percent respectively. With G0.4, forward slip increases uniformly up

Fig. 10.4

404

Rolling Contacts

to reductions of 40 percent when the forward slip is about 7.3 percent.


There are other formulae for forward slip calculation. Dresden (3)
proposed a formula for forward slip in terms of the no-slip angle ,
which is applicable only to cases in which is small. Thus
FSG

D
h2

2 sin2

D 2

2 2h2

(10.16)

This expression applies only to cases where Dh2 is large compared with
unity.
Caswell (4) derived the following expression for the no-slip angle

Gcos1 h2CDJ

h22CD2A2h1(DAh1Ah2)
2D

(10.17)

and, for the velocity V2 with which the bar leaves the rolls, the following
expression is proposed
V2 G

v(h1 cos Ch2 )


2h2

(10.18)

from which
FSG

h1 cos Ah2
2h2

(10.19)

It will be noted that, according to equations (10.17) and (10.19), both


the position of the neutral point and the magnitude of forward slip are
independent of the coefcient of external friction between the rolls and
the bar.

10.4 Friction between the rolls and the material


10.4.1 Role of friction in rolling
The frictional forces acting between the material and the rolls are very
important. It is usual practice to assume constancy of the coefcient of
external friction throughout the arc of contact. It is interesting to examine, in some detail, the state of the art in this area with a view to discovering what is known of the manner in which this quantity varies in
different types of rolling, the factors on which it depends, the means
available for measuring it under rolling conditions, and the accuracy
with which the coefcient of friction may be forecast for any specic
rolling operation.

Rolling in Metal Forming

405

The quantity that determines whether the material will slip on the
rolls or move with them as a result of stick is the value of the product
of the coefcient of friction and the radial roll pressure pr. If pr is
less than the yield stress in shear of the material at the point of the arc
of contact considered, then the rolls slip on the material at this point.
If it is equal to or greater than the yield stress in shear, then the material
moves with the rolls. Once this critical value of pr is reached, the
subsequent changes in become of no consequence provided pr does
not fall below the critical value. In other words, the value of in the
neutral zone is of no particular interest, and attention should be conned to sliding friction. Research on sliding friction has revealed that
its magnitude depends on the material of the surfaces of contact, on the
condition and the area of the surfaces, on their temperature and relative
velocities, and on the pressure with which they are pressed together. In
both hot and cold rolling many of these factors are subject to variation.
Thus, in hot rolling, the presence of scale on the bar, the nature of the
scale, and lms of water or steam between the bar and rolls, as well as
variations in the surface of the rolls, are disturbing factors that may be
present. In cold rolling with a lubricant, the latter may be partially
squeezed out in the regions of high pressure or may collect into patches.
In both hot and cold rolling there is, in general, a non-uniform pressure
distribution between the rolls and material, as well as variation in the
rubbing velocity along the arc of contact, to which reference has already
been made. Friction in rolling is thus by no means a simple matter, and
its investigation presents a number of difculties.

10.4.2 Friction in hot rolling


Experiments show (5) that the coefcient of friction between a hot
steel slab sliding over a smooth cold steel plane is approximately 0.4.
Tafel and Schneider (6, 7) determined for smooth rolls at various
rolling speeds when rolling steel billets 100 mm thick, 180 mm wide, and
1000 mm long at a constant temperature of 1200 C. In order to measure , Tafel and Schneider utilized the fact mentioned earlier that, when
the contact angle equals the friction angle f, the rolls cease to grip
the bar. The billets were carefully rolled to the exact dimensions specied above, and hot sawn to length so that no burrs were left on the
ends. The billets were then heated in a furnace and, when at rolling
temperature, were brought out on to the roller track and up to the rolls,
which had previously been speeded up to the required steady speed, and
the gap between them adjusted so that the billet could not enter. While
the live roller table was kept turning slowly to keep the billet in contact

406

Rolling Contacts

with the rolls, the top roll was slowly raised and the temperature of the
bar was monitored. When the correct gripping angle was reached, the
rolls gripped the bar and rolled it. After cooling, the bar was measured,
and from the difference in thickness before and after the pass, and from
the mean diameter of the upper and lower rolls, the contact angle 1 at
which the rolls gripped was calculated. This angle was taken as equal
to the friction angle f, so that Gtan 1.
The results of these tests are shown in Fig. 10.4, in which the gripping
angle and are plotted against the peripheral speed of the rolls. It is
interesting to note that the rapid decrease in the gripping angle,
amounting to 7, occurs in the range of rolling speeds between 2.2 and
2.7 ms. The cause of this sudden decrease in gripping angle could be
attributed to local melting of the steel by frictional heating. This explanation could be supplemented by the fact that, if a fast-running cold
roll, or a saw blade without teeth, rubs against a piece of hot steel, the
friction melts the steel locally and the coefcient of friction decreases
almost to zero. The practical signicance of the rapid drop in the value
of shown in Fig. 10.3 can be interpreted as follows. In the range of
rolling speed in which the drop occurs, the bar is suddenly gripped, and
such high shock forces are produced as to be a source of danger to the
rolls. Within this speed range, therefore, it is inadvisable to use the
maximum gripping angle. The extent of the danger zone depends mainly
on the rolling pressure, the rolling temperature, and the condition of
the material rolled. For practical purposes, therefore, the maximum
gripping angle can only be used with smooth rolls up to a rolling speed
of about 1.8 ms.
A further point of interest is the value of for very low rolling speeds.
In this connection, Tafel and Schneider made several tests at a speed
that can be described as slightly above 0 ms, and the points corresponding to these tests all fall below the curve in Fig. 10.4. The tests
therefore give no indication of an increase in at, or near, zero velocity.
Tests under dry conditions show that, in general, increases at low
velocities and that stiction is greater than friction. Tests conducted by
Tafel and Schneider were, however, hardly sensitive enough to show
such variations at low velocities.
Ekelund (1) determined for grey cast iron rolls during the rolling
of a soft steel at various temperatures from 720 to 1110 C, using the
same method as Tafel and Schneider, but with some modications in
the technique. Ekelunds results are shown in Fig. 10.5, in which is
plotted against rolling temperature, and the points lie approximately
about a straight line. From this curve the relation between and rolling

Rolling in Metal Forming

407

Fig. 10.5

temperature can be deduced as G1.05A0.0005T, where T is the rolling temperature (C).


The increase in for decreasing temperature, i.e. from 1110 to
700 C, in hot rolling is the reason for the fact that colder steel ingots
are gripped and rolled more certainly than hotter ones, and watercooled rolls grip better than hot dry ones, since the surface of the rolled
material is cooled by the roll surface, and so and the gripping angle
both increase.

10.4.3 Friction in cold rolling


The effect of the condition of the roll surface on was investigated by
Lueng (8) for steel strip rolled between chromium steel rolls of 185 mm
diameter at a roll speed of 36 rmin. Three sets of tests were run,
namely: (a) with rolls articially roughened by sand blasting, (b) with
smooth, dry rolls, and (c) with smooth rolls lubricated with emulsion
oil. Here, smooth means a hardened, ground, and polished surface.
The following values of were obtained:
(a) articially roughened rolls G0.15,
(b) smooth, dry rolls G0.090.11,
(c) smooth, lubricated rolls G0.07.
Trinks (9), summarizing his work on friction in cold rolling, stated that
0.04 was the lowest value of in his experiments, and this was obtained

408

Rolling Contacts

by high-speed rolling and ood lubrication with an emulsion of soluble


oil. The highest value of known to Trinks was 0.15, obtained by
Lueng with articially roughened rolls and reported earlier. From
experiments conducted by Lueng it is possible to conclude that, with
lubrication and slow speed, is hardly lower than for dry rolls, probably because there is time to squeeze out the lubricant. Values of
determined by the coned compression test suggest, however, that may
reach values of 0.4 or 0.5 for rough dry rolls.

10.4.4 Evaluation of friction measuring methods


Basically there are three methods for determining the coefcient of
friction between the material and the rolls:
(a) maximum gripping angle,
(b) maximum forward slip,
(c) measurement of roll load and forward pull on the bar when the
rolls are sliding over the bar.
In all these methods it is assumed that is constant throughout the
surface of contact between the rolls and the material, but this assumption may not be true.
In method (a) the value of determined is that corresponding to the
friction between the upper and lower front edges of the bar and the roll
surfaces at the point of entry only, and for the particular rubbing velocity corresponding to the roll speed at which the test is conducted. This
value may not represent the true mean value of for the whole arc of
contact, as the rubbing velocity decreases from its value at entry to zero
at the no-slip point, is then reversed in direction, and rises to its value
at exit. Also, the pressure distribution may not be uniform. The maximum angle of grip is affected by the magnitude of the force with which
the bar is pressed against the rolls, and unless this force is kept small
and constant, variations in the measured values of may arise because
of this. The method is also not applicable to rolls having non-uniform
surface smoothness, such as may be produced by rough patches.
Regarding method (b), the determination of from forward slip
measurements requires the use of formulae connecting forward slip with
the no-slip angle , and the no-slip angle with . Many of these formulae have already been discussed, and it will be remembered that in all
cases it is assumed that plane vertical sections of the bar before rolling
remain plane during rolling, that the value of is constant along the
arc of contact, and that the pressure distribution is uniform.

Rolling in Metal Forming

409

In rolling a bar in which the draught is small in relation to its height,


the pressure distribution is approximately constant, but plane sections
do not, in general, remain plane. If the error introduced by this departure from the assumption is small, as in many cases it may be, then
forward slip measurements may give a reasonable mean value of .
Method (c) enables to be determined only for the special case in
which the neutral point lies just outside the plane at which the material
leaves the rolls, and the rolls, therefore, slide over the bar. Consequently, it is not representative of normal rolling conditions. It does,
however, enable a value of to be found under conditions of heavy
pressures such as occur in rolling, and, by varying the roll speed, the
effect of the velocity of sliding on can be studied. Use of the method
would appear to be restricted to cold rolling, and it is doubtful whether
the effect on of a lubricant applied to the rolls would be representative
of that occurring in actual rolling, owing to the dissimilar sliding
conditions and pressure distribution.

10.4.5 Coned compression tests


The values of the coefcient of friction so far considered have all been
derived from actual rolling tests. Attention must now be directed to
experiments specially devised to simulate rolling conditions so far as
pressure is concerned, but which are free from other complications
inseparable from rolling tests.
In one form of test commonly used, a cylindrical specimen that has
a conical depression turned in each of its ends is placed between two
conical compression plates which exactly t the conical depression in
the cylinder. The arrangement is shown, schematically, in Fig. 10.6. The
combination of specimen and coned compression plates is then placed
in a testing machine and loaded until the cylinder is deformed plastically, and the shape of the sides of the cylinder after compression is
examined. If the sides remain straight, then the angle (see Fig. 10.6)
is the friction angle between the compression plates and the cylinder,
so that the coefcient of friction Gtan fGtan . That this is so may
be proved as follows. Referring to Fig. 10.6, let P be the normal force
acting on an element of the area on the end of the cylinder. Then the
horizontal component of this force is P sin and acts radially outwards. As compression proceeds, the material ows over the compression plates radially outwards so that a friction force P acts radially
inwards along the interface of the cone plate and the specimen. The
horizontal component of this force equals P cos .

410

Rolling Contacts

Fig. 10.6

Now, if a cylinder with at ends at right-angles to the axis of the


cylinder is compressed between two smooth and frictionless plates, the
compression is homogeneous, as there is no force acting tangential to
the cylinder ends and restricting lateral ow. Hence, the walls of the
cylinder move out parallel to themselves. If there is friction between the
cylinder ends and the plates, the specimen assumes a barrel shape during compression, owing to the restricting force opposing the radial
movement of the cylinder ends.
Reverting again to Fig. 10.6, it will be seen that, if angle is chosen
so that P sin GP cos , the outward force due to P just balances the
inward radial force due to friction, so that there is no resultant horizontal radial force acting on the cylinder ends, and the arrangement
corresponds to compression between frictionless at plates. Hence, if in
the compression test with coned end plates the walls of the cylinder
remain straight and parallel during compression, then the following is
applicable
P sin GP cos
or

Gtan

Rolling in Metal Forming

411

It will be noted that in this method of nding it is assumed that is


constant over the whole surface of contact between the specimen and
the coned plates, notwithstanding the fact that the radial velocity of
ow of the material over the plate varies from zero at the centre to a
maximum at the circumference of the cylinder, which depends on the
rate of compression and the diameter of the cylinder. This type of test
enables to be measured under pressures equal to the yield stress
of the material with free spread, and it is also possible, by treating
the surface of the cone plates in different ways, to simulate various
conditions of roll surface. Thus, they may be lubricated, smooth and
dry, or roughened.
Siebel and Pomp (10) used the coned compression test to determine
the coefcient of friction between cylinders of lead, aluminium, copper,
and wrought iron, and compression plates of chromium steel. The surfaces of the latter were smooth, etched with acid to produce a rough
surface corresponding to a worn roll, and sand blasted to represent a
very rough roll surface. The results of these tests are shown in Table
10.1. As would be anticipated, is much affected by the condition of
the surface of the compression plates, but only slightly by the material
of the cylinder.
Compression tests on cylinders between smooth, polished, coned steel
dies have also been carried out (11) to determine the coefcient of friction during plastic deformation at elevated temperatures. Cylindrical
specimens of 17 mm diameter and 25 mm length were compressed in an
Amsler testing machine at a speed of 6 mmmin to a nal length of
17 mm. The testpieces and the coned compression plates were pressed
together and surrounded by a jacket in which a suitable liquid was
placed, the nature of the liquid depending on the temperature at which
the test was to be conducted. The liquid was then heated to the required
temperature and the load applied to the specimen. The results of a series

Table 10.1 Values of l between cylinders of various metals and


chromium steel coned compression plates (10)
Coefcient of dry friction
Material
Lead
Aluminium
Copper
Wrought iron

Smooth
F 0.1
0.1
0.080.09
0.1

Rough (acid treated)

Very rough

0.3
0.3
0.3
0.25

0.4
0.4
0.4
0.32

412

Rolling Contacts

Table 10.2. Coefcient of external friction at elevated


temperatures
Metal

Temperature (C )

Friction coefcient

Copper
Nickel
Lead
Bismuth
Cadmium
Tin
Zinc
Aluminium

750
900
180
150
180
100
110
375

0.36
0.32
0.33
0.27
0.24
0.18
0.17
0.74

of tests on specimens of various metals are shown in Table 10.2. Since


the ends of the cylindrical specimens and the coned compression plates
were pressed together before the liquid was placed in the jacket, the
values of in Table 10.2 may be taken as referring to the friction
between dry surfaces at the temperatures indicated.
Comparing the values of for smooth surfaces given in Tables 10.1
and 10.2 for copper, lead, and aluminium, it will be noted that is
considerably greater at relatively high temperatures than at room temperature, the values at the high temperatures with smooth compression
plates corresponding approximately to those with rough compression
plates at room temperature.

10.4.6 Friction coefcient variation along the arc of contact


So far it has been assumed that is constant along the arc of contact.
However, it is quite possible that the friction coefcient may vary from
point to point along the arc of contact, even in the ideal case where the
surfaces of the bar and rolls are uniformly smooth, the bar is free from
scale, and the only variables are the rubbing velocity and the pressure
distribution.
Because of the extreme difculty of making measurements of at
various points in the arc of contact, there is apparently no direct experimental information available on the subject. Indeed, few measurements
of under such high pressures as occur in rolling have been made, and
there is little data that can be used as a guide as to the probable variation in in the arc of contact, even if the pressure distribution is
correctly known. In the absence of more positive information, a number
of researchers have assumed a law for the variation in along the
arc of contact, and have investigated the consequences on the pressure
distribution.

Rolling in Metal Forming

413

In the case of hot rolling, Hitchcock (12) proposed such a law. He


stated that experiments showed that between sliding bodies is greatest
at low velocities and decreases as the velocity increases. Hence, he
assumed that it is greatest at the neutral point. This led him to postulate
the following expression for
R2

GvA

(sin Asin )2

(10.20)

where is the coefcient of friction at any point to which the radius


makes an angle to the line joining the roll centres, v is the coefcient
at the no-slip point, and R, , and are as above.
According to this expression, is greatest at the neutral point, i.e.
when G , and decreases as becomes larger or smaller than .
Hence varies with rubbing velocity. There is, however, nothing to
justify the precise form of this expression.
Nadai (13) investigated the consequences of three frictional hypotheses on the distribution of pressure between the rolls and material when
rolling thin material. The three hypotheses are:
(a) Gconst;
(b) the surface frictional force (tangential to the rolls) is constant, i.e.
pr dxGconst;
(c) surface friction is assumed to be proportional to the relative velocity
of slip between the rolls and the material.
Assumption (a) is that usually made in theories of rolling and, in the
case of thin strip or sheet rolling, leads to the well-known sharp peaked
pressure diagram. Assumption (b) also leads to a peaked pressure diagram, but the peak is less sharp than in the previous case because
decreases as the pressure builds up.
In developing assumption (c), Nadai states that little is known about
the surface friction when lubricated sheets are rolled. It is quite certain,
however, that the lubricant must pass through a region of exceedingly
high pressure in its passage between the rolls, and also that its temperature rises on account of contact with the warm rolls and the heat generated in the sheet by the deformation work. Whether, under these
conditions, the oil remains as a lm on the sheet or separates out into
patches, leaving dry areas, is uncertain. The viscosity of the lubricant,
however, will decrease owing to its rise in temperature, and increase on
account of the rise in pressure.
Although observations are not available to substantiate the assumption, it is interesting to examine the shape the pressure line will take

414

Rolling Contacts

if surface friction varies as assumed. The constant that expresses the


proportionality may contain the viscosity and the thickness , of the
oil lm, and may be taken as a mean value of the viscosity at the
high rolling pressures. The surface friction is then expressed by the
relation

G (vAV )

where V is the peripheral velocity of the roll and v is the tangential


velocity of a point on the surface of the bar in the arc of contact.
The pressure distribution resulting from this assumption and the derivation of the equation of the pressure line are discussed later on in this
chapter. The point of interest is that the sharp peak has disappeared
and the top of the pressure diagram is rounded. The pressure diagrams
obtained in the experimental investigation of the pressure distribution
in the region of contact by Siebel and Lueng (14) in all cases gave a
rounded peak. The rounding was due, at least in part, to the nite
dimensions of the measuring pin used, but, even after allowance had
been made for this, the peak remained somewhat rounded. It is therefore possible that the surface friction in the arc of contact varies in a
manner somewhat similar to that assumed in hypothesis (c) by Nadai.
It will be noted that, according to this assumption, the surface friction,
and hence , is zero at the neutral point and greatest at entry, the
reverse of the assumption of Hitchcock for hot rolling.
The analytical study of the variation in the coefcient of friction
along the arc of contact by Brown (15) should be noted. Brown based
his calculations on the test carried out by Siebel and Lueng (14), in
which the pressure distribution along the arc of contact was measured
experimentally. The strip was rolled without lubrication, and the radial
pressure found in this test is reproduced in Fig. 10.6. In order to calculate, from this curve, the value of the coefcient of friction at various
points along the arc of contact, it is assumed that the pressure distribution was correctly given by the theory proposed by von Karman and
described later in this chapter. It is also assumed that the correct value
of was used at each point of the arc of contact. In other words, Brown
assumed that the assumptions on which von Karmans theory is based
held for this particular case, and found at each point in the arc the
value of that must be inserted in von Karmans equation to give the
measured value of the radial pressure pr.
If h is the thickness of the strip at any point in the arc of contact at
a distance x from the line of roll centres and S is the constrained yield

Rolling in Metal Forming

415

stress of the material, then it can easily be deduced from von Karmans
equation that
2 pr GS

dh

Ah

dx

dpr
dx

(10.21)

on the entry side of the rolls and


2 pr Gh

dpr

AS

dx

dh
dx

(10.22)

on the exit side of the rolls.


Since
R(hAh2)Gx2
therefore
dh
dx

2x
R

(10.23)

The value of dhdx was calculated for each point from this equation,
and the values of dprdx were found by measuring the slope of the
measured curve of pr plotted against the contact angle; the value of h
at each point was readily calculated. The quantity S was experimentally
found by Siebel and Lueng (14) and is plotted against contact angle in
Fig. 10.7. By inserting corresponding values of dhdx, dprdx, x, and S
in equations (10.21) and (10.22), was found for the point considered.
The resulting values of are plotted in Fig. 10.7, from which it will
be noted that this method suggests that varies considerably over the
contact arc, having a value of about 0.3 at entry, which falls fairly
steadily to about 0.15 near the centre of the arc, becomes zero at the
neutral point, and reaches a value at the plane of exit of 0.2. The coefcient of friction and the friction drag drop momentarily to zero at or
near the neutral plane, because the measured pressure peak is rounded.
With constant along the arc, von Karmans equation gives a sharp
peak. If, however, there is a non-plastic cone at the neutral plane, then
a rounded peak can be obtained without becoming zero at the neutral
plane, and in lieu of denite evidence to the contrary it seems that this
must be what occurs.
If were constant along the whole arc of contact, it would have to
have a value of about 0.14 if the pressure distribution given by von
Karmans method was to approximate to the experimental pressure
curve.

416

Rolling Contacts

Fig. 10.7

Summarizing the results, it appears that the methods employed for


the determination of the coefcient of friction between the rolls and the
material are all open to certain objections and involve various assumptions that may, or may not, be justied according to the rolling conditions. Experimental values of must therefore be accepted, with
reserve, as the best available at present.

10.5 Theories of metal rolling


10.5.1 Introduction
The thin sheet and strip theories of rolling that are presented in this
section should offer a reasonable explanation of the inuence on rolling
load and mean specic pressure of the various factors playing an
important part in the process. Chief of these effects are:
(a) the increase in rolling load and specic roll pressure with increase
in the coefcient of friction;

Rolling in Metal Forming

417

(b) the increase in rolling load and specic roll pressure with increase
in roll diameter;
(c) the increase in specic roll pressure with decreasing initial thickness
of the strip;
(d) the decrease in rolling load and specic roll pressure with the application of back and front tension.
In addition, the theories should give a pressure distribution between the
rolls and the material that is in reasonable accord with that revealed by
the experimental work of Siebel and Lueng (14), and be capable of
readily allowing for the effect of the nature of the rolled material on
the rolling load and specic roll pressure. In hot rolling, the effect of
rolling temperature and rolling speed, and in cold rolling, the effect
of work hardening, should also be allowed for by the theories.
The theories discussed in this section all full the above requirements
at least qualitatively, and, with the exception of that of reference (16),
are all based on the same assumption, having as their starting point a
differential equation that represents the condition of equilibrium of an
elementary vertical plane section of the strip between the rolls. It is in
the development of this equation that these theories mainly differ.
One of the rst and simplest theories was proposed by Siebel (17, 18)
and has become known as the theory of the friction hill. While this
explains in a general way the effect of the various factors on specic
roll pressure, it is unsuitable for the calculation of roll pressures as it
neglects the accumulative effect of pressure on the friction hill and
hence gives rise, in general, to a peak pressure that is too low.

10.5.2 Equation describing the friction hill


The main differences in the studies of various researchers lie in the
assumptions they make, and the methods they employ, for integrating
the differential equation that represents the condition of equilibrium of
an elementary section of the material between the rolls. In order to
avoid tedious repetition, this differential equation is rst derived and
then used as a common starting point for the consideration of the work
of the various investigators.
Referring to Fig. 10.8, consider a sheet of material of initial thickness
h1, which is rolled between two plain parallel rolls of equal radius R to
a nal thickness h2 without the application of either front or back tension. It is required to nd an expression for the vertical pressure p
between the rolls and the sheet for any point in the arc of contact XYZ.

418

Rolling Contacts

Fig. 10.8

The following assumptions are made:


(1) The sheet does not spread laterally. This condition will be approximately satised if the thickness of the sheet is small compared with
its width.
(2) The coefcient of external friction between the sheet and the rolls
is constant at all points in the arc of contact.
(3) Plane vertical sections of the sheet before rolling remain plane
during rolling.
(4) There is no elastic deformation of the rolls in the arc of contact.
(5) The elastic deformations of the sheet are negligible in comparison
with the plastic deformations, and the material of the sheet is
homogeneous.
(6) The criterion of plastic yielding holds for the material rolled and
for the conditions of rolling.
(7) The compressive strength is constant throughout the contact length.
In cold rolling this requires the material not to undergo work hardening during its passage between the rolls, and in hot rolling it
requires the change in compression rate from point to point along

Rolling in Metal Forming

419

the arc of contact not to have any effect on the magnitude of the
compressive strength.
(8) The peripheral velocity of the rolls is uniform, i.e. the rolls are
neither accelerating nor decelerating.
Referring again to Fig. 10.8, consider the forces acting on a vertical
elemental plane section of a sheet of width dx, measured in the direction
of rolling, and of height h, situated at distance x from the line joining
the roll centres and lying between the plane of entry XX and the neutral
plane YY. Let pr be the radial pressure between the roll and the sheet
acting on the end of the section dx, and let F be the corresponding
tangential frictional force. Let the line joining A to the upper roll centre
make an angle with the line joining the roll centres.
Then, considering the unit width of the sheet, the normal force on one
end of the section dx is pr(dxcos ), and the horizontal component of
this force, which tends to oppose entry of the sheet between the rolls, is
pr

dx
cos

sin Gpr tan dx

The tangential frictional force is


FG pr

dx
cos

where is the coefcient of external friction; hence the horizontal


component of this force, which tends to draw the bar between the rolls
is

pr

dx
cos G pr dx
cos

(10.24)

The horizontal forces acting on the vertical faces of the section dx produce compressive stresses which are assumed to be uniformly distributed over the height of the section. Let these stresses be Cd acting
on a face of the section of height hCdh, and acting on a face of
height h.
Then, for the equilibrium of the section dx, the resultant of the forces
due to these stresses must balance the difference of the horizontal components of the radial force and the frictional force when both ends of
the section dx are considered. Hence, for points in the arc of contact
between XX and YY
2pr tan dxA2 pr dxG(hCdh)( Cd )Ah Gd(h )

(10.25)

420

Rolling Contacts

Now, from Fig. 10.8 the force p dx is the vertical component of the
force pr(dxcos ), and therefore
p dxGpr

dx
cos

cos Gpr dx

hence
pGpr

(10.26)

Substituting p for pr in equation (10.25) and writing Gtan f, where f


is the friction angle, equation (10.25) becomes
p(tan Atan f ) dxGd

(10.27)

Considering now an elemental section of width dx lying between the


neutral plane YY and the plane of exit ZZ, the horizontal component
of the radial force is again pr tan dx, but the horizontal component of
the frictional force now equals pr dx, since the direction of the action
of the tangential frictional force has reversed at the neutral plane.
Hence, for sections between the neutral plane and the plane of exit, the
equation corresponding to equation (10.27) is
p(tan Ctan f ) dxGd

(10.28)

Equations (10.27) and (10.28) may be combined to represent the


conditions of equilibrium throughout the arc of contact. Thus
p(tan tan f ) dxGd

(10.29)

or
p(tan tan f )G

d(h 2)
dx

(10.30)

Now, from Fig. 10.8 it can be seen that the height of the section
increases by dh when x increases by dx, and hence
1 dh
2 dx

Gtan

Putting this expression for tan in equation (10.30) produces


1 dh
d(h 2)
p p tan fG
2 dx
dx

(10.31)

Rolling in Metal Forming

421

In order to express in terms of p, the criterion of plastic yielding can


be used if the mean vertical pressure p* and the mean horizontal stress
acting on the section dx are regarded as principal stresses, in this case
both compressive.
Now the total vertical force on the section dx is
p* dxGp dxCF sin Gp dxC pr tan dx
and from equation (10.26), since pGpr
p* dxGp(1C tan ) dx
Hence
p*Gp(1C tan )

(10.32)

Putting p*Gs1 and Gs3 in the criterion of yielding s1As3 GS produces


p(1C tan )A GS

(10.33)

Equations (10.31) and (10.33) can be combined to give a differential


equation in which p and x are the two variables, since h can be expressed
in terms of x. However, it is usually assumed that the value of tan
is small in comparison with unity, and may therefore be neglected. This
assumption limits the application of the theory to small angles of
contact, especially in hot rolling where may be as large as 0.6.
If tan is regarded as negligible, equation (10.33) simplies to
pA GS

(10.34)

and substituting the value for given by equation (10.34) in equation


(10.31) gives
d[h( pAS )2]
1 dh
p p tan fG
2 dx
dx

(10.35)

Equation (10.35), expressed as above, or in one of its alternative forms,


represents the starting point for analysis which will now be considered.

10.5.3 Theory of von Karman


The theory of rolling proposed by von Karman (16) was claimed to be
the rst in which the horizontal stress was introduced, as all previous
theories had regarded rolling as a squeezing process. In his theory, von
Karman considered the forces acting on an elemental vertical section of
the portion of the sheet between the rolls and, as a rst approximation,
replaced the radial pressure pr acting on the element by the vertical

422

Rolling Contacts

pressure p and assumed that x was small compared with R (see Fig.
10.8). He then obtained the equation
d(h 2)
dx

Gp(sin cos )

which can also be written as


d(h 2)
dx

Gp cos (tan tan f )

He next assumed that cos may be taken as approximately equal to 1,


so that
d

2 Gp(tan tan f ) dx

This last equation is identical to equation (10.30).


The general form of the pressure distribution between the rolls and
the material given by von Karmans equation is shown in Fig. 10.9. At
the planes of entry and exit, the pressure is equal to the constrained
yield stress of the material, for here G0 so that pGS. From the plane
of entry to the neutral plane the pressure increases owing to the horizontal compressive stress set up in the material by the friction between
the sheet and the rolls. Similarly, from the plane of exit to the neutral
plane the pressure increases. The maximum pressure occurs at the
neutral plane and the theoretical pressure diagram has a sharp peak.

Fig. 10.9

Rolling in Metal Forming

423

10.5.4 Simplication of von Karmans equation


In equations (10.31) and (10.35), 12(dhdx) represents the slope of the
arc of contact at any point at distance x from the line of the roll centres.
Now, if it is possible to replace the circular arc of contact with a curve
of another form that very nearly coincides with the circular arc so
that dhdx is a simple function of x, equations (10.31) and (10.35) will
be appreciatively simplied. The obvious choice of such a curve is a
parabola. Assuming a parabolic arc of contact, the equation for this,
referred to the longitudinal axis of the sheet and with the intersection
of the latter with the line joining the roll centres as the origin, is (see
Fig. 10.10)
1
2

hG

1
2

h2Cm

L
x

(10.36)

where mGh1Ah2 the draught, and L is the projected contact length.


Then, from equation (10.36)
1 dh
2 dx

Gm

(10.37)

L2

Inserting this value of 12(dhdx) in equation (10.31) gives


d(h 2) m
mx
G 2 pxp tan fGp 2 tan f
dx
L
L

Fig. 10.10

(10.38)

424

Rolling Contacts

It is interesting to note that equation (10.38) is the form of von


Karmans equation given by Keller (19), except that he gives the righthand side of the equation as

mx
p 2 Ctan f
L

The form given in equation (10.38) appears to be more logical as it is


the frictional force which changes sign at the neutral plane while the
horizontal component of the normal force is always of the same sign.
Since GpAS from equation (10.34) and hGh2Cm (xL)2 from
equation (10.36), equation (10.38) can be written as

d ( pAS )

dx Gp L tan f

1
x
h2C m
2
2
L

mx

which, after differentiation, becomes

h2 mx2 dp mx
mx
C 2
C 2 ( pAS )Gp 2 tan f
2 2L dx L
L

After rearrangement of the terms and simplication, this reduces to

d(pS )
G
d(xL)

2m x p L

tan f
h2 L S m
1C

mx2

(10.39)

h2L2

This form of von Karmans equation introduces an additional assumption to those already made in deriving equation (10.35), namely that
the arc of contact is parabolic.

10.5.5 Modication of von Karmans equation


Tselikov (20) introduced additional assumptions to the eight specied
earlier. As a result, he reduced von Karmans equation to a form that
can be integrated. Also, he suggested a method that enables the rolling
pressure to be easily calculated for small angles of contact. Equation
(10.29), which refers to points in the arc of contact between the plane
of entry and the neutral plane, was changed to the form
2p(tan Atan f ) dxGh d C dh
He also deduced equation (10.33), i.e.
p(1C tan )A GS

Rolling in Metal Forming

425

or

Gp(1C tan )AS

(10.40)

Differentiating equation (10.33) gives


d G(1C tan ) dp

(10.41)

Substituting these values for and d in equation (10.29), assuming


that dh(2 tan )Gdx and dividing by 1C tan , equation (10.29)
becomes
tan( fA )

1C

tan

pA1C tan dhCh dpG0


S

(10.42)

It is then assumed that, for small angles, may be treated as constant


and equal to half the contact angle.
Denoting
tan( fA )
tan

S
1C tan

GS1

and since has been assumed to be a constant, these quantities are also
constant. Then equation (10.42) can be written as
dp
(1C )pAS1

dh

(10.43)

This equation can be integrated and then becomes


1

1
ln[(1C )pAS1 ]Gln CC0
1C
h

(10.44)

Constant C0 is found from the entry conditions where hGh1 and pG


p1, and hence equation (10.44) becomes
pG

p1
h1
S1
(1C )A1
1C S1
h

1C

C1

(10.45)

If the material is rolled without either front or back tension, then, at


the plane of entry, from equation (10.40), G0, or
p1 G

S
1C tan

GS1

426

Rolling Contacts

Hence, for the case of rolling without tension, equation (10.45) reduces
to
pG

p1
1C

C1

h1

C1

(10.46)

For points in the arc of contact between the neutral plane and the plane
of exit, and for rolling without tension, similar equations hold. In this
case
pG

A1

h
p2

A1
h2

A1

(10.47)

where

tan( fC )

p2 G

tan
S
1A tan

It will be noted that p is the vertical pressure (see Fig. 10.8) due to the
vertical component of the radial force between the roll and the sheet
and does not include the pressure due to the vertical component of the
tangential friction force. The true vertical pressure between the roll and
the sheet will be p(1C tan ), but since has been assumed to be
constant, (1C tan ) must be assumed to be constant for all points in
the contact arc between the plane of entry and the neutral plane. This
assumption reduces the applicability of equations (10.46) and (10.47) to
small arcs of contact only, and makes proper representation of the part
played by the vertical component of friction impossible.

10.5.6 Effect of front and back tension on pressure


distribution
In the cases so far considered it has been assumed that the strip is rolled
without either front or back tension, and it has been shown that under
these conditions a vertical elementary section of the strip between the
rolls is subjected to a horizontal compressive stress which has been
denoted by . It follows from the assumptions outlined earlier that at
the planes of entry and exit G0. Hence, at these points, assuming no
work hardening, pGS, that is, the vertical pressure between the roll
and the strip at the plane of entry or exit is equal to the constrained
yield strength of the material rolled.

Rolling in Metal Forming

427

Fig. 10.11

The case that is going to be analysed is shown in Fig. 10.11, where a


front tension, tending to pull the material between the rolls, produces
a uniform tensile stress 2, in the strip. Let 1 be the corresponding
tensile stress set up in the strip by the back tension. Then, under these
conditions a vertical elemental section of the strip will still be subjected
to a horizontal stress and to a vertical compressive stress, p. At the
plane of entry XX this horizontal stress is tensile and is equal to 1. As
the element moves between the rolls, the tension 1 is opposed, or
reduced, by the horizontal compressive stress set up by the action of
friction between the rolls and strip. Hence, from the plane of entry XX
to the neutral plane, the tensile stress on the element decreases from its
initial value of 1 and may become a compressive stress. From the
neutral plane to the plane of exit ZZ it increases again to its nal value of
2, the front tension. The effect of this variation in the horizontal stress
on the vertical stress p can be found from the criterion of plastic yielding
s1As3 GS. At the plane of entry s3 GA 1 (the minus sign is adopted
because tensions are regarded as negative here) and s1 Gp1, so that
p1A( 1)GS
p1C 1 GS

(10.48)

Similarly, at the plane of exit


p1C 2 GS

(10.49)

428

Rolling Contacts

or
p1 GSA 1
and
p2 GSA 2
This means that the effect of tension is to reduce the pressure that must
be exerted on the material by the roll in order to effect reduction in
thickness.
The equation derived by von Karman gave a starting point to many
theories of rolling which are very easily modied to take account of the
effects of front and back tension. This can be seen by referring to equation (10.45), which is an expression for the vertical pressure between
the roll and the strip from the plane of entry to the neutral plane. If a
back tension 1 is applied, then the only modication is that p1 now
equals SA 1 instead of S. Similarly, in the equation for the region of
forward slip p2 GSA 2.

10.5.7 Shortfalls of rolling theories


The theories of rolling discussed above are all based on the common
assumptions that a plane vertical section of the rolled material remains
plane during rolling, and that the rolls slip on the material except at the
neutral plane. It has also been assumed that the yield stress of the material
remains unchanged during rolling, and that the coefcient of friction is
also constant. In most cases of cold rolling the constrained yield stress,
S, increases on account of work hardening and in hot rolling probably
undergoes a change as the material passes between the rolls owing to the
variation in the compression rate from point to point along the arc of
contact. Also, there is a strong probability that is not constant at all
points in the arc of contact, especially during dry rolling.
A theory proposed by Orowan (21) enables the pressure distribution
between rolls and the material to be obtained by a graphical method of
integration of the friction hill equation for the case of variable and
variable S. It goes further and abandons the assumptions of plane
sections remaining plane and of slipping friction.
Practical observations point to the fact that plane sections rarely
remain plane and usually they become curved in a direction opposed to
that of rolling. Although there is no direct experimental evidence that
a similar phenomenon occurs in the rolling of thin sheet and strip
material, there is a strong probability that it does, and the amount of
bending increases as the coefcient of friction increases.

Rolling in Metal Forming

429

In the theories discussed so far, the rolls were assumed always to slip
on the material so that there was a neutral point and not a neutral
zone. This is the case regardless of the magnitude of the product of the
coefcient of friction and the radial roll pressure pr. According to
Nadai (13), this value could not exceed the yield stress of the material
in shear. He considered that, under the high pressures encountered in
the contact arc, ordinary friction rules could not apply, and suggested
the possibility that under these conditions the frictional drag may be
proportional to the relative velocity of slip instead of being equal to
pr. Orowan, on the other hand, assumed the ordinary friction rule to
hold and that, when pr equals or exceeds the yield stress in shear of
the material, the material moves with the roll, or sticks.

10.5.8 Theory of rolling by Orowan


The starting point of Orowans considerations was a study of the compression of a plastic mass between two rough parallel at plates by
Prandtl. The plates were assumed to be of innite length and perpendicular to the plane of drawing shown in Fig. 10.12. It is assumed that
the material of the mass undergoes no work hardening, has a sharp
yield point, and adheres to the surface of the plates, i.e. does not slip
over them. Then, if q is the vertical and t the horizontal pressure, is
the shear stress in the vertical or horizontal plane, h is the distance
between the plates, and S is the constrained yield stress of the material

Fig. 10.12

430

Rolling Contacts

in compression, Prandtl states that the stress distribution is given by


the following equations
S
qGCC x
h

(10.50)

S
tGCC xAS1(1A4y2h2 )
h

(10.51)

S
G y
h

(10.52)

where C depends on the boundary conditions. If qG0 at the edge of


the plate where xGJL, then CGJ(Sh)L. The resulting stress distribution is then as shown in Fig. 10.12.
Since the mouth of the rolls is neither parallel nor plane, an extension
of the above case in which the mass is compressed between plane rough
plates inclined at a small angle 2 (see Fig. 10.13) was elaborated. For
ow towards the apex, the following expressions connecting the stresses
q, t, and , shown in Fig. 10.13, were derived

tGqAS

GA

S
2

2
1A 2

(10.53)

(10.54)

which are analogous to those for parallel plates.


Two points are to be noted about this solution:

Fig. 10.13

Rolling in Metal Forming

431

(a) the solution applies only to ow towards the apex, no mathematical


solution having so far been found possible for ow from the apex.
Orowan, however, assumed that the solution holds for ow away
from the apex;
(b) the solution applies only to small angles of inclination of the plates,
and this limits the application of the resulting theory of rolling to
small angles of contact.
Equations (10.50) to (10.54) inclusive refer specically to the case in
which the material does not slip over the plates, i.e. to sticking. In order
to include the case in which the material slips, it is assumed that the
stress distribution given by these equations applies if a new value of
thickness h is so chosen that the shear stress at the surface equals q
instead of S2 as in the case of sticking. For parallel plates this new
value of h, which is denoted by h*, is
h*G

S
h
2q

(10.55)

For inclined plates the same procedure is adopted by calculating the


semi-angle *, greater than , that a ctitious wedge-shaped plastic
body sticking to the compression plates must have in order for the shear
stress at the surface of a wedge of semi-angle to be equal to qF
(S2). The semi-angle * is determined from
S
*G

(10.56)
2q
Substituting * for in equation (10.54)

q
GA

(10.57)

and from equation (10.53)


tGqAS

2q
S

1A

(10.58)

Hence, equations (10.57) and (10.58) apply to the case of slipping, but
if q is made equal to S2 they reduce to the corresponding equations
for sticking, i.e. equations (10.53) and (10.54).
The above equations (10.57) and (10.58) express the relationship
between the normal pressure q, the radial pressure t, and the shear stress
at any point in the material between inclined plates for the conditions

432

Rolling Contacts

specied. A generalized equation for the equilibrium of a segment of


the rolled material could be derived from the above equations without
assuming homogeneous compression or slipping friction.
Consider a thin vertical section of rolled material of arbitrary shape
bounded by surfaces A and A (see Fig. 10.14) generated by straight
lines moving always parallel to the roll axis and having their intersection
with the rolls in a vertical plane. Let f ( ) be the resulting horizontal
force across surface A per unit width of the material, which intersects
the roll surfaces at an angle to the line of roll centres. Let A intersect
the roll surfaces at an angle Cd so that the horizontal force across A
upon the segment enclosed by them is then (dfd ) d . The horizontal
component of the normal pressure q acting on the ends of the segment
is 2q sin R d ; the frictional drag, denoted by (without assuming
sticking or slipping) is J2 cos R d , where the plus sign refers to the
exit side of the neutral plane.
The sum of these forces must vanish in equilibrium, and hence
df
d

GDq sin JD cos

(10.59)

Fig. 10.14

Rolling in Metal Forming

433

This equation expresses the variation in the horizontal force f ( ) along


the arc of contact in terms of the normal pressure q ( ). The information about the stress distribution in the plastic wedge given by equations (10.53), (10.54) and (10.57), (10.58) makes it possible to obtain
from equation (10.59) a differential equation containing only one
unknown function which is chosen for convenience to be f ( ).
Since the force f ( ) is the same for all surfaces joining points A and
B, it is possible to choose the surface so as to make the calculation most
convenient. Cylindrical surfaces are therefore chosen as shown in Fig.
10.15, where AO and BO are tangents to the roll surfaces at A and B
so that the angle AOBG2 . Let OAGOBGr. It is then assumed that
the distribution of normal and shear stress acting across surface AB is
the same as for a plastic body compressed between two planes tangential to the rolls at A and B. Hence, equations (10.57) and (10.58) are
used to describe the stress distribution over AB.
Considering the unit width perpendicular to the plane of the drawing
shown in Fig. 10.15, the area of an element of surface AB subtending
an angle d at O is
h
dAG
d
(10.60)
2 sin

Fig. 10.15

434

Rolling Contacts

Now, f ( ) is the horizontal component of the force that acts across


surface AB upon the shaded portion of the material. The element df
of f that acts across dA consists of (a) the contribution of the radial
pressure t and (b) the contribution of the shear stress .
Considering rstly contribution (a), this is
t cos dAGt cos

h
2 sin

(10.61)

Substituting for t from equation (10.58) and integrating between G


and G+ , the contribution of t to f ( ) is
ft ( )GhqA

1 S cos d

hS

1A

sin

2q

(10.62)

Denoting
aG

2q
S

w( , a)G

1
sin

1Aa2

2
2

cos d

Slipping or sticking occurs according to whether aF1 or aH1. If w is


plotted against a, then it will be seen that for values of from 0 to 30,
which covers the range of practical rolling, w is virtually independent
of . Hence, equation (10.62) can be written as
ft ( )GhqAhSw

(10.63)

Considering next the contribution (b) due to the shear stress , this is

sin dAG

q
h
sin
d

2 sin

The sign of the expression is positive for the exit side and negative for
the entry side. Hence, integrating by parts gives
ft ( )GJhq

Atan
1

(10.64)

and the total horizontal force on section AB is then

f ( )Gft ( )Cf ( )Gh q 1J

Atan ASw
1

(10.65)

Rolling in Metal Forming

435

For most cases of slipping, f ( ) can be neglected. In cold rolling, if


F0.2, is not greater than 10, and in this case hq (1 A1 tan )
is only about 1 percent of hq. Hence, for cold rolling the following
approximate expression can be used
f ( )Gh(qASw)

(10.66)

f
qG CSw
h

(10.67)

or

Assuming that Gq and substituting the value of q given by equation


(10.67) in equation (10.59), the friction hill equation for slipping will be
obtained
df
d

Gf

(sin J cos )CDSw(sin J cos )

(10.68)

In the case of sticking, q is to be replaced with S2 so that aG1 and,


as approaches zero, w has the constant value of 4. For sticking,
equation (10.65) becomes

1 1
1
f ( )Gh qAS wJ
A
2 tan

(10.69)

or

f
1 1
1
qG CS wJ
A
h
2 tan

(10.70)

and the differential friction hill equation for sticking is then


df
d

Gf

D
h

1 1
1
1
sin CDS wJ
A
sin J cos
2 tan
2

If

sin C2 cos

sin A2 cos

1 1
1
A
mC( )G w( , 1)A
2 tan

and
1 1
1
m( )G w( , 1)C
A
2 tan

(10.71)

436

Rolling Contacts

equation (10.71) can be written as


df
d

Gf

sin CDSm

(10.72)

where m stands for m+( ) on the exit side and m( ) on the entry side.
Equation (10.72) is the friction hill equation for sticking and has the
form
df
d

CA( )f ( )CB( )G0

(10.73)

The solution of the above equation is shown to be


f ( )Gz( )

B( )

z( )

d Cf

(10.74)

where
z( )Gexp

A( ) d

(10.75)

In the case of equation (10.72), A ( )G(Dh) sin and


B( )GDSm

(10.76)
C

On the exit side, B GDSm and on the entry side B GDSm .


If h2 is the thickness of the material leaving the rolls and is the
thickness ratio h2D

ln z( )G

A( ) d G

sin

C1Acos

where
z( )G

C1Acos
h( )
G
C1 Acos 1 h( x )

(10.77)

On the exit side x G 2 G0 and


zC( )G

h
h2

(10.78)

On the entry side x G 1 and


z( )G

h
h1

(10.79)

Rolling in Metal Forming

437

Hence
f ( )Gz( )

DSm

z( )

d Cf

(10.80)

where f is the front or the back tension in the strip.


Function f ( ) can be calculated for various points on the arc of
contact from this equation by graphical integration. The normal roll
pressure is then found for each point in the contact arc from
f
qGpr G C S
h 4

(10.81)

Since w is virtually independent of , equation (10.68) may also be


considered in the form of equation (10.73), i.e.
df
d

CA( ) f ( )CB( )G0

where
A( )G

D
h

(sin J cos )

and
B( )GDSw(sin J cos )
Now, from equation (10.75)
z( )Gexp

D
cos d
x h

D
sin dJ
x h

The rst integral is given by equations (10.77) to (10.79); for the second
integral the following holds, since hG h2CD(1Acos ) and Gh2D

cos

C1Acos

d G (HAH2 )

(10.82)

with
HG

2(1C )
arctan
1[ (2C )]

2C

tan A

(10.83)

10.5.9 Variations of the Orowan equation


If the value of z given by equation (10.79) is substituted in equation
(10.80), and the limits having reference to the entry side of the arc of

438

Rolling Contacts

contact are used, then


fG

DSm
hh1

from which, on the entry side


fhGD

Sm

(10.84)

where mG0.5A0.785 sin .


From equation (10.78), for the exit side zGhh2, and hence

Sm
d
fhGD
(10.85)
h
0

where mG0.5C0.785 sin .


Using appropriate limits for the entry side and substituting in equation (10.86) the value of z given in equation (10.85) gives
fG

h (H 1AH )
e
h1

DSw( cos Asin )


h e (H 1AH )h1

Hence, for the entry side


fhGD e (H 1AH )

Sw( cos Asin )


h e (H 1AH )

(10.86)

Similarly, using the value of z given in equation (10.84), for the exit
side
fhGD e H

Sw( cos Csin )


h e H

(10.87)

10.6 Discussion of metal rolling theories


The theory proposed by Orowan differs from other theories advanced,
for instance, by von Karman in that it does not postulate slipping friction and homogeneous compression. It involves certain assumptions
which may be summarized as follows:
(a) The material does not spread during rolling.
(b) The angle of contact is small. This follows since it is based on the
stress distribution in a plastic mass between plane plates inclined at
a small angle to each other.

Rolling in Metal Forming

439

(c) In deriving the equations for the pressure distribution, it is assumed


that Nadais solution for the stress distribution between inclined
plates when the ow is towards the apex also holds for ow away
from the apex, and that the case for slipping can be derived from
that of sticking by assuming imaginary planes of sticking outside
the boundary plates.
In the rolling of thin wide strip or sheet, the rst two assumptions are
virtually true, but those included under (c) can be justied or refuted
only by a direct comparison of the calculated pressure distribution with
the measured pressure distribution.

10.7 References
(1) Ekelund, S. (1933) The analysis of factors inuencing rolling pressure and power consumption in the hot rolling of steel. Steel, 93.
(2) Dahl, T. (1935) A study of the phenomenon occurring in the surface
of the material in contact with the rolls, with special reference to the
neutral point and forward slip. Arch. Eisenhuttenwes., July, 1521.
(3) Dresden, D. (1925) On forward slip in rolling. Z. Math. und Mech.,
5, 7879.
(4) Caswell, S. J. (1930) The rolling of metals. The Engineer, 149.
(5) Dahl, T. (1937) The determination and magnitude of the coefcient
of friction during rolling. Stahl und Eisen, 57, 205209.
(6) Tafel, W. (1921) Stahl und Eisen, 41, 962963.
(7) Tafel, W. and Schneider, E. (1924) Stahl und Eisen, 44, 305309.
(8) Lueng, W. (1935) The inuence of the rolls on the cold rolling of
strip steel. Stahl und Eisen, 55, 1105.
(9) Trinks, W. (1937) Coefcient of friction and ow in strip rolling.
Blast Furn. and Steel Plant, July, 713715.
(10) Siebel, E. and Pomp, A. (1929) Roll pressure and work on the cold
rolling of metals. Mitt. Inst. Eisenforsch., 11, 7386.
(11) Siebel, E. and Osenberg, E. (1934) The inuence of friction and of
cross-sectional dimensions on the ow of material during rolling.
Mitt. Inst. Eisenforsch., 16, 3350.
(12) Hitchcock, J. H. (1931) Analysis of rolling forces. Rolling Mill J.,
5, 583589.
(13) Nadai, A. (1939) The forces required for rolling steel strip under
tension. Trans. ASME, J. Appl. Mech., 5462.
(14) Siebel, E. and Lueng, W. (1933) Investigations into the distribution
of pressure at the surface of the material in contact with the rolls.
Mitt. Inst. Eisenforsch., 15, 114.

440

Rolling Contacts

(15) Brown, G. M. (191718) Rolling mills and their electrical equipment. Proc. Cleveland Inst. Engrs.
(16) von Karman, T. (1925) On the theory of rolling. Z. Math. und
Mech., 5, 139141.
(17) Siebel, E. (1932) The Plastic Forming of Metals (Stahleisen,
Dusseldorf).
(18) Siebel, E. (1922) Fundamentals necessary to the calculation of
energy and work requirements during forging and rolling. Ber.
Walzw. Aussch. Eisenh., (28).
(19) Keller, J. D. (1937) How thin can strip be rolled. Blast Furn. and
Steel Plant, 25, 1105.
(20) Tselikov, A. I. (1939) Effect of external friction and tension on the
pressure of the metal on the rolls in rolling. Metallurg, (6), 6176.
(21) Orowan, E. (1943) The calculation of roll pressure in hot and cold
at rolling. Proc. Instn Mech. Engrs, 150, 140167.

INDEX
Addendum 265, 267
Adhesion zone 205
Adhesive junctions 57
Adhesive strength 378, 381
Angle of pressure 241, 245
Angular contact ball bearings 354, 358
Antiskid system 72
Apparent area of contact 205
Apparent contact area 52
Arc of contact 412416, 418420, 423,
424, 426, 428, 437
Area of contact 29
apparent 52, 205
nominal 57
real 31, 34
Asperity
heights 35
peak 31, 32
plastic contact 274
Axial deflection 159
Axial load 114, 194
Axial preload 101, 114
Axial rigidity 114, 115
Ball bearing 2, 183, 192, 193
Ball waviness 357, 358
Bearings
ball 2, 183, 192, 193
angular contact 354, 358
ceramic 352355
hybrid 352, 354, 355
crankpin 173
radial 101, 180
rolling 2, 3
thrust 110, 111, 114, 167, 180, 184,
194, 195
Bearing area curve 33
Boundary lubrication 261
Cage 123, 169, 173, 198
Cageless 173

Cam 285286, 288, 289, 296, 297, 300,


302, 303, 305, 306, 307, 309, 310,
311, 312
Camfollower system 285314
Camshaft 297
Centrifugal force 171, 172
Ceramic bearings 352355
hybrid 352, 354, 355
Ceramic materials 350
Ceramic rollers 384
Ceramics 351
Cermet coatings 367, 368
Chemical plating 375, 376
Chemical vapour deposition (CVD) 356,
379383
photo-assisted 382
plasma 382
thermal 381382
Clearance 107, 109
radial 110
Coating 365
cermet 367, 368
processes 365383
Coefficient of cohesion 207
Coefficient of friction 3
Coefficient of rolling friction 64, 76, 78,
339
Coefficient of traction 212
Cold rolling 405, 417, 435
Complex shear modulus 72
Concave curvature 27
Concave flanks 289
Concave surface 162
Conforming contacts 11
Constant pressure 19
Contact
angle 425
area 73
apparent 52, 205
conforming 11

442

Rolling Contacts

ellipse 83, 155


Hertzian 19, 4142, 44
line 11, 156157, 159, 162, 164
non-conforming 11
non-Hertzian 50
point 11, 157, 158, 159, 161, 164
pressure 15, 20
region 30
rolling 2, 55, 318, 319, 329, 330, 331,
365
surface 12
tyreroad 224, 227, 228
zone 24, 35, 39, 53
Convex flanks 289
Convex surface 162
Corrosion fatigue 280
Crankpin bearing 173
Crystalline polymer 322
Cubic mean load 186187
Cumulative damage 342
Dedendum 265, 275, 281
Deflection, axial 159
Deformation
permanent 160163, 182
plastic 17, 18, 22, 31, 3438
Delamination
fatigue 361
wear 388389
Density 39, 264
Destructive pitting 278
Detonation
gun 368, 369
spray 366, 368
Dry road 217, 226
Dynamic hydroplaning 223, 226227
Elastic energy 332334
Elastic hysteresis 3, 5, 57, 59, 80, 82,
84, 146
loss 64
Elastic modulus 57
Elastic strain energy 53
Elastohydrodynamics 7, 8, 195, 196, 197,
222, 251, 252261, 262, 269, 270,
313
isothermal 255
Electroplating 356, 372377, 384

Endurance limit 317


Energy losses 84
Energy of adhesion 332
Engineering ceramics 350351, 353
Enthalpy 334
Entropy 332
Equivalent load 177, 181, 184, 192, 193
Equivalent radial load 194, 195
Equivalent thrust load 194, 195
Fatigue 175, 178, 182, 188, 195, 317, 318,
320, 324, 325, 327, 328, 329, 340,
341
corrosion 280
delamination 361
fracture 279
life 4, 8, 319, 342
resistance 358
rolling contact 344
spalling 362
strength 330, 341, 343
surface 275
Flash temperature 252, 259, 264
Fluid-film lubrication 261
Follower 285286, 288, 289, 290, 291,
295, 297, 299, 300, 302, 305, 306,
307, 309
Forces, surface 57
Fractional film defect 272
Fracture, impact 280
Free rolling 62
Fretting 24
Friction
angle 409, 420
coefficient 7, 15, 74
rolling 4, 55, 57, 5862, 7679
coefficient of 64, 65, 76, 78
static 3, 5
torque 88106, 143, 151
Frictional loss 87
Gears
helical 247, 254255, 265, 267, 284
involute 241, 243
spur 239, 247, 254255, 257, 264, 267,
283
teeth 239243, 248
Gearset 246, 260

Index
Gibbs free energy 334
Glass temperature 321322
Glassy polymers 322
Gyratory moment 174, 175
Hardness 23
Hearth rollers 383, 385
Heat flow 39
Heathcote slip 58
Helical gears 247, 254255, 265, 267, 284
Helix angle 247, 254, 284, 285
Helmholtz free energy 334
Hertzian contact 19, 4142, 44
High-temperature isostatic pressing 384
High-velocity oxyfuel 366
Hot isostatic pressing 356
Hot rolling 391, 392, 405407, 417, 418
Hybrid ceramic bearings 352
Hydrodynamic lubrication 6, 246
Hysteresis
effect 4
loss 59, 65, 66, 79
Impact fracture 280
Inner race 6, 131, 132, 134, 135, 140, 142
Inner ring 100, 117, 123, 165, 168171,
189, 191, 193
Interfacial energy 332
Internal friction 59, 73
Involute curve 240
Involute gears 241, 243
Ion plating 40
process 378379
Isochromatics 17, 18, 22
Isothermal 253
elastohydrodynamics 255
Junction, adhesive 57
Line contact 11, 156157, 159, 162, 164,
270
Load
axial 114, 194
capacity 177, 179
cubic mean 186187
distribution 163
equivalent 177, 181, 184, 192, 193
radial 101, 107, 108, 157, 163, 192, 195
equivalent 194, 195

443

static
equivalent 195
thrust 157, 166, 175, 190, 192, 193
equivalent 194, 195
Load-carrying capacity 178, 180, 189,
191, 192, 193, 194
Lubricant film 6, 87
Lubricating film 95
Lubrication
boundary 261
fluid-film 261
hydrodynamic 246
micro-elastohydrodynamic 225, 226,
227
Maximum shear stress 239240, 241
Mean contact pressure 23
Mechanical equivalent of heat 272
Micro-EHL (elastohydrodynamic
lubrication) 225, 226, 227
Microsliding 149152
Microslip 5, 23, 24, 57, 77
Modulus of elasticity 21
Mohrs circle 12, 17
Moving heat source 39
Neutral plane 422, 424, 426, 427, 428,
432
Neutral point 394, 396, 399, 400, 404,
409, 413, 414, 429
Nominal area of contact 57
Non-conforming contacts 11
Non-Hertzian contact 50
Normal approach 29, 30, 33, 34
Normal load 24
Normal pressure angle 247, 254
No-slip angle 397399, 400, 402, 404,
408
No-slip point 394, 400, 413
Oil inlet temperature 259
Outer race 6, 129, 130, 134, 135, 140, 141
Outer ring 117, 165, 168171, 189
Papermaking rollers 385386
Partial EHL (elastohydrodynamic
lubrication) 227228, 230, 231,
233
Permanent deformation 160163, 182
Physical vapour deposition (PVD) 356,
377379

444

Rolling Contacts

Pinion 262
Pitch 267
circle 247, 253
line 245, 247, 254, 255, 257, 267, 281
point 244, 256
Pitting 274279, 319
destructive 278
progressive 275, 276
Plane strain 35
Plasma spray 366367, 370372
process 370
vacuum 371
Plasma torch 370371
Plasma transferred arc 384
Plastic deformation 17, 18, 22, 31, 3438
Plastic zone 2223
Plasticity index 35, 271, 274
Pneumatic tyre 201, 210
Point contact 11, 157, 158, 159, 161, 164,
270
Polymer 321
crystalline 322
glassy 322
rolling contact 348350
semi-crystalline 322323
Polymeric materials 320, 337
Power loss 105, 245
Preload 107, 109, 125
axial 101, 114
radial 110
Pressure angle 244, 247, 253, 254, 283,
285
normal 247, 254, 284
Pressure coefficient of viscosity 251
Pressureviscosity characteristics 8
Pressureviscosity coefficient 269, 313
Principal radii 26
Printing rollers 385386
Probability density function 33
Probability of scuffing 271, 274
Process rollers 383
Progressive pitting 275, 276
Raceway 86, 117, 123, 159
Radial bearing 101, 180
Radial clearance 110
Radial contact bearing 163
Radial displacement 158

Radial load 101, 107, 108, 157, 163, 192,


195
Radial preload 110, 115
Radial stiffness 106, 111, 120, 121
Rail 201, 207
Railwheel traction 202205
Real area of contact 31, 34
Relative radius of curvature 244245
Residual stress 372, 389
Resisting movement 63
Ribbed tyre, traction of 208
Road
dry 217, 226
wet 222, 225, 226, 229, 231
Rollers
hearth 383, 385
papermaking 385386
printing 385386
process 383
sink 384, 385
Rolling
cold 405, 417, 435
hot 391, 392, 405407, 417, 418
Rolling bearing 2, 3
Rolling contact 2, 55, 318, 319, 329, 330,
331, 365
fatigue 344
polymer 348350
Rolling friction 4, 55, 57, 5862, 7679,
145
coefficient of 64, 65, 76, 78
Rolling motion 2, 89
Rolling process 391, 393
Rolling resistance 2, 4, 5, 57, 58, 59, 60,
61, 69, 145, 147, 150, 209, 210
Rolls 391, 392, 393, 398, 400, 404416,
418, 419, 421, 422, 427, 428, 430
Running-in 260, 262, 263, 274
Scuffing 263274, 279, 313
probability of 271
resistance 268
temperature 265
Semi-crystalline polymers 322323
Service life 175178, 180181, 184, 185,
191, 193
average 177
Shear modulus 50

Index
Shear stress 17, 21
Silicon nitride 353, 357, 358, 359
balls 355, 356, 359, 360, 362
Sink rollers 384, 385
Slip 5, 6, 23, 24, 50, 51, 58, 6872, 125,
132, 135, 197, 198, 201, 205, 214,
215, 216, 217, 218, 221, 223, 236,
347, 389, 400404, 405, 408, 428,
429, 431
Slip band 36, 37
Spalling 276, 278, 359
fatigue 362
Specific bearing capacity 184
Specific capacity 177, 178, 180, 181
Specific damping capacity 6667
Specific film thickness 270
Specific heat 39
Spin 80, 131, 132, 135, 198, 214, 221, 346
Spray gun 366
Spur gears 239, 247, 254255, 257, 264,
267, 283, 318
Sputtering 40
process 379
Static capacity 182, 183
specific 182184
Static equivalent load 195
Static friction 3, 5
Stick 24
Stiffness, radial 106, 111, 120, 121
Strain 338
energy 66, 68
Stress
maximum shear 239240, 241
residual 389
shear 17, 21
tangential 68, 70
yield 17, 68, 240
Surface
concave 162
contact 12
convex 162
energy 52, 53, 54
fatigue 275
forces 57
Tangential load 23
Tangential stress 68, 70
Tappet 286, 297, 309313
Teeth, gear 239243, 248

445

Temperature
flash 252, 259, 264
oil inlet 259
scuffing 265
Thermal conductivity 39, 264, 272
Thermal spray 356, 366, 372, 386, 389
Thrust bearing 110, 111, 114, 167, 180,
184, 194, 195
Thrust load 157, 166, 175, 190, 192, 193
Tooth flank 243
Tooth loading 246
Tooth losses 247
Torch 367
plasma 370, 371
Torque 215
friction 143, 151
Total friction torque 80
Traction 201, 204, 210236, 245, 275,
323
characteristics 215
coefficient 203, 204, 212
of a tyre on a dry road 217
railwheel 202205
ribbed tyre 208
Transverse pressure angle 255
Tribology 1
Tyre 210236
pneumatic 201, 210
Tyreroad contact 224, 227, 228
Tyreroad traction 210, 214
Vacuum plasma spray 371
Viscoelastic 338, 339
flow 327
materials 324, 331, 337
solids 330
Viscoplaning 224227
Viscosity 246
Viscosity grade 259
Viscositytemperature dependence 257
Wear, delamination 388389
Wet road 222, 225, 226, 229, 231
Yield pressure 205
Yield strength 18, 23
Yield stress 17, 68, 240, 396

Вам также может понравиться