1991 Book Tribology PDF
1991 Book Tribology PDF
1991 Book Tribology PDF
Design Applicatio ns
Tribology
Principles and Design Applications
R. D. Arnell
P. B. Davies
J. Halling
T. L. Whomes
M
MACMILLAN
© R. D. Arnell, P. B. Davies, J. Halling and T. L. Whomes 1991
Published by
MACMILLAN EDUCATION LTD
Houndmills, Basingstoke, Hampshire RG21 2XS
and London
Companies and representatives
throughout the world
Preface vi
Notation viii
2. Friction 27
3. Wear 66
References 245
Index 249
v
Preface
In 1975 we and four of our former colleagues wrote the textbook Principles
of Tribology. That book was very successful both in the United Kingdom
and overseas and it is still being sold today. However, 13 years after the first
publication of the book, the publishers suggested that we should consider
revising the text, to include the more important developments which had
taken place in the meantime.
Although we have remained active in tribological research, it was only
when we started work on this revision that we realised that the developments
which have taken place have been so extensive that they could not be
accommodated simply by revision of our existing text. As a result, although
we have reused some of our original material, the present book has resulted
from a complete reappraisal of our ideas on the necessary content of a modern
introductory textbook on tribology. As the title of the book implies, it has
been our intention to write a book which will provide a thorough grounding
in the principles of our subject for undergraduates and new research workers
and will also act as a basic reference book for practising designers.
In the areas of friction and wear, although there has been a great deal of
high-quality research on the underlying physical and mechanical processes,
there are no current, or indeed foreseeable, models which will allow
quantitative predictions of friction coefficients or wear rates from basic
physical, chemical and mechanical properties. Such predictions will continue
to be based on experimental measurements. Therefore, in these areas, we
have attempted to set out as clearly as possible the underlying processes and
thus to assist the reader in making sensible initial choices of materials.
In the areas of fluid film lubrication, as well as covering the fundamental
mechanisms, we present simple design procedures for hydrodynamic bearings
and incorporate the more important aspects of the growing amount of
information on elastohydrodynamic lubrication, with examples of its use in
design.
vi
PREFACE vii
May, 1990 R. D. A.
P. B. D.
J. H.
T.L.W.
Notation
Symbol Definition
A Real area of contact
AF Area of filter
a Semiaxis of contact ellipse in transverse direction, or radius of
contact circle (Chapters 1, 7)
Width of axial supply groove (Chapter 6)
Width of bearing
Length of journal bearing (Chapter 6)
Semiaxis of contact ellipse in direction of motion (Chapter 7)
Dynamic load capacity of rolling bearing (Chapter 8)
Static capacity of rolling bearing
k,/k
Specific heat
Diametral clearance
Radial clearance
Diameter of gear wheel
Diameter of journal bearing
Mean diameter of rolling bearing
E' .
Elastic modulus
Elastic modulus of contact= 2 I( -
El
vi + -
1- - v~ )
1- -
E2
1-v 2 1-v2)
E* Elastic contact modulus = 1 /( - -1 + - -2
Et E2
Shaft eccentricity
Shear or friction force
Axial force
Rolling friction
viii
NOTATION ix
=ar
-ars/3
0112 =(Jl
ar rectangular elliptical
=on contact =(~Yn contact
-arJ/2 G-ar 3
= 0112 G = ----rJ2
H Hardness (Chapters 2, 3)
H Power to rotate shaft (Chapter 6)
H Dimensionless power coefficient
n Dimensionless film thickness ( = hI R)
h Dimensionless surface separation (Chapter 1)
h Film thickness (Chapters 6, 7, 8)
he Central film thickness
hm Film thickness at maximum pressure
hmax Maximum film thickness in a journal bearing
hmin Minimum film thickness in a journal bearing
h0 Minimum film thickness
hP Film thickness at pitch line
h. Starved film thickness
h1 Outlet- film thickness of slider
h2 Inlet film thickness of slider
h 00 Film thickness with fully flooded inlet
I Tou
Inlet shear heating parameter = ( 11
2
)
Subscripts
1, 2 Refer to the two surfaces of the contact
Chapter 1
Surface Properties and
Surface Contact
1.1 INTRODUCTION
Primarily because of their high strength and toughness, metals have been
the most common tribological materials for many years. For this reason,
their tribological behaviour has been studied in greater detail, and is
consequently more fully understood, than that of other materials. Therefore,
we shall start by concentrating on the nature of the metal surface, but in
later chapters we shall extend all our discussions to include more recent
tribological materials such as ceramics, elastomers and polymers.
However, apart from this particular case, other surfaces which we describe
as clean will still retain the oxide films which are of paramount importance
in determining tribological behaviour.
There are now many techniques available for studying the chemical
nature of the surface and near-surface layers of materials. However, for the
purposes of this text it is only necessary to make use of the relevant results
of such studies: it is unnecessary to describe the actual techniques, and readers
interested in these should consult specialised texts such as Walls ( 1989) and
Woodruff and Delchar (1988).
Stylus
(a)
Profile trace
(b)
Figure 1.2 (a) Schematic view of profilometer; (b) typical surface profile
SURFACE PROPERTIES AND SURFACE CONTACT 5
Stylus radi~ I
x200 ~ X 200
Figure 1.3 Illustration of the profile distortion due to unequal vertical and
horizontal magnifications
6 TRIBOLOGY
Until recent years, the analogue output of the profilometer was usually
analysed within the instrument to give a graphical record of the profile and
a single-parameter description of the surface roughness, which was displayed
on an analogue meter. The two parameters in most common use were: ( 1) the
roughness average (RA) value, defined as the mean vertical deviation of the
profile from the centre line, treating deviations both above and below the
centre line as positive; (2) the root mean square (RMS) value, defined as the
square root of the mean of the square of these deviations. The RA has
traditionally been the most common parameter in use in the UK, while the
RMS value has been the most common in the USA.
These parameters are seen to be concerned only with relative departures
from the centre line in the vertical direction; they do not provide any
information about the shapes, slopes and sizes of the asperities or about the
frequencies of their occurrence. It is, therefore, possible for surfaces of widely
different profiles to give the same RA and RMS values, as shown in Figure 1.4.
These single-parameter descriptions are, therefore, mainly useful for
comparing surfaces which have been produced to different standards but by
similar methods: for example, lapped surfaces, differing only in the grade of
lapping compound with which they have been finished.
Surface profiles often reveal both periodic and random components in their
geometric variation, and such components are not revealed by the RA and
RMS values. In recent years it has become common to digitise the surface
profile, to give a record which is more amenable to statistical analysis and
to separation of its periodic and random elements. In this technique the
analogue output from the amplifier is digitised at discrete intervals, as shown
in Figure 1.5, to give a sequence of height readings relative to the profile
centre line. A typical record would consist of 4000 digital height readings
SURFACE PROPERTIES AND SURFACE CONTACT 7
taken at intervals of 2 Jim along the surface. The sequence can then be
analysed by an inbuilt microcomputer, to extract any chosen statistical
information. The two most common types of information to be extracted are
the height distribution of the surface texture and the amplitudes and
wavelengths of periodic variations along the surface; we shall consider these
in turn.
distribution curve
Periodic height variations along the surface can be displayed by plotting the
variation of the autocorrelation function R{l) of the statistical data against
the sampling interval l. R(l) is obtained by delaying the profile relative to
itself by some fixed interval I, multiplying the delayed profile by the original
one, and averaging the resulting product value over a representative length
of the profile. Thus,
1 N-1
R(l)=- L z(x)·z(x+l) ( 1.3)
N-Ix= 1
where l is the distance between the original and the delayed profiles and
N is the total number of ordinates in the total sample length L.
SURFACE PROPERTIES AND SURFACE CONTACT 9
~:Jww, F-l.~~
~------- ~---------
Surface A Surface B
"'~"'L,
Figure 1. 7 Typical surfaces and the resulting autocorrelation functions
It can then be seen that when l = 0, R(l) reduces to the square of the u value
of the profile. The autocorrelation function is therefore usually plotted in its
standardised form r( /), where
and the r(l) value always has its maximum value of unity when l = 0.
Typical plots of autocorrelation functions for two different surfaces are
shown in Figure 1.7. The shapes of these plots reveal both the random and
the periodic characteristics of the profiles. The general decay indicates a
decrease in correlation as l increases, and is an indication of the random
component, while the oscillatory component of the function reflects a similar
periodicity in the profile. Figure 1.8 shows results from some machined
surfaces, where the plots can be seen to demonstrate the general features of
the profiles.
It can thus be seen that it is possible to describe the general features of
any surface profile by two characteristics: the height distribution function
cf>(z) and the standardised autocorrelation function r(l).
When we come to discuss the nature of contact between two surfaces,
we shall also see that the heights, curvatures and frequencies of distribution
of the surface asperities are of critical importance. All these data can be
extracted from the digitised profile.
In practice, it is often found that both the all-ordinate distribution and
the peak height distribution for engineering surfaces are good approximations
to Gaussian distribution curves - i.e. curves of the form
(1.6)
10 TRIBOLOGY
Shaped surface,
RA 1.6 ~m
¢ (z) A r(/) ~
~_l JJJ'J/JJV I
'···~
I
5 L
100
¢~
r4.....,_ ... ,A.o_ •
r(I)L
_..~...... AJ
••-"V'Yfv-v.,.........-.- .,
.-.A Ground surface,
RA 1.0 ~m
¢(z) ( \ r(l)~
ll_~ I
Figure 1.8 Examples of engineering surfaces, their distributions and their auto-
correlation functions
Both friction and wear are due to the forces which arise from the contact of
solid bodies in relative motion. Therefore, any understanding of friction or
wear must be based on an understanding of the mechanics of contact of solid
bodies.
Solid bodies subjected to an increasing load deform elastically until the
stress reaches a limiting value, known as the yield stress ay, and at higher
stresses they then deform plastically. In most contact situations some
SURFACE PROPERTIES AND SURFACE CONTACT 11
asperities are deformed elastically, while others are deformed plastically: the
loads will induce a generally elastic deformation of the solid bodies but at
the tips of the asperities, where the actual contact occurs, local plastic
deformation may take place.
In this section we shall start by considering the elastic deformation of
cylinders and spheres. This study is valuable for two reasons: first, many
engineering situations involve the non-conforming contact of bodies defined
by smooth curves; second, all solid bodies have surface asperities which can
be considered as small spherically shaped protuberances, so that the contact
of two macroscopically flat bodies can be reduced to the study of an array
of spherical contacts deforming at their tips.
It is clear that a detailed study of surface contacts would involve a detailed
understanding of the elastic and plastic deformation of contacting surfaces.
However, it is the purpose of this chapter simply to set out the basic knowledge
which we need to understand the material of the following chapters: it is not
necessary fully to develop the theory, and readers requiring more detail are
referred to the definitive book by Johnson (1985).
We shall be concerned with effects within the outermost layers of bodies,
typically within a millimetre of the surface, and can ignore effects at greater
depths. This allows us to treat the surfaces as surfaces of semi-infinite bodies,
ignoring overall shape, and leads to mathematical simplification.
We are particularly interested in problems of contact between bodies
whose geometries are defined by circular arcs. The problems of elastic contact
between such bodies were first solved by Hertz in 1882, and such situations
are referred to as Hertzian contact. We shall initially concentrate on the
stresses in such contacts.
(a) (b)
loaded with the pressure distribution given by Equation (1.7) acting over the
footprint extending from -a to + a.
The maximum shear stress in plane strain conditions is given by
rmax={(ax-O"z) 2 /4+r;z}t (1.9)
where expressions for ax, az and rxz can be found in Johnson (1985). Therefore,
Equation ( 1.9) defines the values of 'max at all points. The evaluation of this
expression allows us to plot the maximum shear stress distribution beneath
the footprint, and the rather surprising result is shown by the isochromatics
of the photoelastic stress pattern of Figure 1.10. The greatest value of rmax
occurs at a distance of0.79a below the surface, and we find that it will attain
the critical value of k when the maximum pressure, Pmax• at the centre of the
footprint is 3.3k. If we had simple compression, we should expect the surface
material to yield when Pmax attained a value of 2k, but this does not happen
in the present case, because the surface elements are subjected to compressive
stresses in all three orthogonal directions, giving a hydrostatic stress
component which cannot contribute to plastic deformation.
This very important result means that higher loads than expected can
be carried elastically in these non-conforming contacts. Furthermore, even
when sub-surface yielding has taken place, very little plastic deformation can
occur, because the plastic zone is constrained by the surrounding elastic
material.
As the load is increased further, the plastic zone will increase in size and
ultimately spread to the surface, so that plastic indentation can occur. This
happens when the mean pressure is approximately 6k - i.e. approximately
twice the pressure at which yield first occurred. The mean pressure at this
point is essentially the indentation hardness value, H , of the material, which
p..,.,
L.nes ?I
._~-i-- constant
r....,.
Figure 1.10 Actual isochromatics obtained for the contact of a cylinder and a
plane due to normal load alone
14 TRIBOLOGY
Fully plastic
behaviour
Plastic behaviour
constrained by
surrounding elastic matetial
Puft-ly ekJsr•c
behovoour
Figure 1.11
is why, for metals, we find that H ~ 6k ~ 3ay, where O"y is the uniaxial tensile
yield strength of the material.
The behaviour described above is summarised in Figure 1.11.
We have, as yet, considered the effects of normal loads only, and to
understand tribological phenomena, we must consider the effects of both
normal and tangential loads. As we shall see in the next chapter, the tangential
force, F, required to cause sliding under a normal load, W, is given by J1. W,
where J1. is a constant known as the coefficient of friction. The stress field due
to a tangential force can be derived by a method similar to that described
above for a normal force, and combining the stress distributions due to the
normal and tangential loads leads to a distribution of rmax of the type shown
in Figure 1.12. 1t can be seen that the location of the point of maximum shear
stress is now much closer to the surface, so that plastic deformation can take
place more readily than in the previous case. Thus, macroscopic plastic
deformation becomes easier in the presence of friction forces. A full solution
of this problem has been produced by Hamilton ( 1983 ).
In many practical situations, where friction is used to prevent slip
between mating components, contacting bodies are subjected to tangential
forces which are less than JLW. It can then be shown that within the contact
area there exists a central area in which no slip occurs, while outside this
area a small amount of slip can take place. This coexistence of zones of
sticking and microslipping is possible because the contacting materials are
deformable and the stress distribution is such as to allow slipping in the outer
regions of the contact zone. The subject of microslip will be covered in more
detail in the next chapter, when we discuss the nature of friction in rolling
contacts.
SURFACE PROPERTIES AND SURFACE CONTACT 15
Lonn of
Qli'ISianl
r,...
Figure 1.12 Actual isochromatics obtained for the contact of a cylinder and a
plane due to combined normal and tangential loads, where T = 0.5 1-1P
(1.10)
a= (3WR/8E*)t (1.11)
Maximum contact
pressure -(P'E')l
Po- - -
nR
_1 (6WE'
Po--
n
---
R 2
2 )t
('Hertz stress')
Approach of
centres ~_2P(1-vf(
u -- - - ln4R
- + -1) 1
n E1 a 2
+ 1- v~ (In
E2
4R 2
a
-~))
2
Although the footprint for dissimilar spheres is not a plane circular area,
Equation ( 1.10) still holds with substantial accuracy, but in this case we have
a=(3WR'/4E*)1 (1.12)
For contact between a sphere and l!plane, R' is simply the radius of the sphere.
For general contact between two curved surfaces, the footprint is an
ellipse. This is the case in many important practical situations, particularly
the contact between the balls and bearing races in rolling contact bearings.
In such situations the equations become more complex; it is beyond the scope
of this text to derive them or to state them in full, and readers requiring
more information should consult either Johnson ( 1985) or Halling ( 1975).
Fortunately, for mildly elliptical footprints, the equations for circular
footprints can be used in a slightly modified form.
In Table 1.1 we summarise the most commonly required formulae for
rectangular footprints, circular footprints and mildly elliptical footprints.
or
W = (4/3)E*RW (1.15)
Combining Equations ( 1.12) and ( 1.15), we see that the area of contact, A,
will be given by
Equation ( 1.16) indicates that the surface outside the footprint is displaced
in such a way that the area of contact is exactly half the area 2nRD which
would be obtained by plastic flattening of the sphere.
Equation ( 1.16) also shows that for spherical contacts on a plane
D= a 2 jR.
As before, plasticity first occurs when the maximum elastic Hertzian
contact pressure reaches 0.6H, and we know that for this geometry
Pmax = 1.5Pmean• so that, at the onset of plasticity, (3/2) W jna 2 = 0.6H.
Substitution for Wand na 2 from Equations (1.15) and (1.16) shows that the
maximum normal approach before the onset of plastic deformation is given by
(1.17)
We shall see later that this result is of major significance in the contact
of rough surfaces.
The results described above apply to perfectly elastic/plastic materials,
while many real materials, particularly metals, exhibit strain hardening.
Although no analytic solution exists for such materials, Halling and Nuri
(1974) have developed a simple treatment which shows the effect of strain
hardening. They consider the deformation behaviour of any material to be
SURFACE PROPERTIES AND SURFACE CONTACT 19
2.5
Film thickness (J..!m)
£'"
1.4 K = __!_
E'"
2
0.05
1.2
1.0
.
0
~
0.8
0.6
0.4
10-3 10-2
tlao
Figure 1.16 Effect of film thickness on the Hertzian contact of coated surfaces
Although all surfaces are rough, we can simplify the study of two rough
surfaces in contact if we consider the situation to be that of a single rough
surface in contact with a perfectly smooth surface. We can further simplify
the problem by modelling the asperities on the rough surface as sectors of
spheres, so that their elastic deformation characteristics can be modelled by
the Hertz theory for the contact between a sphere and a plane. Finally, we
shall also assume that there is no interaction between the separate asperities,
so that vertical displacement due to a load on one asperity does not affect
the heights of the neighbouring asperities.
We consider first the simple, but unrealistic, model illustrated in Figure
1.17, where all the asperities on the plane of unit area have the same height,
z, relative to the reference plane XX'. As the surfaces are loaded together,
we see that the normal approach is (z- d), where dis the current separation
of the smooth surface and the reference plane in the rough surface. Each
asperity is deformed equally, and carries the same load, W;, so that for '1
asperities per unit area the total load W will be '7W;. For each asperity, the
load, W;, and the area of contact, A;, are known from the Hertzian theory.
Thus, if p is the asperity radius, we have
W; = (4/ 3)E*P(z- d)t
and
A;= nP(z- d)
The total load will then be given by
W = (4/ 3)E*{P(A;/np)!
so that the load is related to the total area of contact, A(= '7A;), by
W = (4E/ 3nt'ltp) ·A! ( 1.21 )
This result shows that, for this particular model, the real area of contact is
proportional to the two-thirds power of the load when the deformation is
elastic.
_x t-"F-~~ ·-¥3-
Refer~e
plone on
rrugh surface
Figure 1.17 Contact between a smooth plane and an idealised rough surface
22 TRIBOLOGY
If the loads are such that the asperities are deforming plastically under
the constant flow pressure, H, each individual contact area, A', will be given
by 2npo. The individual load, w;, will then be given by
w; = HA; = 2Hnp(z- d)
Thus,
W' = 11 w; = 11 HA; = HA' = 2HA (1.22)
so that the real area of contact is proportional to the load.
We shall see in Chapter 2 that it is basic to observed frictional behaviour
that the real area of contact of rough surfaces should be proportional to the
load. While our simple model gives this proportionality for plastic deformation,
it does not do so for elastic deformation.
However, we know that our model of the rough surface is unrealistic.
As pointed out earlier in this chapter, on real surfaces asperities have different
heights following some probability distribution, and we must, therefore,
modify our surface model to take this into account.
Therefore, we consider the contact of a rough surface, having a height
distribution cf>(z), with a smooth reference plane, as shown in Figure 1.18. If
the separation of the two surfaces under load is d, then there will be contact
of any asperity with an original height greater than d. The probability that
any asperity has a height between z and (z + dz) above the reference plane
will be cf>(z) dz. Thus, the probability of contact for any asperity of height z
is given by
If we consider unit area of the surface containing '1 asperities, the number of
contacts, n, will be given by
Since the normal approach is (z- d) for any asperity and W; and A; are
known from Equations (1.15) and (1.16), the total area of contact and the
and
We shall see in the following two chapters that both friction and wear are
very strongly influenced by the degree of plastjc interaction of the contacting
24 TRIBOLOGY
the probability of finding an asperity fell to zero at some finite distance, ba,
above the mean line. This is open to some reservations; in particular, the
expected height of the highest asperity, as measured by b, will not have a
constant value but will increase as the surface area increases. However, even
for Gaussian surfaces, about 99% of all the asperities lie within a distance
3a of the centre line and, as we shall see, truncation of the distribution at a
value of 3a - i.e. at a value of b = 3 - gives a very good fit of theory to
experimental observations.
Halling et al. then considered the situation shown in Figure 1.19. It can
be seen that the asperities from d to (d +be) are deforming elastically, while
the asperities from (d +be) to ba are deforming plastically. By integrating
the distribution curve between these two sets of limits, it is then possible to
define the ratio between plastically deforming area and elastically deforming
area for different contact pressures and different values of the plasticity index,
1/J. The results for truncation at 3a are shown in Figure 1.20; where R is the
ratio between plastic area and elastic area, and p is the non-dimensional
nominal pressure (N I AH). For R = 0, we have the result for purely elastic
deformation. It can be seen that this approach allows us to determine the
relative amount of plastic deformation for different values of the plasticity
index and different nominal pressures. As both friction and wear are very
dependent on the degree of plastic deformation, it is likely that this approach
will prove to be very useful in the future in helping us to predict friction and
Figure 1.19 Elastic-plastic contact of a smooth surface and a rough plane with
a peak height truncated at bu
Figure 1.20 Ratio between plastic and elastic contact area as a function of plasticity
index, t/J, and normalised pressure, jj
26 TRIBOLOGY
1.6 PROBLEMS
(In solving these problems, the following properties may be assumed. Young's
modulus( GPa): steel, 200; copper, 124; tungsten carbide, 550. Poisson's ratio:
0.3 in each case.)
1 A surface profile is sinusoidal, with unit amplitude and wavelength A..
The profile is sampled at equal intervals, with the origin on the centre
line at a position of zero amplitude. Calculate the computed RA
values for this profile for sampling intervals of A.j2, A./4, A./8 and A./16.
Calculate also the RA value which would be derived from the analogue
signal for the same profile.
[Computed RA values: zero; 0.5; 0.603; 0.628. Analogue value: 0.637]
2 A hard steel cylinder 100 mm long and 10 mm in diameter is loaded
under an increasing normal force against a flat copper plate having a
hardness of 6000 MPa. At what load does the material of the plate first
yield; what is the corresponding contact width; and what are the mean
and maximum pressures in the contact zone?
[Py = 2200 N; 2a = 0.084 mm; Pmean = 261 N/mm 2 ; Po= 393 N/mm 2 ]
3 A 2 mm diameter tungsten carbide ball is pressed into a 5 mm diameter
hemispherical recess in a steel plate under a load of 50 N. Calculate
(a) the minimum hardness of the steel plate for the contact to remain fully
elastic; (b) the diameter of the contact zone.
[(a) H=5245 MPa; (b) 2a=0.056 mm]
4 Two similar, macroscopically flat steel surfaces are ground to give an
RMS value of 0.2 11m and a mean asperity radius of 300 Jlm. Calculate
(a) the maximum value of hardness at which the deformation of the
asperities is predominantly plastic and (b) the minimum value of hardness
at which the deformation of the asperities remains predominantly elastic.
Take the mixed elastic-plastic range to be defined by 0.5 < 'P < 1.
[(a) H=4750 MPa; (b) H=7920 MPa)]
Chapter 2
Friction
2.1 INTRODUCTION
Friction is the resistance to motion which occurs whenever one solid body
slides over another. The resistive force, which acts in a direction directly
opposite to the direction of motion, is known as the friction force. The friction
force which is required to initiate sliding is known as the static friction force,
while that required to maintain sliding is known as the kinetic friction force.
Kinetic friction is usually lower than static friction.
It has been observed experimentally that there are two basic 'laws' of
friction, which are usually obeyed to a good approximation. The laws are
entirely empirical and no physical principles are violated in those cases where
the laws are not obeyed.
The first law states that the friction is independent of the apparent area
of contact of the two bodies, and the second states that the friction force is
proportional to the normal load between them. Thus, a brick can be slid as
easily on its side as on its end, and if the load between two bodies is doubled,
then the friction force is also doubled. These laws are known as Amon ton's
laws, after the French engineer who first presented them in 1699.
Coulomb in 1785 proposed a third law, that the kinetic friction is almost
independent of the speed of sliding, but this law has a smaller range of
applicability than have the first two.
Amonton's second law states that the friction force, F, is proportional to the
normal load, W - i.e.
( 2.1 )
27
28 TRIBOLOGY
where F is the friction force, A is the real area of contact and s is the
friction force per unit area.
(2) The real area of contact, A, is proportional to the normal load, W- i.e.
A=q·W (2.3}
where q is the constant of proportionality.
In any apparatus for measuring friction, two specimens are placed together
under a known normal load, one of the specimens is caused to slide relative
to the other, and the tangential force resisting sliding is measured. There are
many methods available, of varying degrees of complexity and cost.
Probably the simplest method is to use the inclined-plane test illustrated
in Figure 2.1. In this test a plate of one of the chosen materials is securely
30 TRIBOLOGY
w
w
+
(a) (b)
I w~
-~- I
--_)
w
(c) (d)
Figure 2.2 Measurements of friction using conformal geometries: (a) flat pin on
disc; (b) thrust washer; (c) pin on cylinder; (d) shaft in bush
w
w
+
(a) (b)
The purpose of this section is not to give a state-of-the-art review of the large
amount of research that has taken place in this area, but rather to give a
clear understanding of the concepts involved.
We know that the friction force is due to interactions between the
opposing asperities of the two sliding surfaces. Each asperity interaction will
contribute to the friction force, so that the total friction force at any time will
be the sum of the forces at the individual contacts. Furthermore, as the
surfaces move relative to one another, the energy supplied by the forces
causing the motion is continuously dissipated at or near the sliding surfaces
and, again, the energy dissipated in unit time will be the sum of the energies
32 TRIBOLOGY
dissipated in the individual processes occurring during the same time. The
observation that friction forces are often almost constant is due to the fact
that the number of asperity interactions taking place at any time is so large
that the statistical distribution of the contact processes is almost constant.
Clearly, in equilibrium the energy dissipated per unit time is equal to
the product of the friction force and the sliding velocity, so we can approach
the problem of explaining the nature of friction by considering either the
origins of the forces at the contacts or the processes of energy dissipation. In
what follows it will be convenient to use both approaches.
In considering asperity interactions, it is conventional to describe two
types of interaction: adhesion and deformation. This convention will initially
be followed here, but as the discussion is developed, it will be stressed that
their contributions to friction are not simply additive, but interactive.
2.3.1 Adhesion
When two surfaces are loaded together, they can adhere over some part of
the true contact area, to form friction junctions, and these junctions must
then be broken if relative sliding is to take place. The break will occur at
the weakest part of the junction, which may be either at the original interface
or in the weaker of the two materials. If the break occurs at the original
interface, the interaction has simply subjected the two asperities to a stress
cycle, although, as we shall see in Chapter 3, the accumulation of such cycles
may ultimately cause the creation of a wear particle by a fatigue mechanism.
If the break occurs in the softer of the two materials, then a fragment of this
material will be transferred to the harder counterface. In either case it is clear
that adhesion is one form of surface interaction which would cause frictional
resistance.
Before considering the contribution of adhesion to friction, it is useful
to describe the adhesion of two solids placed in static contact. In doing this,
use is made of the thermodynamic concepts of surface and interfacial energies.
The surface free energy of a solid is defined as the thermodynamically
reversible energy involved in creating unit area of new surface. Similarly, if
two surfaces are brought into contact, the interfacial free energy is defined
as the thermodynamically reversible energy involved in creating unit area of
interface. Thus, if we take two surfaces having surface energies y1 and y2 and
place them in contact, to form an interface with interfacial energy y1 , 2 , then
the energy released is
ll.y = 1'1 + Y2- 1'1,2 (2.5)
This quantity is known as the thermodynamic work of adhesion as, in
principle, a similar amount of work is needed to separate the bodies.
Although Equation (2.5) has been applied very successfully to liquids,
it has been shown by Tabor (1985) that it cannot be applied precisely to
FRICTION 33
even ideally elastic/brittle solids. For solids of other types, the energy
calculated from Equation (2.5) is only a small fraction of the energy actually
required to separate the surfaces; the reasons for this will be made clear later.
However, although the thermodynamic work of adhesion is of little relevance
in many contacts, the actual work and force needed to separate the surfaces
are highly relevant. Tabor (1985) -proposed for these the apposite terms
'pull-off work' and 'pull-off force' and these will be adopted here.
2.3.2 Deformation
If no adhesion takes place, then the only alternative interaction which would
result in a resistance to motion would be one in which material is deformed
and displaced during relative motion. We need consider only two interactions
of this type: the microscopic interaction illustrated in Figure 2.4(a), where
deformation and displacement of the interlocking surface asperities are
required, and the more macroscopic interaction illustrated in Figure 2.4(b ),
where the asperities of the harder material plough grooves in the surface of
the softer. Other examples of the second type of interaction would involve
...
- Ia)
....
(b)
Figure 2.4 Sliding interactions of surfaces: (a) asperity interaction only; (b)
macroscopic ploughing interaction
34 TRIBOLOGY
This theory, due to Bowden and Tabor (1950), was the first modern
explanation of the existence of friction. It was developed for ideal elastic-
plastic metals and the same approach will be followed here, with the
importance of the theory to other types of materials being covered later in
the chapter.
The theory can be summarised as follows.
( 1) When surfaces are loaded together, they actually make contact only at
the tips of asperities.
(2) Even at low loads, the real contact pressures are so high that the asperity
tips of the softer material deform plastically.
(3) This plastic flow causes the total contact area to grow, both by growth
of the individual initial contacts and by initiation of new contacts, until
the real area of contact is just sufficient to support the load elastically.
(4) Under these conditions, for an ideal elastic-plastic material
W=A·p 0
where W is the normal load, A is the real area of contact and p0 is the
yield pressure of the softer of the two materials. The yield pressure is
very nearly the same as the hardness, H, measured in an indentation
test, so that this can be rewritten as
W=A·H (2.6)
( 5) As a result of the severe plastic deformation, the asperity junctions cold
weld - i.e. strong adhesive bonds are formed.
(6) The specific friction force, s, is then simply the force required to cause
shear failure of unit area of asperity junction, so that
F=A·s (2.7)
FRICTION 35
In the simple adhesion theory the effects of the normal and tangential loads
were considered separately; the true area of contact was assumed to be
36 TRIBOLOGY
determined by the normal load only (Equation 2.6) and the friction force
was then taken to be the force required to cause shear over this area (Equation
2.7). However, the very high friction coefficients observed between clean metal
surfaces indicate that the true area of contact must be far greater than that
predicted by the simple adhesion theory. This was explained when Bowden
and Tabor considered the combined effect of the normal and the shear
stresses.
To illustrate this, we consider first the simple two-dimensional stress
system shown in Figure 2.5(a), and assume that yielding occurs when the
maximum shear stress attains the critical value k. At this point k is given by
the radius of the Mohr's circle shown in Figure 2.5(b ). (The Mohr's circle
construction is an illuminating technique for illustrating compound stresses,
and readers who are not familiar with it should consult any standard text
on strengths of materials, e.g. den Hartog, 1952). Using this construction it
can be seen that yield occurs when
(p/2)2 + !2 = k2
or
(2.1 0)
We can now examine the effects of the combined normal and shear stresses
on the real area of contact in an asperity junction.
Consider first a single asperity junction under a normal load W, having
an area of contact A, defined by A = WI H. Note that the material is at its
yield point, defined by H, and that the maximum shear stress in the material
is equal to k, so that any increase in either the normal stress or the shear
stress in the junction will cause further plastic flow. Thus, if we impose an
increment of tangential force,f, some plastic flow will occur and this will lead
Shear
stress
llllllll
Normal
s srress
(a)
(b)
Figure 2.5 Mohr's circle construction for finding the maximum shear stress for
an idealised two-dimensional junction under normal and tangential stress
FRICTION 37
has remained constant, while the area over which it acts has greatly increased,
but it is worth emphasising that junction growth continues under the
increasing shear stress, even when the normal stress is negligible.
This revised treatment was still based on the assumption that all junctions
initially deform plastically under the applied normal load. However, this is
an unnecessary restriction, as the same type of reasoning can also be applied
to the many asperity contacts which, as shown in Chapter 1, deform elastically
under the applied normal load. In such contacts, although the initial contact
may be well within the elastic range, the same shear stress will still be required
to shear the junction, and for clean surfaces this will be the full shear strength
of either the interface or the weaker of the two materials. Thus, a contact
which was initially elastic will be subjected to plastic shear at a low level of
normal stress - a condition which is similar in all respects to the shearing
of an initially plastic contact after some junction growth.
We have now explained the very high friction coefficients which can be
obtained with clean metals in vacuum. We have not yet explained why such
high coefficients are not observed in normal atmospheres, but it is clear that
junctions formed in such atmospheres must be weaker in shear than those
formed between clean surfaces, and that this relative weakness must be due
to the presence of the contaminant films which are always present in normal
atmospheres. Following Bowden and Tabor, we therefore assume that at any
junction there is a thin contaminant film of shear strength r., and that r. = ck,
where c is less than unity. While the frictional stress, F I A, is less than r.,
junction growth will proceed as described above; but when F /A= r., the
contaminating film will shear, junction growth will end and gross sliding will
occur.
Thus, the condition for sliding is
p2 + ar; =H 2
But it has already been shown that
H 2 =ak 2
Therefore,
or
Hence,
(2.14)
FRICTION 39
20
Figure 2.6 Variation of Jl with c for different values of IX. It can be seen that, except
at large values of c, the exact value of IX is not of major importance
40 TRIBOLOGY
It is worth noting that this modified theory will also apply when the
interface is inherently weaker than either of the asperity materials, even in
the absence of contaminating films: in this case -r. is simply the shear strength
of the interface.
In this final modification to the adhesion theory, unlike the case for
clean surfaces, there can be a very marked difference in behaviour between
the junctions which are initially plastic and those which are initially elastic.
In the former case there will always be some continued plastic deformation
and junction growth until the interface shears, whereas in the latter case the
interface may shear while the deformation of the asperities is still within the
elastic range. The importance of this fact will emerge in Chapters 3 and 4.
In the foregoing discussion we have presented the various versions of
the adhesion theory· as first put forward by Bowden and Tabor. Johnson
(1968) has suggested that Equation (2.11) should be modified to
(2.15)
so that Equation (2.14) becomes
c
11 - -----=-·
]l
-[(J(2-(J(1C 2
( 2.16)
and
Therefore,
p2 + O(C2k2 = O(k2
where
(2.17)
FRICTION 41
(a)
a,
(b)
a,p
Hence,
fJ. = ckjp = c/[1X(1- c 2 )]! (2.18)
It can be seen that Equation (2.18) is of the same form as Equation (2.14),
but the constant IX has been replaced by the variable given by Equation (2.17 ).
The values of IX for the extreme cases of c = 0 and c = 1 can be calculated
from the Mohr's circles shown in Figures 2.8(a) and 2.8(b), respectively. It
can be seen,' for a rigid-plastic material in plane strain, that when r = 0,
t.
a 1 = a 2 = while when r = k, a 1 = a 2 = 1. Inserting these values into
Equation (2.17) shows that IX has a minimum value of 36 when c = 0 and
tends to infinity as c tends to 1. The variation of IX between these limiting
conditions is not known, and the equations would need to be modified for
the more realistic case of general triaxial stress and for either fully elastic
deformation or plastic deformation with work hardening. However, this
treatment does bring out the important point that a friction coefficient will
depend on both the interfacial shear strength, as represented by ck, and the
deformation characteristics of the material.
42 TRIBOLOGY
p=H
(a)
(b)
Figure 2.8 Mohr's circles for extreme values of surface shear stress: (a) c = 0;
(b) c = 1
2.4.3 Ploughing
W;
Similar expressions can be calculated for other asperity shapes. For most
engineering surfaces, the asperity angles are large and the ploughing
component of friction due to asperity interaction is correspondingly small.
For example, Equation (2.17) shows that for an asperity angle of 85°, which
corresponds to an unusually rough surface, the ploughing coefficient of
friction would be only 0.056. However, abrasive material, including work-
hardened or oxidised wear debris may be very angular, and when significant
amounts of such material are present between the rubbing surfaces, the
ploughing component of friction may be much higher.
In the adhesion theory described above the normal and yield stresses on a
single asperity were assumed to be representative of the stresses on all
asperities. In deformation theories, it is recognised that the normal and shear
stresses on the asperities will vary during the lifetime of a junction. The
physical basis of the analysis is that, in the sliding of macroscopically flat
surfaces, motion is parallel to the interface and the separation of the surfaces
remains constant. This must be so, to maintain the area of contact at the
constant level which will support the constant normal load, and the significant
consequence of this is that the contacting asperities must deform to allow
movement to continue.
This type of theory was first advanced by Green (1955) and has been
extended by other workers. To illustrate the principles of such theories, the
treatment of Edwards and Halling (1968) is summarised here.
Edwards and Halling considered two wedge-shaped asperities of
semiangle ()moving as illustrated in Figure 2.10. By assuming that the solids
are ideally elastic-plastic and that the asperities deform plastically, it is
possible to calculate the instantaneous values of shearing force Fi and normal
load ltj over the complete life of the junction. This was done both for the
case of intimate contact, where the shear strength of the junction was taken
to be k, and for the case of asperities separated by a film of shear strength
ck, where c and k have the same meanings as in the Bowden and Tabor
theory. Calculations were made for various values of() and c. The variations
44 TRIBOLOGY
Figure 2.10 The idealised wedge-shaped asperities studied in the 'plastic interaction
theory'
\
\
,... _ _ c•0·8
/
"' "' ..............
..........
I '.._
I ' ...
',J
Figure 2.11 Variation of the normal force, P, and the friction force, F, throughout
the junction life for a junction of angle 10°
FRICTION 45
2<l
1·5
,,,
\
\
June lion
0·5 \ angle
'' 30"
--- --- --
.....
c ..........
li = .....,,(...,-1--c""2~)•
o•
1·0 0·5 0
c
Figure 2.12 Variation of Jl with c for various junction angles
The general equation for J.l. given by the Edwards and Halling theory is
where cf> is a function of c and the geometry of the junction, and is equal to
zero when () is zero.
It is interesting to note that when cf> is zero, this equation reduces to a
form which is identical with the modified Bowden and Tabor theory
(Equation 2.14).
It is a characteristic of all asperity interaction theories of friction that
the energy consumed in plastic deformation increases with the sharpness of
the asperities. This is illustrated in Figure 2.12 for the Edwards and Halling
model, and similar trends can be shown to occur with, for example,
hemispherical asperities of decreasing curvature. In the case of elastic
deformation of asperities, there is no deformation component of friction, but
the real contact area of a junction and, hence, the force required to shear it
46 TRIBOLOGY
increase with asperity slope. These two facts account for the now widely
accepted observation that friction between macroscopically smooth surfaces
increases with the mean absolute surface slope.
There are only three basic mechanisms which could possibly cause appreciable
energy dissipation at asperity junctions - i.e. fracture, elastic deformation
and plastic deformation.
New fracture surfaces will be created whenever an asperity junction is
broken, whether at the original interface or within one of the asperities, and
whenever a wear particle is formed. However, as will be shown in Chapter 3,
except in the case of very severe wear, asperities are normally subjected to
many thousands of stress cycles before a wear particle is formed. Therefore,
the energy losses associated with wear particle formation will, in general, be
negligible, and the contribution of the fracture mechanism to friction will be
largely determined by the energy dissipated in breaking the adhered junctions,
either at the original interface or within one of the contacting asperities. For
brittle materials this will correspond to the energy required to create the two
new surfaces, but for ductile metals, which we are considering here, it will
be almost entirely determined by the energy consumed in plastic deformation
prior to fracture.
Elastic hysteresis and viscoelastic losses can be dominant factors in the
friction of elastomers and polymers, and they will be discussed later in this
chapter. However, the above discussion of friction theories was restricted
to ideal elastic-plastic solids and there can be no energy loss during the
elastic deformation of such materials. As many metals are almost perfectly
elastic, it is normally assumed that elastic hysteresis can also be ignored as
a contributory factor to the observed sliding friction of metals.
Therefore, we must conclude that plastic deformation is the principal
cause of metallic friction. At first sight this is a rather surprising conclusion,
since, in general, the outermost asperities will always undergo fully plastic
deformation during, at most, the first few contacts; they will then have
flattened sufficiently to remain nominally elastic. However, on subsequent
contact the required elastic stress will be similar to that attained on the first
contact - i.e. close to the yield pressure. If we regard the asperities as having
spherical geometries, we can see that we now have highly loaded Hertzian
contacts, and the surface pressure will be sufficiently high for the maximum
shear stress to exceed the critical shear stress for the onset of plasticity. Under
normal loading only, the critical shear stress will occur below the surface
and the plastic zone will be constrained. This will also be the case if the local
FRICTION 47
friction coefficient - i.e. the ratio between shear stress and normal stress on
the asperity - is less than 0.25. In these cases plastic deformation must be
very limited, but the energy consumption will still make some contribution
to the total friction force. However, if J1 is greater than 0.25, the plastic zone
will be at the surface, the deformation will not be constrained and it is likely
that a more significant amount of plastic work will be done on each pass of
an opposing asperity.
Thus, even when the contact geometry is such that deformation
would be expected to be elastic, there may be a significant plastic
deformation contribution to the observed frictional work, and, as stated in
sub-section 2.4.2, this must be the case when the interfacial shear strength
is high, even when the direct stress is very low.
The above arguments will apply to some extent to any asperity which
is subjected to an element of plastic deformation (i.e. to a normal pressure
greater than 0.6H), so that a high proportion of the contacting asperities
will be subjected to the repeated plastic strain described above.
To assess the relative importance of the various contributions to the
total friction, we remember that it takes many thousands of contacts to
produce a wear particle. Only very few of the early contacts involve gross
plastic deformation and only one - the final one - involves fracture, so that
at any time the overwhelming majority of asperity contacts will be nominally
elastic, with about half ofthem experiencing stresses close to the yield pressure.
It is inconceivable that the very small fraction of fully plastic contacts will
make a significant contribution to the overall friction force, so the deformation
contribution to friction will arise almost entirely from localised plastic
deformation of the nominally elastic contacts.
The energy dissipation mechanism described above is identical with that
described by Johnson for rolling friction under high stresses. We shall refer
to it again in Section 2.8, when we discuss rolling friction.
As we know, when any two surfaces are placed together under a normal
load, the surfaces approach each other until the elastic reactive force at the
interface is just sufficient to support the normal load. During sliding, a similar
situation obtains, with the reactive force being required to support both the
normal and the tangential forces.
Force equilibrium demands that each ofthe asperities forming a junction
must be subjected to the same stress, and behaviour during sliding is therefore
very dependent on the relative yield pressures of the two materials. If the
materials are identical, then they will be strained in the same way, either
elastically or plastically, depending on their degree of overlap. However, if
their yield pressures are significantly different, the softer material may deform
plastically but the deformation of the harder one must be nominally elastic.
48 TRIBOLOGY
As deformation continues, the softer material may work harden; but as long
as its current yield pressure is lower than that of the opposing surface, the
deformation of the harder material must remain nominally elastic.
It is important to note that the softer material may sometimes cause
localised plastic deformation in the harder material. This is because the
pressure to cause full plasticity of an asperity is approximately twice as great
as that required for initiation of subsurface plasticity. Thus, an interfacial
pressure which is sufficiently high to cause full plasticity of the softer asperity
will also initiate plastic deformation in the harder asperity as long as the
hardnesses do not differ by a factor of more than about 2. This will be
seen to be particularly important when we come to discuss fatigue wear in
Chapter 3.
If one material is very much harder than the other, the condition of
stress similarity can only be obtained if the harder asperity penetrates the
softer surface, with a consequent ploughing contribution to the friction force.
Ploughing is the extreme form of plastic deformation: in what follows we
shall consider the less extreme forms of plastic deformation during asperity
contact.
For clarity, we shall consider the total life of an asperity junction to
comprise three stages: formation, shearing and separation. This is a rather
unrealistic picture, because the surfaces are macroscopically in continuous
relative motion; therefore, the processes of formation and shear will overlap,
as will the processes of shear and separation. Nevertheless, it is a useful way
of illustrating the way in which adhesion and deformation interact. The three
stages can be characterised by the sense of the direct stress normal to the
interface: during junction formation the direct stress is compressive; during
shear it progressively falls to zero; and during separation, in the presence of
adhesion, it becomes tensile.
During junction formation the only significant process of energy
dissipation will be plastic deformation of one or both asperities; the actual
amount of energy dissipated will depend on the yield pressures and
work-hardening characteristics of the plastically deformed asperities, and on
the plastic strains which occur. This plastic deformation will result from
the combined effects of the normal and shear stresses, as described in
sub-section 2.4.2, and must continue until the total deformation is sufficient
to allow the asperities to pass at constant surface separation. The plastic
deformation can result in significant junction growth, and may also disrupt
any protective surface films, so that strong adhesive bonds are formed over
part of the junction.
If the interface is weak, sliding may occur at the interface and the only
adhesive contribution to the friction will be the small force required to shear
this weak interface. However, if the interface is strong, owing to the formation
of strong adhesive bonds, plastic deformation may continue, initially under
compression, then in pure shear and finally, as the asperities begin to separate,
FRICTION 49
Ductile
Figure 2.13 Comparison of pull-off work for brittle and ductile materials having
the same pull-off force
50 TRIBOLOGY
2.6.1 Elastomers
2.6.2 Polymers
shear angle correlates with the energy-to-rupture of the polymer, and the
frictional work is largely accounted for by the associated energy input.
2.6.3 Ceramics
of the lubricant other than bulk viscosity'. It is, therefore, a rather ill-defined
regime which covers the operating range between dry sliding and elasto-
hydrodynamic lubrication: a fluid film is present but it is not sufficiently
thick to prevent contact between the higher asperities.
The situation can be characterised by the value of the A ratio, hfa,
where h is the separation of the mean planes of the two surfaces and a is
the composite roughness (af +ant. As we saw in Chapter 1, for surfaces
having Gaussian or near-Gaussian asperity height distributions, surface
separation is virtually complete if A~ 3. For values of A less than 3, some
asperity contact will occur, and both friction and wear will be expected to
increase as A decreases.
It should be clear from the earlier parts of this chapter that, if we wish
to ameliorate the effects of this asperity contact, we need to interpose a film
of low shear strength between the sliding surfaces. Therefore, oils intended
for boundary lubrication contain additives which will deposit such films on
the surfaces.
The most common type of additive is the aliphatic 'fatty' acid of the
general formula CnHln+ 1 COOH, a typical example being stearic acid. Each
molecule of such an acid consists of a long covalently bonded hydrocarbon
chain with a single reactive acid group at one end. The acid groups of these
molecules react with surface metal atoms to form a strongly bonded metallic
soap, and the hydrocarbon chains align themselves normal to the surface,
thus providing a very effective barrier to metal-metal contact. The process
is illustrated schematically in Figure 2.14.
H
I
H- C -H
H- C- H
I
I
I
I
H- C- H
0 # " OH
(a)
(b)
Figure 2.14 Boundary lubrication by long-chain fatty acid molecules: (a) molecular
structure; (b) surface protection by reacted layer
54 TRIBOLOGY
Common
normal
The frame of reference as defined above moves with the linear velocity
of the contact point 0 and rotates with an angular velocity which maintains
its orientation relative to the tangent plane and the common normal at 0.
We can then state the following definitions.
Sliding is the relative linear velocity between the two surfaces at 0; as
there can be no component of relative velocity along the normal, sliding is
always in the tangent plane.
Rolling is the relative angular velocity between the two bodies about an
axis lying in the tangent plane.
Spin is the relative angular velocity between the two bodies about the
common normal through 0.
Any relative motion between two contacting bodies can be regarded as
a combination of sliding, rolling and spin. In the previous sections we have
been concerned only with sliding, but in the remainder of this section we
shall describe frictional effects due to combinations of sliding, rolling and spin.
To understand such effects, we must also understand the stresses which
can arise in the contacting bodies due to the forces acting at the interface.
In Chapter 1 we considered in some detail the stresses which arise at
contacting non-conforming surfaces due to normal forces acting at the
interface, but we referred only in passing to the stresses which arise due to
tangential forces. We now need to consider in rather more detail the combined
effects of normal and tangential forces.
The resultant force transmitted across any interface can always be
resolved into a normal force, P, acting along the common normal and a
tangential force, Q, acting in the tangential plane. The tangential force Q is
reacted by frictional resistance and its magnitude must therefore be less than
or equal to Jl-P. When Q attains the limiting value of Jl-P, gross sliding occurs
and the direction of Q is directly opposed to the direction of sliding.
The interfacial forces lead to a contact area of finite size, and the
distribution of forces over this area can, therefore, lead to the transmission
of a moment across the interface. The components of this moment about
axes lying in the tangent plane are known as rolling moments, while the
component about the common normal is known as the spin moment.
Foil owing Johnson ( 1985 ), we define free rolling as rolling motion in
which the tangential force Q is zero, and tractive rolling as rolling motion
in which the tangential force Q is non-zero but less than the limiting value
of Ji-P at which gross sliding will occur.
The interfacial forces and moments described above are transmitted
across the interface by surface tractions, the normal traction, which is simply
the contact pressure, being denoted by p and the tangential traction by q.
In general, both p and q vary over the contact area, so they can be written,
respectively, as p(x, y) and q(x, y), where (x, y) denotes position in the tangent
plane. It is important to realise that tangential tractions can be present even
56 TRIBOLOGY
when the tangential force Q is zero; in such cases it is simply necessary that
the tractions integrated over the contact area should sum to zero.
As p(x,y) andq(x,y) have different distributions over the contact area,
the ratio q(x, y)/p(x, y) also varies with position within the contact area.
As a result, two different contact conditions may obtain: ( 1) wherever
q( x, y) < Jl.P( x, y ), local sliding is prevented by friction and the corresponding
regions of the interface are said to be regions of 'stick'; (2) wherever
q( x, y) = Jl.P( x, y ), local sliding is possible and the corresponding regions are
said to be regions of 'slip' or 'microslip'.
A difference between the tangential strains in the two bodies in the
'stick' area leads to a small apparent slip known as creep. Creep can be
understood by considering the example of a deformable cylinder rolling under
load on a relatively rigid surface. If the tangential strain in the cylinder is
tensile, the cylinder is stretched in the 'stick' area. The cylinder then behaves
as though it has an enlarged circumference, and therefore in one revolution
it moves forward by a distance greater than its actual circumference. The
ratio between the distance traversed in one revolution and the actual,
unstretched circumference is known as the creep ratio.
The actual location of the boundaries between the areas of 'stick' and
'slip' is quite difficult and need not concern us here. Readers requiring a
fuller treatment of such problems are referred to Johnson ( 1985). In the
following sections we shall simply describe in qualitative terms the locations
and sizes of the zones of stick and slip for various rolling situations, starting
with the most simple situation of free rolling between identical elastic bodies
and progressively introducing more variables.
The simplest form of free rolling occurs between two bodies which have the
same elastic properties and are geometrically identical. When such bodies
roll freely under the action of a normal force, no tangential tractions or slip
can occur and the contact stresses and deformations can be calculated from
the Hertz equations for normal loading, which are given in Chapter 1.
So far we have considered situations where the contact area lies substantially
in a single plane. However, in many engineering situations, particularly ball
bearings of all types, this situation does not obtain. For a ball rolling along
a grooved track, as illustrated in Figure 2.16, the contact area will be an
A A
ellipse and, provided that the angle pis less than 30°, the dimensions of the
ellipse are still predicted with reasonable accuracy by the Hertz equations.
If we now consider the situation when the ball has rolled forward through
one complete revolution, we can see that in the centre of the groove the ball
will have measured out its full diameter along the track, while points on the
ball at the edges of the contact zone will have measured out a smaller distance
corresponding to the circumference of the circle of radius ri. These
differences can be accommodated by the ball slipping in its track, and we
might expect the contact area to consist of three zones of slip, as shown in
Figure 2.16- i.e. a single central zone of backward slip and two outer zones
of forward slip. In this situation the axis AA, which passes through the two
boundaries between the regions offorward and reverse slip, can be considered
to be the instantaneous axis of rotation. This situation, which is known as
Heathcote slip, was first put forward by Heathcote (1921) as it is described
above. However, Johnson (1985) has shown that some of the tendency to
slip can be accommodated by elastic deformation of the contact areas.
For problems involving both rolling and spin, as in thrust ball bearings,
and for problems of materials having dissimilar elastic properties, the
solutions become much more complex than those described above. The
problems have been treated in detail by Johnson ( 1985), but are beyond the
scope of this book.
Expons1on
f- 2a -j
Figure 2.17
contacting material, the normal force per unit length of the roller will be
p dx, and this will exert a resistive moment about the centre of the contact
zone of magnitude px dx. Thus, the total moment exerted by all the elements
on the front half of the roller will be
M =I: px dx = 2Pa/3n
If the cylinder rolls forward through a distance x, the elastic work done by
this couple is given by
cp = Mx/ R = 2Pax/3nR
Under ideal conditions the cylinder would be subjected to an equal and
opposite moment arising from the pressure distribution in the rear half of
the contact. However, owing to elastic hysteresis losses, not all the work is
recovered and there is some dissipation of energy. If we define the hysteresis
loss by a coefficient e, we can see that the energy dissipated in rolling a
distance x will be scp, and if this loss represents the rolling resistance, it
follows that
Fx = scp = s·2Pax/3nR
where F is the force required to overcome the resistance. By analogy with
sliding friction, the ratio F I P can be defined as the coefficient of rolling
resistance, A, where
A= 2ea/3nR (2.21)
Using a method similar to the above, it can be shown that the coefficient of
rolling resistance for a sphere rolling on a plane is given by
A= 3ea/16R (2.22)
60 TRIBOLOGY
Equation (2.20) can also be used for elliptical contacts if a is taken as the
half-width of the contact in the direction of rolling.
If the contact pressure on first contact exceeds a critical value, then, as
described in sub-section 1.5.1, some subsurface plastic deformation takes
place and the energy dissipated in such deformation can be substantially
higher than that due to elastic hysteresis. However, almost all rolling contacts
are subjected to repeated stress cycles and, owing to the accumulation of
residual stresses, the degree of plastic deformation can decrease with repeated
cycling until, after a relatively small number of cycles, the deformation has
become fully elastic. This process is referred to as shakedown. However, if
the contact stress exceeds a second critical value, which is higher than the
value for first yield, shakedown does not occur and there is some plastic
deformation on each stress cycle. Johnson ( 1985) has calculated the ratio
between the maximum shakedown pressure (p 0 ). and the pressure for first
yield (p 0 )y for a variety of situations. For free rolling of an elastic cylinder
on an elastic-perfectly plastic half-space, this ratio is 1.29, and in the presence
of tangential tractions the ratio decreases. For rolling of a sphere on an
elastic-plastic half-space, the ratio is 2.2.
When the load exceeds the shakedown limit, some plastic deformation
occurs on each cycle and the surface layers are displaced forward relative to
the deeper layers. The rolling resistance due to this effect can then be much
greater than that due to elastic hysteresis. At even higher loads, when the
plastic zone extends to the surface, the rolling resistance increases to even
higher values.
M, I
u~--f--o,
X
0 X -
(3) The torque MZ' the self-aligning torque which acts to straighten the
course of the vehicle.
When the vehicle is travelling on a straight line, both QY and M z are
zero, but if we consider the car being driven round a left-hand turn, the force
system and the corresponding microslip velocities are as shown in Figure 2.19.
Now the path of motion lags behind the centre line of the tyre by the slip
angle (), which can be interpreted as the angular microslip due to the
self-aligning torque Mz. The relationships between QY, Mz and () are then
shown in Figure 2.20 (Gough, 1954). In the region ABC, the self-aligning
torque and the slip angle increase with the cornering force. In the region CD
there is little change in the self-aligning torque as the cornering force and
the slip angle increase, and the feel of the steering is therefore lost. In the
Vehicle
polh
Figure 2.19
62 TRIBOLOGY
I
I '
c ,'
"
Self-aligning torque ~
Figure 2.20 Relation between the major variables during cornering- Gough plot
region DE, which approaches the point of skidding, there is a reversal in the
feel of the steering as the car drifts around the corner. Although the effects
of the normal load and the friction coefficient are modest in the region ABC,
they become increasingly important as the point E is approached.
Almost all frictional energy is dissipated in the form of heat. As this heat
release is a continuous process, temperature gradients will be set up in the
contacting bodies, with the highest temperature at the heat source- i.e. the
contact surface. The overall temperature rise can, in principle, be calculated
from the rate of input of frictional energy and the thermal properties of the
contacting bodies and the environment; a theoretical treatment and applications
to many practical problems have been discussed by Carslaw and Jaeger
(1959). However, as we now know, the bodies make contact only at the
asperity tips, and all the heat must be generated at these small areas of
contact. As a result, the local temperatures may be instantaneously very
much higher than the bulk temperatures of the bodies. These high temperatures
of short duration are known as 'flash temperatures', and they can be very
important in changing, for example, the microstructures of alloys or the rate
of oxidation and the types of oxide formed. A full analysis of the magnitudes
of flash temperatures has been made by Jaeger (1942), but this analysis is
very complicated and we shall summarise the much simpler treatment given
by Archard ( 1959).
The temperature effects will depend on whether a body is stationary or
moving with respect to the heat source. Archard (1959) based his analysis
of the various problems on the model shown in Figure 2.21, with a
proturberance on the surface of body B forming a circular area of contact
A = na 2 and moving with velocity V over the flat surface of body C. Body B,
therefore, receives heat from a stationary heat source and body C from a
FRICTION 63
(B)
Figure 2.21
moving heat source. Then, for the two bodies, Q8 and Qc are the quantities
of heat supplied per second, K 8 and Kc are the thermal conductivities, c 8
and cc are the specific heats, p 8 and Pc are the densities, and XB and Xc
(given by X = K I pc) are the thermal diffusivities. (]mean is the mean temperature
of the contact area, taking the temperature of a point far removed from the
contact as zero.
Then Archard derives the following temperature rises for the stationary
heat source (body B) and a slow-moving heat source (body C) to be,
respectively,
( 2.23)
and
(}mean = Qc/ 4aKc ( 2.24)
At higher speeds Equation (2.24) does not apply, as there is insufficient time
for the temperature distribution of a stationary contact to be established in
body C. The speed at which this occurs is determined by the dimensionless
parameter
L= Va/2x (2.25)
Equation (2.24) applies to a reasonable approximation for L < 0.1. For L > 5
the temperature rise is given by
(]mean= (0.31Q/ Kca)·(xc/Va)t (2.26)
We can define a parameter N = nq/ pc V, where q is the rate of heat supply
per unit area. Then, for a stationary heat source or a very slow-moving heat
source (L < 0.1 ),
(]mean= 0.5NL (2.27)
64 TRIBOLOGY
10
10-1
10 10 2 103 104
2.10 PROBLEMS
(In solving these problems, assume that the friction coefficient is given by
Jl = f.ladhesive + J.iploughing ·)
1 When a hard steel ball of diameter D is rubbed across a soft metal surface,
it ploughs out a groove of width d. If d « D, derive an expression for the
ploughing contribution to the friction coefficient.
[Jl = d/nD]
2 A hard steel ball is slid across a soft metal surface under two different
loads. In one case the friction coefficient is 0.22 and the groove width is
FRICTION 65
0.3 mm; in the other case the friction coefficient is 0.25 and the groove
width is 0.8 mm. Calculate the diameter of the slider and the adhesive
contribution to the coefficient of friction.
[d = 5.3 mm; Ji = 0.202]
3 Two hard conical sliders of sell}iangles 75o and 80° are slid across an
unlubricated metal surface and the measured friction coefficients are in
the ratio 11:10. The experiment is then repeated with the surfaces
lubricated and the measured friction coefficients are then in the ratio
13:10. Find the coefficients of adhesive friction in the two cases. If the
effect of the lubricant is to reduce the surface shear stress by a factor
of 3, what are the values of c and oc for the unlubricated surface?
[Jidry = 0.5938; JiJub = 0.0866; C = 0.905; OC = 13.96]
Chapter 3
Wear
3.1 INTRODUCTION
66
WEAR 67
tests, and then field tests. Fortunately, because of the economic importance
of wear, many results of such tests have already been published (see, for
example, Peterson and Winer, 1980), and potential users can refer to these
to make an initial, or sometimes final, choice of materials.
Although there is much current research which is increasing our
understanding of wear processes, it is probable that material selection will
remain empirical for the foreseeable future. It is true that there are theoretical
wear equations, some of which will be described later in this chapter, but
they always contain factors which have to be determined by experiment.
Furthermore, even these experimentally determined factors are inherently
imprecise, and in many cases the user must be satisfied if an observed wear
rate is within an order of magnitude of that predicted on the basis of empirical
data. The reason for this lack of precision can be described as the dependence
of wear rates on extreme values, rather than mean values, of experimental
variables. Examples are the domination of the surface chemistry by elements
which are only minor impurities in the bulk (Miyoshi and Buckley, 1985);
the presence of a few relatively hard inclusions in what is nominally a softer
material; the probable disproportionate importance of the heights of the few
outermost asperities of a hard material, rather than its average surface
roughness; and the presence or absence of contaminants from the environment,
particularly in the form of abrasive dust.
However, there are now generally accepted qualitative explanations of
the various processes underlying wear behaviour, although there is much
that remains to be clarified, particularly the relative importance of the various
processes. It would not be possible in a book of this length critically to
review the vast amount of work which has taken place in these areas, and
in discussing the relative importance of the various processes the views
expressed must necessarily be the subjective views of the author. No claim
is made for the infallibility of these views, but it is hoped that they do
represent a clear and consistent picture of wear, which will provide the
interested reader with a background on which to base further reading.
We shall initially concentrate on metals, as they have been subjected to
more detailed analysis than have other materials and have a wide range of
mechanical properties. In Chapter 4 we shall show how this basic understanding
can be extended to the behaviour of other types of materials, particularly
ceramics, polymers and elastomers, how it then allows an informed selection
of candidate pairs of material to be made and, where possible, how wear
behaviour can be predicted on the basis of published data.
In a recent review paper Godfrey ( 1980) states that there are about twelve
kinds of wear, each with symptoms which must be recognized for the wear
68 TRIBOLOGY
The theory of adhesive wear has the same basis as the adhesion theory of
friction which was described in sub-section 2.3.1. We shall initially follow
the historical development of the theory, as it was originally proposed by
Bowden and Tabor and formulated as a semi-empirical law by Archard ( 1953),
but in Section 3.4 we shall examine the theory rather more critically and
show that some phenomena currently ascribed to an adhesive wear process
are explained rather more satisfactorily by a fatigue mechanism.
The basic mechanism underlying the theory has already been explained
in sub-section 2.3.1 - i.e. strong cold welds are formed at some asperity
junctions and these welds must be sheared for sliding to continue. The
amount ofwear then depends on where the junction is sheared: if shear takes
place at the original interface, then wear is zero, whereas, if shear takes place
away from the interface, a fragment of material is transferred from one surface
to the other. This transfer of material is observed in practice, normally from
the softer material to the harder, but occasionally from the harder to the
WEAR 69
softer. The theory at this stage has explained only material transfer and not
the formation ofloose wear debris; in the original theory it was simply stated
that subsequent rubbing causes some of the transferred particles to become
detached. We shall return to this point in Section 3.4.
Here we follow the approach of Archard ( 1953 ), who first derived a theoretical
expression for the rate of adhesive wear.
It is assumed that the area of contact comprises a number of circular
contact spots each of radius a. The area of each contact spot is na 2 and,
assuming, as before, plastic deformation of an ideal elastic-plastic material,
each contact supports a load of na 2 H, where H is the yield pressure, or
hardness. The opposing surface will pass over the asperity in a sliding distance
2a and it is assumed that the wear fragment produced from each asperity is
hemispherical in shape and of volume 2na 3 /3.
Then the wear volume !5Q produced by one asperity contact in unit
sliding distance is given by
!5Q = (2na 3 /3 )/2a = na 2 /3
and the total wear volume, Q, per unit sliding distance is Q = nna 2 /3, where
n is the total number of contacts. But each contact supports a load of na 2 H,
so the total load is W = nna 2 H, and nna 2 = W/ H. Hence,
Q= W/3H (3.1 )
This equation suggests that there should be three 'laws' of adhesive wear.
( 1) The volume of material worn is proportional to the sliding distance.
(2) The volume of material worn is proportional to the load.
(3) The volume of material worn is inversely proportional to the hardness
of the softer material.
The first of these 'laws' is found to be true for a wide range of conditions.
The second 'law' is often true from very low loads up to some point at
which the rate of wear increases catastrophically. At the point where this
transition occurs the apparent pressure (the load divided by the apparent
area of contact) is approximately equal to or slightly less than one-third of
the hardness. We know from Chapter 1 that H/3 is approximately equal to
the bulk yield stress. Therefore, this transition clearly corresponds to the
point at which the whole surface begins to deform plastically, so that the
asperities can no longer be considered to be independent. We also know
from Chapter 1 that in the presence of frictional tractions the material will
yield at a lower normal stress, and we should, therefore, expect the wear
transition to occur at a pressure somewhat lower than H /3; this has been
confirmed experimentally (Amell et al., 1975).
70 TRIBOLOGY
The term 'abrasive wear' covers two types of situation, known, respectively,
as two-body abrasion and three-body abrasion, in each of which a soft surface
is ploughed by a relatively hard material. In two-body abrasion a rough
WEAR 71
____ _L
can be simply piled up at the sides of the grooves and not lost to the surface.
As a result, typical values of k are ~ 0.1.
Two-body abrasion is typified by the action of a file or emery paper
on a softer metal surface. There are two requirements for this process to take
place: one surface must be harder than the other; and the asperities on the
harder surface must be 'sharp' - that is, have high values of tan fJ. The
roughness of the surface does not appear in Equation (3.7) but, although
the worn volume does not depend on roughness, in practice it is also necessary,
if the effects of abrasive wear are not to be too severe, that the hard surface
should have a low roughness; this ensures that the abraded surface has a
myriad of very fine abrasion marks rather than relatively few very deep and
severe ones. Luckily, most finishing processes which reduce asperity slopes
also reduce surface roughness.
Severe two-body abrasive wear of lubricated surfaces has been largely
eliminated from modern machinery, owing to greater awareness of the
importance of material properties and surface finish, and the availability of
surface profilometers to ensure that engineering surfaces are of the required
quality. Studies of this form of wear are now more related to polishing and
grinding processes, where two-body abrasion is deliberately employed.
However, in the absence of effective lubrication, when adhesive transfer can
take place, both the worn surface and the transferred material can be
roughened and work hardened so that two-body abrasion becomes a
significant problem.
Three-body abrasion is still an important cause of wear. Many moving
parts operate in environments containing significant amounts of adventitious
dirt and dust; the products of corrosive wear are more often than not abrasive
in character, and metallic wear particles are often fully work hardened, and
thus abrasive to their parent surfaces. The practical solution to the problem
of three-body abrasion is to flush away abrasive particles and permanently
to remove them by filtration.
Rolling Contact
Adhesive and abrasive wear depend on direct solid-solid contact and these
wear processes operate throughout a period of rubbing. However, if the
moving surfaces can be separated by a lubricating film, as in hydrodynamic
and hydrostatic bearings, these mechanisms of wear cannot operate. In the
WEAR 73
Sliding Contact
1959) fatigue equation, which relates fatigue life to the plastic strain increment
during each cycle, and others (Kragelsky, 1965; Soda et al., 1977) have
suggested fatigue as a predominant process at low wear rates. In what follows
we hope to show that fatigue is probably one of the more important
contributory factors in almost all mechanical wear processes.
To do this, we return to the phenomenon of shakedown (Johnson, 1985)
which we described in Chapter 2. We recall that, in rolling contact of
non-conforming solids, the type of deformation which occurs on repeated
loading is critically dependent on the stress level. At stresses below the
shakedown stress, contact will become entirely elastic within a few stress
cycles, whereas, at any higher stress level, there will be an increment of plastic
deformation on each loading cycle.
We should, clearly, expect qualitatively similar behaviour in the sliding
contact of non-conforming asperities. There will be a limiting stress below
which deformation on repeated cycling will be entirely elastic. However, this
stress may be closer to the stress for first yield than is the case in macroscopic
rolling contact, because of the difficulty of accumulating in the microscopic
contact zones the residual stresses which are necessary for shakedown. At
stresses only marginally above the shakedown stress, plastic deformation will
be subsurface and constrained, but at higher stresses the zone of plastic
deformation will extend to the free surface and be unconstrained. Again,
owing to the much higher friction tractions which occur in sliding contacts,
the zone of plastic deformation may reach the free surface at much lower
stresses than those required in rolling contacts. Indeed, as we showed in
Section 2.4, in the presence of strong interfacial bonds plastic flow may occur
in shear under very low normal stresses.
In any event, each cycle of plastic deformation will correspond to
one fatigue cycle, and the accumulation of such cycles will eventually cause
a fatigue fracture and the generation of a wear particle.
Clearly, in assessing the contribution of fatigue to mechanical wear, we
need an estimate of the proportion of the contacting asperities which will be
subjected to stresses above the fatigue limit, which, for simplicity, we take
to be equal to the stress for first yield. This estimate can be made from a
more detailed examination of the data summarised in Figure 1.20, which
shows the ratio between plastic and elastic contact area, under normal loading
only, for different values of the plasticity index and different normalised
contact pressures. A selection of the data which were computed to plot these
graphs is tabulated in Table 3.1, where R is the ratio between plastic and
elastic contact area; ljJ is the plasticity index (E/H)·(ujp}!; and pis the
normalised pressure, pI H.
The conclusions to be drawn from the data in Table 3.1 are similar to
those of Greenwood and Williamson ( 1966) - i.e. for all realistic loads the
deformation of surfaces is elastic when ljJ is less than 0.6, and at least partially
plastic when ljJ is significantly greater than unity. However, the additional
76 TRIBOLOGY
Table 3.1
p= w-4 p=5·10- 4 p=l0-3 p=5·10- 3 p=l0-2
1{1 R 1{1 R 1{1 R 1{1 R 1{1 R
2 0 2.3 5.3 5 200
1.3 0 1.5 2 5 500
1.2 0.5 1.3 3.6 1.8 13
1.1 0 1 0 1.2 3
1 2
0.6 0
The delamination theory of wear was first put forward by Suh ( 1973) and
has since been elaborated by Suh (1977) and others (Jahanmir et al., 1974).
The theory involves detailed consideration of subsurface dislocation interactions,
and readers interested in the details should consult the original papers. For
the purposes of this text it is sufficient to set out the physical basis of the
theory, which may be summarised as follows (Barwell, 1983).
( 1) When two sliding surfaces interact, the asperities on the softer surface
are flattened by repeated plastic loading, until the sliding condition
corresponds to that of a harder, rough surface sliding against an
approximately plane softer surface. Thus, at each point on the softer
surface there is repeated subsurface cyclic shear loading, as already
described in discussing fatigue wear.
(2) As this cyclic loading continues, voids and cracks are nucleated below
the surface, crack nucleation at the surface being prevented by the triaxial
compressive stress which exists immediately below the surface.
(3) Further cycling causes these voids and cracks to link up, to form long
cracks at an approximately constant distance below the surface.
(4) When the cracks reach some critical length, the stress situation at the
crack tip causes the crack to break through to the free surface, resulting
in the formation of the thin plate-like wear particles which are often
observed in practice.
Clearly, the highest wear coefficients are found under conditions of abrasive
wear and, under these conditions, the value of K would be unity if every
hard asperity removed a chip of material on each contact with the opposing
78 TRIBOLOGY
1•1
lbl
Figure 3.2 Effect of asperity shape on the form of abrasive wear (courtesy, Professor
T. H. C. Childs)
WEAR 79
lei
surface. In sub-section 3.3.2 we simply stated the fact that abrasive wear
coefficients are normally much lower than this, because much of the displaced
material is simply piled up at the sides of the groove and not lost to the
surface. In a recent paper Childs ( 1988) has presented a much more realistic
picture of abrasive wear, and this is summarised below.
Studies of metals scratched by model hard asperities generally show
that the surface damage is of three types, illustrated in Figure 3.2. The blunter
asperities allow the softer metal to flow round them as a wave, as shown in
Figure 3.2(a); this does not cause any removal of metal on the first pass,
although, as we shall see later, metal may be removed on subsequent passes.
The sharper asperities cause chips to be machined from the softer surface,
as shown in Figure 3.2(b ). Asperities of intermediate shapes cause a wedge
or prow of deformed metal to form ahead of them at the start of sliding, as
shown in Figure 3.2( c). This wedge effectively blunts the asperity, so that
further metal may flow round the wedge without removal (cf. Figure 3.2a ),
or it may intermittently displace the wedge, to form a new one. Childs mapped
the form of material displacement in terms of two variables: the surface slope,
(),and the interfacial shear strength parameter, r:./k, where r:, is the interfacial
shear strength and k is the plastic yield stress in shear of the softer material.
(Note that r:,/ k is identical with the parameter c defined in sub-section 2.3.3,
but in this discussion we shall use the form chosen by Childs.)
Childs cites the work of Challen and Oxley ( 1979), who have developed
plane strain slip line field models of wave-forming and cutting flows in friction
80 TRIBOLOGY
and wear. They showed that, for a wedge-shaped asperity of slope (), wave
flow should occur when
()::::;; 0.5 cos - 1 ( •.! k)
while chip formation should occur when
() ~ n/2- 0.5 cos - 1 ( r./ k)
The limiting conditions represented by the equalities in these equations are
mapped as the lines AB and BC, respectively, in Figure 3.3. Within the region
ABC no steady flow is possible and this is the region of initial wedge formation
as illustrated in Figure 3.2( c).
Further work by Challen and Oxley ( 1984) has shown that the situation
is far more complicated than that illustrated in Figure 3.3. However,
discussion of this work would not be appropriate here, and readers requiring
more information are referred to the original papers.
Childs expresses the severity of wear by a wear coefficient K in an
equation similar to the Archard equation for adhesive wear - i.e.
V=KWs/H (3.9)
where V is wear volume, W is the normal load, s is the slid distance and H
is the hardness of the softer material. However, in this case the K value
cannot be interpreted as the probability of forming a wear particle, which
would have a limiting value of unity: it is simply a coefficient relating the
volume of material removed per unit slid distance to the load and the hardness,
and can have a value greater than unity.
When metal is removed as a chip, the wear process is described as
e (deg)
Figure 3.3 Map of flow regimes (after Childs, 1988)
WEAR 81
1.0
0.8
0.6
!:!.
k
0.4
0.2
0
10 20 50
9 (deg)
Figure 3.4 Calculated values of abrasive wear coefficient K for various values of
r:sfk and fJ (after Childs, 1988)
V=fKWs/H (3.1 0)
where p is the nominal applied pressure; b defines the height, ba, of the
highest asperity; sis the non-dimensional height, z/ a, of an asperity; his the
non-dimensional separation, d/ a, of the surfaces; and ljJ is the plasticity index.
This criterion was developed for surface contact under normal load only,
but it would also apply to frictional loading with friction coefficients less
than 0.1. The criterion would need to be modified for situations of higher
friction; nevertheless, it still provides a useful insight into the factors which
must be considered in choosing materials to resist wear. In particular, it
stresses the great importance of the plasticity index.
WEAR 83
We know from experiment that wear is proportional to the normal load and
slid distance, and, although we have shown that this is predicted by theories
of adhesion and abrasion, we have not shown that it would be expected if
fatigue were the predominant wear process. Fatigue wear equations predicting
such relationships have been derived by Halling ( 1975) and Kragelsky ( 1965).
However, the following simple argument shows that the observed linear
relationships of wear rate to both normal load and slid distance derive simply
from surface topography and would be expected for any form of mechanical
wear involving asperity contact only.
When two materials having different hardnesses are loaded together, the two
opposed asperities involved in any particular contact must be loaded, and
therefore stressed, to the same level. If the hardnesses are significantly different,
the softer material may deform plastically, but deformation of the harder
material must remain at least nominally elastic. As the load is increased, the
plastically deforming material may work harden; but as long as its current
hardness is lower than that of the opposing surface, the harder asperity will
remain nominally elastic. However, we now know that subsurface plastic
deformation will occur when the pressure at the surface is about half the
hardness. Therefore, a softer material can cause subsurface plastic deformation
in a harder one as long as its hardness, in the fully work-hardened condition,
is more than about halfthe initial hardness of the harder surface. To emphasise
this important point, it can be seen that the harder surface must be at least
twice as hard as the softer if it is to remain fully elastic.
Thus, behaviour during sliding is very dependent on the difference in
hardness between the two materials.
If one material is very much harder than the other, the condition of
stress equality can only be obtained if the hard asperity penetrates the softer
surface, as in a hardness test. Relative sliding then requires the hard asperity
to plough a groove in the opposing softer surface. Hence, the softer surface
is subjected to abrasive wear. However, it must be remembered that if the
hardness ratio is less than about 2, the abrasive surface may still be subjected
to subsurface plastic deformation and fatigue wear.
If the two hardnesses are such that both surfaces can remain nominally
elastic (a condition that often obtains after running in), it is probable that
each asperity of any contacting pair will be subjected to subsurface plasticity
and, hence, fatigue. Indeed, both will be subjected to identical cycles offatigue
stress, and repeated contacts will cause fatigue failures at both surfaces.
However, the harder asperity will require more stress cycles to produce a
wear particle, and will therefore wear at a lower rate than the softer one.
This is illustrated schematically in Figure 3.5, which shows the S-log N fatigue
curves (stress-log number of cycles to failure) for the two materials. For the
example given, a horizontal line through the stress S 0 shows that the numbers
of cycles to failure for the harder and softer materials at the stress S0 would
be in the ratio N HI N s and the relative wear rates would thus be in the ratio
Nsf NH (remembering that N is shown on a log scale). This oversimplifies
the true situation, where the stress cycles would be of varying amplitude, but
it clearly illustrates the principle involved.
Thus, fatigue wear explains the observation that a harder material can
be worn by a softer one, a phenomenon which has never been explained
satisfactorily by adhesion theory.
By invoking fatigue as the basic wear process we can also explain the
formation of loose wear debris.
WEAR 85
Ns
log N
Figure 3.5 Illustrating simultaneous fatigue ofboth hard and soft rubbing surfaces
0
E'
logW
The rate of oxidation and, hence, the film thickness attained between
successive asperity collisions are very dependent on the local temperature,
and this temperature, in turn, depends on the input of frictional work and
thus on the load. Therefore, all other things being equal, the rate of oxidation
increases approximately exponentially with the load, while the time between
collisions decreases only linearly. Thus, at the second transition load, T2 , the
temperature effect becomes predominant and the wear process returns to
one of 'mild' oxidative wear.
Steel surfaces examined after rubbing above T2 also have a hardened
surface layer, due to the repetitive rapid cooling from high flash temperatures.
This hardened layer also favours the mild wear regime, owing to the reduced
incidence of plastic deformation and the improved support given to the oxide
film. This effect of hardness is illustrated more dramatically by rubbing
steel-steel pairs which have been previously hardened and tempered to
different bulk hardnesses. As the initial hardness is increased, it is observed
that the severe wear regime becomes narrower until, at some critical value
of hardness, it disappears entirely and mild wear persists over the full load
range.
In the above discussion we have considered only the effect of increasing
the load. However, it is clear that similar effects would be seen as a result
of increasing the speed. Increasing speed would both increase flash temperatures,
owing to the increased input of frictional work, and decrease the time available
for reoxidation between collisions.
It has been pointed out by Endo and Fukuda ( 1965) that the presence
of oxygen at crack tips will tend to increase the rate of fatigue crack growth
and, thus, the fatigue wear. This is a further example of the interactions
between, and complementary effects of, wear processes which are often
described as though they are quite independent.
Finally, Barwell (1983) has pointed out that the adjectives 'mild' and
'severe' can be very misleading, in that the rate of material loss in a 'mild'
wear regime can be substantially higher than that in a 'severe' wear regime.
The terms 'mild' and 'severe' apply to the form of surface damage rather
than the rate of material loss, and Barwell suggests that the term 'smooth
sliding wear' should be substituted for 'mild wear'.
As mentioned earlier, the above discussion also applies to corrosive
media other than oxygen. For example, we have already described, in
Section 2.7, the use of extreme-pressure additives to deliberately generate
significant corrosive wear in preference to the much more damaging severe
wear which would occur in their absence.
Recently Lim and Ashby ( 1987) have suggested that the various regimes of
mechanical and corrosive wear for any particular pair of rubbing materials
90 TRIBOLOGY
Seizure
10-1
J!
H
V (m/s)
Figure 3.7 Wear map for soft carbon steels (after Lim and Ashby, 1987, and
Childs, 1988)
3.7.1 Fretting
3.7.2 Erosion
Fluid Erosion
There are two basic types of fluid erosion - i.e. liquid impact erosion and
cavitation erosion. The current state of understanding of these processes has
recently been comprehensively reviewed in excellent papers by Hammitt
( 1980) and Schmitt ( 1980 ), and it is the purpose of this sub-section
to give only an outline of the more salient features of the processes.
The processes of removal of material are rather less well understood
than are the processes occurring in sliding wear, but many studies have been
made of the damage which occurs. It appears that there are different damage
processes for brittle solids, ductile metals and thermoplastic polymers,
thermosetting polymers and elastomers.
Liquid impact on a brittle material generates momentary stresses which
are sufficiently high to cause cracking in the form of initially unconnected
annular ring segments. With continued impact, further cracking occurs and
the cracks eventually join, so that material is removed in the form of chips.
The erosion damage of ductile materials, such as many metals and
thermoplastic polymers, initially takes the form of surface depressions with
92 TRIBOLOGY
upraised edges. These edges are then subjected to high stresses during the
radial flow of subsequent impacting drops, until they eventually fail, with
consequent surface roughening. The surface roughening itself then changes
the flow patterns of impacting liquid, with a possible exacerbation of the
erosion.
The erosion behaviour of thermosetting polymers is similar to that
already described for brittle materials, with the coalescence of ring cracks
leading to the formation of relatively large wear particles.
Elastomers are often used in the form of surface coatings to reduce
erosion damage. Some of these coatings fail by erosive wear which is
phenomenologically similar to that described for thermosetting plastics, while
others remain superficially almost undamaged but suffer loss of adhesion to
their substrates.
Cavitation damage occurs when bubbles entrained in a liquid become
unstable and implode against the surface of a solid. This is thought to produce
a combination of shock waves in the liquid and liquid microjet impact on
the solid surface. Cavitation damage closely resembles damage by liquid
impact erosion, but it is on a finer scale. It can be a very important and
damaging form of attack on solids which are moving in a liquid, examples
being ships' propellers and pump impellers.
Solid Erosion
'incubation' period during which no loose debris is formed, and this suggests
that erosive damage of ductile materials is a cumulative process, with fatigue
again being a likely cause.
It should now be clear that the wear resistance of any surface is dependent
on both the hardness and the elastic modulus of the contacting surface. We
have already shown in Section 1.5 that a relatively thin surface film can
modify both the hardness and the modulus of a surface, and thus have a
very significant effect on its tribological properties. This has recently been
treated in detail by Halling ( 1986), who shows that the critical factors in
determining the tribological properties of any substrate-coating-counterface
combination are the mechanical properties of the three materials, the
roughnesses of the substrate and counterface, and the thickness of the surface
film. As both film thickness and surface roughness are very likely to change
with continued rubbing, there is unlikely to be a quantitative method of
predicting the long-term wear behaviour of coated surfaces in the near future.
However, the principles established by Halling are very useful and can guide
the selection of coated materials for tribological applications. We shall return
to this topic in Chapter 4.
94 TRIBOLOGY
Fluid lubricants have three important effects in reducing both friction and
wear of sliding surfaces.
( 1) They provide a film of low shear strength between the surfaces, thus
mitigating the effects of adhesion. This effect is particularly marked if
boundary lubricants or extreme-pressure additives are present in the
lubricant.
(2) They cool the surfaces, both by reducing the input of frictional energy
and by removing heat, thus reducing such deleterious thermal effects as
surface softening.
(3) In recirculating systems they carry the products of wear away from the
interface, thus preventing three-body abrasion of the surface by work-
hardened or oxidised debris.
Thus, wear coefficients are almost invariably reduced if the surfaces can
be continuously lubricated with a suitable oil. However, the prediction of
wear rates, if anything, is made more difficult than in dry rubbing. This is
because the oil film will normally protect some unknown fraction of the
surfaces, with the remaining fraction, comprising the higher asperities, making
solid contact through the film. Thus, the effective wear coefficient is the
resultant of one coefficient on the unprotected fraction of the surface, and
of a second coefficient on the protected fraction.
As a result of this additional complexity, hardly any wear coefficients
have been published for conditions of lubricated sliding. Recently Rowe
(1980) has suggested that wear coefficients can remain approximately
constant under some sliding conditions, and has made a plea for the
publication of more measured coefficients, with precise descriptions of the
conditions under which they were measured. However, it is likely that wear
prediction in lubricated sliding will rely on realistic laboratory tests and tests
under field conditions for the foreseeable future.
One particularly damaging type of wear under lubricated conditions is
scuffing, a form of wear in which localised solid-phase welding occurs between
sliding surfaces, resulting in damage and roughening of the surfaces. The
original area of scuffing is almost invariably roughened to such an extent
that it is likely to penetrate the oil film on subsequent contacts. Some scuffed
areas are self-healing - i.e. they undergo a period of mild wear which results
in smoothing of the roughened surface. However, in other cases the problem
becomes rather worse on each contact, with both the severity and the extent
of roughening increasing. In this latter case the damage can lead very rapidly
to gross damage and fracture of the sliding components. Thus, it is
characteristic of scuffing failure that it occurs very rapidly: mechanisms which
have been running very smoothly can suddenly become noisy, and failure
often ensues within minutes.
WEAR 95
3.10 PROBLEMS
4.1 METALS
Much of the material in the previous chapters has been concerned with the
tribological properties of metals, and it is not necessary to reiterate it here.
In particular, the effect of hardness on resistance to wear was mentioned
frequently in Chapter 3 and the importance of oxide films was stressed in
Section 3.5. However, in covering the basic mechanisms of wear some other
properties which can have a marked effect on wear behaviour were mentioned
only in passing, or not at all. The effects of these properties are described in
the following sections.
96
TRIBOLOGICAL PROPERTIES OF SOLID MATERIALS 97
1·6
1·2
g
c
£
0
c!!. 0·8 Wear rate
~ 3·5xi0- 7 mm 3/m
8
u
0·4
Figure 4.1 Friction coefficient and wear rate for cobalt on cobalt in vacuum at
various temperatures: pressure, 10- 7 N/m 2 ; sliding speed, 2 mjs; load, 9.81 N
in Figure 4.1 the ambient temperature at the transition is less than the quoted
temperature of 417 oc; this is because frictional heating causes the actual
surface temperature to be higher than the ambient temperature.) Fortunately,
this allotropic transformation temperature can be increased by additions of
suitable alloying elements. For this reason many wear-resistant materials are
based on cobalt-30% chromium, which retains the CPH structure up to a
temperature of > 800 oc.
Rabinowicz ( 1970) has suggested that pairs of materials with a high degree
of mutual solubility will have poor tribological properties, as a consequence
of high interfacial adhesion resulting from their mutual interdiffusion and
bonding. He attempted to correlate mutual solubility, as indicated by binary
alloy equilibrium diagrams, with adhesion, friction and wear. The analysis
showed zero correlation with adhesion; some positive correlation with
friction; and a greater correlation with wear. This suggests that pairs of
metals for boundary lubricated or unlubricated sliding should be chosen to
have low mutual solubilities. This is a generally accepted rule of thumb but,
as with most tribological rules, there are some notable exceptions. For
example, the rule suggests that the same material should not be specified for
each component of a rubbing pair, but some materials, particularly the
cobalt-based superalloys mentioned above, have excellent tribological
properties against themselves. Recently Tabor ( 1985), Buckley ( 1985) and
Ferrante et al. ( 1985) have shown that the situation is much more complex
100 TRIBOLOGY
We use the term 'bearing alloys' to describe alloys developed for use in
well-lubricated journal bearings. As will be shown in Chapter 6, under normal
running the bearing and its counterface- often a crankshaft- are completely
separated by a thin oil film which is generated by hydrodynamic lift. Under
such conditions the tribological properties of the solid material are, of course,
irrelevant. However, the oil film is not present for a short time after each
start of the engine, or when the engine is running slowly at high loads. Under
these circumstances the surfaces are protected only by boundary lubrication,
as described in Chapter 2, and as boundary lubrication involves some
solid-solid contact and high asperity stresses, the tribological properties of
the materials are then very important.
When we use bearing alloys, we are not usually in the fortunate position of
being simply concerned with two chosen materials rubbing together in the
presence of a film of clean oil containing boundary lubricants. For example,
the inside of an internal combustion engine is a very hostile environment,
containing corrosive combustion products, abrasive wear debris and usually
-particularly in new engines- cast-iron dust left behind after machining the
TRIBOLOGICAL PROPERTIES OF SOLID MATERIALS 101
engine block, and wear debris from the piston, cylinder and piston rings
formed during running in. The rubbing materials must, therefore, be able to
withstand most of the wear processes described in Chapter 3. Furthermore,
the bearing must sustain very high fluctuating dynamic loads, which are
inherent in the operation of an engine, without suffering fatigue failure. These
various, and often conflicting, requirements have led to the development of
ranges of bearings which often contain several components. The compositions,
properties and methods of manufacture of the various types of alloy cannot
be described in detail here and the reader requiring more detail should see,
for example, the excellent review by Morris ( 1967) or The Tribology Handbook
(Neale, 1973). We shall simply describe the general requirements of such
bearings and briefly mention some examples. The required characteristics
are described in turn below.
Embeddability
Conformability
So far, all our requirements have led to the choice of a soft material for the
bearing shell and, at first sight, this conflicts with the need for the bearing
to carry very high fluctuating loads without gross plastic deformation or
fatigue failure. However, we can have all the properties listed above and good
mechanical strength by making the soft metal very thin. The reason for this
is that when a very thin film of soft metal is supported on a hard base, it
can carry very high normal pressures without yielding. For example, El-Shafei
et al. ( 1983) have shown that thin lead films on hard steel substrates can
sustain the same pressure as the uncoated steel before yielding. This is because
the film is trapped between two extensive hard surfaces and it can deform
plastically only by flowing outwards so that some material leaves the contact
zone. As the film becomes thinner, the pressure required to cause this outward
flow increases very rapidly, and tends to infinity at zero film thickness. This
effect of plastic constraint is used in bearing design by supporting a thin and
soft bearing material on a relatively strong and hard steel backing strip.
The two metals which combine best the required properties of softness, low
friction against steel and low melting point are tin and lead. However, in
order for these materials in the unalloyed form to have the required
load-carrying capacity they would need to be no more than a few microns
thick, and this would both limit the lifetime and restrict the embeddability
and conformability of the bearing. These metals are therefore universally
used in two types of alloy: first, as lead-based or tin-based alloys, in which
properties such as hardness and strength are enhanced considerably by small
alloying additions; second, as dispersions in a matrix of a stronger material.
The two most common materials of this second type are copper-lead and
aluminium-tin alloys; in each of these the zones of the softer material
interconnect, so that any lead or tin lost from the rubbing surface can be
continuously replenished from the underlying material. Typical micro-
structures of alloys of such materials are shown in Figure 4.2.
The copper-lead and aluminium-tin alloys are stronger and harder
than those based principally on lead and tin. However, the lead component
TRIBOLOGICAL PROPERTIES OF SOLID MATERIALS 103
There is no doubt that the ideal way to reduce both friction and wear of
sliding surfaces is to arrange for them to be completely separated by a fluid
film, and the various ways in which this can be done will be described in the
following chapters. However, even the best fluid lubricants have limitations
which preclude their use under certain circumstances. Their main limitations
are described below.
The basic requirement of any solid lubricant is exactly the same as that for
a liquid lubricant. It should completely separate two solid surfaces, even
under a very high normal load, while allowing the surfaces to slide over each
other under a relatively low shearing force: in other words, a solid lubricant
must meet the twin requirements of having a high normal load capacity and
a low shear strength.
We have already described one type of solid lubricant in sub-section
4.2.2, where we showed that these twin requirements can be met by a thin
film of a soft metal supported on a relatively hard substrate. However, this
type of solid lubrication is a property of the multicomponent system, rather
than an inherent property of a single material, and in this section we intend
to describe single materials which can act as solid lubricants.
It is clear that the twin requirements described above are unlikely to be
met by any isotropic material, in which the shear strength will be about 16%
of the yield pressure, and it has been found that the best solid lubricants are
very anisotropic materials known as lamellar solids.
A lamellar solid is one in which the atoms are bonded together in parallel
and comparatively widely spaced sheets, the two best-known being graphite
and molybdenum disulphide, which have the crystal structures shown in
Figures 4.3 and 4.4, respectively. Under many circumstances these are
excellent lubricants, as are other less well-known lamellar solids such as
tungsten diselenide and cadmium iodide. However, not all lamellar solids
are good solid lubricants, and there is as yet no theory which will tell us
whether or not any particular lamellar solid will be an effective lubricant.
The lamellar solids which can lubricate effectively do have certain
characteristics in common. First, they form a strongly adherent transfer film
on the surface being lubricated, so that, after the short running-in period
during which this film is formed, the actual interface is between lubricant
and lubricant. Second, the lubricant on both surfaces develops the preferred
orientation shown schematically in Figure 4.5, and this reduces the mechanical
106 TRIBOLOGY
-tf~~?1,
~~~,,r ll
f
I
tf
I I I I I I I
I I I I 1 I
~. ~Jff-:;t.
ljl
I1 1I I
I
I
1
I I
11 I
I
I I
I
I I I . I I I
~r?tit 7:J
Figure 4.3 Crystal structure of graphite
e Molybdenum
0 Sulphur
Figure 4.5 Schematic illustration of the surface orientation developed during the
rubbing of lamellar solid lubricants
Solid lubricants can be used in several different ways, the most common ones
being described below.
Solid Blocks
Graphite and graphitic carbons are often bonded together in solid blocks
which can, for example, be made into solid thrust bearings and journal
bearings. The blocks are usually made from a mixture of finely divided coke
and a carbonaceous binder such as pitch. This mixture is fired to a very high
temperature, to cause it to graphitise, and by varying the heat treatment a
wide range of materials can be made. These materials range from highly
crystalline electrographite, which is generally used for low-load applications
such as electrical generator brushes, to almost amorphous mechanical
carbons, which are used for bearings.
Resin-bonded Films
z.o
ME 1·5
E.
1!
g
~
ill
~
...~
10 15 20 25
Weight % MoS 2
Metal-based Composites
Polymer-based bearings can have all the advantages of solid lubricants listed
in the introduction to Section 4.3, plus the following additional advantages.
( 1) They absorb vibrations well and are quiet in operation.
(2) They readily deform to conform to mating parts, so that machining
tolerances and accuracy of alignment are usually less critical than for
metal parts.
110 TRIBOLOGY
(3) They are easily formed into complex shapes, by either machining or
moulding.
( 4) They are cheap.
There is a very wide range of commercially available polymer-based
bearing materials, and they cannot be described in detail here. The aim of
this section is to describe in principle the tribological properties of polymer-
based materials and the various operating factors which can influence these
properties.
There are only five basic polymers in common use in tribology: PTFE,
polyacetal, Nylon, polyethylene and polyimide, although other materials are
TRIBOLOGICAL PROPERTIES OF SOLID MATERIALS 111
now coming into use. However, many reinforcements and fillers have been
incorporated into these basic materials, so that a wide range of bearings is
commercially available.
As their name implies, the purpose of reinforcements is to increase the
mechanical strength of the bearing. Reinforcement is carried out in one of
three basic ways. First, a mechanically strong, normally fibrous material is
mixed with the polymer before moulding, the most common example being
chopped glass fibre, which increases strength, stiffness and creep resistance.
Second, the polymer is incorporated into a woven matrix; with thermosetting
resins the woven matrix is soaked in the resin, which is then allowed to
harden, whereas with thermoplastic poly~ers, such as PTFE, the polymer
fibre and reinforcing fibre are woven together in such a way that the actual
bearing surface is predominantly the lubricating polymer. Finally, the polymer
can be supported as a relatively thin layer on a metal backing plate, to give
the plastic constraint and increased load-bearing capacity already described
for bearing alloys in sub-section 4.2.2. If the polymer is also supported around
its edges, so that it is effectively held in a metal container, its load-carrying
capacity can be increased enormously. Pure PTFE has very poor creep
resistance and, if unconstrained, will creep under very low loads, but if fully
constrained, as described above, it can be used to carry very high loads -
for example, in civil engineering applications such as bridge bearings.
A metal liner on a thin polymer bearing has the additional advantage
of greatly assisting heat dissipation from the sliding interface. In general,
polymers have low thermal conductivities and high thermal expansion
coefficients when compared with metals, and, therefore, if solid bushes of
pure polymer are used against steel journals, frictional heating can cause
bearing seizure. To avoid this, it would be necessary to make the polymer
bearing to a much looser fit than the designer would otherwise wish. However,
the problems of low conductivity and high expansion coefficient can both
be avoided by using the polymer as a thin film on a metal liner.
Fillers, which are particulate materials, are used to improve strength,
thermal properties and frictional behaviour. Metal powders of high conductivity
and relatively high strength are used to improve thermal and mechanical
properties, while graphite, molybdenum disulphide and soft metals such as
lead are used to improve frictional behaviour.
The rate of wear of a polymer will obviously be some function of the load
and sliding speed, and it is found that we can define for any polymer bearing
a design criterion known as the PV factor.
The theoretical derivation of the relationship between the wear rate and
the PV factor starts from the reasonable assumption that the rate of wear
will be proportional to the rate of energy dissipation at the sliding interface.
On this basis, we can then derive a relationship between the rate of wear
112 TRIBOLOGY
-
Load
Shdtng veiOCtly V
Beanng
, , ..
area A
, ,,
,,
Figure 4. 7 Schematic diagram of flat bearing for derivation of P V factor
Shaft
Load supported
on shaded hoff
!al of bearor>g
(b)
Figure 4.8 Schematic diagram of solid lubricated journal bearing and forces acting
Therefore R = wI Dl. But the rate of energy dissipation per unit area of
interface is p.PV, so that the total rate of energy dissipation is given by
p.RV·nD/12
= (p. WI Dl)· V·( nDl/2)
= p.WVn/2
Therefore, the volume wear rate, Q, is given by Q = kWVn/2, where, as
before, p. has again been subsumed into the constant.
114 TRIBOLOGY
But the radial wear rate, Qd, is again given by the volume wear rate
divided by the contact area - i.e.
Qd = Q/(nDl/2) = kWV I Dl
so that Qd = kPV ( cf. Equation 4.2 ), where Pis the load per unit projected
area.
Thus, we can see that we expect the same proportionality between linear
wear rate and the PV factor for both types of bearing, provided that P is
taken as load per unit projected area, where the constant of proportionality,
k, is the specific wear rate of the material.
Equation (4.2) suggests that the rate of wear of a polymer bearing should
be proportional to the product PV, and to a good approximation it is, indeed,
often possible to specify a PV value which must not be exceeded, to give a
certain wear rate or life expectancy. Also, for any material there is a limiting
PV factor above which the material would fall very rapidly, owing to melting
or thermal decomposition, and this is known as the PV limit of the material.
The use of single PV factors is open to some criticism, since it assumes
the same sensitivity of wear rate to changes of P and V. This assumption is
frequently unfounded, and a more exact practice is to give plots of acceptable
wear rates on PV diagrams, as shown in Figure 4.9. This diagram shows
that the PV factor for the chosen wear rate is almost constant, except at the
extremes of load and speed. This is often the case, although some materials
have PV diagrams which are continuous curves over the whole PV range.
It is also found that there is a limited range of low pressures and
temperatures over which k can be taken to be constant, and within this
region k is given the symbol k 0 . Thus, within its range of applicability,
Equation (4.2) can be used to predict wear rates; it is only necessary to
measure the wear rate at one value of PV to determine the value of k 0 , and
this value can then be used to predict wear under other PV conditions.
At higher temperatures or pressures the value of k increases, and for
accurate data recourse must be had to detailed design manuals. In particular,
the ESDU Guide to the Design and Material Selection for Dry Rubbing Bearings
( 1987) should be regarded as an indispensable guide to any designer involved
in the specification of such components. This guide gives not only PV data
of the type described above, but also detailed information on the effects of
other variables such as the friction coefficient, which we initially assumed to
be constant, the counterface composition and roughness, the presence of
lubricants, and the thermal properties of the bearing assembly.
The PV factors or PV curves which are published or supplied by bearing
manufacturers normally refer to unlubricated sliding against a counterface
of a specified material having a specified surface finish. When the sliding
conditions depart from these standard conditions, the designer must take
into account such changes when specifying a bearing, and to do this it is
essential that a more detailed guide, such as the ESDU guide, be consulted.
TRIBOLOGICAL PROPERTIES OF SOLID MATERIALS 115
Velocity lm/s)
Figure 4.9 Typical limiting PV curve of PTFE-based material for wear rates of
25~tmin 100 h
The materials used in the manufacture of brakes and clutches are known as
friction materials. Braking is usually by far the more demanding of these
applications, and in what follows we shall concentrate on brakes, although
the same principles will also apply to clutches.
In contrast to all other tribological applications, high friction coefficients
are desirable when using friction materials, and the materials are normally
developed to give friction coefficients, against the chosen counterface, of the
order of 0.35-0.55. The principal frictional requirement is that the friction
coefficient should be stable with respect to sliding speed, time and temperature,
to avoid the brakes fading during a prolonged application, but a high friction
coefficient ensures that the required braking force can be obtained under a
relatively low normal load.
116 TRIBOLOGY
The present state of knowledge on the friction and adhesion of ceramics has
already been discussed in sub-section 2.6.3, where it was made clear that, as
yet, little is known of the nature of interatomic bonding across ceramic-metal
and ceramic-ceramic interfaces. This is because, particularly with covalently
bonded ceramics, it is unclear how many valences are fully saturated and
how many are free to form interfacial bonds. However, it has been observed
that friction coefficients for ceramic-ceramic sliding are not particularly low,
being of the same order as those between metals in normal atmospheres.
On the other hand, it is well known that ceramics can exhibit very low
wear rates, compared with metals, and this property is causing ceramics to
be used increasingly in tribological situations. Furthermore, ceramics are
refractory materials which retain qualitatively similar properties up to high
temperatures, and they are also chemically very unreactive, so there is no
118 TRIBOLOGY
doubt that they will find increasing use in hostile environments, particularly
those involving high temperatures.
We shall discuss specific examples of the uses of ceramics a little later
in this section, but before doing that it is useful to discuss the reasons for
the good wear behaviour. This is best done by referring again to the plasticity
index, t/t, where
t/t=(E*IH)·(ulfJ)! ( 1.28)
As we stated in sub-section 3.4.2, surface contact between metals will be
almost entirely elastic when t/t < 0.6 and almost entirely plastic when t/t > 1.0.
However, while ceramics have similar elastic moduli (E) to those of metals,
they have very much higher hardness values (H). Thus, for similar surface
finishes (represented by u and /3) ceramics are much more likely than metals
to be subjected to fully elastic surface contact.
To clarify this point, it is useful to consider the ratio HIE from a different
standpoint. As H is a simple multiple of the yield stress of a material, this
ratio is a measure of the yield strain of the material - i.e. the strain which
can be accommodated by fully elastic deformation. Thus, materials with a
high HIE ratio, such as ceramics, can accommodate much higher deformations
elastically than can those with lower HIE ratios, such as metals; this is
obviously critically important in determining the mode of deformation as
asperities interact.
There is another crucial difference between ceramics and metals: ceramics
are brittle, whereas most metals are ductile. Thus, when most metals reach
the limit of elastic deformation, they can accommodate additional deformation
plastically, whereas ceramics are likely to fracture. It is true that ceramics
are capable of undergoing some plastic deformation when there are local
hydrostatic stresses, and such conditions often obtain at regions of asperity
contact. Nevertheless, if a ceramic asperity is subjected to an increasing load1
it is likely to fracture without significant plastic deformation. The higher
asperities on a sliding ceramic surface will, of course, be subjected to the
highest forces, and if the resulting wear debris can escape from the interface,
the surfaces should therefore become progressively smoother, the surface
interactions should become more elastic and the wear rate should decrease.
Despite the elastic deformation processes described above, the wear rates
of ceramics are never zero and we must ask what causes the residual wear.
The first point to note is that we must consider the situation in terms of
fracture mechanics. Ceramics are not homogeneous materials and they
invariably contain surface cracks and internal flaws. Any flaw above a certain
critical size is able to extend if the stress exceeds a critical level, which may
be well below the theoretical fracture stress of the material, and thus a short
stress cycle may cause a small incremental crack extension. On repeated
stress cycling, such incremental extensions can cause the flaw to extend to
the point where a loose wear particle is formed. Furthermore, this process
is likely to be exacerbated by the presence of a corrosive environment, which
TRIBOLOGICAL PROPERTIES OF SOLID MATERIALS 119
causes more interatomic bonds at the tip of the flaw to be broken on each
stress cycle.
The propagation of a crack as described above may be responsible not
only for the creation of local fractures, but also for large-scale fracture of
the bulk material, and it is this which has, until now, restricted the use of
bulk ceramics in tribology. However, the situation is now being improved
in three ways.
First, ceramics technology is improving rapidly, so that materials which
are more homogeneous and have fewer and smaller flaws are now being
produced. Such bulk materials will be used increasingly in future and are
already being actively investigated for various applications. For example,
manufacturers of automobile engines are aiming to design all-silicon nitride
engines in which the engine block, pistons and connecting rods will be made
from bulk silicon nitride. If a suitable ceramic with the necessary mechanical
properties can be developed, the resulting benefits will be: excellent wear
resistance at the piston-cylinder interface; the possibility of a higher operating
temperature and therefore higher efficiency; and lower inertial forces, owing
to the low density of the ceramic.
Second, problems of bulk fracture are being avoided by the use of cermets
- i.e. materials which consist of mixtures of metal and ceramic. An example
of this is metal-bonded tungsten carbide- a mixture which is approximately
90% tungsten carbide with a metal binder. The ceramic grains are coated
with the metal, usually cobalt or cobalt-chromium alloy, and are then
sintered, to form a dense solid. The metal at the grain boundaries then
prevents any crack being propagated across a ceramic-ceramic grain
boundary and leads to the composite material being much tougher than a
pure ceramic. Such materials are already used extensively as, for example,
sealing rings in mechanical seals.
Third, methods have been developed in recent years for depositing
ceramic coatings onto metal surfaces; systems of this type will be described
in more detail in Section 4.7.
In closing this section, we should note that, even when the stresses are
very low and the flaws are unimportant, ceramics, in common with other
materials, will still wear at finite rates. It is thought that this residual wear
probably takes place on an atom-by-atom basis, as a combination of
mechanical, chemical and thermal energies makes it energetically favourable
for an atom to leave the surface.
they have the required beneficial effect, which may then last the life of the
component. The predominant film of this kind is formed by 'phosphating'.
Many commercial processes exist for forming layers of metal phosphate
crystals on the surfaces of metal, particularly steel and cast-iron, and these
films give good corrosion resistance, good short-term wear resistance and
good oil retention. The oil retention is particularly good on grey cast-iron
surfaces, as the phosphate film forms only on the metal parts of the surface
and leaves oil-retaining channels above the graphite flakes.
A second short-term film in very common use is the very thin film of
electroplated tin which is used on aluminium pistons to assist running in.
There are two basic ways of changing the composition of the surface layers
of a material: by diffusion or implantation.
In diffusion treatments the component is held at high temperature in a
specified environment, so that atoms from the environment diffuse into the
122 TRIBOLOGY
outer layers of the component. The principal atoms which are used to modify
the surfaces of steels are carbon and nitrogen.
Carburisation involves holding low-carbon steel in a carbon-rich
environment at high temperature. Carbon from the environment then diffuses
into the outer layer and transforms it into high-carbon steel, which can then
be heat treated to improve its hardness. Nitriding involves the diffusion of
nitrogen into a steel surface to form small crystals of metal nitrides, and
nitriding steels usually contain alloying elements which react readily with
nitrogen. The nitrogen can be diffused in from gaseous atmospheres, fused
salts or argon-nitrogen plasmas (Bell et al., 1983 ), and nitriding temperatures
are usually much lower than those used for carburising. After nitriding, the
component often does not require any further heat treatment. Cyaniding and
carbonitriding cause bpth carbon and nitrogen to diffuse into the surface.
Cyaniding is a salt-bath process carried out at a nitriding temperature,
whereas carbonitriding is a gaseous process carried out at higher temperature.
Many commercial processes exist under particular trade names, for
carrying out the processes described above and for diffusing in other elements
such as boron, aluminium and chromium. Readers requiring more detail of
these are referred to the review by Wilson ( 1975).
In ion implantation processes, ions with very high energies, of the order
of MeV, are fired into the component surface and, as a result of their high
energies, penetrate the surface up to distances of ~ 1 JJ.m. Very high
concentrations of the implanted ions can be created in this way and the
resulting surfaces often have excellent wear resistance (Dearnaley and Grant,
1985). One beneficial, and as yet unexplained, effect is that the improved
wear resistance often persists up to depths well below the original depth of
implantation.
also to deposit the lead alloy overlays on bearing alloys. Electroless deposition
can be used to deposit a range of hard and wear-resistant nickel-based
coatings. Both plasma spraying and detonation-gun plating can be used to
deposit a wide range of ceramic and cermet coatings, which are finding
extensive use as wear- and corrosion-resistant materials in hostile environments
such as nuclear reactors. Both types- of coating are slightly porous, the
plasma-sprayed coatings being rather worse in this respect, and this can
allow some aggressive corrosive species to penetrate the coating and cause
it to separate from its substrate.
Finally, coatings only a few micrometres thick are being increasingly
deposited by physical vapour deposition (Teer and Arnell, 1985) and chemical
vapour deposition (Hitchman, 1985). Both these techniques can allow the
deposition of fully dense and highly adherent coatings of a wide range of
compositions, including metastable compositions which cannot be formed
by any other means. Such coatings are already being spectacularly successful
in a variety of applications, which include: the coating of cutting tools with
titanium nitride to give a fivefold increase in life; coating of the races of
satellite bearings to give a coating ~0.5 .urn thick which provides soft metal
film lubrication for the full life of the satellite; and coating of other satellite
bearings with highly adherent films of molybdenum disulphide.
4.8 PROBLEMS
No problems have been set for this chapter. This is because exhaustive design
procedures for dry rubbing bearings have been set out recently in ESDU
Item No. 87007: Design and Material Selection for Dry Rubbing Bearings.
For commercial reasons neither the data nor the procedures can be
reproduced here, and under these circumstances it is not possible to set
meaningful design problems. However, the reader is strongly recommended
to study the ESDU unit and to work through the example design problems
therein.
Chapter 5
Fluid-lubricated Thrust
Bearings
5.1 INTRODUCTION
124
FLUID-LUBRICATED THRUST BEARINGS 125
Toc-
au or
au
t=l'f- (5.1)
oz oz
where the constant of proportionality, l'f, is known as the coefficient of dynamic
viscosity. The dimensions of this coefficient may be found as '1 = t/(oujoz)
and it follows that t /( ou/ oz) has dimensions of
(F/L 2 ) FT
(L/T)/L =u
In the SI system the coefficient of dynamic viscosity has units of
(newton·second)/(metre) 2 , but as a newton is defined as (kilogram)
(metrejsecond 2 ), the units of dynamic viscosity may also be given in
dimensions of (ML/T 2 )T/L2 = M/LT or (kilogram)/(metre·second). For
example, water at room temperature and atmospheric pressure has a dynamic
viscosity of about 10- 3 Ns/m 2 or 10- 3 kg/m·s.
Another common and important unit of dynamic viscosity is the poise,
which has units of (dyne· second)/( centimetre )2 or (gram)/( centimetre·second)
in the CGS system of units, but a more useful unit is the centipoise (cP).
The relationships between these units is as follows:
dynamic viscosity (kg/m·s)= / 0 dynamic viscosity (P)
= 10100 dynamic viscosity (cP) (5.2)
zL;
Velocity U Area (A)
L,.c;:: :~ . Force F
f 777777777777777777777
Film thicknessh
X
_ F l u i d film
Figure 5.1
126 TRIBOLOGY
The kinematic viscosity of a fluid is simply defined as the ratio between its
dynamic viscosity, 1J, and its density, p. So the kinematic viscosity, v, is simply
v = 11/ p and its dimensions in the SI system are (metre) 2 /(second). The
corresponding unit in the CGS system has units of (centimetre) 2 /(second)
and is known as the stoke, although a more useful unit is the centistoke
(eSt). This latter unit is very important, because most of the internationally
3.0 22
4.0 37
46
5.0
68
7.0 100
9.0
12.5
20
~ 30
40
·~0
&I 75
·;
·~ 150
E
~ 250
c
i2 400
750
1500
3000
10000
50000
200000
-20 0 20 40 60 80 100 120
Temperature (°C)
22 37 46 68 100
200
60 80 100
Temperature (°C)
Consider the flow of a viscous fluid (Massey, 1989) through a wide parallel
rectangular gap of relatively small, but uniform, clearance h under the action
of a pressure drop from p 1 to p 2 over a length L and the action of a moving
surface which has a velocity Uc• as shown in Figure 5.4. The axes x, y and z
are set up as before and it is assumed that the edges at y = 0 and y = B,
corresponding to the width of the gap, are sealed so that the flow of fluid
proceeds only in the direction of the x axis.
Consider the equilibrium of an element of fluid oflength ()x at a distance x
from the origin. Let this element have a thickness (jz and be positioned at
a height of z above the lower surface which moves to the right with a velocity
Uc. The local pressure at x = 0 is p 1 , and at x = L, p2 •
The element of fluid of Figure 5.4 is shown enlarged in Figure 5.5,
together with the pressures and shear stresses which are assumed to act on
it. Pressure p is assumed to increase with x and the shear stress r with z.
In the steady state the sum of the forces acting on the element is zero.
In each case the local pressure, or shear stress, may be multiplied by the
appropriate area, to give the corresponding force. The sum of the forces
acting to the right on the element is thus:
Figure 5.4
T+-·/jz
OT
az j z
~- op
P~j:~j~ p+a;·6X
x 6x
Figure 5.5
FLUID-LUBRICATED THRUST BEARINGS 129
or
or op (5.3)
oz ox
This is the equation of equilibrium for the element of fluid. Development
of the solution now proceeds on the basis that the flow throughout the
gap is laminar, or viscous, so that the relationship between the shear
stress and velocity gradient given in Equation ( 5.1) may be substituted
into Equation (5.3) above, to yield the following expression:
o u op
2
'1 oz 2 =ox
u=(~op)z 2 +C 1 z+C 2
I]OX 2
where C 1 and C 2 are constants to be determined. Now, when z = 0, we know
that u = u c> and when z = h, u = 0.
Thus
c = -(~op)~- uc
1
I]OX 2 h
and C 2 = Uc. These values for the constants may be back-substituted into
the above expression, to give the following solution for the velocity of the
fluid in the gap:
U=- 1(
21]
-- op)
ox (hz-z )+Uc
2 ( 1--
h
z) (5.4)
It is clear that the pressure gradient must be negative if the flow is to proceed
to the right and that the total velocity at any value of z is given by the sum
of the pressure-induced, or Poiseuille, velocity and the shear velocity induced
by the movement of the lower surface.
The total volumetric flow rate of fluid through the gap is given as
Q = J~uBt5z. The expression for u, a function of z, may be substituted from
Equation ( 5.4) into this expression to give
or
In the case of the geometry of Figure 5.4, it is clear that the volumetric flow
rate given by Equation (5.5) does not vary with x, so that
op =
OX -12n(g-
., B 2~u)
c
p= 1211(~- ~·Uc )x + C 3
P2= -12'7(~-~uc)L+pl
In this case the pressure gradient, opjox, is a constant equal to (p 2 - p 1 )/ L;
note that this is negative, so that for this special case Equation ( 5.5) becomes
This elementary geometry is shown in Figure 5.6 and it is assumed that the
width of the bearing, B, is much greater than its length, L. This restriction
has the effect of ensuring that most of the flow through the gap between the
FLUID-LUBRICATED THRUST BEARINGS 131
Figure 5.6
bearing and the moving surface is 1n the direction of the x axis. It will only
be near the ends of the bearing, where y = 0 and B, that there will be significant
flow in the direction of the y axis. This restriction is discussed further in
sub-section 5.6.1.
There is no change of gap or film thickness in the direction of the
y axis, but an uniform inclination of the undersurface of the bearing is
assumed to exist in the direction of the x axis, as shown in the cross-section
of Figure 5.7. The gap is shown to vary uniformly from a minimum of h1
to a maximum of h2 over the length, L, of the bearing. The stationary origin
is taken to be at the intersection of the produced upper stationary and the
lower moving surfaces of the bearing.
Now let there be a local pressure p, a pressure gradient op I ox and a
film thickness h at the section of the bearing located by the ordinate x. If
the fluid is incompressible, then the volumetric flow rate of fluid through
the bearing must be a constant. This is the continuity of flow statement,
which means that as much fluid flows into the bearing at one end as flows
out at the other, so the flow rate, Q, is a constant and independent of x;
thus, oQ/ ox = 0. Substituting for Qfrom Equation ( 5.5) in this expression gives
!__{~(- op)+~uc}=o
ox 121] ox 2
This expression may be put in the following form, as U cis not a function of x:
.. Uo
x 2 = L (K+ 11/K
Figure 5.7
132 TRIBOLOGY
oh = h 2 -h 1 = h1(h2 _ 1) = h1·K
ox L L h1 L
!__{(h x) xop} = _ 6u (h K)
1 3 3
0
1
ox L '7 ox L
or
Note that the terms in the first parentheses may be assumed to be constant
for a given bearing if the variation of viscosity with temperature is ignored
at this stage. The above expression may now be integrated twice with respect
to x, to obtain the following expressions:
(5.8)
(5.9)
This is the solution for the pressure distribution in the bearing, and the two
constants of integration, C 3 and C4 , may be determined by putting p = 0
(atmospheric) at x 1 = L/ K and x 2 = (L/ K)(K + 1). The constants have the
following values:
x =
m
{~(K +
K K+2
l).}L
and that the corresponding value of the film thickness is hm, where
The rate of flow of fluid through the bearing may be found conveniently
at xm, as there is no pressure-induced flow here, because the pressure gradient
is zero. The velocity- or shear-induced flow is then given by the second term
of Equation (5.4) as Q=(U 0 /2)Bhm and putting
h =2(K+l)h
m K+2 1
gives the flow rate as:
Last, the lower sliding member experiences a shear force due to the
shear stress r which is distributed over its surface as it passes under the
inclined pad. Remembering that the shear stress is given by r = ~ ou/ ox and
that u is given by expression (5.3), it follows that, at z = 0,
ou hop U0
r=~-= ---+~-
ox 2ox h
134 TRIBOLOGY
F= f x,
x1
rBdx=B
f(L/K)(K + 1)
L/K
( h Op U 0)
---+17- dx
2 OX h
Integrating the first term by parts and the second by putting h = (h 1 K I L) x
gives
The detailed analysis presented in the previous section has shown that, within
the inherent assumptions, it has been possible to derive expressions for four
of the main parameters which would be of interest in the design of a bearing
of the geometry under discussion. These parameters are the maximum
pressure in the film, the volumetric flow rate of fluid required to operate the
bearing, the load capacity of the bearing and the shear force necessary to
move the sliding member under the bearing.
It will be noticed that the fluid film converges in the direction of sliding
and that all of the four parameters listed above are proportional to the
sliding velocity. Thus, the pressures under the bearing are developed by the
action of the sliding member dragging fluid through the converging gap -
i.e. the bearing is self-acting.
The four expressions are reproduced in a convenient form in expressions
(5.10):
r}[;
Q = (Bh 1 U0 )(K + 1 )
K+2
(5.1 0)
dimensional information for the particular bearing, while the second term
contains information which depends only on the shape or geometry of the
bearing in dimensionless form.
In this simple case the geometry of the bearing (Figure 5.6) is described
by a single dimensionless variable, K, which is a function of the ratio between
the entry and exit film thicknesses, h2 and h1 . Therefore, the following four
dimensionless parameters or coefficients are defined and used to simplify the
presentation of the corresponding expressions of (5.10):
Pm={2(K+~~K+2J
( 5.11 )
- { Kloge(K+1)-
F= 4 6 }
K+ 2
(5.12)
Maximum pressure
Flow rate
ILl..
0.6
I~
..
-g
0.410:
E
lcl: 1<>.
h2
- I 1.25 1.5 1.75 2.0 2.1889 2.25 2.5 2.75 3.0 3.25 3.5 3.75 4.0
hl
,r-
K 0.25 0.5 0.75 1.0 1.1889 1.25 1.5 1.75 2.0 2.25 2.5 2.75 3.0 c
0.1333 0.2000 0.2338 0.2500 0.2555 0.2564 0.2571 0.2545 0.2500 0.2443 0.2381 0.2316 0.2250 0
Pm '
r-
Q 0.5556 0.6000 0.6364 0.6667 0.6804 0.6923 0.7143 0.7333 0.7500 0.7647 0.7778 0.7895 0.8000 c
OJ
:0
w 0.0884 0.1312 0.1511 0.1589 0.1602 0.1601 0.1577 0.1533 0.1479 0.1420 0.1360 0.1300 0.1242 (")
)>
F 0.9036 0.8437 0.8028 0.7726 0.7542 0.7488 0.7292 0.7122 0.6972 0.6836 0.6711 0.6594 0.6494 -1
m
0
-1
:I:
:0
c
en
-1
OJ
m
)>
:0
z
C)
en
~
w
-..J
138 TRIBOLOGY
Load capacity
Shear force
where pis the density of the oil, CP its specific heat and AT the temperature
rise of the oil. Most mineral oils have a specific heat of about 1880 Nmlkg·°C.
Hence,
Figure 5.9 has been produced to summarise the discussion of this bearing.
FLUID-LUBRICATED THRUST BEARINGS 139
960 N
0.923
N/mm 2
0.1 mm 0.219 mm
Figure 5.9
Other useful coefficients may be derived from those which have already been
presented. For example, the power balance may be modified by recognising
that only a proportion, k, of the power is convected out of the bearing by
the oil. Then
SO,
Putting T = PI Q,
ilT=kT('1UoL)=1xl.0988{ 0.4x3(50x10-3) }
pCPhi 880 X 1880(0.1 X 10- 3)2
= 0.4 oc, as before
Values of T for the bearing in question have been added to Figure 5.9.
140 TRIBOLOGY
L\T=T.-T.=kT
o I
-c-
0-{rtU L}
h2
p p 1
"=>
~
c.
b 40 E
.e !!!
~ 30
~ ~--+-~~_,~~~
·~ Effective viscosity= 0.0263 kg/m·s
~ 20 ...._..___,___.__.....____._.-J
35 40 45 50
Mean temperature rcl
Figure 5.10
and
and
2 = 40 +2- = 44.3 C
T.e = T1 +-
142 TRIBOLOGY
- ....
118 kN'
4.87 N/mm 2
I '
I \
- 6 m/s
39.1 N
O.Q18 mm
I
0.055 mm
Figure 5.11
60
55
~ 50
E
"'~"' 45
0
::
-~ 40
·;;
u
-~
~ 35
0
30
40 45 50 55 60
Oil temperature (°C)
Temperature ( C) 0
v (eSt) 'I (cP) 'I (kg/m·s)
40 68 59.7 0.0597
45 56 49.2 0.0492
50 45 39.5 0.0395
55 36 31.6 0.0316
60 30 26.1 0.0261
144 TRIBOLOGY
Table 5.2
K h 1 (mm) w F Q T=F/Q
0.5 0.06 0.1312 0.8437 0.6000 1.4062
0.75 0.04 0.1511 0.8028 0.6364 1.2615
1.00 0.03 0.1589 0.7726 0.6667 1.1589
1.25 0.024 0.1601 0.7488 0.6923 1.0816
1.50 0.02 0.1577 0.7292 0.7143 1.0209
2.00 0,015 0.1479 0.6972 0.7500 0.9296
2.50 0.012 0.1360 0.6711 0.7778 0.8626
3.00 0.010 0.1242 0.6484 0.8000 0.8105
Now the geometry of the oil film is described by the variable K, where
K = (h 2 - h 1 )/h 1 , but in this example only h 2 - h 1 is known as 0.03 mm, so
we may write
h 1 = 0.03 x 10- 3 m
or
K
This relationship may be used in conjunction with Figure 5.8 to produce
Table 5.2. Thus, the dimensionless coefficients Q, W, F and T, which describe
the operation of the bearing, are given as functions of the as yet unknown
dimensional film thickness h 1 . It is useful to plot the values of W and T
against the film thickness h 1 as shown in Figure 5.13.
0.16
1.6
1.5
1>-
0.15 ,,.
1.4
~
·u ~
"'~ ·u
ill
1.3 "'~
·~
]
~ 1.2
0.14
0
-'
~
c.
,_"E 1.1
1.0
0.13
0.8 IL----'---------'---L----'---------'---L----'---------'--___jL_______J
0.01 0.02 0.03 0.04 0.05 0.06
Minimum film thickness of oil (h 1 mm)
Now the known values may be inserted into the third expression of
(5.12), to give
or
Similarly, putting k = 0.8, p = 878 kg/m 3 and cp = 1880 J/kg·OC into (5.12)
gives
Again
dT
T.e = T+- (c)
1
2
Thus, there are six unknown quantities to be found from the three
derived expressions (a), (b) and (c) above, together with the three relationships
given in Figures 5.12 and 5.13. The following procedure may be used to
obtain solutions for these six relationships:
( 1) Guess an initial value of h1 .
(2) Obtain values of Wand T from Figure 5.13.
( 3) Evaluate 17 from expression (a).
( 4) Evaluate T. from Figure 5.12.
(5) Evaluate dT from expression (b).
(6) Evaluate I;, from expression (c).
Now, if the initially guessed value of h 1 is close to the actual value, then
the effective temperature, I;,, found in step (4) above will be very close to
the value of I;, found from step (6). Usually further adjustment to h 1 will be
necessary in order to obtain a consistent set of values for all of the six
unknowns listed above.
Take an initial guess of h 1 = 0.03 mm = 0.03 x 10- 3 m and proceed
through the six steps described above:
(1) h 1 =0.03 x to- 3 m
(2) W = 0.1589 and T = 1.1589 from Figure 5.15
( 3) 17 = 0.0909 from expression (a)
( 4) I;, = 34 oc from Figures 5.13 and 5.2
146 TRIBOLOGY
0.09( 30 X 10 - 3 )
3 2
W = 0.1586{0.0443 X 10 X }
0.021 X 10-
= 12905 N
which is close to the given load of 13 000 N.
Again the temperature rise may be confirmed from expressions ( 5.12) as
L.lT=08· 1
A Q4{ 0.0443 X 10 X 30 X 10- 3 }
. . 878 X 1880(0.021 X 10- 3 ) 2
= 15.2oC
and
T. = 40 + 15·2 = 47.6oC
2
Table 5.3
T., dT T.,
h 1 (mm) w f 11 (kg/m·s) ("C) (OC) (OC)
7.74 N/mm 2
0.021 mm 0.051 mm
Figure 5.14
(a)
Figure 5.15
the bearing it has been assumed that the one-dimensional form of Reynolds's
equation is applicable and that the load capacity, for example, may be
calculated by neglecting the reductions in pressure which occur near the ends
of the bearing. Effectively, this means that the pressure profile deduced from
the one-dimensional form of Reynolds's equation is assumed to exist over
the full width, B, of the bearing.
This is a remarkably good assumption, as in the case of a bearing of
B I L ratio equal to 2 the actual load capacity is typically only about '30%
less than that calculated on the basis of one-dimensional theory. However,
if a square bearing, of B I L ratio equal to unity, is considered, then the flow
of fluid in the direction of they axis may not be ignored and a two-dimensional
solution is necessary. Such two-dimensional solutions are obtained from the
following two-dimensional form of Reynolds's equation (Cameron, 1981):
i_{h3op}+i_{h3op}=6Uooh (5.14)
ox 11 ox oy 11 oy ox
The addition of the second term on the left-hand side involves derivatives
with respect to y and therefore includes the effects of flow in the direction
of the y axis.
The flow pattern of fluid within a square bearing will be of the form
shown in Figure 5.16( a), while the corresponding pressure distribution will
be as indicated in Figure 5.16( b). It will be noticed that there is no extended
FLUID-LUBRICATED THRUST BEARINGS 149
(a) (b)
Figure 5.16
W= w{ ~U0 B(~Y}
F=F-(~U 0 BL)
hl
'!1.
h,
L-----~----_L_____ J_ _ _ _ _ _L-----~-----L------L---~0
0 4
K
Similar sets of data have been obtained for bearings of other width-to-
length ratios, but as most of the individual bearings used in practical designs
of thrust bearings are of an approximately square form, the information
presented in Table 5.4 and Figure 5.17 is particularly useful.
25000)1
L=B= ( - - =72mm
4.8
or
(a)
Again
1]56 80 X 10- 3 }
X
11 T = 0.8 X 0.6579 { 2
ci)
880 X 1880hl
T. = 40 + 17·8 = 48.9°C
2
and the outlet temperature, 7;,, follows as 57.8 oc.
The effective viscosity of Tell us 68 at the temperature of 48.9°C may be
obtained from Figure 5.13 as 0.0413 kg/m · s. Substitution of this value into
152 TRIBOLOGY
( 1) Bearings of length less than 53.5 mm operate with values of h1 less than
0.025mm.
(2) Bearings of length greater than 69.3 mm require flow rates in excess of
10 1/min.
( 3) Bearings of length less than 44 mm generate outlet temperatures greater
than 100°C.
Table 5.5
16 120 0.07
14 0.06
12 100 c 0.05
44
eE
c
10 p 0.04 ~
>-.:'
~ ~
9: "'
~ ~
0 8
rn
80 0.03 ~
0 "'
c.
§
"'rn ~ "'E
~
0
u::
6 "
;
0
0.02 .§
~
<:
~
4 60 0.01
0.08
0 40~----~----~----~----~L-----~-----L----~----~ 0.006
40 50 60 70 80
It is, therefore, clear from Figure 5.18 that an acceptable design of bearing
must have a length of between 53.5 mm, line A, and 69.3 mm, line B. If the
boundary of line A is not violated, then that of line C is also satisfied.
As the relative importance of the limits set on h 1 and Q, by lines A
and B, has not been stated, it is reasonable to select a length of bearing
approximately midway between the lengths corresponding to lines A and B
- say 60 mm, as indicated by line D in the figure.
The values of h 1 , Q and 1;, corresponding to a length of 60 mm may be
read directly off Figure 5.18 or recalculated as a check in the manner employed
previously. It then follows that h 1 has a value of 0.032 mm, Q a value of
6.561/min and 1;, a value of 71.6°C. The table of Figure 5.22 gives the
coefficient F as 0.6633, so that the shear force becomes, from expressions
(5.12),
7777777
. 56 m/s
1
777777/r/777~77777 127 N
6.561/min
0.032 mm 0.080 mm
Figure 5.19
Consider the use of a number of square thrust bearing pads of the type and
size discussed in the previous section to support the axial thrust of 250 kN
which acts on a shaft of 200 mm diameter rotating at 3600 rev /min. The
tangential velocity of a circle of 300 mm diameter at this rotational velocity
is n(3600/60)0.3 =56 m/s, and if ten pads of length 60 mm are spaced
around a circle of 300 mm, then the arrangement of Figure 5.20 evolves.
There is adequate space for the shaft of 200 mm diameter to pass through
the 'ring' of separate thrust pads.
As ten pads of the final design deduced in sub-section 5.6.3 are used in
the complete bearing of Figure 5.20, its specification is as follows:
Load capacity= 25 x 10 3 x 10 = 250 kN
Flow rate of oil= 6.56 x 10 = 65.6 1/min
Inlet temperature = 40°C
Outlet temperature= 71.6°C
Lubricant = Tellus 68 oil
Shaft velocity= 3600 rev /min
Mean velocity over pads= 56 mfs
Power consumed= 7112 x 10 = 71 kW
Minimum film thickness = 0.032 mm
Length and width of pads = 60 mm
The analysis presented here is typical of the methods used in bearing
design, but the schematic arrangement shown in Figure 5.20 is a gross
simplification, and details of the construction of more realistic bearings are
shown in Figure 5.21. An important feature of practical thrust bearings is
that the necessary convergent geometry of the pads is created by allowing
them to tilt about a fixed pivot rather than by attempting to build a known
taper into the assembly. Consequently, the analysis and design of such
practical thrust bearings is more complicated than outlined here, and more
details are available elsewhere (Martin, 1970).
FLUID-LUBRICATED THRUST BEARINGS 155
156 TRIBOLOGY
Section A-A
5.8 PROBLEMS
W= i{ '1~r}LK + ~) 3 + ~}
where K = h 2 /h 1 -1.
(a)
(b)
Figure 5.22
FLUID-LUBRICATED THRUST BEARINGS 157
A bearing of this type lies with one. of its lands in contact with its
stationary working surface, as shown in Figure 5.22(b ), under a load of
3000 N. The gap between the other land and the stationary surface is
0.07 mm, the length of the bearing is 90 mm and its width is 240 mm. If
the working fluid has a dynamic viscosity of 0.05 kg/m · s, estimate the
sliding velocity necessary to separate the bearing from the working surface.
[0.15 m/s]
2 The inclined hydrodynamic bearing pad shown in Figure 5.23(a), of
length L and width B, operates with a fluid of dynamic viscosity 17 and
is positioned relative to a plane slider, which has a velocity U in the
direction shown, so that the film thickness, h2 , at entry is k times the
film thickness, h1 , at exit. During operation the pressure in the fluid at
exit is maintained at a value of p 1 avoe atmospheric, while that at
entry, p2 , remains at the atmospheric value. On the assumption that the
width, B, is much greater than the length, L, solution of Reynolds's
equation shows that the load capacity of the bearing, W, is given by the
expression
(b)
0.05 mm 0.15 mm
(c)
Figure 5.23
158 TRIBOLOGY
and the volumetric flow rate, Q, of fluid through the bearing is given by
the expression
Bhip 1 }
Q=a 3 { BUh1 } -a4 { ~
(a) The thrust pad shown in Figure 5.23 (b) has a length of 30 mm and
a width of 100 mm, and is machined so that a uniform taper ofO.l mm
exists over a length of 15 mm, while a parallel step of height 0.05 mm
exists over the remaining length of 15 mm. A load of 4000 N is carried
by the bearing when the slider has a velocity of 21.5 mjs and an oil
of dynamic viscosity 0.05 kg/m · s is used. Show that the bearing will
operate with a minimum film thickness of 0.05 mm, as shown in
Figure 5.23(b ).
(b) If the load of 4000 N is imposed on .the bearing when it starts from
rest, as shown in Figure 5.23 (c), estimate the sliding velocity at which
the bearing will separate from the sliding member.
[13.72 m/s]
3 The inclined hydrodynamic thrust bearing shown in Figure 5.24(a) of
length L and width B operates with a fluid of dynamic viscosity 11 and
is positioned relative to a plane slider, which has a velocity U0 in the
direction shown, so that the film thickness, h2 , at entry is 1.5 times the
film thickness, h 1 , at exit. During operation the pressure in the fluid at
exit is maintained at a value of p 1 above atmospheric, while that at entry
is maintained at a value of p 2 above atmospheric. Solution of Reynolds's
equation on the assumption that the width, B, is much greater than the
length, L, shows that the load capacity of the bearing is given b-y
and that the volumetric flow rate, Q, of fluid through the bearing is given by
Q =~(U Bh )+ 3Bhi(P2-Pl)
5 ° 1
2011 L
Monitored pressure
h,
0.08 mm 0.12mm 0.18mm
Ia) (b)
Figure 5.24
and that the volumetric flow rate, Q, of fluid through the bearing is given by
~ (a)
1 r~- 49 pads
"""E'l~--.7.r 117.~
r I
' '
L 2m dia.
(b)
~
0.07 mm
~~ (c)
Figure 5.25
Three different oils of dynamic viscosity 0.02, 0.04 and 0.06 kg/m · s
are available. Select one of these oils so that the pads operate with a
minimum film thickness of about 0.07 mm.
[0.06 kgjm·s]
Chapter 6
Fluid-lubric ated Journal
Bearings
6.1 INTRODUCTION
In Chapter 5 bearings of plane form only were considered and these were
shown to be of use in supporting loads which act at right angles to their
plane. Bearings of this type are generally recognised as thrust bearings.
However, another important type of bearing is the journal bearing. Journal
bearings are used to support rotating circular shafts which are loaded in a
radial direction.
In many applications it is found that a hydrodynamic journal bearing
is most suitable for the range ofloads and speeds involved, and it is fortunate
that the shape of the oil film which is formed between the bore of a circular
bearing and an eccentric rotating shaft is conducive to the generation of
hydrodynamic pressures which are, in practice, able to support considerable
radial loads.
6.1 .1 Geometry
161
162 TRIBOLOGY
(AO) R1
cos ct = { 1- (;J 2
sin 2 () }t = 1- ~(;J 2 sin 2 ()
A C = R 2 + h = e cos () + R 1 { 1 - ~(; J
2
sin 2 ()}
or
FLUID-LUBRICATED JOURNAL BEARINGS 163
or
h = c( 1 + e cos()) ( 6.1 )
where e = ejc is known as the 'eccentricity ratio'. This is a dimensionless
geometric parameter which describes completely the shape of the fluid film
which exists between the surfaces of the journal bearing and shaft.
Remembering that the radial clearance, c, and therefore the film thickness,
h, are very much smaller than the radius of either the journal bearing or the
shaft, then the shape of the film between the two surfaces may be developed
into the equivalent orthogonal form shown in Figure 6.2(a). The clockwise
rotation of the shaft shown in Figure 6.1 produces a tangential surface
velocity, U 0 , which appears as the velocity of the lower surface in
Figure 6.2(a). In Figure 6.2(a) the upper surface is fixed in space and
corresponds to that of the journal bearing. The coordinate () varies from
zero, where th,e film thickness is c + e, to n, where the film thickness is c - e,
and then to 2n, where the film thickness is again c + e.
Inspection of Figure 6.2(a) shows that a convergent film exists between
() = 0 and () = n, while a divergent film exists between n and 2n. As the
direction of sliding of the lower surface, at a velocity U0 , is in the direction
of convergence between () = 0 and () = n, it is clear that hydrodynamic
pressures will be developed in the fluid as it flows through the bearing. This
important deduction follows from the general nature of the results which
were obtained in Chapter 5 for the plane convergent geometry of Figure 5.7.
It must be remembered that the sections corresponding to () = 0 and
() = 2n in Figure 6.2(a) do, of course, define the same section because of the
actual circular shape shown in Figure 6.1.
164 TRIBOLOGY
Bore of journal
R,
(a)
Circumferential profile
of hydrodynamically
Pressure generated pressure
(b)
(c)
Figure 6.2 Film shape and pressure profile between shaft and bore of a hydrodynamic
journal bearing
Surface of bore of
journal_bearing
Line of centres
By combining the geometry of Figure 6.1 with the deduced pressure profile
shown in Figure 6.2(b ), the general features associated with the operation
166 TRIBOLOGY
of a practical journal bearing may be seen in Figure 6.3. Fluid is drawn into
the converging clearance at E by the rotating shaft and some of this fluid
flows completely through the bearing, to emerge at G, where the pressure in
the fluid has returned to the ambient value.
The pressure distribution described in Figure 6.2 does, of course, act
around the surface of the shaft and integrates to support the radial load, W,
which is applied to the shaft as shown in Figure 6.3. As the pressure
distribution acts over the arc between E and G, the direction of the load W
which acts on the shaft cannot coincide with the line of centres EOCF. The
relationship between the direction of the load and the line of centres is given
by the 'attitude angle', which is shown as rjl in Figure 6.3. As the load
increases, the position adopted by the centre of the shaft within the available
radial clearance, c, is as indicated in Figure 6.3. The relationship between
the attitude angle, rjl, and the eccentricity ratio, e, is such that a trajectory
of approximately semicircular shape is formed (Cameron and Wood, 1949).
Fluid may be introduced into a bearing thnmgh various configurations
of supply grooves. The usual position of supply groove adopted in many
common bearings is shown schematically in Figure 6.3, although many other
variations exist (Martin, 1983).
The analysis of the flow of fluid within the confines of a journal bearing
is clearly much more complicated than in the case of plane thrust bearings.
Consequently, a two-dimensional form of Reynolds's equation must be solved
(Cameron, 1981 ), and additional complications are encountered in the
establishment of the correct boundary conditions to be applied at the
termination of the pressure profile in the divergent part of the clearance,
section G in Figure 6.3.
As, in general, flow of fluid into a convergent clearance causes an increase
in pressure, it follows that its continued flow into a divergent clearance-will
lead to a decrease in pressure. However, when the pressure approaches the
ambient value, air may be drawn into the divergent clearance, as there will
be insufficient fluid available to fill the clearance, which increases in the
direction of shaft rotation. Similarly, dissolved air or other gases may come
out of solution in the fluid as the pressure decreases. Again, if the pressure
falls below the vapour pressure of the fluid, then cavities composed of the
fluid's vapour will form in the film. It is, of course, physically impossible for
the pressure in the film to fall below zero absolute (Dowson and Taylor, 1975).
These physical limitations are generally referred to as cavitation, and
result in streamers or striations of fluid being formed in a mixture of air and
vapour in the divergent clearance (Cole and Hughes, 1956). Much effort has
resulted in useful sets of boundary conditions being postulated (Dowson and
Miranda, 1975; Etsion and Pinkus, 1975).
A more complete description of a typical complete journal bearing is
given in Figure 6.4, where a shaft of diameter d is supported in a bearing of
axial length b. The shaft rotates with an angular velocity N and supports a
FLUID-LUBRICATED JOURNAL BEARINGS 167
Lubricant
supply groove
Stationary journal
H w bearing
load W, while a power H is required to rotate the shaft against the viscous
resistance which is produced in the clearance of the bearing. Fluid is supplied
at a flow rate Q with viscosity '1r through a supply groove of axial length 1
and width a, which is machined in the wall of the bearing, but the bearing
operates with an effective viscosity '1· The difference between the diameters
of the shaft and the journal bearing is Cd, so that the previously defined
radial clearance, c, is simply Cd/2.
Other important parameters associated with the arrangement of
Figure 6.4 are listed in the Notation section (pp. viii-xii) and introduced in
the following sections.
W= w{ qNbd(~J 2 } ( 6.2a)
Q= Q(NbdCd) (6.2b)
or
Load capacity
W= 312900 N
Flow rate
Power
Temperature rise
rJNK(_!_)2 = (11T)
AT Cd
Multiply this expression by expression (6.2a ), to obtain
WK b -- -
d 2 AT = d(WQIH)
Substitute for rJN from expression (6.2c) into expression (6.2d ), to get
Nd 3
Q(d)
cd = d(Q)
b_
The coefficients on the right-hand sides of the above expressions are functions
of only the length-to-diameter ratio, bId, and the operating eccentricity ratio,
e, for a particular bearing. These three coefficients are now redefined as
follows:
- b -- -
W0 =d(WQIH)
T0 = (11T) (6.3)
- b -
Qo = d(Q)
172 TRIBOLOGY
(6.4)
HK
Nd 3 ilT Cd
(d) = - Qo
1.0 bid-f -
8 1.5
6
5 == --
1.
-
-
v
1-· f-. 0.9
- - · -- - - -~0.8
2
3
- = 0.7
"'~ 0.6
0.
..
~
-·
0.4
10- '
8
0.3 -
6
5
4
-
3
0.
-
I-=
10- 2 = _, == - /
,-; 8
~ 6
;g 5
4 0.1 f-
] 3
1--
E
c0
·~ ..
e
0 10- l
8
6
5
4
3
1--
1o-•
8
6 - -
5
4
3
2
=
0.1 0.2 0.3 0.4 0.5 0.6 0 .7 0.8 0.9 1.0
Eccentricity ratio, c:
be at least 0.7, it is seen from this figure that a value of e = 0.71 occurs when
bId= 0.4. It then follows that the length of the bearing is 0.4 x 50= 20 mm.
The diametral clearance now follows from expression ( 6.1) by putting
e = 0.71 and h = 1.5 x 10- 2 mm at 0 = n; thus,
2(1.5 X 10- 2 )
Cd= =0.103mm
1-0.71
174 TRIBOLOGY
0.03
= - - i...--
- - - - -
--- - - -
- -- -
- ~
- - - - -
0.02
-
- -
//./. ~
' ' """'...... -
-
~- ~
-
r--
·"""' f\\ - 1--
- """
0 -~
1>- ~~ ~
~
0.0
f-- [0..'\.\ \\\\\ -
/
............ ~ ~\W · b /d -
- - ~ ~ -
~""
~-:~-
~~
- -
~ 1\~~ '1- -
-
- - - - - \. \'~\
g:~ - -
\~
O.QI
0.7= --
n ,_ -
_1.0=
0.008
0.007
1-------j
f--- - -
--1
t--
- f- 1-
--
n rvv:
0 0.1 0.2 0.3 0.4 0.5 0.6 07 0.8 0.9 1.0
Eccentricity tatio, E
Figure 6.6 Variation of dimensionless coefficient 7'0 with eccentricity ratio e and
length-to-diameter ratio b / d
The type of oil necessary may now be found from Figure 6.6 by
determining the value of T0 corresponding to bId = 0.4 and a = 0.71; thus,
2
T_0 =rJNK(
-- -
d ) = 17.6 x 10 _ 3
!iT cd
The angular velocity is 10 000 I60 = 167 rev Is and the required dynamic
viscosity, rJ, follows as
3.0 F - -
-- -
= 1- - -
2.0
b/ d -
- -
>----- - - 1.5-
f.-
- -
1.5 ~
>--- - 1-- 1 - -- L-
1.0_
;:::::;:;;;- ~ 0,9 _ -
-~ 0 .8 -
1.0
~ -~ O.J -
0 .9 0.6
::;;;
0.8
0.5 ====::
0.7
I~ 0.4
-~ 0.6
u 0.5 +---
"§ 0.3
E o.4
c0
·~ 0.2 -
E 0.3
i5
0.2
-
0.2
0.1 5
v
~
~ = - -""'
0.1 _
0.1
0 .09
...........
0.08
0.0 7
0.06
0.05
0 0.1 0.2 0.3 0.4 0.5 0.6 0. 7 0.8 0.9 1.0
Eccentricity ratio, t
It is seen from Figure 6.7 that when b/d = 0.4 and e = 0.71, the value
of Q0 is 0.518. Thus, from the fourth expression of (6.4),
-
Qo= Hk (d)- =0.518
Nd 3 AT Cd
or
- 0 =Q
Q -3 -
Nd cd
(d)
=0.518
or
Q = 0.518{ Nd 3 ( ;J}
= 0.518{ 167(50 X 10- 6 ) 3 ( 0· 103 X 10 - 3 ) }
50 X 10- 3
Q = 22.28 x 10- 6 m 3/s = 1.341/min
~
-~-
~0"
~)1
~~-
rectangular groove are not suitable for shafts which are able to rotate in
both directions or for loads which vary significantly in direction.
Provision of two rectangular grooves at right angles to the load line
overcomes these limitations to some extent, as shown in Figure 6.9( c). The
arrangement of Figure 6.9( c) is commonly adopted for constructional reasons,
as bearings are often split along a line at right angles to the load line in
manufacture.
Further consideration of oil-supply grooves will be limited to a
discussion of the flow of fluid from and the design of a single rectangular
groove in the locations shown in Figures 6.9(a) and (b).
180 TRIBOLOGY
In parallel with the second expression of (5.12), analysis shows that the
velocity-induced flow rate of fluid from a rectangular groove may be expressed
as
(6.5)
where the dimensionless coefficient Qu is dependent on the ratio between the
length of the supply groove and the length of the bearing, ljb and the
operating eccentricity ratio of the bearing, e, as shown in Figure 6.10.
Similarly, the pressure-induced flow from a rectangular groove may be
given as follows:
Qp=Qp{p~~~} (6.6)
Many assumptions have been made in the analyses which are summarised
by expressions (6.5) and (6.6 ), as the geometry of a rectangular groove results
in complicated flow patterns. Simplifications have made the curves of
Figure 6.10 independent of the length-to-diameter ratio of the bearing, bjd,
and the angular position of the groove, but the length of the groove, l, does,
of course, appear in the ratio ljb.
The value of the dimensionless coefficient QP is conveniently expressed
as the product of two other coefficients, Q1 and Q2 , which are given by the
curves of Figures 6.11(a) and 6.11(b ), respectively. Thus, QP = Q1 Q2 , so that
QP = Ql Qz { p~~~} (6.7)
FLUID-LUBRICATED JOURNAL BEARINGS 181
2.0
1.5 1/b=
0.9 -
_;.... 0.~=
1.0 0.7=
0.8 0.6=
o.s=:
0.6 0.4::::=
0.5
0.3::;.::;:
0.4 _...,..,
1---C 0.25=
0.3
-- - -A < - / / . /
_.... - .2=
"'
~
10
~
;g 0 .2
f - - 1- //
~E o.25 =
f--
~c
e 0.08 =0.1 7
0 -
0.06
0.0 5
=
0.04 1 - - -
·-
0.0 3 F==
f- 1-
0.02
0 .025 1--
1------ - 1---
0.1
1.5
r-- 1/ b
1.0
- ·- ·-
-=F=- 0.9- -
O.B o.8 =
10 0.7 =
~ 0.6
o.s:
;g 0.5 0.5 - -
] 0.4 - -
~
0.4
-- .. ~
0.3 - · -
"
-~
0 .3 0.2
e 0 .2 --
c
·:
0 - c:-
. - --
1-
0.15
-- - - ·-
0.1
0 0.1 0.2 0.3
1.0
0.8 · -1-- ·
10 0.6
~
·c;
0.5
;;: 0.4
8 · -1-···
"~ 0.3 __ _._-
c - r-- - f--- - .
·~
e 0.2
c
0
0. 15
Eccenuicity ratio, t
(b)
a thrust bearing as that at the effective or mean temperature of the oil during
its passage through the bearing. Most of the oil which enters a thrust bearing
passes completely through the converging clearance and is nearly all available
to convect heat from the bearing. Consequently, the effective temperature
was defined as the sum of the inlet temperature and a half of the temperature
rise as given by the power balance (sub-section 5.5.4 ).
However, in most journal bearings the width of the load-bearing film
is much less than its length. For example, in a journal bearing with the
FLUID-LUBRICATED JOURNAL BEARINGS 183
Qu = iJu(NbdCd)
Qp = Ql Q2( p~~J)
= 0.59 X 0.62{(0.4 X 106)(0.103 X 10-3)3}
14.3 xw- 3
Qp= 11.2 X 10- 6 m 3ls=0.67llmin
Thus,
Qu + QP = 0.72 + 0.67 = 1.39llmin
This flow rate is greater than the required value of 1.34llmin and is therefore
acceptable.
The length of the axial groove is therefore
l = 20 x 0.65 = 13 mm
and its width is
a = 20 x 0.25 = 5mm
AT
AT
T, AT
T,
k 0.8 6 2
K=-= =0.484x 10- m ·°CjN
pCP 880 X 1880
2(0.02 x w- 3 ) o.o4 x w- 3
Cd= = m (b)
1-B 1-B
0 AT Cd '1 AT Cd
or
-
T0 = 0.967 x 10 -6(-ATq- '1 ) (c)
FLUID-LUBRICATED JOURNAL BEARINGS 187
In addition, the effective temperature of the oil will be (50+ AT)°C. The
sequence of calculations outlined in Section 6.5 is now carried out for
various values of AT until convergence occurs between Figure 6.5 and
Figure 6.6 and the viscosity temperature characteristics for Tellus 46 oil given
in Figure 5.2. Take AT= 20°C; then, from expression (a) above,
- 2.418
W0 = - - = 0.129
20
Reference to Figure 6.5 shows that this value corresponds to bid= 0.7 and
6 = 0. 76 or bId = 0.5 and 6 = 0.81. Take the first pair of values and proceed
to find cd from (b) as
o.o4 x w- 3
C - 0.167 X 10- 3 m
d- 1-0.76
Put this value in (c) and recall that, as both bId and 6 are now known, the
value of T0 may be read off Figure 6.6 as 14.52 x 10- 3 ; hence,
W - 2-
-0 - 18 -
.4- - 0 •096 7
25
which corresponds to bid= 0.6 and 6 = 0.77 in Figure 6.5.
The value of Cd then follows from (b):
o.o4 x w- 3
C= =0.174x10- 3 m
d 1-0.77
With bid= 0.6 and 6 = 0.77, the value of T0 may be read off Figure 6.6
as 15 X 10- 3 , SO that (c) becomes
15 x w- 3 = o.967 x w- 6 { 25(0.17417x w- 3 )2 }
for which 11 = 0.0117 kglm · s or 13.3 eSt.
The effective temperature of the oil is (50+ 25) = 75°C and Figure 5.2
shows that the kinematic viscosity of Tellus 46 at 75°C is about 13.5 eSt, so
that the values corresponding to a temperature rise of 25°C are consistent.
188 TRIBOLOGY
With b/d = 0.6 and e = 0.77, the value of Q0 may be read off Figure 6.7
as 0. 78, so that the power loss follows from the fourth expression of( 6.4) as
QP = Q1Q2{Pr~:}
Now the viscosity of the oil at an inlet temperature of 50°C is, from
Figure 5.2, 30 eSt or 30(0.88)/1000 = 0.0263 kg/m·s. The supply pressure
required to provide the necessary flow rate of oil may now be found from
the above expression as
move down in the direction of the load but also moves around the bearing
in the direction of rotation (Figure 6.3). In some cases this motion can
continue, so that the shaft centre describes a circular orbit. If this rotation
takes place at half the rotational speed, it will coincide with the mean
rotational speed of the lubricant and no relative motion between the film
shape and the lubricant will occur, so that the hydrodynamic mechanism
will be destroyed.
This phenomenon will also take place at a certain threshold speed but,
unlike synchronous whirl, there is no possibility of running through half-speed
whirl, further increases in speed only accelerating the failure of the bearing.
6.8 PROBLEMS
1 For a range of eccentricity ratio 6 between 0.2 and 0.7, the dimensionless
load coefficient, W (defined in expression 6.2a ), is found to be related to
the eccentricity ratio by the following expressions:
log 10 (W) = 2.24966- 1.194, when bjd = 0.3
log 10 (W) = 1.9186-0.347, when bjd = 0.8
A complete journal bearing has a diameter of 80 mm, a length of 64 mm
and a diametral clearance of 0.08 mm, uses an oil of effective dynamic
viscosity 0.04 kg/m ·sand supports a shaft which rotates at 600 rev /min.
(a) The radial load imposed on the bearing in service varies between
3000 Nand 18 000 N. Estimate the corresponding values of eccentricity
ratio.
[0.267 and 0.673]
(b) The bearing is now modified by machining in it a complete centrally
located circumferential groove of width 16 mm, as shown in
Figure 6.8(b ). Find the value of the diametral clearance required to
ensure that this bearing operates at an eccentricity ratio of 0. 7 under
the full load of 18 000 N.
[0.036 mm]
2 It is found that in the range of eccentricity ratio 6 between 0.2 and 0.7
the dimensionless load coefficient, W (defined in expression 6.2a ), may
be related to the eccentricity ratio and length-to-diameter ratio of the
bearing, bId, by the following expression:
W = 0.68 (d
b)1.76 e 4 · 2'
192 TRIBOLOGY
(a) The load carried by the bearing in service varies between 3000 N
and 15 000 N. Estimate the corresponding range of eccentricity ratio
and the minimum film thickness under the larger load.
[0.276-0.66; 0.014 mm]
(b) What should be the length of the bearing if it is to operate at an
eccentricity ratio of0.5 uader a load of 3000 N? Its other dimensions,
the viscosity of the oil and the speed of rotation remain unchanged.
[45.6mm]
(c) The length of the bearing is now fixed at 48 mm while carrying a
load of 3000 N. Investigate the variation of eccentricity ratio and
minimum film thickness when the diametral clearance varies between
0.050 mm and 0.070 mm. Hence select the diametral clearance which
gives a minimum film thickness of 0.02 mm.
[0.058 mm withe= 0.312]
W = 0.18e 4 · 8 '
(a) Show that the stiffness of the bearing for small displacements about
any steady operating eccentric position defined by the eccentricity,
e, may be found as:
7.1 INTRODUCTION
194
THE LUBRICATION OF HIGHLY LOADED CONTACTS 195
(b)
---Uz
The film shape between the cylinder and plane has the form
h = h0 + R( 1 -cos fJ)
Solution of the Reynolds equation must clearly include the possibility of
cavitation in the divergent outlet region, as described in Section 6.13. An
approximation to the film shape is given by assuming a parabolic profile
xz
h=h 0 +-
'2R
F,
- u,'
Figure 7.2 Cylinder and plane
The value of f3 depends on R/h 0 , but for R/h 0 greater than about 10 4 may
be taken as 2.45.
The sideways force on the cylinder due to the horizontal component of
pressure is given by
R
Fx2 =-F
Rz x
where R 1 and R 2 are the radii of the original cylinders.
The viscous drags along the plane surface and around the cylinder
circumference are
(7.3)
Fx
F 2 =--+(U 1 -U 2 )AB
2
The value of A, assuming a full contribution from the cavitated region, is
found to be 3.84 for cases where R/ h 0 is large.
The forces described above are shown in Figure 7.2.
v=-
1 fx, p(s)log(x-s)·ds (7.5)
E' x,
(7.6)
If isothermal conditions no longer prevail, the energy equation and the
conduction equation for the heat passing to the surfaces must be incorporated.
The cavitation boundary condition must also be employed as described in
sub-section 6.1.2.
Although a full solution had to wait until the introduction of the
high-speed digital computer, inspired assumptions produced very good
198 TRIBOLOGY
7.6 RESULTS
Results from the foregoing analyses permit the plotting of the shape of the
distorted cylinder and the pressure distribution" for a given set of conditions.
Typically, for a relatively slow-speed, rectangular footprint these will have
the form illustrated in Figure 7.3.
Considering first the film shape, we see the convergent inlet zone, followed
by a virtually parallel section. At outlet there is a constriction which can
amount to a reduction of 25% of the parallel film thickness.
The pressure curve is very close to the Hertzian dry contact pressure
shown in Chapter 2. There is a build-up in the inlet and towards the outlet
there is a pressure spike and the cavitation boundary. The spike is difficult
to find experimentally, because it is very narrow (Kannel, 1965), but it can
be shown theoretically to exist and is important because of the possibility
of high subsurface stresses. There is evidence that, if a more accurate
pressure-viscosity relationship is used (Roelands, 1966), the calcUlated
severity of the spike may be reduced. Of course, the total area under the
curve must be the same as that under the Hertzian half-ellipse. The length
of the parallel section will be approximately 2{(8RW/nBE')}l and the
maxirimm pressure (WE' /2nR )f.
As the speed increases, the pressure distribution will depart more and
more from the Hertzian, as shown in Figure 7.4.
Pressure , - - ,
/ \ ,,P'
II '''
·.1
I 'j
I I
I ~
-/i
I
.
--- r:
For the elliptical footprint the results are very similar, as shown in
Figure 7.5, which is an interference pattern for a circular EHL footprint.
There is still an essentially parallel central section, while the constriction at
outlet now extends around about 180° of the contact circle (Gohar and
Cameron, 1966).
Ex11 ~m
Qron<;~e
Purple
Green
I• Gr~:~
1 Yellow
Yellow
Formulae for film thickness are the result of fitting an algebraic expression
to a large number of computed results. Such a formula only applies to the
particular range of results from which it was derived and must not be
extrapolated into wildly different areas. For this reason we need to define
'regimes' for film thickness calculation and to use the formula appropriate
to each regime (Johnson, 1970). The regimes are defined as follows.
( 1) Rigid-isoviscous In this case the pressures are not high enough to cause
appreciable viscosity change or elastic deformation.
(2) Rigid-piezoviscous The elastic deformation is negligible, but there is
significant viscosity increase with pressure.
(3) Elastic-isoviscous Considerable elastic deformation, but negligible
viscosity change.
( 4) Full EHL solution, including deformation and viscosity change
In order to determine the appropriate regime, it is necessary to refer to
the relevant chart (Figures 7.6-7.9). The horizontal axis of the charts is the
criterion for elastic deformation, g., and the vertical axis is the criterion for
significant viscosity increase, gv. To evaluate these criteria, it is necessary to
examine the dimensionless groupings of the next sub-section.
The results from the computer analyses are conveniently presented in the
form of dimensionless groups. This has the advantage of ease of plotting,
independence of the system of units used and, if the groups are chosen
intelligently, the possibility of physical insight.
The most popular scheme is to use the following four groups:
the dimensionless film thickness fl = h0 1R or fl = h0 1Rx for elliptical
contacts
the load parameter W= WI E' RB for rectangular contacts or
W = WI E' R; for elliptical contacts
the speed parameter 0 = 1'/o U IE' R for rectangular contacts or
0 = 1'/o U IE' Rx for elliptical contacts
the materials parameter G = rx.E'
For the elliptical contact we also need the ellipticity parameter
k=alb (7.7)
where a is the semiaxis of the contact ellipse in the tranverse (y) direction
and b is the semiaxis in the ( x) direction of motion. It is sufficiently accurate
THE LUBRICATION OF HIGHLY LOADED CONTACTS 201
10000 9h ~ 500
9h ~ 300
I
I
9h ~ 200 I
1000
I
9h ~ 100 Rigid
9h 70 piezoviscous
~
I
9h ~50 I
I
., 100
9h ~ 30
I
~
9h ~ 20
E I I
~ I
Q_
~
Uh = 10 I
'iS I
I;!
> 10
9h ~ 7 I
;,~
~c¥0
·~0
I I ~v
9h = E -5'
0
I
1.0 I
Elasticity parameter 9e
10 7
10"
g.~ 10000 /
10• g. ~84000
.;; Piezoviscous /
.;; 10' Piezoviscous 40000 /
~.,-
!
~
rigid 2500 /
rigid 20000 0..>'
~
·,r.,V /
105
1000 / 0~ 106 10000 ~"" /
~'~?1
/
~'~·11
~
Q_
5000
~ Q_
-~
10 4 105 1832 /
0
I;! .8~ /
> I;! / / lsoviscous
10 3 > 104
lsoviscous
rigid rigid
10 2 103
10 10 2 10 3 104 10 5 10•
10 2 10 3 104 105 10• 10 7
I
rigid 150000 /
/
10 7 65000 / /
0.
20000 /
~
/
'§ 10" /
/
~ 6562/ / lsoviscous
> 105
104
10 3 104 10 5 106 10 7 10 8
Elasticity parameter Ue
Rectangular Footprints
w
ge = O!
the film thickness may now be expressed by the equation
(7.9)
for all the regimes, the regime being identified from Figure 7.6. However, the
values of the constant Z and the exponents m and n are different for each
regime, as indicated in Table 7.1.
THE LUBRICATION OF HIGHLY LOADED CONTACTS 203
Table 7.1
Regime z m n
(1) 2.45 0 0
(2) 1.05 l. 0
3
(3) 2.45 0 0.8
(4) 1.654 0.54 0.06
Note the very low value of the exponent n in the full EHL solution.
This indicates the effect of elastic modulus, since g. is the only term containing
E'. This suggests that the film thickness is relatively insensitive to changes
of elastic modulus and it is found to be relatively insensitive to changes of
load. This is confirmed in practice, where an increase in load or decrease in
modulus will produce a resulting increase in the flattening of the surfaces
and a bigger effective load-carrying area. The above values are for the
minimum film thickness at the constricted outlet. The film thickness over
the essentially parallel region is given by the ratio
ho/hc = 0.72-0.81
The error incurred by assuming a ratio of 3/4 is negligible for most cases.
Elliptical Footprints
The 'constants' Z, m and n will be different for each regime, as will the
function f(k). The regime is identified from Figures 7.7-7.9. The values of
Z, m and n are shown in Table 7.2.
Equation (7.10) gives the value of the minimum film thickness, i.e. at
the outlet constriction. The ratio between this thickness and that in the
parallel region shows considerably more variation than the elliptical case.
204 TRIBOLOGY
Table 7.2
Regime z m n f(k)
A valuable manual for the calculation of EHL film thickness in line contact
is published as ESDU Item No. 85027 ( 1985). A similar publication is in
preparation for three-dimensional (point) contacts.
The linear speed of the two surfaces is the same, assuming no slip, and so
h = 2.45( U 1 + U 2 )17RBjW
= 2.45 X 2 X 6.81 X 3 X 10- 3 X 18.9 X 10- 3 X 35 X 10- 3 /3000
h=0.022J.lm
THE LUBRICATION OF HIGHLY LOADED CONTACTS 205
This is, of course, a very small film thickness, much smaller than the
surface roughness of the roller, and this theory does not predict full film
lubrication.
Now we shall introduce elastohydrodynamic theory into the
problem.
Since the two surfaces are steel, E' = E/(1- v2 ), and assuming
E = 208 x 10 9 Njm 2 and v = 0.3 gives E' = 228 x 10 9 N/m 2 • Calculating
the operating parameters,
W= 3000/(228 X 10 9 X 18.9 X 10- 3 X 35 X 10- 3 ) = 1.99 X 10- 5
0 = 3000 X 6.81/(228 X 10 9 X 18.9 X 10- 3 ) = 4.74 X 10- 12
G= E' = 2.3 X 10- 7 X 228 X 10 9 = 524 X 10 2
The criteria for the effect of viscosity change and elasticity are as
follows
= (1.99 X 10- 5 )~524 X 10 2 = 2137
gv (4.74 X 10- 12 )!
g.=l.99x 10- 5 /(4.74x 10- 12 )!=9.14
Plotting these values on Figure 7.6 shows a point just in the rigid-
piezoviscous regime. Using the appropriate values from Table 7.1 in
Equation (7.9),
gh = 0.99(gv)f = 0.99(2137)f = 164.2
0 4.74 X 10- 12
11=-;;::-gh= 5 x 164.2=391 x 1o-
7
w 1.99 x 1o-
h0 = 391 X 10- 7 X 18.9 X 10- 3 = 0.739 ,Uffi
Note that the film thickness has been increased by a factor of more
than 30 by the inclusion of variable viscosity. This is a much more
realistic thickness.
(2) A standard deep-groove ball bearing has balls of 8.5 mm diameter. The
inner race rotates at 5000 rev /min. The groove in which the balls roll
on the inner race has a radius in the direction of motion of 10 mm and
a radius of curvature of 4.6 mm at right angles to the motion. The load
on the bearing is 300 N and it is lubricated with an oil of viscosity
0.6 Nsjm 2 and viscosity index 2.1 x 10- 8 m 2 jN. Estimate the film
thickness.
The load on the most heavily loaded element in a rolling bearing
is given as a pessimistic estimate by 5IN times the total load, where N
is the number ofrollers or balls (Stribeck, 1901). If in this case there are
12 balls,
W = 5(300/12) = 125 N
206 TRIBOLOGY
{J =
gh
( 0)2 = 2.305
W
X 105(4.629
6.174
X lQ-9)2
X 10-S
= 1.292 X 10-3
As the fluid is dragged into the inlet zone, the shearing which takes place
increases the temperature of the lubricant and, therefore, reduces its viscosity.
This results in a reduction of film thickness, the magnitude of which is
dependent on the parameter I, given by
Thermal
1.0~
reduction 10-1
factor
10-2 L
_ __J__ _L . __ _
_ J __ __.__ __J
7.7.3 Starvation
All the results so far have assumed that the inlet to the contact is full of
lubricant. If this is not the case, the contact is said to be starved. This results
in a delay to the start of pressure build-up and a consequent reduction of
film thickness.
For rectangular footprints the reduced film thickness, h., can be
expressed approximately in terms of the fully flooded film thickness, h 00 , by
the formula
h. 2
=- arctan[l.37(J + 0.5) ]
2
-
h 00 n
where J = b1x;/(2Rh 0 )l and X; is the distance from the edge of the Hertzian
zone to the inlet boundary, which is normally considered to be the position
where the films of lubricant on the two surfaces come together; and b is the
Hertzian half-width (Wymer and Cameron, 1974).
For the elliptical case we first determine the dimensionless distance m;.
As the inlet position is moved progressively inward from infinity, it reaches
a position where it begins to affect the film thickness. This distance from the
centre, made dimensionless by dividing by b, the Hertzian half-width, is m;.
If h 00 is the minimum film thickness in the fully flooded case,
!2_ = (~)0.25
hoo m;-1
where m is the actual dimensionless distance to the inlet boundary (Hamrock
and Dowson, 1981).
Clearly, in a case of a starved contact the effect of the shear heating in
the inlet is reduced. The two effects are, therefore, interdependent. A useful
review of the combined effect can be found in Gohar ( 1988).
208 TRIBOLOGY
b= [8WR]t
nBE'
8 X3000 X 18.9 X 10- 3 Jl
[
= 1t X 35 X 10- 3 X 228 X 10 3
b = 1.35 X 10- 4 m= 0.135 mm
Therefore, the distance of the inlet boundary from the edge of the Hertzian
zone is
X;= 5-0.135 = 4.865 mm
J= 0.1351 X 4.865 I=
19 _3
(2 X 18.9 X 1.23 X 10- 3 )3
h. 2 2
-=-arctan( 1.37 x 19.8 ) = 0.9988
hoo 1t
In other words, the film thickness is virtually unaffected by the delay in
pressure build-up.
Only if the contact suffers much more starvation than this will the
film thickness be diminished. For example, if the inlet meniscus were
0.5 mm from the centre line,
X;= 0.5- 0.135 = 0.365 mm
0.1351 X 0.365
j = I = 1.448
(2 X 18.9 X 1.23 X 10- 3 ) 3
h 2
-• =- arctan(1.37 x 1.948 2 ) = 0.88
hoo 1t
Even in this case only 12% of the film is lost.
(2) Using Example (2) on p. 205, let us suppose that there is a film
adhering to each surface of thickness 1 Jim as they enter the contact. If
we neglect the film thickness and the flattening at the contact, simple
geometry shows that the film will be complete at a distance x from the
centre line, given by x = (2Rh)1, where his the film thickness at the inlet
meniscus - in this case 2 x 10- 3 mm. This is, of course, only a guess at
where the pressure build-up begins.
X=(2 X 2.98 X 2 X 10- 3 )i=0.109mm
THE LUBRICATION OF HIGHLY LOADED CONTACTS 209
b= [6ZWR"]
nkE'
where
z= 1.0003 + 0.5968
Ry/Rx
7.8 TRACTION
~
2
• 1000 rev/min
45kN/m
21kN/m
9·6kN/m
2·8kN/m
0·5 1·0
Slop 2 cu,-u2 1
u,tu2
(a)
W = 44·8 kN/m
Rolling speed
(RS) • U,+U2
2
2500
0·5f-l---t---+---l------+----l
Figure 7.11 Traction curves for rectangular contact: four-disc machine results
(Dowson and Whomes, 1967)
212 TRIBOLOGY
10-4
H=8oo·10- 6
10-5
----------
10-6
:::.
.gto:
10-7
"
10-B
10--9
0 4 8 12 16 20 24 28 32 36 40 44 48
O<W
Figure 7.12 Regimes of traction: oil, HVI 650; steam rollers, R = 20 mm (after
Evans and Johnson, 1986)
4K
ij= 2 [cxW +2bln(U 2 -Ud+bln("·(j)]
M(U 2 - Ud B 2K
where {J is the temperature coefficient of viscosity, K is the thermal
conductivity and 'Is is the viscosity of the lubricant at the temperature of the
surfaces at ambient pressure. This approximation is valid for
.:(U2-Ut)2 1
'lxu »
{JK
The discussion so far has been concerned with the sort of contact encountered
in rolling element bearings, wheels, etc. The situation with cams and gears
can be somewhat different, with the inclusion of large velocities of approach
THE LUBRICATION OF HIGHLY LOADED CONTACTS 213
and other dynamic effects. Some of the film thickness formulae suggested for
these cases are given below.
For involute gears the contact at the pitch line may be modelled as two
rollers, having radii of curvature D 1 sin r/J/2 and D 2 sin r/J/2, with the rollers
rotating at the same angular velocities as the gearwheels themselves.
Holmberg (1982) gives the pitch-line film thickness, hP in microns, for
gears on parallel shafts as
hp = 1.7 X w- 3 ('7o Vp)i[t
where VP is the velocity at the pitch line in m/s, 'lois the viscosity in cP and
l is the distance between centres in mm. In this case a standard value has
been taken of the pressure- viscosity coefficient and the piezoviscous regime
has been assumed.
An earlier approximation is given in The Tribology Handbook as
hp=3.5~
where V. is the entraining velocity in m/ s, R is the relative radius of curvature
on the pitch line in metres and '1 0 is the viscosity in poise.
D 1 D 2 sin rjJ
R = ---------=--
2(D1 + D 2 ) cos 2 rjJ
and
V. = 0.4 V sin rjJ cos u
where rjJ is the normal pressure angle and u is the helix angle.
The minimum film thickness encountered during the contact is hP for
gears of approximately equal size, but reduces to around 50% of hP for large
gear ratios.
Experience suggests that, in order completely to avoid surface distress,
we need a specific film thickness, A., of more than 2. However, the situation
is so complicated by running-in and transient effects that results appear like
those shown in Figure 7.13, from Tallian ( 1967).
7.10 PROBLEMS
8.1 INTRODUCTION
216
BEARING SELECTION 217
- logW
llogW
W =canst.
log V log 1
Strength limit lnert10 Ioree 11m1t lnstablhly 11m1t
(a) (b) (c)
log W log W ~
~
log V log V log V
Thermalllm11 Wear llm1t FotiQUe hmtt
(d) (el (f)
log W log W
HydrostatiC, ,
------~ , ...,
,, , '
,,
, 'HydrodynamiC
log V log V
HydrodynamiC f1lm Hydrostatic and hydrodynamic
ltmlt limit
(g) (h)
Wear limit A permitted wear rate will be defined for the system. As
shown in Chapter 4, the wear rate in many wear mechanisms such as adhesive
or abrasive wear is proportional to the product of load and velocity.
Fatigue limit In some mechanisms, notably rolling element bearings,
it is often fatigue failure which limits the life of the bearing. Practical tests
show that the relationship between load, W, and life, L, in revolutions for a
ball bearing is L oc 1I W3 and L oc 1I wlf for a roller. Since Lis clearly related
to speed, V, we can define the fatigue limit by W3 V =constant for ball
bearings and wlf V = constant for rollers.
Hydrodynamic limit When load is carried by the pressure generated
by hydrodynamic action, the film thickness is some function of '7 VI W, where
'1 is the viscosity. In such cases the operational limit is defined by the need
for a continuous film to be present to carry the load at a given speed using
a particular lubricant. This defines the limit as a line VI W =constant.
However, at higher sliding velocities the heat generated due to viscous
shearing results in a reduction of viscosity of the fluid. This causes a departure
from linearity of the limit at high velocity.
Elastohydrodynamic limit As we have seen in Chapter 7, in the
lubrication of very highly loaded contacts the dependence of film thickness
on load is very slight and so a nearly vertical characteristic is appropriate.
Hydrostatic limit The load capacity of a hydrostatic bearing is
determined entirely by the external pressure available. This leads to a limit
in the form of a horizontal line, with a slight reduction at higher speeds due
to the effect of shear heating on viscosity. However, it is almost inevitable
that, as the velocity increases, some hydrodynamic action will take place and
so the total load capacity will be the sum ofthe hydrodynamic and hydrostatic
effects.
In order to see how the above limits may be combined to produce the
characteristic limit curve for a bearing solution, we consider three journal
bearing types: the dry rubbing bearing based on PTFE compounds, the
rolling contact bearing and the hydrodynamic journal bearing. These
characteristics are shown in Figure 8.2.
If we plot these characteristics for different shaft sizes we obtain a chart
as shown in Figure 8.3. These curves cover the majority of the range of
engineering applications. From the plot we can see clearly the advantages
of rolling bearings at speeds in the range 1000-2000 rev /min, which explains
their use in such situations - e.g. small electric motors. With larger shafts
the enhanced load capacity of hydrodynamic bearings is clearly demonstrated.
This explains why these bearings are so often used in such applications as
BEARING SELECTION 219
In W In W
Max.
speed
limit
Max.
speed
limit
In V In V In V
(rev/m1n) (rev/min) (rev/min}
- - Hydrodynamic bearing
- - - Rolling bearing
- .. - Dry rubbing bearing
-g
0
...J
'-·---
d= 50 mm
. "'-·;--
·,·~
_..-->:( I . )'!
•,.--d= 5mm
Speed
10 7 r---,-----,-~~--~--,-----,----,----,
"- d• 500 mm
105
10
0.01 0.1 10 100 1000 10000
Shalt speed (rev/s)
- - Rolling bearings
- · - · - Dry rubbing bearings
- - - - Hydrodynamic bearings
10 7
106
10 5
~
a:
0
104
..J
103
102
/
/
/ /
10
w-' w-' 10 10 2 104
- - Rolling bearings
· - · - Dry rubbing bearings
- - - - Hydrodynamic bearings
It is most unlikely that the reader will ever be called on to design a rolling
element bearing, but the selection of such a bearing is a common design
task. Each bearing manufacturer produces a guide to the selection and use
of his particular range of bearings. For more than a decade all the major
manufacturers world-wide have conformed to the recommendations of
ISO 281/1:1977 (BS 5512 Pt 1:1977). This has been reviewed recently to
reflect the widespread availability of cleaner steels, but the selection procedures
222 TRIBOLOGY
remain very similar for the major manufacturers. The material in the
discussion that follows is drawn from the RHP data published in the reference
given (RHP, 1977).
The type of bearing to be used will depend on three major criteria: the radial
load, the axial load and the degree of axial location to be provided by the
bearing.
The loads on the bearing may be calculated by normal means, although
help is provided in the manufacturers' guides for certain common situations,
such as shafts with heavy rotors and ~ear trains of different types. The degree
of axial location is, of course, a matter for the designer to decide. The most
common bearing types are shown in Figure 8.6.
R ~ ~
!al (bl (C)
~ (d )
~ ~
!e) (f)
~
(g)
~~
(h) Iii
A (j)
~
(k)
~ (II
(m)
Radialload Low.
Axial load Low.
Axial location One direction only.
Comments The bearing races are detachable and interchangeable. They are
often used in pairs, one bearing being adjusted against the other.
Comments The rollers have a high length/diameter ratio. They are often
used in situations where there is not sufficient space for the equivalent ball
or roller bearing.
Having decided on the bearing type, the size must be determined. For each
bearing the manufacturer will quote:
C - the dynamic load capacity of the bearing
C0 - the static load capacity of the bearing
Each of these must be checked against the operating conditions.
Dynamic Capacity
The dynamic failure of rolling bearings takes place through various fatigue
mechanisms. It is in the nature of fatigue that the life of samples shows a
large scatter and can only be expressed statistically. The 'life' of the bearing
is usually defined by the L 1 0 life, which is the number of millions of revolutions
that 90% of the bearings are expected to exceed before failure. If P is the
dynamic equivalent load, which is in most cases the applied load while
running (but check with the manufacturer's recommendations), the life is
r
given by
16667(c)-'
L 10 = -n- p f- hours for roller beanngs
.
Modifications to this life should be made to get a more realistic life, L, by
L=a 1 a 2 a 3 L 10
Reliability factor a 1 There may be some applications where 90%
reliability is not good enough. In this case we introduce the factor a 1 , as given
226 TRIBOLOGY
90 1.0
95 0.62
96 0.53
97 0.44
98 0.33
99 0.21
n = speed (rev/s)
d = bearing bore+ o/d (mm)
m 2
104 ~-------------L--~-L-L----~-----U----~
10 10
8 8
~ ~
.J .J
v,
4
v,
4
+
3 3
3 4 5 6 8 10 3 4 5 6 8 10
V, v,
Figure 8.8 Viscosity ratio at operating temperature
3.5
3.0
V = actual lubricant viscosity {eSt)
1 recommended viscosity (CSt)
2.5 at the operating temperature
2.0
;::
~
1.5
1.0
0.5
0
0.1 0.5 1.0 5.0 10.0
v,
Figure 8.9 Material and lubrication life adjustment factor, a 2 •3
228 TRIBOLOGY
Static Loading
The manufacturers' tables give a static load capacity C0 . This is the load
which, when applied to the stationary bearing, produces a permanent
deflection of 0.0001 of the rolling element diameter. This is known as
brinelling. For satisfactory performance the maximum equivalent static load,
which is in most cases the actual value of the static load, must not exceed C 0 .
Other Considerations
The preceding data assume that the bearing is installed correctly, with the
appropriate tolerances, and is adequately lubricated. Instructions on fitting
and lubrication are to be found in the bearing manufacturers' catalogues
and technical literature, together with maintenance schedules. Additional
information is to be found in The Tribology Handbook (Neale, 1973), which
includes an interesting section on the identification of bearing failures. This
is also dealt with in Section 9.9.
8.6.3 Example
Determine the life to give 97% reliability for the deep-groove ball bearing
described below.
8.7 PROBLEMS
9.1 INTRODUCTION
In the preceding chapters we have seen the beneficial effect of the presence
of the lubricant on the bearing performance. The lubricant has three main
functions: ( 1) to provide a coherent film between the surfaces; (2) to remove
heat from the contact; ( 3) to prevent the ingress of dirt and other contaminants.
The following sections deal with the volume and means of lubricant supply,
the monitoring of the system and the lubricant, and the likely results of poor
performance in this area.
230
Table 9.1
fo
Z = (P 0 / C 0 ). Lower values are for light series, higher for heavy series. -<
(/)
--1
m
s::
(/)
N
~
232 TRIBOLOGY
This section is confined to the supply of grease and oil to rolling bearings.
These form the vast majority of lubricants used today. Which is used will
depend upon the circumstances. Grease is good for keeping out contamination,
operating in any attitude, and for low speeds, and is extensively used in
rolling element bearings, especially in sealed units where continuous lubricant
supply is not possible or is undesirable. Oil lubrication is suitable for heat
removal and for hydrodynamic bearings operating at higher speeds.
"'tlc
~
c
8
c 106
0
·~ 10 3
107
~w
10"
~
"'
105
10 2 L----------------L--------~L---~
10' 105
dmn
(mm x rev/s)
For rolling element bearings the initial charge of grease should reach
the working surfaces, but the housing should not be overcharged as this can
cause excessive churning and high temperatures.
The life of the bearing may be limited by the degradation of the grease
unless it is replenished occasionally. The interval between regreasing is given
by RHP ( 1977), T = K 6 R h, where K 6 is the grease lubrication constant
derived from Figure 9.1 and R is the relubrication factor from Table 9.1.
There are many methods of supplying oil to a bearing. The most common
are illustrated in Figure 9.2. Any of these methods can be found in use for
rolling bearings, but for plain journal bearings and sliders, oil mists, capillaries
and ring oilers are unlikely to provide sufficient flow rate.
Pumped Supply
In this case the oil is supplied to the bearing by a pump under pressure. The
flow rate can be as large as the pump can produce and therefore the heat
removal can be very effective. The capital cost, maintenance costs and initial
filling cost can be high and, of course, energy is required to operate the
pump. However, where a large, reliable flow is required, this system is most
suitable.
234 TRIBOLOGY
Pumped
8
Oil mist
i
g
Capillary
Gravity
Ring oiler
Dipping
Gravity Supply
Capillary Systems
These systems use a wick or pad of felt in which the capillary action draws
the fluid from a reservoir on to the bearing surfaces. The oil flow increases
LUBRICATING SYSTEMS 235
as the cross-sectional area of the wick, and decreases as the length of the
wick increases and with viscosity increase. The design and construction of
such systems is very simple and the costs small. However, the supply rate is
quite limited and is only sufficient for lightly loaded bearings.
Dipping Systems
9.4 FILTRATION
VD
Re=-
v
D being usually taken as the diameter of the filter inlet port. K F is a function
of the filter construction and is determined by the Kozeny -Carman relation:
e is the void fraction or porosity, Sis the specific surface of solid/unit volume
(m 2 jm 3 ) and k is a constant for the filter material (for cellulose k = 5.55).
The relationship between flow rate and pressure drop is shown for a
typical case in Figure 9.4, which represents the results from tests to BS 6277.
This graph is for the case where the filter is equipped with a bypass which
opens when the pressure drop through the filter reaches a predetermined
value. This could happen if the filter were blocked.
As the pressure drop increases with Q, if the flow rate is high, it is often
advantageous to use bypass filtration. In this case the filter is connected in
parallel with the bearing system and the pressure differentials are arranged
so that only a small fraction of the lubricant passes through the filter on
each circuit. This means that not all the fluid is cleaned on every pass, but
filtration is still quite efficient.
A more common arrangement is a combined full-flow /bypass filter,
where all the lubricant is filtered to some degree, but a small proportion still
goes through a bypass to remove finer contamination. Cartridges including
both types of filtration in parallel in one element are available.
Since most filters operate by trapping particles within the medium, there
will come a time when the trapped material will seriously interfere with the
performance. With fibrous filters, where filtration operates through the depth
of the element, this will be noticed as a reduction in the degree of filtration
and the possible release of previously collected material due to pressure
surges. For a surface filter, which operates by trapping the particles in
well-defined openings on the surface, trouble is indicated by a decrease in
flow rate or increase in pressure drop. In either case the filter must be serviced,
by either replacing the element or cleaning it in a solvent, by steam backwash
or ultrasound.
238 TRIBOLOGY
1.4
1.2
~ 1.0
e
Q_
'0
~ 0.8
~
6::
0.6
0.4
0.2
0 10 20 30 40 50 60 70 80
Figure 9.4 Pressure drop plotted against flow rate for a filter with integral bypass
In this procedure a small sample of oil, typically 0.1 ml, is vaporised, and its
atomic constitution is determined both qualitatively and quantitatively by
spectroscopic analysis. Three types of analysis are in common use: differential
infrared spectroscopy; atomic emission spectroscopy; and atomic absorption
spectroscopy. All give essentially the same information.
SOAP gives the total concentrations of the elements in the sample but
it gives no information on whether the atoms were previously combined in
a chemical compound or an alloy. Furthermore, it does not give any
information on the sizes or shapes of th~ particles, and it detects particles
only in the size range 0-3 Jlm. Its major advantage lies in the detection and
measurement of wear debris produced by corrosive or oxidative wear, which
is normally of submicron size.
9.6.2 Ferrography
into which the portable transducer can be fitted reproducibly. In this type
of monitoring, the results are usually recorded and plotted manually, and
there is an interval of some hours, or even days, between the measurements
being taken and any changes being noted and acted upon.
At the second level of sophistication, transducers are permanently
mounted at critical points on machinery and the output of each transducer
is monitored continuously. This output is then used in two ways. First, alarm
levels are set on each output so that immediate warning will be given if any
potentially dangerous vibration develops; second, the output levels are
sampled at frequent intervals and trends identified by the data acquisition
and processing system.
At the most sophisticated level, vibration levels are both recorded and
trended continuously so that at any time the maintenance engineer has at
his disposal a fully up-to-date display of the machine condition.
Dirt can be introduced into the bearing either during assembly or with the
lubricant during running. This will be shown by scoring and scratching of
the surface in the direction of motion. Larger particles introduced during
assembly can cause localised overheating and distortion.
242 TRIBOLOGY
Any of these will cause the film thickness to be too small for satisfactory
performance. Severe damage to the bearing may result, with surface melting
of the bearing material, especially if a low-melting-point alloy is used, and
smearing or wiping of the surface.
This will permit small oscillatory movements between the surfaces of the
bearing or bearing and housing. This is a common cause of fatigue failure
and is characterised by the production and oxidation of fine debris, usually
reddish-brown in colour for ferrous metals, with welding or pick-up of
material at the joint between the housing and bearing shell.
Fatigue Failure
Corrosion
Fatigue Failure
Insufficient Interference
This causes fretting between the race and the housing, with the production
of the characteristic reddish-brown debris. It is very important to get right
the diametral interference of raceways, as axial clamping is an inadequate
substitute.
This will cause the normal track of dulled material on the raceways to be
skewed. It does not necessarily lead to immediate failure, but can be damaging
if allowed to continue. In cylindrical roller journal bearings this can lead to
excessive loading on the flange, which has only very limited axial load
capacity.
Hard particles will be rolled into the bearing surface to produce dents. Grit
will be crushed to powder, making the raceways become frosted in appeatance.
Softer dirt will produce marking of the tracks and may interfere with the
free running of the elements. This will produce noise and possible wear.
Since raceways are an interference fit on the shaft and in the housings, there
is a great temptation to abuse them by overenthusiastic use of hammers.
This can dent the races or the rolling elements and in extreme cases fracture
the flanges. This damage is often visible after the bearing has run.
Inadequate Lubrication
bearings the cage pockets and rims may be worn and the grease will be hard
and dry.
Brine/ling
Dents may be present in the rolling tracks, which conform to the shape of
the rolling elements. This is symptomatic of a sudden overload being applied
to the bearing. This can happen during assembly, in which case the grinding
marks will remain visible or when the bearing is stationary (false brinelling),
which will obliterate the grinding marks.
Corrosion
9.8.3 Gears
Mechanical Failure
The most dramatic failure is when a tooth breaks off. The obvious caQse of
this is that the tooth load produces bending stresses at the tooth root in
excess of the tensile or fatigue strength of the material. For brittle materials
a sudden shock load may cause the failure. The tooth may also be weakened
by other forms of surface damage.
Surface Fatigue
This is caused by the load on the tooth being too high. Pits form in the
contact zone which may join together to affect large areas of the tooth.
Scuffing
245
246 REFERENCES
Sasada, T., Norose, S. and Mishina, H. (1979). Proc. Int. Corif. Wear Materials,
Dearborn, MI, ASME, p. 72
Savage, R. H. ( 1948 ). J. Applied Physics, 19, 1
Schmitt, G. F. (1980). In The Wear Control Handbook, New York, ASME
Scott, D., Seifert, W. W. and Westcott, V. C. (1975). Wear, 34, 251
Sherbiney, M. A. and Halling, J. ( 1976). Wear, 40, 325
SKF (1974). Rolling Bearing Damage: A Morphological Atlas, King of Prussia, Pa.,
The Centre
SKF ( 1984 ). Principles of Bearing Selection and Application, King of Prussia, Pa.,
The Centre
Soda, N., Kimura, Y. and Tanaka, A. (1977). Wear, 43(2), 165
Stribeck, R. (1901), Z. ver. D. I., 45(3), 73-125
Suh, N. P. (1973). Wear, 25, 111
Suh, N. P. (1977). Wear, 44, 16
Tabor, D. (1985). In Loomis, W. R. (ed.), New Directions in Lubrication, Materials,
Wear and Surface Interactions, New Jersey, Noyes Publications, p. 2
Tallian, T. E. (1967). Trans. ASLE, 10, 418
Tavernelli, J. F. and Coffin, L. F. (1959). 1rans. Am. Soc. Metals, 51, 438
Teer, D. G. and Arnell, R. D. ( 1985). In Recent Developments in Surface Coating and
Modification Processes, London, Mechanical Engineering Publications, p. 21
Walls, J. M. (ed.) (1989). Methods of Surface Science, Cambridge, CUP
Waterhouse, R. B. ( 1988 ). Fretting Corrosion, Oxford, Pergamon Press
Weber, C. and Sallfeld, K. (1954). Z. angew. Math. Mech., 34, Nos. 1-2, 54
Wells, R. M. (1967). Proc. I. Mech. E., 182, Pt. 3A, 443
Wilson, R. W. (1975). Proc. Eurotrib. 1, J. Mech. Eng., 165
Woodruff, D.P. and Delchar, T. A. (1988). Modern Techniques of Surface Science,
Cambridge, CUP
Woolacot, R. G. (1965). National Engineering Laboratory Report No. 194, East
Kilbride, NEL
Woolacot, R. G. and Macrae, D. (1967). National Engineering Laboratory Report
No. 315, East Kilbride, NEL
Wymer, D. G. and Cameron, A. (1974). Proc. I. Mech. E., 188, 221
Zum Gahr, K. H. (1987). In Microstructure and Wear of Materials, Amsterdam,
Elsevier, Chapt. 5
Index
249
250 INDEX