Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Home Search Collections Journals About Contact us My IOPscience

Friction, lubrication and wear: a survey of work during the last decade

This content has been downloaded from IOPscience. Please scroll down to see the full text.

1966 Br. J. Appl. Phys. 17 1521

(http://iopscience.iop.org/0508-3443/17/12/301)

View the table of contents for this issue, or go to the journal homepage for more

Download details:

IP Address: 134.151.40.2
This content was downloaded on 29/08/2014 at 18:31

Please note that terms and conditions apply.


BRIT. J. APPL. PHYS., 1966, VOL. 17

lubrication an wear : a survey 0%work


e last deca
F. P. BOWDEN and D. TABOR
Surface Physics, Cavendish Laboratory, Cambridge
M.S. received 20th July 1966
Abstract. This article describes the progress that has been made during the last
decade in our understanding of the processes involved in friction, lubrication and wear.
It does not pretend to be an encyclopedicsurvey-that would be a bore-but summarizes
work that seems to us to be of greatest significance. The first section deals primarily
with the broader issues underlying the friction of unlubricated surfaces. The second
section deals with specific items of research, many of which have opened up new lines
of attack. The third, for completeness, deals with lubrication and wear, but this part
is brief; it merits a full article in its own right.

1. Basic mechanisms of friction


1 .1. IIztroduction
During the last two or three decades most workers in the field have recognized that
friction between unlubricated surfaces arises from two main factors (e.g. Bowden and
Tabor 1950, 1964). The first, and usually the more important, factor is the adhesion which
occurs at the regions of real contact: these adhesions, welds or junctions have to be sheared
if sliding is to occur. Consequently, if A is the true area of contact and s the average shear
strength of the junctions, this part of the friction may be written Faahesion = As. The second
factor arises from the ploughing, grooving or cracking of one surface by asperities on the
other. We may call this the deformation term P. Then if there is negligible interaction
between these two processes we may add them and write
F = Fadhesion T Fdeformation = As $- P. (1)
It is clear that the factors of major importance here are the area of real contact A , the strength
of adhesion between the surfaces, the shear strength s of the interface, possible interactions
between A and s, the deformation component P and possible interactions between this and
the adhesion component of friction. I n all these areas considerable work has been carried
out during the last decade and in some cases new and useful ideas have been introduced.

1.2. Area of contact: analjtical models


The two basic laws of friction state that the frictional force is proportional to the load
and independent of the size of the bodies. These laws hold approximately for a wide
range of materials and over a wide range of experimental conditions. If the friction is due
primarily to adhesion it implies that the area of true contact A should be proportional t o the
load and independent of the size of the bodies. Such a view seemed completely untenable
to the early workers (especially Coulomb, see e.g. Coulomb 178.5) and for this reason they
rejected the adhesion mechanism. They attributed friction primarily to the dragging of one
surface up the asperities on the other. If the average slope is B the coefficient of friction is
then ,U = tan 0 and the load and size of the bodies do not come into the picture. Of course,
as Leslie pointed out in 1804, such a mechanism is non-dissipative since the sliding body, on
the average, remains at a constant horizontal level. The problem was resolved, at least
for metal surfaces, twenty or thirty years ago in the following way (Bowden and Tabor 1950,
Holm 1958). Metal surfaces are never smooth on an atomic scale. When placed in
contact the tips of the asperities are deformed, zt first elastically; for loads exceeding more
12 1521
I522 F. P. Bowden and D.Tabor.
than a minute value the elastic limit is exceeded and plastic flow occurs. It turns out that
the local plastic yield pressurep is very nearly constant and is comparable to the indentation
hardness of the metal. Under these conditions the area of contact for any one asperity
bearing a load w1 is A , = wJp so that for an assembly of asperities the total area is

where W is the total load. The area is thus proportional to the load and independent of
the size or smoothness of the surfaces. This conclusion, based on the idea that the contact
pressures over the individual asperities are sufficient to produce local plastic flow, was
supported by measurements of electrical contact resistance (Bowden and Tabor 1950,
Holm 1958). Further, it was in harmony with the observation that the sliding of metals is
generally accompanied by plastic deformation at the interface and by tearing and shearing
of the metals.
During the last decade or so a great deal of interest has been shown in other models
involving other types of deformation, particularly elastic deformation. Archard (I 95 1)
considered for example the contact between surfaces covered with asperities of spherical
shape. The two extreme types of deformation are purely plastic and purely elastic. For
purely plastic deformation the area of contact is proportional to the load, as we saw above.
For purely elastic deformation the area of each contact will be proportional to 1 1 . ~ ’ ~ . There
are now two distinct ways in which elastic deformation can occur:
(i) The number of asperity contacts remains constant so that an increase in load increases
the elastic deformation of each contact. The area of real contact is then proportional
to w213.
(ii) The average area of each deformed asperity remains constant and increasing the load
increases the number of regions of contact proportionally. Clearly the real area of contact
is directly proportional to W.
Archard concluded that in any real situation where elastic deformation occurs the area of
contact will be proportiozal to VP” where ?7z lies between $ and 1. This is supported by a
more detailed analysis. For example, Lodge and Howell (1954) considered a spherical
surface (radius R ) covered with a close-packed array ( n per cmz) of hemispheres of radius
v. For elastic deformation (see figure 1) the true area of contact turns out to be
A = ,7.+/3nl/3R2/9 W8/9. (3)
If the asperities are covered with still smaller asperities then, as Archard has shown? A
becomes proportional to W26:2i,Thus a multiple asperity model involving purely elastic

W
1

k
- 20-

Figure 1. Lodge and Howell’s model of a spherical surface of radius R covered with hemispherical
asperities of radius Y (n asperities per cm2)when pressed on to a hard flat surface with a normal load
W. If the deformation of the upper surface is elastic the pressure distribution (inset) is Hertzian
an6 the area of contact is given by equation (3).
Friction,iubricatioii and Itsear: a survey of work duviiig the last decade 1523
deformation tends to give an area almost linearly proportional to the load. Such a model
would not give a power of Wgreater than unity. However, Malgaard (1962) has shown that
such a situation is possible over a limited load range if the contour of the surfaces has a
suitable geometry. The contour is such that, in effect, there is a collapse from one asperity-
covered level to a neighbouring level as the load is increased.
These models all assume purely elastic deformation. A much more penetrating analytical
study has recently been given by Greenwood and Williamson (1966). First they take as their
model surfaces which have asperities closely resembling those found on real surfaces (the
levels usually show a Gaussian distribution). They then consider how these will deform if
there is a finite elastic limit. They show that for very smooth contours and modest loads
the asperities may still be within the elastic limit of the material. (A similar conclusion was
described by Blok (1952) for surfaces with grooves in one direction.) For rougher surfaces
or higher loads plastic deformation gradually takes over. In both situations the area of
rea! contact is very nearly proportional to the load. They conclude that a linear increase of
area of contact with the load is not proof that plastic deformation is occurring. Another
very interesting point concerns the relation between line contour traces (profilometer
records) and areal calculations. If a line contour shows that say 10% of the surface is
above a given level, what is the area of contact if the surfaces are deformed to that level?
Kraghelsky and Demkin (1960) suggest that it will be 10 % of IO%, i.e. 1 %, of the geometric
contact area. Wi!!iamson and Greenwood show that this is not so. For perfectly random
surface features the 10 % figure must apply whatever the direction of the contour line. This
can be true only if the contact area itself at this level is 10 <
; of the nomina! geometric area.
These papers cover work similar to that described by Dyachenko e f al. (1964). but the
model is physically more realistic and the treatment is less heavily burdened with
mathematics.
These papers assume either elastic or plastic deformation. With polymers neither
approach is valid. They are viscoelastic materials and the deformation is time dependent as
well as being dependent on load and geometry. However, over a fairly wide range the
contact area between a hard sphere and a smooth flat specimen of polymer (or between a
smooth sphere of polymer and hard fiat surface) is given by A = W m where n7 is generally
between 0 . 7 and 0.8. It is possible, by making a few reasonable assumptions (Cohen and
Tabor 1966), to incorporate this behaviour into the Lodge and Howell model described
above. The area of contact A is then given by

It is seen that for ni = 5 (elastic case) this reduces to the Lodge and Howell equation.
For H I = 1 it reduces to an area of contact independent of R and proportional to W :
A = k( WIE). (5)
This corresponds to plastic deformation, and E corresponds to the hardness or yield pressure
of the solid.

1.3. Area of Confact: experimental


With metals some idea of the area of contact can be deduced using electrical resistance
methods (Bowden and Tabor 1950, Holm 1958). The main uncertainty arises from the
fact that the electrical conductance of a junction is proportional to the diameter of the
junction-not to its area. With non-metals the only other available experimental approach
is the optical one. If the surfaces are molecularly smooth the optical area deterinined by
interference methods for example will be identical with the true area of contact. The only
material for which this seems to be applicable is mica (Bailey and Courtney-Pratt 1955).
For other surfaces there is always some uncertainty as to the difference between the optical
and real areas of contact. Adams (1963) has carried out a very detailed study of the contact
between a 'smooth' hemisphere of nylon and a fiat glass surface using interference. He
1524 F, P.Bowden and D.Tabor
finds that A is proportional to W m where Pn = 0.70. Further, the area of contact does not
change when sliding occurs, i.e. there is no junction growth (see below) during sliding.
Other optical methods include phase contrast (Dyson and Hirst 1954) and total internal
reflection. Kraghelsky and Sabelnikov (1957) have used this method to study the area of
contact and the friction between polymers and other transparent solids. Their results with
silver chloride specimens are puzzling: they find that the area of contact increases more
rapidly than the load. This implies (as Mdgaard’s analysis shows) a rather unlikely type of
surface contour. It is possible that the difficulty arises from the difference between the
true area of contact and the optical area deduced by this method.
It is evident that more criiicai techniques for studying the true area of contact could be
of great value in friction studies. Some attempt has been made to use thermal conductivity
in a manner similar to that involved in electrical contact-resistance measurements, but it is of
limited value.

1 .4. Interaction of A and s


We first consider the behaviour of metals. According to the simple adhesion theory of
friction the coefficient of friction should be given by
F As s
p=-=-=-
w AP P‘
If the metal does not work harden appreciably the shear strength of the interface is roughly
equal to the critical shear stress T of the metal. The yield pressure or hardness p is generally
found to be about 57 (Tabor 1951); consequently p N 0.2. In practice most metals in air
give a value of p of about 1. The reason for this has been revealed by the experiments of
Courtney-Pratt and Eisner (1957). They studied the contact between a hemispherical
slider and a flat surface where, because of the geometry, the static contact is essentially a
single circular region. This could be determined optically or with electrical resistance
measurements. For platinum surfaces where oxide contaminants are not present there is
very good agreement between the two determinations. They then applied a tangential
force, gradually increasing it until gross sliding occurred. They then found that during the
initial microdisplacement of the slider there was a large reduction in the contact resistance.
By the time gross sliding took place the area of contact had increased three or four fold
(see figure 2). This is because the plastic yielding of the junction is determined by the

0 10 1.5 20
A /A0
Figure 2. The behaviour of platinum surfaces in contact when subjected to a gradually increasing
tangential force. Ordinate C is the ratio of the tangential force F to the normal load W. (It
corresponds to the coefficientof friction only when sliding actually occurs.) Abscissa is the area of
contact A expressed as a ratio of the initial static area Ao. The ratio A/Ao is calculated directly
from the electrical resistance measurements, assuming metallic contact. 0 clean surfaces,
@ lubricated surfaces. The full curve is the theoretical curve plotted according to equation (7)
using a value of 12 for a.
Friction, lubrication and wear: a survey of work during the last decade 1525
combined effect of the normal and the tangential stress. If the normal stress on the
junction is p and the tangential stress is s, the yield criterion is of the form
p2 + as2 = po2 (7)
where a is a constant with a value of about 10 andp, is the static contact pressure (McFarlane
and Tabor 1950). When the normal load is first applied the contact region flows plastically
under the normal load itself (s = 0) and p = p o . As soon as a small tangential stress is
applied the equality in equation (7) can be satisfied only if p diminishes. This is brought
about by the surfaces sinking together (see figure 3) so that the area of contact is increased
with a corresponding drop in p . This process continues indefinitely if the surfaces are
perfectly clean so that a very high coefficient of friction is attained ( p = 10 or more). If,
however, as Tabor (1959a) pointed out there is some surface contamination so that the
1

Microd!splacement

X Y Horizontal
(U) (b)

Figure 3. ERect of combined normal and tangential stresses on the behaviour of a soft asperity in
contact with a hard flat surface. (a) Under a normal load alone, contact occurs across XY; when
a tangential stress s is applied the junction grows as shown by the broken curve and the point Qo
moved to Q. (b) Schematic diagram showing locus of Q as a tangential stress is increased.

shear strength of rhe interface is a little less than 7, junction growth ceases at an earlier
stage and the friction has a smaller value. The way in which the coefficient of friction
depends on the strength of the interface is shown schematically in figure 4. It is seen that
for an interface with a shear strength about 95 % of 'T slip occurs for a value of p = 1, and
the area of contact is about three times the static value. These results show that with
ductile materials such as metals (see also Rubenstein 1958-9a, Spurr 1961, 1964, 1965) the

3l- /

:21
0
I

A IA,

Figure 4. Growth in area of junctions for different values of the interfacial shear strength s, where
s is expressed in terms of the critical shear strength t of the metal (s = kz). As the tangential force
coefficient 4 (= F/W)is increased, junction growth proceeds until at some critical stage determined
by the value of k gross slip occurs. For k = 1 junction growth proceeds indefinitely; for k = 0.95
slip occurs when the area of contact has grown about three-fold and the force coefficient (now equal
to the coefficient of friction) has a value of about 1.
1526 F. P. Bo,t.deu arid D.Tabor
yielding of the junctions depends both on the normal stress and on the tangential stress
itself. They cannot be treated as independent parameters.
This conclusion has a direct bearing on the contact studies described above which suggest
that metal surfaces may deform elastically if the surface geometry is suitable. Although
this is possible under conditions of normal loading it is clear that a small amount of tangen-
tial stress may lead readily to plastic deformation of the interface. Consequently elastic
deformation is possible under conditions where tangential stresses are small as in static
contact, rolling contact or lubricated sliding. It is less likely to be possible if there is strong
interfacial adhesion. Under these conditions: as we have already seen, appreciable junction
growth and plastic distortion can occur. In some sliding systems, as the more detailed
studies of Cocks (1958, 1965, 1966) and Antler (1963) S ~ O W this , can lead to the formation
of a large wedge of welded metal which moves with one of the surfaces over the other until
at some stage it becomes detached and a new wedge begins to form. The shape and shear
pattern of these wedges closely resemble those described in earlier work by Greenwood and
Tabor (1957) on the shearing of model junctions. This work illustrates the general
phenomena cf junction growth and junction failure, but the failure criterion appears to be
far more complicated than that suggested in figure 4; more work on this is desirable. How-
ever, although the behaviour is not understood in detail it is clear that surfaces which rub
in this way are repeatedly roughened by the sliding process; this favours plastic deformation
at the contact region even after prolonged rubbing.
In some cases, particularly if adhesioii is weak, repeated traversals may lead to Eattening
and smoothing of the surfaces; this will favour elastic rather than plastic deformation in the
contact zone. Midgley (1963) has shown that such a situation can, for example, be achieved
in the sliding of low-graphitic carbon. As the surfaces grow smoother the area of true
contact, resulting from elastic deformation, gradually increases with an accompanying rise
in friction. In this regime the friction is high although the specific tangential stress in the
interface remains small; the wear is extremely low until at some crucial stage the surface
layers fragment, probably as a result of fatigue. Earlier work by Tabor (1948) showed that
if metal surfaces are first loaded plastically and then unloaded the recovery is primarily
elastic. Archard (1957) carried out friction experiments under similar conditions and
showed that during the unloading part of the cycie the friction varied in a manner that
corresponded to elastic deformation. It is significant, however, that in Archard’s experi-
ments the surfaces were deliberately lubricated so that the interfacial shear stress was
always small. In a similar type of experiment O’Connor and Johnson (1963) showed that
surface asperities may be deformed plastically when initially stressed but that cyclic stressing
leads to elastic deformation. In this work and in the analytical model used by Greenwood
and Williamson (1966) the asperity contact pressure, even in the elastic range, is not vastly
less than that which occurs under plastic conditions. It seems reasonably safe to conclude
that the real contact pressure for both elastic and plastic contacts, over a wide range of
conditions, will lie between say +p and p where p is the local plastic yield pressure.
We now consider the interaction of A and s in the case of polymers. The most thorough
study is due to Adams (1963). Using optical interference methods he found that the
optical contact area is the same during sliding as under static conditions. This implies that
the normal stress and the tangential stress can be treated as independent parameters.
Consequently, if the area varies as Wm we should expect the friction to vary in the same way.
However, Adams finds that the area varies as whereas the friction varies as
Detailed considerations show that this difference is significant and that it is due to the fact
that the shear strength of the interface increases slightly with pressure. Kraghelsky and
Sabelnikov (1957) also deduce an increase of s with pressure, but the effect seems far too
large to be acceptable. As mentioned above, this is probably caused by the difference
between the optical and real areas of contact. Certainly most other workers in the field
find that i f s is treated as a constant there is a fairly close parallelism between the way in which
both A and F vary with load and with the geometry of the surfaces (Lodge and Howell
1954, Pascoe and Tabor 1956, Rubenstein 1958-9b, Cohen and Tabor 1966). Future
work will, however, need to consider carefully the possible effect of pressure on s.
Friction, lubrication and wear: a survey of work during the last decade 152’7
Finally we refer briefly to the interaction of A and s with brittle solids. The frictional
behaviour resembles that of metals since the hydrostatic stresses around the regions of
contact inhibit cracking and the deformation is often dominantly plastic. If the material
has some ductility (e.g. rock-salt) junction growth will occur as for metals (King and Tabor
1954). However, there is a limit to this. As the area of contact increases the normal
pressure falls off and the ability of the hydrostatic stress to inhibit brittle failure diminishes.
The properties are shifted towards the brittle range and further junction growth becomes
impossible. Even if the surfaces are rigorously cleaned the friction never exceeds
p N 0.7-1. We shall discuss the fracture and deformation properties in greater detail
below.

1 .5. Interaction of normal and tangential stress$elds


In the previous section we considered how the strength properties of the interfacial
juxtions are affected by the interaction of the normal and tangential stresses. There is,
of course, an allied problem of a slightly different nature, namely, how the stress fields
interact without regard to the frictional mechanism itself. Consider, for example, a sphere
loaded on to a hard surface under conditions where the overall deformation is assumed to
be elastic. If a tangential stress is applied, what happens at the contact region before
gross slip occurs? A theoretical analysis by Mindlin (1949) shows that the shear stresses
around the periphery of the contact region readily exceed the normal stress mukiplied by
the coefficient of friction. Consequently, when a modest tangential stress is applied
‘slip’ occurs over the peripheral band although the surfaces remain ‘locked’ together in the
central region. The larger the tangential force the wider the band, until for a tangential
force equal to the limiting friction the band fills the whole contact area and gross slip occurs.
This analysis, as already pointed out, is based on purely elastic deformation, but Johnson
(1955, 1961) has shown experimentally that it holds extremely well even though the
asperities within the elastic contact zone are plastically deformed. The explanation, in a
sentence, is that the individual junctions are sheared sequentially from the edge of the
contact region inwards, at a rate determined by the bulk strains in the elastic ‘hinterland’;
consequently the overall bulk tangential-force-tangential-displacementrelation remains
essentially the same as that predicted by the ideal elasticity solution of Mindlin. Apart
from its theoretical interest this work has a direct bearing on the mechanism of fretting.
In a later paper Hamilton and Goodman (1966) have analysed more generally the problem
of superposing a tangential surface traction on the Hertzian distribution for normally
loaded elastic surfaces. Their calculations show that the tangential traction can greatly
modify the distribution of shear stresses and tensile stresses within the solid. The latter
observation is particular!y important in considering the sliding of brittle solids where
fracture and cracking are readily initiated by tensile stresses, espectially those operating in
the free surface. It turns out that quite small tangential tractions can produce far larger
surface tensile stresses than those arising from the normal load itself. Experiments do,
in fact, confirm that tangential stresses play a vital part in generating fracture during the
sliding of brittle solids (Billinghurst et al. 1966).

1 .6. Adliesion of sugaces


The strength of the interface depends crucially on the adhesion between the surfaces.
During the last decade there has been a vcluminous literature on adhesion and adhesives,
much of which is summarized in books by Weiss (1962), Eley (1962) and Houwink and
Salomon (1963). We shall refer only to those aspects which are relevant to the frictional
mechanisms. For two perfectly clean specimens of the same metal (e.g. clean gold) the
adhesion must necessarily be very strong and the interface will have a strength comparable
to that of the metal itself. For dissimilar metals the adhesion is more difficult to specify.
It has been suggested that metal pairs which are mutually soluble will be of the right atomic
dimensions and electronic configuration to give strong adhesion unless the compounds
they form are basically brittle, and that insoluble pairs should give poor adhesions (Roach
1528 F. P. Bowden and D.Tabor
et al. 1955). However Anderson et al. (1957) have found very strong adhesion between
gold and germanium, Moore and Tabor (1952) found very strong adhesion between indium
and diamond, whilst Machlin and Yankee (1954) reported strong welding between metals
which are mutually insoluble, Recently Spalvins and Keller (1962) have found that the
very clean surfaces of many metals will stick strongly in a high vacuum but that certain
pairs such as vanadium-iron and iron-silver show very small adhesion. They attribute this
to unfavourable surface energy conditions, but the relation between surface energy of solids
and strength properties of the interface is not yet satisfactorily resolved. Further, some
friction experiments by C. A. Brookes (private communications) show strong adhesion and
metallic transfer for iron sliding on vanadium. It may however be noted that in a practical
mechanism, in the absence of deliberate lubrication, it has been found that iron and silver
constitute a very good anti-wear combination (Qsborn 1964). On the other hand de Gee
(1965) finds that the wear of iron and silver is crucially dependent on the atmosphere.
More work on this is clearly desirable. Another factor involved in the adhesion of clean
metals is the mutual crystallographic orientation of the clean surfaces. Semenoff (1958,
1961) suggests that if there is marked mis-match the adhesion will be small; the adhesion
can, in his view, be increased by providing energy at the interface, either thermally or by
the work of plastic defamation. This idea too merits further study.
Anderson (1960) and Sikorski (1964) have studied the adhesion between metals in air.
If heavy deformation is imposed on the surfaces to break up surface films, strong adhesions
c21; be obtained with rilaiiy combinations. Soiiie pairs such as iron-vanadium gave poor
adhesion (see above). Work by Rowe suggests that practically all metal combinations can
give strong adhesion: if strong adhesion is not observed in practice he suggests that this is
due either to contaminant films or to the effect of elastic stresses which are released as the
joining force is removed. This would be consistent with the observation that many
practical adhesices fail not because the adhesion is poor but because stress concentrations
in the glue-line provide a region of weakness (de Bruyne 1956).
Strong adhesion can occur between non-metals. The most striking example is that of
freshly cleaved mica : two such surfaces placed in contact have an interfacial strength almost
as great as that of the originai marerial. Contaminant films of air or water vapour greatly
reduce the adhesion (Bailey and Courtney-Pratt 1955) and if a long-chain organic molecule
is adsorbed on the surfaces there is a further large reduction in adhesion (Bailey and Kay
1967). Many other non-metals adhere when placed in contact under pressure, for example
two pieces of rock-salt can be pressed together at room temperature to form a very strong
bond (King and Tabor 1954). Polymers can adhere strongly under pressure, particularly
if the temperature is sufficient to encourage diffusion (Voyutskii 1963). All these results
are consistent with the observation that in the sliding of clean surfaces there is generally
transfer of material from one surface to the other, showing that the interface is at least as
strong as one of the sliding pair. This applies to a wide range of materials and indicates
that the adhesion component of friction can be estimated on the assumption that s is of the
same order of magnitude as 7.
There are two interesting exceptions. Rubber specimens may be slid over hard smooth
surfaces without any sign of transfer of rubber even if the most refined techniques for study-
ing transfer are used (Grosch 1963). Presumably the interface is weaker than the rubber
itself (see below). Secondly the friction of polytetrafluoroethylene (PTFE;Teflon or fluon)
is very low ( p < 0.1) and corresponds to an interfacial strength only about one-fifth of the
bulk shear strength of the polymer. I n the past this was generally attributed to poor
adhesion and it is indeed true that it is very difficult to glue any material to PTFE (Sharpe
and Schonhorn 1964a, b, 1965). However, recent experiments show (Makinson and
Tabor 1964) that marked adhesion can occur between a PTFE slider and clean flamed
glass: a thin film of PTFE about 300 A thick is laid down on the glass surface during sliding
although the friction is still below p = 0.1. Presumably the low friction in this case is
due to the anisotropic yield properties of the polymer.
Although the greater part of recent work supports the concept of interfacial adhesion
between clean surfaces it is ciear that there are some contradictions that remain to be
Friction, lubrication and wear: a survey of work during the last decade 1529
resolved. I n a later section we shall discuss specifically the effect of contaminant films on
adhesion and friction.

1.7. Deformation term


If a hard asperity ploughs its way through the surface of a softer metal the ploughing
force may be easily calculated if adhesion between the surfaces is negligible. Taking
horizontal and vertical components it follows that the normal load is supported by the
appropriate component of the area of contact whilst the horizontal force is 'supported' by
the material ahead of the slider. The horizontal force is then equal to the cross-sectional
area of the groove multiplied by the yield pressure of the metal.
We may consider two simple examples-a hard spherical slider of radius pi and a hard
conical slider of semi-apical angle ,6. If the sphere is sunk into the softer metal to produce
a grooved track of width d (see figure 5(a)) the area supporting the load is A, = +=d2
whilst the cross-sectional area of the groove is A , = +R2(20- sin 28) where 0 is the semi-
angle subtended by the track. If there is no piling up of material ahead of the slider and
adhesion can be ignored, we may write W = PA, and F = p A , ~ Therefore
F 4R'
P = p = X , (20 - sin 26).

Consequently, for d = R(0 = 30") we obtain ,U N 0.23, for d = 2R (the ball immersed ta
its diameter) p = 1.
W

L -

Figure 5. Diagram showing the grooving of a soft metal when a hard indenter slides over it.
(a) Spherical indenter, (b) conical indenter. If interfacial adhesion is negligible and there is no
pile-up of metal the component A , of the area of contact supports the normal load W and the
component A , resists tangential displacement with a force F. In this simple model W = p A l and
F = p A , wherep is the yield pressure of the softer metal. The effective coefficient of friction due to
the grooving mechanism is ,U = F/ W = APIA1.

For a cone the result is independent of the size of the groove (because of geonetric
similarity) and depends only on the semi-apical angle p. For a groove of width d (see
figure 5(b)) we have A , = id2and A , = $d2cot p. Hence

For a semi-apical angle ,f3 of 60" we obtain p = 0.32; for ,8 = 30" (a rather pointed cone)
P Y 1 * 1. (For a more elaborate treatment see Goddard and Wilman 1962.)
1530 F. P. Bowden aiid D.Tabor
If the adhesion is small this deformation term may then be added to the adhesion
component of the friction. This approach was first used by Bowden and Tabor (1943)
in their original studies of the mechanism of friction. It has been much extended and
elaborated by Wilman and his associates in their studies of the action of abrasive particles
(e.g. Avient, Goddard and Wilman 1960). The results show that the rate of removal of
metal by abrasion is inversely proportional to the hardness of the metal, a conciusion in
agreement with that of Kruschov ~ h investigated
o the wear and abrasion of metals over a
wide range of technological conditions (Kruschov and Babichev 1960). It is interesting to
note that if one assumes that 10 % of the volume of the grooves appears as removed metal
there is a very good quantitative agreement between the calculated and observed abrasive
wear rate. Muihearn and Samuels (1962) have suggested that this can be explained in
terms of the average orientation of the abrasive particles. Wilman has also studied, using
electron diffraction, the orientation of surface layers produced by the abrasive process
(Scott and Wiiman 1958). This and later papers show that an oriented texture is obtained
similar to that produced in compression or in rolling.
Several papers have appeared in the last few years emphasizing-perhaps overemphasiz-
ing-the role of the deformation term in sliding friction (Walton 1963, Kraghelsky 1965).
4s the analysis in the previous paragraphs shows, it is possible to carry out model experi-
ments which may involve relatively large ploughing terms. For example, a sharp pointed
cone or a deeply embedded sphere could give coefficients of friction of the order of unity
from the ploughing mechanism alone. However, with repeated traversals the track width
gradually grows (Eldredge and Tabor 1955) and the metal becomes increasingly work
hardened. In general this will lead to a steady decrease in the ploughing term. With
extended surfaces the contribution from the ploughing of the softer surface by asperities on
the harder will, in general, be considerably smaller. This is becsuse the asperities on
polished surfaces rarely have an effective slope greater than 5" or 6", and even on abraded
surfaces the slopes do not appear to be very much greater. If one considers an equivalent
cone of semi-apical angle 85" the contribution to the friction from the ploughing term will
be less than p = 0.1. This is in agreement with the more reaiistic assessment of Eiu (1964)
who suggests that in his sliding experiments on copper surfaces the ploughing term can
make a contribution of the order of p = 0 . 2 to the overall coefficient of friction. Of
course if the sliding surface is a file or a hard-backed assembly of abrasive particles the
deformation term can be considerable.
The simple model discussed above gives a reasonable estimate of the ploughing term when
interfacial adhesion is small. If, however, the adhesion is not negligible the ploughing
term will interact with the adhesion term and this may well lead to an appreciable increase
in the deformation component itself. Some attempt has been made to study this theoret-
ically under conditions of plastic deformation by Johnson and Jefferis (1964) and Kraghelsky
(1965), but it cannot be said as yet that the results are available in a simple quantitative
form. Courtel (1965) has shown experimentally that interactions of this type may play an
important part in the generation of intermittent motion. A more detailed study of this is
now overdue.
In most sliding situations the deformation term is small compared with the adhesion
term. It can however be important in the abrasion of metals as mentioned above, in the
grinding and pulping of wood (Atack and May 1958, Atack and Tabor 195&),in the sliding
friction of PTFE where the adhesion component is generally small (Tabor and Wynne-
Williams 1961) and in lubricated friction of materials for which the deformation term may be
relatively large. This can be particularly marked with rubber. If a hard asperity traverses
a weli-lubricated rubber surface it may be shown that 80-90 % of the friction arises from the
energy lost in deforming the rubber (see figure 6). The energy loss is large, that is the
friction is greater, the larger the hysteresis losses in the rubber (Tabor 1959b). This has
stimulated the use of high-hysteresis-loss rubber on automobile tyres. On wet or greasy
roads the friction is presumably due to the ploughing of the tyre tread by the road asperities.
Consequently the friction is increased (that is the tyre is more skid-resistant) if the tread
has a high hysteresis loss.
Friction, lubrication and \year: a survey of \cork during the last decade 1531
The deformation term is also of primary importance in rolling friction. We shall discuss
this in more detail below.

0 200 400 600


Mean c o n t a c t p r e s s u r e p (Ib in-*)

Figure 6. Coefficient of friction of steel spheres rolling and sliding over well-lubricated rubber as a
function of the mean contact pressure p over the circle of contact. The broken curve is the
theoretical curve calculated on the assumption that the friction is due solely to elastic hysteresis
losses in the rubber. It is seen that over a wide range of pressures the friction for lubricated sliding
is primarily due to hysteresis losses (Tabor 1959b, c).

2. Specific researches
2.1. Friction at cery high speeds
The development of high-speed turbines, aircraft and projectiles has led to a growing
interest in friction at very high speeds. One of the earliest studies under these conditions
was described by Shugarts (1953), who pressed a thin rod against the rim of a rapidly
revolving disk, the whole being maintained in a !ow-pressure atmosphere. Krairr (1955)
studied the friction between a rotating bullet and a steel target and found evidence for
melting during penetration. A more technological technique has involved the use of a
rocket sledge (SORT). More recently, laboratory studies have been carried out using the
elegant ultra-centrifuge technique developed by Beams (1947) for suspending and spinning
rotors magnetically. One method (Bowden and Freitag 1958) is based on the continuous
deceleration of a rapidly rotating ball dropped between three friction pads. I n later

0 200 400 600


Sliding speed [msec-I)

Figure 7. Friction of steel on copper (load 10 g) and on bismuth (load 30 g) as a function of sliding
speed (continuous deceleration experiments). The drop in friction is due to the formation, as a
result of frictional heating, of a thin molten surface film. This drop occurs at a much lower speed
for bismuth than for copper (Bowden and Freitag 19.58).
1532 F. P. Boilden and D.Tabor
experiments (Bowden and Persson 1961) sliding takes place during the oblique impact of a
spinning ball on a plate. Sliding speeds up to 2000 ft sec-I were achieved.
The results show that with metals the friction decreases as the speed increases (see figure 7).
This is due to the generation, by frictional heating, of a thin film of molten metal at the
interface. If large-scale melting occurs the friction again rises because of the large increase
in the area of contact. The rate of removal of metal is determined by the rate at which
the molten front diffuses into the metal ahead of it. With non-metals there is a similar
decrease in friction with increasing speed. With brittle materials thermal shock may lead
to fragmentation and excessive near. With polymers on the other hand the friction and
wear remain low even at the highest speeds. This is because a considerable part of the
frictional energy may be expended in raising the temperature of the molten interface rather
than in expanding the molten zone. In addition there is some evidence that decomposition
of the polymer may provide some measure of gas lubrication, but more work on this is
necessary. The most striking result is that at sliding speeds of 200 ft sec-I the wear of
nylon may be less than that of copper or aluminium and scarcely greater than that of
molybdenum.

2.2. Friction in Gery high cacuum


The friction of solids in air is greatly affected by the presence of oxide layers or adsorbed
gases or other contaminant films. Earlier work on the roie of oxide films has been described
by Whitehead (1950) and also by Wilson (1955); these films generally, but not always,
reduce the friction. They may be broken down if the substrate is soft and the deformation
severe. Some recent studies of the part played by oxides on aluminium have been described
by Courtel (1965).
Oxides and other contaminants may be removed in two ways. The surfaces may be
heated in a high vacuum until the surface films evaporate or decompose. Experiments
along these lines have been carried out by Young (Bowden and Young 1951), Rowe (Bowden
and Rowe 1956) and Kenyon (1956 Ph. D. Dissertation, University of Cambridge): they
sho~7that gross seizure generally occurs between surfaces cleaned in this way. Alternatively
surfaces may be cleaned by rubbing repeatedly in a high vacuum: the surface films are
ruptured or worn away and are not able to reform. The conditions resemble those occurring
between rubbing surfaces in outer space. Broadly speaking the results are the same as
those obtained in the older method: there is a steady increase in friction with repeated

/ 1 ~ 1 ! l 1 1 1 1 1

0 200 400 600 800 l00%0


Number of cycles
Figure 8. Effect of repeated sliding on the friction of diamond on diamond in a high vacuum.
E, vacuum of torr; 11, vacuum of 5 x lo-'' torr. The rise in friction is due to the gradual
attrition of surface films: it is accompanied by marked fragmentation of the diamond surfaces
(Bowden and Hanwell 1964).
Friction, lubrication and wear: a survey of work during the last decade 1533
traversals of the friction track. One interesting difference is observed with diamond surfaces.
These are difficult to clean by heating in a vacuum since prolonged heating produces
graphitization of the diamond surface; in fact the friction in these experiments does not
exceed about p = 0.4. If the diamond surface is repeatedly traversed in a vacuum of
torr at slow sliding speeds so that frictional heating is negligible and graphitization
does not occur, the friction rises to a much higher value, about p = 1 (see figure 8). In
addition the strong adhesion and the increased tangential stress in the interface lead to
catastrophic fragmentation of the diamond (Bowden and Hanwell 1964).
Euckley and Johnson (1966) have carried out friction and wear experiments in a vacuum
of torr after baking out at 200‘c. The surfaces are repeatedly slid over one another
for periods of the order of one hour and it is found that certain alloys and metals give
relatively low friction and wear even under these severe conditions. For example, a
iitanium alloy containing 21 % aluminium sliding on a stainless-steel surface gives a friction
coescient p of only about 0.4; with pure titanium sliding on the same surface there is gross
>
seizure ( p 1.6). The authors attribute this to structural properties of the materials
although a detailed explanation is not possible. The results however are of considerable
practical value.

2.3. Friction at cery high and cery IOW temperatures


2.3.1. High temperatures. The growth of nuclear power has led to a considerab!e
interest in the friction of solids at very high temperatures. Most of the fundamental
studies have been carried out in a vacuum or in an inert atmosphere in order to avoid
complications arising from rapid oxidation. With pure metals the main effect of high
temperature is to increase the ductility of the material at the interface. This greatly
facilitates junction growth and readily leads to gross seizure. For this reason the use of
pure metals and even steels is impractical above about 1000ccc.
In this higher temperature range studies have been made of the friction of the carbides,
borides and nitrides of the transition metals. With most of these materials, even when
cleaved in a vacuum, the friction is low at room temperature. This may be became of tbeir
relatively low ductility. Above about 12OO”c, however, a marked increase in volume
diffusion takes place, sintering at the interface is facilitated and the friction rises to a high
value (Brookes 1966).
At moderately high temperatures (say up to 400”c) the sliding surfaces may be lubricated
by applying solid film lubricants such as graphite or MoS, or by forming in situ a suitable
layer-lattice compound (I, on titanium, see Rowe 1957, Roberts and Owens 1963). At
higher temperatures these films decompose but it is possible that lubricating layers may be
developed which are stable at considerably higher temperatures.
Many technological studies have also Seen carried out at elevated temperatures on metals
immersed in liquid sodium or in an atmosphere of CO,. We shall not discuss this work
here.

2.3.2. Low temperatui es. During the last decade experiments have been carried out on
the friction of solids down to liquid-helium temperatures. To a large extent this work has
been stimulated by specific problems arising from space travel. Two types of laboratory
investigation have been undertaken. In one the rubbing surfaces are immersed in liquid
hydrogen or nitrogen. The results show that with metals the natural protective oxide film
is gradually worn away and is unable to reform; this results in considerable adhesion and
metallic transfer (Bison and Anderson 1964). In the presence of liquid oxygen (Hardy
et al. 1963) the friction and wear of steel surfaces are reduced as the sliding speed increases.
presumably because frictional heating enables an oxide film to form. Under these conditions
the friction and wear are less severe than in the presence of hydrogen or nitrogen. In the
second type of experiment the surfaces are maintained in a vacuum or in an atmosphere of
helium gas. With this arrangement little change in friction with decreasing temperature
has been observed (Burton et al. 1962). One of the major difficulties at the lowest tempera-
1534 F. P. Bowden and D.Tabor
tures is to prevent condensation of small amounts of gaseous impurity. Another important
factor is the possible embrittlement of many metals at these low temperatures.
More work of a fundamental nature is still needed on the friction of solids at very high
and at very low temperatures. On the other hand, at a technological level, the National
Aeronautics and Space Adlninistration (U.S.A.) has made a most successful attack on the
production of friction surfaces that will operate under fairly extreme conditions. This
has been achieved largely by the fabrication of composite materials. For a recent survey
article see Amateau and Glaeser (1965).

2,4. Friction and orientation


A number of investigations have recently been carried out on the friction of hard sliders
over the surface of single crystals of metals and non-metals. With clean metals (Bailey and
Gwathmey 1962, Dyer 1961, Courtel 1965) the friction depends markedly on Orientation.
This is largely due to the way in which slip occurs and to the way in which the surface

\
8

t f

topography changes as a result of the sliding process itself. In some directions the
displacement is such that a ‘hill’ of work-hardened metal is raised ahead of the slider
increasing both the area of contact (see Takagi and Tsuya 1961) and the ploughing or
deformation term: this corresponds to the high friction direction. In other directions this
displacement occurs on either side of the slider and is associated with low friction. With
non-metals similar effects have been observed (Halaunbrenner 1958-9, Steijn 1962, 1964,
Duwell 1962, Riesz and Weber 1964) under conditions where the deformation is primarily
plastic. When the deformation is predominantly brittle, frictional anisotropy also occurs
and again is largely due to an increase in the ploughing term. However in this case the
high friction in certain crystallographic directions is due not so much to a piling up of
material as to an increase in penetration by the slider into the brittle crystal (Bowden and
Brookes 1966). It is possible that this mechanism also accounts for the frictional aniso-
tropy observed earlier on diamond (Seal 1958). Typical results from his work are shown
in figure 9.

2.5. Friction of rubber


Grosch (1963) has studied the friction of rubber over a wide range of temperature and
sliding speeds, and by working at speeds below about 1 cm sec-1 he was able to avoid
complications due to frictional heating. His results show that on rough surfaces there is a
component of friction due to ploughing or deformation and another component due to
adhesion. Both these factors vary with speed and temperature in a manner which directly
Friction, lubrication and wear: a survey of work during the last decade 1535
parallels the known viscoelastic properties of the rubber, and they can be fitted to a ‘master’
friction curve by the use of a WLF transform (see figure IO). The adhesion component,
at room temperature, reaches its maximum value at a sliding speed U of about 1 cm sec-l.
The maximum viscoelastic loss at this temperature occurs at a frequency v of about IO6 cjs.
If these are to be correlated it implies that, in the sliding process, there is a significant
distance h given by h = c/v. This quantity has the value of about 100 i, which is approxi-
mately the length of a segment of the rubber molecule which other experiments show to be
involved in relaxation processes. Similar results at much higher sliding speeds have been
quoted by Bulgin et al. (1962).
Grosch’s results have led to a number of theoretical models which mark an appreciable
advance on the earlier models of Schallamach (1953) and Bartenev (1954). In a recent
paper Schallamach (1963) has described a molecular theory involving bond formation
between the rubber molecule and the other surface, dwell-time and stress-activated detach-
ment of the molecule. Quantitative agreement is reasonably good, but some discrepancies
remain that a more detailed analysis may resolve. Other models have been proposed by
Bulgin et al. (1962), Savkoor (1965) and Kummer (1965 Ph. D. Thesis, Pennsylvania State
University). A different approach by Ludema and Tabor (1966) analyses the sliding
behaviour by applying the simple adhesion theory of friction and considering the way in
which A and s individually vary with temperature and rate of deformation. Good quali-
tative agreement is obtained but the calculated friction F (= As) turns out to be about ten
times larger than the observed friction. This is probably connected with the fact that for
rubber the interface appears to be weaker than the bulk, so that over a wide range of
conditions shearing occurs at the interface itself rather than in the rubber specimen.

2.6. Friction of rigid polymers


Similar studies have been undertaken on the friction of rigid polymers. Earlier work
was carried out at ‘engineering’ speeds so that the interpretation has been greatly complicated
by the effects of frictional heziting (McLarea and Tabor 1953). This not only influences
the strength properties at the contact regions but may also have a profound effect on
interfacial diffusion. However, as a broad generalization it appears that the more rigid
the polymer the less its friction depends on speed and temperature (McLaren and Tabor
1963). Frictional heating has been used as a means of ‘welding’ polymers. At extremely
high sliding speeds surface melting may lead to a marked reduction in friction and wear
(see above).
More recent studies at very low sliding speeds (Eudema and Tabor 1966) where fricticnal
heating is negligible show that, ai temperatures below the glass transition temperature, there
is generally a very poor connection between the frictional behaviour and the viscoelastic
properties of the polymer as conventionally determined. The results suggest that this is
because the conventional study of viscoelasticity is carried out at very small strains and at
ambient pressure. By contrast the material around the friction junctions is subjected to a
very high local pressure and the shear strains, before sliding occurs, are enormous.
Consequently although the sliding friction often varies markedly with speed and temperature
there is no reason why this should reflect viscoelastic properties measured under much
gentler conditions.
With polymers the adhesion at the interface is generally strong. Shearing occurs within
the polymer itself and there is marked transfer. Under carefully controlled conditions strong
interfaciai adhesion can occur even with PTFE,though the friction remains small (Makinson
and Tabor 1964). The low friction must be attributed to the anisotropic yield properties
of the crystallites within the PTFE. The details are not certain but it is interesting to note
that PTFE of low crystallinity gives a friction that is generally 10-15% higher than that
observed with highly crystalline material (Ludema and Tabor 1966). It is of course true
that in many practical situations the adhesion between PTFE and other surfaces is small
(Sharpe and Schonhorn 1964a, b, 1965) and this is accompanied by a further small but
significant drop in friction.
1536 F. P. Bowden and D.Tabor

0 1 1 1 1
-4 -2 0 -4 -2 0

1 I
O i l I I I I I I I I I I I I I I
-4 0 4 e
l o g a,V

(b)
Figure 10. (a) Variation of friction with speed for acrylonitrite-butadiene rubber sliding over a
hard surface. (b) Master curve for the friction data in (a) deduced from the Williams-Landel-
Ferry transform using the known viscoelastic characteristics of the rubber (&os& 1963).
Friction, lubrication and wear: a survey of work during the last decade 1537
2 . 7 . Friction of lainellar solids
During the last decade there have been two main lines of work concerned with the
friction of lamellar solids. The first concerns the basic mechanism. It is now generally
agreed that the low friction of graphite is associated with the presence of adsorbed films.
When these films are removed adhesion is strengthened, the friction rises and there is an
enormous increase in wear (Savage 1948). There is still some dispute as to whether the
cleavage face in graphite is intrinsically a low-energy surface. Some experiments by
Bryant et al. (1964) suggest that this is not so, which would support the view of Rowe (1960)
that the friction of graphite is due to the fracture or cleavage of crystals and that absorbed
gases penetrate to some depth into the lattice and sc reduce the cleavage strength of the
graphite crystallites. Most other workers consider that the cleavage face is a low-energy
surface and that friction involves the sliding of the crystallites themselves over one another.
The edges are considered to be the high-energy sites where adsorption of gases and vapours
readily occurs (Deacon and Goodman 1958). In this view any interaction involving edges
gives a high friction when the edges are clean, a low friction when the edges are covered with
adsorbed films (see figure 11).

Figwe 11. The mechanism of graphite friction according to Deacon and Goodman (1958). The
cleavage face is assumed to be one of IOWsurface energy, the edges of high surface energy. Sliding
occurs between crystallites. In the presence of adsorbed gases on the edges the friction is always
low. If the adsorbed gases are removed by outgassing the graphite, the only high-friction
combinations will be those involving edge interactions.

With MoS, it is now generally agreed that the low friction is essentially a property of the
structure of the material itself, MoS, being extremely weak along the shear planes. The
presence of adsorbed vapour and the other impurities tends, if anything, to increase the
friction (Haltner 1964). For a recent survey of the frictional properties of MoS, see
Winer (1966).
These friction studies have been accompanied by other investigations including the
orientation of surface layers as a result of sliding (Scott and Wilman 1958), the formation
of blisters (Salomon et al. 1964, Swinnerton and Turner 1966), the electrical characteristics
of carbon contacts during sliding, the packing and adhesion of surface films (Johnston
and Moore 1964) and the wear and durability of these films in various environments.
The second line of work is concerned with the formation in situ, by chemical attack of
surface films possessing a lamellar structure. This was first clearly studied in detail by
Rowe (1957) who showed, for example, that titanium could be effectively lubricated by
forming titanium iodide on the surface, tungsten carbide by forming tungsten chloride on
the surface. This has led to a number of practical applications, in particular the work of
Roberts and Owens (1963, 1964). They showed that a lubricant containing a n activated
complex of iodine is a very successful lubricant for titanium and also for chromium.
2.8. Rolling fiiction
Rolling friction arises from two main factors: (i) interfacial slip due to differential
stretching of surface elements, (ii) sub-surface deformation by the rolling member.
1538 F. P. Ban-den and D.Tabor
A great deal of progress has been made in recent years, mainly by Johnson and his
colleagues, in analysing the microslipping process. It turns out that, in general, this is
only a small part of the rolling resistance unless there is very close conformity between, say,
a rolling ball and a grooved track. Interfacial slip can of course become much more
important if there is appreciable spinning of the ball. Otherwise the energy losses due to
sub-surface deformation may constitute the main part of the rolling resistance even for hard
metals. With rubber and polymers this factor is even more important: the deformation
component is almost invariably the main cause of rolling resistance.

2 8.1. Rubber. and rigid polymers. When a steel ball rolls over a rubber surface it com-
~

presses the rubber on the front half of the region of contact and does work on it. The
rubber on the rear half of the region of contact recovers elastically and urges the ball
forjvard. If as much energy was released on the rear half as IS expended on the front half
no net work would be done during rolling. Because rubber is an imperfect elastic material
this balance is never achieved. Energy is lost in the loading-unloading cycle by elastic
hysteresis and the work done in rolling is primarily due to these losses (Tabor 1955).
The energy dissipated appears as heat. Since rubbers and polymers generally have rela-
tively high hysteresis-loss characteristics this factor dominates the rolling friction behaviour.

h
VI PTFE 92 ' / o crystallinity PTFE 48% crystallinity ,/*I,
e
~

80 ib0 240 320 400 480 323 400 486


Temperature (OK) Temperatuie (OK)

Figure 12. Comparison between rolling friction ,uR and viscoelastic losses for two types of PTFE.
The viscoelastic loss is deduced from torsional damping experiments at about 1 cis and allows for
changes in both the loss factor and the elastic modulus. The rolling friction is for a hard sphere
rolling over the ?TEE surface at such a speed that the time to pass through the circle of contact is of
the order of 1 sec. The deformation frequency thus matches the conditions in the viscoelastic
experiments: in the latter experiments the PTFE is at atmospheric pressure, whilst in the rolling
experiments the pressure on the ?TFE in the contact zone is of the order of 1000 atmospheres. In
spite of this difference there is good agreement between the rolling friction and the calculated
viscoelastic loss (Ludema and Tabor 1966).

During the last decade a number of papers have been published to analyse the hysteresis
losses involved in rolling on viscoelastic solids. The problem is basically difficult since the
stress cycle is extremely complex. Papers of increasing sophistication have appeared by
Tabor (1955), Flom and Bueche (1959), May et al. (1959), Greenwood et al. (1961), Hunter
(1961) and Morland (1962).
Flom and Bueche (1959) have shown that the rolling friction on rubber varies with speed
and temperature in a manner which closely parallels the variation of hysteresis loss with
frequency of deformation and temperature. Recently Ludema and Tabor (1944) have
found a similar correlation for rigid polymers below the glass-transition temperature (see
figure 12). This shows that, even at the relatively high pressures occurring at the contact
region, the viscoelastic properties are not greatly changed (see Zosel 1964). This supports
their view that in sliding friction with these materials where no such correlation exists the
high shear strains occurring in the junctions involve properties very different from those
occurring in conventional viscoelastic measurements.
Friction, lubrication and wear: a survey of work during the last decade 1539
2.8.2. Metals. When a hard sphere or cylinder rolls over a metal surface at a high load
the resistance t o rolling arises from the plastic displacement of metal ahead of the roller
(Eldredge and Tabor 1955). As Johnson has shown residual stresses are left below the
surfxe so that during subsequent traversals the metal is subjected to the combined effect of
these residual stresses and the applied contact stresses. These together may not reach the
condition for plastic yielding so that the system may shake down to a state where the
deformations are entirely elastic. The losses will then be due mainly to elastic hysteresis
losses in the metal (which are generally small) and to interfacial microslip. It is probable
that ball and roller races operate in this rigime. At higher loads subsurface plastic flow
may continue, leading t o cumulative plastic flow in the forward direction (Merwin and
Johnson 1963). This elegant mechanism (see figure 13) explains the experimental results of
Crook (1957) and Hamilton (1963) on the deformation accompanying the rolling of
cylinders; it probably has a bearing on the fatigue of metals under repeated rolling.
Shear
stress

Shear

1
Shear
stress

(4 (b) (c)
Figure 13. (a) The rolling of a hard cylinder (or sphere) over an elastic solid showing the deforma-
tion of elements A, B, C, D, E below the surface. (b) The rolling friction of a hard rigid cylinder
over the flat surface of iL softer metal of Young’s modulus E, Poisson’s ratio v and critical shear
stress 5 . The friction F is expressed as the dimensionless parameter FE/2(1 1.XI)@, where a is
half the width of the band of contact. The load is expressed as the ratio of the contact pressure
p to 3. For p < 3. It there is no plastic work even for the first traversal. For repeated traversals
a ‘shake-down’limit occurs for p < 47. For p > 47 sub-surface plastic deformation occurs with
each traversal. The broken lines show the rolling friction parameter if only hysteresis losses are
involved, where U is the fraction of energy lost per cycle by hysteresis. (c) Stress-strain diagrams
showing how an element passing under the zone of contact (stages A, B, C, D, E of figure 13(a))
is first stressed in one sense and then to a greater extent in the reverse sense so that it is left
with a permanent forward shear strain AE.

2.8.3. Ball and roller bearings. There has been an extensive technological study of the
behaviour of ball and roller bearings particularly under severe conditions of operation
(Bisson and Anderson 1964). The behaviour is complicated by interfacial slip, spinning of
balls, ball-cage friction, fatigue and pitting. One of the newer ideas introducsd by Kapitza
(1955) is that, in normal operation, lubricated roller bearings often function under conditions
of almost complete hydrodynamic (or rather elastohydrodynamic) lubrication, a suggestion
developed in much greater detail in recent years by Tallian and his colleagues (Tallian
et ai. 1964).

3. Lubrication and wear


We refer here briefly to recent advances in lubrication and wear.

3.1. Lubrication
Under suitable conditions of load and speed, sliding surfaces can operate with a continu-
ous filin of lubricant between them. The friction is entirely due to the viscous resistance
1540 F. P. Bowden and D.Tabor
of the fiuid and, in principle, no wear of the surfaces occurs. This is the rCgime of classical
hydrodynamic lubrication. As the sliding speed falls or the load increases the hydro-
dynamic film decreases in thickness and the surfaces may be separated by a film which is
thinner than the average height of the surface asperities. The contacting asperities may
now be separated by films only a few molecules in thickness, the properties of which are
radically different from the bulk properties of the lubricant. The friction is higher and
some wear of the asperities may occur. This rCgime is known as that of boundary
lubrication.
The most exciting developments of the last decade have been concerned with the elucida-
tion of a third rkgime of lubrication which occurs between classical hydrodynamic and
boundary, namely eiastohyaroaynamic lubrication. This is most simply explained by
considering the behaviour, as the load is increased, of a sliding system operating under
conditions of classical hydrodynamic lubrication. According to the classical theory, above
a critical load the lubricant film should be squeezed out and metal-metal contact (or
boundary contact) should occur. Careful experiments show, however, (Crook 1957) that
for smooth and carefully aligned surfaces the lubricant film may not be extruded. The
reason is as follows. The oil dragged into the interface is subjected to very high pressures;
with most hydrocarbon oils these pressures produce a prodigious increase in viscosity.
The oil film is thus trapped between the surfaces and behaves virtually like a solid separating
layer (see e.g. Dowson and Higginson 1966). This prevents metal-metal contact and
explains why many mechanisms, in practice, operate unc?er much Eore severe conditions
than the classical theory would ailow. It probably also explains why certain lubricants such
as silicones. which have low pressure-viscosity characteristics, are generally poor lubricants
(Tabor and Winer 1965).
One of the difficulties of elastohydrodynamic lubrication is that it is now difficult to
explain why lubricants ever fail, since the harder they are compressed the more difficult
they are to extrude. It has been suggested that high local temperature flashes are respons-
ible, or that high rates of shear can actually fracture the lubricant film (Hutton 1965), but
more work on this is necessary. It is fairly clear that a good deal of the earlier observations
which were classified as ‘boundary’ hbrication are probably expiicabie in terms of elasto-
hydrodynamic lubrication. Nevertheless there still remains an area in lubrication where
the behaviour is not explicable solely in terms of bulk viscosity or pressure-viscosity
characteristics of the lubricant but is due to some type of liquid-solid interactions at the
surface (Dowson 1965). This may involve physical adsorption or chemical reaction. In
view of the massive ezort being devoted to elastohydrodynamic lubrication a more critical
approach is now required to the rCgime of boundary lubrication.
There are three further points of applied interest that have emerged during the last
decade. The first is the observation of McCutchen (1959) that porous materials may extrude
lubricant when subjected to pressure, thus providing a ’weeping’ type of lubrication: he
has suggested that this may be one of the ways in which the cartilege and the synovial fluid
lubricate bone joints. The second is somewhat similar but operates under much higher
pressures. If a netal surface is lightly shot blasted before lubrication it can be subjected to
very heavy deformations in subsequent rolling or drawing without the occurrence of
scuffing; the lubricant is squeezed out of the cavities during the metal working operation
and prevents metal-metal contact (Lancaster and Rowe 1958-9). In this way even titanium
may be successfully worked. The third concerns the role of chemical reaction in producing
protective surface films. The reactions range from oxygen that may exist fortuitously as
gas dissolved in the lubricant (see e.g. Vinogradov et al. 1961, 19631, to the splitting of
unsaturated bonds in an organic additive (St. Pierre et al. 1964).
3.2. Kramer and Rehbinder efects
We refer here briefly to two areas of work involving surfaces which ten years ago attracted
considerable attention. The first, generally named the Kramer effect, is concerned with the
emission of electrons from surfaces as a result of deformation such as occurs during abrasion
or frictional sliding. A very helpful survey by Grunberg (1 958) describes the main mechan-
Friction, lubrication and wear: a survey of work during the last decade 1541
isms involved. One would expect the Kramer effect to play a part in the formation, as a
result of physical or chemical attachment to the activated surface, of surface films possessing
same lubricating or protective property. However, although a small number of papers
have appeared on the Kramer effect since Grunberg’s survey was published, little systematic
work seems to have been done on its role in frictional and lubrication processes.
The second effect is associated with the name of Rehbinder; it is concerned with the
influence of surface active materials on the mechanical strength of solids. Although
Rehbinder’s observations are not all fully confirmed it is now clear (see recent surveys by
Kramer and Demer 1961, Westwood 1964) that in many cases the adsorption of surface
active materials, especially at surface flaws or cracks, can produce marked reductions in
strength. This may well play a significant part in such processes as lubrication both in
sliding and in metal cutting (P. E. Barlow 1965 private communication) as well as in
Rehbinder’s classical field of rock-drilling. More attention to the role of the Rehbinder
effect in these and allied fields is now overdue.
3.3. Wear
U’c mention here four main areas of work. The first is the abrasive wear produced by
hard particles or asperities (Avient et al. 1960, Kruschov and Babichev 1960, Mulhearn
and Sainuels 1962). This is of considerable practical importance in grinding, boring and
lapping, as well 2s in the wea; of metals by hard oxides and other inclusions. It is also
related to the wear of brittle solids, though in this case local cracking may greatly complicate
the wear process. The second main area is the wear produced by adhesion and transfer
of friction junctions (Antler 1963, Cocks 1962). This is pardcularly marked with metals,
especially if lubricant films are initially absent or worn away by the wear process itself.
It is also important in the friction and wear of polymers. Here the detailed behaviour
greatly depends on the running conditions especially load and speed since these can lead to
changes in the surface layers; with metals there may even be important phase changes in the
sub-surface material (Welsh 1957, de Gee 1965). Thirdly, studies of the wear of rubber
s h m that in some situations the wear does not resemble the junction-fracture mechanism of
metals. and other processes have been suggested (Schallamach 1957-8). Fourthly a deter-
mined attack has been made to explain the freeing of transferred wear fragments of metals
and polymers in terms of surface energy; on this view the fragments are released when the
stored energy exceeds the necessary surface energy (Rabinowicz 1964, 1965). 4 somewhat
different approach connects the freeing of wear fragments with fatigue (Kerridge and
Lanczster 1956). Indeed a good deal of recent work on wear (particularly under conditions
where interfacial adhesion is small) emphasizes the importance of fatigue in many wear
processes as a result of repeated asperity contact (Rozeanu 1963, Kraghelsky and
Nepomnyashchi 1965).
4. Conclusion
This survey has omitted a number of border-line fields in which work of considerable
practical and theoretical interest has been carried out during the last decade. These include
the eRect of electric current on the contact between stationary and sliding metals, the effect
of current on the friction, wear and conductivity of brush-commutator combinations, the
immense reactivity of fresh metal surfaces to organic compounds, a subject of vital interest
to those studying the action of boundary lubricants.
In his remarkable trilogy Lectures on Physics, Professor Feynman remarks that ‘;friction
is a very complicated matter” and that “in view of all the work that has been done it is
surprising that more understanding of this phenomenon has not come about” (Vol. I, 12-5).
This survey in some ways supports these conclusions; but it is also clear that a great deal of
valuable clarification has been carried out during the last decade. We do indeed understand
much better the basic mechanisms of friction and allied phenomena even if quantitative
predictions still elude us. The next decade should see further advances in the more
quantitative interpretation of friction and wear and in a better understanding of the
breakdown of lubricant and other surface films.
1542 F. P. Boirden and D.Tabor
References
ADAMS,N“, 1963, J. Appl. Polymer Sci., 7, 2075-103.
AMATEAU, M. F., and GLAESER, W. A., 1964, Wear, 7, 385-418.
ANDERSOS? D. L.,1960, Wear, 3, 253-73.
ANDERSON, D. L.,CHRISTENSEN, H., and ASDRE.ATCH, P., 1957, J. Appl. PIiys., 28, 923.
ANTLER,M., 1963, J. Appl. Phys., 34, 438-9.
ARCHARD? J. F., 1951, Nature?Lond., 172, 918.
-1957, Proc. Roy. Soc. A, 243, 190-205.
ATACK,D., and MAY,W. D., 1958, Pulp Paper Magazine of Canada. Conference Paper entitled
‘Frictional Mechanisms in the Grinding Process’.
ATACK,D., and TABOR,D., 1958, Proc. Roy. Soc. A, 246, 539-55.
AVIENT,B. W. E., GODDARD, 3.. and WILMAN,H., 1960, Proc. Roy. Soc. A, 258, 159-80.
BAILEY,A. I., and COURTNEY-PRATT, J, S., 1955, PVOC. ROY. SOC. A, 227 500-15.
BAILEY,A. I., and KAY,S., 1967, Proc. Roy. Soc., in the press.
BAILEY,J. M., and GWATHMEY-, A, T., 1962, Trans. Amer. Soc. Lub. Engrs, 5, 45-56.
BARTENEV, 6. M., 1954, Dokl. Akad. Nauk SSSR, 96, 1161-4.
BARTENEV, G. M., and EL’KIN,A. I., 1965, Wear, 8, 8-21.
BEAMS,J. W., 1947, J. Wash. Acad. Sei., 37, 221-41.
BILLINGHURST, P. A., BROOKES, C. A., and TABOR,D., 1966, Physical Basis of Yield and Fracture:
Conf. Series No. 1 (London: Institute of Physics and Physical Society), chap. 4.
BISSON,E. E., and AKDERSON, W. J., 1964, Advanced Bearing Technology (Washington: N.S.A.A.)
BLOK,H., 1952, Proc. Roy. Soc. A, 212, 480-2.
BOWDEN,F. P., and BROOKES, C. A., 1966, Proc. Roy. Soc. A, 295, 244-58.
BOWDEN,F. P., and FREITAG,E. H., 1958, Proc. Roy. Soc. A, 248, 350-67.
BOWDEN,F. P., and HANWELL, A. E., 1964, Nature, Loud., 201, 1279-81.
BOWDEN,F. P., and PERSSON, P. A., 1961, Proc. Roy. Soc. A, 260, 433-58.
BOWDEN,F. P., and ROWE,G. W., 1956, Proc. Roy. Soc. A, 233,429-42.
BOWDEN,F. P., and TABOR,D., 1943, J. Appi. Phys., 14, 141-51.
-1950, Friction and Lubrication of Solids, Part 1 (Oxford: Clarendon Press).
-1964, Friction and Lubrication of Solids, Part 2 (Oxford: Clarendon Press).
BOWDEN,F. P., and YOUNG,J. E., 1951, Proc. Roy. Soc. A, 208, 311-25.
BROOKES, C. A., 1966: Wear, 9, 103-17.
BRYANT, P. J., 1962, Trans. Ninth P’acuum Symposiuni, U.S.A.
BRYAKT,P. J., GUTSHALL, P. L., and TAYLOR, L. H., 1964, Wear, 7, 118-26.
DE BRUYNE, N. A., 1956, J. Appl. Chem., 6, 303-10.
BUCKLEY, D. H., and JOHNSON, R. E., 1966, N A S A Rep. TND-3235.
BULGIN, D., HUBBARD, G. D., and WALTERS, M. H., 1962, Proc. 4th Rubber Technology Coif.,
London, May 1962 (London: Inst. Rubber Industry), pp. 173-88.
BURTON,R. A., RUSSELL, J. A., and Ku, P. M., 1962, Wear, 5, 60-8.
COCKS, M., 1958, J. Appl. Phys., 29, 1609-10.
-1962, J. Appl. Phys., 33, 2152-61.
-1966, Wear, 9, 320-8.
COHEN,S . C., and TABOR,D., 1966, Proc. Roy. Soc. A, 291, 186-207.
COULOMB, C. A., 1785, Theorie des Machines Simples (Memoire de Mathematique et de Physique
de 1’Academie Royale, Paris).
COURTEL, R., 1961, C. R. Acad. Sci., Pcris, 253, 1758-63.
-1965, C. R. Acad. Sei., Pari.s, 261, 3962-5 ; Bull. d’hformations Sei. Techn. Comm. a 1‘Energie
Atomique, No. 90, 1-29.
COURTNEY-PRATT, J. S., and EISSER,E., 1957, Proc. Roy. Soc. A, 238, 529-50.
CROOK,A. W., 1957, Proc. Instn Mech. Engrs, 171, 187-214.
DEACON, R. F., and GOODMAN, J. F., 1958, Proc. Roy. Soc. A, 243, 464-82.
Dowsox, D., 1965, Annual Meeting Amer. Soc. Mech. Engn., Chicago.
DOWSON, D., and HIGGINSON, G. R., 1966, Elasfohydrodjvzamic Lubriccfion (Oxford: Perganlon].
DUWELL, E. J., 1962, J. Appl. Phys., 33, 2691-8.
DYACHENKO, P. E., TOLKACHEVA, N. N., ANDREEV, G. A., and KARPOVA, T. M., 1964, The Actitnl
Contact Area Between Touching Surfaces (New York: Consultants Bureau).
DYER,L., 1961, Acta. Metall., 9, 928-36.
DYSON, J., and HIRST,W., 1954, Proc. Phys. Soc. B, 67, 309-12.
ELEY,D. D., 1962, Adhesion (London: Oxford University Press).
ELDREDGE, K. R., and TABOR, D., 1955, Proc. Roy. Soc. A, 229, 181-98.
Friction, lubrication and wear: a sumey of work during the last decade 1543
FLOM,D. G., and BUECHE, A. M., 1959, J. Appl. Phys., 30, 1725-30.
DEGEE,A. W. J., 1965, Wear, 8, 121-32.
GODDARD, J., and WILMAN, H., 1962, Wear, 5, 114-35.
GREENWOOD,J. A., MINSHALL, H., and TABOR, D., 1961, Proc. Roy. Soc. A, 259, 480-507.
GREESWOOD,J. A., and TABOR,D., 1957, Insrr? Mech. Engrs. Conference on Lubrication and Wear
(London: Instn Mech. Engrs), pp. 314-7.
GREENWOOD, J. A., and WILLIAMSON, J. P. B., 1966, Proc. Roy. Soc. A, 295, 300-19.
GROSCH, K. A., 1963, Proc. Roy. Soc. A, 274, 21-39.
GRUKBERG, L., 1958, Brit. J . Appl. Phys., 9, 85-93.
HALAUSBRENNER, J. B., 1958-9, Wear, 2,423-7.
HALTNER, A. J., 1964, Wear, 7, 102-17.
HAWLTOS,G. M., 1963, P ~ o cInstii . Mech. Eilgrs, 177, 667-75.
HAMILTON, G. M., and GOODMAN, L. E., 1966, J. Appl. Mech., 33, 371-6.
HARDY,W. F.?ALLES:G. P., and JOHXSON, R. L., 1963, N A S A Rep. TND-1580.
HOLM,R., 1958, Electrical Contacts (Berlin: Springer).
HUXTER, S . C., 1961, J. Appl. Mech., 28, 611-7.
HUTTOS,J. F., 1965, Proc. Roy. Soc. A, 287, 222-39.
HOUWINK, R., and SALOMON, G., 1963, Adhesion and Adhesives (Amsterdam: Elsevier).
JOHNSON, K. L., 1955, Proc. Roy. Soc. A, 230, 531-48.
-1961, J. Mech. Eng. Sci., 3, 362.
JOHNSOS,K. L., and JEFFERIS, J. A., 1964, Instn Mech. Engrs Symposium on Fatigue in Rolling
Contact; 1963 (London: Tnstr, Mech. Engrs), pp. 54-65.
JOHNSTON, R. R. M., and MOORE,A. J. W., 1964, Wear, 7,498-512.
KAPITZA,P. L., 1955, Zh. Tekh. Fiz., 25, 747-62.
KERRIDGE, M., and LAKCASTER, J. K., 1956, Proc. Roy. Soc. A, 236, 250-64.
KISG, R. F., and TABOR,D., 1954, Proc. Roy. Soc. A, 223, 225-38.
KRAFFT, J. M., 1955, J. Appl. Phys., 26, 1248-53.
KRAGHELSKY, I. V., 1965, Friction and Wear (London: Butterworths).
KRAGHELSKY, I. V., and DEMKIK, N. B., 1960, Wear, 3, 170-87.
KRAGHELSKY, I. V., and NEPOMNYASHCHI, E. F., 1965, Wear, 8, 303-19.
KRAGHELSKY, I. V., and SABELNIKOV, V. P., 1957, Instn Mech. Eiigrs Conference on Lubrication
a i d Wear (London: Instn Mech. Engrs), pp. 247-51
KRL\MER, I., and DEMER,L. J., 1961, Progr. Mater. Sci., 9, 131-99.
KRUSCHOV, M. M., and BABICHEV, M. A., 1960, Investigations of the Wear Process in Metals
(Moscow: USSR Acad. Science) (in Russian).
LASCASTER, P. R., and ROWE,6 . W., 1958-9, Wear, 2, 428-37.
LESLIE,J., 1804, An Experimental Inquiry into the Nature and Propagation of Heat (London:
J. Mawman).
LIU,T., 1964, Wear, 7, 163-74.
LODGE, A. S., and HOWELL, H. G., 1954, Proc. Phys. Soc. B, 67, 89-97.
LUDErVfA, K. C., and TABOR,D., 1966, Wear, 9, 329-48.
MCCUTCHEN, C. W., 1959, Nature, Lond., 184, 1284-5.
-_ 1962, Wear, 5, 1-17.
MCFARLAKE, J. S., and TABOR,D., 1950, Proc. Roy. Soc. A, 202,244-53.
MACHLIN, E. S . , and YANKEE, W. R., 1954, J. Appl. Phys., 25, 576-81.
MCLAREN, K. G., and TABOR,D., 1963, Nature, Lond., 197, 856.
-1965, Wear, 8, 79-83.
MAKINSON, K. R., and TABOR, D., 1964, Proc. Roy. Soc. A, 281,49-61.
M.~Y,W. D., MORRIS,E. L., and ATACK,D., 1959, J. Appl. Phys., 30, 1713-24.
MERWIN, J. E., and JOHNSON, K. L., 1963, Proc. Instn Mech. Engrs, 177, 676-85.
MIDGLEY, J. W., LOSGLEY,R. I., STRASG,A., and TEER,D. G., 1963, Proc. Instn Mech. E ~ g r s ,
Lubrication and Wear Convention (London: Instr, Mech. Engrs), pp. 198-209.
MISDLIN,R. D., 1949, J . Appl. Ji4ech., 71, 259-68.
MBLGAARD, J., 1962, Proc. Phys. SOC., 79, 516-34.
MOORE,A. C., and TABOR,D., 1952, Brit. J. Appl. Phys., 3, 299-301.
MORLAND, L. W., 1962, J. Appl. Mech., 29, 345-52.
MULHEARX, T. O., and SAMUELS, L. E., 1962, Wear, 5,478-98.
~‘COSSOR, J. J., and JOHNSON, K. L., 1963, Wear, 6, 118-39.
OSBORN, A. B., 1964, R.A.E. Tech. Note No. CPM 69.
PASCOE, M. W., and TABOR,D., 1956, Proc. Roy. Soc. A, 235, 210-24.
RABINOWICZ, E., 1964, Wear, 7, 9-22.
1544 F. P. Bowden and D.Tabor
R.~BINOWICZ, E., 1965, Friction and Wear of Materials (New York: John Wiley).
RIESZ,C. PI., and WEBER,H. S., 1964, Wear, 7, 67-81.
ROACH, A. E., GOODZEIT, C. L., and HLXVICUT, R. P., 1955, Mech. Engng, 77, 350-60.
ROBERTS, R. W., and OWENS,R. S., 1963, Wear, 6,444-56.
-1964, Wear, 7, 513-5.
ROWE,G. W., 1957, Instn Mech. Engrs. Confei*ence on Lubrication and Wear (London: Instn
Mech. Engrs), pp. 333-8
-1960, Wear, 3, 274-85, 454-62.
ROZEAXU, L., 1963, Wear, 6, 33740.
RUBENSTEIN, C., 1958-9a, Wear, 2, 85-95.
-1958-9b, Wear, 2, 296-310.
ST. PIERRE,L. E., OWEKS,W. S . , and KLIXT,R. V., 1964, Nature, L o n 4 202, 1204-5.
SAVAGE, R. H., 1948, J. Appl. Phys., 19, 1-10.
SALOMOK, G., DE GEE,A. W. J., and ZAAT,J. H., 1964, Wear, 7, 87-101.
SAVKOOR, A. R., 1965, Wear, 8, 222-37.
SCHALLAMACH, A., 1953, Proc. Phys. Soc. B, 66, 386-92.
-1957-8, Wear, 1, 384-417.
-1963, Wear, 6, 375-82.
SCOTT,V. D., and WILMAN, H., 1958, Proc. Roy. Soc. A, 247, 353-68.
SEAL,M., 1958, Proc. Roy. Soc. A, 248, 379-93.
SEMENOFF, A. P., 1958, Sxvativanye Metallov (Moscow: Mashgiz).
-1961, Wear, 4, 1-9.
SHARPE, L. H., and SCHONRORN, H., 1964a, Advances in Chemistry No. 43 (Washington: Amer.
Chem. Soc.), p. 189.
-1964b, J . Polymer Sci., B2, 719.
-1965, J. Polymer Sci., A3, 3087.
SHUGARTS, W. W., 1953, J. Franklin Inst., 256, 187-9.
SIKORSKI,M. E., 1964, Wear, 7, 144-62.
SPALVINS, T., and KELLER,D. V., 1962, Trans. 5th Vac. Techn. Conf, New York, pp. 149-55.
SPURR,R. T., 1961, Wear, 4, 150-3.
-1964, Wear, 7, 330-3.
-1965, Wear, 8, 264-9.
STEIJN,R. P., 1961, J. AppL Phys., 32, 1951-8.
-1964, Wear, 7,48-66.
Swmxwrox, B. P. G., and TURNER, M. J. B., 1966, Wear, 9, 142-59.
TABOR,D., 1948, Proc. Roy. Soc. A, 192, 247-74.
-1951, The Hardness of Solids (Oxford: Clarendon Press).
-1955, PUOC.ROY.SOC.A, 229, 198-220.
-1959a, Proc. Roy. Soc. A, 251, 378-93.
-195913, Proc. First Znt. Skid Prevention Conf , Virginia (Charlottesville, Virginia: Virginia
Council of Highway Investigation and Research), pp. 21 1-18.
-1959c, Rev. Gen. Caolrtchoiie, 36, 1401-8.
TABOR, D., and WINER, w. o., 1965, Trans. Amer. SOC.Lub. Engrs, 8, 69-77.
TABOR,D., and WYNNE-WILLIAMS, D. E., 1961, Wear, 4, 391400.
TAKAG:,R., and TSUYA,Y., 1961, Wear, 4, 216-27.
TALLIAN, T. E., et al., 1964, Trans Amer. Soc. Lub. Eqgrs, 7, 109-26.
VINOGRADOV, G. V., ARKAROVA, V. v., and PETROV, A. A., 1961, Wear, 4,274-91.
VINOGRADOV, G. V., PAVLOVSKAYA, N. T., and PODOLSKY, Y. Y . , 1963, Wear, 6, 202-25.
VOYUTSKII, S. S., 1963, Auto-adhesion and Adhesion of High Polymers (New York: John Wiley).
WALTON,D., 1963, Wear, 6, 257-61.
WEISS,P., 1962, Adhesion and Cohesion (Amsterdam : Elsevier).
WESTWOOD, A. R. C., 1964, Ind. Eng. Chem., 56, 14-25.
WHITEHEAD, J. R., 1950, Proc. Roy. Soc. A, 201,109-24.
WILSON,R. W., 1955, Proc. Phys. Soc., 68, 625-41.
WINER,W. O., 1966, Wear. in the press.
WELSH,N. C., 1957, J. AppI. Phys., 28, 960-8.
ZOSEL,A., 1964, Koll. Zeit. U . Z e i t f . Polym., 199, 113-25.

You might also like