You are on page 1of 11

PHYSICAL REVIEW B, VOLUME 65, 134106

Theory of rubber friction: Nonstationary sliding


B. N. J. Persson and A. I. Volokitin
IFF, FZ-Jülich, 52425 Jülich, Germany
共Received 31 May 2001; published 19 March 2002兲
When rubber slides on a hard, rough substrate, the surface asperities of the substrate exert oscillating forces
on the rubber surface leading to energy ‘‘dissipation’’ via the internal friction of the rubber. In this paper we
extend an earlier published theory 关B.N.J. Persson, J. Chem. Phys. 115, 3840 共2001兲兴 to nonstationary sliding,
and present a discussion of how the area of real contact and the friction force depend on the nature of the
substrate surface roughness and on the history of the sliding motion. We consider in detail the case when the
substrate surface has a self-affine fractal structure.

DOI: 10.1103/PhysRevB.65.134106 PACS number共s兲: 81.40.Pq, 62.20.⫺x

I. INTRODUCTION and the surfaces of many cleaved, brittle materials tend to be


self-affine fractal with the fractal dimension D f ⫽3⫺H
The nature of the friction when rubber slides on a hard ⬇2.2⫺2.5. In practice there is always a lower, ␭ 1 , and up-
substrate is a topic of considerable practical importance, e.g., per, ␭ 0 , cutoff length, so that the surface is self-affine fractal
for the construction of tires,1 wiper blades,1 and in the cos- only when viewed in a finite length-scale interval ␭ 1 ⬍␭
metics industry.2 Rubber friction differs in many ways from ⬍␭ 0 . For surfaces produced by brittle fracture, the upper
the frictional properties of most other solids. The reason for cutoff length ␭ 0 is usually identical to the lateral size L of the
this is the very low elastic modulus of rubber and the high fracture surface. This may also be the case for many surfaces
internal friction exhibited by rubber over a wide frequency of engineering importance 共see, e.g., Refs. 13–17兲. However,
region. for road surfaces the upper cutoff ␭ 0 is of the order of a few
The pioneering studies of Grosch3 have shown that rubber millimeters, which corresponds to the size of the largest sand
friction in many cases is directly related to the internal fric-
particles in the asphalt. Less is known about the short-
tion of the rubber. Thus experiments with rubber surfaces
distance cutoff ␭ 1 , but below, it will be assumed to be of the
sliding on silicon carbide paper and glass surfaces give fric-
order of a few micrometers, so that the length-scale region
tion coefficients with the same temperature dependence as
that of the complex elastic modulus E( ␻ ) of the rubber. In where the road surface may be assumed to be fractal may
particular, there is a marked change in friction at high speeds extend over ⬃three orders of magnitude.
and low temperatures, where the rubber’s response is driven In spite of its great practical importance, very few detailed
into the so-called glassy region. In this region, the friction experimental studies of the nonstationary frictional dynamics
shows marked stick slip and falls to a level of ␮ ⬇0.4, which of rubber on a hard rough substrate have been performed.
is more characteristic of plastics. This proves that the friction Recently Ronsin and Coeyrehourcq18 have studied 共experi-
force under most normal circumstances is directly related to mentally兲 the state-, rate-, and temperature-dependent fric-
the internal friction of the rubber, i.e., it is mainly a bulk tion of elastomer, but focused mainly on the glassy side of
property of the rubber.3 the dynamical response of the polymer. An early study by
Almost all real surfaces have roughness on many different Roberts and Thomas focused on some simple nonstationary
length scales which must be taken into account when calcu- sliding problems involving rubber.19
lating the rubber sliding friction force. This was considered When rubber slides on a hard, rough surface with rough-
in the work by Klüppel and Heinrich.4 However, in Ref. 4 ness on the length scales ␭, it will be exposed to fluctuating
the deformation of the rubber surface in response to the sur- forces with frequencies ␻ ⬃ v /␭. Since we have a wide dis-
face roughness is only included in some average way. One of tribution of length scales ␭ 1 ⬍␭⬍␭ 0 , we will have a corre-
us has developed a theory of rubber friction where the defor- sponding wide distribution of frequency components in the
mations of the rubber are taken into account on all relevant Fourier decomposition of the surface stresses acting on the
length scales.5 The theories in Refs. 4 and 5 consider rubber sliding rubber block. The contribution to the friction coeffi-
friction when a rubber block is slid at constant velocity over cient ␮ from surface roughness on the length scale ␭ will be
a hard, rough surface. Other studies of this topic are pre- maximal when v /␭⬇1/␶ , where 1/␶ is the frequency when
sented in Refs. 1 and 6 –10. In this paper we extend the Im E( ␻ )/ 兩 E( ␻ ) 兩 is maximal, which is located in the transi-
theory in Ref. 5 to nonstationary sliding. The theory is valid tion region between the rubbery region 共low frequencies兲 and
for arbitrary 共random兲 surface roughness, but explicit results the glassy region 共high frequencies兲. We can interpret 1/␶ as
are presented for self-affine fractal surface profiles.11,12 Such a characteristic rate of flips of molecular segments 共configu-
surfaces ‘‘look the same’’ when magnified by a scaling factor rational changes兲, which are responsible for the viscoelastic
␨ in the xy plane of the surface and by a factor ␨ H 共where properties of the rubber. Since the flipping is a thermally
0⬍H⬍1) in the perpendicular z direction. We note that activated process it follows that ␶ depends exponentially 共or
many materials of practical importance have 共approximately兲 faster兲 on the temperature ␶ ⬃exp(⌬E/k B T), where ⌬E is
self-affine fractal surfaces. Thus, for example, road surfaces the barrier involved in the transition. In reality, there is a

0163-1829/2002/65共13兲/134106共11兲/$20.00 65 134106-1 ©2002 The American Physical Society


B. N. J. PERSSON AND A. I. VOLOKITIN PHYSICAL REVIEW B 65 134106

wide distribution of barrier heights ⌬E and hence of relax- We now make the basic assumption that during sliding the
ation times ␶ , and the transition from the rubbery region to whole rubber interfacial surface area moves forwards accord-
the glassy region is very wide, typically extending over three ing to ⫺r(t), i.e., we assume that no Schallamach wave
orders of magnitude in frequency. propagation or local interfacial stick-slip motion 共where dif-
ferent interfacial areas slip at different times兲 occurs 共see
II. SLIDING FRICTION Ref. 20兲. At this point it is convenient to introduce a coordi-
nate system with the xy plane fixed in the 共undeformed兲
This section is based on the formalism developed in Ref. bottom surface of the rubber block, and consider the sub-
5. Using the theory of elasticity 共assuming an isotropic elas- strate as moving with the velocity ṙ(t). We first assume that
tic medium for simplicity兲, one can calculate the displace- the rubber block is in complete contact with the substrate
ment field u i on the surface z⫽0 in response to the surface during sliding. Thus, the rubber block surface displacement
stress distributions ␴ i ⫽ ␴ 3i . Let us define the Fourier trans- field induced by the substrate roughness is
forms


1 u共 x,t 兲 ⫽u关 x⫺r共 t 兲兴 ,
u i 共 q, ␻ 兲 ⫽ d 2 x dt u i 共 x,t 兲 e ⫺i(q•x⫺ ␻ t) ,
共 2␲ 兲3
where r(t)⫽x(t)x̂ and

u i 共 x,t 兲 ⫽ 冕 d 2 qd ␻ u i 共 q, ␻ 兲 e i(q•x⫺ ␻ t) ,
u共 q, ␻ 兲 ⫽
1
共 2␲ 兲3
冕 d 2 xdt u关 x⫺r共 t 兲兴 e ⫺i(q•x⫺ ␻ t)
and similar for ␴ i (q, ␻ ). Here x⫽(x,y) and q⫽(q x ,q y ) are
two-dimensional vectors. We have5

1
共 2␲ 兲3
冕 d 2 xdt u共 x兲 e ⫺i 兵 q•[x⫹r(t)]⫺ ␻ t 其
u i 共 q, ␻ 兲 ⫽M i j 共 q, ␻ 兲 ␴ j 共 q, ␻ 兲
⫽u共 q兲 f 共 q, ␻ 兲 , 共3兲
or, in matrix form,
where
u共 q, ␻ 兲 ⫽M 共 q, ␻ 兲 ␴共 q, ␻ 兲 ,
where the matrix M is given in Ref. 5.
We now assume that the surface stress ␴ (q, ␻ ) only acts
u共 q兲 ⫽
1
共 2␲ 兲2
冕 d 2 x u共 x兲 e ⫺iq•x
in the z direction 共see below兲 so that
and
u z 共 q, ␻ 兲 ⫽M zz 共 q, ␻ 兲 ␴z 共 q, ␻ 兲 . 共1兲
Since ␻ ⬃ v q 共where v is a typical sliding velocity兲 we get
␻ /c T q⬃ v /c T Ⰶ1 in most cases of practical interest. Thus,
f 共 q, ␻ 兲 ⫽
1
2␲
冕 dt e ⫺i[q•r(t)⫺ ␻ t] . 共4兲

we can expand M (q, ␻ ) to leading order in ␻ /c T q. This


If ␴ f (t) denotes the frictional shear stress acting on the bot-
gives5
tom surface of the rubber block, then the instantaneous
Eq power absorption P(t)⫽ ␴ f (t)A 0 ẋ(t) must be given by the
共 M zz 兲 ⫺1 ⫽⫺ . 共2兲 rate of work by the substrate surface asperities on the rubber
2 共 1⫺ ␯ 2 兲
block:
It is interesting to note that if instead of assuming that the
surface stress acts in the z direction we assume that the dis-
placement u points along the z direction, then ␴ f 共 t 兲 A 0 ẋ 共 t 兲 ⫽ 冕 d 2 x 具 u̇共 x,t 兲 • ␴ 共 x,t 兲 典 , 共5兲

␴ z 共 q, ␻ 兲 ⫽ 共 M ⫺1 兲 zz 共 q, ␻ 兲 u z 共 q, ␻ 兲 , where the terms inside the angular brackets 具 ••• 典 denotes


the ensemble average, and where A 0 is the surface area. But
where in the limit ␻ /c T qⰆ1,

2Eq 共 1⫺ ␯ 兲 u̇共 x,t 兲 ⫽u̇关 x⫺r共 t 兲兴 ⫽⫺ṙ共 t 兲 •ⵜu关 x⫺r共 t 兲兴


共 M ⫺1 兲 zz ⫽⫺ ,
共 1⫹ ␯ 兲共 3⫺4 ␯ 兲 ⳵
⫽⫺ẋ 共 t 兲 u关 x⫺r共 t 兲兴 . 共6兲
which differs from Eq. 共2兲 only with respect to a factor ⳵x
4(1⫺ ␯ ) 2 /(3⫺4 ␯ ). For rubberlike materials ( ␯ ⬇0.5) this
factor is of order unity. Hence, practically identical results Thus, using Eqs. 共5兲 and 共6兲,

冕 冓 ⳵⳵ 冔
are obtained independently of whether one assumes that the
interfacial stress or displacement vector are perpendicular to 1
the nominal contact surface. In reality, neither of these two ␴ f 共 t 兲 ⫽⫺ d 2x u共 x,t 兲 • ␴ 共 x,t 兲 . 共7兲
A0 x
assumptions holds strictly, but the result above indicates that
the theory is not sensitive to this approximation. Using Eqs. 共1兲 and 共7兲 gives

134106-2
THEORY OF RUBBER FRICTION: NONSTATIONARY SLIDING PHYSICAL REVIEW B 65 134106

␴ f 共 t 兲⫽
共 2␲ 兲2
A0
冕 d 2 qd ␻ d ␻ ⬘ 共 ⫺iq x 兲 e ⫺i( ␻ ⫹ ␻ ⬘ )t 冕 d ␻ ⬘ f 共 ⫺q, ␻ ⬘ 兲
E共 ␻⬘兲
2 共 1⫺ ␯ 2 兲
e ⫺i ␻ ⬘ t

⫻ 具 u共 q, ␻ 兲 • ␴共 ⫺q,⫺ ␻ ⬘ 兲 典


共 2␲ 兲2
冕 d qd ␻ d ␻ ⬘ 共 ⫺iq x 兲 e
2 ⫺i( ␻ ⫹ ␻ ⬘ )t

1
2␲
冕 dt ⬘ e iq•r(t ⬘ )
E 共 t⫺t ⬘ 兲
2 共 1⫺ ␯ 2 兲
. 共15兲
A0
Using these equations, Eq. 共13兲 gives
⫻ 共 M zz 兲 ⫺1 共 ⫺q, ␻ ⬘ 兲 具 u z 共 q, ␻ 兲 u z 共 ⫺q,⫺ ␻ ⬘ 兲 典 .
共8兲 ␴ f 共 t 兲⫽
1
2␲
冕 d 2 q q 2 cos ␾ C 共 q 兲
Substituting Eqs. 共3兲 in 共8兲 gives

␴ f 共 t 兲⫽
共 2␲ 兲2
冕 d qd ␻ d ␻ ⬘ 共 ⫺iq x 兲 e
2 ⫺i( ␻ ⫹ ␻ ⬘ )t
⫻ 冕 dt ⬘ e ⫺iq•[r(t)⫺r(t ⬘ )]
iE 共 t⫺t ⬘ 兲
2 共 1⫺ ␯ 2 兲
. 共16兲
A0
If we define
⫻共 M zz 兲⫺1 共⫺q, ␻ ⬘ 兲 f 共 q, ␻ 兲 f 共 ⫺q, ␻ ⬘ 兲 具 u z 共 q兲 u z 共 ⫺q兲 典 .

Let us now consider sliding on a randomly rough surface


共9兲
F 共 q,t 兲 ⫽⫺
1
2␲
冕 dt ⬘
E 共 t⫺t ⬘ 兲
␴ 0 共 1⫺ ␯ 2 兲
e ⫺iq•[r(t)⫺r(t ⬘ )] , 共17兲

described by the function z⫽h(x) 关where x⫽(x,y)]. As-


we get
sume first that the rubber is able to deform and completely


follow the substrate surface profile and that 兩 ⵜh(x) 兩 Ⰶ1. 1
Thus we can approximate u z ⬇h(x). Using Eq. 共9兲 gives ␴ f 共 t 兲⫽ ␴0 d 2 q q 2 cos ␾ C 共 q 兲 Im F 共 q,t 兲 . 共18兲
2

␴ f 共 t 兲⫽
共 2␲ 兲2
A0
冕 d 2 qd ␻ d ␻ ⬘ 共 ⫺iq x 兲 e ⫺i( ␻ ⫹ ␻ ⬘ )t The friction coefficient ␮ can be obtained by dividing the
frictional shear stress 共18兲 with the pressure ␴ 0 :
⫻ 共 M zz 兲 ⫺1 共 ⫺q, ␻ ⬘ 兲 f 共 q, ␻ 兲 f 共 ⫺q, ␻ ⬘ 兲 具 h 共 q兲 h 共 ⫺q兲 典 ,

where we assumed that 具 h 典 ⫽0. Now, note that


共10兲
␮共 t 兲⫽
1
2
冕 d 2 q q 2 cos ␾ C 共 q 兲 P 共 q,t 兲 Im F 共 q,t 兲 . 共19兲

具 h 共 q兲 h 共 ⫺q兲 典 ⫽
A0
共 2␲ 兲4
冕 d 2 x 具 h 共 x兲 h 共 0兲 典 e ⫺iq•x
In Eq. 共19兲 we have introduced an additional factor P(q,t),
defined as the fraction of the original macrocontact area
where contact remains when we study the contact area on the
A0 length scale ␭⫽2 ␲ /q 共see below兲. In principle, ␯ depends
⬅ C共 q 兲, 共11兲 on frequency but the factor 1/(1⫺ ␯ 2 ) varies from 4/3
共 2␲ 兲2
⬇1.33 for ␯ ⫽0.5 共rubbery region兲 to ⬇1.19 for ␯ ⫽0.4
since 具 h(x)h(x⬘ ) 典 depends only on the difference x⫺x⬘ . The 共glassy region兲 and we can neglect the weak dependence on
spectral density C(q) is defined by frequency.
Since C(q) and P(q,t) only depend on the magnitude of
C共 q 兲⫽
1
共 2␲ 兲2
冕 d 2 x 具 h 共 x兲 h 共 0兲 典 e ⫺iq•x. 共12兲
q, from Eq. 共19兲,

Substituting Eqs. 共11兲 in 共10兲 and using Eq. 共2兲 gives ␮共 t 兲⫽


1
2
冕 dq q 3 C 共 q 兲 P 共 q,t 兲 冕 d ␾ cos ␾ Im F 共 q,t 兲 .

␴ f 共 t 兲⫽ 冕 d 2 q q 2 cos ␾ C 共 q 兲 冕 d ␻ d ␻ ⬘ e ⫺i( ␻ ⫹ ␻ ⬘ )t f 共 q, ␻ 兲
Note that the factor cos ␾ in the integrand vanishes when
共20兲

iE 共 ␻ ⬘ 兲 ␾ ⫽ ␲ /2, while it is maximal when ␾ ⫽0. This has a simple


⫻ f 共 ⫺q, ␻ ⬘ 兲 , 共13兲 but important physical origin: Consider two cosine-surface
2 共 1⫺ ␯ 2 兲 corrugations, where the ‘‘wave vector’’ points 共i兲 along the x
where we have used polar coordinates so that q x ⫽q cos ␾ . axis 共the sliding direction兲, and 共ii兲 along the y axis. The
Now, note that former case corresponds to ␾ ⫽0, and in this case the rubber
block will experience pulsating deformations during sliding

冕 d ␻ f 共 q, ␻ 兲 e ⫺i ␻ t ⫽ f 共 q,t 兲 ⫽e ⫺iq•r(t) 共14兲


along the x axis. The second case corresponds to ␾ ⫽ ␲ /2,
where the elastic deformations of the rubber do not change
during sliding along the x axis, and this type of surface
and roughness will therefore not contribute to the friction. The

134106-3
B. N. J. PERSSON AND A. I. VOLOKITIN PHYSICAL REVIEW B 65 134106

function P(q,t) was derived in Ref. 5 for the case of station- 具 ␴ z2 共 t 兲 典 q


ary sliding, and the same derivation is valid in the present G 共 q,t 兲 ⫽ . 共22兲
2 ␴ 20
case. Thus,
Here 具 ␴ z2 (t) 典 q is the average of the square of the interfacial
P 共 q,t 兲 ⫽
2

冕 0

dx
sin x
x
exp关 ⫺x 2 G 共 q,t 兲兴 , 共21兲
stress, where the average only includes the roughness wave
vectors with magnitude smaller than q, and where it is as-
sumed that contact with the substrate occurs everywhere.
where Now, note that

具 ␴ z2 共 t 兲 典 ⫽
共 2␲ 兲2
A0
冕 d 2 qd ␻ d ␻ ⬘ e ⫺i( ␻ ⫹ ␻ ⬘ )t 具 ␴ z 共 q, ␻ 兲 ␴ z 共 ⫺q, ␻ ⬘ 兲 典

⫽ 冕 d 2 qd ␻ d ␻ ⬘ e ⫺i( ␻ ⫹ ␻ ⬘ )t M zz
⫺1 ⫺1
共 q, ␻ 兲 M zz 共 ⫺q, ␻ ⬘ 兲 C 共 q 兲 f 共 q, ␻ 兲 f 共 ⫺q, ␻ ⬘ 兲


1
4
冕 d 2 qd ␻ d ␻ ⬘ q 2 e ⫺i( ␻ ⫹ ␻ ⬘ )t
E共 ␻ 兲 E共 ␻⬘兲
1⫺ ␯ 2 1⫺ ␯ 2
C 共 q 兲 f 共 q, ␻ 兲 f 共 ⫺q, ␻ ⬘ 兲

冕 冏 冕 冏
2
1 1 E 共 t⫺t ⬘ 兲
⫽ d 2q q 2C共 q 兲 dt ⬘ e iq•r(t ⬘ )
4 2␲ 1⫺ ␯ 2
1
⫽ ␴ 20
4
冕 dq q 3 C 共 q 兲 冕 d ␾ 兩 F 共 q,t 兲 兩 2 . 共23兲

Thus r共 t ⬘ 兲 ⬇r共 t 兲 ⫹ṙ共 t 兲共 t ⬘ ⫺t 兲 共29兲

G 共 q,t 兲 ⫽
1
冕 q
dq q 3 C 共 q 兲 冕 d ␾ 兩 F 共 q,t 兲 兩 2 . 共24兲
so that


8 qL 1 E 共 t⫺t ⬘ 兲
F 共 q,t 兲 ⫽⫺ dt ⬘ e ⫺iq•ṙ(t)(t⫺t ⬘ ) . 共30兲
2␲ ␴ 0 共 1⫺ ␯ 2 兲
Let us summarize the basic results obtained above. The
friction coefficient ␮ (t) for a flat rubber surface sliding on a Substituting this in Eqs. 共25兲–共27兲 gives ␮ (t)⫽ ␮ k 关 ẋ(t) 兴 ,
nominally flat substrate in the most general case is given by i.e., the friction coefficient depends only on the instantaneous
sliding velocity. The necessary condition for the validity of
␮共 t 兲⫽
1
2
冕 dq q 3 C 共 q 兲 P 共 q,t 兲 冕 d ␾ cos ␾ Im F 共 q,t 兲 ,
the expansion 共29兲 in Eq. 共28兲 is that

q 兩 ẍ 共 t 兲 兩 共 ␶ * 兲 2 Ⰶ1, 共31兲
共25兲
where ␶ * is the memory time of the kernel E(t) 关see Eq.
where the 共normalized兲 area of contact A(␭)/A 0 , on the 共38a兲兴.
length scale ␭⫽2 ␲ /q, is given by We consider now the limit ␴ 0 ⰆE( ␻ ⫽0), which is satis-
fied in most applications. In this case, for most q values of
P 共 q,t 兲 ⫽
2

冕 0

dx
sin x
x
exp关 ⫺x 2 G 共 q,t 兲兴 , 共26兲
interest, G(q,t)Ⰷ1, so that only xⰆ1 will contribute to the
integral in Eq. 共26兲, we can approximate sin x⬇x, and

G 共 q,t 兲 ⫽
1
冕 q
dq q C 共 q 兲
3
冕 d ␾ 兩 F 共 q,t 兲 兩 .
2
共27兲
P 共 q,t 兲 ⬇
2


0

dx exp关 ⫺x 2 G 共 q,t 兲兴 ⫽ 关 ␲ G 共 q,t 兲兴 ⫺1/2.
8 qL 共32兲
In the equations above Thus, within this approximation, using Eqs. 共27兲 and 共32兲 we
get P(q,t)⬀ ␴ 0 so that ␮ is independent of the nominal


1 E 共 t⫺t ⬘ 兲 stress ␴ 0 . Similarly, note that if we scale E( ␻ )→ ␣ E( ␻ ),
F 共 q,t 兲 ⫽⫺ dt ⬘ e ⫺iq•[r(t)⫺r(t ⬘ )] . 共28兲 then from Eqs. 共27兲 and 共32兲, P(q,t)⬀1/␣ , so that ␮ de-
2␲ ␴ 0 共 1⫺ ␯ 2 兲 pends only on the frequency variation of the complex elastic
modulus, but not on its magnitude. For tires, the condition
Note that if E(t) changes slowly with t, then in Eq. 共28兲 we ␴ 0 ⰆE( ␻ ⫽0) is usually satisfied. Consequently, on a dry
can expand road track one expects the same friction for wide and narrow

134106-4
THEORY OF RUBBER FRICTION: NONSTATIONARY SLIDING PHYSICAL REVIEW B 65 134106

tires, assuming that the rubber-road adhesional interaction


and temperature effects are unimportant.
In order to take into account that P(q,t)→1 when
G(q,t)→0, we use the interpolation formula

P 共 q,t 兲 ⬇ 共 1⫹ 关 ␲ G 共 q,t 兲兴 3/2兲 ⫺1/3. 共33兲


Numerical evaluation of Eq. 共26兲 shows that Eq. 共33兲 is an
accurate representation of P(q,t) for all q 共or, equivalently,
all G).
If we assume that the substrate surface is self-affine frac-
tal on all length scales between an upper and lower cutoff,
␭ 0 ⬅2 ␲ /q 0 and ␭ 1 ⬅2 ␲ /q 1 , we have C(q)⫽0 for q⬍q 0 ,
while for q⬎q 0 ,

C 共 q 兲 ⬇k 共 q/q 0 兲 ⫺2(H⫹1) , 共34兲


where H⫽3⫺D f 共the fractal dimension 2⬍D f ⬍3). If we
define 具 h 2 典 ⫽h 20 /2, then Eq. 共11兲 gives k⫽(h 0 /q 0 ) 2 H/2␲ .
Using Eqs. 共25兲 and 共34兲 with q⫽q 0 ␨ gives

␮共 t 兲⫽
H
共 q h 兲2
4␲ 0 0
冕 1
q 1 /q 0
d ␨ ␨ 1⫺2H P 共 ␨ ,t 兲

⫻ 冕 d ␾ cos ␾ Im F 共 q,t 兲 , 共35兲

where P( ␨ ,t)⫽A(␭)/A 0 is given by Eq. 共26兲 with

G 共 q,t 兲 ⫽
H
共 q h 兲2
16 0 0
冕 1
q 1 /q 0
d ␨ ␨ 1⫺2H 冕 d ␾ 兩 F 共 q,t 兲 兩 2 .
共36兲
Note that since, to a good approximation, P(q,t)
FIG. 1. 共a兲 Rheological model corresponding to Eq. 共37兲. 共b兲
⬃ 关 G(q,t) 兴 ⫺1/2, it follows that P⬃1/q 0 h 0 , and thus ␮ The complex elastic modulus E( ␻ ) as a function of frequency. The
⬃q 0 h 0 . logarithm has 10 as the basis. E 1 ⫽109 N/m2 and a⫽1000.


III. NUMERICAL RESULTS 1
dt ⬘ E 共 t⫺t ⬘ 兲 e ⫺iq x [x(t)⫺x(t ⬘ )]
2␲
As an example, assume that E is given by the model
shown in Fig. 1共a兲. This model is, in fact, not a good descrip-
tion of real rubbers, since the transition with increasing fre-
quency from the rubbery region to the glassy region is too
⫽E 1 1⫺ 冋 冕 a

t

⫺⬁

dt ⬘ e ⫺(1⫹a)(t⫺t ⬘ )/ ␶ e ⫺iq x [x(t)⫺x(t ⬘ )] .

abrupt, leading to a too narrow 共and too high兲 ␮ k ( v ) peak. 共38b兲


Nevertheless, the model gives a qualitatively correct E( ␻ ). In what follows we consider two different histories for x(t):
The model in Fig. 1共a兲 corresponds to the complex elastic 共i兲 Sliding velocity step-change:
modulus
x共 t 兲⫽ v 0t for t⬍0,
E 1 共 1⫺i ␻ ␶ 兲
E共 ␻ 兲⫽ . 共37兲
1⫹a⫺i ␻ ␶ x共 t 兲⫽ v 1t for t⬎0.
This function is shown in Fig. 1共b兲. Note that E(⬁)⫽E 1 and Substituting this in Eq. 共30兲 and using Eqs. 共38b兲 gives


E(0)⫽E 1 /(1⫹a) so that E(⬁)/E(0)⫽1⫹a. Since typi-
cally E(⬁)/E(0)⬇1000 we take a⫽1000 in all numerical a 1 ␪ 共 ⫺t 兲
F⫽⫺ E 1 1⫹ ⫺
calculations presented below. We assume E 1 Ⰷ ␴ 0 , in which 1⫹a a 1⫹iq x v 0

冋 册冊
case ␮ (t) is independent of E 1 and ␴ 0 . Note that
e ⫺t⫺iq x v 1 t 1⫺e ⫺t⫺iq x v 1 t

E共 t 兲⫽ 冕 d␻ e ⫺i ␻ t
冋 a
E 共 ␻ 兲 ⫽2 ␲ E 1 ␦ 共 t 兲 ⫺ ␪ 共 t 兲 e⫺(1⫹a)t/ ␶
␶ 册 ⫺␪共 t 兲
1⫹iq x v 0

1⫹iq x v 1
where t is measured in units of ␶ * ⫽ ␶ /(1⫹a), v in units of
, 共39兲

共38a兲
1/(q 0 ␶ * ), and q is measured in units of q 0 . 共ii兲 Stop and
so that start:

134106-5
B. N. J. PERSSON AND A. I. VOLOKITIN PHYSICAL REVIEW B 65 134106

x共 t 兲⫽ v 0t for t⬍0,

x 共 t 兲 ⫽0 for 0⬍t⬍T,

x 共 t 兲 ⫽ v 1 共 t⫺T 兲 for t⬎T.


Substituting this in Eq. 共30兲 and using Eq. 共38b兲 gives

F⫽⫺
a
1⫹a
1

E 1 1⫹ ⫺
␪ 共 ⫺t 兲
a 1⫹iq x v 0

⫺ ␪ 共 t 兲 ␪ 共 T⫺t 兲 冋 e ⫺t
1⫹iq x v 0
⫹1⫺e ⫺t 册
⫺ ␪ 共 t⫺T 兲 冋 e ⫺t⫺iq x v 1 (t⫺T)
1⫹iq x v 0
⫹e ⫺(t⫺T)⫺iq x v 1 (t⫺T)

⫺e ⫺t⫺iq x v 1 (t⫺T) ⫹
1⫺e ⫺(t⫺T)⫺iq x v 1 (t⫺T)
1⫹iq x v 1 册冊 , 共40兲

where again t is measured in units of ␶ * , wave vector q in


units of q 0 , and velocity v in units of 1/(q 0 ␶ * ).
Note that ␮ (t) depends on H and q 0 h 0 . In what follows
we use H⫽0.8 and q 0 h 0 ⫽1. Since ␮ ⬃q 0 h 0 , the friction
coefficient for other q 0 h 0 can be obtained from direct scal-
ing.
Figure 2共a兲 shows the kinetic friction coefficient as a
function of the logarithm 共with 10 as the basis兲 of the sliding
velocity. The maximal friction occurs at a velocity v c where
the substrate surface asperities give rise to fluctuating forces
acting on the rubber with frequencies occurring in the tran- FIG. 2. 共a兲 Kinetic friction coefficient ␮ k , and 共b兲 relative area
sition region between the rubbery region and the glassy re- of contact P⫽A( ␨ )/A(1), as a function of the logarithm 共with 10 as
gion in the mechanical response of the solid. For real rubber the basis兲 of the sliding velocity. The relative area of contact is
shown at two different magnifications, ␨ ⫽3 and ␨ ⫽100. The slid-
the peak maximum of ␮ k ( v ) is smaller than in Fig. 2共a兲, and
ing velocity is in natural units 1/(q 0 ␶ * ) 共see text兲. H⫽0.8, q 0 h 0
␮ k does not decrease towards zero for ‘‘large’’ and ‘‘small’’
⫽1, a ␴ 0 /E 1 ⫽0.03, a⫽1000, and ␨ max⫽100.
velocities as in Fig. 2共a兲, but levels off at ␮ k ⬇0.2⫺0.4. This
is, at least in part, a result of the fact that for real rubber
Im E( ␻ ) does not decrease as rapidly towards zero as it does ␶ * and 1/(q 0 ␶ * ), respectively. Note that v 0 and v 1 are both
for the model elastic modulus 共37兲. to the left of the peak maximum in Fig. 2 共i.e., v 0 ⬍ v c and
Figure 2共b兲 shows the logarithm of the relative area of v 1 ⬍ v c ); in this case the friction coefficient changes mono-
contact P⫽A( ␨ )/A(1) 关where A(1) is the nominal contact tonically between the steady-state values ␮ k ( v 0 ) and
area兴 as a function of the logarithm of the sliding velocity. P ␮ k ( v 1 ). Similarly, for t⬎0 the relative contact area P de-
is shown for two different magnifications, ␨ ⫽3 and ␨ ⫽100. creases monotonically towards the steady-state value taken at
Note that P decreases from a constant value in the rubbery the velocity v 1 . A similar effect is observed when the sliding
region to another much smaller value in the glassy region. velocity changes abruptly from v 0 ⬎ v c to v 1 ⬎ v c , both lo-
The drop in magnitude corresponds to the factor cated to the right of the peak maximum in Fig. 2共a兲. This is
E(⬁)/E(0)⫽1⫹a⫽1001. The physical reason for this is illustrated in Fig. 4, which shows the same as Fig. 3, but now
clear: at low sliding velocities the perturbing frequencies act- when the velocity changes 共at t⫽0) from v 0 ⫽0.01⬎ v c to
ing on the rubber surface from the surface asperities occur in v 1 ⫽0.1. The situation is drastically different, however, if the
the rubbery region where E( ␻ )⬇E(0) so the rubber is very velocity changes from v 0 ⬍ v c to v 1 ⬎ v c 共or vice versa兲. This
soft. At high sliding velocities the perturbing frequencies are is shown in Fig. 5 for the case when v 0 ⫽10⫺4 and v 1
very high and correspond to the glassy region where E( ␻ ) ⫽0.1. In this case the sliding friction exhibits a characteristic
⬇E(⬁)⫽(1⫹a)E(0). Since the area of real contact is peak 共stiction spike兲, with the height ⌬ ␮ ⬇ ␮ k ( v c )⫺ ␮ k ( v 0 )
roughly proportional to 1/E the observed results follow. and the width ⌫⬃ ␶ * . This result is easy to understand: As-
Figure 3 shows 共i兲 the friction coefficient ␮ (t), and 共ii兲 sume that the sliding velocity changes abruptly from v 0
the relative area of contact P⫽A( ␨ )/A(1) as a function of ⬍ v c to v 1 ⬎ v c . Thus, the velocity of the solid at the inter-
time when the sliding velocity changes abruptly from v 0 face will, during a very short time period, increase from v 0
⫽10⫺4 to v 1 ⫽10⫺3 at t⫽0. The relative area of contact is to v 1 . If we assume that the frictional shear stress is given
shown at two different magnifications, ␨ ⫽3 and ␨ ⫽100. The 共approximately兲 by the instantaneous value, ␮ (t)⬇ ␮ 关 ẋ(t) 兴 ,
time and the sliding velocity are measured in natural units, then it follows that there must be a friction spike of height

134106-6
THEORY OF RUBBER FRICTION: NONSTATIONARY SLIDING PHYSICAL REVIEW B 65 134106

FIG. 3. 共a兲 Friction coefficient ␮ (t), and 共b兲 relative area of


contact P⫽A( ␨ )/A(1), as a function of time when the sliding ve- FIG. 4. 共a兲 Friction coefficient ␮ (t), and 共b兲 relative area of
locity changes abruptly from 10⫺4 to 10⫺3 at t⫽0. The relative contact P⫽A( ␨ )/A(1), as a function of time when the sliding ve-
area of contact is shown at two different magnifications, ␨ ⫽3 and locity changes abruptly from 0.01 to 0.1 at t⫽0. The relative area
␨ ⫽100. The time and the sliding velocity are in natural units, ␶ * of contact is shown at two different magnifications, ␨ ⫽3 and ␨
and 1/(q 0 ␶ * ), respectively 共see text兲. H⫽0.8, q 0 h 0 ⫽1, a ␴ 0 /E 1 ⫽100. The time and the sliding velocity are in natural units, ␶ * and
⫽0.03, a⫽1000, and ␨ max⫽100. 1/(q 0 ␶ * ), respectively 共see text兲. H⫽0.8, q 0 h 0 ⫽1, a ␴ 0 /E 1
⫽0.03, a⫽1000, and ␨ max⫽100.
⌬ ␮ ⫽ ␮ ( v c )⫺ ␮ ( v 0 ) as indeed observed. This argument also
共36兲 gives G(q,t)⫽G 0 (q) for t⬍0 and G(q,t)
predicts that no friction spike should occur if both v 0 and v 1
are below or above v c , again in agreement with the calcula- ⫽G 0 (q)exp(⫺2t) for t⬎0. As long as the area of real con-
tact is small compared to the nominal contact area, we have
tions 共see Figs. 3 and 4兲.
Figures 6 and 7 show the results of two stop-start calcu- 关see Eq. 共32兲兴 P⬇ 关 ␲ G(q,t) 兴 ⫺1/2 so that the area of contact
lations. Figure 6 shows a case where the sliding velocity is A(t)⫽A(0)exp(t), or in real time units, A(0)exp(t/ ␶ * ), for
abruptly changed from v 0 ⫽10⫺3 ⬍ v c to zero, t⫽0, and then t⬎0.
increased back to v 0 at t⫽T⫽5. Note that ␮ (t) decreases Figure 7 shows a case where the sliding velocity is
continuously during ‘‘stop’’ 共relaxation兲, and then increases abruptly changed from v 0 ⫽0.1⬎ v c to zero 共at t⫽0兲 and
monotonically back to ␮ k ( v 0 ) for t⬎5. Similarly, the area of then switched back to v 0 at t⫽T⫽5. Note that the friction
contact increases monotonically during ‘‘stop,’’ and then at coefficient again decreases monotonically during ‘‘stop,’’ but
the onset of sliding decreases back towards its original 共for exhibits a stiction spike when v is switched back to v 0 . This
t⬍0) value. It is interesting to note that the area of contact at is in sharp contrast to the case when v 0 ⬍ v c 共Fig. 6兲, and the
the magnification ␨ ⫽100 initially increases exponentially origin of this difference is the same as presented before in
with time as ⬃exp(t/ ␶ * ) giving a straight line in the t the context of Fig. 5. Figure 7共b兲 shows that the area of
contact increases monotonically during ‘‘stop’’ and then at
⫺log P diagram with the slope 1/log(10). This result is easy
the onset of sliding decreases back to its original 共for t⬍0)
to prove analytically: First note that when v ⫽ v 0 for t⬍0
value. In this case the contact area for both ␨ ⫽3 and 100
and v ⫽ v 1 ⫽0 for t⬎0, from Eq. 共39兲 we get
initially increases exponentially with time as ⬃exp(t/ ␶ * )
1 iq x v 0 giving a straight line in the t⫺log P diagram with the slope
F⫽ ⫺ 关 1⫺ 共 1⫺e ⫺t 兲 ␪ 共 t 兲兴 . 共41兲 1/log(10). This result can be explained in the same way as in
a 1⫹iq x v 0
the context of Fig. 6.
Since in the present case 1/a⫽10⫺3 while q v 0 ⬎0.1 we can Figure 8 shows the results of several stop-start calcula-
neglect the 1/a term in Eq. 共41兲. Substituting F(q,t) in Eq. tions with T⫽0.1, 0.3, 0.5, 1, 2, 3, 5, and 10. An analysis of

134106-7
B. N. J. PERSSON AND A. I. VOLOKITIN PHYSICAL REVIEW B 65 134106

FIG. 5. 共a兲 Friction coefficient ␮ (t), and 共b兲 relative area of FIG. 6. 共a兲 Friction coefficient ␮ (t), and 共b兲 relative area
contact P⫽A( ␨ )/A(1), as a function of time when the sliding ve- of contact P⫽A( ␨ )/A(1), as a function of time when the sliding
locity changes abruptly from 10⫺4 to 0.1 at t⫽0. The relative area velocity changes abruptly, reducing from v 0 ⫽10⫺3 to 0 at t⫽0 and
of contact is shown at two different magnifications, ␨ ⫽3 and ␨ then returning to v 0 at t⫽5. The relative area of contact is shown
⫽100. The time and the sliding velocity are in natural units, ␶ * and at two different magnifications, ␨ ⫽3 and ␨ ⫽100. The time and
1/(q 0 ␶ * ), respectively 共see text兲. H⫽0.8, q 0 h 0 ⫽1, a ␴ 0 /E 1 the sliding velocity are in natural units, ␶ * and 1/(q 0 ␶ * ), respec-
⫽0.03, a⫽1000, and ␨ max⫽100. tively 共see text兲. H⫽0.8, q 0 h 0 ⫽1, a ␴ 0 /E 1 ⫽0.03, a⫽1000, and
␨ max⫽100.
the height ⌬ ␮ of the stiction spikes shows that ⌬ ␮ (T) in-
creases logarithmically with stop time T 关see Fig. 9共a兲兴 as
long as the height of the stiction spike is much smaller than time. Figure 9共b兲 shows ⌬ ␮ as a function of ⌬A(T)/A 0 for
the limiting 共saturation兲 level, ⌬ ␮ (⬁)⬇2.8, observed for ␨ ⫽3, as obtained from Fig. 8.
large stopping time. Thus the results in Fig. 8 for T⬍3 are
well described by
IV. DISCUSSION


⌬ ␮ ⫽ ␮ max⫺ ␮ k 共 v 0 兲 ⬇ 共 ⌬ ␮ 兲 0 ln 1⫹
T
3␶*冊, 共42兲
The theory developed above can be used to estimate the
friction coefficient for nonsteady sliding of rubber on a
rough, hard substrate. The input for the calculation, namely
where (⌬ ␮ ) 0 ⬇4. This curve is given by the solid line in Fig. the complex elastic modulus E( ␻ ) and information about the
9共a兲 while the circles are the data from Fig. 8. This type of substrate roughness 关spectral function C(q)] can be obtained
logarithmic time dependence of ⌬ ␮ on T is often observed in directly from relatively simple experiments.
experiments,22 but the origin of the effect for nonrubber ma- The theory relates the friction force to the coordinate x(t)
terials is likely to be different from the present case 共see Sec. of the bottom surface of the rubber block. However, the ex-
IV兲. One contribution to the increase in the height ⌬ ␮ (T) ternal driving force does not act directly on the bottom sur-
with increasing stopping time T results from a corresponding face of the block, but usually on the top surface or at some
increase in the contact area ⌬A(T). If the average 共over the other distant area. When studying the motion of the rubber
contact area兲 shear stress ␴ that is needed to start sliding block it is, in general, necessary to include the elastic defor-
would be independent of the stopping time, then one expects mation of the rubber between the bottom surface of the block
⌬ ␮ (T)⬃⌬A(T). This formula is often assumed to hold, but and the area where the external forces act. Consider, for ex-
Fig. 9共b兲 shows that this relation does not hold accurately in ample, a rectangular elastic block. Assume that the upper
the present case; thus, ␴ also will depend on the stopping surface of the block is ‘‘glued’’ to a thin rigid sheet as indi-

134106-8
THEORY OF RUBBER FRICTION: NONSTATIONARY SLIDING PHYSICAL REVIEW B 65 134106

FIG. 7. 共a兲 Friction coefficient ␮ (t), and 共b兲 relative area of FIG. 8. 共a兲 Friction coefficient ␮ (t), and 共b兲 relative area of
contact P⫽A( ␨ )/A(1), as a function of time when the sliding ve- contact P⫽A( ␨ )/A(1), as a function of time when the sliding ve-
locity changes abruptly, reducing from v 0 ⫽0.1 to 0 at t⫽0 and locity changes abruptly reducing from v 0 ⫽0.1 to 0 at t⫽0 and then
then returning to v 0 at t⫽5. The relative area of contact is shown at returning to v 0 at t⫽T⫽0.1, 0.3, 0.5, 1, 2, 3, 5, and 10. The relative
two different magnifications, ␨ ⫽3 and ␨ ⫽100. The time and the area of contact is shown at two different magnifications, ␨ ⫽3 and
sliding velocity are in natural units, ␶ * and 1/(q 0 ␶ * ), respect- ␨ ⫽100. The time and the sliding velocity are in natural units, ␶ *
ively 共see text兲. H⫽0.8, q 0 h 0 ⫽1, a ␴ 0 /E 1 ⫽0.03, a⫽1000, and 1/(q 0 ␶ * ), respectively 共see text兲. H⫽0.8, q 0 h 0 ⫽1, a ␴ 0 /E 1
and ␨ max⫽100. ⫽0.03, a⫽1000, and ␨ max⫽100.

cated in Fig. 10. If a parallel force is applied to the rigid to both the static and kinetic friction forces. Finally, since
sheet, either directly, or, as is typical in many applications, most real surfaces are contaminated with a few monolayers
via an external spring 共spring constant k s ), then the block of physisorbed organic molecules, the contamination layer
will deform elastically as indicated in the figure. This elastic will also contribute the friction force as discussed in detail in
coupling between the sliding interface, where the friction many recent publications.21,7 All these additional contribu-
force is generated, and the point or area where the driving tions to the rubber friction give a contribution to the friction
force acts must be taken into account when studying sliding coefficient which is typically of order ␮ ⬃0.2.
dynamics of the block, and is particularly important for ma- Let us comment on the concept of the static friction force.
terials with a low elastic modulus such as rubber. We will If there would be no interfacial pinning processes of the type
study this problem in another publication. described above, then, strictly speaking, the static friction
Let us point out that in addition to the contribution to force would vanish. However, assume that the E( ␻ ) function
rubber friction from the internal friction of the rubber studied has the form shown in Fig. 11共a兲, giving rise to a kinetic
above, there will in general be other contributions arising friction coefficient ␮ k ( v ) of the form shown in Fig. 11共b兲.
from pinning effects at the interface. Thus, for a clean rubber Assume now that we start to pull the block with some speed
surface 共if that ever exists兲 in contact with a hard substrate, v 1 indicated in Fig. 11共b兲. In this case we would observe a
the rubber molecules at the interface will rearrange them- stiction spike 共or static friction coefficient兲 of height
selves to bind as strongly as possible to the substrate surface. ⬃ ␮ k ( v a ), see Fig. 11共c兲. Thus, if there are very low-
Because of the lateral corrugation of the substrate potential frequency 共long-time兲 relaxation processes in the solid 关cor-
this will in general give rise to an energy barrier towards responding to the low-frequency peak in Fig. 11共a兲兴, they
sliding. If the rubber is in contact with another polymer sur- may show up as a static friction force under most normal
face, e.g., rubber in contact with rubber, chain interdiffusion sliding friction experiments. However, if the sliding velocity
may also occur at the interface which will give a contribution is extremely small ( v ⬍ v a ) there would be no stiction spike,

134106-9
B. N. J. PERSSON AND A. I. VOLOKITIN PHYSICAL REVIEW B 65 134106

FIG. 9. 共a兲 The stiction ⌬ ␮ (T) as a function of the stopping


time T 共from Fig. 8兲. The solid line is given by Eq. (42). 共b兲 The
stiction ⌬ ␮ (T) as a function of the increase of the relative area of
contact ⌬A( ␨ )/A(1) 共for ␨ ⫽3) 共from Fig. 8兲. The time T is in units FIG. 11. 共a兲 The ‘‘loss spectra’’ Im E( ␻ )/ 兩 E( ␻ ) 兩 as a function
of ␶ * . H⫽0.8, q 0 h 0 ⫽1, a ␴ 0 /E 1 ⫽0.03, a⫽1000, and ␨ max⫽100. of frequency ␻ . 共b兲 The kinetic friction coefficient. 共c兲 The time-
dependent friction coefficient 共defined as the time-dependent fric-
which could be interpreted as the absence of a static friction tion force divided by the load兲.
force. Thus, there is no single value of the static friction
coefficient—it depends upon the initial dwell time and rate
for rubber. However, we believe that the origin of the ⌬ ␮
of starting 共which, according to Fig. 10, also depends on the
⬃ln T relation observed for nonrubber materials22 共e.g.,
rubber modulus兲.19
stone, paper, or plastic兲 has a different origin, unrelated to
In many sliding friction experiments it has been observed
the internal friction of these materials: During steady sliding
that the height ⌬ ␮ (T) of the stiction spike in stop-start ex-
at low sliding velocities a wide distribution of tangential
periments increases logarithmically with the stopping time. A
stresses will occur at the sliding interface. Thus, surface
similar effect was observed above in our model calculations
asperities23 or stress domains24 in the contact area form a
wide distribution of elastically deformed states which during
‘‘stop’’ will slowly relax 共by thermal excitation over the bar-
riers兲 towards the equilibrium 共unstressed兲 state. This process
can be shown24 to give a ⌬ ␮ (T) which depends logarithmi-
cally on the stopping time T. The same process gives rise to
a logarithmic dependence of the kinetic friction coefficient
␮ k on the sliding velocity 共for low sliding velocities兲, again
in agreement with experiments. This latter effect is not ob-
served for rubber unless the rubber is probed well into the
glassy region 共which requires ‘‘low’’ enough temperature
compared to the rubber glass transition temperature兲, where,
in fact, rubber behaves like most other solid materials, and
FIG. 10. A rubber block on a rough substrate. where the friction no longer is dominated by the internal

134106-10
THEORY OF RUBBER FRICTION: NONSTATIONARY SLIDING PHYSICAL REVIEW B 65 134106

friction, but rather by interfacial processes as described has been developed for arbitrary E( ␻ ), e.g., the experimen-
above. tally measured E( ␻ ) can be directly used as an input. The
theory relates the friction force to the coordinate of the bot-
V. SUMMARY AND CONCLUSION tom surface of the rubber block. In most practical applica-
tions the external driving force acts some distance away from
There is at present a strong drive by tire companies to the sliding interface, and it is then necessary to include in the
design new rubber compounds with lower rolling resistance, equations of motion the elastic coupling between the bottom
higher sliding friction, and reduced wear. At present these surface of the rubber block and the area where the external
attempts are mainly based on a few empirical rules and on driving force acts.
very costly trial-and-error procedures. We believe that a fun-
damental understanding of rubber friction and wear may help
in the design of new rubber compounds for tires and other ACKNOWLEDGMENTS
rubber applications, e.g., wiper blades or products for the
cosmetic industry. This work has been supported by a research and develop-
In this paper we have presented a general theory of the ment grant from Pirelli Pneumatici. We thank O. Albohr,
hysteretic contribution to rubber friction for nonstationary G. Heinrich, M. Klüppel, U. Kuhlmann, G. Luengo, and
sliding. The theory has been developed for rubber sliding on V. Peveri for useful discussions. B.P. also thanks BMBF for
a hard self-affine fractal surface, e.g., a tire on a road surface. a grant related to the German-Israeli Project Cooperation,
Numerical results were presented for a simple rheological ‘‘Novel Tribological Strategies from the Nano to
model of the complex elastic modulus E( ␻ ), but the theory Meso-Scales.’’

1
D.F. Moore, The Friction and Lubrication of Elastomer 共Perga- 13
J.F. Archard, Proc. R. Soc. London, Ser. A 243, 190 共1957兲.
14
mon Press, Oxford, 1972兲; M. Barquins, Mater. Sci. Eng. 73, 45 J.A. Greenwood, in Fundamentals of Friction, Macroscopic and
共1985兲; A.D. Roberts, Rubber Chem. Technol. 65, 673 共1992兲. Microscopic Processes, edited by I.L. Singer and H.M. Pollack
2
G. Luengo 共private communication兲. 共Kluwer, Dordrecht, 1992兲.
15
3
K.A. Grosch, Proc. R. Soc. London, Ser. A 274, 21 共1963兲; see S. Roux, J. Schmittbuhl, J.P. Vilotte, and A. Hansen, Europhys.
also, A.D. Roberts, Rubber Chem. Technol. 65, 3 共1992兲; S.P. Lett. 23, 277 共1993兲.
Arnold, A.D. Roberts, and A.D. Taylor, J. Nat. Rubber Res. 2, 1
16
A. Majumdar and B. Bhushan, J. Tribol. 113, 1 共1991兲.
17
共1987兲; M. Barquins and A.D. Roberts, J. Phys. D 19, 547 J.A. Greenwood and J.B.P. Williamson, Proc. R. Soc. London,
共1986兲. Ser. A 295, 300 共1966兲.
18
4 O. Ronsin and K.L. Coeyrehourcq, Proc. R. Soc. London, Ser. A
M. Klüppel and G. Heinrich, Rubber Chem. Technol. 73, 578
457, 1277 共2001兲.
共2000兲. 19
A.D. Roberts and A.G. Thomas, Rubber Chem. Technol. 50, 266
5
B.N.J. Persson, J. Chem. Phys. 115, 3840 共2001兲.
共1977兲.
6
B.N.J. Persson, Surf. Sci. 401, 445 共1998兲.
7
20
B.N.J. Persson, Phys. Rev. B 63, 104101 共2001兲.
B.N.J. Persson, Sliding Friction: Physical Principles and Appli- 21
B.N.J. Persson, J. Chem. Phys. 113, 5477 共2000兲; M.H. Müser, L.
cations, 2nd ed. 共Springer, Heidelberg, 2000兲. Wenninger, and M.O. Robbins, Phys. Rev. Lett. 86, 1295 共2001兲;
8
B.N.J. Persson and E. Tosatti, J. Chem. Phys. 112, 2021 共2000兲. J. Gao, W.D. Luedtke, and U. Landman, Tribol. Lett. 9, 3
9
A. Schallamach, Wear 6, 375 共1963兲; Y.B. Chernyak and A.I. 共2000兲.
Leonov, ibid. 108, 105 共1986兲. 22
P. Berthoud, T. Baumberger, C. G’Sell, and J.M. Hiver, Phys.
10
G. Heinrich, Kautsch. Gummi Kunstst. 45, 173 共1992兲; Rubber Rev. B 59, 14 313 共1999兲.
Chem. Technol. 70, 1 共1997兲. 23
C. Caroli and P. Nozieres, in Physics of Sliding Friction, edited
11
J. Feder, Fractals 共Plenum Press, New York, 1988兲. by B.N.J. Persson and E. Tosatti 共Kluwer, Dordrecht, 1996兲.
12
E. Bouchaud, J. Phys.: Condens. Matter 9, 4319 共1997兲. 24
B.N.J. Persson, Phys. Rev. B 51, 13 568 共1995兲.

134106-11

You might also like