Você está na página 1de 47

Tire road frictional interaction

Physical approach

B.G.B. Hunnekens
DCT 2008.088

TU/e Bachelor Final Project


June, 2008

Supervised by:
dr.ir. I. Lopez Arteaga (TU/e)
ir. R. van der Steen (TU/e)
Eindhoven University of Technology
Department of Mechanical Engineering
Dynamics and Control Group

Abstract
The friction of rubber plays an important role in practical situations. Friction between the road and
the tire is such a practical application. Models, developed by B.N.J. Persson, predict the kinematic
friction coefficient as a function of the slip velocity of the rubber. The basic theory of the models include
viscoelastic energy dissipation on many different length scales.
The first model that is implemented does not include the local heating of the rubber. Results are
shown for different power spectral densities of road surfaces and for different ranges of length scales. The
obtained (v) curves are validated with the results published by Persson.
The viscoelastic modulus of rubber strongly depends on temperature. Therefore, the second model
addresses the local heating of the rubber. An additional heat diffusion equation is solved to account for
this raise in temperature. The results obtained are in qualitative agreement with the results presented
by Persson. An increase in temperature softens the rubber and, in practical tire-road situations, leads
to less energy dissipation of the rubber. Therefore the friction coefficient will drop compared to the
temperature-independent case. Quantitative validation of the temperature-dependent model was not
possible: Persson uses measurement data in his calculations that is not published.
The power spectral density of the road surface plays a significant role in the calculation of the friction
coefficient. Therefore, measurements on a real road surface (surface 1) are performed with an optical
imaging profiler. The data is used to calculate the psd and the resulting friction coefficient. This is also
done for a fitted psd and it is concluded that a fitted psd will lead to similar results. Therefore a fitted
spectrum can be used to calculate the (v) curve. Additionally, road surface data was available from
another project (surface 2). The spectra and the resulting friction coefficients are compared. Although
the two surfaces are quite different, the two (v) curves are comparable.
It is widely known that rain decreases the friction between the tire and the road. The measured road
profile is flutted numerically which leads to a new corrected surface. This leads to a different psd of
the road surface and results in a different (v) curve. As observed in practice, the friction coefficient
decreases a significant amount when the road is wet.

Samenvatting
De wrijving van rubber speelt een belangrijke rol in praktische situaties. De wrijving tussen het wegdek
en de band is zon praktische toepassing. Modellen, ontwikkeld door B.N.J. Persson, voorspellen de
wrijvingscoefficient als functie van de slip snelheid van het rubber. De basis van de modellen bestaat uit
visco-elastische energie dissipatie op vele verschillende lengteschalen.
Het eerste model dat is gemplementeerd bevat niet de locale opwarming van het rubber. Resultaten worden gegeven voor verschillende power spectral densities van wegdekken en voor verschillende
ranges van lengteschalen. De berekende (v) curves zijn gevalideerd met de resultaten die Persson heeft
gepubliceerd.
De visco-elastische modulus van rubber hangt sterk af van de temperatuur. Daarom behandeld
het tweede model de locale opwarming van het rubber. Een extra warmte diffusie vergelijking wordt
opgelost om de verhoging van de temperatuur in rekening te brengen. De resultaten komen kwalitatief
overeen met de resultaten van Persson. Een verhoging van de temperatuur versoepelt het rubber en,
in praktische band-weg situaties, leidt tot minder energie dissipatie van het rubber. Om deze reden zal
de wrijvingscoefficient afnemen in vergelijking met het temperatuur-onafhankelijke geval. Kwantitatieve
validatie van het temperatuur-afhankelijke model was niet mogelijk: Persson gebruikt gemeten data in
zijn berekeningen die hij niet heeft gepubliceerd.
De power spectral density van het wegdek speelt een belangrijke rol in de berekening van de wrijvingscoefficient. Om deze reden zijn metingen verricht op een echt wegdek (wegdek 1) met een optical
imaging profiler. De data is gebruikt om de psd en de resulterende wrijvingscoefficient te berekenen. Dit
is ook gedaan voor een gefit psd en er kan geconcludeerd worden dat een gefit psd leidt gelijke resultaten.
Daarom kan een gefit spectrum gebruikt worden voor het berekenen van de (v) curve. Ook was extra
meetdata van een wegdek beschikbaar van een ander project (wegdek 2). De spectra en de resulterende
wrijvingscoefficienten zijn vergeleken. Hoewel de twee oppervlakken van elkaar verschillen, zijn de (v)
curves erg vergelijkbaar.
Het is bekend dat regen de wrijving verminderd tussen band en wegdek. Het gemeten wegdek is
numeriek vol laten lopen met water wat leidt tot een nieuw gecorrigeerd wegdek. Dit leidt tot een
ander psd van het wegdek en resulteert in een andere (v) curve. Net als in de praktijk neemt de
wrijvingscoefficient significant af als het wegdek nat is.

Acknowledgements
I would like to express my gratitude to my supervisor, dr. ir. Ines Lopez Arteaga, for the support
throughout the project. I value the weekly meetings with room for questions, discussions and tips. I
am also grateful for the support of ir. Rene van der Steen working on the CCAR project about FEM
Tire modelling. He helped me a lot during the project and was always available to answer my questions.
Furthermore I would like to thank Marc van Maris for measuring the road surface in such a short amount
of time and providing me with excellent data.

Contents
1 Introduction
2 Rubber friction without temperature dependence
2.1 Background of the friction model . . . . . . . . . . .
2.2 The friction model . . . . . . . . . . . . . . . . . . .
2.3 Numerical results . . . . . . . . . . . . . . . . . . . .
2.3.1 Implementation of the model . . . . . . . . .
2.3.2 Friction coefficient for different slip velocities
2.3.3 Friction coefficient for different road profiles .

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

3
3
5
6
6
7
7

3 Rubber friction with temperature dependence


3.1 Background of the friction model . . . . . . . . . . . . . . . . . . . . .
3.2 The friction model . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3 Numerical results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.3.1 Implementation of the model . . . . . . . . . . . . . . . . . . .
3.3.2 Friction coefficient and temperature for different slip velocities

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

.
.
.
.
.

11
11
11
13
13
13

4 Results for real road surfaces


4.1 The power spectral density of a real asphalt road surface
4.2 Friction coefficient for a real asphalt road surface . . . .
4.3 Comparison of two real road surfaces . . . . . . . . . . .
4.4 Influence of rain . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

.
.
.
.

17
17
19
21
24

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.
.
.

.
.
.
.

.
.
.
.

5 Conclusion and recommendations


27
5.1 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
5.2 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
A Self-affine fractal surfaces

29

B The rheological model

31

C Used parameter values

33

Symbols
A
a
a1
aT
C
Cv
D
E
F0
Ff
H
h
L
P
q
q0
q1
Q
Rasp
T
t
t0
T0
Tg
Tq
v
x, y, z

area, m2
lattice spacing, m
ratio moduli rheological model
shift factor WLF-equation
power spectral density road, m4
heat capacity, J/kgK
heat diffusivity
viscoelastic modulus, Pa
nominal force, N
shear force, N
Hurst exponent
height, m
length, m
apparent area of contact
wave vector, 1/m
lower cut-off
upper cut-off
heat production per unit volume and unit time, W/m3
radius asperity region, m
temperature, K, C
time, s
half time a asperity region is in contact with the road
background temperature, K, C
glass transition temperature, K, C
temperature on length scale q, K, C
slip velocity, m/s
rectangular coordinates, m

Greek Letters

length scale, m
th
thermal conductivity, W/mK

kinematic friction coefficient

Poisson ratio

frequency of oscillation, rad/s

angle between sliding direction and wave vector q, rad

density, kg/m3
0
nominal stress, Pa
f
shear stress, Pa

magnification
max maximum magnification

Chapter 1

Introduction
Rubber friction plays an important role in many practical situations. One practical application is for
instance the rubber friction between a tire and the road. It is important to know how the kinematic
friction coefficient depends on the slip velocity of the tire, in order to make good models. The friction of
rubber on a hard rough substrate differs form most other solids due to the very low elastic modulus and
the high internal friction exhibited by the rubber over a wide region of frequencies. The rubber exhibits
viscoelastic behavior and will therefore dissipate energy. The energy dissipation can be related to the
friction coefficient. When slipping, the temperature of the tire increases. The temperature increase has
a significant effect on the rubber friction. From a physical basis, models have been developed by Persson
that predict the friction coefficient as a function of the slip velocity [4], [5].
The goal of this project can be split into three parts:
implement a temperature-independent friction model that calculates the kinematic friction coefficient. Investigate the effect of different parameters on the friction coefficient.
implement a temperature-dependent friction model and compare the results with the temperatureindependent model. Investigate the effect of heating of the rubber.
investigate the effect of a real measured road surface on the friction coefficient.
In the first chapter the temperature-independent model will be discussed. In the subsequent chapter
the temperature-dependent model will be addressed. A comparison between the two models will be made
in this chapter. The effect of a real, measured road surface on the model outcome will be discussed in
chapter three. A comparison of two real road surfaces will be made. Also, the influence of rain on the
friction coefficient will be addressed in this part.

Chapter 2

Rubber friction without


temperature dependence
2.1

Background of the friction model

Rubber friction has got two fundamental contributions: hysteric and adhesion. The hysteric contribution
results from internal friction of the rubber. The road surface exerts oscillating forces on the rubber. Due
to internal damping of the rubber this leads to viscoelastic energy dissipation. The adhesion component
of the rubber friction is only important for very clean rubber surfaces. The adhesion force is a result
from the van der Waals interaction between the interfaces. In [3] it is shown that adhesion on a rough
surface becomes important for velocities in order of magnitude v < 108 m/s. This is due to the fact
that the actual contact area is only 1% of the nominal contact area. The internal or hysteric friction is
for that reason dominant in all practical cases. The adhesion component is only relevant for very clean
surfaces and is therefore not included in the here presented model.
The theory that Persson developed [4] starts with the energy dissipation E of the rubber due to
the oscillating forces exerted by the road surface. This energy dissipation can be related to the frictional
shear stress f exerted on the rubber. The kinematic friction coefficient can be related to the frictional
shear stress by dividing it with the nominal stress 0 . This is illustrated in figure 2.1.

Figure 2.1: Definition of the kinematic friction coefficient as the ratio between shear stress and nominal
stress

The oscillating forces occur on many different length scales. This is due to the many irregularities
present on a road surface, which is illustrated in figure 2.2. The length scales that are present in the
road are defined as and the corresponding wave vector is:
q

On a road surface sample of length L the longest possible length scale is L. The smallest possible
length scale could be determined by road contamination (or a thin layer of thermally degraded rubber)
and is defined as 1 .

Figure 2.2: Simplified road surface with length scales . The longest length scale possible is L

The length scales that are considered in the analysis are 1 < < L. In terms of wave vectors this
can be defined as:
2
2
= q0 < q < q 1 =
,
L
1
where q0 is called the lower distance cutoff and q1 the upper distance cutoff.
A magnification when looking at a length scale is defined as qq0 . The maximum magnification
corresponds to a length scale 1 and is defined as = max = qq01 .
To describe the wavelengths that are present in a road surface, the power spectral density C(q) can
be used. A special kind of surface is a self-affine fractal surface. From experiments it is known that a
typical road surface can be approximated as being self-affine fractal. The power spectral density can
then be described by:
C(q) q 2(H+1) ,
with H the Hurst exponent. More information about self-affine fractal surfaces can be found in Appendix A.
The road surface exerts oscillating forces on the rubber with certain frequencies. This frequency can
be related to the sliding velocity v and wavelength :
=

The viscoelastic modulus of the rubber material depends on this frequency. The energy dissipation
in the material is maximal if the loss tangent is maximal. This frequency is located in the transition
region between the low frequency rubbery region and the high frequency glassy region. A broader range
of length scales will lead to more energy dissipation and hence, this will broaden the (v) curve and will
increase the peak maximum.
When rubber is squeezed onto a hard solid surface of different length scales it will fill out part of the
cavities. Such a situation is sketched in figure 2.3. It is clear that the rubber pressure is high enough to
make the rubber follow the large length scale roughness of the hard solid. The small cavities however
are not completely filled. Especially at the bottom of a cavity it is hard for the rubber to follow the solid
surface. This is due to the smaller pressure at the bottom of the cavities. The small-scale roughness
does contribute to the friction force but only at the top of the big asperities. This is taken into account
in this friction model. Due to the above described phenomenon the apparent area of contact is smaller
than the nominal area of contact. This apparent area of contact at magnification (length scale ) is
defined as:
P () =

A(
A(L)

Figure 2.3: Only part of the cavities of the hard solid are filled with rubber. This is due to the smaller
pressure at the bottom of the cavities

2.2

The friction model

This paragraph contains the equations that make up the temperature-independent model. The friction
coefficient depends on many different parameters, the most important are = (v, P (q), C(q), E()).
The resulting equations are the following:

=
P (q)

G(q)



Z 2
E(qv cos())
q 3 C(q)P (q)dq
cos()Im
d
(1 2 )0
q0
0


1

erf
2 G

Z q
Z 2
E(qv cos())
1
3


q C(q)dq
(1 2 )0 d
8 q0
0
1
2

q1

(2.1)
(2.2)
(2.3)

where is the Poisson ratio. The frequency has been written in terms of the slip velocity:
v
= = qv cos(),

where the factor cos() accounts for the fact that the sliding direction may differ from the direction of
the wave vector q.
5

2.3
2.3.1

Numerical results
Implementation of the model

Equations 2.1-2.3 for the friction model and equation A.1 for the road surface are implemented in Matlab.
The rheological model used for the viscoelastic modulus can be found in Appendix B. The calculations
can be performed in many different ways. Especially the integrals can be determined with different
methods. It is possible to calculate the integrals numerically with a trapezoid integration method.
Another possibility is calculating the integrals explicitly as an analytical expression. This could lead to
more accurate results and less calculation time. The calculation time is especially important because
the model that includes temperature dependence (chapter 3) needs to be solved iteratively. This leads
to much longer calculation times so it is important that this simple model is calculated as efficiently as
possible.
The friction model is implemented in three different ways. The first method uses numerical integration
(trapezoid method) to evaluate the integrals. The second method performs the integration by analytical
integration. After this, the corresponding values are substituted in the symbolic expression for the
integral (using subs(...) command in Matlab). At a certain point it is not possible to calculate every
integral analytically because of the complexity of the expression. From that point on, one has to rely
on numerical integration. The advantage of this second method is that the integrals are calculated
more accurately. The disadvantage of this method lies in the roots of the programming language. The
substitution is performed not by Matlab itself but by Maple. Calling Maple functions requires a lot of
time and is therefore inefficient. The last method that is implemented does not differ very much from the
second method: the substitution is no longer performed by Maple. Special function files are used that
contain the analytical expressions of the integrals. When the functions are called they only calculate the
numerical value of the integral. This way calling Maple is avoided.
The implemented models calculate the kinematic friction coefficient for a certain range of slip velocities. As mentioned above, some integrals are necessarily evaluated numerically with the trapezoid
method. A certain number of points has to be chosen to calculate the integrals. Table 2.1. shows the
time required to calculate the friction coefficient for the different methods. The number of velocities
solved for is 50 and the number of points used for numerical integration is 100. It is clear that the third
method reduces calculation time significantly and that this method is therefore used.
Table 2.1: Calculation time for different methods of integration, 100 integration points, 50 velocities
Method
Calculation time [s]
1st method: all numerical integration
72
2nd method: analytical + numerical integration with substitution
108
3rd method: analytical + numerical integration without substitution 1.6
The calculation of the friction coefficient with method 3 can be divided into the following steps:
for one certain velocity the following calculations are performed

R 2
cos())
calculation of q 3 C(q) 0 E(qv
(1 2 )0 d analytically

Rq
R 2
cos())
calculation of G(q) = 18 qL q 3 C(q)dq 0 E(qv
2
(1 )0 d by means of the trapezoid integration
method


calculation of P (q) = erf 21G
calculation of

R 2
0

cos()Im

calculation of =
integration method

1
2

R q1
q0

E(qv cos())
(1 2 )0

q 3 C(q)P (q)dq

d analytically

R 2
0

cos()Im

these steps are repeated for every velocity v

E(qv cos())
(1 2 )0

d by means of the trapezoid

2.3.2

Friction coefficient for different slip velocities

With the implemented model the kinematic friction coefficient can be calculated for different ranges of
length scales considered. The range of length scales is determined by the upper and lower distance cutoff:
q0 < q < q1 = q0
The friction coefficient curve is calculated for different values of the maximum magnification max .
These results are shown in figure 2.4. As discussed in paragraph 2.1, a broader range of length scales
(larger max ) leads to a broadening of the curve and a raise of the peak maximum. The values for the
parameters used in the calculations are listed in Appendix C. These results are validated with the results
obtained by Persson in [4].
Kinematic friction model without flash temperature
9
max = 10
max = 100

max = 1000

kinematic friction coefficient

7
6
5
4
3
2
1
0
4
10

10

10
slip velocity [m/s]

10

10

Figure 2.4: Kinematic friction coefficient for different maximum magnifications

In paragraph 2.1 the influence of the apparent contact area has been discussed. When shorter length
scales are considered the apparent area of contact decreases. In other words, the value of P (q), which
describes the ratio between apparent area of contact and nominal contact area, decreases with increasing
magnification. This leads to a strong reduction of the contribution to the friction force from the smallscale roughness of the road. The reduction in apparent area of contact is visualized in figure 2.5 with
P () for different slip velocities. These curves are qualitative in agreement with the curves published
by Persson in [4]. However, the heights of the curves differ with those of Persson. Probably, the curves
published by Persson have been produced with other parameter values.

2.3.3

Friction coefficient for different road profiles

The characterization of a road profile is discussed in Appendix A. The power spectral density C(q) can
be used to describe the road surface. For self-affine fractal surfaces the Hurst exponent H determines
the slope of C(q), it tells us the presence of different length scales found in the road surface. A low Hurst
exponent means that all length scales (in the considered domain) are all present in a same amount. This
will lead to more energy dissipation and hence to a higher friction coefficient. The kinematic friction
coefficient has been calculated for different values of the Hurst exponent. The corresponding power
7

Apparent contact area

10

10

v = 0.00005 m/s
v = 0.013 m/s
v = 0.316 m/s
v = 7.85 m/s
v = 187 m/s

P []

10

10

10

10

20

40

[]

60

80

100

Figure 2.5: Ratio of apparent area of contact to nominal contact area for different slip velocities

spectral densities are shown in figure 2.6. The results are visualized in figure 2.7, where max = 100 was
used for the maximum magnification. The other parameter values are those tabulated in Appendix C.
These results are again identical to the results published in [4].
This model does not include the effect of the raise of temperature of the rubber due to the energy
dissipation of the rubber. In reality the rubber will heat up and the characteristics of the rubber will
change. This effect of local warming of the rubber, referred to as the flash temperature, will be considered
in the next chapter.

Power Spectral Density of asphalt road surfaces

14

10

H = 0.2
H = 0.4
H = 0.6
H = 0.8

15

10

16

10

17

C [m4]

10

18

10

19

10

20

10

21

10

22

10

10

10

10

10

Wavevector q [1/m]

Figure 2.6: Power spectral densities of road surfaces with different Hurst exponents

Kinematic friction model without flash temperature


25
H = 0.2
H = 0.4
H = 0.6
H = 0.8

kinematic friction coefficient

20

15

10

0
4
10

10

10
slip velocity v [m/s]

10

10

Figure 2.7: Kinematic friction coefficient for different Hurst exponents

10

Chapter 3

Rubber friction with temperature


dependence
3.1

Background of the friction model

In this chapter the friction model with temperature dependence will be discussed. In practical situations
the friction of the tire with the road will raise the temperature of the tire. Especially for rubbers, this
has an effect on the energy dissipation due to internal friction, because the viscoelastic modulus strongly
depends on the temperature and therefore the modulus can be written as:
E = E(, T )
Just below the surface of the rubber tire (small length scale , i.e. a large wave vector q) the temperature will be higher than deep inside the tire. Therefore, the local temperature (flash temperature)
has to be calculated for different length scales. Persson developed a theory in [5] that provides the basis
of the model discussed in this chapter.
The basis for the theory starts from the heat diffusion equation, based on stationary sliding:
D2 T =

Q
,
Cv

with D = th /Cv the heat diffusivity, the mass density, th the thermal conductivity and Cv the
heat capacity of the rubber. Q is the heat production per unit volume and unit time due to the internal
friction of the rubber.

3.2

The friction model

This paragraph contains the equations that make up the temperature dependent model. In [5] Persson
shows that the heat diffusion equation can be reduced to the following equations that determine the
temperature for a velocity v and length scale = 2/q:
Z

= T0 +
g(q, q 0 )f (q 0 )dq 0
0
Z

1 1 
4q 0
4q 2
0
Dk2 t0
g(q, q ) =
1

e
0 Dk 2
k 2 + 4q 02 k 2 + 4q 2


04
0 Z q1
vq
E(qv cos()), Tq
0
0 P (q )
f (q ) =
C(q )
cos()Im
d
Cv
P (qm ) q0
(1 2 )
Tq

(3.1)
(3.2)
(3.3)

where T0 is the background temperature and t0 is roughly half the time a rubber patch is in contact with
a so called macro-asperity contact region. When a rubber is squeezed against a hard surface it looks as
11

if complete contact occurs at many regions. These regions are called marco-asperity contact regions and
these can be observed at a magnification m = qm /q0 . Typically this magnification appears to be in the
range 2 < m < 5. If the magnification is increased the macro-asperity contact regions break up into
much smaller and closely separated micro-asperity contact regions. The separation of the macro-regions
is relatively large and thus the interactions between these regions can be neglected.
The half-time a rubber patch is in contact with a macro-asperity contact area can be approximated
by:
t0 =

Rasp
,
v

with:
Rasp =

a + bP (qm )c
,
qm

where a, b and c are constants.


The friction coefficient can be determined in a way similar to the temperature-independent model.
However, the visco-elastic modulus now depends on the flash temperature of the rubber:

=
P (q)

G(q)



Z 2
E(qv cos(), Tq )
q 3 C(q)P (q)dq
cos()Im
d
(1 2 )0
q0
0


1

erf
2 G

Z
Z 2
E(qv cos(), Tq )
1 q 3


q C(q)dq
(1 2 )0 d
8 q0
0
1
2

q1

(3.4)
(3.5)
(3.6)

The relation between the visco-elastic modulus and the temperature can be approximated by the
temperature-frequency shift described by the Williams-Landel-Ferry equation (WLF-equation) [2]:
E(, Tq ) = E(aT , T0 )

(3.7)

where the shift factor aT is given by:


log10 (aT ) 8.86

Tq Tg 50
,
51.5 + Tq Tg

here Tg is the glass transition temperature of the rubber. The WLF-equation states that at a certain
temperature Tq and frequency the modulus is equal to the modulus considered at a temperature T0
and frequency aT .

12

3.3
3.3.1

Numerical results
Implementation of the model

Equations 3.4-3.6 determine the temperature of the tire rubber for a certain velocity v and length scale
. It is not possible to obtain an explicit expression for the flash temperature due to the complexity
of the equations. Therefore the temperature and corresponding shift factor aT has to be calculated in
an iterative way. The iterative method that is applied is the bisection method. This method uses a
predefined range that has to contain the answer Tq . The range is divided into two sections. The section
that doesnt contain the answer is discarded and a new range is defined. This method is repeated until
the solution is below a certain specified tolerance. The calculation of the temperature and the shift factor
can be divided into the following steps:
for one velocity v calculate P (m ) and the half-contact time t0

R 1 
2
4q 0
4q 2
1 eDk t0 k2 +4q
loop over different wave vectors q and calculate g(q, q 0 ) = 1 0 Dk
2
02 k 2 +4q 2 dk
by means of the trapezoid integration method
start the iteration loop until error < tolerance
T T 50

q
g
calculate the shift factor aT with log10 (aT ) 8.86 51.5+T
for the last calculated temperature
q Tg
Tq

calculate G(q) analytically and calculate P (q 0 )




R q1
0
04
E(qv cos()),Tq
0 P (q )
calculate f (q 0 ) = vq
cos()Im
C(q
)
d analytically
Cv
P (qm ) q0
(1 2 )
calculate the new temperature Tq = T0 +
method



T

1
calculate the error: error = Tq,new

q,old

R
0

g(q, q 0 )f (q 0 )dq 0 by means of the trapezoid integration

apply the bisection method and discard one half of the range of temperatures
end the iteration loop if error < tolerance and store the temperature Tq and shift factor aT
perform this iteration loop for all wave vectors q and velocity v
The nature of the calculation of the friction coefficient doesnt differ very much from the temperatureindependent model. The difference lies in the fact that the shift-factor aT is included in the analysis.
For each velocity a range of shift factors aT is calculated for a range of wave vectors q. These are used
to calculate the viscoelastic modulus at the calculated temperatures through the use of the temperaturefrequency shift. The subsequent steps are identical to those discussed in paragraph 2.3.1.

3.3.2

Friction coefficient and temperature for different slip velocities

With the method described above, the kinematic friction coefficient can be calculated for a range of
velocities. The temperature at magnification is also calculated. The same rheological model is used in
this temperature-dependent model (see appendix B). The parameter values can be found in appendix
C. These are different from the ones used in the temperature-independent model because Persson uses
different values in [5].
The results for the friction coefficient are depicted in figure 3.1. For comparison the results without
flash temperature are shown too (constant background temperature). The friction coefficient lowers
when the flash temperature is included in the analysis. This can be easily understood: when the rubber
warms up, the visco-elastic modulus shifts to higher frequencies (see WLF-equation 3.7). This leads to
a more elastic, less viscous, rubber. The frequency where the loss tangent reaches its maximum is in
practical applications in tire friction always higher then the excitation frequency. Therefore, a shift to
higher frequencies will lead to less energy dissipation. This is visualized in figure 3.2.

13

The effect of a decrease of the friction coefficient is also found by Persson. However the quantitative
results presented here are not in good agreement with the results published by Persson. He uses measurement data for the visco-elastic modulus and for the shift factor aT . This data is not published in
the paper and could therefore not be used to validate the model. A good description of the tire rubber
would require a broad range of relaxation times. However, the rheological model used here only uses one
relaxation time. This could explain the difference in the results obtained.

Kinematic friction model comparison


8
without flash temperature
with flash temperature

kinematic friction coefficient

7
6
5
4
3
2
1
0
8
10

10

10
10
slip velocity [m/s]

10

10

0.8

0.8
loss tangent

loss tangent

Figure 3.1: Kinematic friction coefficient for models with and without flash temperature

0.6
0.4
0.2
0

0.6
0.4
0.2

10

10

10

[s1]

10
[s1]

Figure 3.2: A shift to higher frequencies due to a higher temperature will lead to less energy dissipation
in practical tire applications

14

The temperature at magnification is also calculated for different velocities. The results are presented
in figure 3.3 for a magnification of = 1 and maximum magnification = max = 400. At a length
scale a rubber volume 3 heats up. The length scale corresponding to = 1, therefore equals
= 2/q0 4 mm. Thus, the temperature at magnification = 1 corresponds to the temperature
4 mm below the surface of the tire. Equivalently, a magnification of = 400 corresponds to a length
scale 2/q0 10 m. The rubber just below the surface can reach temperatures of 145 C. Deeper
inside the rubber, the temperature is of course lower. This can also be concluded from figure 3.3: the
maximum temperature a few millimeters inside the tire reaches only 70 C. The raise in temperature leads
to a frequency shift of the visco-elastic modulus. This leads to less energy dissipation in the rubber as
discussed above. Again, the qualitative results are the same as the results obtained by Persson. However,
the quantitative agreement of the results is relatively poor. As explained above, the difference in the
results could originate from the fact that Persson uses measurement data, and here a simple rheological
model is used.

Flash temperature
150
140

=1
= 400

flash temperature C

130
120
110
100
90
80
70
60
50
8
10

10

10
10
slip velocity [m/s]

10

10

Figure 3.3: Flash temperature for magnification = 1 and maximum magnification max = 400

15

16

Chapter 4

Results for real road surfaces


In the previous chapters the numerical data that was calculated with power spectral densities (psd) as
depicted in figure A.1. The psd is calculated through the use of formula A.1, as proposed by Persson.
It could be questioned if the use of such an idealized surface will generate realistic results. Therefore,
measurements have been made on a real asphalt road surface. From this data, a spectrum of the
road is calculated and compared with the spectrum used in chapter 3. This is discussed in paragraph
4.1. The influence of a measured psd on the friction coefficient will be discussed in paragraph 4.2. In
the subsequent paragraph two real road surfaces are compared since data of another road surface was
available from another project. Finally, the influence of rain on the spectrum and the friction coefficient
will be addressed in paragraph 4.4.

4.1

The power spectral density of a real asphalt road surface

In order to calculate the power spectral density of a road surface, the height profile of the road has to be
measured. Because small length scales are considered in these models, down to the m range, a detailed
scan of the road surface is required.
A piece of asphalt with dimensions 56 x 61 mm2 has been scanned with an optical imaging profiler
(Sensofar). A rectangular grid of 978 x 1088 points has been scanned with a spacing of 55.4 m. Only
small areas can be scanned during one measurement. Therefore, subsequent measurements are partially
mapped with previous measurements. A plot of the measured road surface is shown in figure 4.1.
In [1] Persson describes a way to calculate the power spectral density from surface height data. First,
a two-dimensional Fast Fourier Transform (FFT) is performed to calculate the spatial frequencies that
are present in the surface:
hA (q)

a2
Hm ,
(2)2

where a is the lattice spacing of the data points (55.4 m) and Hm is the two-dimensional FFT of the
surface height. The FFT is performed on discrete data, which leads to leakage: the fourier transform of
the actual frequencies present is smeared out onto neighboring frequencies. Also, the data is not periodic
which leads to aliasing. The most common way to prevent leakage and aliasing is to make use of a
window. The height data is windowed in two dimensions by means of tukey-windows. The advantage
of a tukey window is that the shape of the window can be altered. The ratio between the tapered and
constant section can be set by means of the factor . A value of = 0 corresponds to a rectangular
window and a factor of = 1 corresponds to the hamming window. The use of a hamming window
would alter the height data to much. A rectangular window would not satisfy the periodicity of the data.
A factor = 0.25 is used in the calculation and leads to satisfactory results. The used window and the
windowed data are shown in figure 4.2.

17

Figure 4.1: Asphalt road surface measured with an optical imaging profiler

Figure 4.2: Used 2-dimensional tukey window with = 0.25 (left) and surface data windowed (right)

18

The power spectral density can easily be calculated as:


C(q) =

(2)2
2
h|hA (q)| i,
A

where A is the surface area of the asphalt piece that is studied and h...i stands for surface averaging.
It is assumed that the psd is independent of the direction of q. This makes it possible to calculate
an angular average of C(q) which means C(q) = C(q) with q = |q|.
Through the above described procedure a psd of the measured road surface is calculated. The
resulting spectrum is shown in figure 4.3. The smallest wave vector is determined by the dimensions of
the measured surface: q0 = 2/L = 103 m1 . The largest possible wave vector is determined by the
Nyquist frequency stated by the Nyquist-Shannon sampling theorem: the highest possible reconstructible
frequency equals half the sample frequency. Therefore q1 = /a = 5.6 104 m1 . It can be concluded
from the figure that the calculated spectrum is very similar to the spectrum used for the temperatureindependent model. The Hurst exponent H can be calculated from the slope of the curve. It appears to
be 0.86, only a deviation of 7% with the value that was used in the calculations of the friction coefficient in
the previous chapter (0.8). Persson also published spectra of road surfaces calculated from measurement
data. The results are in excellent agreement with the results presented here.
Power spectral density of the road

10

Calculated spectrum
Spectrum temperatureindependent model
10

10

12

C [m4]

10

14

10

16

10

18

10

20

10

10

10

10

10

Wave vector q [m1]

Figure 4.3: Calculated power spectral density of measured road surface and the psd used in chapter 3

4.2

Friction coefficient for a real asphalt road surface

With the spectrum discussed in paragraph 4.1, the friction coefficient can be calculated for different slip
velocities. All the measured length scales are considered so there will not be any extra cut-off length
scales. A significant difference with the spectra used in the previous chapters is the range of wave vectors
q. The psd of figure 4.3 is limited to 102 < q < 5.6 104 m1 . The psd used in for example the
temperature-dependent model is defined for 1.5 103 < q < 6 105 m1 .

19

In figure 4.4 the calculated psd is shown together with a complete fitted spectrum. The slope of the
curve determines the Hurst exponent. The roll off of the curve is determined by q0 and the height of the
curve is determined by the rms value of the surface height h0 . The resulting friction coefficient for these
two spectra can be found in figure 4.5. Here the flash temperature is not included in the analysis, because
the temperature-independent implementation of the model is validated and the temperature-dependent
model is not. As figure 4.5 indicates, the deviation between the friction coefficients is small. This makes
it possible to make use of a fitted model for a road surface when calculations are performed. However,
one should bear in mind that the upper cut off is now determined by the spacing of the measurement
points. A more detailed analysis of the road surface would yield a larger cut-off wave vector. In practice,
the upper cut off wave vector is determined by contamination of the tire or due to a thermally degraded
layer of tire rubber. Extrapolation of the power density spectrum could therefore lead to more realistic
results. For example, if the spectrum is extrapolated to a upper cut-off of q1 = 6 105 m1 , the friction
coefficient will increase with 58%.
Comparison of calculated and fitted psd

12

10

Calculated spectrum
Fitted spectrum
14

10

16

C [m4]

10

18

10

20

10

22

10

10

10

10

10

Wave vector q [m1]

Figure 4.4: Calculated and complete fitted power spectral density of the measured road surface

20

Kinematic friction model without flash temperature


2.5
Calculated spectrum
Fitted spectrum

kinematic friction coefficient

1.5

0.5

0
4
10

10

10
slip velocity [m/s]

10

10

Figure 4.5: Kinematic friction coefficient for the calculated and fitted spectrum. Flash temperature is
not included

4.3

Comparison of two real road surfaces

In the previous paragraph the results for the friction coefficient for a real road surface (surface 1) has
been discussed. Another road surface has been measured with the same optical imaging profiler during
another project (surface 2). The data of the two roads will be used here to compare the two roads
and the resulting friction coefficients with each other. Road surface 2 consisted of much coarser stone
particles than road surface 1. Therefore direct measurement with the Sensofar was not possible. A
clay impression was made in order to get good quality data. The resulting road surface (negative of
the clay impression) is shown in figure 4.6. With the data from surface 2 a power spectral density is
calculated. The two spectra are shown in figure 4.7. The psd of surface 2 is slightly larger at small wave
vectors. This is due to the lager stone particles that are present in the second road surface. As can be
concluded from the figure, the Hurst exponent of road surface 2 is slightly lower (smaller slope of the
curve) and the smallest length scales (large q) are more present). Calculations have been made for the
temperature-independent model. The resulting friction coefficient is shown in figure 4.8 for different slip
velocities. The two spectra lead to different (v) curves, the difference can be as high as 50%. Road
surface 2 has got a higher psd at small wave vectors, therefore the friction curve is higher for surface
2 at low slip velocities (corresponding to low frequencies and low wave vectors). At high slip velocities
surface 1 is dominant and will therefore exhibit more friction at these velocities. However, caution has
to be taken when interpreting these results: the clay impression that was made of the road surface did
not fill out the deep cavities of the road. Therefore the clay impression is not the exact negative of the
road surface. Furthermore it has not been validated that the clay method will lead to accurate results.
It could be that the clay impression does not resemble the road surface very well, especially on small
length scales. One way to check the validity of this method is to make a clay impression of road surface 1
and scan that with the optical imaging profiler. If the results are identical to the original measurements
(figure 4.1) it is allowed to use a clay impression in the analysis.

21

Figure 4.6: Asphalt road surface (surface 2) measured with an optical imaging profiler

22

Comparison of calculated and fitted psd

12

10

Spectrum surface 1
Spectrum surface 2

13

10

14

10

15

C [m4]

10

16

10

17

10

18

10

19

10

20

10

10

10

10

10

10

Wave vector q [m1]

Figure 4.7: Comparison of the power spectral densities of two real road surfaces

Kinematic friction model without flash temperature


2.5
Spectrum surface 1
Spectrum surface 2

kinematic friction coefficient

1.5

0.5

0
4
10

10

10
slip velocity [m/s]

10

10

Figure 4.8: Comparison of the friction coefficient for two real road surfaces

23

4.4

Influence of rain

It is widely known that the friction between the tire and road decreases when the road is wet. A possible
explanation for the decrease in friction could be the following: rain fills up some of the cavities present in
the road surface. When the rubber is moving over these cavities it seals them off, preventing the rubber
from filling the deep holes in the road surface. The rubber will therefore feel a different corrected road
surface. It is possible to calculate numerically a flutted surface. The function imfill(...) in Matlab
is normally used for imaging processing of pictures with missing data. However, it can be used perfectly
to calculate the corrected road surface. In figure 4.9 a detailed view of a part of the surface (surface 1)
is shown with the original road and the filled road.

Figure 4.9: Detailed view of original road surface and filled road surface

The filling of the cavities will lead to a more smooth road surface. The rms-value of the surface
height will therefore be lower for the wet surface. The power spectral density for the dry surface and
for the wet surface is shown in figure 4.10. The psd of the wet surface is about four times lower then
that of the dry surface. This will therefore lead to less energy dissipation in the rubber. As a result,
the friction coefficient will be lower for the wet surface. A comparison of the friction coefficient for the
spectra of figure 4.10 is shown in figure 4.11. The figure shows what is widely known in practice: the
friction between a tire and the road is lower for a wet surface.

24

Comparison of psd of wet road and dry road

12

10

Wet road
Dry road
14

10

16

C [m4]

10

18

10

20

10

22

10

10

10

10

10

Wave vector q [m1]

Figure 4.10: Comparison of the power spectral densities for the dry and wet road

Kinematic friction model without flash temperature


2.5
Wet road
Dry road

kinematic friction coefficient

1.5

0.5

0
4
10

10

10
slip velocity [m/s]

10

10

Figure 4.11: Comparison of the friction coefficient for the dry and wet road

25

26

Chapter 5

Conclusion and recommendations


5.1

Conclusion

The implementation of the temperature-independent model has been validated with the results presented
by Persson. The influence of different road surfaces and the effect of the range of length scales has been
investigated. A lower Hurst exponent H of the road surface leads to a flatter spectrum and leads to
higher peaks of the (v) curve. If the range of length scales is increased, more length scales will lead to
more energy dissipation in the rubber. Therefore, a larger max will lead to a broadening and increase
of the (v) curve.
The temperature-dependent model includes the heat diffusion equation to account for the fact that
the rubber heats up due to the energy dissipation of the rubber. The qualitative results are as expected:
an increase in temperature softens the rubber and, in practical tire-road situations, leads to less energy
dissipation of the rubber. Therefore the friction coefficient will drop compared to the temperatureindependent case. The temperature of the rubber at different length scales is in relatively good agreement
with the temperatures calculated by Persson. However, quantitative validation of the model was not
possible. Persson uses measurement data for the viscoelastic modulus E() and for the shift factor aT ,
but does not publish this data in his papers.
Successful measurements have been made with an optical imaging profiler on a real road surface. With
this data, the psd of the measured road has been calculated (surface 1). The results are in very good
agreement with the psds published by Persson. The kinematic friction coefficient has been calculated
for this spectrum and for a fitted one. The resulting (v) curves are similar and therefore it is concluded
that fitted psds can be used in the calculations. Comparison between two different real road surfaces
has been made with data available from another project (surface 2). The surfaces are quite different, and
this has an effect on the resulting (v) curves. Surface 2 has got a higher psd at low wave vectors and
therefore exhibits more friction at low slip velocities. At high slip velocities surface 1 is mostly dominant
and therefore surface 1 results in a higher friction coefficient at these high sliding velocities. Caution has
to be taken when interpreting the results because measurements have been made of a clay impression of
the road surface. Another interesting feature that is investigated is the influence of rain on the friction
coefficient. A flutted surface makes the road more smooth and will lead to a lower psd of the road. A
numerical procedure is used to fill the surface with water. The result is what is expected from practice:
the friction coefficient decreases a significant amount when the road is wet.

27

5.2

Recommendations

The main drawback of this report is the lack of validation of the temperature-dependent friction model.
This is due to the fact that the measurement data that Persson uses in his calculations is not published.
Furthermore it would be interesting to perform experiments with rubber blocks and compare the results
with the theory presented here.
The four recommendations for future study on this subject are therefore:
Validate the temperature dependent model
Use measurement data for the viscoelastic modulus and for the shift factor aT . In order to use
measurement data in the implemented model, some modifications have to be made: the analytical
integrals that are used have to be adapted: either numerical integration should be used or the
measurement data has to fitted and integrated analytically. The second method is preferable, if
possible, because it needs much less calculation time. The results obtained from real data will be
more realistic. The rheological model used in this report is a very simple one: only one relaxation
time is considered. A real rubber, however, can only be appropriately described by the use of a
broad range of relaxation times.
Perform experiments to obtain the friction coefficient for a range of slip velocities. This way the
theoretical model can be compared with the experimental data. If possible, temperature should be
considered in the experiments because it has a great impact on the friction coefficient.
Perform measurements on a clay impression of road surface 1. If the results are identical to the
original measurements performed directly on the road surface, the clay method can be used and
the data from road surface 2 is valid.

28

Appendix A

Self-affine fractal surfaces


In practice it has been found that many real surfaces can be described as self-affine fractal surfaces.
A somewhat different surface is a self-similar surface. A self-similar surface has the property that the
statistical properties do not change if a magnified version of the surface is considered. If the equation of
the surface is given by z = h(x, y), then it cannot be distinguished from its magnified version:
z = h(x, y) = h(x/, y/)
A self-affine surface is a bit different. The statistical properties of such a surface do not change if
the magnification in the z-direction is chosen differently. The magnification in the x- and y-direction is
and in the z-direction H . The Hurst exponent H defines the scaling in the z-direction. The Hurst
exponent is related to the fractal dimension Df as H = 3 Df . For a self-affine fractal surface the
following relation holds:
z = h(x, y) = H h(x/, y/)
For a self-affine fractal surface the following relation holds for the power spectral density of the road:

C(q) = k
with k =

h0
q0

2

H
2 .

q
q0

2(H+1)

Here h0 is the rms-value of the height of the road surface.

29

(A.1)

In figure A.1. the power spectral density is shown for a road surface with rms-value h0 = 0.5 mm
and Hurst exponent H = 0.8 is shown.
Power Spectral Density of an asphalt road surface

14

10

15

10

16

10

17

C [m4]

10

18

10

19

10

20

10

21

10

22

10

10

q0

10

10
Wavevector q [1/m]

10

Figure A.1: Power spectral density of an asphalt road surface. The rms surface height h0 = 0.5 mm and
the Hurst exponent H = 0.8

30

Appendix B

The rheological model


In order to use equations 2.1-2.3 to compute the friction coefficient, a rheological model has to be chosen
for the behavior of the rubber. The response of a rubber (like many polymers) depends on the excitation
frequency. In other words, rubbers exhibit viscoelastic behavior.
A simple model that can be used to describe viscoelastic behavior is the Standard Linear Solid (SLS)
model. This rheological model can be visualized as in figure B.1 with a spring in series with a Maxwell
element (spring and damper parallel). The springs have spring constants E1 and E2 and the damper has
a time constant of . The (complex) elastic modulus of this system is described by:
E() =

E1 (1 i )
,
1 + a1 i

here a1 = E()/E(0) is the ratio between the elasticity modulus in the high frequency (glassy) region
E1 and the modulus in the low frequency (rubber) region E2 . Figure B.2 shows the real and imaginary
part of the elastic modulus for the values shown in table B.1.

Parameter
E1
a1

Value
109 Pa
1000
103 s

Table B.1: Parameter values used in the SLS model

Figure B.1: SLS model

31

Complex Emodulus for rheological model

10

10

Emodulus [Pa]

10

10

Re(E)
Im(E)
4

10

10

10

10

10

10
10
frequency omega [rad/s]

10

Figure B.2: Real and imaginary part of the viscoelastic modulus

32

10

Appendix C

Used parameter values


The table below shows the values that are used for the parameters in the temperature-independent model
and in the temperature-dependent model.

Parameter
H
q0
h0
E1
a1

0
Tg
T0

Table C.1: Parameter values used in the friction models


Value temperature-independent model Value temperature-dependent model
0.85
0.8
2000 m1
1500 m1
4
5 10 m
6.67 104 m
9
10 Pa
109 Pa
1000
1000
103 s
103 s
0.5
0.5
0.2 106 Pa
0.4 106 Pa
-30 C
60 C

33

34

Bibliography
[1] U. Tartaglino A.I. Volokitin E. Tosatti B.N.J. Persson, O. Albohr. On the nature of surface roughness
with application to contact mechanics, sealing, rubber friction and adhesion. Journal of Physics:
Condensed Matter, 17:R1R62, 2005.
[2] J.D. Ferry M.L. Williams, R.F. Landel. The temperature dependence of relaxation mechanisms in
amorphous polymers and other glass-forming liquids. Journal of the American Chemical Society,
77:3701, 1955.
[3] B.N.J. Persson. On the theory of rubber friction. Surface Science, 401:445454, 1998.
[4] B.N.J. Persson. Theory of rubber friction and contact mechanics. Journal of Chemical Physics,
115:38403861, 2001.
[5] B.N.J. Persson. Rubber friction: role of the flash temperature. Journal of Physics: Condensed
Matter, 18:77897823, 2006.

35

Você também pode gostar