You are on page 1of 26

H. R.

Dorfi1

A Study of the In-Plane Force Transmission


of Tires2

REFERENCE: Dorfi, H.R., “A Study of the In-Plane Force Transmission of Tires,” Tire
Science and Technology, TSTCA, Vol. 32, No. 4, Oct.-Dec. 2004, pp. 188-213.

ABSTRACT: A better understanding of the tire force transmission mechanisms from footprint
to spindle due to cleat impact and other road disturbances is important to improve the ride per-
formance of the tire/vehicle system. In this study, the in-plane force transmission of tires rolling
over cleats is studied using various approaches: analytical, numerical and experimental. The tire
force response at the spindle is compared and several important observations are made. Results
reported in the literature are critically reviewed as well.

KEY WORDS: tire ride, tire force transmission, cleat envelopment, FTIRE, belt ring equations

One of the important properties of a pneumatic tire is its ability to envelope


obstacles and to reduce the force transmitted into the suspension. Quite often the
analysis of the force transmission can be limited to forces within the wheel plane
(vertical and longitudinal direction). Lateral forces are, in general, small unless
camber or asymmetric disturbances such as oblique cleats are being considered.
A good understanding of the mechanisms of the force transmission due to road
disturbances is important in order to improve the vehicle ride performance. In
this paper, the cleat envelopment and resulting spindle forces are compared using
experimental data, numerical models and analytical equations and compared to
results reported earlier in the literature.
Early experimental work on the subject was done by Gough [1] and Julien
and Paulsen [2], who measured the normal and longitudinal force variation of
tires rolling over cleats. It was seen that for some load and cleat shape combina-
tions a force minimum arises when the cleat is located underneath the wheel cen-
ter. At certain load conditions the tire “swallows” the cleat without any signifi-
cant rise in the static force due to cleat envelopment. Pacejka [3] developed a
simple enveloping model to explain the occurrence of a minimum during cleat
envelopment. Later work by Takayama and Yamagishi [4] was based on a simple
5 degrees of freedom (DOF) tire model: a rigid ring with 2 translations and in-
plane rotation plus 2 additional degrees of freedom to represent the footprint
deflection and cleat envelopment. Their model required the use of an assumed
enveloping function. Correlation with experimental data was good, and both stat-
1 Tire Engineering Technology, Hankook Tire, Akron Technical Center, 3535 Forest Lake Drive,
Uniontown, OH 44685.
2 Presented at the twenty-second annual conference of the Tire Society, Akron, Ohio, September 23-
24, 2003.
186
DORFI ON IN-PLANE FORCE TRANSMISSION 189

ic and dynamic force responses were predicted well with the model. Due to the
advent of high performance computing, more recent work such as the paper by
Zhang et al. [5] focuses on using explicit finite element techniques to integrate
the rolling tire model in order to determine its response to cleat impacts and other
road inputs in the time domain.
Scavuzzo et al. [6] studied experimentally measured tire modes and related
these modes to vehicle interior accelerations. The effect of tire size, load and
inflation pressure are also studied with emphasis on the changes in tire natural
frequencies. In a subsequent study [7], Richards et al. emphasized the effect of
boundary conditions on tire natural frequencies. He compared frequencies as
obtained from experiment and FE models under various boundary conditions and
showed how modal frequencies differed due to boundary conditions, particular-
ly for the lower order modes.
In the present work, the objective is to study tire dynamic behavior related
to in-plane force transmission using analytical tire models and a time integrated
flexible belt tire model, FTIRE, and relate the results to experimental data. The
focus is on both understanding low and high speed cleat envelopment and the
role the tire plays during the force transmission.

Analytical Tire Models

Many analytical belt models have been developed and studied over the years
[8,9,10,11]. For a derivation of these models, the reader is referred to the relevant
literature. Rather than rederiving these results, the relevant equations are pre-
sented as needed to interpret the tire in-plane vibration modes.
Several authors have developed equations of motions for circular rings sup-
ported by a sidewall support stiffness. The basic assumptions are:
• The sidewall stiffness can be represented by springs in the radial, tan-
gential (and lateral) direction. The stiffness of these springs is a result
of the pressure stiffening of the carcass and the sidewall rubber shear.
• The belt can be idealized as a pretensioned ring. The stiffness of this
ring is dependent on the belt bending stiffness, which may be zero for a
membrane belt model, and the belt tension stiffness, which is induced
by the inflation pressure in the tire.
• The belt extensibility of the belt ring is sometimes considered; howev-
er, for typical radial tires belt extensibility is only relevant above 300Hz
and is therefore frequently ignored.
A sketch of an analytical tire model is shown in Fig. 1. The belt is modeled
as a membrane or curved beam under inflation loading. The sidewall provides a
linear distributed support stiffness in the radial (kr) and tangential (kt) direction.
190 TIRE SCIENCE & TECHNOLOGY

FIG. 1 — Conceptual tire ring model with sidewall support. The sidewall provides a linear founda-
tion stiffness while the belt ring is modeled as a curved beam.

Following the work by Soedel [10] and Kim [12], the equation of motions
for the tire model can be derived in a reference frame rotating with the tire (body-
fixed). General solutions to the equations of motion have the form of a traveling
wave. If we denote the radial belt motion by w(θ,t) , where θ is a body-fixed
angle on the belt and t is the time, the two traveling waves associated with wave
number n are given by

wn1, 2 (θ , t ) = An1, 2 ei ( nθ +ω 1,2 t ) (1)

These waves describe complex modes; they are characterized by the cou-
pling of both the space variable θ and time variable t. Unlike normal modes, the
belt motion does not reach an extremum at all locations θ at the same instant in
time. A requirement for solutions to the equations of motion to exist is that the
angular wave number n and the in-plane modal frequencies ωn have to satisfy an
equation. This dispersion equation for the ring model is given by

2Ωn ) 2 (n 2 − 1)2 Ω 2
ω nB1, 2 = ± ω n
− (2)
n2 + 1 (n 2 + 1)2
DORFI ON IN-PLANE FORCE TRANSMISSION 191

where ωnB1,2 are the two angular frequencies associated with the circumferential
wave number n; and the superscript B denotes that it is measured with respect to
a body fixed (rotating) coordinate system; Ω is the rotational velocity of the tire
and ω̂n is given by

) n 2 (n 2 − 1)2 (n 2 − 1)2 n2 1
ωn = c B
+ c I
+ cR + 2 c (3)
2
n +1 2
n +1 2
n +1 n +1 T

where the bending coefficient cB , inflation coefficient cI , radial support coeffi-


cient cR , and tangential support coefficient cT are given by

2π EI 2π TI kR AR kT AR
cB = ; c I
= ; c R
= ; cT
= (4a,b,c,d)
mb R3 mb R mb mb

In the above equations, EI represent the belt bending stiffness; TI the cir-
cumferential tension force in the belt; AR is the belt surface area (AR = 2π R b);
and mb is the belt mass.
Alternatively, the equations of motion can be derived in a non-rotating ref-
erence frame [9], using an absolute angle ψ rather than a relative angle θ. The
absolute angle is simply the angle due to the rotational velocity Ωt plus the rela-
tive angle θ

ψ = Ωt + θ (5)

Replacing θ by ψ in Eq. (1) yields

wn1, 2 (ψ , t ) = An1, 2 ei ( nψ − nΩt +ω 1,2 t ) = An ei ( nψt + (ω 1,2 − nΩ )t ) (6)

The term associated with the time is now the angular frequency ωI1,2 in a
non-rotating fixed coordinate system.

ω1I, 2 = ω1B, 2 − nΩ (7)

Thus, the inertia fixed angular frequency differs from the frequency meas-
ured in the rotating coordinate system by -nΩ, the rotational velocity multiplied
by the wave number. With Eq. (7), the dispersion Eq. (2) can be expressed in
terms of ωI1,2 as
192 TIRE SCIENCE & TECHNOLOGY

Ωn(n 2 − 1) ) 2 (n 2 − 1)2 Ω 2
ω nI1, 2 = − ± ω n
− (8)
n2 + 1 (n 2 + 1)2

Note that for Ω = 0 ωn1,2 equal ± ω̂n , the frequencies of the non-rotating tire.
For non-zero Ω, the frequencies ωn1,2 are no longer equal in magnitude. This
bifurcation effect can be explained by the fact that the disturbance travelling in
the direction of the rotating belt mass moves faster compared to the one travel-
ing in the opposite direction.

In-Plane Tire Vibration Modes


Many of the vibrational properties of tires can be understood from the non-
rotating tire. Therefore, we will focus initially on Eqs. (2) and (3) for Ω = 0.
The eigenmodes associated with the non-rolling tire simplify to normal
modes, where the time and space variables are separated

)
wn (θ , t ) = An sin nθ sin ωˆ nt (9)

FIG. 2 — In-plane vibration modes for a non-rolling belt ring model. Circumferential wave numbers
n=1, 2, 3, 4 are shown. The orientation of these modes is arbitrary.
DORFI ON IN-PLANE FORCE TRANSMISSION 193

The vibration modes for in-extensible belts are stationary sinusoidal waves;
the number of peaks on the belt represents the wave number n. The first ring
vibration modes are shown in Fig. 2 for the wave numbers of n = 1, 2, 3, 4. Note
that because of the axi-symmetry of the tire, the orientation of the modes is arbi-
trary,
Closer inspection of Eq. (3) shows that four different terms are associated
with the belt bending stiffness, tension stiffness and sidewall support. These
coefficients together with the wave number n determine the in-plane frequencies.
The relative importance of each term in the above equation will depend on the
relative magnitude of the coefficients cs. These coefficients can be determined
from experimental data. Since only four coefficients are required, the frequencies
of four tire modes have to be measured. The coefficient associated with the side-
wall torsion (cT) can be determined from another in-plane mode, the in-plane tor-
sion mode, which is simply a rigid belt rotation against the sidewall torsion stiff-
ness. The torsion mode frequency is related to the torsion coefficient simply by

ω Tor = cT (10)

With Eqs. (3) and (10), the four coefficients cB, cI, cR and cT can be calcu-
lated using measured or predicted frequencies of the torsion mode and the first
three in-plane belt modes. If more than three in-plane belt modes are available, a
least squared approach can be used to determine the coefficients.
It is insightful to study the individual terms in Eq. (3). The bending stiffness
coefficient is multiplied with a different factor compared to the inflation coeffi-
cient. This implies that bending stiffness and inflation stiffness have a different
effect on the vibration response. The terms dependent on the wave number n can
be written in terms of the order O(n p) where p represents the power of each term

[ 4 2 0
ω n1, 2 ≈ O(n ) cB + O(n ) cI + O(n ) cR + O(n ) cT −2
] 2 (11)

From the above equation, we notice that the contribution of cB to the fre-
quency will be of O(n2) and the contribution of the inflation will be of order
O(n1). Experimental data shows that there is essentially a linear relationship
between wave number and frequency for n = 1 through 8, which implies that the
bending contribution is rather small compared to the inflation stiffening below
200Hz.

Experimental Data and Comparison to Analytical Ring Equations


To compare the analytical equations to experimental data eigenfrequency
data, a tire of size P235/70R17 was obtained. The tire was mounted on a standard
194 TIRE SCIENCE & TECHNOLOGY

rim, inflated to 240 kPa (35 psi) and the fixed spindle unloaded eigenmodes and
frequencies were measured (see Table 1). In addition, an FE model was built and
correlated to the experimental data. Its predicted eigenfrequencies are shown in
Table 1 as well.

TABLE 1 — Experimentally measured and FE predicted frequencies of P235/70R17 tire.


Circum. Wave Nr Exp. Freq. FE Freq.
Mode [Hz] [Hz]
Torsion N/A 45.9 47.8
Belt-Bend 1 1 70.8 71.3
Belt-Bend 2 2 92.9 93.5
Belt-Bend 3 3 113.6 113.2
Belt-Bend 4 4 133.5 133.4
Belt-Bend 5 5 153.7
Belt-Bend 6 6 173.5
Belt-Bend 7 7 192.5
Belt-Bend 8 8 210

Using Eqs. (3) and (10), the four belt ring model coefficients can be deter-
mined using either the experimental data or the FE data. Since there are four
coefficients to be determined, using more than four frequencies over-constrains
the system of equations and requires a least squared approach. The results are
shown in Table 2 using the first four and all five experimental modes and using
all nine FE modes. For all curve fits, the ring bending stiffness coefficient cB is
found to be negative. This is physically not meaningful. However, the coefficient
is rather small and is even smaller when a larger number of modes are used.
Therefore, all curve fits were repeated by ignoring the coefficient cB in Eq. (3)
(cB = 0). This is equivalent to assuming that cB, the bending stiffness of the belt,
is small enough to be ignored for the lower modes and that a membrane model
of the belt is adequate for lower order modes.

TABLE 2 — Ring model coefficients calculated from experimental and FE data.


cb cI cR cT
Nr of modes used [(rad/sec)2]
to fit data
4 exp. modes -1,353 46,518 312,609 83,174
4 exp. modes, CB = 0 0 41,107 312,609 83,174
5 exp. modes -511 38,200 319,787 82,163
5 exp. modes, CB = 0 0 29,089 339,062 78,681
9 FE modes, CB = 0 0 22,722 406,329 67,089
DORFI ON IN-PLANE FORCE TRANSMISSION 195

Figure 3 compares the experimental and FE frequencies to the predicted fre-


quencies based on the ring model using four or five experimental modes and
either setting cB to zero prior (▲,* no cB) or after the curve fit (*,●, cB→ 0). The
best correlation between the ring model and the experimental and FE modes is
achieved if all five experimental modes are used to determine the coefficients and
cB is ignored prior to the curve fit. This result again suggests that the bending
stiffness is small enough and can be ignored for the lower modes (wave number
n ≤ 8).

FIG. 3 — Comparison of predicted belt-ring frequencies based on experimental data.

Figure 4 shows similar results, but with the curve fit performed based on the
predicted FE frequencies (total of nine modes). Good correlation is achieved for
all curve fit approaches. If the negative bending stiffness is considered in the
curve fit, the best correlation is achieved. However, the negative bending stiffness
is not considered meaningful. Recall that we assumed that the belt is inextensi-
ble. This assumption may introduce an error at the higher frequencies.
If the negative bending stiffness is ignored initially, the modes are fit well,
with errors mainly in the first three modes (n = 0, 1, 2). If the bending stiffness
is set to zero after the curve fit, good correlation is found up to n = 5.
It is worth noting that the FE model tends to under-predict the higher fre-
quencies. The reason for this trend appears to be due to a finite element dis-
cretization error. A comparison of tire modes with different refinement shows
that models with increased circumferential refinement have higher frequencies.
196 TIRE SCIENCE & TECHNOLOGY

Therefore, the ring model may actually be more accurate than it appears based
on the comparison to the FE model.

FIG. 4 — Comparison of predicted belt-ring frequencies based on FE data.

In summary, the membrane belt ring tire model correlates quite well with
both the experimental data and the FE results up to about n = 6 (170Hz), if bend-
ing stiffness is neglected.

Relative Contribution of the Stiffnesses to the Eigenfrequencies


To get a better understanding of how the different tire stiffness parameters
contribute to the eigenfrequencies, it is instructive to look at each part of Eq. (3).
For this purpose, we set the coefficients cs to the values which yielded the best
fit based on the experimental data:
CB = 0; CI = 29089; CR = 339062; CT = 78681

The units for the above ring coefficients are [rad/sec] 2. Equation (3) togeth-
er with fn = ωn/2π yields:

(n 2 − 1)2 cI n 2 cR 1 cT
fn = 2 2
+ 2 2
+ 2 (12)
n + 1 4π n + 1 4π n + 1 4π 2

This equation can be interpreted graphically. If fn is interpreted as the length


of three-dimensional vector, then each of the three terms in the above equation
represents the length along one axis:
DORFI ON IN-PLANE FORCE TRANSMISSION 197

fn = xn 2 + yn 2 + zn 2 (13)

Using this interpretation and the values for the coefficients cs from above, a
graphical vector representation of each mode in the inflation, radial and torsion-
al stiffness coefficient “space” can be obtained. The results are shown in both
Fig. 5 as a vector plot and in Fig. 6 as a bar chart. In Fig. 5, the length of each
line represents the frequency of the mode and the number indicates the wave
number. The dashed lines indicate the projected vector onto the x-y plane
(Inflation-Radial Stiffness). The torsion mode frequency depends only on the tor-
sion stiffness. The first in-plane mode is driven by both the radial stiffness and to
a lesser extent the torsional stiffness. For the higher frequencies, the contribution
of the radial stiffness remains constant while the inflation stiffness contribution
increases linearly. For the 4th in-plane mode, the contribution of the inflation
stiffness and the radial sidewall stiffness are about equal, and the torsional stiff-
ness is now a minor contributor.

FIG. 5 — Graphical interpretation of stiffness contribution to ring vibration modes.

It is important to note that with only the in-plane sidewall coefficients and
the belt tension coefficient, a relatively large number of modes can be described
and the contribution of each stiffness varies significantly for different modes.
Thus, using only three parameters, the in-plane tire vibration up to 200Hz can be
computed. Vice versa, by measuring the torsion mode and the first two in-plane
modes, the sidewall coefficients and belt tension can be estimated.
198 TIRE SCIENCE & TECHNOLOGY

FIG. 6 — Comparison of stiffness contribution to ring vibration modes.

Comparison of Ring Model Coefficients


Using the aforementioned technique, belt ring coefficients were determined
based on the best fit to the experimental data for a P235/70R17 at two inflation
pressures (240 and 260kPa) and a P175/70R14 at 240kPa. The data is shown in
Table 3. For the P235/70R17 tire, all coefficients increase in magnitude due to an
increase in inflation pressure, which is to be expected. The belt tension coeffi-
cient cI increases by about 7%, which is comparable to the increase in inflation
pressure of 8%. The radial coefficient cB increases by 5% and the torsion coeffi-
cient cT increases by 3%. This is also plausible, since the torsion coefficient is
only partly dependent on the tension in the sidewall cords but also depends on
the sidewall rubber shear. The coefficients of the P175/70R14 tire are similar in
magnitude to the P235/70R17 tire.

TABLE 3 — Comparison of ring model coefficients.


Size IP cB cI cR cT
[kPa] (psi)
P235/70R17 240 (35) 0 29,089 339,062 78,681
P235/70R17 260 (38) 0 31,143 357,520 81,010
P175/70R14 240 (35) 0 36,784 414,739 123,257
DORFI ON IN-PLANE FORCE TRANSMISSION 199

With Eqs. (3), (4) and (10) as well as the belt mass and geometric tire param-
eters, the belt ring coefficients can be converted into belt tension TI and sidewall
stiffness coefficients kR and kT. The belt tension can now be compared to the ten-
sion force in a hoop under pressure (without sidewalls). This force should be sim-
ilar to the belt tension. Since the sidewalls carry some of the pressure load, the
predicted tension force should be smaller than the tension force calculated from
a simple inflated hoop model. The force in the belt using the simple inflated hoop
model is found from

THoop = InflPressure * RadiusBelt * Widthbelt (14)

Table 4 shows the predicted belt tension force and the hoop force together
with the sidewall coefficients. The interpretation of these coefficients is of inter-
est. The coefficient cI is proportional to the belt tension and inversely propor-
tional to the belt mass times belt radius. Thus, we write

2π TI 2π THoop 2π pRb p
cI = ≈ = 2
= (15)
mb R mb R 2π R bhρ B R(hρ B )

where
p K Inflation Pressure h K Belt Height
b K Belt Width ρb K Average Belt Density

TABLE 4 — Comparison of tire stiffness coefficients.


Tire Belt Belt Belt Belt Hoop kR kT
Mass Radius Width Tension Tension [N/m3] [N/m3]
[kg] [m] [m] TI [N] THoop [N]
P235/70R17 8.75 0.371 0.175 15,030 15,670 7.27x106 1.69x106
P235/70R17 8.75 0.371 0.175 16,090 16,980 7.67x106 1.74x106
IP a)

P175/70R14 5.25 0.285 0.13 8,760 8,940 9.35x106 2.78x106


a) Increased inflation pressure.

The result of Eq. (15) is a simple estimate of the belt tension coefficient cI ;
it is proportional to the inflation pressure and inversely proportional to the belt
radius and (hρB), the weight of the belt package per unit surface area. This is
indeed observed for the P235/70R17 tire. By increasing the inflation pressure by
about 8%, the coefficient cI increased by an equivalent percentage.
200 TIRE SCIENCE & TECHNOLOGY

The coefficients kR and kT listed in Table 4 are to be interpreted as bedding


stiffness and represent the radial and torsion stiffness per unit surface area of the
belt. These coefficients can be multiplied by the belt width to convert them into
a stiffness per unit length of the belt in circumferential direction. It is interesting
to note that the radial and tangential stiffness are larger for the P175/70R14 tire.
The tangential stiffness kT is primarily driven by the rubber shear in the sidewall
and the inflation induced shear stiffness. The radial stiffness kR is dependent on
the pressure stiffening and the inflated shape of the tire.

Rolling Effect
As discussed earlier, the eigenfrequencies satisfy the dispersion Eqs. (2) or
(8) depending on whether a rotating or non-rotating coordinate system is used to
describe the modes. With the ring coefficients determined earlier, the dispersion
equation can be plotted in graphical form (see Fig. 7). The theoretical dispersion
relation shows the locus of all possible frequencies and wave numbers. The solid
line in Fig. 7 shows the non-rolling condition discussed earlier. Because of the
continuity of the deformation at [0,2π], only integer values of n satisfy the con-
tinuity condition. Two distinct branches exist for the rolling condition: the upper
branch reflects a wave traveling with the rolling direction; and the lower branch
represents a wave traveling against it. Also, the observed frequency is a function
of the reference frame, an important distinction with respect to the interpretation
of test data. The dispersion equation is plotted as a continuous line because a
rolling tire allows non-integer wave numbers. This is because of the fact that now

FIG. 7 — Dispersion equation for non-rolling and rolling condition (Ω = 5Hz). The dispersion equa-
tion shows the locus of all possible frequencies/wave numbers.
DORFI ON IN-PLANE FORCE TRANSMISSION 201

two distinct solutions exist for each wave number in the dispersion relation. The
permissible wave numbers are dependent on the boundary conditions; several
examples are given in Ref. [9]. It is interesting to note that for n = 1, the disper-
sion equation for a fixed reference frame shows no dependence of the wave speed
on rolling speed.

Numerical Tire Models – FTIRE

Model Concept
FTIRE [13], or Flexible belt TIRE, has recently been developed as an alter-
native to classical modal dynamics tire models. Instead of linearizing the tire
model around a base state and a set of boundary conditions, FTIRE is a fully
time-integrated non-linear tire model. However, because of its relatively small
number of degrees of freedom (around 500 DOF) and its fast integrator, typical
simulation times are only 10 times slower than real time.
This model strikes a compromise between a detailed finite element model
and simplified tire models, which do not adequately represent the physics of the
rolling pneumatic structure. FTIRE retains some of the simplicity of simpler
models by using a simplified geometry description but includes a flexible dis-
cretized belt. Since FTIRE models the physics of the tire structure, its input
parameters have a direct physical interpretation, which makes this model quite
useful for parametric studies.
A typical FTIRE model has on the order of 100-200 belt segments. The belt
is constrained to the wheel by non-linear sidewall springs and dampers. Contact
to the road is established via tread blocks, which allow shear deformation and
compression. The frictional model allows for pressure and sliding velocity
dependent friction factors. FTIRE is very similar to the tire ring concept shown
in Fig. 1 with a discretized belt and tread blocks. For further details on the mod-
eling concept of FTIRE the reader is referred to Ref. [13] or the Web site
www.ftire.com.
FTIRE makes use of the fact that the lower order tire vibration modes are
closely related to the tire stiffness and mass properties. Therefore, instead of
using the sidewall and belt stiffness parameters as inputs, FTIRE preprocesses
the modal frequency data of several low order modes to determine the sidewall
stiffness and damping properties and several of the belt stiffness properties. It is,
therefore, a very attractive model to use since many of its inputs are readily deter-
mined.
In this study, FTIRE was used to study its effectiveness to simulate cleat
envelopment. Both low and high speed cleat envelopment were studied. Results
and an interpretation of their significance are presented in subsequent chapters.
202 TIRE SCIENCE & TECHNOLOGY

FTIRE Model Input Data


One of the challenges with any tire model is determining the input parame-
ters. In fact, much of the usability of a model hinges on the fact of whether its
input parameters are easily determined. One of the attractive properties of FTIRE
is the fact that its input parameters are relatively easily measured in the lab or
predicted via FEA models. Simple stiffness and vibration measurements are used
to characterize the tire properties. To characterize the sidewall stiffness and
damping properties and also some of the belt properties, tire eigenfrequency data
is needed. This data is readily obtained with simple vibration testing of the tire
on a fixed hub, inflated and unloaded. For the tire used in this study, the fre-
quency data shown in Table 5 was obtained. F1, F2 and F4 are used to define the
sidewall stiffness and damping properties. F5 and F6 define the in-plane and out-
of-plane bending stiffness of the belt. However, as mentioned earlier, FTIRE is
not a modal model. Eigendata is only used during preprocessing to identify the
stiffness and damping characteristics of the tire model.

TABLE 5 — P235/75R17 Tire: typical fixed spindle vibration modes.


Mode / FTIRE Label Freq. [Hz] Damp [%]
In-plane torsion / F1 41 3
In-plane translation / F2 64 2
Lateral translation / F3 34 1
Out-of-plane rotation / F4 39 1
1st in-plane belt bending / F5 86 2
1st out-of-plane belt bend / F6 69 2

FIG. 8 — FTIRE model during cleat impact simulation. Footprint contact forces are shown.
DORFI ON IN-PLANE FORCE TRANSMISSION 203

Several other parameters are needed to generate a complete FTIRE model:


basic tire geometric data, tread stiffness data and friction properties. For further
information, the reader is referred to Ref. [13]. An example of an FTIRE model
during cleat impact is shown in Fig. 8.
It is important to note that FTIRE is not a 2-D tire model; rather, it is capa-
ble of producing cornering forces as well as in-plane forces. However, in the
present study, we will limit ourselves to evaluate the model performance to cleat
impact. Because of the symmetry of the cleat and tire model with respect to the
wheel plane, only the in-plane forces will show any significant response.

FIG. 9 — Simple tire enveloping model.

Slow Rolling Quasistatic Cleat Envelopment

Measurement and FTIRE Model


Cleat envelopment at low velocities permits the study of the enveloping
effect of the tire without regard to tire dynamics. Of interest is the rise in normal
force as a disturbance rolls through the footprint. It had been observed in Ref.
[1,2] that the enveloping force is dependent not only on the size of the distur-
204 TIRE SCIENCE & TECHNOLOGY

bance but also on the loading condition of the tire. To study the effect of tire load-
ing, slow speed cleat envelopment force measurements were performed on the
reference tire (P235/75R17) as it rolled over a 14 mm cleat. The static preload on
the tire was varied from 2670 to 9790N (600-2200 lbf) in increments of 1780N
(400 lbf). The preload was applied under displacement control prior to rolling the
tire over the cleat. Figure 10 shows the result of both the experiment and the sim-
ulation as the tire envelopes the cleat. The cleat is centered at about x = 250 mm.
It is apparent from the data that at loads above 6230N (1400lbf) a local force min-
imum exists as the cleat is centered over the tire. The agreement between FTIRE
and the experimental data is very good.

FIG. 10 — Measured and predicted normal force during quasistatic cleat envelopment for preloads
ranging from 2670 to 9790N (600-2200 lbf). The line width in the graphs increases with preload.

FIG. 11 — Measured and predicted normalized normal force during quasistatic cleat envelopment
for preloads ranging from 2670 to 9790N (600-2200 lbf).The line width in the graphs increases with
preload.
DORFI ON IN-PLANE FORCE TRANSMISSION 205

The data shown in Fig. 10 is normalized with respect to the static preload
and plotted in Fig. 11. The rise in normal force as a portion of the preload is
largest at low loads, at 2670N (600 lbf) the normal forces rises to more than
180% of its preload. At high loads, the cleat center normal force nearly drops
back to the static preload; the tire “swallows” the disturbance.
It is important to note that for quasistatic rolling FTIRE is modeled with a
linear sidewall support stiffness. Therefore, the cleat envelopment versus load
effect cannot be explained with a non-linear sidewall stiffness effect since the lin-
ear sidewall stiffness model shows the same nonlinear enveloping response as the
actual tire.

Simple Enveloping Model


Pacejka [3] originally suggested a simple enveloping model to explain the
occurrence of a minimum force during cleat envelopment. His model is extend-
ed here to suggest a mechanism for the observed phenomenon. The following
assumptions are made:
• The belt is inextensible.
• The tire maintains its inflated radius outside the footprint.
• The sidewall provides a linear radial stiffness support.
• Belt stiffness can be neglected.
• The cleat envelopment near the cleat can be approximated by an envel-
opment factor γ.
Based on these assumptions, a simple enveloping model can be constructed.
A sketch of the tire enveloping model considered here is shown in Fig. 9. The
assumption of belt inextensibility requires that the undeformed and deformed cir-
cumference are equal. This yields the geometric constraint conditions for the tire
circumference:

R0π = R1 (π − φ1 ) + a + γ h (16)

where R0 and R1 are the radii of the undeformed and loaded tire, respectively; φ1
is the angle at which the footprint starts; a is the footprint length; h is the cleat
width and height; and γ is a cleat envelopment factor.
The assumed geometry relates the loaded radius RL and footprint length a to
φ1 and R1 through the equations

RL = R1 cos φ1 a = RL tan φ1 (17)

With the assumption that the sidewall provides a linear radial stiffness sup-
port, the contact force is assumed to be proportional to the deflection d(φ) from
the undeformed configuration of the tire
206 TIRE SCIENCE & TECHNOLOGY

d = R0 cos φ − RL (18)

RL
The line element along the footprint is given by ds = dφ
cos φ .

With the assumption of a linear radial stiffness kr the vertical force Fz in the
footprint can thus be estimated by summing the forces in the footprint

φ1
R
Fz = kr ( Ro − RL + h) h + 2 kr ∫ d L dφ
1442443 φ0
cos φ (19)
underneath cleat
14 42443
outside cleat

where φ0 is h/(2RL). The above equations can be solved numerically to determine


the vertical force Fz as a function of loaded radius RL and input variables radius,
R0 cleat dimension h and cleat envelopment factor γ.
Using the P235/75R17 tire as an example, the simple envelopment model
can predict the normalized center force versus loaded radius. Figure 12 shows the
result of the theoretical model compared to the measured data and the FTIRE
model using a cleat envelopment factor of γ = 1/2. The simple model clearly
reflects the correct trend, which indicates that the observed cleat envelopment
behavior versus load can partly be explained based on the belt inextensibility
constraint, which affects the footprint length during enveloping. Sidewall non-
linearity and belt bending stiffness are not considered in the simple model.

FIG. 12 — Comparison of normalized cleat center force versus loaded radius. Measurement and
FTIRE agree very well. The simple theoretical envelopment model captures the trend correctly.
DORFI ON IN-PLANE FORCE TRANSMISSION 207

Dynamic Tire Response on Cleat Drum Test

As a second study of the model capability, the simulated spindle force


response of an FTIRE model rolling over a cleat was compared to the measured
response for a tire size P235/75R17. The correlation was performed for several
tire constructions, test conditions and rolling speeds. For details on this study the
reader is referred to Ref. [14].
Figure 13 compares the cleat force response at one test condition for the
P235/75R17 tire and 30 kph rolling velocity. The correlation between the FTIRE
simulation and the experimental data is very good with respect to force magni-
tude, damping and frequency content. The data also clearly shows that there are
very distinct frequencies in the force responses of the longitudinal and vertical
direction and that the response is strongly dependent on the test velocity. At 30
kph the vertical direction has a dominant oscillation at about 68Hz. This fre-
quency is the 1st vertical resonance of the tire. While this mode is similar to the
rigid belt translation mode (F2 in Table 5), the slightly higher frequency indicates
an increase in stiffness due to the footprint boundary condition and the reduction
in the participating modal mass. This frequency shift is correctly predicted by the
FTIRE model.

FIG. 13 — Comparison of cleat force response of FTIRE simulation and measurement data at 30 kph
(P235/75R17, 10 mm semi-circle cleat).

In the longitudinal direction, the force response has a significant signal


around 28Hz. Comparing this frequency to the eigenmodes in Table 5, it is appar-
ent that this resonance is not associated with a particular mode of the tire char-
208 TIRE SCIENCE & TECHNOLOGY

acterization test. This is due to the fact that the cleat test involves different bound-
ary conditions, the tire is now loaded against the ground and, therefore, its modal
frequencies have changed.
Figure 14 shows the same test data at 40 kph. Now both the vertical and lon-
gitudinal force response have changed significantly compared to 30 kph. Again
the FTIRE simulation correlates very well with the experimental data. At 40 kph,
the strong vertical resonance at 68 Hz is no longer dominant. A second resonance
at 76 Hz is now visible in both the vertical and longitudinal direction. The source
of this resonance will be addressed later in this paper.

FIG. 14 — Comparison of cleat force response of FTIRE simulation and measurement data at 40 kph
(P235/75R17, 10mm semi-circle cleat).

Simple Rigid Belt Model in Ground Contact


It was seen from Fig. 13 that the tire produces a longitudinal resonance at 28
Hz during the cleat test. The physical significance of this resonance will be dis-
cussed here with the help of a simple rigid belt model of three degrees of free-
dom. Richards [7] showed both experimentally and with FE models, how modal
frequencies change due to boundary conditions. Here, we use a simple model to
show the effect of the footprint and pinned wheel boundary condition on the
fore/aft modes of the tire.
The model is shown in Fig. 15. The degrees of freedom are the belt transla-
tion in the longitudinal direction x, and the belt and wheel rotation θB and θW.
The kinetic and potential energy of the system are given by
DORFI ON IN-PLANE FORCE TRANSMISSION 209

x˙ 2 θ˙W 2 θ˙B 2
T = mB + IW + + IB (20)
2 2 2

1 2 1 2 1 ⎛ K + KR ⎞ 2
V=
2
K FP ( x + Rθ B ) + KT R2 θ B − θ W
2
( ) + ⎜ T
2⎝ 2
⎟x

(21)

where KT and KR are the total stiffness of the sidewalls. By applying the Hamilton
Principle, we can obtain the equation of motion as
⎡ ⎤
⎡ IW 0 0 ⎤ ⎧θ˙˙W ⎫ ⎢ KT R2 − KT R 2 0 ⎥ ⎧θ W ⎫
⎢ ⎥⎪ ⎪ ⎥ ⎪θ ⎪ = 0
⎢0 IB 0 ⎥ ⎨θ˙˙B ⎬ + ⎢− KT R2 ( K FP + KT ) R2 K FP R
⎢ ⎥⎨ B ⎬
⎢⎣ 0 0 mB ⎥⎦ ⎪⎩ x˙˙ ⎪⎭ ⎢ 0 K FP R K FP +
K R + KT ⎥ ⎪ x ⎪
⎩ ⎭
⎢⎣ 2 ⎥⎦
(22)

From the above equation of motion, the eigenfrequencies and eigenvectors


can be found. While closed form solutions to the 3 × 3 eigenvalue are possible, a
numerical solution is obtained with the stiffness and mass values for the
P235/75R17 tire. The three mode shapes have frequencies of 28 Hz, 61 Hz and
109 Hz. A sketch of the mode shapes is shown in Fig. 16. The mode shapes show

FIG. 15 — Rigid ring-tire model with 3 DOF and a footprint constraint. The model is limited to belt
motion in the longitudinal direction and belt and wheel rotation. The first two eigenmodes can be
seen in the spectrum of cleat impact data.
210 TIRE SCIENCE & TECHNOLOGY

that the 28Hz frequency signal represents a vibration mode, where the belt rocks
fore/aft while grounded in the footprint. The wheel inertia rotates with the belt
motion. At the second fore/aft mode, the wheel rotates out-of-phase with the belt.
Belt flexibility becomes relevant beyond 60 Hz; therefore, the third mode is
unlikely to exist in this form in real tires.

FIG. 16 — Eigenmodes of rigid ring tire model with footprint constraint. The first eigenmode repre-
sent an in-phase rotation of the wheel and belt against the footprint constraint. In the 2nd mode, the
wheel and belt rotate out-of-phase. The 3rd mode is not likely to be similar to a real tire mode, because
its frequency is beyond 100 Hz, in a frequency range, where flexible belt modes occur.

It is important to note the distinction between the modes discussed earlier


(fixed spindle, unloaded) and these tire modes. The observed frequency is now
not only dependent on the tire mass and stiffness properties but also on the foun-
dation stiffness (footprint shear stiffness) and wheel inertia. Any components
rotating with the wheel have to be considered part of the wheel inertia. The sen-
sitivity of these fore/aft vibration modes with respect to the tire parameters is
now easily determined.

Speed Dependence of Cleat Impact Force Transmission


As was seen in Fig. 13 and Fig. 14, the tire force transmission is a function
of rolling speed. To compare the spectrum across a wider range of velocity, a sur-
face plot was created, which shows the spectrum amplitude versus speed (see
Fig. 17). The comparison between the FTIRE and the experimental data again
shows very good correlation. The spectrum shows the very noticeable 1st vertical
resonance for most speeds. Also there is no shift in frequency with rolling speed.
Since this mode has a wave number of n = 1, this is to be expected according to
DORFI ON IN-PLANE FORCE TRANSMISSION 211

Eq. (8). However, at 45 kph the vertical resonance virtually disappears. This
coincides with a resonance at 76 Hz, which is both noticeable in the longitudinal
and normal direction. Figure 18 shows the force orbit of the spindle force in the
Fx-Fz plane filtered with a band path filter of 60-80Hz. It shows that the 76 Hz
resonance is a counter-clockwise rotating force, which travels against the tire
rolling direction. This is one of the traveling resonances, which can be excited in
the rolling tire. This phenomenon can also be observed at different loads and is
discussed in more detail in Ref. [14].

FIG. 17 — Surface plot of force spectrum of normal and longitudinal force. A dominant resonance
in the vertical direction at 68Hz is apparent for most speeds. At about 45 kph, the vertical resonance
disappears and a 76 Hz signal is noticeable in both the longitudinal and vertical direction. In the lon-
gitudinal direction, the 28 Hz resonance is only noticeable in the lower speed range.
212 TIRE SCIENCE & TECHNOLOGY

FIG. 18 — Spindle force orbit at 45 kph. The spindle force rotates against the direction of tire rota-
tion with a frequency of 76 Hz.

Conclusion

Tire in-plane force transmission and cleat envelopment are important con-
tributors to the ride performance of automobiles. This paper attempts to con-
tribute to this subject by providing additional insight into the vibratory response
of tires during cleat envelopment. It is demonstrated how some simple models
can be used to understand tire vibration phenomena, how tire mass and stiffness
affect the modal frequencies and how modal data can be used to estimate stiff-
ness properties.
This “know-how” has been used in FTIRE, a new numerical time integrated
flexible belt tire model. FTIRE shows excellent correlation to measured data for
both the static and dynamic force response due to cleat impact. FTIRE’s ability
to accurately predict the spindle force, enveloping behavior and modal response
as a function of speed makes it a valuable simulation tool for both the automo-
tive and tire engineer.

Acknowledgements
The author would like to thank the Hankook Tire Co. for permission to pub-
lish this work. Further thanks go to Dr. Marvin Janssen, Manager, Tire
Engineering Technology for several helpful discussions.
DORFI ON IN-PLANE FORCE TRANSMISSION 213

References

[1] Gough, V. E., Tires and Air Suspension, Advances in Automobile Engineering, Tidbury, Ed.,
Pergamon Press, 1963.
[2] Julien, M. M. and Paulsen M. J., “Methode Experimentale de Mesure et Definition du Pouvoir
Absorbant du Pneumatique,” J. de la S.I.A., janvier 1953, p. 33.
[3] Pacejka, H. B., “Tire In-plane Dynamics,” Mechanics of Pneumatic Tires, Clark, S. K. Ed., pp.
695, Nat. Bur. Stand. Monogr. 122.
[4] Takayama, M. and Yamagishi, K., “Simulation Model of Tire Vibration,” Tire Science and
Technology, Vol. 11, 1984, pp. 38-49.
[5] Zhang Y., Palmer T. and Farahani, A., “A Finite Element Tire Model and Vibration Analysis: A
New Approach,” Tire Science and Technology, Vol. 26, 1998, pp. 149-172.
[6] Scavuzzo R. W., Richards T. R. and Charek, L. T., “Tire Vibration Modes and Effects on Vehicle
Ride Quality,” Tire Science and Technology, Vol. 11, 1984, pp. 38-49.
[7] Richards T. R., Charek, L. T. and Scavuzzo R. W., “The Effects of Spindle and Patch Boundary
Conditions on Tire Vibration Modes,” SAE Paper 860243, SAE Congress, Feb. 24-28, 1986.
[8] Boehm F., “Mechanik des Gürtelreifens,” Ingenieur Archiv. 35, 1966, pp. 82-101.
[9] Vinesse, E. P., “Substructuring Method and Eigenfunction Representation for a Rolling Tyre
Coupled to a Secondary System,” Proc Instn Mech Engrs, Vol. 210, pp. 313-325.
[10] Soedel, W. “Vibrations of Shells and Plates,” 2nd Edition, 1993.
[11] Kung, L. E., Soedel, W. and Yang, T. Y. “Free Vibration of a Pneumatic Tire-Wheel Unit Using
a Ring on an Elastic Foundation and a Finite Element Model, Journal of Sound and Vibration,
Vol. 107, no. 2, 1986, pp 181-194.
[12] Kim, Y.K. “Forced Response of Tires with Mass Nonuniformities Using Ring Models,”
Dissertation, Purdue University, 2000.
[13] Gipser, M., “FTIRE, a New Fast Tire Model for Ride Comfort Simulations,” International
ADAMS User Conference, Berlin, 1999.
[14] Dorfi, H., “Tire Cleat Impact and Force Transmission: Modeling Based on FTIRE and
Correlation to Experimental Data,” Paper to be presented at the SAE 2004 Congress and
Exhibition, March 2004.

You might also like