Download as pdf or txt
Download as pdf or txt
You are on page 1of 33

Engineering rubber bushing stiffness

formulas including dynamic amplitude


dependence

Maria-José Garcia

Stockholm 2006

Licentiate Thesis

Royal Institute of Technology

School of Engineering Sciences

Department of Aeronautical and Vehicle Engineering

The Marcus Wallenberg Laboratory for Sound and Vibration Research


Akademisk avhandling som med tillstånd av Kungliga Tekniska Högskolan i Stock-
holm framläggs till offentlig granskning för avläggande av teknologie licentiatexa-
men torsdagen den 15:e juni 2006, kl 10.30 i sal MWL 74, Teknikringen 8, KTH,
Stockholm.

TRITA-AVE -2006:36
ISSN -1651-7660

c
°Maria-José Garcia, June 2006
Abstract Engineering design models for the torsion and axial dynamic stiff-
ness of carbon black filled rubber bushings in the frequency domain including
amplitude dependence are presented. They are founded on a developed ma-
terial model which is the result of applying a separable elastic, viscoelastic
and friction rubber component model to the material level. Moreover, the
rubber model is applied to equivalent strains of the strain states inside the
torsion or axial deformed bushing previously obtained by the classical linear
theory of elasticity, thus yielding equivalent shear moduli which are inserted
into analytical formulas for the stiffness. Therefore, unlike other simplified
approaches, this procedure includes the Fletcher-Gent effect inside the bush-
ing due to non-homogeneous strain states. The models are implemented in
Matlabr . In addition, an experimental verification is carried out on a
commercially available bushing thus confirming the accuracy of these mod-
els which become a fast engineering tool to design the most suitable rubber
bushing to fulfil user requirements. Finally, they can be easily employed in
multi-body and finite element simulations.
Licentiate Thesis
This licentiate thesis consists of this summary and two appended papers
listed below and referred to as Paper A and Paper B.

Paper A
M.J. Garcı́a Tárrago, L. Kari, J. Viñolas and N. Gil-Negrete, 2006:
“Torsion stiffness of a rubber bushing: a simple engineering design formula
including amplitude dependence”. Submitted to Journal of Strain Analysis
for Engineering Design.

Paper B
M.J. Garcı́a Tárrago, L. Kari and J. Viñolas, 2006: “Axial stiffness
of carbon black filled rubber bushings including frequency and amplitude
dependence”. Submitted to Kautschuk Gummi Kunststoffe.
Contents
1 Introduction 1

2 Objectives 3

3 Methodology 4
3.1 Theoretical formulation . . . . . . . . . . . . . . . . . . 4
3.1.1 Rubber material model . . . . . . . . . . . . . . 5
3.1.2 Equivalent strain . . . . . . . . . . . . . . . . . 8
3.1.3 Engineering formulas for the stiffness . . . . . . 10
3.2 Experiments . . . . . . . . . . . . . . . . . . . . . . . . 11
3.2.1 Test object . . . . . . . . . . . . . . . . . . . . 12
3.2.2 Setup . . . . . . . . . . . . . . . . . . . . . . . 13
3.2.3 Type of measurements . . . . . . . . . . . . . . 13

4 Results 14
4.1 Implementation in Matlabr . . . . . . . . . . . . . . 14
4.2 Experimental verification . . . . . . . . . . . . . . . . . 18

5 Conclusions 23

Acknowledgments 24

References

Paper A

Paper B
1 Introduction
Rubber isolators play an important role in the noise and vibration con-
trol. They isolate the car body from the vibration source: wheel and tire
unbalance, road input, powertrain, etc. There is a wide variety of designs de-
pending on the geometry of the component and the disposition of the metal
layers, chosen to fulfil user requirements. In particular, rubber bushings are
the object of this study. They consist of a rubber tube bonded on their
outer and inner surfaces to rigid metal layers. Their high axial and torsion
flexibility, large radial and conical rigidity and their inherent damping make
them interesting for being used in primary suspensions, pivot arms and sev-
eral types of mechanical linkage. Therefore, due to the increasing interest
in predictions performed by multi-body simulations of complete vehicles or
subsystems, it is important to develop simple and effective models to repre-
sent the dynamic stiffness of these rubber bushings. Simplicity is necessary
as the rubber bushing represents only a small part connecting components
of more complex structures.
The main characteristic of rubber materials is their elasticity and capac-
ity for dissipating vibration energy which is represented by the hysteresis
loop enclosed by the loading and unloading curves in a stress–strain dia-
gram. Furthermore, rubber materials are strongly dependent on frequency
and amplitude, as reviewed by Medalia (1). The latter dependence, known
as Fletcher–Gent effect (2), involves a non-linearity in the dynamic stress-
strain behavior and it becomes more pronounced with the presence of rubber
fillers, generally one of the many kinds of carbon black, as shown by Payne
and Whittaker (3). But also this presence produces an increase in the shear
modulus and damping. For that reason, it is important to include the am-
plitude dependence in the representation of the dynamic behavior of filled
rubber bushings.
Firstly, the behavior of these components has been defined by their static
stiffness as calculated by finite element analysis (4), by using truncated

1
Fourier and Bessel functions (5) or through principal mode approaches (6).
Lately, Horton et al. have developed formulas for radial, conical and torsion
static stiffness (7; 8; 9) based on three-dimensional static elastic theory.
Moreover, the static stiffness in many directions is provided in standard
textbooks (10; 11; 12; 13; 14). In addition, the structure-borne properties
of a long rubber bush mounting in all directions are considered in Kari
(15). However, none of these methods takes into account the amplitude
dependence.
To date, several models have been devised to represent this property, like
Kraus (16) who explains the amplitude dependence as due to the continuous
breaking and reforming of van der Waals bonds between carbon-black ag-
gregates, while modifications and experimental application are carried out
by Ulmer (17) and Lion (18) presenting a time domain formulation of the
Kraus model. Moreover, the Fletcher–Gent effect is also characterized us-
ing friction models, like Gregory (19), Coveney et al. (20) and Kaliske and
Rothert (21) proposing the Prandtl element, a Coulomb damper in series
with a elastic spring; a model expanded by Bruni and Collina (22), Olsson
and Austrell (23) and Brackbill et al. (24).
Furthermore, several authors have employed these material models in
order to calculate the dynamic stiffness of rubber bushings. Some of them
insert the models into finite element systems, like Austrell et al. (25; 26; 27)
who represent frequency and amplitude dependencies adding integer deriva-
tives to stick-slip friction components, similar to Gil–Negrete (28) except
for the use of fractional derivatives. However, finite element procedures re-
quire a long-time overlay process to calculate the stiffness. In contrast, other
authors work directly at the component level, like Berg (29) who presents
a five-parameters model which gives a good resemblance to the smoother-
friction rubber behavior, also used by Sjöberg and Kari (30) together with a
rate-dependent part using fractional derivatives, and Misaji et al. (31) who
take into account amplitude dependence as the parameters of an ordinary
Kelvin–Voigt model are updated continuously for every oscillation cycle.
However, the latter methods neglect the amplitude dependence inside the
rubber bushing due to non-homogeneous strain states.

2
This work overcomes previous limitations by the development of sim-
ple engineering formulas to represent the dynamic stiffness of rubber bush-
ings in the frequency domain including amplitude dependence. They are
achieved by applying a separable elastic, viscoelastic and friction material
model to equivalent strains of the non-homogeneous strain states inside a
rubber bushing, when it is subjected to torsion, and axial deformations, as
shown in Paper A and B respectively, thus giving equivalent shear modulus
which is inserted into analytical formulas for the stiffness. The novel material
model is presented in Paper A and results from applying a sound compo-
nent level to the material level. Consequently, unlike other simplified models
these formulas include the amplitude dependence due to non-homogeneous
strain states and therefore they become a fast engineering design tool to
determine the most suitable rubber bushing. Moreover, they can be eas-
ily employed in multi-body and finite element simulations. Experimental
verification has been carried out in Paper B for the axial stiffness of a com-
mercial bushing, showing a good agreement between measured and model
stiffness over a frequency range at several amplitudes. Furthermore, for the
torsion case it is shown in Paper A—by dividing the bushing into several
slices and consequently each equivalent shear modulus is closer to the true
value—that the approach of working with only one equivalent modulus for
the whole bushing is accurate enough.

2 Objectives
The main purpose of the presented research is to contribute to the work done
in the field of modelling rubber isolators by developing simple engineering
design models for the dynamic stiffness of rubber bushings including am-
plitude and frequency dependence when they are subjected to harmonic
displacement amplitudes in the 1-100 Hz frequency range.
This objective involves the following tasks:

• Create a rubber material model by extending a sound component


model to the material level.

3
• Apply the classical theory of elasticity in order to obtain the strain
states inside a bushing subjected to torsion and axial deformations,
to subsequently calculate an equivalent strain for each state. Fur-
thermore, the rubber model is applied to these strains, thus giving
equivalent shear moduli that are inserted into engineering formulas
for the stiffness.

• Models implementation in Matlabr .

• Experimental verification on commercially available bushings.

3 Methodology
The methodology to achieve the objects presented in previous section is
subsequently described. Two main procedures are carried out: firstly, a the-
oretical formulation of the model and secondly, an experimental verification.

3.1 Theoretical formulation


Static engineering formulas are widely used to predict the behavior of rub-
ber bushings. Nevertheless they neglect the dynamic properties of rubber
leading to errors in the stiffness prediction. The solution presented in this
work is based on the idea that the dynamic properties can be introduced
into engineering stiffness formulas by means of a dynamic equivalent mod-
ulus which is the result of applying a rubber model to a harmonic strain.
However, this modulus depends on the strain amplitude, due to the non-
linear relation between stress and strain for rubber. Therefore, in order to
represent the behavior of the whole bushing, a global value for the shear
modulus is required.
The model is developed as follows: Firstly, a rubber material model is
created by extending a sound component model to the material level. Sec-
ondly, the classical theory of elasticity is applied to obtain the strain states
inside the bushing when it is subjected to torsion and axial deformation,
while considering that the relation between stress and strain contains only
elastic and fractional derivative components. Subsequently, the calculation

4
of energy density balances between the bushing subjected to each defor-
mation and a simple shear specimen made of the same material gives the
equivalent strains. Finally, the rubber model is applied to those strains,
thus giving the equivalent complex shear moduli which are inserted into the
formula for the torsion and axial stiffness calculated previously with the
classical theory of elasticity.

3.1.1 Rubber material model

The simplest representation of rubber materials is the Kelvin–Voight model


where an elastic frequency independent spring is coupled in parallel with a
viscous dashpot, as discussed by Knothe and Grassie (32). However, this
model strongly overestimates both stiffness and damping for higher frequen-
cies. Furthermore, the three-parameter Maxwell model, which is the result
of including a spring in series with the dashpot, represents better the stiff-
ness at high frequencies. However, a better representation of the frequency
dependence can be obtained by using a summation of Maxwell models, as
discussed by Lodhia and Esat (33) and Betz (34). Nevertheless, this rep-
resentation increases the complexity and number of parameters needed to
accurately describe material properties. An alternative to obtain a good de-
scription of the frequency dependence while reducing the required number
of parameters consists of using time derivatives of non-integer order, known
as fractional derivatives, a method used by Sjöberg and Kari (30). However,
the fractional derivative models neglect the amplitude dependence, an effect
which appears when including fillers, such as carbon black, in a rubber mix-
ture, thus introducing additional bonds within the material. These bonds
exhibit a friction-like behavior as discussed in Ref. (35) and modelling it
demands for friction models. Austrell et al. (25) employ stick-slip com-
ponents, an approach also used by Bruni and Collina (22). Furthermore,
Berg (29) presents a model giving better resemblance to the smoother rub-
ber friction behavior, a representation also used by Sjöberg and Kari (30).
The material model employed in this work is developed in Paper A by
applying to the material level the component model presented by Sjöberg
and Kari (30), where the amplitude dependence is represented by a smooth

5
friction component while the frequency dependence is modelled by a frac-
tional Kelvin–Voigt model, thus giving an effective model while using only
five parameters.

s to ta l( t) e (t)

m ,a
2 m s f m a x ,e 1 / 2

Figure 1: Rubber material model.

Elastic stress The first branch in Fig. 1 represents the elasticity by a


linear relation between the elastic stress σelast and strain ε

σelast (t) = 2µε(t), (1)

where µ is the elastic shear modulus and t is time.

Frequency dependent stress The second component in Fig. 1 indicates


the frequency dependence and it is modeled by using fractional derivatives
which increases the ability to adjust to measured characteristics while keep-
ing the number of parameters to only two: a proportionality constant m
and the time derivative order α

σfract (t) = mDα ε(t), 0 < α ≤ 1, (2)

where σfract is the viscoelastic stress and Dα denotes the fractional time
derivative of order α, defined through an analytical continuation of a frac-
tional Riemann–Liouville integration (36). Numerically, the viscoelastic

6
component can be evaluated as
n−1
∆t−α X Γ(j − α)
σfract (t) ≈ m εn−j , (3)
Γ(−α) Γ(j + 1)
j=0

where εn−j = ε((n − j)∆t), εn is the final strain at time tn = n∆t, ∆t is


a constant time step applied in the estimation process and Γ denotes the
Gamma function (37) defined as
Z ∞ ∞
X
β−1 −x (−1)j
Γ(β) = x e dx + , β 6= 0, −1, −2, ... (4)
1 j!(β + j)
j=0

In addition, the temporal Fourier transformation of the fractional time


derivative branch leads to

σ̃fract (ω) = m(iω)α ε̃(ω), (5)

which is an expression numerically more easy to deal with while omitting


the Γ function, where (˜·) represents the temporal Fourier transform, ω the
angular frequency and i the imaginary unit.

Frictional stress The third branch in the rubber model represents the
amplitude dependence by a smooth friction component which enables a very
good fit to measured curves using only two parameters σf max and ε1/2 . The
frictional stress σfrict develops gradually following the equation
h ih i
ε(t) − εs σf max − sign(ε̇) σf s
σfrict (t) = σf s + h i h i, (6)
σf s
ε1/2 1 − sign(ε̇) σf max + sign(ε̇) ε(t) − εs

where the material parameters σf max and ε1/2 are the maximum friction
stress developed and the strain needed to develop half of that stress, with
sign(ε̇) denoting the sign of the strain rate. The values of σf s and εs are
updated each time the strain changes direction at ε̇ = 0 as σf s ← σfrict |ε̇=0
and εs ← ε |ε̇=0 .

7
Total stress Finally, the rubber behavior is represented by a total stress
σtotal which is the sum of the three stresses

σtotal (t) = σelast (t) + σfract (t) + σfrict (t). (7)

This Equation, with Equations (1), (2) and (6), represents a nonlinear rela-
tion between the total stress and the strain and a value of the shear modulus
cannot be obtained directly as would be the case with the elastic component.

3.1.2 Equivalent strain

The strain state inside the rubber bushing of length L in Fig. 2 bonded to
two cylindrical metal layers at inner and outer radii a and b and subjected to
a harmonic torsion deformation is studied in Paper A while the axial case
is analysed in Paper B. The classical linear theory of elasticity provides

Ø 2 a

u j
r
z
u
r
q
L r
z u
r
r

q y

x
Ø 2 b

Figure 2: Rubber cylindrical bushing subjected to torsion deformation.

the strain states inside the rubber bushing while considering firstly, that the

8
constitutive equations between the stress and strain have only elastic and
frequency dependence components, as the amplitude dependence is taken
into account later on when the rubber model is applied to the equivalent
strains. Furthermore, the strain states with torsion and axial deformations
have only one shear component and therefore the constitutive equation in
both cases reads

σt/a (t) = 2µεt/a (t) + mDα εt/a (t), (8)

where σt/a is the torsion/axial shear stress and εt/a is the torsion/axial shear
strain. In addition, working in the frequency domain as result of applying the
temporal Fourier transform makes calculations easier because the derivation
disappears and the total stress is proportional to the strain
h i
σ̃t/a (ω) = 2µ + m(iω)α ε̃t/a (ω) = 2µ̂(ω)ε̃t/a (ω). (9)

Secondly, the rubber is assumed homogeneous and isotropic and displace-


ment gradients are considered small enough to apply the theory of elastic-
ity. Furthermore, over the frequency range of interest wave effects within
the rubber bushing are negligible. Consequently, when the bushing is sub-
jected to a harmonic torsion angle at the outer surface relative to the inner
ϕ(t) = ϕ sin(ωs t) with ωs the excitation frequency and ϕ the amplitude, the
strain state reads
b2 a2 1
ε̃t (r, ω) = ϕ̃(ω), (10)
b2 − a2 r2
where r is the radius. Whereas if a harmonic axial displacement d(t) =
d sin(ωs t) is applied at the outer surface the strain state becomes
1 ˜
ε̃a (r, ω) = d(ω). (11)
2r loge ( ab )

Subsequently, in order to calculate the equivalent strains one idea would be


to perform volume averages of the strain states. However, in the torsion and
axial deformation modes the strain states have only one shear component,
but in other deformation modes it might be quite complex as it happens in
the radial case where there are both shear and traction/compression strains

9
inside the bushing, consequently performing a volume average of the strain
state is not a general methodology that could be extended to other defor-
mation modes. Therefore, the solution here presented consists in working
with equivalent strains obtained from energy density balances between the
torsion and axial deformation in a bushing and a simple shear specimen. For
instance, as shown in Paper A the deformation energy in the torsion case
is equal to the distortion energy because the volume remains constant. Fur-
thermore, the temporal Fourier transform of the volume averaged distortion
energy reads
Z Z
1 1
Ũd bushing ∝ σ̃t∗ ε̃t dV = 2µ̂|ε̃t |2 dV, (12)
V V

where V = π(b2 − a2 )L is the volume of the bushing and µ̂ is the frequency


dependent shear modulus presented in Equation (9). In addition, the tem-
poral Fourier transform of the distortion energy density in case of a simple
shear specimen made of the same material as the rubber bushing is

Ũd specimen ∝ 2µ̂ε̃∗sh ε̃sh = 2µ̂|ε̃sh |2 , (13)

where ε̃sh is the temporal Fourier transform of the homogeneous shear strain
inside the specimen. Consequently, the balance between both energy densi-
ties gives the equivalent strain in the frequency domain which is transformed
into the time domain by applying the inverse Fourier transform
ba
εequiv t (t) = ϕ(t). (14)
b2 − a2
Moreover, if the same methodology is conducted for the axial case the equiv-
alent strain becomes
d(t)
εequiv a (t) = q ϕ(t). (15)
2(b2 − a2 ) loge ( ab )

3.1.3 Engineering formulas for the stiffness

In order to obtain the equivalent shear modulus, the rubber model in Fig. 1
is applied to the equivalent strains thus giving the total stresses σtotal t/a (t)

10
made up of three components: elastic, frequency dependent and frictional
stresses. Furthermore, the last component introduces the amplitude depen-
dence and makes the rubber model nonlinear.
In order to switch to the frequency domain the excitation frequency
of the input signals ϕ(t) and d(t) is varied as ωs = s∆ω with ∆ω a con-
stant frequency step and s ∈ N, and therefore s different equivalent strains
εsequiv t/a (t) and their corresponding total stresses σtotal
s
t/a (t) are obtained.
Furthermore, the temporal Fourier transform is applied to those signals,
thus giving the frequency functions ε̃sequiv t/a (ω) and σ̃total
s
t/a (ω) which are
evaluated at ωs leading to the equivalent complex moduli in the frequency
domain
s
σ̃total t/a (ωs )
µ̂equiv t/a (ωs ) = s . (16)
2 ε̃equiv t/a (ωs )

This linearization process takes into account the non-linear relation between
stress and strain at frequency ωs considering the first order response while
omitting the less important overtones (stress response at 3ωs , 5ωs ,...). Fi-
nally, the moduli are inserted into the stiffness formulas calculated previously
with the classical linear theory of elasticity, thus giving a formula for the
torsion dynamic stiffness
4πa2 b2 L
Kdyn torsion (ωs ) = µ̂equiv t (ωs ), (17)
b2 − a2
while the axial stiffness reads
2πL
Kdyn axial (ωs ) = µ̂equiv a (ωs ). (18)
loge ( ab )
Consequently, two engineering formulas to represent in the frequency domain
the dynamic behavior of rubber bushings in the torsion and axial directions
including the amplitude dependence have been achieved.

3.2 Experiments
The accuracy of the developed models representing the dynamic stiffness
of rubber components while including amplitude dependence is experimen-
tal verified. Specifically, the model for the axial stiffness is validated by a

11
commercially available bushing in Paper B. Firstly, the material parame-
ters involved in the formula of the stiffness, see Equation (18), are achieved
from two measurements: elastic and amplitude dependence parameters are
obtained from a large deformation quasi-static test while the viscoelastic
parameters are determined by a steady state harmonic dynamic test per-
formed over a frequency range from 1 to 100 Hz at an amplitude of 0.1 mm.
Secondly, five new dynamic tests at 0.013, 0.04, 0.17, 0.25 and 0.35 mm are
conducted in order to demonstrate that the formula for the axial stiffness
predicts well the bushing behavior including the amplitude and frequency
dependence.

3.2.1 Test object

The test object is a commercial bushing manufactured by Trelleborg AVS


under the trade name of VP 40/70120 with a rubber hardness of 60 IRH.
The rubber type is a mixture between SMR 10 and SMR GP and the carbon
black filler is a combination of N660 and N774.
L = 9 0 m m

b = 3 5 m m
a = 2 0 m m

Figure 3: Test object.

12
3.2.2 Setup

The experiments are performed in a servo-hydraulic test machine consisting


of a lower body with a hydraulic piston inside, two columns to the sides
and an upper rigid crosshead, as shown in Fig. 4. Measurements are per-
formed with the inner surface of the bushing fixed and joined to the upper
crosshead through a shaft and one load sensor, while the outer surface of the
bushing is subjected to axial displacements produced by the lower hydraulic
piston. The motion of the bushing is measured by two displacement sensors
symmetrically positioned on the outer surface of the bushing. The electrical
signals are conditioned in a 6-channels amplifier producing electrical inputs
to the frequency analyzer which are processed subsequently in a personal
computer. In order to achieve constant amplitude, the electric signal that
comes from the displacement sensor of the test machine and goes through
the amplifier is automatically adjusted in a control loop in the analyzer to
supply the exciter.

3.2.3 Type of measurements

The test component is mechanically conditioned before measurements to


eliminate influence of the Mullins’ effect (38) which is shown when previ-
ously unstrained rubber is subjected to strain cycles at constant peak value
reducing peak stress values at the first few oscillations.

Quasi-static test Measurements at 0.1 Hz and 1 mm displacement ampli-


tude are used to fit the elastic and amplitude dependence parameters. The
large deformation assures that all friction has been developed in the loop
and the low frequency guarantees that all viscous effects are eliminated.

Dynamic test The bushing is excited by stepped sine displacements start-


ing at 1 Hz and increasing with a constant frequency step of 1 Hz to a
maximum frequency of 100 Hz, with amplitude held constant at 0.1 mm.
Viscoelastic parameters are the results of an estimation iterative minimiza-
tion process between the measured and the model complex stiffness.

13
F IX E D S E R V O -H Y D R A U L IC
T E S T M A C H IN E

6 -C H A N N E L S A M P L IF IE R

L O A D S E N S O R

D IS P L A C E M E N T
S E N S O R

R U B B E R B U S H IN G

IN P U T O U T P U T

F R E Q U E N C Y A N A L Y Z E R

P IS T O N

P E R S O N A L C O M P U T E R

D E S K C O N T R O L L E R

Figure 4: Measurement setup.

4 Results
The methodology to obtain the torsion and axial dynamic stiffness is imple-
mented in Matlabr , thus giving a fast engineering design tool to predict
the dynamic behavior of rubber components. Furthermore, the calculation
of the torsion stiffness for a typical commercial bushing is carried out in
Paper A, thus showing the expected behavior for a rubber bushing, while
Paper B presents an experimental verification for the axial stiffness per-
formed on a commercially available bushing.

4.1 Implementation in Matlabr


The implementation in Matlabr is carried out for each deformation mode
as follows: firstly, the equivalent strains in the time domain are calculated

14
as function of the geometry of the bushing and the boundary conditions, as
shown in Equations (14) and (15). Secondly, the rubber model is applied to
those strains and the steady state of the resultant stresses are converted into
frequency functions through the temporal Fourier transform. Furthermore,
the value of the frequency dependent shear moduli result from dividing the
stress frequency functions by the strain frequency signals when both func-
tions are evaluated at the excitation frequency, see Equation (16). In addi-
tion, the amplitude of the displacement excitations is varied, as well as, the
frequency range that goes from 1 to 100 Hz, to check how well the model
represents both dependencies. Finally, the frequency and amplitude depen-
dent shear moduli are inserted into the formulas of the stiffness, Equations
(17) and (18).

Table 1: Typical geometric and material data for a commercial rubber bush-
ing.

Geometric data Material data


L = 22.0 × 10−3m µ = 2.0 × 106 N/m2
−3
a = 5.0 × 10 m m = 2.0 × 105 Nsα /m2
b = 11.0 × 10−3 m α = 0.40
σf max = 4.0 × 103 N/m2
ε1/2 = 1.0 × 10−3

Paper A presents the torsion stiffness for a typical commercial avail-


able bushing whose data is given in Table 1. The torsion angle is varied
harmonically with amplitudes going from 1.74 × 10−6 to 1.74 × 10−1 rad
over a frequency range from 10 to 100 Hz. Fig. 5 displays the variation in
magnitude and loss factor of the torsion stiffness versus frequency at three
angle amplitudes while Fig. 6 shows the variation in magnitude and loss
factor of the stiffness versus angle amplitude at three frequencies. Both fig-
ures demonstrate that the stiffness calculated with the presented formula
represents the expected behavior for a rubber bushing: the magnitude and
the loss factor increase with frequency, while versus the angle amplitude the

15
magnitude of the stiffness decreases and the loss factor initially increases
and then decreases.

50
Magnitude (Nm/rad)

40

30

20 6.3 x 10−6 rad


2.9 x 10−4 rad
10
1.3 x 10−2 rad
0
Loss Factor

0.2

6.3 x 10−6 rad


0.1
2.9 x 10−4 rad
1.3 x 10−2 rad
0
10 20 30 40 50 60 70 80 90 100
Frequency (Hz)

Figure 5: Torsion stiffness magnitude and loss factor versus frequency at


three amplitudes.

Consequently, the displayed frequency and amplitude stiffness depen-


dency highlights the need of models which take into account both effects.
Moreover, the use of static models to represent the behavior of rubber com-
ponents, a procedure commonly used in engineering design, may induce
errors. Therefore, the developed engineering formulas become a fast and
effective tool to predict the behavior of rubber bushings including frequency
and amplitude dependence.
In addition, in Paper A the sample bushing is divided into one, two

16
Magnitude (Nm/rad) 50

40

30

20
20Hz
10 50Hz
80Hz
0
Loss Factor

0.2

0.1 20Hz
50Hz
80Hz
0 −4 −2
10 10
Amplitude (radians)

Figure 6: Torsion stiffness magnitude and loss factor versus amplitude at


three frequencies.

and three slices to show that even if there are several slices and therefore
the equivalent shear moduli are closer to the true value than dealing with
only one equivalent modulus for the whole bushing, there are only small
differences in the magnitude and loss factor of the dynamic stiffness between
all cases, and therefore working with only one equivalent shear modulus for
the whole bushing is accurate enough.
The dynamic stiffness is the result of an iterative process which starts
with an estimation of the torsion angles at the inner and outer radius of each
slice to, subsequently calculate the complex moment at each radius following
the methodology explained previously. Furthermore, in order to fulfil the

17
continuity of the moment along the radius, the values of the torsion angles
are modified until the moment differences at each radius are negligible.

4.2 Experimental verification


The model for the axial stiffness is experimental validated for a commercially
available bushing in Paper B. Calculations and graphical representations
are carried out in Matlabr . Elastic and amplitude dependence parameters
are varied until the model hysteresis loop enclosed by the resultant force
in the bushing when is subjected to axial displacement at 0.1 Hz and 1
mm amplitude fits well with the hysteresis loop obtained from one quasi-
static test performed on the bushing at the same displacement conditions.
The good agreement between the two loops displayed in Fig. 7, is achieved
with the values: µ = 2.42 × 106 N/m2 , σf max = 1.92 × 104 N/m2 and
ε1/2 = 4.0 × 10−3 .
Moreover, viscoelastic parameters are obtained from the error minimiza-
tion iteration process between the complex stiffness obtained with one dy-
namic test performed from 1 to 100 Hz at 0.1 mm constant amplitude and
the stiffness calculated by Equation (18) at the same conditions. In par-
ticular, when the viscoelastic parameters are assigned the values: m =
6.0 × 105 Nsα /m2 and α = 0.20, the measured stiffness fits well in mag-
nitude and loss factor with the model stiffness, as shown in Fig. 8.

In addition, five new dynamic tests are performed at amplitudes of 0.013,


0.04, 0.17, 0.25 and 0.35 mm and the results are compared to those of
the model when using the five material parameters calculated previously,
displaying a good agreement over the whole frequency range, as presented
in Figs. 9 and 10, which verifies the accuracy of Equation (18) representing
the axial dynamic stiffness of a rubber bushing including amplitude and
frequency dependence.

18
3000
Model
Measurement
2000

1000
Force (N)

−1000

−2000

−3000
−1.5 −1 −0.5 0 0.5 1
Displacement (m) −3
x 10

Figure 7: Comparison between measured and model hysteresis loop at 1 mm


amplitude and 0.1 Hz.

19
6
x 10
5
Magnitude (N/m)

4
3
2
1 Model
Measurement
0

0.15
Loss Factor

0.1

0.05
Model
Measurement
0
0 20 40 60 80 100
Frequency (Hz)

Figure 8: Comparison between measured and model dynamic stiffness in


magnitude and loss factor over a frequency range from 1 to 100 Hz when
amplitude is held constant at 0.1 mm

20
6
Magnitude (kN/mm)

Model 0.04mm
4 Meas. 0.04mm
Model 0.1mm
Meas. 0.1mm
2 Model 0.35mm
Meas. 0.35mm

0.15
Model 0.04mm
Loss Factor

Meas. 0.04mm
0.1 Model 0.1mm
Meas. 0.1mm
Model 0.35mm
0.05 Meas. 0.35mm

0
0 20 40 60 80 100
Frequency (Hz)

Figure 9: Measured and model dynamic stiffness in magnitude and loss


factor versus frequency at three amplitudes.

21
6

5.5
Magnitude (kN/mm)

Model 10Hz
5 Meas. 10Hz
Model 20Hz
4.5
Meas. 20Hz
4 Model 40Hz
Meas. 40Hz
3.5

0.18

0.16 Model 10Hz


Loss Factor

Meas. 10Hz
0.14
Model 20Hz
Meas. 20Hz
0.12
Model 40Hz
0.1 Meas. 40Hz

0.08
0.05 0.1 0.15 0.2 0.25 0.3 0.35
Amplitude (mm)

Figure 10: Measured and model dynamic stiffness in magnitude and loss
factor versus amplitude at three frequencies.

22
5 Conclusions
Simple engineering models for the torsion and axial dynamic stiffness of filled
rubber bushings including amplitude and frequency dependence have been
presented. They have been developed by applying a rubber material model
to the equivalent strains of the strain states inside the deformed bushing,
thus providing equivalent shear moduli which are inserted into the analytical
formulas for the stiffness.
Paper A presents the model for the torsion stiffness and illustrates the
results in magnitude and loss factor at several amplitudes over a frequency
range from 1 to 100 Hz when the model is applied to typical geometric
and material data of rubber components, showing the expected behavior
of a rubber bushing. Furthermore, in order to check the accuracy of the
approach the same bushing has been divided into several slices thus using
an equivalent shear modulus closer to the true value for each slice. The
small differences in magnitude and loss factor between the dynamic stiffness
obtained when the bushing is divided into a single, two and three slices, show
that the amplitude dependence is enough well modelled while using a single
slice. In addition, an experimental verification of the axial stiffness model
is carried out on a commercially available bushing in Paper B. Firstly, the
model parameters are obtained from two tests: a quasi-static and dynamic.
Subsequently, five new dynamic experiments are performed on the bushing
to compare the results with those obtained with the engineering formula
at the same amplitude and frequency conditions, showing good agreements
which verifie the accuracy of the model.
The simplicity of this model is very convenient for the dynamic analysis
of complex structures in which rubber bushings are connecting components.
Several advantages can be outlined. Firstly, a novel model has been pre-
sented to represent the dynamic behavior of the rubber at the material level
including frequency and amplitude dependence with only five parameters.
Secondly, the time invested in performing the calculations: Matlabr has
been used to carry out the process without requiring large amount of mem-
ory due to discretization process, just by applying the novel rubber model

23
to one equivalent strain for the whole bushing. Additionally, unlike other
simplified models these formulas depend on the geometry of the bushing
and the material properties and therefore they are useful and fast predic-
tion tools to determine the most suitable rubber bushing to connect to other
structures to fulfil user requirements.

Acknowledgments
The European Community is gratefully acknowledged for the financial sup-
port through contract No: MEST-CT-2004-503675, for a research train-
ing project European Doctorate in Sound and Vibration Studies (EDSVS)
within the framework of the European Community’s Scheme Improving Hu-
man Research Potential. I would like to express my special thanks to my
supervisor Leif Kari for excellent and always encouraged guidance through
the course of this project. My thanks also to Kent Lindgren and Danilo
Prelevic for professional assistance in my experiments and to my colleagues
at MWL. Finally I thank my family and friends for being always there.

References
[1] Medalia, A. I. Effects of carbon black on dynamic properties of rub-
ber. Rubber Chem. Tech., 1978, 51, 437–523.

[2] Fletcher, W. P. and Gent, A. N. Non-Linearity in the dynamic


properties of vulcanised rubber compounds. I.R.I. Transactions , 1953,
29, 266–280.

[3] Payne, A. R. and Whittaker, R. E. Low strain dynamic properties


of filled rubbers. Rubber Chem. Tech., 1971, 44, 440–478.

[4] Morman, K. N. and Pan, T. Y. Application of finite-element analy-


sis in the design of automotive elastomeric components. Rubber Chem.
Tech., 1988, 61, 503–533.

24
[5] Hill, J. M. Radical deflections of rubber bush mountings of finite
lengths. Int. J. Eng. Sci., 1975, 13, 407–422.

[6] Adkins, J. E. and Gent, A. N. Load-deflexion relations of rubber


bush mountings. Brit. J. Appl. Phys., 1954, 5, 354–358.

[7] Horton, J. M., Gover, M. J. C. and Tupholme, G. E. Stiffness


of rubber bush mountings subjected to radial loading. Rubber Chem.
Tech., 2000, 73, 253–264.

[8] Horton, J. M., Gover, M. J. C. and Tupholme, G. E. Stiffness


of rubber bush mountings subjected to tilting deflection. Rubber Chem.
Tech., 2000, 73, 619–633.

[9] Horton, J. M., Gover, M. J. C. and Tupholme, G. E. Stiffness


of spherical rubber bush mountings. Int. J. Solids Struct., 2005, 42,
3289–3297.

[10] Lindley, P. B. Engineering design with natural rubber , 1992 (The


Malaysian Rubber Producers’ Research Association, Brickendonbury).

[11] Gent, A. N. Engineering with rubber , 1992 (Carl Hansen Verlag, Mu-
nich).

[12] Göbel, E. F. Rubber springs design, 1974 (Newnes-Butterworths, Lon-


don).

[13] Payne, A. R. and Scott, J. R. Engineering design with rubber , 1960


(Interscience Publishers, New York).

[14] Freakley, P. K. and Payne, A. R. Theory and practice of engineering


with rubber , 1978 (Applied Science Publishers, London).

[15] Kari, L. Dynamic stiffness matrix of a long rubber bush mounting.


Rubber Chem. Tech., 2002, 75(4), 747–770.

[16] Kraus, G. Mechanical losses in carbon black filled rubbers. J. Appl.


Polym. Sci., 1984, 39, 75–92.

25
[17] Ulmer, J.D. Strain dependence of dynamic mechanical properties of
carbon black-filled rubber compounds. Rubber Chem. Tech., 1998, 69,
637–667.

[18] Lion, A. Phenomenological modelling of strain-induced structural


changes in filler-reinforced elastomers. Kautsch. Gummi Kunstst., 2005,
58(4), 157–162.

[19] Gregory, M.J. Dynamic properties of rubber in automotive engineer-


ing. Elastomerics, 1985, 117(11), 19–24.

[20] Coveney, V. A., Johnsson, D. E. and Turner, D. M. A triboelastic


model for the cyclic behavior of filled vulcanizates. Rubber Chem. Tech.,
1995, 68, 660–670.

[21] Kaliske, M. and Rothert, H. Constitutive approach to rate-


independent properties of filled elastomers. Int. J. Solid. Struct., 1998,
35(17), 2057–2071.

[22] Bruni, S. and Collina, A. Modelling of viscoelastic behaviour of elas-


tomeric components : An application to the simulation of train-track
Interaction. Vehicle Syst. Dyn., 2000, 34, 283–301.

[23] Olsson, A. K. and Austrell, P. E. 2001, A fitting procedure for a


viscoelastic-elastoplastic material model. Second European Conference
on Constitutive Models for Rubber.

[24] Brackbill, C.R., Lesieutre, G.A., Smith, E.C. and Ruhl, L.E.
Characterization and modelling of the low strain amplitude and fre-
quency dependent behavior of elastomeric damper materials. J. Am.
Helicopter Soc. , 2000, 45(1), 34–42.

[25] Austrell, P. E., Olsson, A. K. and Jonsson, M. 2001, A Method


to analyse the non-Linear dynamic behaviour of rubber components
using standard FE codes, Paper no 464. Conference on Fluid and Solid
Mechanics.

26
[26] Olsson, A. K. and Austrell, P. E. 2003, Finite element analysis of
a rubber bushing considering rate and amplitude dependence effects.
Third European Conference on Constitutive Models for Rubber.

[27] Olsson, A. K.,Austrell, P. E. and Sandberg, G. 2005, Approx-


imate viscoelastic FE procedures in frequency and time domain to
account for the Fletcher-Gent effect. Fourth European Conference on
Constitutive Models for Rubber.

[28] Gil-Negrete, N. 2004, On the modelling and dynamic stiffness pre-


diction of rubber isolators. University of Navarra. Spain.

[29] Berg, M. A Non linear rubber spring model for rail vehicle dynamics
analysis. Vehicle Syst. Dyn., 1998, 30, 197–212.

[30] Sjöberg, M. and Kari, L. Non-linear behavior of a rubber isolator


system using fractional derivatives, Vehicle Syst. Dyn., 2002, 37, 217–
236.

[31] Misaji, K., Hirose, S. and Shibata, K. vibration analysis of rubber


vibration isolators of vehicle using the restoring force model of power
function type, Jsme Int. J. C-Dyn. Con., 1995, 38(4), 679–685.

[32] Knothe, K.L. and Grassie, S.L. Modelling of railway track and vehi-
cle/track interaction at high frequencies, Vehicle Syst. Dyn., 1993, 22,
209–262.

[33] Lodhia, B.B. and Esat, I.I. Vibration simulation of systems incorpo-
rating linear viscoelastic mounts using Prony series formulation, ASME,
Engineering system design and analysis conference, 1996, 81(9), 171–
176.

[34] Betz, E. and Spanier, J. Spring and dashpot models and their appli-
cations in the study of the dynamic properties of rubber, Mechanical
and Chemical Engineering Transactions, 1996.

27
[35] Sjöberg, M.Rubber isolators - measurements and modelling using
fractional derivatives and friction, SAE PAPER No. 2000-01-3518 ,
2000.

[36] Oldham, K. B. and Spanier, J. The Fractional Calculus, 1974 (New


York and London: Academic press).

[37] Abramowitz, M. and Stegun., I. A. Handbook of Mathematical


Functionss, 1972 (New York: Dover Publication).

[38] Mullins, L. Softening of rubber by deformation. Rubber Chem. Tech.,


1969, 42, 339–362.

[39] Spencer, A. J. M. Continuum Mechanics, 1980 (New York: Longman


Publishers).

28

You might also like