You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/305826803

DYNAMIC PROPERTIES OF FILLED RUBBER

Conference Paper · October 2008

CITATION READS
1 244

4 authors, including:

Tod Dalrymple Hamid Ahmadi


Dassault Systèmes Tun Abdul Razak Research Centre
8 PUBLICATIONS 13 CITATIONS 19 PUBLICATIONS 94 CITATIONS

SEE PROFILE SEE PROFILE

Alan Muhr
Tun Abdul Razak Research Centre
57 PUBLICATIONS 449 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Integrated calibration tool for hyperelastic+Prony viscoelastic materials. View project

Multi-scale Material Modeling View project

All content following this page was uploaded by Tod Dalrymple on 04 August 2016.

The user has requested enhancement of the downloaded file.


Paper No. 78

DYNAMIC PROPERTIES OF FILLED RUBBER – Part III: USE OF NONLINEAR


KINEMATIC HARDENING PLASTICITY MATERIAL MODELS

Hamid Ahmadi**, Tod Dalrymple *, John Kingston**& Alan Muhr**

* Dassault Systèmes Simulia Corp., Great Lakes Region, USA


**Tun Abdul Razak Research Centre-TARRC, UK

1
ABSTRACT

In part I, a simple “viscoplastic” model capable of implementation in existing


commercial FEA packages, with the scope to capture those aspects of the stress-strain
behaviour of filled rubber that are most significant in engineering applications was
proposed. In order to achieve the appropriate elastoplastic contribution to the stress
response of filled rubber a number of elastic-perfectly-plastic material models had to be
overlaid. This increased the time required to model a component as well as the
computation time
In this paper two plasticity models, recently implemented in Abaqus, with non-
linear hardening rules appropriate to filled rubber were examined. These are Multilinear
Kinematic Hardening Plasticity (MKHP) model and Multiple Back-Stress Kinematic
hardening (MBSK) model The predictions of the viscoplastic models consisting of a
viscoelastic material in parallel with either of these two plasticity models were compared
with those predicted using the original “Overlay” model. It is shown that there is a very
good agreement between all three models when predicting the dynamic properties of
filled rubber in simple shear mode of deformation. The MKHP model predicted identical
results to the Overlay model.
Finally, the predictions of the “Fletcher-Gent” or “Payne” effect in a bi-axial
deformation such as the cyclic response of filled natural rubber in pure shear, when
superimposed on a mean strain, for all three models were examined and compared with
the experimental data. The MKHP model was identified as the better of the two models
examined; providing the user with a more computationally efficient approach than the
Overlay model previously reported.

2
INTRODUCTION

In parts I and II,1,2 a simple “viscoplastic” model was proposed, capable of


implementation in existing commercial FEA packages, with the scope to capture those
aspects of the stress-strain behaviour of filled rubber that are most significant in
engineering application, in particular the Payne or Fletcher-Gent effect. Attention was
given to assembling the model from separately identified physical contributions, so that
not only are the number of parameters small but also they may be at least semi-
quantitatively related to the formulation of the elastomer. It was confirmed that the
proposed “viscoplastic” approach captured the essence of the behaviour when examined
in several modes of deformation and for two materials, one filled NR and the other filled
SBR.

However, the hardening rules for the plasticity models available in commercial packages,
required for predicting the elastoplastic contribution to the stress response, were not
capable of producing the unloading stress-strain paths observed with filled rubber. This
was overcome by decomposing the elastoplastic component into a number of elastic-
perfectly-plastic components in parallel, each represented as a separate FE mesh in an
overlay. Although this approach has been successful, it led to an increase in the
computation time.

Two plasticity models have recently been implemented in the Abaqus FE code that could
provide the appropriate stress-strain hysteresis loops. These are Multilinear Kinematic
Hardening Plasticity (MKHP) model and Multiple Back-Stress Kinematic hardening
(MBSK) model. This paper begins by examining in cyclic simple shear the predictions of
the overlaid “viscoplastic” model, described in I, having its overlay of elastic-perfectly-
plastic materials replaced with either one MKHP or one MBSK model. The results are
compared with the predictions of the original viscoplastic model (i.e one viscoelastic
material plus overlaid models of elastic-perfectly-plastic materials) and the experimental
data from which the original model was fitted The models are then subjected to a
benchmark test to establish whether they are capable of predicting the behaviour of filled
rubbers in uniaxial and biaxial stress configurations. This was prediction of the cyclic
stress response of filled rubber in the orthogonal directions when subjected to cyclic pure
shear strain superimposed on a mean strain.

THE BASIC MODEL

The basic model, consisting of hyperelastic, viscoelastic and elastoplastic contributions in


parallel, is described in a previous paper1 (Ahmadi et al., referred to hereafter as Part I),
where it was also shown that the model is cast in a suitably general form for 3-D
implementation in Finite Element Analysis. According to the model the shear stress τ is
given by

τ = τ he + τ ep + τ ve (1)
where the subscripts he, ep and ve stand for hyperelastic, viscoelastic and elastoplastic
respectively.

3
For filled rubber the elastoplastic stress contribution is considered to be governed by two
curves2 (see Figure 1):
Loading dτ epL d 2τ epL
τ ep = τ epL (γ ) : ≥ 0 and → 0 as γ → ∞; ≤0
curve dγ dγ 2
τ

Δγ

Δτ
τepL(γ) = γ 0.5

Δτ γ

Δγ
Δτ ≡ Δτ epR(γ ) = 2τ epL( )
2
Retraction
Δγ curve

FIG 1- Schematic diagram showing hysteresis loop for an elastoplastic material with the loading curve
given in equation 4, when A = 1 and b = 0.5

1) The Loading curve


The loading of the pre-conditioned material from a stress and strain free state is given by
a function:
dτ epL d 2τ epL
τ ep = τ epL (γ ) : ≥ 0; <0 (2)
dγ dγ 2
where L stands for loading.

2) The Retraction curve


The change in stress following a change in direction of straining is calculated
from an incremental relationship:
Δγ
Δτ ≡ Δτ epR (γ ) = 2τ epL ( ) (3)
2
where R stands for retraction. The relationship is only valid as long as there is no further
reversal in the direction of straining and the deformation is appropriately limited.

In Part I it was proposed that


−b / 2
⎛γ2 ⎞
τ epL (γ ) = Aγ (γ + c ) ≡ Gepγ ⎜ 2 + 1⎟
2 2 −b / 2
(4)
⎝c ⎠
−b
where Gep ≡ Ac is the elastoplastic shear modulus
contribution in the small strain limit

4
as suggested by Davies et al3. for fitting the high stiffness at low strain, but rapid fall with
increasing strain, characteristic of rubbers with reinforcing filler. From Equation (4) the
elastoplastic contribution to the secant shear modulus, equal to the contribution to the
dynamic storage modulus because there is no rate dependence in the elastoplastic model,
is given by
τ epL
′ =
Gep = A(γ 2 + c 2 ) −b / 2 (5)
γ

Whence, the ratio of the excess of storage modulus at strain γ over that at high strain to
the maximum value of the excess is given by:
−b / 2
G′ − G∞′ Gep′ (γ ) ⎛ ⎛ γ ⎞ ⎞
2

≡ = ⎜1 + ⎜ ⎟ ⎟ (6)
G0′ − G∞′ Gep′ (0) ⎝⎜ ⎝ c ⎠ ⎠⎟

PLASTICITY MODELS

Two plasticity models, recently implemented within Abaqus, were considered to have the
hardening rules required to construct the hysteresis loops exhibited by filled rubber.
These are:

a) Multilinear Kinematic Hardening Plasticity (MKHP) model.


This constitutive model is based on the original work of Besseling (1958)4, also known as
the “overlay” model5,6,7,8 or “Fraction” model9. It is provided in the form of a built-in
UMAT user subroutine within Abaqus and therefore Fortran compilers are not required.
The model instead of using “conventional” kinematic hardening rule, uses an overlay of
elastic-perfectly plastic material properties, in principal, very similar to the plasticity
model described in part 1 but with the exception that only one set of FE mesh is required.

The material is assumed to have various subvolumes, which have different yield strengths
but the same elastic modulus and are subjected to the same total strain. When multiple
subvolumes are combined together, complex material behaviours, such as multilinear
hardening and the Bauschinger effect9, can be modeled. The multilinear kinematic
hardening model takes piece-wise linear stress-plastic strain curves as input (see Figure
2). The number of subvolumes used in the model is equal to the number of stress-plastic
strain pairs given on the curve.

The yield strength of the ith subvolume is calculated as4:


σ Yi = σ i + 3μip (7)
and the weight of the ith subvolume is:
i −1
1
wi = − ∑ wk (8)
H
1+ i k

where μ is the shear modulus of the material and Hi is the hardening modulus between the
ith data point and the (i +1)th. It is assumed that the hardening modulus beyond the last

5
data point is zero; from the above equation, this ensures that the sum of all subvolume
weights is one.

σ
σ5
σ4 H4
σ3 H3

σ2 H2

σ 1 H1

ε 1p ε 2p ε 3p ε 4p ε 5p ε p
FIG 2: Stress-plastic strain curve for the multilinear kinematic hardening (MKHP) model

In a given solution (time) increment, all subvolumes are subjected to the same strain
increment. For each subvolume, the increment in plastic strain and the total stress at the
end of the increment are computed using a von Mises yield criterion with the associated
flow rule. The total stress and total plastic strain for the entire volume is then calculated
as:
N
{σ } = ∑ wi {σ i } (9)
i
N
{ε } = {ε } + ∑ w {Δε }
p p
0 i i
p
(10)
i =1

where N is the number of subvolumes and {ε } is the total plastic strain at the
p
0

beginning of the increment.

b) Multiple Back-Stress Kinematic hardening (MBSK) model


This is a nonlinear isotropic/kinematic hardening model based on the work of Lemaitre
and Chaboche (1990)10.

The evolution law of this model consists of two components: a nonlinear kinematic
hardening component, which describes the translation of the yield surface in stress space
through the backstress, α ; and an isotropic hardening component, which describes the
change of the equivalent stress defining the size of the yield surface, σ 0 , as a function of
plastic deformation. For our purposes in modelling the Payne effect in elastomers, the

6
size of the yield surface is held constant and the hardening behaviour reduces to a purely
kinematic form.
The kinematic hardening component is defined to be an additive combination of a purely
kinematic term (linear Ziegler hardening law11) and a relaxation term (the recall term),
which introduces the nonlinearity. In addition, several kinematic hardening components
(backstresses) can be superposed, which may considerably improve results in some cases.
When temperature and field variable dependencies are omitted, the hardening laws for
each backstress are
1
α k = ck ε p 0 (σ − α ) − γ k α k ε p (11)
σ
and the overall backstress is computed from the relation
N
α = ∑ αk (12)
k =1

where N is the number of backstresses, and Ck and γk are material parameters that may be
calibrated from cyclic test data. Ck are the initial kinematic hardening moduli, and γk
determine the rate at which the kinematic hardening moduli decrease with increasing
plastic deformation. When Ck and γk are zero, the model reduces to an isotropic hardening
model. When all γk equal zero, the linear Ziegler hardening law11 is recovered.

MATERIAL AND TESTS

Two double shear testpieces1, moulded from a filled natural rubber material (Table I),
were used to characterize the material. All tests were carried out using an Instron 1271
servo hydraulic test machine under displacement control.

TABLE I
FORMULATION AND CURE DETAILS
SMR CV60 Natural rubber 100
N 330 Carbon Black 45
ZnO 5
Stearic Acid 2
HPPD 3
Sunproof Improved Wax 2
CBS 0.6
Soluble Sulphur 2.5
Cure Time @ 150°C / Minutes 18
Shore A Hardness 60

One “virgin” test piece was subjected to 30 cycles at 1 and 10 Hz, in a sequence of
increasing amplitudes from 0.1% dynamic shear strain to 200%. The testpiece was rested
for 5 seconds between cycling at every frequency and amplitude. This testing sequence
was repeated to eliminate the Mullins effect- i.e. the semi-permanent stress-softening that
occurs during the primary loading. The force and deflection were recorded using the
testing machine’s internal sensors and the dynamic properties were calculated using the

7
secant method12. The dynamic properties data for the second test sequence were used to
find the parameters of the overlay model described in Part I.

Figure 3 shows an apparatus for applying a pure shear deformation to rubber, enabling
measurement of the longitudinal extensional force as well as the lateral force developed
in the clamps, preventing departure of the lateral extension ratio of the rubber from
unity13.

Measurements were made of the dynamic longitudinal and lateral forces when a sheet of
rubber, made from the material described in Table 1, was subjected to a pre-extension
(i.e. mean extension) of λ = 1.6 and superimposed cyclic deformations of strain
amplitudes (λmax-λmin) from +/- 0.1% (of 200mm) up to +/- 50% at a frequency of 1 Hz.
The rubber was scragged to 10 cycles of λ=2.41 at a frequency of 0.1 Hz prior to these
tests. The test sequence was as followed: an extension of 12 mm was applied to the sheet
in 20 seconds, along the direction parallel to 20mm side, and held for 300seconds before
applying the cyclic tests. This was then followed by 18 dynamic amplitude tests ranging
from 0.1% to 50% each at 1 Hz and for 20 cycles. A period of 20 seconds wait was
allowed between the cyclic tests.

FIG 3: Split pure shear apparatus (rubber sheet nominal dimensions: width=200, height=20 and thickness
=1mm.)

FIT OF THE PARAMETERS TO EXPERIMENTAL DATA

a) The overlay model


TABLE II.
FITTED MODEL PARAMETERS
Model part Parameter Value
Hyperelastic C1 /MPa 0.3619
Linear viscoelastic H0 /MPa 0.0152
A /MPa 0.1936
Elastoplastic b 0.4647
Offset σep(γmax) /MPa 0.0567

8
Table II and III show the parameters of the overlay model fitted to the filled natural
rubber used in this paper-see Table I. The description of the model and the methodology
for fitting the parameters to the experimental data are described in Part I.

TABLE III.
PARAMETERS OF DISCRETISED MODEL
Model part Parametersa Values
Finite Linear C1 /MPa 0.4884b
Viscoelasticity G1 /MPa; τ1/s 0.035; 10-2
G2; τ2 0.035; 10-1
G3; τ3 0.035; 1
G4; τ4 0.035; 10
G5; τ5 0.035; 102
G6; τ6 0.035; 103
G7; τ7 0.035; 104
Elastoplastic 1 T1/MPa; E1/MPa 0.0106; 3.6550
Elastoplastic 2 T2; E2 0.0040; 0.7573
Elastoplastic 3 T3; E3 0.0055; 0.5766
Elastoplastic 4 T4; E4 0.0076; 0.4345
Elastoplastic 5 T5; E5 0.0105; 0.3275
Elastoplastic 6 T6; E6 0.0595; 1.0121
Note: (a). The bulk modulus is an additional parameter. It is fixed in the Elastoplastic models by choice of a
Poisson ratio of ν= 0.49999. (b). In Abaqus the user can use the value (0.5G∞) of C1 at infinite time, or the
value (0.5G(0)) at infinitesimal time, being greater by 0.5ΣGi (see Equation (6) in Part I).

b) MKHP material model


0.10
Resultant Element 1
Element 2 Element 3
0.08 Element 4 Element 5
Element 6
Tensile stress (Mpa)

0.06

0.04

0.02

0.00
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07
Total Strain
FIG 4 The variation of the tensile stress with total strain for the six elements of the Overlay model. The
solid line is the total contribution of all six elements and is required to construct the yield tensile stress
versus plastic strain curve. This is then submitted to Abaqus in the form of an input file.

9
The elastic-perfectly-plastic components of the overlay model described in Table III was
used to reconstruct a piecewise linear relationship between the tensile stress and the total
strain for the elastoplastic part of the model, see Figure 4. This information was then used
to define the piecewise linear yield stress-plastic strain curve and supplied to the solver
via Abaqus’s input file.

For the MKHP material model, the “user interface” in the Abaqus input file (.inp file) is
the general purpose UMAT interface. The user is required to define Young’s modulus (E)
and Poisson’s ratio (ν ), and then provide a set of points defining the piece-wise linear
yield stress – plastic strain curve. By definition Abaqus knows that this model is only a
purely kinematic hardening model. An extract of the input file is given below:

EXTRACT FROM MKHP MATERIAL MODEL INPUT FILE


*MATERIAL,NAME=ABQ_MULTILIN_KINHARD_1
*USER MATERIAL, CONSTANTS = 64
0.49999, 6.7630,,
,,
0.000000, 0.0196
0.001287, 0.0270
0.004064, 0.0370
0.009931, 0.0511
0.021615, 0.0706
0.044342, 0.0977
*DEPVAR
48

c) MBSK material model


For this material model, the user defines several parameters that control the size and
shape of the yield curve. The yield curve may be defined by the summation of N
exponential functions.
N C
σ 0 = σ 0 + ∑ k =1 k (1 − e−γ k ε )
p
(13)
γk
The curve is defined by parameters Ck and γk . N is the number of backstresses, Ck and γk
are the initial kinematic hardening moduli and the rate at which the kinematic hardening
moduli decrease with increasing plastic deformation respectively. The kinematic
hardening law can be separated into a deviatoric part and a hydrostatic part; only the
deviatoric part has an effect on the material behaviour. When Ck and γk are zero, the
model reduces to an isotropic hardening model. When all γk equal zero, the linear
hardening law is recovered. It must be mentioned that for the nonlinear kinematic
hardening model used in this paper, σ 0 , the initial size of the yield surface, was set to a
constant and does not change. The yield surface only translates in stress space.

10
0.12

0.10
Yield Stress (MPa)

0.08

0.06 Target Curve


Fit
0.04
Term 1

0.02

0.00
0.00 0.02 0.04 0.06 0.08 0.10
Plastic Strain
FIG 5 The variation of the yield tensile stress with plastic strain. The “Fit” curve is Eq. 13 fitted, with N=1,
to the piecewise linear data, shown as “target curve”. The “Term 1” line shows the contribution of the
summation part of the Eq. 13 with only one term.

For the filled natural rubber tested in this paper, an Excel spreadsheet was used to find
the best fit to the yield curve. The construction of the yield curve from the elastoplastic
component of the overlay model was discussed in the above section describing MKHP
model. This curve is shown as “target curve” in Figures 5 and 6. Two functions were
tried; one with N=1 and the other N=4. The quality of fit is shown in Figures 5 and 6. It
was decided that the 1 term MBSK model (one set Ck and γ k terms) followed the “target
curve” reasonably closely. The resulting Abaqus input file definition of the MBSK model
is:

EXTRACT FROM MBSK MATERIAL MODEL INPUT FILE


*Material, Name=MBSK
*Elastic
6.7630, 0.49999
*Plastic, Hardening=Combined, Data type=Parameters, Number=1
0.0196, 4.0, 50.0

where 0.0196, 4.0, 50.0 are σ 0 , C1 and γ 1 respectively.

It is worth mentioning that a general approach would have been to define the MKHP or
MBSK material models using equation 4 for the shear stress-strain relationship and use
this function to calculate the tensile yield stress-plastic strain curve.

11
0.14
Fit Term 1
0.12 Term 2 Term 3
Term 4 Target Curve
0.1
Yield Stress (MPa)

0.08

0.06

0.04

0.02

0
0 0.02 0.04 0.06 0.08 0.1
Plastic Strain
FIG 6 The variation of the yield tensile stress with plastic strain. The “Fit” curve is Eq. 13 fitted, with N=4,
to the piecewise linear data, shown as “target curve”. The “Term 1” to “Term 4” lines shows the
contribution of each of the terms in the summation part of the Eq. 13.

PREDICTIONS OF THE DYNAMIC PROPERTIES

The above three material models were used to simulate the dynamic response of filled
natural rubber in simple shear and pure shear modes of deformation. Simulation of the
behaviour of the material in simple shear was used as a rapid exercise to establish if there
are significant differences in the predictions of the three models. For this purpose a single
2D plane strain model was used to model the test piece. The Overlay model had seven
superposed elements one defined by a viscoelastic material and 6 by elastic-perfectly-
plastic materials. The analysis involved subjecting the model to dynamic shear strain
amplitudes from 0.1% to 100 at 1 Hz. This analysis was repeated for where the material
is first subjected to pre-strains of 10, 30 and 50% and then the cyclic inputs are
superimposed on this deformation. The purpose of this was to investigate the effect of
prestrain on the dynamic response of the material.

The pure shear simulations were carried out with two FE models. The first model
exploited the symmetry of the pure shear test piece and used one linear brick element,
with reduced integration (8 noded, C3D8RH), to represent 1/8 symmetry model of the
test piece. In the second model the number of elements was raised to 1000. The former
does not allow for the edge effects shown by the test piece, where as the multi-element
model clearly showed this behaviour by predicting that the free edges of the sheet curve-
in towards the centre of the sheet

For many of the MKHP analysis it was found that at the 50% dynamic amplitude
hourglassing (keystoning) appeared in the deformed shapes. This was rectified when
using a full integration element, C3D8H that by definition will not exhibit keystone
behaviour

12
3.5
MBSK
3 MKHP

2.5 Overlay
Modulus (MPa).

1.5

0.5

0
0.001 0.01 0.1 1
Dynamic Strain Amplitude
FIG 7 The predicted dynamic shear modulus for filled natural rubber, at zero prestrain, by the three models.
The MKHP and Overlay model data are coincident.

0.3
MBSK

0.25 MKHP

Overlay
0.2
sin ( δ )

0.15

0.1

0.05

0
0.001 0.01 0.1 1
Dynamic Strain Amplitude
FIG 8 The predicted sin(δ) for filled natural rubber, at zero prestrain, by the three models. The MKHP and
Overlay model data are coincident

Figure 7 and 8 show the simulated shear modulus and sin(δ) results for Filled Natural
Rubber (FNR) for each of the three models. The Overlay and MKHP models results
coincide and are identical. However, there is a slight difference in the modulus data and a
more significant difference for the sin(δ) results between the MBSK and the other two
models. The presence of the pre-strain does not seem to have an effect on the dynamic

13
properties predicted by all three models, shown in figure 9 and 10. The experimental data
has been presented14,15,16 for the case of dynamic excitations superimposed on static
strains and a cylinder of rubber subjected to combined axial and torsion which showed
that the dynamic behaviour of filled rubber is not affected by the general state of strain
within the material.
3.5

2.5
Modulus (MPa).

1.5

0.5

0
0 0.1 0.2 0.3 0.4 0.5
Pre-strain
FIG 9 The effect of pre-shear strain on the dynamic shear modulus for filled natural rubber, as predicted by
the three models. The Overlay model data are not shown as they coincide with the MKHP data. MKHP,
MBSK, 0.002 strain, 0.01 strain, 0.2 strain, 0.05 strain, 1
strain
0.6

0.5

0.4
Sin( δ )

0.3

0.2

0.1

0
0 0.1 0.2 0.3 0.4 0.5
Pre-strain
FIG 10 The effect of pre-shear strain on sin(δ) for filled natural rubber, as predicted by the three models.
The Overlay model data are not shown as they coincide with the MKHP data. MKHP, MBSK,
0.002 strain, 0.01 strain, 0.2 strain, 0.05 strain, 1 strain

14
The predictions of the longitudinal and “apparent” lateral moduli as a function of
dynamic longitudinal strain amplitude are shown in Figures 11 and 12. These figures also
show the experimental data. As reported previously, 14,15,16 the “Fletcher&Gent” or
“Payne” effect observed in the cyclic simple shear or cyclic tensile deformation appears
to be present in the transverse response of the FNR. All three models have also predicted
this behaviour.

The data for the one element models predict slightly higher moduli than those predicted
by 1000 element model. This is expected, as one-element models are not capable of
catering for the edge effect. The Overlay simulation runs, for the 1000 element model,
had to be truncated at 10% dynamic shear strain amplitude due to excessive computation
time. Although this had not been experienced before but during the course of this work
computation time was excessive, due to large number of iteration required to reach
convergence.

One-element planar tension models were used to compare solver statistics. Four material
models were used:

TABLE IV
Material model Comment Cut backs Solver passes
MKHP Results OK, best performance 6 22441
MBSK Cyclic relaxation behaviour 6 26399
Overlay ν =0.49999 Results OK, run time long 7756 49442
Overlay ν =0.499 Results OK, run time short 6 24944

The specified time step during the harmonic loadings was 0.025, or 40 time steps per
period, but cut backs in the time step size was allowed if needed to achieve convergence.

The models were subjected to the total loading sequence i.e. ramping up to the mean
stress in 20 seconds holding for 5 minutes and then applying cyclic loads for 20 cycles,
followed by 20 seconds wait between each cyclic test. The mean load and cyclic loads
were then examined for each model during the quasi-static and relaxation phases to
monitor the effect of viscoelastic component of the model on the predicted response. The
Overlay and MKHP models both showed qualitatively the expected behaviour. However,
the MBSK model exhibited higher relaxation rate than the other two models and more
importantly the mean stress during the cyclic inputs were changing per cycle. The reason
for this behaviour is not clear even though this model does not have an isotropic
component.

The predicted longitudinal modulus for all three models is in good agreement with the
experimental data-see Figure 11. However, it is interesting to note that this is not the case
for the transverse modulus at small dynamic strain amplitudes shown in Figure 12. The
reason for the disagreement would require further investigation.

15
6
MBSK 1000 element
MKHP 1000 element
5 Overlay 1000 element
MBSK 1 element
4 MKHP 1 element
Modulus (MPa).

Overlay 1 element
Experimental
3

0
0.001 0.01 λ max-λ min 0.1 1

FIG 11 Longitudinal dynamic modulus as a function of dynamic strain amplitude (λmax -λmin). The
simulated data are shown for all three material models using a) one solid element 1/8 symmetry FE model
and, b) 1000 solid elements 1/8 symmetry FE model

3.5
MBSK 1000 element
MKHP 1000 element
3
Overlay 1000 element
MBSK 1 element
2.5
MKHP 1 element
Modulus (MPa).

Overlay 1 element
2
Experimental

1.5

0.5

0
0.001 0.01 λ max-λ min 0.1 1
FIG 12 Transverse dynamic “modulus” as a function of dynamic strain amplitude (λmax -λmin). The
simulated data are shown for all three material models using a) one solid element 1/8 symmetry FE model
and, b) 1000 solid elements 1/8 symmetry FE model

16
SUMMARY & CONCLUSIONS
In this paper, two recently implemented plasticity models within Abaqus have been
assessed for their capability to provide the appropriate kinematic hardening rule that
governs the stress-strain behaviour of filled natural rubber under time varying inputs.
These were Multilinear Kinematic Hardening Plasticity (MKHP) model and Multiple
Back-Stress Kinematic hardening (MBSK) model. The MKHP model, when used in
parallel with a viscoelastic model, predicted identical dynamic properties response in
simple shear to the viscoplastic model, described in Part I, consisting of a viscoelastic
model in parallel with a number of overlaid elastic-perfectly-plastic materials. The
MBSK model gave similar results, however, it was noticed that under cyclic loads the
mean stress response predicted by this model decayed gently during the excitation. The
reason for this behaviour is not clear. Both models showed that the magnitude of the state
of the strain in the material does not eliminate the manifestation of the “Fletcher-Gent” or
“Payne” effect by the material. This is in line with published observations13, 14,15. The
cyclic pure shear experiment was used as a benchmark test to examine the capability of
the two nonlinear kinematic hardening models to predict the “Fletcher-Gent” or “Payne”
effect in bi-axial mode of deformation. This was exhibited by both models and agreed
well with the overlay model predictions. The longitudinal dynamic modulus predicted by
all models agreed remarkably well with the experimental data. However, this was not
observed for the transverse dynamic “modulus”. The reason for this discrepancy is not
clear and would require further investigation.

The MKHP model appears to be a close representation of the Overlay model, and would
therefore provide a much more computationally efficient approach than the Overlay
model.

ACKNOLEDGEMENT
Assistance of Ms J. Picken in preparing this manuscript is very much appreciated.

REFERENCES
1. AHMADI, H.R, KINGSTON J G R & MUHR A H, (2008) Dynamic properties
of filled rubber. I – Simple model, experimental data and simulated results.
Rubber Chemistry and Technology, 81, 1-18.
2. AHMADI, H.R, & MUHR A H, (2007) Dynamic properties of filled rubber. II -
physical basis of contributions to the model. Paper 39 of Fall 172nd Technical
Meeting of the Rubber Division of the American Chemical Society, Cleveland,
Ohio.
3. DAVIES, C.K.L., DE, D.K., & THOMAS, A.G., (1994) Characterization of the
behaviour of rubber for engineering design purposes (I) stress and strain relation
Rubber Chemistry and Technology, 67, 716-729.
4. BESSELING, J.F., (1958) A theory of elastic, plastic and creep deformations of
an initially isotropic material showing anisotropic strain-hardening, creep
recovery and secondary creep, Journal of Applied Mechanics, 529-536
5. MROZ, Z. (1973) Mathematical models of inelastic material behaviour, Solid
Mechanics Division, University of Waterloo, Ontario, Canada.

17
6. OWEN, D.R.J., PARAKASH, A., & ZIENKIEWICZ, O.C., (1974) Finite element
analysis of nonlinear composite materials by use of overlay systems, Comp. &
Struct., 4, 1251-1267
7. HARPER, P.G., (1978) An Analysis of Sublayer Model for Plasticity under
Multiaxial Stress, Report No. RD/B/N4270, CEGB, Berkeley Nuclear
Laboratories.
8. ZIENKIEWICZ, O.C., NAYAK, G.C., & OWEN, D.R.J., (1973) Composite and
overlay models in numerical analysis of elasto-plastic continua, Foundations of
Plasticity, ed. A. Sawcczuk, Noordhoff, Leiden, pp.107-123.
9. BESSELING, J.F., & van der GIESSEN, E., (1994) Mathematical Modelling of
Inelastic Deformation, Applied Mathematics and Mathematical Computation 5,
Chapman & Hall- London.
10. LEMAITRE, J., & CHABOCHE, J.L., (1990) Mechanics of Solid Materials, CUP
Cambridge, UK
11. ZIEGLER, H., (1959) A modification of Prager’s hardening rule., Q. Appl. Math.,
17, 55-65.
12. AHMADI, H.R. & MUHR, A.H., (1997) Modelling the dynamic properties of
filled rubber. Plastics Rubber, & Composites, 26, 451-46
13. GOUGH, J., GREGORY, I.H., & MUHR, A.H., (1999) Determination of
constitutive equations for vulcanised rubber, Boast, D. and Coveney, V.A (eds.)
Finite Element Analysis of Elastomers, 5-26 Bury St Edmonds: Professional
Engineering Publishing
14. AHMADI, H.R., KINGSTON, J.K.R., & MUHR, A.H., (2005) Modelling
dynamic properties of filled rubber in uniaxial and biaxial stress configurations,
Austrell, P.E, and Kari, L. (eds.), Constitutive Models for Rubber IV, 299-304.
15. GREGORY, M.J., (1985) Dynamic properties of rubber in automotive
engineering. Elastomerics, 117,19-24.
16. AHMADI, H.R., GOUGH, J, MUHR, A.H., & THOMAS, A.G. (1999) Bi-axial
experimental techniques highlighting the limitations of a strain-energy
description of rubber, Dorfmann, A.L., and Muhr, A.H. (eds.) Constitutive
Models for RubberI, 65-71 Rotterdam: Balkema.

18

View publication stats

You might also like