Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Mechanics of Materials 52 (2012) 40–57

Contents lists available at SciVerse ScienceDirect

Mechanics of Materials
journal homepage: www.elsevier.com/locate/mechmat

Inelastic material behavior of polymers – Experimental


characterization, formulation and implementation of a material model
Markus Kästner ⇑, Martin Obst, Jörg Brummund, Karin Thielsch, Volker Ulbricht
Institute of Solid Mechanics, TU Dresden, D-01062 Dresden, Germany

a r t i c l e i n f o a b s t r a c t

Article history: In this contribution an experimental procedure based on displacement controlled tensile
Received 6 April 2011 tests at different rates of loading, relaxation experiments and deformation controlled load-
Received in revised form 12 April 2012 ing and unloading processes with intermediate relaxations to experimentally characterize
Available online 26 April 2012
and classify the nonlinear, inelastic mechanical behavior of polymers is presented. These
experiments provide data for a structured approach to parameter identification. In line
Keywords: with the experiments, a small strain uniaxial viscoplastic material model is derived, subse-
Polymer
quently generalized to multiaxial loadings and implemented into a finite element program.
Polypropylene
Nonlinear viscoelasticity
The combination of the experimental procedure and the proposed material model is then
Viscoplasticity used to characterize and model the mechanical behavior of the thermoplastic polypropyl-
ene. After the identification of the necessary material parameters, stress–strain curves have
been computed for different uni- and multiaxial loadings and are compared with experi-
mental results.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction behavior considered in terms of a phenomenological con-


stitutive model.
Polymeric materials generally exhibit a nonlinear, To this end, the following paper focuses on the experi-
inelastic material behavior. One of the dominating charac- mental characterization and the constitutive modeling of
teristics is the strain rate dependent deformation behavior the material behavior of polymers. The developed proce-
which in turn leads to relaxation and creep phenomena dures and material models are applied to a typical thermo-
that can impair the long-term performance of structural plastic – namely polypropylene. After a short introduction
components. to the used classification of the mechanical material
On the other hand, the dissipative deformation pro- behavior and existing modeling approaches in Section 1,
cesses allow for the development of energy absorbing experimental results obtained from monotonic tensile
structures for crash and impact loadings. The design and tests, relaxation experiments and loading/unloading pro-
the structural analysis of these components necessitate cesses are outlined in Section 2. These data are used to for-
the formulation and numerical implementation of consti- mulate appropriate uniaxial viscoplastic constitutive
tutive models that capture the inelastic material behavior equations based on rheological models (Section 3) which
for multiaxial loading conditions. In addition, a structured are subsequently generalized to multiaxial loadings. The
approach to the identification of the material parameters developed material model is able to represent a rate inde-
from experimental observations is required since the data pendent hysteresis as well as rate dependent material
obtained from monotonic tensile tests do not allow for a behavior. A time discretization of the constitutive equa-
clear distinction of different sources of inelastic material tions provides the basis for the numerical implementation
of the material model in terms of a stress update algorithm
⇑ Corresponding author. Tel.: +49 351 46332656. and an algorithmic consistent tangent stiffness which al-
E-mail address: markus.kaestner@tu-dresden.de (M. Kästner). low for the application of the model in a nonlinear finite

0167-6636/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.mechmat.2012.04.011
M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57 41

element analysis. In Section 4 aspects of parameter identi- In the case of strain rate independent material behavior
fication and several alternative modeling approaches are every stress–strain curve is an equilibrium relation. On
discussed. The material behavior is simulated for both – the other hand the existence of a hysteresis in the stress–
uniaxial and multiaxial isothermal loadings and compared strain curve does not allow for a clear classification of
to experimental observations. The paper is finally closed by the material behavior in the case of strain rate dependence.
some concluding remarks and an outlook to future work in Therefore, strain rate dependent and independent phe-
Section 5. nomena have to be separated by a suitable experimental
procedure and the shape of the equilibrium relation has
to be determined. As outlined in Section 2.2, this is accom-
1.1. Classification of material behavior plished by a set of monotonic tensile tests, relaxation
experiments as well as loading/unloading process with
Concerning the classification of the observed phenom- and without intermediate holding times.
ena we follow a proposal of Haupt (Haupt, 1993) who sug-
gests four categories of mechanical constitutive behavior 1.2. Material models for polymers
based on the strain rate dependence of the material behav-
ior and on the examination of an equilibrium relation. The modeling of the mechanical behavior of polymers is
Examples for equilibrium points are the terminal points very often based on phenomenological material models, for
observed in relaxation (e_ ¼ 0) or creep (r_ ¼ 0) experi- instance motivated from experimental observations. In
ments. A set of equilibrium points for different strain states many cases simple uniaxial rheological models are the basis
forms the equilibrium relation req (Fig. 1). for the formulation of constitutive equations which are sub-
If the stress–strain curves of experiments carried out at sequently generalized to multiaxial loadings or augmented
various strain rates differ from each other, the material to capture special characteristics of the material behavior.
behavior is named strain rate dependent. The difference This section gives a brief overview on existing material
between the stress measured in an experiment at a strain models which influenced the development of the presented
rate je_ j > 0 and the equilibrium stress is called overstress model or demonstrate alternative approaches.
rov . The mathematical representation of inelastic, process
The second essential feature of the material behavior is dependent material behavior involves the definition of
the shape of the equilibrium relation obtained from load- internal constitutive variables and the formulation of rate
ing and unloading processes. The classification distin- equations which govern their evolution. Many material
guishes between equilibrium relations with and without models, (e.g. Kletschkowski et al., 2002a; Lion, 1997c; Reese
hysteresis (Fig. 1). According to these criteria, the following and Govindjee, 1998), use internal variables of strain type
classification can be made: which can be identified as viscous or plastic strains in the
related rheological models. On the other hand, the utiliza-
 Elasticity: strain rate independent without hysteresis tion of the overstress in Holzapfel (1996), Lion (1996) and
(Fig. 1(a)), Simo (1987) corresponds to internal variables of stress type.
 Plasticity: strain rate independent with hysteresis In addition to these two principal approaches, it is generally
(Fig. 1(b)), possible to define rate and deformation dependent material
 Viscoelasticity: strain rate dependent without equilib- properties as applied in Drozdov (2009) and Drozdov and
rium hysteresis (Fig. 1(c)), Christiansen (2007a,b) to model the cyclic material behav-
 Viscoplasticity: strain rate dependent with equilibrium ior of different polymers.
hysteresis (Fig. 1(d)). One of the most important characteristics of polymers
that has to be captured by the constitutive equations is
the rate dependent material behavior which also results
(a) (b) in creep and relaxation phenomena for constant applied
stress or strain states, respectively. Starting from simple
uniaxial rheological models of viscoelasticity (Tschoegl,
1989), e.g. the generalized Maxwell model, both – small
and large strain formulations – of linear (Kaliske and Rot-
hert, 1997) and nonlinear viscoelasticity (Reese and Gov-
indjee, 1998) have been developed by generalizing the
(c) (d) constitutive equations. Compared to these standard ap-
proaches, models of fractional viscoelasticity (Bagley and
Torvik, 1983; Müller et al., 2011) have been found to de-
scribe pronounced creep and relaxation behavior with less
parameters but generally require a more extensive numer-
ical treatment.
Although the strain rate dependence can be considered
to be the predominant feature of the material behavior,
Fig. 1. Four categories of material behavior according to (Haupt, 1993;
viscoelastic models will not be sufficient to quantitatively
Haupt, 2002) – (a) elasticity, (b) plasticity, (c) viscoelasticity, (d) describe the mechanical behavior of polymers. Therefore,
viscoplasticity. more complex material models are composed of different
42 M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57

branches corresponding to a parallel connection of rheo- measurement with ARAMIS has shown a homogeneous
logical models which account for elastic behavior, rate multiaxial deformation state in the middle of the
independent hysteresis and rate dependent effects. These specimen.
models use an additive split of the total stress and hence
of the free energy into the contributions of the individual
model branches. In line with suitable experiments this also 2.2. Experimental procedure
allows for a structured approach to parameter identifica-
tion. For instance, the large strain behavior of elastomers The testing procedure applied is chosen in line with the
has been realistically modelled by combining models of proposed classification of the material behavior. In the first
nonlinear elasticity, viscoelasticity and plasticity (Lion, instance, monotonic tensile tests at constant strain rates
1997a,c; Miehe and Keck, 2000). Since thermoplastics gen- (Fig. 2(a)) are performed. Any differences between the
erally exhibit inelastic material behavior even at moderate stress–strain curves obtained for different strain rates
strain levels also small strain formulations are of interest. e_ 1 > e_ 2 > e_ 3 indicate the strain rate dependence of the
Corresponding viscoplastic material models (Hartmann, material behavior. The quantification of the strain rate
2006; Kletschkowski et al., 2002a, 2005) even imply con- dependence is then accomplished by relaxation tests at a
stitutively nonlinear elastic equilibrium relations. constant strain e and holding time th (Fig. 2(b)).
Microstructural changes of the polymer network occur- The results obtained in tensile tests and relaxation
ring during the deformation are frequently taken into ac- experiments at a single strain level e do not allow for an
count in a phenomenological way by process dependent unambiguous classification as no unloading processes are
viscosity functions. Different dependencies on strain, strain considered. Hence, there is no information on the existence
rate and overstress (Krempl and Khan, 2003; Lion, 1997c; of an hysteresis. The testing procedure is for this reason ex-
Müller et al., 2011) have been proposed. The evolution of tended by loading and unloading processes with interme-
the microstructure including a recovery of the polymer diate holding times at discrete strain levels in order to
network can also be captured by an additional internal var- separate strain rate dependent and independent effects
iable (Haupt and Sedlan, 2001) which controls a shift of (Fig. 3(a)). Cyclic loadings with different strain amplitudes
relaxation times. Eventually, several combinations with (Fig. 3(b)) are used for verification purposes and can pro-
damage models (Govindjee and Simo, 1992; Miehe and vide information on special phenomena observed in cyclic
Keck, 2000; Simo, 1987) have been used to model complex processes, e.g. the Mullins effect.
inelastic material behavior observed under cyclic loading,
e.g. the Mullins effect. 2.3. Material behavior – strain rate dependence
In addition to the phenomenological modeling of the
inelastic material behavior mentioned so far, several ap- In order to characterize the rate dependence, monotonic
proaches employing kinetic theory of polymer chains and tensile tests have been carried out at different constant
micromechanical considerations to derive macroscopic strain rates Fig. 2. Due to the displacement controlled
constitutive equations (Arruda and Boyce, 1993a,b; experiments three distinct rates of loading of 1, 10 and
Bergström and Hilbert, 2005; Drozdov, 1998; Heinrich 100 mm/min have been chosen. The corresponding strain
and Kaliske, 1997) have been used. rates are summarized in Table 1 and three characteristic
nominal stress–strain curves are shown in Fig. 4.
With higher velocities a clear non-proportional increase
2. Experimental characterization of the material in the stress level and the stiffness can be observed. Simul-
behavior taneously, the strain to failure and the amount of perma-
nent deformation decrease. At a loading rate of 100 mm/
2.1. Experimental set-up min the specimen fails at a maximum strain of 15%. In con-
trast to this, no failure can be observed at a velocity of
All of the displacement controlled experiments were 1 mm/min. In this case a neck develops and propagates
carried out on a standard electromechanic testing machine through the specimen at a virtually constant load level.
with a maximum force of 5 kN. The loading rate can be var- The increasing deformation is concentrated in two process
ied in a range of v ¼ u_ ¼ 0:005 . . . 500 mm/min. An air con- zones at both ends of the neck. As expected, tests at a
ditioning system provided approximately constant
ambient temperatures of # ¼ 296 K. The tensile specimens
of the polypropylene homopolymer Moplen HP500N have (b)
been cut from injection molded plates. Size and geometric
(a)
properties of the specimen are chosen in accordance with
DIN EN ISO 3167 (specimen type A). The nominal stress
has been computed from the measured force F and the
cross section of the specimen A0 in the undeformed state.
In the same manner the nominal strain has been derived
from the ratio of the displacement of the testing machine
u and the reference length lref ¼ 110 mm in the unde-
formed state. The strain state in the specimen was mea- Fig. 2. Monotonic tests – (a) tensile tests at different constant strain rates
sured by digital image correlation (ARAMIS). The and (b) tensile relaxation experiments.
M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57 43

strength of the fibres and the fibre matrix interface respec-


(a)
tively, only small strains to failure ef 6 2:5% are observed.
Therefore, the following experiments as well as the formu-
lation of the material model will focus on the small strain
domain jej 6 5%. From Fig. 5, which shows the magnified
small strain domain of the tensile tests, it can be observed
that the characteristic strain rate dependence dominates
the deformation behavior.
Relaxation experiments allow for the separation of the
strain rate dependent and independent fractions of stress
and are therefore of great importance for the quantification
of the strain rate dependent material behavior. Fig. 6 shows
(b)
the stress–time and temperature–time curves of a relaxa-
tion test with a holding time t h ¼ 48 h ¼ 1:73  104 s per-
formed at a strain level of e ¼ 4:5%. After a pronounced
relaxation of stress at the beginning, the rate of stress de-
creases until the relaxation seems to stop after approxi-
mately 48 h. This stress state is assumed to be an
equilibrium state. The figure also illustrates the influence
of the slightly varying temperature during the experiment.
An increase of the temperature level leads to an enhanced
Fig. 3. Cyclic tests – (a) loading and unloading experiments with relaxation.
intermediate holding times at different strain levels and (b) cyclic
loadings with different strain amplitudes.
2.4. Material behavior – equilibrium relation

Table 1
So far the strain rate dependence of the material behav-
Monotonic tensile tests – applied rates of loading and corresponding strain
rates. ior has been demonstrated in tensile tests, relaxation
experiments allow for its quantification. However, for the
Rate of loading v Strain rate e_ ¼ v =lref ð1=sÞ
clear classification of the material behavior the shape of
0.005 mm/min 7:5  107 the equilibrium relation has to be assessed. Theoretically,
1 mm/min 1:5  104 an estimation of the equilibrium relation can be found
10 mm/min 1:5  103 from continuous tensile tests at a very low strain rate. In
100 mm/min 1:5  102 this kind of experiment every point of the stress–strain
500 mm/min 7:5  102 curve has to be a proper approximation for the equilibrium
stress in order to obtain meaningful results.
In Haupt and Lion (1995) the stress–strain curve ob-
tained for a very low strain rate is compared to tensile tests
35 100 mm/min
with higher strain rates and subsequent stress relaxation
10 mm/min
1 mm/min for a stainless steel. The stress in the terminal points of
30
σ [MPa]

the relaxation is very close to the stress–strain curves of


25

20 35

15 30

10
25
Stress σ [MPa]

5
20
0
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 15
ε [-]
10
Fig. 4. Comparison of characteristic stress–strain curves measured in
monotonic tensile tests at three distinct rates of loading. 5

0
0 0.01 0.02 0.03 0.04 0.05
velocity of 10 mm/min produce intermediate results
Strain ε [−]
regarding the stress level as well as the failure mechanism.
The material model to be developed will be applied in Fig. 5. Comparison of average stress–strain curves measured in mono-
the multiscale analysis of fibre reinforced polymers. As tonic tensile tests at three distinct rates of loading for the small strain
the failure of the composite material is governed by the domain.
44 M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57

two monotonic tensile tests with intermediate relaxations.


It can be seen that four equally spaced holding times of
three hours produce the same terminal point as a single
holding time of ten hours. In order to generate as many ter-
minal points as possible, closely spaced intermediate relax-
ations with short holding times are preferred.
Hence, discrete points of the equilibrium relation are
determined by intermediate relaxation processes every
De ¼ 0:45% with a reduced holding time t hir ¼ 3 h. All load-
ing and unloading processes are carried out at a velocity of
10 mm/min. Fig. 9(a) shows the resulting stress–strain
curve for this experiment. Although the relaxation of the
Fig. 6. Stress–time and temperature–time curves measured in a relaxa-
overstress is not completely finished after the reduced
tion experiment at e ¼ 4:5%, holding time th ¼ 48 h. holding time of 3 h (Fig. 9(b)), the experiment indicates
an equilibrium hysteresis as the terminal points at a cer-
tain strain level do not coincide for the loading and unload-
the continuous experiments. In contrast to this, due to the ing path. At least the existence of an equilibrium relation
pronounced relaxation behavior and the long relaxation req – 0 is proven.
times of polypropylene, the equilibrium relation cannot
be identified from continuous tests in the present case. 2.5. Material behavior – cyclic loadings
This fact is illustrated in Fig. 7. It shows three stress–strain
curves – two for continuous loading at 10 mm/min and Finally, cyclic tension tests without intermediate hold-
0.005 mm/min – the latter representing the smallest possi- ing times have been carried out. As shown in Fig. 3(b),
ble rate of loading for the used equipment. In addition, a the specimen has been loaded with 10 cycles at a strain
monotonic tensile test at 10 mm/min with intermediate amplitude of ^e1 ¼ 2:3%. The measured stress–strain curve
relaxations (holding time t hir ¼ 10 h) is depicted. is plotted in Fig. 10(a). It is noticeable that the stress level
It can be recognized that the stresses obtained for during the first cycle is significantly higher then in the fol-
0.005 mm/min are considerably larger than the terminal lowing ones. Furthermore, the hysteresis becomes smaller
points of the intermediate relaxations. Therefore, it has to with each cycle. It is likewise that the material behavior
be concluded that only discrete equilibrium points can be approaches a stationary hysteresis.
measured. To this end, loading and unloading processes In a second test case two additional amplitudes of
^e2 ¼ 3:2% and ^e3 ¼ 4:5% with ten cycles each have been
with intermediate holding times according to Fig. 3 are em-
ployed. For practical reasons holding times of 48 h and more applied. The stress–strain curve in Fig. 10 (b) shows, that
as used in the relaxation tests are not reasonable in these at each strain level the stresses in the first cycle are again
experiments. A reduced holding higher then in the following ones.
 time
 for the intermediate
relaxations is defined by Drr thir ¼ 0:9  Drr ðt h ¼ 48 hÞ,
i.e. 90 percent of the stress relaxation observed in a relaxa- 2.6. Summary
tion experiment with a holding time of 48 h relaxes during
thir . For polypropylene the condition above leads to holding From the various results obtained from the depicted
times of t hir  3 h (compare Fig. 6). This choice can be fur- experimental procedure it can be concluded that the mate-
ther motivated by Fig. 8 which shows the comparison of rial shows both – strain rate dependence and equilibrium

30 30
Stress σ [MPa]

Stress σ [MPa]

20 20

10 10
10 mm/min 10 mm/min
10 mm/min, thir = 10 h 10 mm/min, t hir =10h
0.005 mm/min 10 mm/min, t hir =3h
0 0
0 0.01 0.02 0.03 0.04 0.05 0 0.01 0.02 0.03 0.04 0.05
Strain ε [−] Strain ε [−]

Fig. 7. Comparison of stress–strain curves measured in continuous Fig. 8. Comparison of stress–strain curves measured in a continuous
tensile tests (minimal rate of loading v ¼ 0:005 mm/min and 10 mm/ tensile test and monotonic tests with two different intermediate holding
min) and a tensile test with intermediate relaxations (intermediate times of thir ¼ 3 h and thir ¼ 10 h, loading is performed at a velocity of
holding times thir ¼ 10 h). 10 mm/min.
M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57 45

(a) (a)
20 20

Stress σ [MPa]
Stress σ [MPa]

10
10

0
0

-10
-10
0 0.01 0.02 0.03 0.04 0.05 0 0.005 0.01 0.015 0.02 0.025
Strain ε [−] Strain ε [−]

(b) (b) 30
20

Stress σ [MPa]
20
Stress σ [MPa]

10

10
0
0

-10
-10

0 0.5 1 1.5 2 0 0.01 0.02 0.03 0.04 0.05


Time t [s] 5 Strain ε [−]

Fig. 9. Results measured in loading and unloading processes with Fig. 10. Stress–strain curves measured in cyclic experiments without
intermediate holding times thir ¼ 3 h – (a) stress–strain and (b) stress– intermediate holding times – (a) 10 cycles with constant strain amplitude
time curve. and (b) 10 cycles at three different strain amplitudes, loading and
unloading are performed at a constant velocity of 10 mm/min.

hysteresis. According to the presented criteria and differ- 3.1. Uniaxial material model
ent types of material behavior it has to be classified as
viscoplastic. In the following a general material model will In order to model both – equilibrium relations with and
be proposed which can capture all features observed in the without hysteresis, the proposed material model for the
experiments. However, depending on the intended appli- equilibrium stress
cation the small equilibrium hysteresis could be neglected
compared to the dominating strain rate dependence. req ¼ re þ rend ð2Þ

consists of a nonlinear elastic


3. Modeling of the material behavior
E1
re ¼ e ð3Þ
In line with the presented experimental procedure, we 1 þ ajej
propose constitutive equations for modeling the inelastic
and an endochronic model (Valanis, 1971a,b) of plasticity
material behavior of polymers which can be characterized
as a viscoplastic material model based on an overstress for-  
rend ¼ E2 e  qend ð4Þ
mulation. The stress
r ¼ req þ rov ð1Þ The stress–strain relations involve the material parameters
E1 ; a and E2 . As illustrated in the corresponding rheological
is a combination of the strain rate independent equilibrium model (Fig. 11), the total strain e is additively decomposed
stress req and the strain rate dependent overstress rov . To-
gether with the experiments performed, this allows for a e ¼ ee þ qend ð5Þ
clear identification of the necessary material parameters.
Similar approaches are applied for instance in Hartmann into an elastic ee and a plastic strain qend which serves as
(2006), Haupt and Lion (1995) and Lion (1996). internal constitutive variable. The superscript end refers
Initially, a uniaxial formulation of the material model to the endochronic model. The internal variable is used
will be developed. The model for the equilibrium relation to phenomenologically account for the process dependent
is addressed at first. In the following the linear and nonlin- plastic deformation and hence the hysteresis in the equi-
ear viscoelastic overstress models are outlined. Both mod- librium relation. Its temporal change is governed by the
els are subsequently generalized to multiaxial loadings. evolution equation
46 M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57

Fig. 11. Rheological model – equilibrium relation.

 
q_ end ¼ b eðzÞ  qend ðzÞ z_ ð6Þ

where b is an additional material parameter which con-


trols the nonlinearity of the stress–strain relation and the
size of the equilibrium hysteresis. The required strain rate
independence of the plasticity model is ensured by the use
of a generalized arclength z, whose rate is defined by Fig. 12. Rheological model – equilibrium relation and overstress.
z_ ðtÞ ¼ je_ ðtÞj. The generalized arclength can be interpreted
as an internal time of the material model. It is therefore
X
nv
named endochronic. Combining Eqs. (4) and (6) yields rov ¼ rov ð9Þ
i
the differential equation i¼1

r_ end ¼ E2 e_  brend je_ j ð7Þ is the sum of the individual stresses


 
rov ov
i ¼ c i e  qi ð10Þ
for the stress r end
in the endochronic model which forms
the basis for the numerical implementation of the material in each Maxwell element. The evolution equations
model. Comparable models have previously been applied ci  
to model the mechanical (Kletschkowski et al., 2002a,b, q_ ov
i ¼ e  qov
i ð11Þ
gi
2004) and thermomechanical (Kletschkowski et al., 2005)
behavior of filled polytetrafluorethylen. describe the temporal change of the internal variables qov i
The set fE1 ; E2 ; a; bg of material parameters (compare during a deformation process. The individual Maxwell
Table 2) is identified from fitting the model prediction to models are characterized by the relaxation strength ci
the terminal points of the intermediate relaxations ob- and the viscosity gi . Their ratio si ¼ gc i is known as relaxa-
i
tained from experiments according to Fig. 3(a). If no equi- tion time. The set {ci ; si } of parameters represents a dis-
librium hysteresis is observed, only the elastic part of the crete relaxation spectrum which can be determined from
stress–strain relation can be used. Linear elastic material the stress–time curves of relaxation experiments, e.g. using
behavior is included in this model for a ¼ 0. the window algorithm of Emri and Tschoegl (Emri and
The constitutive formulation for the rate dependent Tschoegl, 1993; Tschoegl and Emri, 1992).
part of stress is a generalized Maxwell model (Tschoegl, Analogous to the plastic material model a single differ-
1989) with nv parallel elements. From the uniaxial rheo- ential equation
logical model depicted in Fig. 12 it can be observed that c
i ov
the strain in the ith Maxwell element is additively r_ ov
i ¼ ci e 
_ r ð12Þ
gi i
decomposed
can be derived from combining Eqs. (10) and (11). The lin-
e ¼ eei þ qov
i ð8Þ ear viscoelastic Maxwell model is directly related to the
endochronic plasticity model when the time t is replaced
into an elastic eei and a viscous fraction qov
i which serves as by the generalized arclength z. This is known as correspon-
internal variable. The nonequilibrium overstress dence principle (Haupt, 2002).

Table 2
Material parameters of the uniaxial and the multiaxial material model as well as experiments used for identification (nv is the number of used Maxwell
models).

Model branch Uniaxial model Multiaxial model Identification (cf. Sections 4.1, 4.3)
Parameter Number Parameter Number
Nonlinear elasticity E1 ; a 2 G1 ; K 1 ; aG ; aK 4 Equilibrium relation, loading path
Endochronic plasticity E2 ; b 2 G2 ; K 2 ; bG ; bK 4 Equilibrium relation, hysteresis
Nonlinear viscoelasticity ci ; gi ¼ ci si 2nv Gov ov G
i ; K i ; gi ; gi
K 4nv Relaxation test
Nonlinear viscoelasticity s0 ; k0 2 s0 ; k0 2 Tensile test
M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57 47

Nonlinear viscoelastic behavior, which can be observed In order to obtain a generally applicable material model
for many polymers, requires an extension of the linear over- and to be able to capture experimental results from torsion
stress model. To this end, a number of conceivable possibil- and combined tension–torsion experiments, the multiaxial
ities exists. On the one hand some of them can be excluded constitutive equations are decomposed into deviatoric and
by modeling or implementation considerations – e.g. the volumetric parts. This allows for the modeling of the signif-
desired separation of rate dependent and independent ef- icantly different mechanical response for bulk and shear
fects. Therefore, the equilibrium stress state is assumed to loadings which is a typical feature of the material behavior
have no influence on the strain rate dependent overstress. of polymers. Hence, the symmetric strain tensor
On the other hand a detailed analysis of the experimental
1
results can give more insight into the material behavior. A eij ¼ eij þ emm dij ð15Þ
comparison of the reloading after the end of the intermedi- 3
ate relaxations at strain different levels in Fig. 9(a) reveals is split into the deviatoric strain eij and the volumetric
no significant differences in the slopes of the stress–strain strain 13 emm dij . This applies also for the internal constitutive
curve. For this reason, the viscosity should be independent variables qend
ij and qov;k
ij
of the current strain level. Since different strains are associ-
ated to different equilibrium stresses this observation also ðÞ ðÞ;dev 1
qij ¼ qij þ qðÞmm dij ð16Þ
confirms the assumed independence from req . 3
More reasonable influences can be seen in the depen- representing inelastic strains as illustrated in the uniaxial
dency of the viscosity on the strain rate e_ or the current rheological model (Fig. 12). In the same manner the sym-
overstress rov . Here, we introduce a nonlinear strain rate ðÞ
metric stress tensors rij of each model branch can be split
dependence by replacing the constant viscosity gi in the ðÞ ðÞ
into the deviators sij and the hydrostatic stresses 13 rmm dij
evolution Eqs. (11) and (12) by a function according to
!
jrov jk0 1
rðÞij ¼ sðÞij þ rðÞmm dij ð17Þ
g~ i ¼ gi exp  ð13Þ 3
s0
In line with the uniaxial model, the stress
depending on the overstress and two additional parame-
ters s0 and k0 . Hence, the model predicts a decline of the X
nv
rij ¼ reij þ rend
ij þ rov;k
ij ð18Þ
viscosity driven by the current overstress level which also k¼1
results in a shift of the relaxation spectrum to smaller
relaxation times is additively decomposed into the contributions from the
! nonlinear elastic, endochronic and viscoelastic models,
jrov jk0 whereat the former two contribute to the equilibrium rela-
s~i ¼ si exp  ð14Þ tion and the later is a nonequilibrium overstress. In order
s0
to obtain a generalization of the viscoplastic material mod-
The idea of an overstress dependent viscosity function has el, each model branch is extended to multiaxial loadings.
previously been used for modeling of metals (Haupt and A possible hyperelastic generalization for the stress–
Lion, 1995) and elastomers (Lion, 1996, 1997c). strain relation of the nonlinear elastic model is given by
The proposed uniaxial material model involves 2nv þ 6 2G1 K1
material parameters, where nv is the number of parallel reij ¼ eij þ enn dij ð19Þ
1 þ aG jjekl jj 1 þ aK jemm j
Maxwell models. The set of parameters is summarized in
Table 2. The modular structure of the constitutive equa- The nonlinear elastic behavior of the deviatoric and volumet-
pffiffiffiffiffiffiffiffiffi
tions provides the necessary flexibility to model viscoplas- ric parts depends on the norms of the deviator jjeij jj ¼ eij eij
pffiffiffiffiffiffiffiffiffiffiffiffiffiffi
tic material behavior. The remaining three types of and the volumetric strain tensor jemm j ¼ emm enn , respec-
mechanical behavior are included as limit cases. Depend- tively. The material parameters E1 and a of the uniaxial mate-
ing on the particular experimental findings for different rial model find their multiaxial counterparts in the shear and
materials, certain branches of the model can be neglected bulk moduli G1 ; K 1 as well as the parameters aG and aK sum-
or substituted by other formulations. In Section 4 the marized in Table 2. Other small strain formulations of consti-
material model will be used to simulate the material tutively nonlinear elasticity were for instance proposed by
behavior of polypropylene. Hartmann (2006) and Kletschkowski et al. (2002a, 2005) to
model the mechanical material behavior of polyoxymethy-
len and polytetrafluoroethylene, respectively.
3.2. Multiaxial generalization
The formulation of the multiaxial endochronic model
starts from the generalization of the addititve decomposi-
In the following, the uniaxial model defined in the pre-
tion of the uniaxial strain (5) to the symmetric strain
vious section will be generalized to the case of multiaxial
tensor
loadings. To this end, isotropic material behavior is as-
sumed. The anticipated isotropy has been verified by ten- eij ¼ eeij þ qend ð20Þ
ij
sile tests of specimens which were cut from the cast
material plates under different orientations (Kästner, The elastic and inelastic component can each be split into
2009). deviatoric and volumetric components according to Eqs.
48 M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57

(15) and (16). Then, an isotropic generalization of the uni- for the isothermal case. It states that the specific internal
axial stress–strain relation (4) is given by dissipation power d which is equal to the difference of
    the total stress power and the change of the
 free energy  f
rend ¼ 2G2 eij  qend;dev þ K 2 emm  qend
mm dij ð21Þ should not be negative. The free energy f eij ; qend ov;k
ij ; qij of
ij ij
the proposed viscoplastic material model depends on the
with shear G2 and bulk modulus K 2 . The generalization of
strain and the internal constitutive variables. It can be
the evolution Eq. (6) reads
additively decomposed
h     i
q_ end
ij ¼ bG eij  qend;dev
ij þ bK emm  qend
mm dij z
_ ð22Þ     X
nv  
pffiffiffiffiffiffiffiffiffi
f ¼ fe eij þ f end eij ; qend
ij þ f ov;k eij ; qov;k
ij ð29Þ
whereat the rate of the generalized arclength z_ ¼ e_ ij e_ ij is k¼1

computed from the norm jje_ ij jj of the strain rate tensor. The
into the contributions of the individual model branches.
material parameters bG and bK define the evolution of the
Using the relations
deviatoric and volumetric plastic strains in the endochron-
ic model, respectively. For the numerical implementation, @f e
reij ¼ ð30Þ
Eqs. (21) and (22) are combined in order to obtain the dif- @ eij
ferential equations @f end @f end
rend
ij ¼ ¼  end ð31Þ
s_ end
ij ¼ 2G2 e_ ij  bG send
ij z
_ ð23Þ @ eij @qij
r_ end _ end _
mm ¼ 3K 2 emm  3bK rmm z ð24Þ @f ov;k @f ov;k
rov;k
ij ¼ ¼  ov;k ð32Þ
@ eij @qij
for the deviator and trace of rend
ij which are the basis for the
plastic stress update algorithm.
between the stresses and free energies and the fact, that
The generalization of the nonlinear viscoelastic model is
the Clausius–Duhem inequality has to be fulfilled for arbi-
performed in the same way as for the endochronic model
trary loadings, Eq. (28) reduces to
and yields two differential equations
X
nv
2Gov rend _ end þ
ij qij rov;k _ ov;k P 0
ij qij ð33Þ
s_ ov;k
ij ¼ 2Gov _
k eij  Gk sov;k ð25Þ
g~ k ij k¼1

ov 3K ov Considering the evolution equations proposed for the


r_ ov;k
mm ¼ 3K k emm 
_ k
rov;k ð26Þ
g~ Kk mm internal variables, the dissipation power of the inelastic
for deviator and trace of the overstress rov;k in the kth non- models (33) yields
ij
linear viscoelastic element. Again Gov ov
k and K k are the elas- bG end end b
s s z_ þ K rend rend z_
tic shear and bulk moduli of each Maxwell element. The G2 ij ij K 2 mm nn
multiaxial versions of the process dependent viscosity Xnv 
ov;k 1 ov;k ov;k 1
functions read þ sov;k
ij s ij þ r mm r nn P0 ð34Þ
k¼1
g~ Gk g~ Kk
k0 k0
krov
kl k krov
kl k
g~ Gi ¼ gGi e s0 ~ Ki ¼ gKi e
and g s0
ð27Þ As the scalar products of the stress deviators, the products of
their traces, the rate of the generalized arc length and the
with the constant viscosities gGi and gKi for the deviatoric
overstress dependent part of the viscosity functions are al-
and volumetric response of each Maxwell element. As the
ways nonnegative, the choice gGi > 0; gKi > 0; bG > 0; bK > 0
viscosity
  functions ffi g ~ Ki and g
~ Gi are each related to the norm
 ov  qffiffiffiffiffiffiffiffiffiffiffiffiffiffi and G2 > 0; K 2 > 0 will be a sufficient condition for thermo-
rij  ¼ rov ij rov
ij of the overstress, Eqs. (25) and (26) result
dynamic consistency of the material model.
in a coupled system of ordinary nonlinear first order differ-
ential equations.
In this multiaxial generalization inelastic deformations 3.3. Numerical implementation
are allowed for the deviatoric and volumetric parts. The
common assumption of incompressible viscous and plastic In order to use the material model in a nonlinear finite
material behavior can subsequently be enforced by the element analysis, the multiaxial constitutive equations
appropriate choice of the material parameters. However, have been discretized with respect to time. This defines
in Section 4.3 the material parameters of the multiaxial the stress algorithm for the update of the internal vari-
model will be computed from the uniaxial ones (Table 2) ables. An algorithmic consistent material tangent stiffness
by the assumption of a constant Poisson’s ratio and the has subsequently been derived. The implementation of the
constraint that the uniaxial constitutive relations are numerical constitutive model into the finite element pro-
recovered. This results in Eqs. (65)–(68) which directly re- gram MSC.Marc was realized using user subroutines and
late the corresponding parameters. can be adopted to other finite element codes.
To conclude the formulation of the material model, its Applying a finite element discretization to a weak form
thermodynamic consistency is to be checked by the exam- of boundary value problems involving any constitutive and
ination of the Clausius–Duhem inequality geometric nonlinearities yields a system of nonlinear alge-
braic equations
1
d¼ rij e_ ij  f_ P 0 ð28Þ g ¼ fe  fi ¼ 0 ð35Þ
q
M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57 49

given in the typical vector–matrix notation. With respect and the derivation of the material tangent stiffness are re-
to the mechanical field problem, f e is the vector of external quired (Simo and Hughes, 1997). For convenience, the rela-
nodal loads. The vector of internal loads f i results from the tions will be given in tensor notation and the iteration
stresses in the finite elements. Generally, both vectors will index i will be dropped for clarity. Generally, quantities
change with the deformation expressed by the vector of from the previous iteration step i  1 will be used to com-
nodal degrees of freedom u. The solution of the discrete pute the algorithmic tangent stiffness while for the stress
global equilibrium Eq. (35) over a time interval ½0; t is update the new strain increment for iteration i is available.
accomplished by an incremental procedure associated Due to the additive decomposition of the total stress
with a partition of the analysis into subintervals ½t n ; t nþ1  (18) and the split into deviatoric and volumetric compo-
with t nþ1 ¼ t n þ Dt. For each time increment Dt it is as- nents according to (17), the relations
sumed that the equilibrium is satisfied for tn and hence
Xnv @ t nþ1 rov;m
@ tnþ1 rij @ nþ1 rij @ nþ1 rij
t e t end
all quantities are known for this point of time. In a stan- ij
¼ t þ t þ ð40Þ
dard finite element framework the new values at t nþ1 are @ nþ1 ekl @ nþ1 ekl
t @ nþ1 ekl m¼1
@ t nþ1 e
kl
obtained from an iterative solution procedure by a Newton ðÞ ðÞ
@ tnþ1 rij @ tnþ1 sij 1 @ tnþ1 rmm
ðÞ
method. The linearization ¼ þ dij ð41Þ

@ tnþ1 ekl @ tnþ1 ekl 3 @ tnþ1 ekl
t  t  @g

g nþ1 i
u ¼g nþ1
u i1
þ

tnþ1 i
t nþ1
dui hold for the tangent stiffness and allow for a modular
@u tnþ1 ui1  K
|fflfflfflfflfflffl{zfflfflfflfflfflffl} structure of the implementation. In the following, the con-
 2  tributions of all model branches to the stress and tangent
þ O tnþ1 dui ¼0 ð36Þ
stiffness will be specified.
The current stress in the nonlinear elastic branch of the
of the nonlinear system of equations at the previous itera-
material model for the equilibrium stress is directly ob-
tion step i  1 yields the tangent stiffness matrix tnþ1 Ki .
tained from the stress–strain relation since the current
Assuming that nonlinearities are only due to the material
strain tnþ1 eij is known from the global Newton iteration.
behaviour, it is obtained from a linearization of the consti-
To this end, Eq. (19) is evaluated separately for the devia-
tutive equations according to
toric and volumetric fractions

ne Z
[

t nþ1 @f i

T @ r

Ki ¼
¼ B BdV ð37Þ 2G1
@u tnþ1 ui1 j¼1 Xe;j @ e
tnþ1 ui1 t nþ1 e
sij ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi t nþ1
eij ð42Þ
|fflfflfflfflfflffl{zfflfflfflfflfflffl} 1 þ aG t nþ1 e t nþ1 e
kl kl
tnþ1
Ci
t nþ1 3K 1
remm ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi t nþ1
epp ð43Þ
1 þ aK t nþ1 emm nþ1 enn
t
where tnþ1 Ci is known as material tangent stiffness which is
evaluated at each integration point. Multiplication of tnþ1 Ci Using the partial derivatives with respect to the symmetric
by the matrix B containing partial derivatives of the finite strain tensor
element shape functions and integration over the element
domain Xe yields the element tangent stiffness matrices. @ tnþ1 eij 1   1
¼ dik djl þ dil djk  dij dkl ¼ wijkl ð44Þ
The tangent stiffness of the complete system follows from @ tnþ1 ekl 2 3
the assembly of all ne elements in the mesh, indicated by @ tnþ1 enn 1
Se ¼ ðdnk dnl þ dnl dnk Þ ¼ dkl ð45Þ
the operator nj¼1 . The matrix tnþ1 Ki is then used to predict @ tnþ1 ekl 2
the deviation tnþ1 dui of the displacement vector for the cur-
the contribution of the nonlinear elastic model
rent iteration step i
t  @ tnþ1 Dseij 2G1 wijkl
t nþ1
Kitnþ1 dui ¼ tnþ1 f e  f i nþ1
ui1 ð38Þ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi  
@ tnþ1 Dekl 1 þ aG tnþ1 emn tnþ1 emn
The total displacement increment after i iterations is ob- 2G1 aG t nþ1
eij tnþ1 ekl
P
tained from tnþ1 Dui ¼ ik¼1 tnþ1 duk . In the next step the equi-  pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2 pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
t
ð46Þ
1 þ aG tnþ1 emn tnþ1 emn nþ1 emn tnþ1 emn
librium is checked for the new displacement state. To this
end, a new vector of internal loads
@ tnþ1 Drepp 3K 1 dkl
ne Z
[ ¼ ð47Þ
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi 2
t nþ1 i
fi ¼ BT tnþ1 ri dV ð39Þ @ tnþ1 Dekl 1 þ aK tnþ1 emm tnþ1 enn
j¼1 Xe;j

to the algorithmic tangent stiffness can be obtained.


is computed from the current state tnþ1 ri at each integra-
The application of the midpoint rule
tion point which necessitates a suitable update procedure
t nþ1
for all internal variables known as stress update algorithm. y  t y tnþ1 Dy
_t
yj  ¼ ð48Þ
This local procedure is completely strain driven, as the nþ1
2 Dt Dt
strain increment tnþ1 ei in each integration point is known 1
from the global iteration step. The iteration is stopped yjt  t y þ tnþ1 Dy ð49Þ
nþ1
2 2
when some stop criterion is fulfilled.
Regarding the finite element implementation of the to the time discretization of the the evolution Eqs. (23)–
proposed material model, the time discretization of the (26) of the endochronic and the viscoelastic model branch
constitutive equations defining the stress update algorithm yields equations of the form
50 M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57

_t
yj _ yÞjt
¼ r ðx; x; ð50Þ represents a nonlinear system of equations which is a con-
nþ1 nþ1
2 2
sequence of the overstress dependent viscosity functions
and ensures a second-order accurate and stable computa- 0 P t  1
 nv nþ12 ov;j k0
tion of the stress increments tnþ1 Drend and tnþ1 Drov;k
ij . Algo-
t   r  C
ij
nþ1 B j¼1 kl
rithmic consistent material tangent stiffnesses are g~ Gi 2 rov
kl ¼ gGi exp @ A ð60Þ
s0
subsequently obtained from the linearization
0 P t  1
@ tnþ1 rij
ðÞ ðÞ
@ tnþ1 Drij  nv nþ12 ov;j k0
t 
B  j¼1 rkl  C
¼ ð51Þ g~ Ki nþ1
rov ¼ gKi exp @ A ð61Þ
@ tnþ1 ekl @ tnþ1 Dekl 2
kl
s0
of these stress update algorithms, where tn rij ¼ const: and
tn
eij ¼ const: have been used. It is solved for the deviatoric and volumetric components
The application of the implicit midpoint rule to the dif- of the current overstress increment tnþ1 Drov ij by an addi-
ferential Eqs. (23) and (24) defining the endochronic model tional Newton iteration at each integration point. The re-
yields quired contribution of the overstress model to the
t nþ1
t 1 algorithmic material tangent stiffness is obtained from
Dsend
ij ¼ 2G2 tnþ1 Deij  bG nþ2 send
ij
t nþ1
Dz ð52Þ
the partial derivatives of the overstress increment with re-
t
t nþ1 end tnþ1 nþ1 end t nþ1
Dr pp ¼ 3K 2 Demm  3bK 2 rmm Dz ð53Þ spect to tnþ1 Deij which in turn results in a linear system of
@ rov;m
ij
with the increment of the generalized arclength equations for the computation of @ Dekl
formed by
t nþ1
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi "
Dz ¼ tnþ1 Deij tnþ1 Deij . Due to the linear evolution equa-
@ tnþ1 Dsov;m
ij ov 2Gov
m Dt
t ov;m
1 @ nþ1 Dsij
tions, the discretized version of the endochronic model ¼ 2G w
m ijkl  t 1
 þ 
@ Dekl
t nþ1
g~ mG nþ2 rov 2 @ Dekl
t nþ1
can be explicitly solved for the new stress increments rs
#
k0 tnþ1 ov;m tnþ1 ov tnþ1 ov  20 1 tnþ1 ov @ tnþ1 Drov
k

t nþ1
2G2 tnþ1 Deij  bG tn send
ij
t nþ1
Dz þ 2s 2r 2r 2r
op
Dsend
ij ¼ ð54Þ 2s0 ij rs rs op t nþ1
@ Dekl
1 þ 12 bG tnþ1 Dz
ð62Þ
t nþ1 end 3K 2 tnþ1 Demm  3bK tn rend mm
tnþ1
Dz
Dr pp ¼ ð55Þ and
1 þ 32 bK tnþ1 Dz
tnþ1 ov;m
@ tnþ1 Drov;m 3K ov Dt 1 @ Drpp
The contributions of the deviatoric and volumetric compo-
nn
¼ 3K ov
m dkl 
t m  þ  
@ Dekl
t nþ1
g~ m nþ2 rov 2 @ nþ1 Dekl
1 t
K rs
nents to the algorithmic tangent stiffness of the endocronic #
model are obtained from the differentiation of the discret- k0 tnþ1 t t
k20 1 @ tnþ1 Drov
nþ1 nþ1 tþ12Dt uv
ized evolution Eqs. (54) and (55) with respect to the cur-
þ 2 rov;m
qq
2 rov
rs
2r
ov
rs rov
uv
2s0 @ tnþ1 Dekl
rent strain increment
ð63Þ
 t
Deij tnþ1DDezkl
nþ1
@ tnþ1 Dsend 2G2 wijkl bG tn send
ij þ G2
tnþ1
The total algorithmic material tangent stiffness is obtained
ij
¼    ð56Þ from the sum of the contributions of the individual model
@ tnþ1 Dekl 1 þ 2 bG tnþ1 Dz
1
1 þ 12 bG tnþ1 Dz
2

  t De branches according to Eq. (40). For the numerical imple-


@ tnþ1 Drend
pp 3K 2 dkl 3bK tn rend 3
mm þ 2 K 2
tnþ1
emm nþ1 t nþ1
kl
Dz mentation the expressions for the stress increments and
¼    ð57Þ
@ tnþ1 Dekl 1 þ 32 bK tnþ1 Dz 3 t
1 þ bK nþ1 Dz
2
tangent operators have been transformed to the vector–
2
matrix notation usually applied in finite element codes.
It can be seen that the elastic moduli 2G2 wijkl and 3K 2 dkl are
reduced due to the influence of the stress state and the 4. Results and discussion
strain increment. While the first term in Eq. (56) is sym-
metric, the second one represent a nonsymmetric contri- The presented viscoplastic material model will now be
bution to the tangent stiffness matrix with respect to the applied to simulate the nonlinear material behavior of
permutation of the index pairs fijg and fklg. Finite element polypropylene. The following sections show how the mate-
codes in general provide solution procedures for nonsym- rial parameters can be identified using data obtained from
metric stiffness matrices. On the other hand, we have ob- tensile and relaxation tests as well as loading and unload-
served that accurate computational results are even ing experiments with intermediate holding times. In order
achieveable, if only the first summands in (56) and (57) to illustrate the capabilities of the formulated model, sev-
are used which prevents nonsymmetric stiffness matrices. eral test cases have been simulated and are compared with
In contrast to the endochronic approach, the time dis- the experimental data.
cretization of the overstress model
2Gov Dt tnþ1 ov;k 4.1. Identification of material parameters
t nþ1
Dsov;k
ij ¼ 2Gov
k
t nþ1
Deij  t k  2 sij ð58Þ
nþ1
g~ Gk 2 rov
ij The structured identification procedure starts with the
3K ov Dt tnþ1 ov;k parameters of the equilibrium relation. They are identified
t nþ1
Dr ov;k
pp ¼ 3K ov
k
t nþ1
e
D mm  t k  2 rmm ð59Þ from adopting the model prediction to the terminal points
nþ1
g~ Kk 2 rov
ij
of the intermediate relaxations. After the equilibrium
M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57 51

relation can be simulated with sufficient precision, the rate using the maximum strain rate emax ¼ 7:5  102 1/s, the
dependent effects are quantified using relaxation minimal relaxation time smin ¼ 101 s. Therefore, the set
experiments. of relaxation times
From the experimental observations it was concluded n o
that the material shows a small equilibrium hysteresis. fsi g ¼ 101 ; 100 ; 101 ; 102 ; 103 ; 104 s
Therefore, at least the endochronic part of the model pro-
posed for the equilibrium stress will be required to repro- corresponding to a total number of six Maxwell elements
duce the observed material behavior. It can be adapted to in the overstress model has been chosen. The application
the terminal points of the intermediate relaxations in the of the window algorithm of Emri and Tschoegl to the
loading path. However, the hysteresis will be significantly experimental data yields the set of relaxation strengths
overestimated. In order to obtain a satisfying approxima- fci g ¼ f4:81; 55:17; 91:81; 88:52; 83:89; 70:52gMPa
tion of the equilibrium relation also the nonlinear elastic
model has to be used. By combining both models, the load- Using the linear viscoelastic model defined by the relaxa-
ing and unloading path of the equilibrium hysteresis can tion spectrum above together with the previously identi-
be captured (Fig. 13). The identified parameters are fied model for the equilibrium relation, the material
E1 ¼ 400:0 MPa, a ¼ 15:0; E2 ¼ 350:0 MPa and b ¼ 60:0. behavior in relaxation experiments and monotonic tensile
As mentioned above, the parameters of the strain rate tests has been simulated. From Fig. 14 it can be seen that,
dependent overstress model are quantified using data from as expected, the relaxation test used for identification can
relaxation experiments. To begin with the identification be reproduced by the simulation. Both major characteris-
procedure, linear viscoelastic material behavior is as- tics of the material behavior – the equilibrium stress state
sumed. Several techniques exist for the determination of at the end of the relaxation and the time dependent stress
the discrete spectrum defined by the relaxation strengths relaxation – are accurately covered by the material model.
ci and the associated relaxation times si . Here, the straight In contrast, the prediction of the material behavior in
forward window algorithm of Emri and Tschoegl (Emri and monotonic tensile tests performed at constant strain rates
Tschoegl, 1993, 1994, 1995; Tschoegl and Emri, 1992) is (Fig. 15) reveals two major shortcomings of the identifica-
applied to iteratively compute the relaxation strengths tion procedure and the material model. Although the stress
for a given set of relaxation times using only the most sen- level at e ¼ 4:5% is appropriate, the material model fails to
sitive subset of data. display the instantaneous modulus as well as the nonlin-
Defining a set of relaxation times corresponds to the earity of the stress–strain curve. While the first drawback
specification of the number of Maxwell elements in the is strongly related to the practical realization of the relax-
overstress model. Commonly one relaxation time and ation experiment and the missing information for relaxa-
hence one Maxwell element per decade is used. The mini- tion times s 6 101 s, the latter is due to the nonlinear
mum and maximum relaxation times are determined by viscoelastic properties of the investigated material.
the data available from the relaxation experiment. During In order to solve these problems a modified version
the experimental characterization procedure relaxation (Kästner, 2009) of the Emri-Tschoegl window algorithm
experiments with holding times of 10 to 48 h have been was applied to use data from tensile tests for the identifi-
performed. Since the maximum relaxation time must not cation of the relaxation spectrum. The overstress model
exceed the holding time, smax ¼ 104 s has been chosen. was additionally extended to nonlinear viscoelastic mate-
On the other hand, a lower boundary for the time spectrum rial behavior by using the viscosity function (13). However,
is given by the resolution of the measured stress–time since the properties fci ; si g have lost their clear interpreta-
curve and first of all by the fact that in any real relaxation tion due to the shift of relaxation times (14) caused by the
experiment no ideal jump in the strain level can be applied. variable viscosity, the set of parameters of the nonlinear
As approximately 0.5 s are needed for the load application viscoelastic overstress model fci ; si ; k0 ; s0 g has to be identi-

Experiment
20
30 Model
Stress σ [MPa]
Stress σ [MPa]

10
20

0
10
Experiment
-10 Model eq
0
0 0.01 0.02 0.03 0.04 0.05 0 0.5 1 1.5 2 2.5
4
Strain ε [−] Time t [s] x 10

Fig. 13. Identification of the material parameters of the equilibrium Fig. 14. Identification of the material parameters of the overstress model
relation from the terminal points of intermediate relaxations. from relaxation experiments.
52 M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57

n o
fsi g ¼ 101 ; 100 ; 101 ; 102 ; 103 ; 104 s
30
fci g ¼ f62:18; 107:54; 133:27; 288:59; 283:88; 173:51gMPa
Stress σ [MPa]

a significantly improved approximation for the monotonic


20
tensile tests can be achieved (Fig. 17).

100 mm/min 4.2. Simulation of the material behavior – uniaxial loading


10
10 mm/min
In the previous section the parameters of the elastoplas-
1 mm/min tic equilibrium relation have been determined. Subse-
0 quently, the linear viscoelastic properties of the
0 0.01 0.02 0.03 0.04 overstress model were computed using information from
Strain ε [−] relaxation experiments. As the prediction of monotonic
tensile tests was not satisfactory, the model was extended
Fig. 15. Prediction of the strain rate dependence for monotonic tensile
tests using linear viscoelasticity and the parameters identified from the
to nonlinear viscoelastic material behavior. Since this
relaxation experiments (symbols: experimental results, solid lines: model extension is related to a shift in the relaxation times, the
prediction). influence on the simulation of relaxation tests has to be
verified. From Fig. 18 it can be concluded that the shift of
the relaxation times has only a minor, acceptable influence
fied iteratively. To this end the influence of the parameter on the simulation of the relaxation experiments which can
s0 of the viscosity function has been studied. From Fig. 16 it still be recovered with sufficient accuracy.
can be found that suitable values for the parameter can be The material model will now be applied to loading and
expected in the interval s0 2 ½1:0; 10:0. A refined analysis unloading experiments with intermediate relaxations and
in this region delivered an optimal value of s0 ¼ 5:2. To- cyclic loadings. Fig. 19 shows the comparison of experi-
gether with k0 ¼ 1:25 and the spectrum mentally and numerically determined stress–strain curves
for a loading and unloading experiment with intermediate
relaxations whereat the loading and unloading was accom-
(a) plished at a constant velocity of 10 mm/min. Except for the
final part of the unloading the agreement of experimental
30
and numerical results is very good as both – the terminal
points of relaxation as well as the reloading and unloading
Stress σ [MPa]

are accurately predicted.


20
In contrast to the simulation results obtained for relax-
ation tests and experiments with intermediate relaxations,
10 100 mm/min the mechanical behavior under cyclic loading is captured
10 mm/min only qualitatively. With an increasing number of cycles
the experimental (Fig. 10) and numerical (Fig. 20) stress–
1 mm/min strain curves both show a decreasing stress level in the
0 reversal points. However, the material model overesti-
0 0.01 0.02 0.03 0.04
Strain ε [-]

(b) 30
30
Stress σ [MPa]
Stress σ [MPa]

20
20
100 mm/min
100 mm/min 10
10
10 mm/min
10 mm/min
1 mm/min
1 mm/min 0
0 0 0.01 0.02 0.03 0.04
0 0.01 0.02 0.03 0.04 Strain ε [−]
Strain ε [−]
Fig. 17. Prediction of the strain rate dependent material behavior for
Fig. 16. Influence of the parameter s0 on the model prediction for monotonic tensile tests using the nonlinear viscosity function (14) and
monotonic tensile tests (k0 ¼ 1:25, symbols: experimental results, solid the modified spectrum (symbols: experimental results, solid lines: model
lines: model prediction) – (a) s0 ¼ 10:0, (b) s0 ¼ 1:0. prediction).
M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57 53

was used to simulate the mechanical behavior of polyeth-


Experiment ylene (Drozdov and Christiansen, 2007a).
30
Model In summary, it can be concluded that the material mod-
el is able to account for the major phenomena observed in
Stress σ [MPa]

the experiment – the nonlinear strain rate dependence and


20
the equilibrium hysteresis. While the model prediction for
monotonic loadings is satisfactory, only qualitative agree-
ment was found for cyclic loadings. As the focus of the
10 application is on the simulation of the effective inelastic
material behavior of fibre reinforced polymers in tensile
and relaxation tests the quality of the material model is
0 sufficient. In addition to fact that the presented uniaxial
0 0.5 1 1.5 2 2.5
4 material model performs well for different load cases, it of-
Time t [s] x 10 fers a structured procedure for the identification of the
material parameters.
Fig. 18. Prediction of the relaxation behavior using the nonlinear
viscosity function (14) and the modified spectrum. In literature several alternatives for modeling the linear
and nonlinear viscoelastic material behavior have been
proposed. Due to the pronounced relaxation behavior six
parallel Maxwell elements are required to model the mate-
rial behavior in the considered time scale. This number
20 could be reduced by power law models (Haupt, 2002) or
the application of fractional viscoelasticity (Lion, 1997b;
Müller et al., 2011). Compared to the classic Maxwell ele-
Stress σ [MPa]

10 ment which predicts a transition from the glassy to the


equilibrium state over a period of two decades in the vicin-
0
ity of its relaxation time, the active time is significantly ex-
Experiment tended by the application of power law or fractional
Model models, respectively. This results in a reduced number of
-10 Model eq material parameters. On the other hand the nonlocal char-
acter of fractional models requires the storage of the com-
0 0.01 0.02 0.03 0.04 0.05 plete loading history.
Strain ε [−] Regarding the modeling of the nonlinear strain rate
dependence, an alternative could be seen in a single Max-
Fig. 19. Prediction of the material behavior (r ¼ req þ rov and req ) for
loading and unloading experiments with intermediate holding times. well model modified by a nonlinear viscosity function as
proposed in Krempl and Khan (2003). However, a single
nonlinear Maxwell element is not able to recover the pro-
nounced relaxation behavior of polypropylene but could be
30 considered for thermoset resins.
Another interesting approach uses an additional inter-
20 nal variable to account for the deformation induced
Stress σ [MPa]

changes in the microscopic polymer network. From a phe-


10
nomenological point of view this results in process depen-
dent structural viscosities (Barnes, 1997). An appropriate
model was applied by Haupt and Sedlan (2001) to simulate
0
the material behavior of elastomers. Changes in the poly-
mer network are modelled by the so called structural var-
-10
iable qs which is used to control the relaxation time s ~ðqs Þ
s
or viscosity gðq Þ of a Maxwell element. Its temporal evolu-
~
0 0.01 0.02 0.03 0.04 0.05 tion is governed by
Strain ε [−]
1
Fig. 20. Prediction of the material behavior for cyclic loadings, compa- q_ s ¼ cs je_ jð1  qs Þ  qs ð64Þ
sq
rable experimental result are shown in Fig. 10.

The first term in the evolution Eq. (64) describes the grad-
ual deterioration of the polymer network by increasing val-
mates the hysteresis. Better predictions for cyclic loadings ues of the structural variable qs 6 1. The second term
of polypropylene have been reported by Drozdov and accounts for a recovery of cross links between the individ-
Christiansen, 2007b. Nevertheless, the proposed pseudo- ual polymer chains. The time scale for recovery is given by
elastic material model involves different sets of material sq . If sq is significantly larger the time of the time scale of
parameters for loading and unloading and a rate depen- the considered experiment, no recovery can be observed
dent scaling of the material properties. A similar approach and hence Eq. (64) describes a deformation dependent
54 M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57

evolution of the structural variable. Such a model could re- Presuming a constant Poisson’s ratio m, the material
place the overstress dependence of the viscosity function. parameters of the multiaxial model can be computed from
the properties of the uniaxial model. The values
4.3. Simulation of material behavior – multiaxial loading E1 E1
K1 ¼ ; G1 ¼
3ð1  2mÞ 2ð 1 þ m Þ
After the successful modeling of the material behavior E2 E2
in uniaxial tensile tests and relaxation experiments, the K2 ¼ ; G2 ¼ ð65Þ
3ð1  2mÞ 2ð 1 þ m Þ
material parameters for the multiaxial generalization have ci ci
to be determined. To this end, the multiaxial strain state in K ov
i ¼ ; Gov ¼
3ð1  2mÞ i 2ð1 þ mÞ
the specimen has been analyzed by digital image correla-
tion for some monotonic tensile tests and relaxation exper- are obtained from the elastic constants E1 and E2 as well as
iments. The evaluation of the longitudinal and transversal the relaxation strengths ci . All other parameters are calcu-
strain for tensile tests with two different velocities indi- lated from the requirement that the multiaxial model has
cates a Poisson’s ratio that increases from m ¼ 0:2 to to predict the same results for uniaxial experiments as
m ¼ 0:4 with the applied longitudinal strain e (Fig. 21(a)). the original uniaxial formulation. This constraint yields
On the other hand, the analysis of relaxation experiments the following relations
in Fig. 21(b) shows an approximately constant value of rffiffiffi
m ¼ 0:43. a 3 a
aK ¼ ; aG ¼ ð66Þ
However, as the extent of the digital image correlation 1  2m 21 þ m
measurements is very limited and since no experimental
b b
results for pure hydrostatic or deviatoric loadings are cur- bK ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ; bG ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð67Þ
rently available, assumptions are required to define the 3 1 þ 2m2 1 þ 2m 2
material parameters of the multiaxial model from tests
gi gi
with uniaxial stress states. Until further significant experi- gKi ¼ ; gGi ¼ ð68Þ
1  2m 1þm
mental results have been obtained from combined ten-
sion–torsion loadings, we will use a constant Poisson’s between the material parameters of the uniaxial and the
ratio of m ¼ 0:4 representing a simplification of the multi- multiaxial material model (compare Table 2).
axial material model. Similar assumptions are found in The following results for multiaxial loadings are ob-
Kolarik and Pegoretti (2006); Tschoegl et al. (2002). tained from the finite element analysis of a single

(a) (a) 0.06

0.4
ν [−]

0.04
Strain [-]

0.3

0.02
0.2
100mm/min
10mm/min ,
0.1 0
0 0.05 0.1 0.15 0 500 1000 1500 2000
Strain ε [−] Time t [s]

(b) 0.5 (b) 30


pure tension
ν [ -]

Stress [MPa]

0.4 20
combined tension-shear
pure shear
10
0.3
σ
τ
0
0.2 0 500 1000 1500 2000
0 500 1000 1500 2000 Time t [s]
Time t [s]
Fig. 22. Simulation of combined tension and shear relaxation tests – (a)
Fig. 21. Experimental investigation of Poisson’s ratio – (a) monotonic strain–time curve and (b) comparison of stress–time curves of combined
tensile tests and (b) relaxation experiments at e ¼ 4:5%. tension and shear to pure tension and shear tests.
M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57 55

hexagonal trilinear finite element. The material behavior is


(a) 0.06
modelled by the implemented user subroutine containing
the stress update algorithm and the algorithmic tangent
stiffness derived from the presented material model as

Shear strain γ [-]


0.04
outlined in Section 3.3. The respective displacement
boundary conditions are chosen to apply different combi-
nations of tensile e and shear strain c and to allow for free
lateral contraction. 0.02
Fig. 22 shows the applied strain history and the com-
puted stress–time curves for the simulation of a combined
tension and shear relaxation test. The loading is performed 0
at a constant strain rate e_ ¼ 1:5  103 1/s. The same value 0 0.02 0.04 0.06
is used for the rate c_ ¼ 1:5  103 1/s of the shear strain. Tensile strain ε [−]
Since no multiaxial experiments have been performed in
this study, the numerical results can only be checked for (b) 15
plausibility. From the comparison of the results obtained
for the combined loading to the simulation of pure tension

Shear stress τ [MPa]


10
and shear relaxation tests, it can be seen that the stress le-
vel is reduced due to the multiaxial overstress state which 5
leads to smaller viscosities. In addition to the lower stress
levels, the overstress relaxes faster as the relaxation spec- 0
trum is shifted to shorter relaxation times. This effect can-
not be observed in linear viscoelastic material models. -5
In the next numerical experiment shown in Fig. 23 the
tensile and shear stresses are studied for a sudden change -10
of the strain rates. Experimental results for polypropylene -20 -10 0 10 20 30
under similar loading conditions have been obtained by Tensile stress σ [MPa]
Kitagawa and Takagi (1990) and can be used to verify the
Fig. 24. Simulation of combined cyclic tension and shear loading – (a)
e  c curve and (b) r  s curve.

(a) 0.06
model predictions. The tensile strain rate is reduced from
e_ ¼ 1:5  103 1/s to e_ ¼ 1:5  104 1/s when a strain of
0.04 e ¼ 2:5% is reached. At the same time the shear strain rate
is increased from c_ ¼ 1:5  104 1/s to c_ ¼ 1:5  103 1/s.
Strain [-]

Although the strain e is still extended, the tensile stress re-


laxes due to the reduced strain rate. On the other hand the
0.02
ε shear stress is very low at the beginning of the loading and
increases strongly after the jump of the shear strain rate.
γ These simulated stress–time curves are in good agreement
0 with experimental results presented in Kitagawa and Tak-
0 20 40 60 agi (1990) for combined loading of polypropylene speci-
Time t [s] mens. As both – the type of polypropylene and the
applied strain rates – are not identical, the numerical re-
(b) 30 sults can only be assessed qualitatively.
Finally, the material behavior is simulated for cyclic mul-
tiaxial loadings. To this end, a rectangular load path is ap-
plied in the e  c plane (Fig. 24(a)). Due to the relaxation
Stress [MPa]

20
σ of the stresses that occurs when the respective strain is kept
τ constant, a distorted shape is obtained in the r  s plane.
Although the same shape is obtained for linear viscoelastic
10 material behavior, the amount of distortion is significantly
larger for the overstress dependent viscosity function.

0 5. Conclusions
0 20 40 60
Time t [s]
In order to classify and model the inelastic material
Fig. 23. Simulation of combined tension and shear loading according to
behavior, an experimental testing procedure has been pre-
(Kitagawa and Takagi, 1990) – (a) strain–time curve and (b) stress–time sented. It includes monotonic tensile tests, relaxation tests
curves. as well as more complex loading and unloading processes
56 M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57

with and without intermediate holding times. This proce- Barnes, H.A., 1997. Thixotropy – a review. Journal of Non-Newtonian Fluid
Mechanics 70 (1–2), 1–33.
dure was applied to experimentally characterize the
Bergström, J., Hilbert Jr., L., 2005. A constitutive model for predicting the
mechanical behavior of polypropylene. The experimental large deformation thermomechanical behavior of fluoropolymers.
observations indicate a viscoplastic material behavior with Mechanics of Materials 37 (8), 899–913.
a strong rate dependence and a small equilibrium hysteresis. Drozdov, A., 2009. Mullins effect in semicrystalline polymers.
International Journal of Solids and Structures 46 (18–19), 3336–3345.
Inline with the proposed experiments, a uniaxial visco- Drozdov, A.D., 1998. A model for the nonlinear viscoelastic response in
plastic material model based on an overstress formulation polymers at finite strains. International Journal of Solids and
has been derived, was subsequently generalized to multi- Structures 35 (18), 2315–2347.
Drozdov, A.D., Christiansen, J.d.C., 2007a. Cyclic viscoplasticity of high-
axial loadings and implemented into a finite element code. density polyethylene: experiments and modeling. Computational
The parameters of this material model can be determined Materials Science 39 (2), 465–480.
using the stress–time curves obtained from relaxation Drozdov, A.D., Christiansen, J.d.C., 2007b. Cyclic viscoplasticity of solid
polymers: the effects of strain rate and amplitude of deformation.
experiments and the stress–strain curves of loading and Polymer 48 (10), 3003–3012.
unloading processes with intermediate holding times. Emri, I., Tschoegl, N.W., 1993. Generating line spectra from experimental
To verify the formulated material model, a set of simu- responses. Part I: Relaxation modulus and creep compliance.
Rheologica Acta 32 (3), 311–322.
lations has been carried out and the results are compared Emri, I., Tschoegl, N.W., 1994. Generating line spectra from experimental
with experimental data. It is demonstrated that the model responses. Part IV: Application to experimental data. Rheologica Acta
is capable of simulating the long term (relaxation test) as 33 (1), 60–70.
Emri, I., Tschoegl, N.W., 1995. Determination of mechanical spectra from
well as the short term (tensile tests) rate dependent mate-
experimental responses. International Journal of Solids and Structures
rial behavior at different strain rates. The comparison to 32 (6–7), 817–826.
experiments with cyclic loading has shown that the model Govindjee, S., Simo, J.C., 1992. Mullins effect and the strain amplitude
overestimates the hysteresis and partially fails when it dependence of the storage modulus. International Journal of Solids
and Structures 29 (14–15), 1737–1751.
comes to pressure loading. Reasonable results are obtained Hartmann, S., 2006. A thermomechanically consistent constitutive model
for combined tensile and shear loading. for polyoxymethylene experiments, material modelling and
The development of the presented material model was computation. Archive of Applied Mechanics 76, 349–366.
Haupt, P., 1993. On the mathematical modelling of material behavior in
mainly motivated by the need to predict the effective continuum mechanics. Acta Mechanica 100 (3), 129–154.
material behavior of fibre-reinforced polypropylene in Haupt, P., 2002. Continuum Mechanics and Theory of Materials. second
terms of a multiscale approach. To this end, the authors ed.. Springer Verlag, Heidelberg, Berlin.
Haupt, P., Lion, A., 1995. Experimental identification and mathematical
have applied multiscale modeling techniques based on modeling of viscoplastic material behavior. Continuum Mechanics
the finite element simulation of a representative volume and Thermodynamics 7 (1), 73–96.
element to predict the macroscopic deformation behavior Haupt, P., Sedlan, K., 2001. Viscoplasticity of elastomeric materials:
experimental facts and constitutive modelling. Archive of Applied
of the composite material using only the constitutive prop- Mechanics 71 (2), 89–109.
erties of its constituents and the geometrical arrangement Heinrich, G., Kaliske, M., 1997. Theoretical and numerical formulation of a
of the reinforcing fibres on the micro- and mesoscale. molecular based constitutive tube-model of rubber elasticity.
Computational and Theoretical Polymer Science 7 (3–4), 227–241.
Effective linear elastic stiffness properties are outlined in
Holzapfel, G.A., 1996. On large strain viscoelasticity: continuum
Kästner et al. (2008). Using the viscoplastic material model formulation and finite element applications to elastomeric
presented in this article, the macroscopically nonlinear structures. International Journal for Numerical Methods in
material behavior of textile-reinforced polypropylene is Engineering 39 (22), 3903–3926.
Kaliske, M., Rothert, H., 1997. Formulation and implementation of three-
predicted in Kästner (2009, 2011) and compared to exper- dimensional viscoelasticity at small and finite strains. Computational
imental observations. Mechanics 19 (3), 228–239.
Future work regarding the material behavior of poly- Kästner, M., 2009. Skalenübergreifende modellierung und simulation des
mechanischen verhaltens von textilverstärktem polypropylen unter
mers will focus on multiaxial loadings whereat also ther- nutzung der xfem. Dissertation, TU Dresden <http://nbn-resolving.de/
momechanical loading conditions will be considered in urn:nbn:de:bsz:14-qucosa-27515>.
both – experiments and numerical models. Kästner, M., Haasemann, G., Brummund, J., Ulbricht, V., 2008.
Computation of Effective Stiffness Properties for Textile-reinforced
Composites Using X-FEM. Springer (Chapter 13).
Acknowledgements Kästner, M., Haasemann, G., Ulbricht, V., 2011. Multiscale xfem-modelling
and simulation of the inelastic material behaviour of textile-
reinforced polymers. International Journal of Numerical Methods
The present study is supported by the German Research and Engneering 86, 477–498.
Foundation (DFG) within the Collaborative Research Centre Kitagawa, M., Takagi, H., 1990. Nonlinear constitutive equation of
polypropylene and combined tension and torsion. Journal of
(SFB) 639, subproject C2. This support is gratefully
Materials Science 25, 2869–2872.
acknowledged. Kletschkowski, T., Schomburg, U., Bertram, A., 2002a. Endochronic
viscoplastic material models for filled ptfe. Mechanics of Materials
34 (12), 795–808.
References Kletschkowski, T., Schomburg, U., Bertram, A., 2002b. An inelastic
material model for filled polytetrafluorethylen. Archive of Applied
Arruda, E.M., Boyce, M.C., 1993a. Evolution of plastic anisotropy in Mechanics (Ingenieur Archiv) 72 (4), 293–299.
amorphous polymers during finite straining. International Journal of Kletschkowski, T., Schomburg, U., Bertram, A., 2004. An endochronic
Plasticity 9 (6), 697–720. viscoplastic approach for materials with different behavior in tension
Arruda, E.M., Boyce, M.C., 1993b. A three-dimensional constitutive model and compression. Mechanics of Time-Dependent Materials 8 (2),
for the large stretch behavior of rubber elastic materials. Journal of 119–135.
the Mechanics and Physics of Solids 41 (2), 389–412. Kletschkowski, T., Schomburg, U., Bertram, A., 2005. A rate-dependent
Bagley, R.L., Torvik, P.J., 1983. A theoretical basis for the application of endochronic approach to thermoplastic materials: temperature and
fractional calculus to viscoelasticity. Journal of Rheology 27 (3), 201– filler volume fraction dependence. Mechanics of Materials 37 (6),
210. 687–704.
M. Kästner et al. / Mechanics of Materials 52 (2012) 40–57 57

Kolarik, J., Pegoretti, A., 2006. Non-linear tensile creep of polypropylene: Reese, S., Govindjee, S., 1998. A theory of finite viscoelasticity and
time-strain superposition and creep prediction. Polymer 47 (1), 346– numerical aspects. International Journal of Solids and Structures 35
356. (26–27), 3455–3482.
Krempl, E., Khan, F., 2003. Rate (time)-dependent deformation behavior: Simo, J., Hughes, T., 1997. Computational Inelasticity. Interdisciplinary
an overview of some properties of metals and solid polymers. Applied Mathematics, vol. 7. Springer, New York, Heidelberg,
International Journal of Plasticity 19 (7), 1069–1095. Berlin.
Lion, A., 1996. A constitutive model for carbon black filled rubber: Simo, J.C., 1987. On a fully three-dimensional finite-strain viscoelastic
experimental investigations and mathematical representation. damage model: formulation and computational aspects. Computer
Continuum Mechanics and Thermodynamics 8 (3), 153–169. Methods in Applied Mechanics and Engineering 60 (2), 153–173.
Lion, A., 1997a. On the large deformation behaviour of reinforced rubber Tschoegl, N., 1989. The Phenomenological Theory of Linear Viscoelastic
at different temperatures. Journal of the Mechanics and Physics of Behavior. Springer.
Solids 45 (11–12), 1805–1834. Tschoegl, N., Knauss, W.G., Emri, I., 2002. Poisson’s ratio in linear
Lion, A., 1997b. On the thermodynamics of fractional damping elements. viscoelasticity - a critical review. Mechanics of Time-Dependent
Continuum Mechanics and Thermodynamics 9 (2), 83–96. Materials 6 (1), 3–51.
Lion, A., 1997c. A physically based method to represent the thermo- Tschoegl, N.W., Emri, I., 1992. Generating line spectra from experimental
mechanical behaviour of elastomers. Acta Mechanica 123 (1), 1–25. responses. III. Interconversion between relaxation and retardation
Miehe, C., Keck, J., 2000. Superimposed finite elastic–viscoelastic– behavior. International Journal of Polymeric Materials 18 (1), 117–
plastoelastic stress response with damage in filled rubbery 127.
polymers. Experiments, modelling and algorithmic implementation. Valanis, K., 1971a. A theory of viscoplasticity without a yield surface. Part
Journal of the Mechanics and Physics of Solids 48 (2), 323–365. I – General theory. Archives of Mechanics 23, 517–533.
Müller, S., Kästner, M., Brummund, J., Ulbricht, V., 2011. A nonlinear Valanis, K.C., 1971b. A theory of viscoplasticity without a yield surface.
fractional viscoelastic material model for polymers. Computational Part II – Application to mechanical behavior of metals. Archives of
Materials Science 50 (10), 2938–2949. Mechanics 23, 535–551.

You might also like