Sie sind auf Seite 1von 10

Full Paper

Predicting the Long-Term Mechanical


Performance of Polycarbonate from Thermal
History during Injection Molding
Tom A. P. Engels, Lambert C. A. van Breemen, Leon E. Govaert,*
Han E. H. Meijer

applied load [kN]

The influence of the thermal history experienced during injection molding on the mechanical properties of polycarbonate is investigated. Distributions of the yield stress as they
result from inhomogeneous cooling during processing, predicted by a previously developed
50
modeling approach, are validated and are in
good agreement with experiment. Predictions
of the mechanical performance for different
40
mold temperatures during processing are also
validated over a range of applied strain and
30
deformation rates and applied stresses and
forces, for simple tensile bars and an actual
20
(open markers)
(closed markers)
product. Good agreement is found. It is shown
that mold temperature has a tremendous influ10
Tm = 130C
ence on the life time of polymer products,
Tm = 30C
model prediction
which can differ by more than two orders of
0 0
1
2
3
4
5
6
7
magnitude.
10
10
10
10
10
10
10
10
timetofailure [s]

Introduction
Catastrophic failure of polymer objects, either upon impact
(e.g., of protective products such as airbags and helmets) or
after prolonged exposure to load (for instance supporting
structures, high-pressure pipes), limits their ultimate useful
lifetime. Hence, understanding of that process and, ideally,
being able to accurately predict when failure occurs is of
critical importance. This reflects on the selection of the
materials employed, the choice of the proper processing
parameters and the selection of the geometrical design of
T. A. P. Engels, L. C. A. van Breemen, L. E. Govaert, H. E. H. Meijer
Materials Technology (MaTe), Eindhoven University of
Technology, P.O. Box 513, NL-5600 MB, Eindhoven,
The Netherlands
E-mail: l.e.govaert@tue.nl
Macromol. Mater. Eng. 2009, 294, 829838
2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

the product. The problem is complex, since even for simple


single-phase amorphous polymers the final mechanical
properties are generally inhomogeneous due to nonuniform cooling, leading to differences in thermodynamic
state of the material throughout a product.
Modern engineering environments usually integrate
design and manufacturing where all the processes involved
are supported by numerical tools that aid distinct steps in
the total design procedure. Most prominent are the
numerical codes that allow simulation of the processing
of the product (e.g., injection molding), and those developed
for mechanical analysis of the final product under the
desired loading conditions. Until now, there has been no
real interaction between these two fields. This is a serious
drawback, since it is the processing step which largely
determines the mechanical properties of injection molded
products.

DOI: 10.1002/mame.200900227

829

T. A. P. Engels, L. C. A. van Breemen, L. E. Govaert, H. E. H. Meijer

An excellent example are semi-crystalline polymers


where the temperature and flow history experienced by the
material during molding result in an inhomogeneous,
anisotropic crystalline morphology.[1,2] As a result, samples
taken from an injection molded product of a semi-crystalline polymer like polyethylene, display different failure
behavior dependent on position and orientation, e.g., tough
parallel to flow and brittle in perpendicular direction.[3,4]
For amorphous polymers the effects of flow-induced
orientation during injection molding is much less pronounced. Frozen-in molecular orientation causes some
anisotropy,[5,6] but its influence on the yield stress can
be considered small with respect to its absolute value.[7] On
the other hand, the thermal history experienced upon
solidification from the melt has a marked influence on the
mechanical properties of molded polymer glasses. Determining for the evolution of properties is the kinematic
nature of the glass transition, Tg, and the inherent nonequilibrium state below Tg.[8,9] Differences in cooling rate
have large consequences for the yield stress of the product,
and its time-to-failure under static or dynamic load.[1012]
Another important issue is that the change in yield stress
may lead to a change in failure mode from ductile to
brittle.[13,14]
With respect to mechanical analysis, considerable effort
has been directed toward the numerical simulation of
deformation and failure of glassy polymers. A number of 3D
constitutive models were developed and validated, e.g., in
the group of Mary Boyce at MIT,[1517] the group of Paul
Buckley in Oxford,[1820] and in our Eindhoven group.[2123]
These developments enabled a quantitative analysis and
prediction of short- and long-term failure (static and
dynamic),[10,11] and demonstrated the intimate relation
between the value of the yield stress and the lifetime in
long-term static loading. In our Eindhoven Glassy Polymer
model (the EGP-model),[23] changes in the value of the yield
stress, resulting from alterations in the thermodynamic
state, can be accommodated by adapting only a single
parameter Sa. Since the other parameters required are
independent of thermal history, only a single yield stress
measurement will suffice to fully characterize the material
and enable numerical simulation of its mechanical
performance.[10] Unfortunately, however, this might be a
trivial exercise in the case of a standardized test piece with
homogeneous properties; it surely is not in the case of more
complex product geometries, possessing a heterogeneous
distribution of yield stress throughout the product. Moreover, for true product optimization one would like to predict
the final properties of a product in a virtual environment,
without even the need of making a prototype.
In the present study, we aim to develop a predictive,
quantitative methodology that will enable not only
evaluation of this process-induced structural development,
but also provide direct assessment of its influence on the

830

Macromol. Mater. Eng. 2009, 294, 829838


2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

resulting short- and long-term mechanical performance of


the final product. The basis is formed by a recently
developed methodology that allows the evaluation of the
evolution of the yield stress of glassy polymers during
processing.[2426] The method employs standard simulation
tools to obtain the thermal history during cooling from the
melt in each position of a molded product. If the aging
kinetics are known, this information allows calculation of
the yield stress distribution throughout the final object. By
translating this yield stress information into the corresponding Sa-distributions, we open new possibilities for
direct numerical evaluation of the short- and long-term
performance of products in a very early stage of design. This
will provide designers insight into consequences of specific
choices of processing parameters on the time-to-failure of
their products, and provide a means for true product
optimization.

Modeling
Multi-Mode EGP-Model
The basis of the EGP-model is the decomposition of the
Cauchy stress into a hardening stress sr and a driving stress
ss, which is split up into a hydrostatic part (superscript h)
and a deviatoric part (superscript d).[23] The deviatoric part
is modeled as a combination of n parallel linked Maxwell
elements,[22,27,28] which, for isothermal conditions, leads to:
s s r s hs

n
X

s ds;i

i1

~ d kJ  1I
Gr B

n
X

~d :
Gi B
e;i

(1)

i1

~ the isochoric
where Gr is the strain-hardening modulus, B
left Cauchy-Green strain tensor, k the bulk modulus, J the
volume change ratio, I the unity tensor, and G is the shear
modulus. The subscript e refers to the elastic part, the
subscript i refers to a specific mode, i 1; 2; 3; . . . ; n.
The evolution of the elastic and volumetric strains is
given by:
J_ JtrD

(2)





~_ e;i ~L  Dp;i B
~ e;i B
~ e;i ~Lc  Dp;i :
B

(3)

The plastic deformation rate tensors Dp;i are related to


the deviatoric stresses s ds;i by a non-Newtonian flow rule:
Dp;i

s ds;i
;
2hi t; p; Sa

(4)

DOI: 10.1002/mame.200900227

Predicting the Long-Term Mechanical Performance of Polycarbonate from Thermal History

where t, the total equivalent stress, and p, the hydrostatic


pressure, depend on the total stress, according to:
r
1 d d
1
t
s : s s ; p  trs :
2 s
3

(5)

The viscosities are described by an Eyring flow rule,


which has been modified[21,23] to take pressure dependence
and intrinsic strain softening into account:
 
t=t 0
mp
exp
hi h0;i
expS:
sinht=t 0
t0

(7)

r1 r2 1=r1

 
1 r0 exp g p
R gp
r 1=r1
1 r0r1 2

(8)

where r0, r1, and r2 are the fitting parameters. The


evolution of the plastic strain, g p , is governed by the
mode with the longest relaxation time:
t1
g_ p
h1

where

r
1 d
s : sd :
t1
2 s;1 s;1

(9)

Parameters used here are adopted from[27,28] and are


given in the (Table 1 and 2). An illustration of the model
response for materials with different thermal histories with
Table 1. Injection molded samples.

a)

a)

Tm

sy

-C

MPa

30

57.1

27.6

90

58.8

29.1

130

62.1

32.2

Measured at _ 1  103 s1 .

Macromol. Mater. Eng. 2009, 294, 829838


2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

t0

Sa

r0

r1

r2

MPa

MPa

MPa

26

3 750

0.7

0.08

0.965

50

3

Table 3. Multi-mode Maxwell spectrum.

Mode

where Sa captures the initial thermodynamic state of the


material, which is hereassumed
to be defined solely by its

thermal history, and R g p is the softening function:


Gr

(6)

where h0;i is the zero viscosity of the ith mode, t0 the


characteristic stress, and S is the state parameter that
captures the effect of thermal history (aging) and strain
softening. Note that the viscosities are defined with
respect to the rejuvenated reference state.[17,23,29]
 
S is related to the equivalent plastic strain g p according to:
 
 
S g p Sa R g p

Table 2. Model parameters.

Sa

h0

Pas1

MPa

2.10  1017

3.50  102

3.48  10

16

5.55  101

2.95  1014

4.48  101

2.84  10

13

4.12  101

2.54  1012

3.50  101

2.44  10

11

3.20  101

2.20  1010

2.75  101

2.04  10

2.43  101

1.83  108

2.07  101

10

1.68  10

1.81  101

11

1.51  106

1.55  101

12

1.40  10

1.37  101

13

1.27  104

1.19  101

14

1.10  10

9.80  100

15

1.23  102

1.04  101

16

2.62  10

2.11  100

17

2.14  100

1.64  101

respect to their rejuvenated reference state, i.e., quenched


and annealed, is given in Figure 1. Effectively an increase in
Sa shifts the relaxation time spectrum to longer relaxation
times (Figure 1 left), leading to an increase in the yield stress
(Figure 1 right).

Evolution of the State Parameter


The model as presented above is capable of giving
quantitative descriptions of various deformation modes
once the initial thermodynamic state of a material, Sa, is
known.[10,23,30] The value of the initial thermodynamic
state is obtained by a single mechanical test, but when
known, the model even accounts for aging during the
experimental investigation.[10] This aging is captured
by introducing the evolution of the state parameter
with an effective time and was determined based on
annealing experiments below Tg.[23] In Figure 2 a schematic

www.mme-journal.de

831

T. A. P. Engels, L. C. A. van Breemen, L. E. Govaert, H. E. H. Meijer

90

comp. true stress [MPa]

relaxation modulus [MPa]

10

10

10

Sa
1

10

10 8
10

10

10

10

10

12

10

16

10

20

10

24

10

80
70
60
50

40
30
20
10
0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

comp. true strain[]

time [s]

where c0 and c1 are the two constants


(which can be derived from the intersect,
i.e., intersection of the curve with the
vertical axis at t1 s, and slope of
the linear relation of yield stress with
the logarithm of time, respectively)
and their values equal 4.41 and 3.3,
respectively, ta is the initial age of the
material, t01 s and teff is the effective
time defined as:

Figure 1. Left: Influence of state parameter on relaxation modulus versus time. Right:
influence of state parameter on true stress versus true strain response. Dashed lines ():
rejuvenated state; solid lines (-): increasing thermal histories, i.e., quenched and
annealed, respectively.

representation is given of the evolution of the yield stress


with time. In this figure a division in two regions, i.e.,
processing and service life, can be seen. The initial height of
the yield stress is directly determined by the processing
conditions, but upon aging (service life) it starts to increase
and will, independently of its initial properties, follow the
same evolution with time. The rate of increase in the yield
stress is independent of the processing conditions, but is
increased by temperature. The initial region in which the
yield stress does not increase with time, however, is related
to the processing conditions. The length of this region can be
defined as the initial age, ta, of the material, i.e., the time at
which the yield stress starts to increase.
The state parameter is directly proportional to the
uniaxial yield stress and captures the aging kinetics. It
is the only parameter in the model which changes with time
and is given by the following relation:

yield stress

(10)

t
0

0
0
a1
T T t dt

(11)

The effective time is thus the time


accelerated by temperature and captures
the thermal history of the material. The
acceleration is governed by an Arrhenius
dependency, given by:

aT T t exp



DUa
1
1

T t Tref
R

(12)

where DUa is the activation energy of aging, R the


universal gas constant, and Tref a reference temperature.
If the tensile tests are performed at the same strain rate,
and it is assumed that the strain at yield is independent of
applied strain rate, the yield stress can be given analogous
to Equation 10 where the only variable is the effective time
of the material, and thus the thermodynamic state of the
material. The experimentally determined yield stress can
now be given by:
s y t; T s y;0 c log



teff t; T ta
Sa t; T c0 c1 log
t0



teff t; T ta
t0

(13)

In this equation s y;0 is the intersect of the linear relation


of yield stress with the logarithm of effective time and c is
the slope. Mind that s y;0 will depend on the strain rate used
for the determination of the master curve. Values for the
parameters used in this study can be found in.[24]

Properties from Processing

processing
related

service
related

log(time)
Figure 2. Yield stress versus life-time: two regions; processing
related and service related.

832

teff t; T

Macromol. Mater. Eng. 2009, 294, 829838


2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

In a previous study we showed that the yield stress of


polycarbonate (PC) can even be predicted if the thermal
history experienced during processing is taken into
account,[24] rendering a mechanical determination of the
thermodynamic state obsolete. To do so, the evolution of the
effective time has to be followed starting from the moment
the glass starts to solidify during processing. In effect, the
effective time previously[23] was used as a parameter which
captures the thermal history of the material from the
moment we start evaluating the properties of the material,

DOI: 10.1002/mame.200900227

Predicting the Long-Term Mechanical Performance of Polycarbonate from Thermal History

i.e., the start of the experiment. The initial age we


introduced takes into account the thermal history the
material already received prior to the experiment and is
mostly unknown. However, if we start our numerical
investigation already during processing, we can quantitatively predict the initial age of the material.[24] Applying
Equation 11 to the thermal history the material receives
during processing gives:
T > Tg : teff;c 0 and
T  Tg : teff;c

Z
0

tc

t_eff;c 0

0
0
a1
T Tc t dt

cular weight and weight-average molecular weight were 9.2 and


25.8 (kgmol1), respectively.

(14)

here teff;c is the effective time which accumulates during


processing and will give the initial age, ta and Tc t is the
thermal history of the material during processing. From ta
the value of the state parameter, Sa can be calculated and used
as input for the investigation of the mechanical performance.
Tg is effectively used as an input parameter and
determined from the maximum of the loss angle of a DMTA
measurement. Surely this is a very limited interpretation of
the complex phenomenon know as the glass transition.[8] In
a second study on processing induced properties[25] we used
the framework of structural relaxation[3133] to predict the
yield stress. This approach used a more physically accurate
description of the glass transition, but did not improve on
the accuracy of the predictions, i.e., both methods giving
quantitative excellent predictions. Of course the evolution of
the thermodynamic state also has a pronounced effect on
other polymer properties, e.g., enthalpy, volume, creep
compliance, and stress-relaxation modulus,[9,3436] but it is
generally accepted that the times scales on which aging
affects different properties are not the same and cannot be
not correlated in a simple manner.[9,36]
The thermal history during processing is determined by
means of numerical simulation of the injection molding
process. This is done using the commercial injection
molding simulation package MoldFlow MPI. For the flat
rectangular plate, from which the tensile bars are taken, a
2.5D approach is used in which the velocity and temperature fields are computed 3D, but for the pressure a 2D
approximation, which is valid in the case of a thin product,
is used. For the actual product full 3D computations are
performed. Of course any other finite element approach/
package can be used.

Experimental Part
Materials
The material used in this study was an injection molding grade of
PC: Lexan 141R, supplied as granules by Sabic Innovative Plastics
(Bergen op Zoom, The Netherlands). The number-average moleMacromol. Mater. Eng. 2009, 294, 829838
2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

Figure 3. Injection molded samples and tensile bars made thereof.

Sample Preparation
All tensile experiments were performed on samples machined from
injection-molded rectangular plates, see Figure 3 (left). The plates
with dimensions 70  70  1 mm3 were molded on an Arburg 320S
all-rounder 500150. The runner of the mold ensured uniform
filling, as proven by several short shot experiments. The only
variable used during the injection molding is the mold temperature. Melt temperature was kept constant at 285 8C, as was the
cooling time at 60 s. From the plates, bars with dimensions
70  10  1 mm3 were cut parallel to the flow direction and fitted
with gauge sections of 33  5  1 mm3, see Figure 3 (right).
To investigate the influence of the temperature history on the
distribution of yield stress over the thickness of a product by means
of micro-indentation, small bars with cross-sections 2  1 mm2
were taken from the centers of the molded square plates by a
precision machining operation. Subsequently, increasingly thinner
layers (1 mm) were removed from the surface of the 2  1 mm2
cross-section by a microtoming operation under cryogenic (liquid
nitrogen) conditions, this to minimize a possible influence of
machining on the thermodynamic state of the sample surface. A
Leica RM 2165 rotation microtome was used; for each sample a
fresh glass knife was taken. To obtain different temperature
histories, three mold temperatures were used 30, 90, and 130 8C.
To verify the performance after injection molding and the
influence of thermal history thereupon, tensile samples were taken
from injection molded plates processed at mold temperatures of 30
and 120 8C.
Finally, as an example of a real product, a thick-walled cupshaped sample, see Figure 4, was used. The bottom ring of the cup
has a diameter of 78 mm. The cup itself starts with an outer
diameter of 65 mm and has a gradually decreasing diameter up till
60 mm at the top. We define Figure 4 (left) as the upright position
(the way in which the samples was loaded in the load-frame). The
thickness was around 3 mm in all cross-sections. The cup-shaped
sample was injection molded on an Arburg 320S all-rounder
500150. The melt temperature was set to 285 8C; the injection
rate to 50 ccms1; and the packing pressure to 500 bars. Mold
temperatures were set to 30 and 130 8C. In both cases a cooling time
of 120 s was used.
All samples used in this study were stored at room temperature
after injection molding. For PC no increase in yield stress was to be
expected at room temperature based on the time-temperaturesuperposition results of Klompen et al.[23] It was also experimen-

www.mme-journal.de

833

T. A. P. Engels, L. C. A. van Breemen, L. E. Govaert, H. E. H. Meijer

Figure 4. Injection molded cup.


Figure 5. Left: force versus indentation depth for a quenched and
an annealed material. Right: residual indents on a sample of mold
temperature 130 8C.

tally shown that PC stored at room temperature did not show an


increase in yield stress for at least a period of 3 years.[37]

Methods
Micro-indentation experiments were performed on a nanoindenter XP (MTS NanoInstruments, Oak Ridge, Tennessee) under
displacement control. The indenter has a flat tip, effectively a flatended cone with a top angle of 708 and a circular contact area with
diameter 10 mm. Correction for tip-sample misalignment was
performed using a specially designed alignment tool. For details on
the experimental technique see;[30,38] and for the alignment tool
see.[38]
Tensile tests were performed on a Zwick Z010 universal tensile
tester at a room temperature of 23 8C. Experiments were performed
_ 0 or engineering
by applying constant linear strain rates "_ x=l
stresses s F=A0 . Unless indicated otherwise, a standard
constant linear strain rate of 103 s1 was used. All tensile yield
stresses listed in the results section are engineering yield stresses,
and taken as the mean value of five experiments.
The cup-shaped samples were tested on a Zwick 1475 tensile
testing machine. The cups were placed in the machine with the
tapered section to the top (see Figure 4 left). The bottom plate used
has a flat circular recess to fit the outer diameter of the bottom ring
of the cup. Samples were loaded with constant displacement rates,
or constant forces. Experiments were performed at a room
temperature of 23 8C and corrected for the finite stiffness of the
setup.

Results
Influence of Thermal History on the Yield-Stress
Distribution
Previously[24,25] distributions of yield stresses throughout a
polymer product were predicted based on the differences in
temperature history experienced during molding. Fast
cooling, e.g., at the mold surface, limits physical aging
leading to a low yield stress, while during slow cooling, e.g.,
in the core of the product, aging is more pronounced as is the
increase in yield stress. Verification was done using a mean
yield stress calculated based on area averaging, and
excellent agreement was found for different mold temperatures and mold thicknesses. Here we will attempt to
also validate the distributions, first by using micro-

834

Macromol. Mater. Eng. 2009, 294, 829838


2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

indentation. Figure 5 (left) shows results of indentation


with the use of a flat-tip and indeed a measurable
distinction is found in force-displacement results for a
quenched (sy  62 MPa)a and an annealed (sy  70 MPa)a
material.[30]
The yield stress distribution was studied on the crosssection of samples produced with different mold temperatures (resulting in different properties, see Table 1). To
achieve a smooth cross-sectional surface with little plastic
deformation of the surface, the samples were first precision
machined, followed by cryogenic microtoming of the
surface. Indents were made over the total width of the
samples, see Figure 5 (right), and the resulting indentation
forces (at an indentation depth of 2 mm) are presented in
Figure 6 (left). Although the calculated yield stress
distributions (Figure 6 (right)) and the yield stress
measurements (Table 1), both suggest differences in
indentation forces, they are not measured (The decrease
in indentation force at the edges is most likely the influence
of a local decrease in stiffness due to the presence of the
surface.) The conclusion could be that, despite our precautions, sample preparation still influences the thermodynamic state of the surface, preventing to measure properties as they result from processing.
We therefore try to obtain samples that are prepared
without the need of a post-processing machining operation,
by stacking twelve (1 mm thick, vacuum dried) sheets of
extruded PC (200  200 mm2), separated by thin sheets of
aluminium (<0.1 mm thick) to prevent sticking. Two
thermocouples are added, one at a surface sheet, one
between the two middle sheets. The stack of sheets is placed
in a hot press at 200 for 10 min, to erase any previous
thermal history and orientation from the sheets, and
subsequently the stack is cooled in a cold press at 18 8C. The
stack is removed from the press once the center temperature is 18 8C, and the sheets are separated. Subsequently
tensile bars are machined from the center of the sheets with
gauge sections of 33  5  1 mm3.
a

Note: listed yield stresses are engineering values measured at


"_ 1  103 s1

DOI: 10.1002/mame.200900227

Predicting the Long-Term Mechanical Performance of Polycarbonate from Thermal History

25

yield stress [MPa]

The 3D constitutive framework of the


predictive tool developed[23] allows to
predict the performance for an arbitrary
sample geometry under any loading
condition.[10] To validate this, tensile
experiments are performed at different
strain rates and stresses for samples
made using mold temperatures of 30
and 120 8C. Next a more complex geometry is analyzed, using a cup-shaped
product.
Figure 8 (left) shows the cooling
profiles over the thickness in the center
of square samples, see inset Figure 8 (left),
as obtained from a numerical simulation
of the injection molding process for mold
temperatures of 30 and 120 8C. Cooling is
Macromol. Mater. Eng. 2009, 294, 829838
2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

center

surface
center

50

0
0

200

400

600

800

55

50

1000

time [s]

10 11 12

sheet #

Figure 7. Left: cooling histories as measured (-) and as calculated (). Right: measured
(*) and predicted (solid drawn line) yield stress distributions.

300
70

temperature [ C]

Influence of Processing on the Final


Properties of a Product

yield stress [MPa]

250
200
surface to center

150
100

Tmold=120C

surface to center

50
Tmold=30C

0
0

10

15

time [s]

20

25

30

yield stress [MPa]

temperature [ C]

ind. force @ 2 m [mN]

faster near the surface than at the center,


and cooling rates are much higher for the
24
60
low mold temperature. Subsequently we
23
compute point wise the resulting local
55
value of Sa and the related value of the
22
yield stress, see Figure 8 right. Near the
50
T = 130C
T =130C
surface we find a much lower yield stress
21
T = 90C
T = 90C
45
compared to that in the center of the
T = 30C
T = 30C
20
plate, related to the higher cooling rates
1
0.5
0
0.5
1
1
0.5
0
0.5
1
normalized thickness []
normalized thickness []
occurring there. To enable comparison of
the numerical predictions with experiFigure 6. Left: measured indentation force distributions over the thickness of samples
with different thermal histories. Right: calculated yield stress distributions over the mentally determined yield stresses, a
thickness-weighted average of the comthickness of samples with different thermal histories.
puted Sa distribution was determined.
This yielded a value of Sa 27.4 for
samples with a mold temperature of 30 8C, and Sa 31.8
Figure 7 (left) shows the calculated and measured
temperatures (heat capacities used as input for the
for those with a mold temperature of 120 8C.
numerical results were taken from the ATHAS Data
Experimental and numerical results on tensile tests at a
Bank[39,40]), and Figure 7 (right) shows the calculated,[24]
strain rate of 103 s1 are compared in Figure 9. Using the
and measured yield stress distributions. The calculated
multi-mode EGP-model the stress responses up to yield are
accurately predicted, while the postyield localization is
distribution overestimates the experimental one by
stronger in the experiments than in the simulations. This
2 MPa, i.e., a deviation of 3%. When the predicted
can be attributed to sample preparation. The tensile bars are
distribution is lowered by 2 MPa (dashed line), an excellent
made by a machining operation and are not polished
agreement between experimental and numerical distribuafterwards giving them a rather rough surface, resulting in
tions is obtained. In conclusion it proves to be difficult to
strong localization, which is not captured by simulation.
measure yield stress distributions caused by inhomogeneous cooling via direct indentation measurements, due to
errors induced by the sample preparation
70
method. But by using samples that did
model
measured
200
model output minus 2 MPa
not require any post-fabrication machincalculated
65
ing, the calculated distribution indeed
150
can be measured, and the results are
surface
60
correct with an error of 3% only.
100
65

60

Tmold=120C

50

Tmold= 30C

center
0

0.25

0.5

0.75

surface
1

normalized thickness []

Figure 8. Left: temperature versus time during the cooling of the injection molded
samples. Right: corresponding predicted yield stress distributions.

www.mme-journal.de

835

T. A. P. Engels, L. C. A. van Breemen, L. E. Govaert, H. E. H. Meijer

70

eng. stress [MPa]

60
50
40
30

Figure 11. Cup-shaped sample (middle) and injection molding


analysis mesh (left) and structural analysis mesh (right) made
thereof.

20
Tm = 120C
T = 30C

10

model prediction

0
0

whether it will also perform well on a more complex


product. To investigate this a cup-shaped sample is chosen,
as discussed in the Experimental Part, and shown in
Figure 9. Experimental and numerical results of tensile tests at a
Figure 11 (middle). For the analysis of the injection molding
strain rate of 103 s1 for samples with mold temperatures 30 and
process, a full 3D mesh is build, see Figure 11 (left), while for
120 8C.
the structural analysis a 2.5D axi-sym70
70
metrical mesh proved to be sufficient, see
Tm = 120C
Tm = 120C
Tm = 30C
Tm = 30C
Figure 11 (right). The dimensions of the
65
65
model prediction
model prediction
mesh for the mold filling analysis are
60
60
based on the dimensions of the mold,
55
55
whereas the dimensions of the mesh for
structural analysis are based on dimen50
50
sions of actual samples, which however
45
45
differ only slightly from the dimensions
of the mold.
40
1
2
3
4
5
10
10
10
10
10
105
104
103
102
101
Analysis of the injection molding
1
timetofailure [s]
strain rate[s ]
process is again performed with the
Figure 10. Left: yield stress versus applied strain rate. Right: applied stress versus time- commercial injection molding simulato-failure. Both for samples with mold temperatures 30 and 120 8C.
tion package (MoldFlow MPI). A full 3D
analysis is used with a mesh which has
eight elements over the thickness and a total of 800 000
However, while it influences the width of the yield peak, the
elements. The large amount of elements is necessary to
height is unaffected.
obtain a sufficiently detailed temperature information over
Figure 10 (left) shows the yield stress versus the strain
the thickness of the cup. In Figure 12 the distribution of the
rate for samples made with mold temperature 30 and
state parameter Sa is given for a quarter cross-section taken
120 8C, and Figure 10 (right) the applied stress versus the
time-to-failure in a creep experiment. Solid lines are
from the center height of the cup. As expected, the predicted
predictions based on our modeling approach and they
values of Sa are much higher for the 130 8C samples than for
are in excellent agreement with experimental results. The difference in mold
40
Tm = 30C
Tm = 130C
30
30
temperature of 90 8C between the two
35
samples, results in an increase in failure
30
time of about a factor 100 if the samples
20
20
25
are loaded with the same stress. A similar
effect was already found and predicted
20
for a quenched versus an annealed
15
material,[23] but mind that here this
10
10
10
effect is predicted based on the tempera5
ture history the material experienced
during its fabrication in the injection
0
0
0
0
10
20
30
0
10
20
30
Sa
molding process.
x [mm]
x [mm]
The model thus far proves to predict
the performance of simple tensile bars
Figure 12. Distribution of Sa over a quarter cross-section for Tm 30 8C (left), and for
Tm 130 8C (right).
very well. Of course the question rises
0.02

0.04

0.06

0.08

0.1

836

Macromol. Mater. Eng. 2009, 294, 829838


2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

y [mm]

y [mm]

applied stress [MPa]

eng. yield stress [MPa]

strain []

DOI: 10.1002/mame.200900227

Predicting the Long-Term Mechanical Performance of Polycarbonate from Thermal History

45
T = 130C
m

40

Tm = 30C
model prediction

force [kN]

35
30
25
20
15
10
5
0
0

displacement [mm]
Figure 13. Experimental and numerical results of compression
tests of a cup at a loading rate of 0.045 mms1 for samples with
mold temperatures 30 and 130 8C.

50

50

40

40

applied load [kN]

max. load [kN]

the 30 8C samples. The distributions are, however, more


pronounced for the last one. These findings correspond well
with the results observed for the yield stress of the tensile
bars molded at different mold temperatures, see Figure 8
(right).
To facilitate the numerical analysis, again a thicknessaveraged value of Sa is used, similar to what was done for
the tensile bars. The calculated area mean values are
Sa  25.0 for the 30 8C mold temperature samples and
Sa  34.9 for the 130 8C samples. The results of the structural
analysis are shown in Figure 13 and 14, where markers
indicate experimental results and solid drawn lines are the
predictions. For the 130 8C samples the loading path as
shown in Figure 13 is predicted accurately. This also holds
for the rate dependence of the maximum force, see Figure 14
(left), and the stress dependence of the time-to-failure
under static load, see Figure 14 (right). For the 30 8C samples
the overall loading path is described well, although we
observe an underestimation of the maximum force of 6%.
This deviation appears to be systematic as it is also observed
in the rate dependence of the maximum force, see Figure 14

30

20

10

0 4
10

Tm = 130C
Tm = 30C
model prediction
3

10

10

10

10

10

(left), and the stress dependence of the time-to-failure


under static load, see Figure 14 (left). Please note that the
slopes of the predictions in Figure 14 in all cases correlate
well with experimental data and all predictions are within
10% accuracy.
On second thought, the predicted Sa 25 for the 3 mm
thick cups appears rather low, especially when compared to
the Sa 27.4 for 1 mm thick tensile samples at the same
mold temperature of 30 8C. Re-examination of the Sa
distributions learned that at the surface Sa values are as
low as 69, which is unrealistic. Such values correspond to
yield stresses normally only encountered after mechanical
rejuvenation. The explanation for this anomaly is found in
the fact that the cooling profiles for the surface nodes
contain insufficient information, i.e., the temperature at the
first time increment is already below Tg, thereby giving an
underestimation of the actual thermodynamic state. [See
also the rough Sa values at surfaces in the inset of Figure 12
(left)]. The use of the commercial injection molding package
here limits the detail of our investigation, since the time
resolved information which can be extracted is limited to
100 data points, apparently posing a problem in the 3D case
but not for the 2.5D case. To demonstrate the impact of this
shortcoming we recalculated the area mean value of Sa
while excluding the surface nodes, resulting in a value of
Sa  30.2 for the 30 8C sample and a small but insignificant
difference for the 130 8C sample. Predictions for Sa 30.2 are
added as dashed lines in Figure 13 and 14 and give a slight
overestimation of the data. A better calculation of the
cooling profiles therefore is expected to result in
25 < Sa < 30.2.
As a final remark we note that the numerically calculated
and experimentally observed failure modes were, thus far,
all ductile. The cup-shaped samples which were injection
molded with a high mold temperature, however, display a
transition from ductile failure [open symbols, Figure 14
(right)] to brittle failure [closed symbols, Figure 14 (right)].
The model still predicts the failure times accurately, since
despite the change in failure mode, the kinetics of plastic
flow determine the onset to failure. This is in agreement
with experiments on loaded poly(vinyl
chloride) pipes[41] and on tensile bars of
PC with different molecular weights.[10]

30

20

Conclusion
(open markers)

(closed markers)

10

0 0
10

loading rate [mm/s]

Tm = 130C
Tm = 30C
model prediction
1

10

10

10

10

10

10

10

timetofailure [s]

Figure 14. Left: maximum load versus loading rate. Right: applied load versus time-tofailure. Both for samples with mold temperatures 30 and 130 8C.
Macromol. Mater. Eng. 2009, 294, 829838
2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

A previously developed modeling


approach which allows the prediction
of the yield stress distribution as it
follows directly from the temperature
history experienced during processing[24]
is validated more extensively. Predicted
yield stress distributions are experimen-

www.mme-journal.de

837

T. A. P. Engels, L. C. A. van Breemen, L. E. Govaert, H. E. H. Meijer

tally validated and found to be in good agreement. For


simple tensile bars the performance under various constant
strain rates and stresses was also found to be in excellent
agreement with experimental results. Finally, the prediction of the performance of a more complex product in the
form of an injection molded cup was investigated for
different loading rates and applied loads and again
excellent agreement was observed, including the tremendous influence of different mold temperatures, that can
change the lifetime of polymer products with more than
two orders of magnitude. Given the good overall agreement
between predictions and experimental results the framework presented in this study facilitates the mechanical
optimization of polymer products without the need for
performing any mechanical tests.
Acknowledgements: Authors are grateful to the Dutch Polymer
Institute (DPI) for financial support (grant 578).

Received: February 13, 2009; Revised: July 30, 2009; Published


online: October 28, 2009; DOI: 10.1002/mame.200900227
Keywords: mechanical properties; modeling; performance;
processing

[1] M. R. Kantz, H. D. Newman, Jr., F. H. Stigale, J. Appl. Polym. Sci.


1972, 16, 1249.
[2] G. Menges, G. Wu
bken, B. Horn, Colloid Polym. Sci. 1976, 254,
267.
[3] B. A. G. Schrauwen, L. C. A. van Breemen, A. B. Spoelstra, L. E.
Govaert, G. W. M. Peters, H. E. H. Meijer, Macromol. Symp.
2002, 185, 89.
[4] B. A. G. Schrauwen, L. C. A. van Breemen, A. B. Spoelstra, L. E.
Govaert, G. W. M. Peters, H. E. H. Meijer, Macromolecules 2004,
37, 8618.
[5] Z. Tadmor, J. Appl. Polym. Sci. 1974, 18, 1753.
[6] L. F. A. Douven, F. P. T. Baaijens, H. E. H. Meijer, Prog. Polym. Sci.
1995, 20, 403.
[7] T. A. P. Engels, L. E. Govaert, H. E. H. Meijer, Macromol. Mater.
Eng. 2009, 294, in print. DOI: 10.1002/mame.200900050.
[8] G. B. McKenna, Comprehensive Polymer Science, Vol 2: Polymer Properties, chapter Glass Formation and Glassy Behavior,
Pergamon Press, Oxford 1989, pp. 311362
[9] J. M. Hutchinson, Prog. Polym. Sci. 1995, 20, 703.
[10] E. T. J. Klompen, T. A. P. Engels, L. C. A. van Breemen, P. J. G.
Schreurs, L. E. Govaert, H. E. H. Meijer, Macromolecules 2005,
38, 7009.
[11] R. P. M. Janssen, D. de Kanter, L. E. Govaert, H. E. H. Meijer,
Macromolecules 2008, 41, 2520.

838

Macromol. Mater. Eng. 2009, 294, 829838


2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

[12] R. P. M. Janssen, L. E. Govaert, H. E. H. Meijer, Macromolecules


2008, 41, 2531.
[13] H. G. H. van Melick, L. E. Govaert, H. E. H. Meijer, Polymer 2003,
44, 3579.
[14] H. E. H. Meijer, L. E. Govaert, Macromol. Chem. Phys. 2003, 204,
274.
[15] M. C. Boyce, D. M. Parks, A. S. Argon, Mech. Mater. 1988,
7, 15.
[16] E. M. Arruda, M. C. Boyce, Int. J. Plast. 1993, 9, 697.
[17] O. A. Hasan, M. C. Boyce, X. S. Li, S. Berko, J. Polym. Sci., Part B:
Polym. Phys. 1993, 31, 185.
[18] C. P. Buckley, D. C. Jones, Polymer 1995, 36, 3301.
[19] J. J. Wu, C. P. Buckley, J. Polym. Sci., Part B: Polym. Phys. 2004,
42, 2027.
[20] C. P. Buckley, P. J. Dooling, J. Harding, C. Ruiz, J. Mech. Phys.
Solids 2004, 52, 2355.
[21] L. E. Govaert, P. H. M. Timmermans, W. A. M. Brekelmans,
J. Eng. Mater. Technol., Trans. ASME 2000, 122, 177.
[22] T. A. Tervoort, E. T. J. Klompen, J. Rheol. 1996, 40, 779.
[23] E. T. J. Klompen, T. A. P. Engels, L. E. Govaert, H. E. H. Meijer,
Macromolecules 2005, 38, 6997.
[24] L. E. Govaert, T. A. P. Engels, E. T. J. Klompen, G. W. M. Peters,
H. E. H. Meijer, Int. Polym. Process. 2005, XX 170.
[25] T. A. P. Engels, L. E. Govaert, G. W. M. Peters, H. E. H. Meijer,
J. Polym. Sci., Part B: Polym. Phys. 2006, 44, 1212.
[26] T. A. P. Engels, B. A. G. Schrauwen, L. C. A. van Breemen, L. E.
Govaert, Int. Polym. Process. 2009, XXIV 167.
[27] L. C. A. van Breemen, E. T. J. Klompen, L. E. Govaert, H. E. H.
Meijer, J. Mech. Phys. Solids 2009, Chapter 2. http://www.
mate.tue.nl/mate/pdfs/10677.pdf.
[28] L. C. A. van Breemen, Contact Mechanics in Glassy Polymers
Ph.D. thesis, Eindhoven University of Technology, 2009, Eindhoven, The Netherland.
[29] H. G. H. van Melick, L. E. Govaert, B. Raas, W. J. Nauta, H. E. H.
Meijer, Polymer 2003, 44, 1171.
[30] L. C. A. van Breemen, T. A. P. Engels, C. G. N. Pelletier, L. E.
Govaert, J. M. J. den Toonder, Philos. Mag. 2009, 89, 677.
[31] A. Q. Tool, J. Am. Ceram. Soc. 1946, 29, 240.
[32] O. S. Narayanaswamy, J. Am. Ceram. Soc. 1971, 54, 491.
[33] C. T. Moynihan, P. B. Macedo, C. J. Montrose, P. K. Gupta, M. A.
DeBolt, J. F. Dill, B. E. Dom, P. W. Drake, A. J. Easteal, P. B.
Elterman, R. P. Moeller, H. Sasabe, J. A. Wilder, Ann. N. Y. Acad.
Sci. 1976, 279, (Glass Transition Nat. Glassy State), 15.
[34] I. M. Hodge, J. Non-Cryst. Solids 1993, 169, 211.
[35] L. C. E. Struik, Polym. Eng. Sci. 1978, 18, 799.
[36] G. W. Scherer, Relaxation in Glass and Composites, Krieger
Publishing Company, Malabar, Florida U.S.A. 1986.
[37] J. C. Bauwens, Plast. Rubber Process. Appl. 1987, 7, 143.
[38] C. G. N. Pelletier, E. C. A. Dekkers, L. E. Govaert, J. M. J. den
Toonder, H. E. H. Meijer, Polym. Test. 2007, 26, 949.
[39] Athas data bank: http://athas.prz.edu.pl/.
[40] U. Gaur, S. Lau, B. Wunderlich, J. Phys. Chem. Ref. Data 1983,
12, 91.
[41] H. Niklas, H. H. Kaush von Schmeling, Kunstoffe 1963, 53,
886.

DOI: 10.1002/mame.200900227

Das könnte Ihnen auch gefallen