Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Composites Part B 97 (2016) 68e83

Contents lists available at ScienceDirect

Composites Part B
journal homepage: www.elsevier.com/locate/compositesb

Creep behavior and modeling of neat, talc-filled, and short glass fiber
reinforced thermoplastics
Mohammadreza Eftekhari, Ali Fatemi*
Mechanical, Industrial and Manufacturing Engineering Department, The University of Toledo, 2801 West Bancroft Street, Toledo, OH 43606, USA

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

Article history: Creep behavior of neat, talc-filled, and short glass fiber reinforced injection molded thermoplastic
Received 29 January 2016 composites were investigated and modeled at room and elevated temperatures. Creep strength
Received in revised form decreased and both creep strain and creep rate increased with increasing temperature. The temperature
6 April 2016
effect was more significant for samples with glass fibers in the transverse to the load direction, as
Accepted 17 April 2016
compared with the longitudinal direction. The LarsoneMiller parameter was able to correlate the creep
Available online 1 May 2016
rupture data of all materials. The MonkmaneGrant relation and its modification were successfully used
to correlate minimum creep rate, time to rupture, and strain at rupture data. The Findley power law and
Keywords:
A. Discontinuous reinforcement
timeestress superposition principle (TSS) were used to represent nonlinear viscoelastic creep curves.
Short glass fiber Long-term creep behavior was also satisfactory predicted based on short-term test data using the TSS
A. Thermoplastic resin principle.
B. Creep © 2016 Elsevier Ltd. All rights reserved.
C. Analytical modelling

1. Introduction When a constant load is applied on the specimen, an initial


elongation occurs which is known as instantaneous strain, ε0. This
The usage of polymeric materials is increasingly rapidly due to is followed by a rapidly decreasing deformation rate known as the
their lightweight, low cost, and capability to be manufactured in primary stage, followed by a steady-state linear deformation stage,
complex geometries at high production rates. Addition of fillers known as the secondary creep stage, for which the experimental
such as talc or reinforcement such as short glass fibers can improve data is typically of most interest. The final stage involves rapid
the mechanical performance of unreinforced polymers to a high deformation at an accelerated rate known as tertiary deformation,
degree. These materials are typically subjected to loading and ultimately leading to the specimen rupture. Some materials do not
environmental conditions causing creep deformation. This time- have the secondary stage, while tertiary creep only occurs at high
dependent behavior results in molecular rearrangement, the de- stresses and for ductile materials [2]. Instantaneous creep strain, ε0,
gree of which depends on factors including the type of material, can consist of elastic (εe) and plastic (εp) strain depending on the
magnitude of the stress, temperature, and time. level of the applied load or stress and can constitute a considerable
For a linear viscoelastic material under creep condition, the fraction of the total strain.
applied stress is proportional to the strain response at a given time, While creep in metallic materials occurs only at elevated tem-
independent of the stress. For viscoelastic creep, the behavior ex- peratures, creep of polymers can be significant at any temperature.
hibits nonlinearity at large strains. While linear viscoelastic Therefore, creep or relaxation are major concerns in using polymer
behavior is represented using physical based constitutive equa- and polymer composites in structural applications. Long-term
tions, such as Maxwell model, constitutive equations for nonlinear creep deformation and strength applicable to service condition
creep or other viscoelastic behaviors are much more complex. As a are normally obtained by extrapolation of short-term test data
results, they require a large number of functions with higher order obtained under accelerated testing conditions, such as higher
stress terms with a large number of material constants and are temperature, stress, and humidity by using prediction models.
typically empirical in nature [1]. A number of studies in the literature have investigated creep
behavior of neat polymers including damage mechanism [3],
nonlinear creep behavior and modeling [4], superposition methods
for long-term creep predictions considering environmental effects
* Corresponding author. Tel./fax: þ1 419 530 8213.
E-mail addresses: mohammadreza.eftekhari@rockets.utoledo.edu (M. Eftekhari),
[5e7], and creep rupture behavior and predictions [8,9]. Seltzer
afatemi@eng.utoledo.edu (A. Fatemi). et al. [10] studied the effect of organoclay on creep behavior of

http://dx.doi.org/10.1016/j.compositesb.2016.04.043
1359-8368/© 2016 Elsevier Ltd. All rights reserved.
M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83 69

35 wt% talc (referred to as PO) and a 40 wt% talc filled poly-


Nomenclature propylene (referred to as PP-T). The short fiber reinforced com-
posites are polypropylene containing 30 wt% short glass fibers
A, B, C, D, G, H, a, b material constants (referred to as PP-G), polyamide-6.6 reinforced with 30 wt% glass
CLMP LarsoneMiller parameter constant fibers (referred to as PA66), 20 wt% short glass fiber reinforced
E elastic modulus modified polyphenylene ether and polystyrene resin (here referred
LMP LarsoneMiller parameter to as PPE/PS), polybutylene terephthalate with 30 wt% glass fibers
m, n Findley power law constants (here referred to as PBT), and polyamide-6 containing 35 wt% glass
S stress fibers and about 10 wt% rubber impact modifier (here referred to as
Su ultimate tensile strength PA6). Common applications of these materials include those in the
t time automotive industry. A summary of the materials used in the
tR time to creep rupture experimental study, including constituents, melting temperature
T temperature (Tm), and glass transition temperature (Tg) is presented in Table 1.
Tg glass transition temperature Mean glass fiber length and diameter for short glass fiber reinforced
Tm melting temperature composites were obtained as 0.25 mm and 10 microns, respectively.
εc creep strain More details on fiber length, diameter, and orientation distribution
ε0 instantaneous strain of studied reinforced composites can be found in Ref. [12]. Tensile
ε_ min minimum creep rate and fatigue behaviors of the mentioned materials were studied
4S shift factor previously in Refs. [12e18].
l, h MonkmaneGrant relation constants Rectangular plates with thicknesses of 2.8 mm (for PP-T, PP-G,
l0 h0 modified MonkmaneGrant relation constants PA66 and PPE/PS) and 3.8 mm (for PO, PP, PA6 and PBT) and di-
mensions of 100  200 mm were injection molded. Rectangular
strips were cut in 90 (transverse) and 0 (longitudinal) directions
with respect to the mold flow direction with a width of about
injection molded polyamide-6 using dynamic mechanical analysis 25 mm according to Fig. 1(a) and (b). A portion of plaques from
(DMA) in cantilever-bending. Eftekhari and Fatemi [11] reviewed injection molding side with the length of 60 mm was discarded to
the studies related to creep behavior and modeling of short fiber prevent end effects. The specimens were machined from the strips
reinforced polymer composites (SFRPCs) at elevated temperatures. using a CNC milling machine with the geometry shown in Fig. 1(c).
Time-temperature-stress superposition (TTSS) procedure has been All the specimens were dried before testing to preserve the dry
used for extending short-term creep data for long-term creep as molded condition and exclude moisture effect. PA66 and PPE/PS
predictions. Empirical power law models have been widely used to specimens were dried for 24 h at 80  C and 4 h at 100  C, respec-
represent nonlinear creep strain curves for a wide range of tem- tively, in a vacuum chamber. PO, PP, PP-T and PP-G specimens were
peratures and stress levels. Prevailing creep damage is crazing dried for 4 h at 80  C, and PBT and PA6 specimens were dried for 6 h
perpendicular to the stress direction where crazes start mainly at at 120  C and 80  C, respectively, in a regular chamber. All the
the fiber/matrix interface independent of the interface quality. specimens were kept in a desiccator prior to testing.
This study investigated creep behavior of several injection
molded neat, talc-filled, and short glass fiber reinforced thermo- 2.2. Experimental method
plastic composites. Creep tests were conducted for different con-
ditions including stress level, temperature, and fiber orientation. Creep tests were conducted using a closed loop servo-hydraulic
Empirical fits were used to represent creep rupture data and relate testing machine which was controlled with a digital controller. A
minimum creep rate data to the applied stress. LarsoneMiller video extensometer was used to measure creep strain. A pair of
parameter was used to correlate creep rupture data and power law pneumatic grips with 5 kN capacity and adjustable pressure system
models were used to represent experimental creep strain curves. which were suitable for testing polymers at low and elevated
Semi-empirical models were used to correlate minimum creep rate, temperatures was used. A strain-gaged specimen was used to align
time to rupture, and strain at rupture together. Creep strain master the load train components prior to testing. An environmental
curves and corresponding shift factor versus stress curves were chamber employing an electronic heating element with accuracy of
developed using short-term creep data according to the timee- better than ±1  C was used for performing tests at elevated tem-
stress superposition (TSS) method. A small number of long-term peratures. A lever arm test machine with dead weights was used for
creep tests were also conducted to verify the capability of the long-term creep tests at room temperature. Some tests which were
developed models for long-term creep predictions based on short- conducted using the servo-hydraulic test machine were repeated
term creep data. The experimental program conducted, the results with the lever arm test machine to verify the repeatability of the
obtained, and the models applied are presented and discussed in results.
the following sections. Tensile creep tests were performed according to ASTM D2990
[2] test standard. Creep tests were conducted in the transverse
direction for all the materials and also in the longitudinal direction
2. Experimental program for PBT and PA6 for a wide range of stress levels at several tem-
peratures, as reported in Table 1. Short-term tests lasted on the
2.1. Materials and specimen order of about one hour to three days. A limited number of longer
time creep tests (up to about 40 days) were also conducted for PP-G.
Neat, talc-filled, and short glass fiber reinforced thermoplastics Two repeat tests were conducted for most of the conditions.
with or without rubber as a modifier were considered for the
experimental study. The neat thermoplastic used is an impact 3. Creep rupture data and modeling
polypropylene copolymer, here designated as PP. The talc-filled
composites used are a polypropylene-elastomer blend (a thermo- Stress versus time to rupture data for all the materials at
plastic polyolefin) with 25% by weight (25 wt%) elastomer and different temperatures are plotted in Fig. 2. Tests that did not fail
70 M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83

Table 1
Summary of the materials with creep test conditions used for the experimental study.

Material Polymer matrix Reinforcement Tm Tg Directions Temperatures Stress range


designation ( C) ( C) studied studied ( C) studied (% Su)

PO Polypropylene þ 25 wt% Rubber 30 wt% Talc >120 <-10 T 23, 85, 125 30e82
PP Impact Polypropylene None 170 4 T 23, 85, 125 40e85
PP-T Polypropylene 40 wt% Talc 165 11 T 23, 85 53e80
PP-G Polypropylene 30 wt% Short Glass Fiber 165 23 T 23, 85, 120 46e85
PA66 Polyamid-6,6 30 wt% Short Glass Fiber 260 55 T 23, 85, 120 75e90
PPE/PS Polyphenylene ether þ Polystyrene 20 wt% Short Glass Fiber 325 135 T 23, 85 60e85
PBT Polybutylene terephthalate 30 wt% Short Glass Fiber 255 50 T, L 85, 125 72e91
PA6 Polyamide-6 þ 10% Rubber 35 wt% Short Glass Fiber 220 40 T, L 85, 125 81e91

L: Longitudinal (fibers along load direction).


T: Transverse (fibers perpendicular to load direction).

Fig. 1. Direction and location of specimens cut from injection molded plaques in (a) the transverse and, (b) the longitudinal direction according to the mold flow direction (arrows).
(c) Specimen geometry used for creep tests (dimensions are in mm).

(run-outs) and are not included in the fits, however, the data points fiber reinforcement for PPE/PS is lower, as compared to the other
are shown in the plots. A significant effect of temperature is SFRPCs, therefore, creep rupture behavior of PPE/PS is more tem-
observed for all the materials, as expected. For both PBT and PA6, perature dependent. A craze type failure was observed for PPE/PS
creep resistant is significantly higher in the longitudinal direction, samples, while no evidence of crazing even at elevated tempera-
as compared to the transverse direction. Also, the effect of tem- tures was observed for PA66, PBT, and PA6.
perature is more accentuated in the transverse direction, where A power fit was utilized to represent the creep rupture experi-
creep resistance is more matrix dependent, as compared with the mental data in the form of:
longitudinal direction. Also.
To be able to compare the creep rupture behavior of the mate- S ¼ AðtR ÞB (1)
rials together, stress versus time to rupture data in Fig. 3 are
normalized with the corresponding ultimate tensile strength, Su, where S is the applied stress level, tR is the time to rupture in hours,
obtained from tension tests conducted at a displacement rate of A is the intercept at tR ¼ 1, and B is the slope of the fitted line. A and
1 mm/min. The data for PO, PP, PP-T, and PP-G are shown in B are temperature and orientation dependent constants and their
Fig. 3(a), while the data for PA66, PPE/PS, PBT, and PA6 are shown in values for all the materials at different temperatures are reported in
Fig. 3(b). The second group of materials are more creep resistant Table 2. Correlations of A and B values with Su are shown in Fig. 4
than the first group, and the Tg of the second group is also higher and expressed by the following equations:
than the first group. PO, PP, PP-T, and PP-G data at 23  C are
correlated very well by normalizing with tensile strength. A ¼ 0:85 Su with R2 ¼ 0:99 (2)
Normalizing with tensile strength also correlates all the creep
rupture data of the second group very well, except for PPE/PS data
at 85  C (see Fig. 3(b)). B ¼ 0:44ðSu Þ0:66 with R2 ¼ 0:8 (3)
Tg and Tm of PO are lower than those of PP, PP-T, and PP-G, These correlations can be used to estimate creep rupture
therefore, more temperature sensitivity in creep rupture behavior properties of thermoplastics and thermoplastic composites at
of PO is expected. The mechanism of creep damage of poly- different temperatures and in different mold flow directions.
propylene, which is mainly due to the crazes perpendicular to the Different correlative parameters based on the Arrhenius rate
stress direction, remains unchanged in the presence of reinforce- equation [19] which relate temperature and time to rupture
ment. PPE/PS has an amorphous matrix, while PA66, PBT, and PA6 together have been developed over the years. LarsoneMiller [20],
have a semi-crystalline matrix. Also, percent weight of short glass MansoneHaferd [21], SherbyeDorn [22], MansoneSuccop [23],
M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83 71

Fig. 2. Effect of temperature on creep rupture of (a) PO, (b) PP, (c) PP-T, (d) PP-G, (e) PA66, and (f) PPE/PS in the transverse direction and, (g) PBT and, (h) PA6 in both transverse and
longitudinal directions. Lines are fits to experimental data. Arrows show tests that did not fail and are not included in the fits.

and MansoneMuralidharan [24] are some of these parameters. where T is the temperature in  C, tR is time to rupture in hours, and
LarsoneMiller Parameter (LMP) has become the most widely used CLMP is a constant. CLMP value was determined by fitting a line to log
because of its simplicity. Comparison of correlations with different tR versus 1/T data for a given material and stress level. Intersection
parameters did not show considerable differences based on the of the lines for different stress levels at 1/T ¼ 0 defines the value of
experimental data in this study. This parameter is expressed in the constant CLMP.
following form: LarsoneMiller master curves for all the materials are shown
in Fig. 5. As can be seen, the correlations are very good. The
ðT þ 273Þðlog tR þ CLMP Þ power law fit is used to represent master curves in the following
LMP ¼ (4) form:
1000
72 M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83

Fig. 3. Normalized creep rupture data with tensile strength for (a) PO, PP, PP-T, and PP-G and, (b) PA66, PPE/PS, PBT, and PA6 at different temperatures.

Table 2
Material, temperature and fiber orientation dependent tensile properties and modeling constants.

Material Direction Temperature ( C) Su (MPa) E (MPa) A B C D n a b

PO T 23 14.7 2123 11.7 0.076 4E-15 14.3 0.500 2E-6 6.33


T 85 4.1 670 2.7 0.133 4E-5 12.5 0.450 2E-2 6.29
T 125 2.1 350 1.0 0.219 2E1 8.6 0.410 2E1 6.89
PP T 23 22.3 1630 21.8 0.070 1E-15 12.2 0.400 1E-5 4.50
T 85 9.5 566 6.8 0.125 7E-9 10.6 0.420 9E-4 4.69
T 125 5 282 3.1 0.205 3E-5 9.8 0.410 1E-1 3.20
PP-T T 23 29.4 4200 23.3 0.052 4E-23 16.6 0.438 1E-9 6.73
T 85 13.7 1550 9.23 0.051 1E-19 19.9 0.318 1E-5 5.56
PP-G T 23 46.8 3446 37.3 0.062 1E-20 12.9 0.375 5E-8 5.00
T 85 21.7 1516 15.8 0.094 2E-13 11.4 0.600 8E-9 7.55
T 120 11.9 966 9.1 0.116 2E-12 13.3 0.762 1E-8 9.51
PA66 T 23 104.7 5250 91.9 0.026 3E-31 15.4 0.237 1E-11 5.71
T 85 58.8 2250 51.0 0.025 5E-52 30.1 0.182 2E-11 6.62
T 120 46.7 1550 40.0 0.028 8E-34 20.7 0.206 6E-11 6.82
PPE/PS T 23 70.8 4400 58.2 0.027 1E-32 17.9 0.288 2E-15 8.39
T 85 39.5 4000 35.3 0.052 5E-20 12.4 0.343 4E-9 5.54
PBT T 85 35.5 2300 31.2 0.023 5E-63 41.6 e
T 125 26.5 1830 22.0 0.036 2E-35 25.6
L 85 64 4600 55.5 0.017 7E-53 30.0
L 125 48.5 3380 41.8 0.023 1E-59 36.3
PA6 T 85 46.5 1800 38.7 0.049 3E-62 38.4 e
T 125 37 1300 32.3 0.022 3E-53 34.6
M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83 73

Fig. 4. Correlations of (a) A and, (b) B values with Su for all the materials at different temperatures and fiber orientations. The conditions associated with the symbols are the same as
those in Fig. 3.

Fig. 5. LMP master curves for (a) PO, PP, PP-T, and PP-G in the transverse direction and, (b) PA66, PPE/PS, PA6 in the transverse direction and for PBT in both transverse and
longitudinal directions.
74 M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83

Table 3 and PP-G the third stage of creep is an important portion of the
Material and fiber orientation dependent modeling constants. creep curve (see Fig. 6(a)), while for the other materials it is a small
Material Direction CLMP G H l h l0 h0 portion of the creep curve (see Fig. 6(b)). Creep strain of PA6 is
PO T 40 4.92E9 8.03 0.60 4.0 0.90 0.53
considerably higher than PA66, PPE/PS, and PBT (see Fig. 6(b)), due
PP T 31 2.66E7 6.36 0.92 4.5 1.12 0.46 to the effect of elastomer additive (10%wt).
PP-T T 42 4.90E6 4.86 1.00 2.1 0.68 0.27 An increase in temperature results in higher macromolecular
PP-G T 26 1.75E6 5.24 0.83 5.3 0.91 0.50 mobility, therefore, higher creep deformation for all the materials,
PA66 T 55 3.41E5 2.95 1.35 1.7 e e
as expected. An example of this is shown in Fig. 7(a) for PP-G at a
PPE/PS T 32 4.07E4 2.90 1.68 0.4 e e
PBT T 50 4.24E5 3.30 1.00 0.6 e e stress level of 75% of Su. Creep curves of PP-T and PP-G are similar
L 60 2.34E5 2.72 // // at both 23  C and 85  C, as shown in Fig. 7(b). This indicates that
PA6 T 22 1.6E3 1.80 0.82 1.2 e e the type of reinforcement (talc or short glass fiber) has the same
effect on creep deformation of polypropylene, similar to that
observed for creep rupture behavior. Reinforcement or filler does
S ¼ GðLMPÞH (5)
not change the creep mechanism and only results in improved
where S is stress in MPa, and G and H are material constants which creep strength (i.e. reduction of both creep strain and rate), as
are listed in Table 3. The LarsoneMiller creep parameter can be used compared to the neat polymer [26]. Creep strain is higher for the
for long-term creep-rupture predictions at different temperatures, samples tested in the transverse to the mold flow direction, as
provided that stress should be in the range of test stresses [25]. compared to the longitudinal direction, as shown in Fig. 7(c) for
PBT. In addition, as can be observed from Fig. 7(c), the effect of
temperature on creep behavior is more significant in the trans-
4. Creep strain-time behavior and predictions verse direction. This is because of the fact that creep behavior is
more matrix dependent for the samples with transversely ori-
Creep strain versus time curves at 85  C in the transverse di- ented short glass fibers.
rection at several stress levels are shown in Fig. 6. For PO, PP, PP-T

Fig. 6. Creep strain versus time curves at 85  C for (a) PO, PP, PP-T, and PP-G and for, (b) PA66, PPE/PS, PBT, and PA6.
M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83 75

Several empirical power law models have been proposed to Predictive creep curves using Equations (6) and (7) for PP-G at
describe the nonlinear creep deformation behavior of polymeric different temperatures are presented in Fig. 8. It can be seen that
materials. Some of these models were reviewed in Refs. [27] and the Findley model is reasonably accurate in representing the short-
[28]. The power law models have evolved over the years in order to term creep strain curves of PP-G. Similar predictions were observed
be applicable to a wider range of temperatures, stress levels, and for other materials at different stress levels and temperatures.
polymeric materials [29e31]. The Findley power law model [32]
has become a widely used model for describing creep deforma-
tion of polymers and their composites and is given by: 5. Minimum creep rate data and correlations

Minimum creep rate, ε_ min , is the slope of the portion of creep vs.
εc ¼ ε0 þ mt n (6)
time diagram corresponding to secondary creep and is an impor-
where εc is creep strain, n is a temperature dependent material tant parameter widely used for creep design of many components.
constant, and m is both stress and temperature dependent con- Norton power law [38,39] is used to relate minimum creep rate to
stant. m is the intercept and n is the slope of the fitted line in the stress as:
logelog plot of (εc- ε0) versus time. n values are reported in Table 2
for different temperatures with t and εc in hour and %, respectively. ε_ min ¼ CðSÞD (8)
n is nearly constant for PP, independent of temperature, but for talc-
where C and D are material, temperature, and orientation depen-
filled and short glass fiber reinforced materials, larger variation
dent constants, with values reported in Table 2 with S and ε_ min in
with temperature is observed. Temperature independency of n for
MPa and %/hours, respectively. Fig. 9 shows ε_ min versus stress data
neat polymers [33,34] and temperature dependency for filled and
and the corresponding Norton power law fits at different temper-
reinforced polymers [35] have also been reported in the literature.
atures for all the materials in logelog plots. As can be seen, ε_ min
Values of ε0 can be estimated using RembergeOsgood equation,
increases with increasing temperature and stress level for all the
which was used in Refs. [13,36,37] to model the tensile stress-strain
materials. Minimum creep rate versus normalized stress using Su
behavior of the studied materials. In order to find the variation of
data are shown in Fig. 10. The data for PO, PP, PP-T and PP-G are
constant m with stress, the following equation, which was sug-
shown in Fig. 10(a), while the data for PA66, PPE/PS, PBT and PA6
gested by Hadid et al. [28], was used:
are shown in Fig. 10(b). Similar to what was observed for creep
rupture data (Fig. 3), PO, PP, PP-T, and PP-G data at 23  C are
m ¼ a Sb (7) correlated very well by normalizing with tensile strength. Also,
normalizing with tensile strength correlates all the minimum creep
where a and b are material constants and their values are reported rate data of the second group very well, except for PPE/PS data at
in Table 2. 85  C (see Fig. 10(b)).

Fig. 7. (a) Effect of temperature on creep strain curves of PP-G under stress level of 75% of Su. (b) Comparison of PP-T and PP-G creep strain curves at 23  C and 85  C under stress
level of 62% of Su. (c) Fiber orientation effect on creep strain of PBT at 85  C and 125  C under stress level of 80% of Su.
76 M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83

Fig. 8. Comparison of predicted creep strain curves using Findley power law model (dash lines) with experimental creep strain curves of PP-G at different stress levels in the
transverse direction at (a) 23  C, (b) 85  C, and (c) 120  C.

PA66, PPE/PS, PBT, and PA6 exhibit small primary and tertiary 0
creep stages and the secondary creep stage is the main part of the ε_ lmin ðtR =εR Þ ¼ h0 (11)
creep strain curve, particularly for tests conducted at lower stress
levels. Therefore, for these conditions creep strain can be estimated where the material constants l0 and h0 are reported in Table 3 with
with a linear line with its slope and intercept to be ε_ min and ε0, tR and ε_ min in hours and %/hours, respectively. As mentioned earlier,
respectively, where ε0 can be estimated using RambergeOsgood a significant portion of creep curves of PO, PP, PP-T, and PP-G is in
relation explained earlier. the third stage of creep and strain at rupture cannot be obtained
The MonkmaneGrant relationship [40] has been widely applied using minimum creep rate predictions (Equation (9)), therefore,
to metallic materials. It relates time to rupture, tR, to the minimum Equation (11) is valuable when strain at rupture is needed for these
or secondary creep rate, ε_ min , given by: materials.
As can be seen in Fig. 11(b), correlations of data using the
modified MonkmaneGrant relation is very good for PP-G and
ε_ lmin tR ¼ h (9) similar good correlations was also obtained for PO, PP, and PP-T. l0 is
suggested to be unity for most metallic materials [42]. For PO, PP,
where l and h are material constants with their values reported and PP-G, l0 values are near to unity and is 0.68 for PP-T. A single fit
in Table 3 for all the materials, with tR and ε_ min in hours and using Equation (11) for PO, PP, PP-T, and PP-G is shown in Fig. 12(b),
%/hours, respectively. The correlations of PP-G data at 23  C, 85  C expressed by:
and 120  C using this relationship is shown in Fig. 11(a).
As can be seen, the correlation are reasonable and similarly good
correlations were obtained for the other studied materials. l and h ε_ 0:89
min ðtR =εR Þ ¼ 0:41 with R2 ¼ 0:92 (12)
were found to be fiber orientation independent for PBT. Correla-
tions of all data using a single fit by MonkmaneGrant relationship
is shown in Fig. 12(a), expressed as:
6. Long-term creep tests and predictions

ε_ 0:86 with R2 ¼ 0:78 6.1. Time-stress superposition principle for long-term creep
min tR ¼ 2:65 (10)
predictions
A modified version of MonkmaneGrant relation was suggested
by Dobes and Milicka [41] which also includes strain at rupture, εR, Creep tests usually take a long time and are tedious and
in the correlations. This model is suitable for materials exhibiting expensive to conduct, therefore, predictive methods for long-term
small secondary creep and large tertiary creep stage and is repre- creep properties based on short-term data are of great interest.
sented by the following form: Accelerated or short-term tests can be conducted at elevated
M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83 77

Fig. 9. Minimum creep rate versus stress data with corresponding power law fits for (a) PO, (b) PP, (c) PP-T, (d) PP-G, (e) PA66, (f) PPE/PS, (g) PBT, and (h) PA6.
78 M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83

Fig. 10. Minimum creep rate versus normalized stress with tensile strength data for (a) PO, PP, PP-T, and PP-G and, (b) PA66, PPE/PS, PBT, and PA6 at different temperatures.

Fig. 11. Correlation of (a) ε_ min and tR using MonkmaneGrant relation and (b) ε_ min , tR and εR using modified MonkmaneGrant relation, for PP-G at 23, 85 and 120  C.

temperatures or under higher stress levels. Time-temperature su- [5,7,28,43,44] and is expressed in the following form for creep
perposition (TTS) is a well-known procedure frequently applied strain:
either to determine the temperature dependence of the viscoelastic
 
behavior of polymeric materials, or to expand the time regime at a t
given temperature at which the material behavior is studied from a
εðS1 ; tÞ ¼ ε S2 ; (13)
fS
reference condition. This procedure is developed for prediction of
linear viscoelastic properties of homogeneous polymers and may where S1 and S2 are applied stresses and 4S is the stress shift factor.
not be accurate when applied to semi-crystalline polymers [28]. In This method was successfully applied to creep strain data of PO, PP,
addition, the change of temperature causes a variation of the vol- PP-T and PP-G at 23  C and 85  C, while the shifting method was not
ume of the material, and consequently material density change. successful at 120  C for PP-G and 125  C for PO and PP. Therefore, it
Due to these limitations, applying this method to experimental data appears that there is a temperature limit for TSS principle, like the
of thermoplastic composites may result in inaccurate predictions. limit for Williams-Landel-Ferry (WLF) equation which is used to
In an analogous manner to the TTS procedure, a timeestress relate temperature shifting factor to temperature in TTS, and is
superposition (TSS) approach has been developed which is used for valid for the temperature range between Tg and Tgþ100  C [45].
prediction of viscoelastic properties at lower stress levels (long- The creep master curve and corresponding shift factors were
term creep) from shifting the resultant data at higher stress levels obtained by shifting the short-time (about an hour) creep strain
(short-term creep) at a constant temperature. This method was curves along the time axis with respect to the reference curve at
been applied successfully for thermoplastic and their composites each stress level and each temperature. The TSS procedure is shown
M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83 79

Fig. 12. Correlation of ε_ min and tR using MonkmaneGrant relation for (a) all the studied materials. (b) Correlation of ε_ min , tR and εR using modified MonkmaneGrant relation for PO,
PP, PP-T, and PP-G.

in Fig. 13(a) for PP-G at 85  C for a reference stress of 11 MPa. The scatter bands of ±2, indicating the capability of TSS method in
master curves and corresponding shift factors versus stress in a prediction of creep rupture data.
logelog plot are shown in Fig. 13(b) and (c) for PP-G at 23  C and
85  C. A line was fitted to shift factors versus stress data in a logelog 6.2. Experimental results versus predictions
plot, the same as that in the study of Hadid et al. [28].
As mentioned earlier, only the first one hour data of creep strain To evaluate the applicability of predictive models for long time
curves were used to generate the master curves and shift factors. It creep condition, which is of interest in real applications, a limited
can be seen from Fig. 13(d) and (e) that the prediction of creep number of long-term (up to 40 days) creep tests were conducted for
strain curves using TSS are very good and better than the Findley PP-G at 23  C. As shown in Fig. 16(a), 10 days and 40 days creep
model (see Fig. 8) for PP-G at 23  C and 85  C for different stress rupture data of PP-G at 23  C are consistent with the extrapolated
levels. Similar good predictions were obtained for PO, PP, and PP-T. creep rupture curve fitted to short-term creep test data. The Lar-
The creep strain master curves for PO and PP-T at 23  C and 85  C soneMiller model (Equations (4) and (5)) predicts the time to
and for PP at 23  C and the corresponding shift factors versus ruptures at 23  C for PP-G to be 197 and 720 h for the stress levels of
normalized stress with Su data are shown in Fig. 14. It can be seen 25.5 and 23 MPa, respectively. As compared with the experimental
that the shift factor versus normalized stress data are correlated at time to rupture of 273 and 1304 h, the predictions are within a
23  C for PO, PP, PP-T, and PP-G. At 85  C, however, PO data are not factor of less than two. The predicted time to rupture data for the
correlated with PP-T and PP-G. two stress levels using TSS included in Fig. 15, indicates that pre-
Obtained time to rupture of the master curve at the reference dictions for long-term tests are also in the scatter band of two.
stress levels can be shifted using shift factors at each stress level to The TSS method developed using the first one hour creep strain
predict time to ruptures at other stress levels. As can be seen in curves of short-term tests not only can predict the creep strain of
Fig. 15, nearly all the data predicted using this procedure are in the rests of those tests, but can also be used for predictions at other
80 M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83

Fig. 13. (a) TSS procedure for generating master curve and shift factors for PP-G at 85  C. (b) Master curves and (c) corresponding shift factor versus stress plots for PP-G at 23 and
85  C. Comparison of predictive creep strain (dash lines) using TSS and experimental creep strain curves for PP-G for different stress levels at (d) 23  C and (e) 85  C.

stress levels. As can be seen in Fig. 16(b), the creep strain pre- The calculated minimum creep rate of long-term tests from
dictions for long-term tests are very good for both stress levels of experimental data are 0.011 and 0.0021%/hour for tests conducted
25.5 and 23 MPa. For long-term creep predictions, Findley [46] under stress levels of 25.5 and 23 MPa, respectively. Predicted
suggested that the constants of the power law model can be ob- minimum creep rate values from Equation (9) are 0.014 and
tained from the early creep strain data. The Findley power law 0.0037%/hour and from Equation (11) are 0.015 and 0.0013 for tests
predictions were not as good as those based on TSS, however. conducted under stress levels of 25.5 and 23 MPa, respectively.
M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83 81

Fig. 14. Creep strain master curves of (a) PO at 23  C and 85  C, (b) PP at 23  C and, (c) PP-T at 23  C and 85  C. Normalized plot of shift factor versus stress using Su for PO, PP, PP-T,
and PP-G at 23  C and 85  C.

or aging effects can be incorporated in the generated TSS master


curves by additionally shifting the curves according to specimen
aging times [6,7,43].

7. Conclusions

Based on the observed experimental behavior and the analysis


performed, the following conclusions can be made:

1) Temperature significantly affected creep behavior of neat, talc-


filled, and short glass fiber reinforced thermoplastics. For short
fiber reinforced thermoplastics, the effect of temperature was
more accentuated in the transverse direction, as compared with
the longitudinal direction.
2) For polypropylene composites, the type of reinforcement (talc or
short glass fiber) had the same effect on creep deformation.
Reinforcement or filler did not change the creep mechanism
noticeably and only resulted in improved creep strength (i.e.
reduction of both creep strain and rate), as compared to the neat
Fig. 15. Predicted versus experimental time to rupture for PO, PP, PP-T and PP-G at polymer.
23  C and 85  C using the TSS technique. 3) The LarsoneMiller parameter widely used for creep rupture
correlations of metallic materials was used to correlate creep
rupture data of all the materials. This parameter could also be
Comparisons between experimental and predictive minimum used to predict long-term creep rupture data, based on short-
creep rate values show that the minimum creep rate correlations term data.
based on short-term creep data give acceptable predictions for 4) The Findley power law model and time stress superposition
long-term tests. (TSS) method represented creep strain curves well. The TSS
As thermal and hygrothermal aging affect the mechanical method could be used to predict long-term creep curves based
behavior of some polymers and polymer composites in a beneficial on short-term creep data. This method could also satisfactory
or a detrimental way [11], they can cause error for long-term creep predict creep rupture data.
predictions based on short-term creep data. To overcome this issue, 5) Minimum creep rate data were very well correlated by the
short-term creep tests can be conducted on already aged samples MonkmaneGrant relation and its modified version. Using the
82 M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83

Fig. 16. (a) Long-term (up to 40 days) creep rupture data superimposed on short-term (up to 3 days) creep rupture curve and, (b) comparison of experimental long-term creep
strain (data points) and predicted creep strain curve using TSS for PP-G at 23  C.

modified version, creep rupture strain, time to creep rupture, [4] Dean GD, Broughton W. A model for non-linear creep in polypropylene. Polym
Test 2007;26(8):1068e81.
and creep rate are related together.
[5] Luo Wb, Wang Ch, Vu-Khanh T, Jazouli S. Time-stress equivalence: application
to nonlinear creep of polypropylene. J Cent South Univ Technol 2007;14(1):
Acknowledgments 310e3.
[6] Zhao RG, Chen CZ, Li QF, Luo WB. Effects of stress and physical ageing on
nonlinear creep behavior of poly(methyl methacrylate). J Cent South Univ
Funding of this study was provided by General Motors. Technical Technol 2008;15(1):582e8.
support of Dr. A.K. Khosrovaneh and Mr. T. Wang is appreciated. [7] Qaiser A, Price J. Estimation of long-term creep behavior of polycarbonate by
stress-time superposition and effects of physical aging. Mech Time-Depend
Seyyedvahid Mortazavian and Stephen R. Mellott conducted creep Mater 2011;15(1):41e50.
tests of PO, PP, PBT, and PA6. [8] Spathis G, Kontou E. Creep failure time prediction of polymers and polymer
composites. Compos Sci Technol 2012;72(9):959e64.
[9] Teoh SH, Cherry BW, Kausch HH. Creep rupture modelling of polymers. Int J
References Damage Mech 1992;1(2):245e56.
[10] Seltzer R, Mai YW, Frontini PM. Creep behaviour of injection moulded poly-
[1] Drozdov AD. Creep rupture and viscoelastoplasticity of polypropylene. Eng amide 6/organoclay nanocomposites by nanoindentation and cantilever-
Fract Mech 2010;77(12):2277e93. bending. Compos Part B Eng 2012;43(1):83e9.
[2] ASTM D2990-09. Standard test methods for tensile, compressive, and flexural [11] Eftekhari M, Fatemi A. Tensile, creep, and fatigue behaviors of short fiber
creep and creep-rupture of plastics. West Conshohocken, PA: ASTM Interna- reinforced polymer composites at elevated temperatures: a literarture survey.
tional; 2009. Fatigue Fract Eng M 2015;38(12):1395e418.
[3] Hamouda HBH, Simoes-betbeder M, Grillon F, Blouet P, Billon N, Piques R. [12] Mortazavian S, Fatemi A. Effects of fiber orientation and anisotropy on tensile
Creep damage mechanisms in polyethylene gas pipes. Polymer 2001;42(12): strength and elastic modulus of short fiber reinforced polymer composites.
5425e37. Compos Part B Eng 2015;72:116e29.
M. Eftekhari, A. Fatemi / Composites Part B 97 (2016) 68e83 83

[13] Eftekhari M, Fatemi A. Tensile behavior of thermoplastic composites including timeetemperatureestress superposition. Compos Part A Appl S
temperature, moisture, and hygrothermal effects. Polym Test 2016;51: 2009;40(6e7):870e7.
151e64. [30] Hadid M, Rechak S, Zouani A. Empirical nonlinear viscoelastic model for in-
[14] Mortazavian S, Fatemi A. Tensile and fatigue behaviors of polymers for jection molded thermoplastic composite. Polym Compos 2002;23(5):771e8.
automotive applications. Mater Werkst 2015;46(2):204e13. [31] Kouadri-Boudjelthia A, Imad A, Bouabdallah A, Elmeguenni M. Analysis of the
[15] Mortazavian S, Fatemi A. Fatigue behavior and modeling of short fiber rein- effect of temperature on the creep parameters of composite material. Mater
forced polymer composites including anisotropy and temperature effects. Int J Des 2009;30(5):1569e74.
Fatigue 2015;77:12e27. [32] Findley WN. Mechanism and mechanics of creep of plastics. SPEJ 1960;16:
[16] Eftekhari M, Fatemi A. On the strengthening effect of increasing cycling fre- 57e65.
quency on fatigue behavior of some polymers and their composites: experi- [33] Gupta Vidya Bhushan, Lahiri J. Non linear viscoelastic behavior of poly-
ments and modeling. Int J Fatigue 2016;87:153e66. propylene and glass reinforced polypropylene in creep. J Compos Mater
[17] Eftekhari M, Fatemi A, Khosrovaneh A. Fatigue behavior of neat and short 1980;14(4):286e96.
glass fiber reinforced polymers under two-step loadings and periodic over- [34] Lai JSY, Findley WN. Elevated temperature creep of polyurethane under
loads. SAE Int J Mater Manuf 2016;9(3). nonlinear torsional stress with step changes in torque. Trans Soc Rheol
[18] Mellott SR, Fatemi A. Fatigue behavior and modeling of thermoplastics 1973;17(1):129e50 (1957-1977).
including temperature and mean stress effects. Polym Eng Sci 2014;54(3): [35] David W, Scott JSL. Creep behavior of fiber-reinforced polymeric composites: a
725e38. review of the technical literature. J Reinf Plast Comp 1995;14(6):588e617.
[19] Arrhenius SA. Über die Dissociationsw€ arme und den Einflusß der Temperatur [36] Mortazavian S. Fatigue behavior and modeling of short fiber reinforced
auf den Dissociationsgrad der Elektrolyte. Z Phys Chem 1889;4:96e116. polymer composites [Doctoral Desertation]. Toledo, Ohio: University of
[20] Larson FR, Miller J. A time-temperature relationship for rupture and creep Toledo; 2015.
stresses. Trans ASME 1952;74:765e71. [37] Mellott SR. Tensile, creep, and fatigue behaviors of thermoplastics including
[21] Manson SS, Haferd AM. A linear time-temperature relation for extrapolation thickness, mold flow direction, mean stress, temperature, and loading rate
of creep and stress-rupture data. Technical Note 2890. NACA; 1953. effects [Master thesis]. Toledo, Ohio: University of Toledo; 2012.
[22] Sherby OD, Dorn JE. Creep correlations in alpha solid solutions of aluminum. [38] Norton FH. The creep of steel at high temperature. New York: McGraw Hill;
Trans AIME 1952;194. 1929.
[23] Manson SS, Succop G. Stress-rupture properties of inconel 700 alloy and their [39] Pang F, Wang CH. Activation theory for creep of woven composites. Compos
correlation with some time-temperature curve parameters. In: Heat-resisting Part B Eng 1999;30(6):613e20.
metallic materials [Russian translation], IL, Moscow; 1958. [40] Monkman CF, Grant NJ. An empirical relationship between rupture life and
[24] Manson SS, Muralidharan U. Analysis of creep rupture data for five multi-heat minimum creep rate in creep-rupture tests. ASTM Proc 1956;56:593e620.
alloys by the minimum commitment method using double het term centering [41] Dobes F, Milicka K. The relation between minimum creep rate and time to
technique. Palo Alto, CA: EPRI CS 317, Electric Power Research Institute; 1983. fracture. Metal Sci 1976;10:382e4.
[25] Wilson DJ, Marrone RE, Freeman JW. Larson-Miller and Manson-Haferd [42] Povolo F. Comments on the Monkman-Grant and the modified Monkman-
parameter extrapolation of rupture data for type 304 (18Cr-8Ni), grade 22 Grant relationships. J Mater Sci 1985;20(6):2005e10.
(2-1/4Cr-1Mo) and grade 11 (1-1/4Cr-1/2Mo-3/4Si) steels. Ann Arbor, Mich- [43] Lai J, Bakker A. Analysis of the non-linear creep of high-density polyethylene.
igan: The University of Michigan, Office of Research Administration; 1968. Polymer 1995;36(1):93e9.
[26] Raghavan J, Meshii M. Creep of polymer composites. Compos Sci Technol [44] Jazouli S, Luo W, Bremand F, Vu-Khanh T. Application of timeestress equiv-
1998;57(12):1673e88. alence to nonlinear creep of polycarbonate. Polym Test 2005;24(4):463e7.
[27] Plaseied A, Fatemi A. Tensile creep and deformation modeling of vinyl ester [45] Malcolm LW, Landel RF, Ferreira JD. The Temperature dependence of relaxa-
polymer and its nanocomposite. J Reinf Plas Comp 2009;28(14):1775e88. tion mechanisms in amorphous polymers and other glass-forming liquids.
[28] Hadid M, Rechak S, Tati A. Long-term bending creep behavior prediction of J Am Chem Soc 1955;77:3701e7.
injection molded composite using stressetime correspondence principle. [46] Findley WN. 26-Year creep and recovery of poly(vinyl chloride) and poly-
Mater Sci Eng A Struct 2004;385(1e2):54e8. ethylene. Polym Eng Sci 1987;27(8):582e5.
[29] Chevali VS, Dean DR, Janowski GM. Flexural creep behavior of discontinuous
thermoplastic composites: non-linear viscoelastic modeling and

You might also like