Sunteți pe pagina 1din 7

Materials Science and Engineering A 480 (2008) 259265

Creep behaviour of layered silicate/starchpolycaprolactone


blends nanocomposites
C.J. Perez, V.A. Alvarez, A. Vazquez
Research Institute of Material Science and Technology (INTEMA), National University of Mar del Plata (UNMdP),
Av. Juan B. Justo 4302, 7600 Mar del Plata, Argentina
Received 14 May 2007; received in revised form 2 July 2007; accepted 12 July 2007

Abstract
The creep behaviour of biodegradable composites based on starch/polycaprolactone commercial blends reinforced with an organo-modified
nanoclay, processed by melt-intercalation was studied. Clay content and temperature effects were also analyzed. The experimental behaviour was
correlated with several models. The master curves were built by means of the timetemperature superposition principle. Findleys power law
correctly predicted the straintime curves and also the complete compliance curves. The Burgerss model (four parameters) was also used allowing
the relationship between the creep behaviour and the composite morphology. All the results showed that the addition of clay to the neat matrix
leads to a significant improvement of creep resistance.
2007 Elsevier B.V. All rights reserved.
Keywords: Creep; Clay; Nanocomposites; Biodegradable polymers; Modelling

1. Introduction
In the last years, biodegradable polymers have been widely
used for packaging with the emphasis on the reduction of
environmental pollution [1,2]. Even tough biodegradable polymers mean ecological benefits, they are required to have
physicalmechanical properties and cost similar to those of the
traditional plastics. The interest on starch-based polymers is constantly increasing due to its low cost but it is well know that it
properties are poor [3,4]. In order to improve the mechanical
properties, small quantities of layered silicates can be added
producing materials with desired properties [5,6].
Creep behaviour is a very important property of a thermoplastic composite that controls its dimensional stability and
especially in applications where the material has to support loads
for long periods of time [79]. This mechanism can drive to an
unacceptable deformation level and in due course to a structural
failure.
On this field, Vlasveld et al. [10] have investigated the
creep and physical aging behaviour of polyamide six (PA6)
nanocomposites. They demonstrated that the creep compliance

Corresponding author. Tel.: +54 223 4816600.


E-mail address: anvazque@fi.mdp.edu.ar (A. Vazquez).

0921-5093/$ see front matter 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.msea.2007.07.031

was diminished by layered silicate incorporation and related this


with the modulus increment, but the shape of the curves was similar to that of unfilled PA6. In addition, Pegoretti et al. [11] have
considered the effect of nanoparticles on the creep behaviour of
polyethylene terephthalate (PET). They showed that nanoparticles improved the creep resistance of the neat matrix. Galgali
et al. [12] have studied the creep properties of polypropylene (PP)/maleated polypropylene (PP-g-Ma) reinforced with
clays. They established that clay produces an enlarged creep
resistance and drive to higher shear viscosity. Ranade et al.
[13] have analyzed the creep performance of polyethylene
(maleated and non-maleated)montmorillenite layered silicate
blown films. They found a decrease on the creep compliance and
an increase on the modulus by clay addition. Yang et al. [14,15]
have characterized the tensile creep resistance of polyamide 66
(PA66) nanocomposites. They proved that the deformation of
the nanocomposites were clearly lower than that of the pure
matrix showing their improved creep performance. They have
demonstrated that the nanocomposites exhibited higher elasticity and that the nanoreinforcements were able to improve the
polymer elasticity. Chiou et al. [16] have evaluated the rheology of starchclay nanocomposites by using several clays. They
confirmed that the decrease on the creep compliance was evidently related with the polymerclay interaction being higher
in the case of higher compatibility. Martucci et al. [17] have

260

C.J. Perez et al. / Materials Science and Engineering A 480 (2008) 259265

studied creep behaviour of glutaraldehyde-crosslinked gelatine.


They demonstrated that chemical crosslinking drives to a lower
viscous creep and an increment on the elastic part.
Related with starch-based blends, Cyras et al. [9] and Alvarez
et al. [18] have investigated the effect of natural fibres on the
creep behaviour of such kind of blends. Both demonstrated that
the addition of fibres improves the creep resistance of polymeric matrices and this improvement strongly depends on the
fibre content, their dimensions, orientation, distribution, and the
adhesion between the fibres and matrix.
The aim of this work was to investigate the influence of
clay incorporation and clay content on the creep behaviour
of starch/polycaprolactone blend. The creep behaviour of this
material is of crucial interest because it is used above its glass
transition temperature. Several models and equations were used
to predict the long-time response of studied materials.
2. Experimental
2.1. Materials
The matrix used in this work was a commercial
starch/polycaprolactone blend called MaterBi Z [19], kindly
supplied by Novamont, Novara, Italy. An organoclay, under the
commercial name of Cloisite 10A (C10A) was used as nanoreinforcement. It was purchased from Southern Clay Products
Inc., USA. The organoclay was a natural montmorillonite modified by quaternary ammonium salt:

2.3.1. Theoretical background


Creep performance is commonly represented by creep compliance, J(t) which is defined as the ratio between creep strain
(t) and the applied stress () as follows:
(t)
(1)

It is known that creep modulus; defined as the relationship


between the applied stress and the deformation at a given time,
decreases as a function of time and temperature. William, Landel
and Ferry [21] have developed the WLF theory showing that, in
the linear viscoelastic range, a viscoelastic curve determined at
an arbitrary temperature T1 can be carried to another one T2
making an appropriate translation on the time axis by applying
a shift factor aT at a reference temperature Tr . When reference
temperature is far from the glass transition temperature (more
than 50 K) the Arrhenius law is preferred. Based on it, the shift
factor can be described as a function of temperature as follows:


Q
1
1
log aT =
(2)

2.303R T
Tr

J(t) =

where Q is the activation energy and R is the universal gas


constant.
2.3.2. Constitutive model
The viscoeslastic behaviour of thermoplastic composites has
been modelled by using a constitutive model based on a power
law equation known as Findley power law [22]. It can be
expressed as follows:
(t) = 0 + At n

at a concentration: 125 mequiv/100 g of clay. This organomodified clay was chosen due to their compatibility with the
biodegradable matrix, obtained in a previous work [20].
2.2. Composite preparation
Composites were prepared by melt-intercalation followed by
compression-moulding. An intensive Brabender type mixer with
two counter-rotating roller was used. Mixing temperature was
100 C; speed of rotation was 150 rpm and mixing time was
10 min. The concentration of clay ranged from 0 to 7.5 wt.%.
Compression moulding was carried out in a hydraulic press for
10 min at 100 C. The thickness of the samples was between 0.3
and 0.5 mm.
2.3. Methods
Creep tests were conducted in a DMA 7-e PerkinElmer under nitrogen atmosphere. Specimen dimensions were
20 mm 2 mm 0.4 mm, according to ASTM 2990 standards.
Tests were carried out at five different temperatures: 3, 10, 20,
25 and 30 C, all of them above the glass transition temperature of the matrix (66 C). These temperatures were selected
because these are typical temperatures of use. A 3 MPa stress
was applied for 30 min and then it was removed and recovery
measurements were recorded.

(3)

where (t) is creep strain at time t, 0 the instantaneous initial


strain, A the amplitude of transient creep strain and n is the time
exponent.
The instantaneous strain, the amplitude of transient creep
strain, and the time exponent are defined as creep parameters.
The relative creep strain, r , is defined as the ratio between the
creep strain value at each time, (t) , and the instantaneous initial
strain value, 0 . Rearranging Eq. (3):
r = A t n

(4)

where r is the relative strain creep in percentage and A is the


slope of the power law. A and n are parameters that can be
obtained by plotting log r versus log t; log A is obtained from
the intersection with log r axe and n from the slope of the curve.
Equation (3) can be also expressed in terms of creep compliance given by the following equation:
J(t) = J0 + J1 t n

(5)

where J0 is the time-independent compliance, J1 is the coefficient of time-dependent term and n is a constant, generally lower
than one, independent on the applied stress.
Differentiating Eq. (3), it is possible to obtain the power law
creep rate, i.e. PL , as follows:
PL =

nA
t 1n

(6)

C.J. Perez et al. / Materials Science and Engineering A 480 (2008) 259265

261

2.3.3. Four parameters model


Four-parameters (or Burgers) model [23], gives the relationship between the morphology of the composites and their creep
behaviour. This model is a series combination of the Maxwell
and KelvinVoigt model [24], the total strain is given by the
general equation:
= 1 + 2 + 3




tE2

=
1 exp
+ t
+
E1
E2
2
1

(7)
(8)

where 1 and 3 are the elastic and viscous strains represented


by Maxwell model and 2 is the viscoelastic strain represented
by the KelvinVoigt model. E1 and E2 are elastic moduli, 2 and
1 are viscosities, the applied stress and t is the creep time.
This model has only one relaxation time, 2 = 2 /E2 . When a
constant load is applied, the initial deformation takes place in
the spring with the modulus E1 . Later deformation comes from
the spring E2 and dashpot 2 , in parallel, and from the dashpot
with the viscosity 1 . In the recovery test, after all the load is
removed at 30 min, the creep is all recoverable except for the
flow that occurred in the dashpot with viscosity 1 .
From this model, it is also possible to estimate the Burgers
creep rate, i.e. B , by differencing Eq. (8). In this case, the creep
rate can be calculated as:



tE2
(9)
B =
+
exp
1
2
2

Fig. 1. Strain (percentage)time curves for the neat matrix (MaterBi Z) at several
temperatures.

At very long time, when creep rate get to a constant value, it is


given by:
B =

(10)

3. Results and discussion


Fig. 1 shows the strain (percentage)time curves for the neat
matrix at several temperatures. In these curves, the creep stages
(instantaneous deformation, primary and secondary creeps) are
evident. On the other hand, there is no evidence of tertiary creep,
i.e. creep rupture, which would require longer times. As it was
expected, strain increases with temperature. Similar behaviour
was displayed by the nanocomposites. Fig. 2 shows the same
curves but at room temperature for the matrix and nanocomposites with different clay contents. It is clear from this figure
that the strain of nanocomposites is lower than that of the neat
matrix and this implies that the creep behaviour was improved
by the presence of organoclay nanoparticles. In order to show
the effect of nanoparticles on the deformation of the pure blend
at different temperatures, Fig. 3a and b summarizes the initial
(instantaneous) and total (after 30 min of applied stress) deformations. Both deformations are noticeably less in the case of
nanocomposites and they are reduced when clay concentration
is increased from 1.0 to 7.5 wt.%. In addition the instantaneous
deformation as well as the total deformation exhibited almost
an exponential variation with the temperature.
The parameters of Findley power law obtained for the matrix
and nanocomposites are showed on Table 1. A parameter was

Fig. 2. Strain (percentage)time curves for the neat matrix (MaterBi Z) and their
nanocomposites at room temperature and different clay contents.

calculated from A and 0 and the obtained values are resumed


on Fig. 4. It is apparent that the parameter increased with temperature and decreased with clay content which indicates an
enhanced creep performance. On the other hand, n was around
0.113 0.015 and it was not affected by nanoclay incorporation. Similar behaviour was observed by Yang et al. [15] in the
Table 1
Average parameters obtained from Findley power law for the matrix (MaterBi
Z) and their nanocomposites with clay
Clay content (wt.%)

A

0.0
1.0
2.5
5.0
7.5

0.661
0.596
0.610
0.599
0.600

0.084
0.127
0.058
0.130
0.070

0.1010
0.1303
0.1046
0.1225
0.1260

0.006
0.025
0.010
0.013
0.014

262

C.J. Perez et al. / Materials Science and Engineering A 480 (2008) 259265

Fig. 5. E1 parameter of matrix (MaterBi Z) and their nanocomposites as a


function of temperature for different clay content.

Fig. 3. (a) Initial or instantaneous deformation (0 ) for matrix (MaterBi Z) and


their nanocomposites as function of temperature, for different clay content. (b)
Final, after 30 min of applied stress (30 min ) deformation for matrix (MaterBi Z)
and their nanocomposites as a function of temperature, for different clay content.

case of polyamide 66 nanocomposites. 0 parameter, which is


related with the instantaneous deformation, decreased with clay
incorporation, showing an improvement on the elastic part.
Experimental curves were fitted by means of Burgers model.
The non-linear curve fit function of the OriginPro 7.5 software
was used and the four parameters (E1 , E2 , 2 and 1 ) were
estimated. The correlation between the experimental data and
model prediction was really good. Fig. 5 summarizes the values
of E1 parameter for the matrix and nanocomposites as a function of temperature. This parameter (associated to the Maxwell
spring) established the instantaneous creep strain that would be
recovered after stress elimination. It is evident that nanocomposites have higher E1 values (increasing with clay content). E1
of each kind of material displayed a decreasing trend with the
temperature. This is related with the softening of the material
at high temperatures that drives to a decrease on the stiffness.
For the range of temperatures studied in the present work, a
pseudo-linear relationship between E1 and temperature can be
established as it is shown on Table 2. The results show that,
although E1 was higher for the nanocomposites in the complete range of temperatures showing that C10A particles were
able to improve the elasticity of the neat matrix, the effect of
nanoparticles on the stiffness (or elastic part) became lower as
temperature is raised. At lower temperatures, big changes in E1
were observed when organoclay was incorporated to starch/PCL
blend, which in turns indicates that instantaneous elasticity modulus was improved by the presence of the nanofillers. The last
Table 2
Relationship between E1 and the temperature for the matrix (MaterBi Z) and
their nanocomposites with clay
Clay content (wt.%)
0.0
1.0
2.5
5.0
7.5

Fig. 4. A parameter of matrix (MaterBi Z) and their nanocomposites as a function


of temperature for different clay content.

r = regression coefficient.

Modulus as a function of temperature


(MPa) = 300.4 4.24 T( C)

E
E (MPa) = 733.7 16.9 T( C)
E (MPa) = 873.6 20.9 T( C)
E (MPa) = 879.2 17.5 T( C)
E (MPa) = 1222.1 23.6 T( C)

r
0.980
0.996
0.993
0.977
0.996

C.J. Perez et al. / Materials Science and Engineering A 480 (2008) 259265

result is in accordance with the increase in Youngs modulus as


the clay concentration increased previously reported [20]. Similar trend on increasing E1 indicating the enhancement of the
tensile modulus was found by other authors [13,15]. It is important to note that all selected temperatures were higher to the glass
transition temperature and, so that, the stiffness of the material
is very low.
The retardant elasticity (E2 ), which is related to the stiffness
of amorphous polymer chains in the short time, decreased as a
function of temperature, as is shown on Fig. 6. This implies that
the deformation in the Kelvin unit turn into higher by the effect of
the temperature. Another important effect was the reinforcement
of the nanofillers on the Kelvin unit.
The other central parameter is 1 which represents the
irrecoverable creep strain. This parameter is plotted as a function of temperature on Fig. 7. Cyras et al. [9] have shown for
the same matrix that the viscosity 1 increased with sisal fibre
content and lower flow was occurred at the dashpot and the
permanent deformation decreased. Above the glass transition
temperature, as in this case, the amorphous chains have noteworthy mobility that encourages irreversible creep deformations.
Other authors [15] have pointed out that 1 could be related
with the damage of crystalline polymer or oriented noncrystalline regions. This process includes chains pulling out, crystal
slipping and irreversible deformation from amorphous regions.
An improved deformation resistance of nanocomposites, i.e.,
the decrease on the permanent creep deformation by nanofillers
addition, should be associated with an alteration of the crystalline region morphology of the neat matrix, as it was observed
by the Avramis exponent of nanocomposites as compared with
the pure starch/PCL blend in a previous work [25] and also with
the immobilization of polymer chains previously establish in
the same work. Increasing temperature, a diminished 1 was
observed. At higher temperatures, polymer chains were thermally activated and large deformation of crystallized regions
as well as irreversible transitions of amorphous regions took
place. Creep data provides information about zero-shear viscosity (associated with 1 ) which was higher for nanocomposites

263

Fig. 7. 1 parameter of matrix (MaterBi Z) and their nanocomposites as a


function of temperature, for different clay content.

than for the neat matrix. There exist two possible explanations
for this behaviour: the larger barrier for flow of confined chains
or the frictional interactions between the anisotropic clay crystallites.
As it is known, elastic materials exhibit no residual deformation after the stress is removed. Opposite, the viscoelastic
one, display a relaxation time. This time is defined as the time
needed to produce 63.2% of the total deformation in the Kelvin
unit (1 e1 ). Table 3 shows the average relaxation time and the
value at room temperature (20 C) for the matrix and nanocomposites; 2 is also included in this table. Both parameters, and
2 seems to decrease with clay incorporation and then, with clay
content. Neither the relaxation time nor the retardant viscosity
displayed a clear tendency with the temperature.
The power law (Findley) and the four parameters model
(Burgers) were suitable to reproduce the strain data. Besides
their parameters, another characteristic factor is the creep rate
which also determines the dimensional stability of materials.
Creep rates (directly fitted from the linear part of experimental curves in the secondary creep stage) are shown on Fig. 8.
Creep rate constantly increased with temperature because more
polymer segments can be thermally activated at elevated temperatures but, nevertheless the temperature, creep rate decreased
with clay content indicating that the used organoclay is a
good option to enhance the dimensional stability of the pure
starch/polycaprolactone blend. Zhang et al. [26] have observed
the same tendency for PA66-TiO2 nanocomposites. The comparTable 3
Relaxation times and 2 parameter of matrix (MaterBi Z) and their nanocomposites with clay

Fig. 6. E2 parameter of matrix (MaterBi Z) and their nanocomposites as a


function of temperature, for different clay content.

Clay content (wt.%)

average (s)

0.0
1.0
2.5
5.0
7.5

1061
1020
1000
722
674

200
300
79
90
163

room T (s)
1680
1368
1091
938
617

2 (GPa s)
966
820
787
633
616

264

C.J. Perez et al. / Materials Science and Engineering A 480 (2008) 259265

Fig. 8. Creep rates of matrix (MaterBi Z) and their nanocomposites as a function


of temperature, for different clay content.
Table 4
Creep rates (experimental and models) of matrix (MaterBi Z) and their nanocomposites with different content of clay
Clay content (wt.%)

experimental (s1 )

B ( 106 s1 )

PL ( 107 s1 )

0.0
1.0
2.5
5.0
7.5

1.1 106
5.4 107
8.8 107
7.2 107
6.5 107

3.17
3.13
2.98
2.24
1.25

8.69
7.12
7.01
6.91
5.48

ison of experimental values and thus predicted from the two used
models (at room temperature) are summarized on Table 4. As a
general rule, B was higher than PL for all kind of materials and
the experimental values were closer to power law values, PL .
B was higher than that directly obtained from the straintime
curve indicating that this model was not able to reproduce the
experimental data of all studied materials but the power law
(Findleys model) was applicable to materials that exhibited a
non-pronounced second creep stage [15].
The creep strains measured at different temperatures
were superposed rearranging the time scale by using the
timetemperature superposition principle. The activation energies are given on Table 5 and the obtained master curves are
shown on Fig. 9a. Master curves were obtained shifting the modulus/compliance data at different temperature, taking 20 C as
the reference temperature because the materials are commonly
Table 5
Activation energies for creep process and Findleys power law parameters (compliance) of matrix (MaterBi Z) and their nanocomposites with different content
of clay
Clay content (wt.%)

Q (kJ/mol)

J0 (GPa1 )

J1 ( 104 GPa1 )

0.0
1.0
2.5
5.0
7.5

133.2
189.7
256.1
236.8
211.1

2.31 103
1.05 103
7.45 104
5.94 104
5.56 104

8.66
7.08
6.16
6.03
4.09

Fig. 9. (a) Master curves of matrix (MaterBi Z) and their nanocomposites


obtained by using the timetemperature superposition principle and (b) compliance as a function of corrected time for matrix (MaterBi Z) and their
nanocomposites, for different clay content.

used at this temperature. The shape of the curves was similar


for un-reinforced and reinforced materials. The quality of the
master curve superposition is acceptable whereas the shift factor exhibited a good linear dependence with the temperature
and the activation energy values showed that nanocomposites
displayed higher creep resistance under the effect of temperature; i.e., that the nanofillers immobilize the polymer chains.
Eq. (5) was used to model the complete curve. An average
value of 0.18 0.03 was obtained for n. The values of the timeindependent compliance and the coefficient of time-dependent
term are summarized on Table 5 and the correlation between
experimental data and model prediction are shown on Fig. 9b.
As expected, both parameters, J0 and J1 , decreased with clay
content showing the enhancement on the creep stiffness. The
lower initial creep compliance is related to the higher modulus
with increasing layered-silicate concentration. Similar results
were observed by Vlasveld et al. [10] for PA6 nanocomposites.
The rise of compliance with the time was clearly diminished by
clay incorporation. In addition, from the last graph, it is quite

C.J. Perez et al. / Materials Science and Engineering A 480 (2008) 259265

evident that the power law correctly predicts the compliance


behaviour of the matrix and nanocomposites for long times.
4. Conclusions
The creep behaviour of polycaprolactone/starch blends and
its nanocomposites with organo-modified layered silicates was
analyzed.
The creep compliance was decreased by the incorporation
of nanofillers. This effect was related with the enhancement of
the modulus and it was related with the mechanical properties
studied in a previous work. In addition, the reduction in the
creep strain due to clay addition drives to a material with higher
dimensional stability. Other creep parameters, such as creep rate,
also demonstrated this tendency.
The parameters of Burgers model also confirmed this trend:
the instantaneous and retardant modulus increased and the permanent viscosity and relaxation time decreased as a function
of clay content. The effect of temperature on such parameters was established showing the lower stiffness and the higher
deformability of matrix and nanocomposites as a function of it.
The efficiency of nanofillers to get better the creep performance of neat matrix was confirmed by the parameters
of Findleys power law which properly represent the real
straintime curves of matrix and nanocomposites in the selected
range of temperatures.
The master curves displayed a tolerable superposition and
the activation energy for the shift factor was also an evidence
for the creep resistance development.
All the results discussed in the present work indicate that
nanofillers contributed to an improvement on the creep resistance which is an important result for the application point of
view.
It is important to point out that, for the range of clay content studied here, the creep resistance is continuously improved
together with the increase of the clay concentration.
Acknowledgements
Authors acknowledged to CONICET (PIP 6254) and
ANPCyT, Project FONCyT-PICT No. 12-15074 and PICT 2225766 for the financial support of this project.

265

References
[1] Y. Orhan, H. Buyukgungor, Int. Biodeterior. Biodegrad. 45 (2000) 49
55.
[2] H.C. Flemming, Polym. Degrad. Stabil. 59 (1998) 309315.
[3] C. De Kesel, C. Vander Wauven, C. David, Polym. Degrad. Stabil. 55 (1997)
107113.
[4] R.P. Singh, J.K. Pandey, D. Rutot, P. Degee, P. Dubois, Carbohydr. Res.
338 (2003) 17591769.
[5] M. Alexandre, P. Dubois, Mat. Sci. Eng. R 28 (2000) 163.
[6] S. Sinha Ray, M. Bousmina, Prog. Mater. Sci. 50 (2005) 9621079.
[7] B.D. Park, J.J. Balatinecz, Polym. Compos. 19 (1998) 377382.
[8] A. Vazquez, V. Dominguez, J.M. Kenny, J. Thermoplast. Compos. Mater.
12 (1999) 477497.
[9] V.P. Cyras, J.F. Martucci, S. Iannace, A. Vazquez, J. Thermoplast. Compos.
Mater. 15 (2002) 253266.
[10] D.P.N. Vlasveld, H.E.N. Bersee, S.J. Picken, Polymer 46 (2005)
1253912545.
[11] A. Pegoretti, J. Kolarik, C. Peroni, C. Migliaresi, Polymer 45 (2004)
27512758.
[12] G. Galgali, C. Ramesh, A. Lele, Macromolecules 34 (2001) 852856.
[13] A. Ranade, K. Nayak, D. Fairbrother, N. Dsouza, Polymer 46 (2005)
73237333.
[14] J.J. Yang, Z. Zhang, A. Scharb, K. Friedrich, Polymer 47 (2006)
27912801.
[15] J.J. Yang, Z. Zhang, A. Scharb, K. Friedrich, Polymer 47 (2006)
67456758.
[16] B.S. Chiou, E. Yee, G. Glenn, W. Orts, Carbohydr. Polym. 59 (2005)
467475.
[17] J.F. Martucci, R. Ruseckaite, A. Vazquez, Mater. Sci. Eng. A 435436
(2006) 681686.
[18] V.A. Alvarez, J.M. Kenny, A. Vazquez, Polym. Compos. 25 (2004)
280288.
[19] C. Bastioli, Polym. Degrad. Stabil. 59 (1998) 263273.
[20] C.J. Perez, V.A. Alvarez, A. Vazquez, Polym. Int. 56 (2007) 686693.
[21] J.D. Ferry, Viscoelastic Properties of Polymers, F. Wiley and Sons, New
York, l961.
[22] W.N. Findley, J.S. Lai, K. Onaran, Creep and Relaxation of Nonlinear
Viscoelastic Materials with an Introduction to Linear Viscoelasticity, Dover
Publications Inx, New York, 1989.
[23] L.S. Nielsen, R.F. Landel, in: L.L. Faulkner (Ed.), Mechanical Properties of Polymers and Composites, Marcel Dekker Inc., New York,
1994.
[24] C. Marais, G. Villoutreix, J. Appl. Polym. Sci. 69 (1998) 1983
1991.
[25] C.J. Perez, A. Vazquez, V.A. Alvarez, J. Therm. Anal. Calorim.,
in press.
[26] Z. Zhang, J.L. Yang, K. Friedrich, Polymer 45 (2004) 34813485.

S-ar putea să vă placă și