Sei sulla pagina 1di 7

Composites Science and Technology 161 (2018) 85–91

Contents lists available at ScienceDirect

Composites Science and Technology


journal homepage: www.elsevier.com/locate/compscitech

Predicting the stress relaxation behavior of glass-fiber reinforced T


polypropylene composites
Numaira Obaida, Mark T. Kortschota,∗, Mohini Sainb
a
Department of Chemical Engineering and Applied Chemistry, University of Toronto, 200 College Street, Toronto, Ontario M5S 3E5, Canada
b
Faculty of Forestry, Centre for Biocomposites and Biomaterial Processing, University of Toronto, 33 Willcocks Street, Toronto, Ontario M5S 3B3, Canada

A R T I C LE I N FO A B S T R A C T

Keywords: It is well established that the addition of short elastic fibers slows the relaxation process in composites, but this
Polymer-matrix composites (PMCs) phenomenon is not well-understood. Our recent study explained changes in the stress relaxation constant by
Short-fiber composites accounting for the time-dependent interfacial shear stress transfer at the fiber-matrix interface. An analytical
Stress relaxation model was developed and was successfully compared to finite-element experiments. This approach represents a
Modelling
significant departure from the previously published literature, where the effect of fibers on viscoelasticity was
Stress transfer
typically attributed to changes in the covalent bonds at the fiber-matrix interface. In the present study, the stress
relaxation behavior of glass fiber-reinforced polypropylene composites was experimentally measured and
compared to analytical model predictions. Further, the effect of additional covalent bonding at the fiber-matrix
interface was studied experimentally by introducing an interfacial coupling agent. Good agreement was obtained
between the experimental data and the analytical model and it was concluded that most of the stress relaxation
behavior of a composite can be predicted using a model that only accounts for the time-dependent matrix
modulus and a time-dependent shear stress transfer efficiency.

1. Introduction observed to slow the rate of stress relaxation of a composite [3–7]. This
has been a subject of interest because elastic fibers do not exhibit time-
Short, elastic fibers are routinely incorporated in polymers to im- dependent behavior, but nevertheless appear to do so when embedded
prove mechanical properties such as modulus and strength. The static in a viscoelastic matrix. The explanation for this observation remains
properties of polymers and their composites are relatively simple to elusive and previous studies in literature have focused on attributing
understand: the addition of a stiff, elastic reinforcing phase into a softer the effect of fibers on stress relaxation to chemical bonding at the fiber/
polymer matrix typically increases both stiffness and strength. There is matrix interface [8–12]. These studies have proposed that covalent
a substantial body of theory to predict these mechanical properties interfacial bonds between the two phases inhibit polymer mobility and
based on the size, shape, and orientation of the reinforcing phase [1,2]. thus slow stress relaxation.
The viscoelastic properties of composites are significantly more In a previous paper, we proposed a novel explanation for the effect
complex. The time-dependent properties of the matrix makes polymer- of elastic fibers on the stress relaxation of polymer matrix composites
based composites prone to creep and stress relaxation, which is a [13]. We proposed that the time-dependent shear stress transfer at the
challenge when considering composites for long-term applications. A fiber-matrix interface, and not increased covalent bonding at the in-
better understanding of composite viscoelasticity is needed in order to terface, was primarily responsible for altering the non-linear viscoe-
provide guidance for optimizing composite structure, and for predicting lasticity of the composite. In that study, a quantitative model was de-
long-term properties. veloped based on composite micromechanics, and was used to predict
Stress relaxation is a straightforward way of characterising polymer the stress relaxation of composites without postulating changes in
viscoelasticity. Since all viscoelastic properties stem from the same structure near the fiber interface.
basic mechanisms, a model for composite stress relaxation would also There have been other attempts to model the behavior of viscoe-
provide insight into composite creep and dynamic mechanical beha- lastic and viscoplastic matrices reinforced with elastic fibers. Several
vior. studies have used a tensor approach to model the full stress field around
The addition of short fibers in a composite has been repeatedly elastic inclusions in viscoelastic matrices [14–16]. This approach,


Corresponding author.
E-mail address: mark.kortschot@utoronto.ca (M.T. Kortschot).

https://doi.org/10.1016/j.compscitech.2018.04.004
Received 18 November 2017; Received in revised form 1 April 2018; Accepted 4 April 2018
Available online 05 April 2018
0266-3538/ © 2018 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/BY/4.0/).
N. Obaid et al. Composites Science and Technology 161 (2018) 85–91

though accurate, does not provide a simple analytical model useful for Table 2
predicting stress relaxation in a short fiber reinforced polymer as a Experimental design for this study.
function of fiber loading and aspect ratio. Sample PP GF MAPP Replicates
There is also an extensive body of literature for high volume fraction
continuous fiber composites, where at high stress, fiber breakage results Set 1: Samples without MAPP S1 100% – – 5
S2 90% 10% – 5
in concentrated matrix shear stresses near the end of the broken fibre,
S3 80% 20% – 3
and the decay of these stresses with time has been modelled [17–19]. S4 70% 30% – 4
Solutions to such problems tend to be quite specific for the particulars
of the geometry chosen [20,21], although the conceptual underpinning Set 2: Samples with MAPP S5 95% – 5% 5
is the same. Fibre fracture can be modelled with a Weibull approach S6 85% 10% 5% 5
S7 75% 20% 5% 5
[22–25], and understanding the causes and evolution of fibre fractures
S8 65% 30% 5% 5
is often the focus of these studies.
The purpose of the present study was to compare the experimental
stress relaxation behavior of glass fiber-reinforced polypropylene 2.3. Fiber characterization
composites to predictions from a simplified analytical model.
Additionally, the hypothesis that there is additional covalent bonding at The fiber orientation in the samples was examined via x-ray tomo-
the fiber-matrix interface on the stress relaxation behavior of compo- graphy. The post-processing aspect ratio of the fibers was measured by
sites was examined experimentally. examining 100 fibers for each fiber content after a matrix burnout at
600 °C for 6 h. The post-processing fiber aspect ratio was not observed
to vary with fiber content significantly.
2. Experimental
2.4. Stress relaxation tests
2.1. Sample preparation
The stress relaxation samples were milled from a region near the
Two batches of glass fiber reinforced polypropylene composites edge of the injection molded flexural modulus samples where the fibers
were prepared: one without any compatibilizer or coupling agent and were well oriented. The final samples had an approximate dimension of
another containing maleic anhydride-grafted polypropylene (MAPP, 3.5 mm (W), 1.3 mm (T) and 26 mm (L). The stress relaxation behavior
5% by weight) to improve interfacial bonding. Polypropylene (with and was evaluated using a TA Q800 Dynamic Mechanical Analyser (DMA)
without MAPP) was first melted in a C.W. Brabender Compounder with a tensile clamp. The tests were conducted at 30 °C with a soaking
(Type R.E.E.6) at 185 °C and 20 RPM for 10 min. Glass fibers were time of 150 min to ensure that the entire thickness of the sample had
added to the mixer and the mixing speed was increased to 60 RPM, and reached the target temperature, followed by the application of a con-
the compound was mixed for five additional minutes. The details of all stant strain of 0.05% for 90 min.
the chemical components used in this study are summarized in Table 1.
The compounded mixture was granulated using a C.W. Brabender
3. Analytical model
Granulator (Model S9-10). The samples were prepared via an Engel ES-
28 injection molder with an injection temperature of 210 °C, an injec-
The analytical model proposed previously was based on the shear-
tion time of 8 s, a cooling time of 38 s, and a mold opening time of 2 s.
lag theory, as defined by Cox [26]. Although the theory has been ex-
The injection mold produced ASTM tensile, flexural and impact test
tensively used to understand and predict the elastic properties of
specimens.
composites, it had not been used to predict their stress relaxation be-
havior before our work [13]. The previously developed analytical
model can be best understood by first examining the elastic case pre-
2.2. Experimental design
sented in Cox's shear-lag theory.
The composites examined in this study consisted of 10%, 20%, and
30% mass fraction of fibers (5%, 10%, and 15% by volume). Table 2 3.1. Elastic properties
summarizes all the samples that were prepared for this study and the
number of replicates conducted for each sample. The stress-relaxation In short-fiber composites, the fiber end faces are assumed to carry
behavior of the composites without MAPP was compared to that of neat no load, and the stress-transfer between the fiber and matrix occurs
polypropylene without MAPP (S1). Similarly, the stress relaxation be- through shear stresses stemming from the significant mismatch in the
havior of composites with MAPP was compared to neat polypropylene modulus of the matrix and fiber. Upon tensile loading the matrix re-
containing MAPP (S5). A comparison between samples S1 and S5 (see mote from the fiber is assumed to experience the global strain applied
Table 2) was also conducted to determine whether or not the presence to the composite. However, the matrix adjacent to the fiber is assumed
of MAPP resulted in any changes to the properties of the base matrix to be bonded to the fiber and thus constrained, creating a gradient of
itself. stress within the matrix, and shear stresses at the interface (Fig. 1).
Based on the ratio of the matrix and fiber moduli, a fraction of the fiber
will experience maximum tensile strain while a fraction near the ends
will experience both shear strain and tensile strain; the tensile strain
Table 1
increases inwards from the fiber ends. Accordingly, only a fraction of
Details of the chemical components used in the study.
the fiber length will experience maximum tensile stresses, while a
Component Company Product Name fraction near the ends of the fiber will experience tensile stresses less
Polypropylene Total Petrochemicals PP3622
than the maximum. The total contribution of a fiber to the composite
Inc. modulus depends on the average tensile stress over its entire length.
E-Glass Fibers Johns Manville StarStrain EC14
738
3.2. Stress relaxation behavior
Maleic anhydride grafted Eastman G-3003 Polymer
polypropylene (MAPP)
In the case of a viscoelastic matrix, where the Young's modulus of

86
N. Obaid et al. Composites Science and Technology 161 (2018) 85–91

(summarized in Equations (1) and (2)) [13].

tanh (n (t ) s ) ⎞
Ec (t ) = Vm Em (t ) + Vf Ef ⎛1 −
⎜ ⎟

⎝ n (t ) s ⎠ (1)

1
2
⎡ ⎤
⎢ 4 ⎥ 1
n (t ) = ⎢ ⎥ [Gm (t )] 2
⎢ Ef ln

() Pf
Vf ⎥
⎦ (2)

Here, the relaxed modulus of the composite at any point in time is


related to the matrix and fiber moduli (Em (t ) and Ef ) scaled by their
respective volume fractions (Vm and Vf ). The time-dependence of the
elastic modulus of the matrix has significant influence on the time-de-
pendence of the composite. However, our model highlights the im-
portance of incorporating a time-dependent reinforcement effectiveness
factor (n (t ) ), which defines the rate at which the stress is transferred
Fig. 1. Shear stresses in a short-fiber composite. between the matrix and fiber. The time dependence of this factor stems
from the time-dependence of the matrix shear modulus (Gm (t )) (see Eq.
the matrix decreases with time, there is an increase in the mismatch (2)).
between the fiber and the matrix moduli. At the same time, the matrix As mentioned previously, there is a significant body of literature
shear modulus must also decrease with time, decreasing the shear presenting a full tensor analysis of viscoelastic materials containing
stresses at the interface, causing the tensile strain and stresses in the elastic inclusions, which are often ellipsoidal bodies for computational
fiber to decrease with time as illustrated in Fig. 2. reasons. While these analyses do, in principle, explicitly account for
In a short-fiber composite, the elastic fibers can, in this way, exhibit changes in shear stress transfer, they do not, unfortunately, provide
a decrease in their effective reinforcement factor with time when em- much physical insight into the mechanics of the process, nor do they
bedded in a viscoelastic matrix, although this pseudo time-dependence lend themselves to experimental verification. The advantage of the
would not be observed if the elastic fibers were tested directly. Our simplified shear lag model used in the current work is that it clearly
model predicts that these underlying changes in the matrix shear identifies the time dependent stress transfer efficiency factor as the
modulus are responsible for the changes in stress relaxation rate in- underlying cause of the changes in stress relaxation behavior with fiber
duced by the addition of elastic fibers in a composite. An analytical loading. If the model is validated by experiments, then it can be used to
model based on these ideas was derived by incorporating the time-de- make very practical predictions, such as the dependence of stress re-
pendent shear modulus of the matrix into Cox's shear-lag equation laxation on particle aspect ratio.

Fig. 2. In a stress relaxation test, stresses in the fiber decrease over time due to the decay in the matrix modulus during stress relaxation.

87
N. Obaid et al. Composites Science and Technology 161 (2018) 85–91

Fig. 3. The glass fibers were well-oriented within the matrix in the direction of loading (Vf = 5%)

Fig. 4. Post-processing fiber aspect ratios were measured via matrix burnout; approximately 100 fibers were measured at each fiber content.

3.3. Fiber orientation and aspect ratio

Since the analytical model assumes the fibers to be completely or-


iented in the loading direction, it was important to validate this as-
sumption for accurate comparison between the experimental results
and the model. X-ray tomography of our samples indicated the fibers to
be well-oriented in the loading direction (See Fig. 3). It was also ob-
served that the fibers were well-dispersed within the polypropylene
matrix.
High shear-forces during compounding and injection molding are
expected to cause fiber degradation. Since the fiber aspect ratio has
significant influence on the expected properties of the composite,
comparison with the analytical model requires accurate measurement
of the average post-processing fiber aspect ratio (shown in Fig. 4).

3.4. Effect of fiber content on stress relaxation behavior of PP (without


MAPP)

The overall stress relaxation of the composites is shown in Fig. 5.


The addition of elastic fibers increases the absolute modulus of the
Fig. 5. Stress relaxation behavior of PP/GF composites reinforced with various
composite at all time periods, which is expected due to the high mod- fiber volume fractions.
ulus of the fibers. It is also important to observe from Fig. 5 that the
fibers not only improve the absolute values of modulus but also delay
analytical model derived in Equations (1) and (2), where Eq. (2) in-
stress relaxation, which agrees with past experimental studies. This is
corporates the effect of changing Gm on the reinforcement efficiency
not predicted by the Rule of Mixtures equation (Equation (1)) if only
factor (Fig. 6). The model predictions were calculated using the post-
the Em is a function of t but the reinforcement efficiency factor is
processing aspect ratios measured via matrix burnout. The relaxing
treated as a constant, but it is predicted if the reinforcement efficiency
shear modulus of the matrix (Gm (t ) ) was calculated using the experi-
is also treated as a function of time.
mentally-measured relaxing elastic modulus (Em (t ) ) of the neat matrix
The experimental data were compared to predictions from the

88
N. Obaid et al. Composites Science and Technology 161 (2018) 85–91

and by assuming an isotropic material with a Poisson's ratio of 0.3.


Fig. 6 shows that the experimental data aligns quite well with the
model predictions. It is important to note that the analytical model is
based on easily measurable properties of the matrix and fiber para-
meters without any curve-fitting parameters. Based on only these
properties, it is remarkable that such a good fit between the experi-
mental data and the analytical model was obtained. The strong fit be-
tween the experimental data and the model predictions at various fiber
contents provides evidence that the underlying assumptions of the
analytical model must be correct, and it is apparent that the influence of
fibers on the stress relaxation behavior of composites can be largely
explained by the incorporation of a time-dependent shear-stress transfer
coefficient.

3.5. Effect of MAPP addition

Previous studies have neglected the time dependence of Gm on the


reinforcement efficiency, and have instead attributed the effect of fibers
on composite viscoelasticity to the formation of additional covalent
bonds at the fiber-matrix interface. The addition of MAPP can result in
two effects: firstly, it can homogeneously mix with the polymer matrix
and thus alter the stress relaxation behavior of this material, which
would, of course, alter the stress relaxation behavior of the composite.
Secondly, the MAPP can promote additional bonding at the fiber-matrix
interface and restrict polymer chain mobility in the surrounding
polymer, as suggested in the literature.
A comparison of the stress relaxation behavior of the neat poly-
propylene films with and without MAPP (S5 and S1, respectively) was
used to understand the effect of 5% MAPP addition on the base matrix
itself (see Fig. 7).
It is evident that the addition of the low-molecular weight MAPP
restricted stress relaxation of the pure polypropylene matrix.
Investigating the reasons for this effect was beyond the scope of this
study but could be related to possible changes in the crystallinity of the
base polymer [27,28]. Thus, when using an analytical model to predict
the stress relaxation behavior of the glass-reinforced polypropylene
samples containing MAPP, it is more appropriate to use the stress re-
laxation behavior of polypropylene with MAPP (S5) as the matrix rather
than comparing composite behavior to polypropylene without MAPP
(S1).

Fig. 6. Comparison of experimental stress relaxation (○) to the analytical


model (–) shows good agreement between the two at all fiber fractions; the
error bars are based on a 90% confidence interval.

Fig. 7. The effect of MAPP addition on the stress relaxation behavior of the base
polymer was evaluated by comparing the behavior of polypropylene without
MAPP [S1 (○)] to polypropylene with MAPP [S5 (●)].

89
N. Obaid et al. Composites Science and Technology 161 (2018) 85–91

3.6. Effect of fiber content on stress relaxation behavior of PP (with MAPP)

The experimental data for the stress relaxation in fiber-reinforced


samples containing MAPP was compared to the predictions from the
analytical model (see Fig. 8).
Fig. 8 shows that the experimental data aligns well with the ana-
lytical model; however, it is apparent from the composites containing
10% and 15% fiber volume fractions that the addition of MAPP does
hinder stress relaxation somewhat, although the effect is quite small.
The effect of MAPP is more pronounced at shorter relaxation times,
evident by decreased agreement between the model predictions and the
experimental data. It is possible that the addition of MAPP results in
increased inhomogeneity in the matrix, causing the matrix closer to the
fiber to crystallize differently than that in the bulk to create an “in-
terphase”, which slows stress relaxation. However, since the analytical
model is based only on the time-dependent shear stress transfer be-
tween the two phases, and does a reasonably good job of predicting the
experimental data, it is still safe to conclude that most of the stress
relaxation behavior of composites can be explained by incorporating
time-dependent shear stress transfer without inferring matrix changes
near the interface.

4. Conclusions

In the first part of this study, stress relaxation experiments were


conducted on glass fiber-reinforced polypropylene. The experimental
results were compared to a previously published analytical model,
which can be used to predict the stress relaxation behavior of short-
fiber composites using a viscoelastic shear-lag approach. The detailed
development of the analytical model is described in our previous work
[13].
A good fit was observed between the analytical model and the ex-
perimental results, indicating that the analytical model was an ade-
quate and accurate tool to predict the stress relaxation behavior of
short, elastic fiber-reinforced composites with various fiber fractions.
The agreement with the model indicates that the stress relaxation be-
havior of a short-fiber composite can be explained by incorporating a
time-dependent shear stress transfer at the fiber-matrix interface.
Previous studies investigating the effect of short fibers on the stress
relaxation behavior of a composite have attributed changes in the re-
laxation time constant to increased covalent bonding at the fiber-matrix
interface. The second part of the study evaluated this hypothesis and
investigated the role of interfacial covalent bonds. This was conducted
by examining the stress relaxation behavior of composites containing
MAPP as a coupling agent, and then comparing the experimental
findings to predictions made using the analytical model. The addition of
MAPP was found to make changes to the properties of the bulk matrix,
and thus, the analytical model predictions were calculated using the
new matrix properties.
It was found that even when MAPP was added to the system to alter
the fiber/matrix interface, the experimental data remained well-aligned
with the analytical model predictions which do not depend on model-
ling interfacial changes. In fact, since the analytical model predictions
align quite closely with the experimental data, it was concluded that
most of the stress relaxation behavior of a composite can predicted
using a simple model incorporating the time-dependent matrix modulus
and the time dependent shear stress transfer efficiency. The fibers are
gradually unloaded during a stress relaxation experiment, even though
they themselves are perfectly elastic.
Fig. 8. Stress relaxation behavior of composites reinforced with 5%, 10%, and
15% fiber volume fractions. These experimental data is for samples containing
The analytical model was able to predict the experimental stress
MAPP (●), and has been compared to an analytical model based on S5 as the relaxation behavior of composites, both with and without MAPP, with a
matrix (-) to understand the effect of covalent bonding only. high level of accuracy. This highlights its value as an accurate tool to
predict the stress relaxation behavior of short-fiber composites.

90
N. Obaid et al. Composites Science and Technology 161 (2018) 85–91

Acknowledgements 437–445.
[13] N. Obaid, M.T. Kortschot, M. Sain, Understanding the stress relaxation behavior of
polymers reinforced with short elastic fibers, Materials 10 (2017) 472.
The authors gratefully acknowledge the National Science and [14] I. Sevostianov, V. Levin, E. Radi, Effective viscoelastic properties of short-fiber re-
Engineering Research Council of Canada (NSERC) for financial support. inforced composites, Int. J. Eng. Sci. 100 (2016) 61–73.
[15] N. Smith, G.A. Medvedev, R.B. Pipes, Viscoelastic shear lag analysis of the dis-
continuous fiber composite, Proceedings of the 19th International Conference on
References Composite Materials, Montreal, QC, Canada, August 2013 28 July–2.
[16] R. Luciano, E.J. Barbero, Analytical expressions for the relaxation moduli of linear
[1] M.R. Piggott, D.M.R. Taplin, Load Bearing Fiber Composites, Pergamon Press, New viscoelastic composites with periodic microstructure, J. Appl. Mech. 62 (1995)
York, 1980. 786–793.
[2] R.M. Jones, Mechanics of Composite Materials, Scripta Book Company, [17] T. Okabe, M. Nishikawa, GLS strength prediction of glass-fiber-reinforced poly-
Washington, D. C, 1975. propylene, J. Mater. Sci. 44 (2009) 331–334.
[3] S.K. Kutty, G.B. Nando, Short Kevlar fiber-thermoplastic polyurethane composite, J. [18] T. Okabe, M. Nishikawa, N. Takeda, Micromechanics on the rate-dependent fracture
Appl. Polym. Sci. 43 (1991) 1913–1923. of discontinuous fiber-reinforced plastics, Int. J. Damage Mechanics 19 (2010)
[4] F. Suhara, S.K. Kutty, G.B. Nando, Stress relaxation of polyester fiber-polyurethane 339–360.
elastomer composite with different interfacial bonding agents, J. Elastom. Plast 30 [19] M. Hashimoto, T. Okabe, T. Sasayama, H. Matsutani, M. Nishikawa, Prediction of
(1998) 103–117. tensile strength of discontinuous carbon fiber/polypropylene composite with fiber
[5] U. Saeed, K. Hussain, G. Rizvi, HDPE reinforced with glass fibers: rheology, tensile orientation distribution, Composites Part A 43 (2012) 1791–1799.
properties, stress relaxation, and orientation of fibers, Polym. Compos. 35 (2014) [20] I.J. Beyerlein, S.L. Pheonix, R. Raj, Time evolution of stress redistribution around
2159–2169. multiple fiber breaks in a composite with viscous and viscoelastic matrices, Int. J.
[6] M.S. Sreekala, M.G. Kumaran, R. Joseph, S. Thomas, Stress relaxation behavior in Solids Structures 35 (1998) 3177–3211.
composites based on short oil-palm fibers and phenol formaldehyde resins, Compos. [21] N. Iyengar, W.A. Curtin, Time-dependent failure in fiber-reinforced composites by
Sci. Technol. 61 (2001) 1175–1188. matrix and interface shear creep, Acta Mater. 45 (1997) 3419–3429.
[7] F. Stan, C. Fetecau, Study of stress relaxation in polytetraflyoroethylene composites [22] Z.Z. Du, R.M. McMeeking, Creep models for metal matrix composites with long
by cylindrical macroindentation, Compos. Part B. Eng 47 (2013) 298–307. brittle fibers, J. Mech. Phys. Solids 43 (5) (1995) 701–706.
[8] J. George, M.S. Sreekala, S. Thomas, S.S. Bhagawan, N.R. Neelakantan, Stress re- [23] C.H. Weber, Z.Z. Du, F.W. Zok, High temperature deformation and fracture of a
laxation behavior of short pineapple fiber reinforced polyethylene composites, J. fiber reinforced titanium matrix composite, Acta Mater. 44 (1996) 683–695.
Reinf. Plast. Compos 17 (1998) 651–672. [24] B. Fabeny, W.A. Curtin, Damage-enhanced creep and rupture in fiber-reinforced
[9] V.G. Geethamma, L.A. Pothan, B. Rhao, N.R. Neelakantan, S. Thomas, Tensile stress composites, Acta Mater. 44 (1996) 3439–3451.
relaxation of short-coir-fiber reinforced natural rubber composites, J. Appl. Polym. [25] D.C. Lagoudas, C.Y. Hui, S.L. Pheonix, Time evolution of overstress profiles near
Sci. 94 (2004) 96–104. broken fibers in a composite with a viscoelastic matrix, Int. J. Solids Structures 25
[10] B. Mirzaei, M. Tajvidi, R.H. Falk, C. Felton, Stress relaxation behavior of lig- (1) (1989) 45–66.
nocellulosic-high density polyethylene composites, J. Reinf. Plast. Compos 30 [26] H.L. Cox, The elasticity and strength of paper and other fibrous materials, Br. J.
(2011) 875–881. Appl. Phys. 3 (1952) 72–79.
[11] L.A. Pothan, N.R. Neelakantan, B. Rao, S. Thomas, Stress relaxation behavior of [27] X. Zhou, Y. Yu, Q. Lin, L. Chen, Effects of maleic anhydride-grafted polypropylene
banana fiber-reinforced polyester composites, J. Reinf. Plast. Compos 23 (2004) (MAPP) on the physico-mechanical properties and rheological behavior of bamboo
153–165. powder-polypropylene foamed composites, Bioresources 8 (2013) 6263–6279.
[12] S. Boukettaya, F. Almaskari, A. Abdala, A. Alawar, H.B. Daly, A. Hammami, Water [28] A. Oromiehie, H. Ebadi-Dehaghani, S. Mirbagheri, Chemical modification of poly-
absorption and stress relaxation behavior of PP/date palm fiber composite mate- propylene by maleic anhydride: melt grafting, characterization and mechanism,
rials, in: M. Chouchance, T. Fakhfakh, H. Daly, N. Aifaoui, F. Chaari (Eds.), Design Inter. J. Chem. Eng. Appl. 5 (2014) 117–122.
and Modeling of Mechanical Systems -II, Springer, Hammamet, Tunisia, 2015, pp.

91

Potrebbero piacerti anche