You are on page 1of 16

Article

Journal of Thermoplastic Composite


Materials
HDPE reinforced with 26(8) 1025–1040
ª The Author(s) 2012
nanoparticle, natural Reprints and permissions:
sagepub.co.uk/journalsPermissions.nav

and animal fibers: DOI: 10.1177/0892705712454867


jtc.sagepub.com

Morphology, thermal,
mechanical, stress
relaxation, water
absorption and
impact properties
Suchart Siengchin1 and Vilai Rungsardthong2

Abstract
Binary and ternary composites composed of high-density polyethylene (HDPE), boehmite
alumina (BA) and different kinds of natural and animal fibers, such as flax, sponge gourd,
palm and pig hair (PH), were produced by hot press technique. Aqueous BA suspensions
were sprayed on the HDPE/flax mat to prepare nanoparticle/natural fiber–reinforced tern-
ary polymer composites followed by drying. The dispersion of the natural and animal fibers
and BA particles in the composites was studied by scanning electron microscopy and dis-
cussed. The thermomechanical and stress relaxation properties of the composites were
determined by the thermogravimetric analysis, dynamic-mechanical thermal analysis and
short-time stress relaxation tests (performed at various temperatures), respectively. The
HDPE-based composites were subjected to water absorption and instrumented falling
weight impact tests. It was found that all the composite systems increased the stiffness and
stress relaxation and reduced the impact toughness. The stress relaxation modulus of nat-
ural and animal fiber composites was higher compared with that of the neat HDPE. This

1
Department of Production Engineering, The Sirindhorn International Thai-German Graduate School of
Engineering (TGGS), King Mongkut’s University of Technology North Bangkok, Bang Sue, Bangkok, Thailand
2
Department of Agro-Industrial Technology, Faculty of Applied Science, King Mongkut’s University of
Technology North Bangkok, Bang Sue, Bangkok, Thailand

Corresponding author:
Suchart Siengchin, Department of Production Engineering, The Sirindhorn International Thai-German Graduate
School of Engineering (TGGS), King Mongkut’s University of Technology North Bangkok, 1518 Pibulsongkram
Road, Bang Sue, Bangkok 10800, Thailand.
Email: suchart.s.pe@tggs-bangkok.org

1025
1026 Journal of Thermoplastic Composite Materials 26(8)

modulus increased greatly with incorporation of BA. The relaxation master curves were
constructed by applying the time–temperature superposition principle. The inverse of
Findley power law could be fairly applicable to describe the relaxation modulus versus time
traces for all systems studied. Incorporation of BA particles enhanced the thermal
resistance, which started to degrade at higher temperature compared with the HDPE/flax
mat composite. The HDPE/flax mat/BA composite could reduce the water uptake.

Keywords
High-density polyethylene (HDPE), flax, natural and animal fibers, composites and stress
relaxation

Introduction
The use of polymer composites has increased substantially in engineering applications
due to their outstanding properties when compared with pure polymers. Polyethylene
(PE) is one of the most widely used polyolefin polymers. High-density PE (HDPE) was
used as the matrix for nano and natural composite materials and has attracted increasing
interest recently due to its excellent chemical resistance, high impact strength, good ther-
mal and mechanical properties.1–3 The utilization of the good thermal and mechanical
properties of boehmite alumina (BA) nanoparticles in the HDPE composites has been
studied.4,5 BA incorporation increased thermal stability and stiffness of HDPE and
enhanced the elongation at a break obviously at the same time, while impact resistance
of the HDPE/BA systems was reduced with increasing BA content. To achieve a fine
nanoscale dispersion of the fillers is a great challenge. Besides the traditional nanocom-
posite techniques for thermoplastics, the water-mediated dispersion of suitable nanofil-
lers has been recommended.6 It was noticed earlier that this preparation technique of the
nanocomposites has a strong impact on the remarkable property improvements.7
A variety of natural fibers such as rice husk,8 jute,9 flax10 and banana11 fibers have
been tested for use in HDPE/fiber composites. Natural fiber is straightforward not only
from the viewpoint of offering significant cost advantages and benefits associated with
processing compared with synthetic fibers but also with respect to the reduction in the
dependency on foreign and domestic petroleum oil. Several research articles deal with
the optimization of natural fiber-reinforced HDPE composites. Some investigations look
for ways to improve the interactions between polymer matrix and natural fibers. The use
of HDPE-based maleated coupling agents was helpful to increase the strength properties
of HDPE–wood-flour composites, due to the increased interaction between the wood
flour and the HDPE matrix.12 Choudhury13 demonstrated that the tensile and flexural
properties of ionomer-treated sisal/HDPE composites were strongly improved compared
with neat HDPE. Mulinari et al.14 studied the effect of modification of sugarcane bagasse
cellulose with zirconium oxychloride/HDPE composites. The modified bagasse reduced
the composites elongation. The increase in stiffness is associated with fibers reinforce-
ment at the cost of the tensile strength and ductility. This reduction is associated
with defects generated in the material after cellulose fibers insertion. The plant- and
animal-based fibers attracted much attention in the engineering and bioengineering

1026
Siengchin and Rungsardthong 1027

industries such as automotive and medical applications and have undergone comprehen-
sive researches for new composites. The goal of this study was to explore the potential of
the water-mediated technique using nanospraying to disperse suitable fillers on nanos-
cale and investigate the potential of flax, sponge gourd (SG), palm and pig hair (PH)
fibers produced reinforced HDPE composites. A further aim of this work was to demon-
strate the feasibility of the production of HDPE/flax mat/BA composites. Accordingly,
binary and ternary composites composed of HDPE, BA and different kinds of natural and
animal fibers were produced by hot press technique and the structure–property relation-
ships of the resulting compounds were evaluated.

Experimental
Materials and preparation of composites
Nonwoven flax fibers (flax mat; 220 g/m2) were supplied by Dittrich Vliesstoffe GmbH
(Ramstein-Miesenbach, Germany). The flax fibers in the nonwoven textile are randomly
orientated. The nonwoven textile is made of individual fibers. PH and oil palm empty
fruit bunch fiber are the byproducts obtained from two food processing plants in Thai-
land (Betagro safety Meat Packing Co., Ltd and Asian Palm Oil Co., Ltd) SG fibers are
commercially available in Thailand. The natural and animal fibers were used as reinfor-
cing fillers. The water dispersible BA (AlO(OH); Dispal111N7-80 of Sasol GmbH,
Hamburg, Germany) served as filler for all composite systems. The nominal particle size
of alumina in water was 220 nm, although that of the alumina powder as delivered was
40 mm. Alumina has Al2O3 content of 80 wt% and specific surface area is 100 m2/g.
HDPE (Finck&Co, Krefeld, Germany) was used as a polymeric matrix for all composite
systems and its density was 0.95 g/cm3.
The HDPE binary and ternary composites were prepared by nanospraying and/or hot
press methods, respectively. First, the HDPE/flax mat/BA ternary composites were
produced by a hot press using nanospraying technique. The BA particles were dispersed
in water at ambient temperature under continuous mechanical stirring for 30 min to
obtain aqueous BA slurry. The BA slurry was sprayed into the nonwoven flax fibers by a
hand spraying-up and dried for 48 h at room temperature (RT) and then for 24 h at 80 C
in oven. The flax mat contents in the binary composites were set for 20 and 40 wt%. The
flax mat and alumina or PH, SG, palm fibers contents in the ternary composites were set
for 20 and 10 wt%, respectively. The composites produced are listed in Table 1. The
HDPE ternary composites followed by hand lay-up process of natural or animal fibers
and then a layer of HDPE sheet. The binary and ternary composites based on HDPE,
after the BA particles spraying on the flax mat and dried, were compression molded into
1-mm-thick sheets at T ¼ 190 C using the hot press (P/O/Weber, Maschienen und Appa-
ratebau, Remschalden, Germany).

Characterization and testing


Morphology detection. The fracture surfaces of compression-molded specimens were
subjected to scanning electron microscopy (SEM) inspection in a Supra™ 40VP SEM

1027
1028 Journal of Thermoplastic Composite Materials 26(8)

Table 1. Recipe and designation of the HDPE-based systems.

Sample designation Flax mat (wt%) BA (wt%) PH (wt%) Palm (wt%) SG (wt%)

HDPE – – – – –
HDPE/flax mat(l) 20 – – – –
HDPE/flax mat(h) 40 – – – –
HDPE/flax mat/BA 20 10 – – –
HDPE/flax mat/PH 20 – 10 – –
HDPE/flax mat/palm 20 – – 10 –
HDPE/flax mat/SG 20 – – – 10

BA: boehmite alumina; HDPE: high-density polyethylene; PH: pig hair; SG: sponge gourd.

(Carl Zeiss GmbH, Oberkochen, Germany). The surface was gold coated prior to SEM
inspection, which was performed at low acceleration voltage.

Thermal and thermomechanical response. Thermogravimetric analysis (TGA) was


performed on a DTG-60 SHIMADAZU device (Kyoto, Japan). TGA experiments were
conducted in the temperature range from 30 to 500 C under nitrogen at a heating rate of
10 C/min.
Dynamic mechanical thermal analysis was performed in tensile mode at 1 Hz, using a
Dynamic mechanical analysis (DMA) Q800 apparatus (TA Instruments, New Castle,
New Jersey, USA). The storage modulus (E0 ) along with mechanical loss factor (tan )
were determined as a function of the temperature (T ¼ 130 C . . . þ130 C). The strain
applied was 0.1% and the heating rate was set for 3 C/min. The specimen dimensions
were 30  10  1 mm3 (length  width  thickness).

Stress relaxation response. Short-time (duration 30 min) stress relaxation tests were
made in single cantilever mode at different temperatures, ranging from 5 to 45 C, using
the above-mentioned DMA apparatus. The strain applied was 0.5%. Isothermal tests
were run on the same specimen in the temperature range of 5–45 C by increasing its tem-
perature stepwise by 5 C. Prior to the stress relaxation measurement, the specimen was
equilibrated for 3 min at each temperature step. The specimen dimensions were
30  10  1 mm3 (length  width  thickness).

Water absorption. Water absorption of the composites was investigated over a period
of 30 days. The composites were cut into specimens (20  20 mm2) and then they were
immersed in water in a bath at RT. Weight gains were recorded by periodic removal of
the specimens from the water bath and weighing on a balance. The percentage gain at
any time t (Mt) as a result of moisture absorption was calculated from the following
equation
Ww  Wd
Mt ¼ 100% ð1Þ
Ww

1028
Siengchin and Rungsardthong 1029

where Wd and Ww denote the weight of dry material (initial weight of materials) and
weight of materials after exposure to water absorption, respectively.

Instrumented falling dart impact. Instrumented falling weight impact (IFWI) tests were
performed on a Fractovis 6785 (Ceast, Pianezza, Italy) using the following settings:
incident impact energy, 20 J; diameter of the dart, 20 mm; diameter of the support rig,
40 mm; weight of the dart, 10.357 kg; drop velocity, 1.97 m/s. IFWI tests were
performed on quadratic specimens of 60  60 mm2 at RT.

Results and discussion


Morphology
The SEM pictures in Figure 1(a) to (e) present the morphological characteristics of the
HDPE-based systems studied. It can be seen in Figure 1(a) to (b) that the BA particles
were homogeneously dispersed in the HDPE/flax mat composite. The individual flax
fiber had diameter in the range of 30–60 mm and displayed clear fiber lumina. However,
the lumina was not filled with HDPE matrix. Considering the present image of Figure
1(c), it can be concluded that due to the very smooth surface, a poor adhesion was present
between the HDPE matrix and the PH. The latter was present in the form of large fibers
and broad fibers size distribution. Furthermore, one may conclude from Figure 1(d) to (e)
that fiber surfaces of palm and SG were rough. This seems to be better stacked with the
related HDPE composites, which may indicate an enhanced adhesion of HDPE/flax mat/
palm and HDPE/flax mat/SG ternary composites. It was the first hint for an upgrade
property in the corresponding composites.

Thermogravimetric analysis
The effects of BA particles and natural and animal fibers on the thermal stability of
HDPE systems are displayed in Figure 2. One can be recognized that thermal degra-
dation of HDPE was a one-step procedure representing depolymerization. The binary
and ternary composite systems exhibited two-step–like transitions. Due to the dehydra-
tion from cellulose unit and thermal cleavage of glycosidic, the decomposition tempera-
ture of HDPE/fibers composites was comparatively lower than that of the HDPE as
reported for jute/HDPE composites.9 One can recognize that the thermal stability
decreased remarkably with increasing of flax mat content in the HDPE at whole tempera-
ture range. For the composites containing 20 wt% flax mat, the weight loss as reaching
370 C, was enhanced by approximately 15% compared with the HDPE/flax mat 40 wt%.
The incorporation of BA particles strongly improved the thermal degradation of the
HDPE/flax mat composite. In addition, it has to be mentioned that BA for improving the
thermooxidative stability has been reported for PEs and Polyoxymethylene (POM).4,15
The HDPE/flax mat/BA composite started to degrade at higher temperature compared
with the HDPE/flax mat composite. Note that the resistance to thermal degradation of
the HDPE/flax mat/PH composite was slightly higher than those of composites contain-
ing palm and SG. One possible explanation is that the high moisture content of natural

1029
1030 Journal of Thermoplastic Composite Materials 26(8)

Figure 1. SEM image from the fracture surfaces of HDPE/flax mat (a), HDPE/flax mat/BA (b),
HDPE/flax mat/PH (c), HDPE/flax mat/palm (d) and HDPE/flax mat/SG (e) composites. BA: boeh-
mite alumina; HDPE: high-density polyethylene; PH: pig hair; SEM: scanning electron microscopy;
SG: sponge gourd.

1030
Siengchin and Rungsardthong 1031

Figure 2. Weight loss versus temperature for the HDPE systems studied. HDPE: high-density
polyethylene.

fibers also caused a great problem for thermal properties. The change in the TGA values
shows the moisture difference between PG and natural composites. This finding is in
agreement with that reported by Cheung et al. who observed that the composite with ani-
mal fiber provided better thermal properties also.16

Dynamic mechanical response


Figure 3(a) and (b) depicts the storage modulus (E0 ) and loss factor (tan ) as functions of
the temperature for the binary and ternary composites systems containing different
amounts of flax mat. It can be observed in Figure 3(a) that the all composites exhibited
markedly higher stiffness than the neat HDPE. This can well be explained by the rein-
forcing effect of the natural fibers leading to increased stiffness. The storage modulus
increased remarkably with the increasing flax mat content. For the binary composites
containing 40 wt% flax mat, the E0 is improved by approximately 25% compared with
the HDPE/flax mat 20 wt%. The incorporation of BA into HDPE/flax mat resulted in a
pronounced stiffness enhancement at the whole temperature range. Poor adhesion of
large PH fibers in the HDPE/flax mat matrix acts as stress concentrations (compare
Figure 1(c)) and causes premature interfacial debonding that was accompanied by low
storage modulus. It is also well resolved that the stiffness of those composites containing
palm and SG was always inferior to that of HDPE/flax mat composite. This can be
explained by the consideration of the surface roughness of the natural fibers in the related
composites, as discussed earlier. Concerning, previous studies on the relaxation

1031
1032 Journal of Thermoplastic Composite Materials 26(8)

Figure 3. E0 (a) and tan  (b) versus T traces for the HDPE systems studied. HDPE: high-density
polyethylene.

processes in pure HDPE describe three transitions, namely, a-, b- and g-relaxations from
higher to lower temperature.17,18 The relaxation transition (g) located at 110 C was
assigned to the glass transition temperature. The b transition of HDPE was invisible due
to the absence of the chain branches. The latter peak in HDPE (a-relaxation at

1032
Siengchin and Rungsardthong 1033

approximately 80 C) was usually attributed to the chain segment mobility in the
crystalline phases and reoriented defect in crystalline phases. One can notice that the
g-relaxation peak of HDPE also exhibited a shift to high temperature regions with
the incorporation of BA particles. The shift of this peak toward higher temperatures due
to alumina indicates for the reinforcing effect. On the other hand, the effect of natural
and animal fibers on the relaxation transitions was marginal.

Stress relaxation response


Theoretical background. Stress relaxation phenomenological models provide another
route to study the time dependence and gain understanding on viscoelastic behavior of
reinforcing composite materials. For stress relaxation under applied constant strain ("0),
the material response is given by
ð t Þ
Er ð t Þ ¼ ð2Þ
"0
where Er(t) is the stress relaxation modulus, and (t) is stress, both as a function of time (t).
In the nonlinear range, the dependence upon the level of the applied deformation can
be expressed by multiplying the linear parameters by so-called nonlinearity factors,
which are deformation-, time- and temperature-dependent. The nonlinear stress relaxa-
tion modulus is given by
ðt; "ðtÞ; T Þ
Er ðt; "ðtÞ; T Þ ¼ ð3Þ
"0
Generally, the basis of the conversion method from the modulus to the creep compliance
can be determined using a linear viscoelastic material principle.19 Relaxation modulus is
defined as the inverse quantity of creep compliance (J) by the following equation
1
Er ¼ ð4Þ
J
Accordingly, the stress relaxation and creep behaviors of material are predicted by
empirical model, such as the Findley power law can well describe the creep compliance
versus time traces. Therefore, in order to obtain a deeper insight into the effect of the
natural fiber and nano-reinforced HDPE systems on the temperature and the long-term
stress relaxation behavior, the applicability of the inverse of the Findley equation has
been investigated following the correspondence relationship. Accordingly, for the time
dependency of the relaxation modulus, the inverse of the Findley power law equation is
given by20

ErF ¼ ðErF0 þ ErF1  tn Þ1 ð5Þ


where the subscript F indicates that the parameters are linked with the Findley
power law, n is a constant independent of strain, ErF0 is the time-independent
relaxation modulus and ErF1 is coefficient of the time-dependent term.

1033
1034 Journal of Thermoplastic Composite Materials 26(8)

Figure 4. Characteristic relaxation modulus–time curves registered for the HDPE systems stud-
ied at testing temperature of 5 C (a) and 45 C (b). HDPE: high-density polyethylene.

Figure 4(a) to (b) display the stress relaxation modulus (Er) as a function of time for
the binary and ternary composites systems containing different amounts of flax mat at
temperature of 5 C and 45 C, respectively. The solid lines appearing in each curve
represents the calculated data using the inverse of the Findley power law model. It is
clear from Figure 4 that the relaxation modulus decreased with increasing temperature.
The Er values of the all composites systems were higher when compared with the HDPE.

1034
Siengchin and Rungsardthong 1035

Figure 5. Characteristic relaxation modulus–time curves registered on HDPE/flax mat composite


at different testing temperatures. HDPE: high-density polyethylene.

For the relaxation modulus of HDPE/flax mat/BA ternary composite after the relaxation
time of 1800 s was increased by 25% and 40% at T ¼ 5 and 45 C, respectively, compared
with the HDPE/flax mat binary composite. This relaxation modulus increased greatly
with incorporation of BA particles due to the development of a restricted molecular
mobility. This phenomenon, observed also in DMA test and reported in Polystyrene
(PS),21 is associated with a pronounced increase in the stiffness. One can also recognize
that the HDPE/flax mat/SG composite exhibited higher relaxation modulus value than
the composites containing PH and palm. The DMA corroborated this fact by indicating
that the storage modulus of roughness surface SG was shifted toward higher values
(compare Figure 3(a)). However, the difference in the relaxation modulus between
natural and animal fibers diminishes with increasing temperature. Figure 5 demonstrates
the effect of increased temperature on the relaxation modulus response of HDPE/flax
mat composite. One can recognize that the shape of relaxation curves of the composites
was similar at all test temperature. The time-independent relaxation modulus (ErF0)
increased remarkably with increasing temperature.
Master curves of the relaxation modulus were produced by superimposing the
relaxation modulus versus time traces using the time–temperature superposition (TTS)
principle. A reference temperature (Tref ¼ 25 C) was used for this superposition
(shifting) process. The related shift factor (aT) used for the generation of relaxation
modulus master curve is aT ¼ Er(T)/(Er(Tref)). Master curves of the relaxation modulus
against time created and simulated fits according to the inverse of the Findley power law
at a reference temperature of 25 C are shown in Figure 6 for the HDPE, HDPE/flax mat

1035
1036 Journal of Thermoplastic Composite Materials 26(8)

Figure 6. Relaxation modulus master curves constructed by considering the TTS and selecting
Tref ¼ 25 C and their fitting by the Findley power law equation (inverse). TTS: time–temperature
superposition.

and HDPE/flax mat/BA composite. The shift factors are linked with TTS via the
Arrhenius function. From the shift factors for the relaxation modulus of the HDPE, the
activation energy of 200 kJ/mol can be calculated. The activation energy of the HDPE/
flax mat and HDPE/flax mat/BA binary and ternary composites was calculated as 157
and 153 kJ/mol, respectively. One can notice that the relaxation modulus increased with
decreasing activation energy and addition of flax and BA reinforcements. On other hand,
the related curves show a good fit between the experimental and the simulated results
according to the Findley power law. This power law fits very well to the relaxation
modulus data for all systems up to approximately 1333 min.

Water absorption. Figure 7 depicts the water uptake as a function of time for the
binary and ternary composites systems containing different amounts of flax mat. The
water uptake was 0.93% after subjected to water absorption. This was due to the light
interaction with water. A rapid water uptake was observed for all binary and ternary
composites within the first 7 days of immersion. The water absorption of composite
systems was reached quasi-equilibrium state after 18 days. One can also recognize that
the penetration of water increased markedly with increasing flax mat content in HDPE
matrix. Note that the incorporation of BA particles in the HDPE/flax mat composite
could reduce the water uptake. The HDPE/flax mat/BA composite recorded water
absorption value at 4.87% upon 30 days. This is in accordance with recent reports claim-
ing that the water absorption was reduced by adding nanofillers. For polyester and poly-
esterimide with silicon dioxide, titanium dioxide and zinc oxide composites when
compared with standard compounds, showed better water resistance.22 On the other

1036
Siengchin and Rungsardthong 1037

Figure 7. Water uptake on the binary and ternary composites systems containing different
amounts of flax mat. Fick’s law was used to calculate the diffusion coefficient of water absorption
(D) after 30 days immersed in water, that is D ¼  (kb/40)2.

hand, the water sorption behavior was slightly considered to depend on the types of rein-
forcing fibers. The water uptake of HDPE/flax mat/PH composite was reduced by
approximately 33% and 14% as reaching 30 days compared with the composites contain-
ing SG and palm, respectively. Fick’s law was adopted to calculate the diffusion coeffi-
cient (D) of water absorption values. It is well resolved that there is a fair agreement
between the water uptake and those calculated by the Fick’s law. By comparing the dif-
fusion coefficient (compare Figure 7), one may conclude that the D of HDPE/flax mat
composite decreased with incorporation of BA. This result suggests that D may be sen-
sitive to the nanoparticles barrier effect. It has also been observed that these nanofillers
reacted with water lead to decrease the rate of diffusion.23,24 Addition of SG and palm
fibers into the HDPE/flax mat also yielded an increase in diffusion coefficient of water
absorption.

Falling dart impact response. The reinforcing effects of the BA particles and natural and
animal fibers on the force versus time curves are demonstrated on the HDPE systems in
Figure 8. Observation of impact energy indicates that the natural and animal fibers
enhanced the stiffness and reduced the toughness of the related composites at the same
time. Addition of the BA in HDPE/flax mat composite was accompanied with a shift of
the force peak toward higher force and longer time. This is due to the nanoparticle
character of the reinforcing effects. Note that the BA reinforcing the matrix in HDPE/

1037
1038 Journal of Thermoplastic Composite Materials 26(8)

Figure 8. Characteristic force–time curves for the HDPE systems studied. HDPE: high-density
polyethylene.

flax mat system leads to increased stiffness and accompanied with reduced ductility.
This is in qualitative agreement in accordance with the DMA results (compare Figure 3(a))
and the reduction in impact toughness of HDPE by BA particles was also discussed in Ref.
[5]. The presence of SG fibers was associated with a reduction in the maximum force peak
compared with the PH and palm reinforcements. Accordingly, the observed impact
resistance is likely less due to the BA particles themselves, but more due to their amount of
flax fibers.

Conclusion
This work devoted to study the morphology, thermal, mechanical, stress relaxation,
water absorption and falling weight impact properties of a HDPE and its PH, palm, SG
fibers and BA reinforced binary and ternary composites. The BA was introduced in the
HDPE/flax mat in 10 wt% via spraying method. Based on this work, the following
conclusions can be drawn:

 Nanospraying technique resulted in homogeneously BA dispersed in the HDPE/flax


mat composite. The flax mat, palm and SG fibers seem to be better stacked with the
related composites compared with the PH.
 Incorporation of BA particles improved the thermal degradation and water uptake of
the HDPE/flax mat composite.
 The storage modulus of the all composites was higher than the neat HDPE at the cost of
the ductility (impact toughness). The stiffness increased with increasing the flax mat
content. Poorly, adhesion of PH fibers in the HDPE/flax mat matrix accompanied
by low stiffness.

1038
Siengchin and Rungsardthong 1039

 The stress relaxation values of the all composites systems were higher compared with
the HDPE. This relaxation modulus increased greatly with incorporation of BA parti-
cles due to the development of a restricted the molecular mobility. With addition SG
exhibited higher relaxation modulus value than the composites containing PH and
palm. Based on isothermal short-term stress relaxation experiments master curves in
form of stress relaxation modulus versus time were constructed by TTS principle. The
related master curves could be well described by the inverse of the Findley power law.

References
1. Minkova L, Peneva Y, Tashev E, Filippi S, Pracella M and Magagnini P. Thermal properties
and microhardness of HDPE/clay nanocomposites compatibilized by different functionalized
polyethylenes. Polym Test 2009; 28: 528–533.
2. Swain SK and Isayev AI. Effect of ultrasound on HDPE/clay nanocomposites: rheology,
structure and properties. Polymer 2007; 48: 281–289.
3. Faruk O and Matuana LM. Nanoclay reinforced HDPE as a matrix for wood-plastic compo-
sites. Compos Sci Technol 2008; 68: 2073–2077.
4. Khumalo VM, Karger-Kocsis J and Thomann R. Polyethylene/synthetic boehmite alumina
nanocomposites: structure, thermal and rheological properties. Express Polym Lett 2010; 4:
264–274.
5. Khumalo VM, Karger-Kocsis J and Thomann R. Polyethylene/synthetic boehmite alumina
nanocomposites: structure, mechanical, and perforation impact properties. J Mater Sci
2011; 46: 422–428.
6. Siengchin S, Karger-Kocsis J, Apostolov AA and Thomann R. Polystyrene-fluorohectorite
nanocomposites prepared by melt mixing with latex precompounding: structure and mechan-
ical properties. J Appl Polym Sci 2007; 106: 248–254.
7. Siengchin S, Karger-Kocsis J and Thomann R. Alumina filled polystyrene micro- and nano-
composites prepared by melt mixing with and without latex precompounding: structure and
properties. J Appl Polym Sci 2007; 105: 2963–2972.
8. Panthapulakkal S, Law S and Sain M. Enhancement of processability of rice husk filled
high-density polyethylene composite profiles. J Thermoplas Compos Mater 2005; 18:
445–458.
9. Mohanty S, Verma SK and Nayak SK. Dynamic mechanical and thermal properties of MAPE
treated jute/HDPE composites. Compos Sci Technol 2006; 66: 538–547.
10. Li X, Tabil LG, Oguocha IN and Panigrahi S. Thermal diffusivity, thermal conductivity, and
specific heat of flax fiber-HDPE biocomposites at processing temperatures. Compos Sci
Technol 2008; 68: 1753–1758.
11. Liu H, Wua Q and Zhang Q. Preparation and properties of banana fiber-reinforced composites
based on high density polyethylene (HDPE)/Nylon-6 blends. Biores Technol 2009; 100:
6088–6097.
12. Li Q and Matuana LM. Effectiveness of maleated and acrylic acid-functionalized polyolefin cou-
pling agents for HDPE-wood-flour composites. J Thermoplas Compos Mater 2003; 16: 551–564.
13. Choudhury A. Isothermal crystallization and mechanical behavior of ionomer treated sisal/
HDPE composites. Mater Sci Eng A 2008; 491: 492–500.
14. Mulinari DR, Voorwald HJC, Cioffi MOH, Silva MLCP, Cruz TG and Saron C. Sugarcane
bagasse cellulose/HDPE composites obtained by extrusion. Compos Sci Technol 2009; 69:
214–219.

1039
1040 Journal of Thermoplastic Composite Materials 26(8)

15. Siengchin S, Karger-Kocsis J, Psarras GC and Thomann R. Polyoxymethylene/polyurethane/


alumina ternary composites: structure, mechanical, thermal and dielectric properties. J Appl
Polym Sci 2008; 110: 1613–1623.
16. Cheung HY, Ho MP, Lau KT, Cardona F and Hui D. Natural fibre-reinforced composites for
bioengineering and environmental engineering applications. Compos Part B 2009; 40:
655–663.
17. John B, Varughese KT, Oommen Z, Potschke P and Thomas S. Dynamic mechanical behavior
of high density polyethylene/ ethylene vinyl acetate copolymer blends: the effects of blend
ratio, reactive compatibilization, & dynamic vulcanization. J Appl Polym Sci 2003; 87:
2083–2099.
18. Vaisman L, GonzÁlez MF and Marom G. Transcrystallinity in brominated UHMWPE fiber
reinforced HDPE composites: morphology and dielectric properties. Polymer 2003; 44:
1229–1235.
19. Ferry JD. Viscoelastic properties of polymers. New York: Wiley, 1980.
20. Siengchin S and Karger-Kocsis J. Mechanical and stress relaxation behavior of santoprene1
thermoplastic elastomer/boehmite alumina nanocomposites produced by water-mediated and
direct melt compounding. Composites Part A 2010; 41: 768–773.
21. Siengchin S, Abraham TN and Karger-Kocsis J. Structure–stress relaxation relationship in
polystyrene/fluorohectorite micro- and nanocomposites. Mech Compos Mater 2008; 44:
495–504.
22. Gornicka B and Prociow E. Polyester and polyesterimide compounds with nanofillers for
impregnating of electrical motors. Acta Phys Pol 2009; 115: 842–845.
23. Becker O, Varley RJ and Simon GP. Thermal stability and water uptake of high performance
epoxy layered silicate nanocomposites. Eur Polym J 2004; 40: 187–195.
24. Chow WS. Water absorption of epoxy/glass fiber/organo-montmorillonite nanocomposites.
Express Polym Lett 2007; 1: 104–108.

1040

You might also like