Você está na página 1de 6

The effect of crystallization of PLA on the thermal and mechanical

properties of microbrillated cellulose-reinforced PLA composites


Lisman Suryanegara, Antonio Norio Nakagaito, Hiroyuki Yano
*
Research Institute for Sustainable Humanosphere, Kyoto University, Uji, Kyoto 611-0011, Japan
a r t i c l e i n f o
Article history:
Received 5 September 2008
Received in revised form 9 January 2009
Accepted 12 February 2009
Available online 26 February 2009
Keywords:
A. Polylactic acid
A. Microbrillated cellulose
A. Nanocomposites
B. Differential scanning calorimetry
B. Thermo-mechanical properties
a b s t r a c t
This paper describes the thermal and mechanical properties of nanocomposites based on polylactic acid
(PLA) and microbrillated cellulose (MFC). The primary objective of this study was to improve the storage
modulus of PLA at a high temperature. MFC and PLA were mixed in an organic solvent with various ber
contents up to 20 wt%, followed by drying, kneading and hot pressing into sheets. The nanocomposites
were prepared in two different states, fully amorphous and crystallized. Differential scanning calorimetry
(DSC) measurements revealed that the presence of MFC accelerates the crystallization of PLA. The tensile
modulus and strength of neat PLA were improved with an increase of MFC content in both amorphous
and crystallized states. The addition of 20 wt% of MFC in PLA improved the storage modulus of crystal-
lized PLA at a high temperature (120 C) from 293 MPa to 1034 MPa.
2009 Elsevier Ltd. All rights reserved.
1. Introduction
During the last decades, polymers have replaced many conven-
tional materials in various applications because of the ease of pro-
cessing, high productivity, and substantial cost reduction that are
possible with polymers [1]. Recently, rising oil prices and a grow-
ing environmental awareness throughout the world have moti-
vated many researchers to develop products from bio-based
materials. Among bio-based polymers, polylactic acid (PLA) has a
great potential to replace petroleum-based plastics because of its
high stiffness and strength, which are comparable to polystyrene.
PLA is a versatile polymer made from renewable agricultural raw
materials that are fermented into lactic acid.
Based on the stereochemical structure, PLA can easily be modi-
ed by polymerizing a controlled mixture of the L- or D-isomers to
yield high-molecular-weight amorphous or crystalline polymers
[2]. PLA resins containing more than 93% of L-lactic acid are
semi-crystalline while PLA with 5093% L-lactic acid is strictly
amorphous [3].
In the industry, PLA can be processed in a way similar to poly-
propylene. However, low heat resistance, brittleness, and a slow
crystallization limit the wider application of PLA.
The reinforcement of PLA using lignocellulosic materials has
been studied with the goal of obtaining fully bio-based composites
[47]. Compared to inorganic llers, the main advantages of ligno-
cellulosics are their renewability, low cost, and low density [5]. The
thermal and mechanical properties of PLA composites have been
investigated using plant and pulp bers as reinforcement [6,7]. In
these studies, even though the tensile modulus of PLA increased
signicantly with an increase of plant ber content, the strength
decreased.
In recent years, microbrillated cellulose (MFC), which consists
of mechanically brillated pulp into nano to submicron wide bers
forming a web-like network, has received signicant research
attention. Zimmerman et al. [8] reported that MFC-reinforced
hydroxypropyl cellulose (HPC) prepared by lm casting has three
times higher tensile modulus and ve times higher tensile strength
compared with the matrix without reinforcement. Nakagaito and
Yano [9] showed that nanocomposites produced by compression
molding of MFC sheets impregnated with phenol formaldehyde
(PF) resin have high bending strength and modulus comparable
to magnesium alloy. Nanocomposite materials based on MFC and
melamine formaldehyde (MF) resulted in high mechanical damp-
ing, showing their potential to be used as loudspeaker membranes
[10], and reinforcing polyurethane (PU) with MFC drastically im-
proved heat resistance [11].
In a previous study, Iwatake et al. [12] developed a technique to
attain uniform dispersion of MFC in PLA matrix by premixing the
bers and the matrix in an organic solvent medium, followed by
kneading. They demonstrated that the tensile modulus of PLA im-
proves without a reduction of yield strain at a ber content of up to
10 wt%. The MFC reinforcement increased the tensile strength of
neat PLA by 25%. In their case, the PLA matrix was a fully amor-
phous grade which will never crystallize. However, the amorphous
PLA/MFC composite has low storage modulus at high temperature,
limiting its application particularly as a structural material.
0266-3538/$ - see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compscitech.2009.02.022
* Corresponding author. Tel.: +81 774 38 3669; fax: +81 774 38 3658.
E-mail address: yano@rish.kyoto-u.ac.jp (H. Yano).
Composites Science and Technology 69 (2009) 11871192
Contents lists available at ScienceDirect
Composites Science and Technology
j our nal homepage: www. el sevi er. com/ l ocat e/ compsci t ech
The main objective of the present study was to improve the
storage modulus of PLA at high temperature by using as the matrix
a semi-crystalline grade PLA. The nanocomposites were prepared
by mixing MFC and PLA in an organic solvent with various ber
contents up to 20 wt%, followed by drying, kneading and hot press-
ing into sheets. The mechanical properties of the nanocomposites
were evaluated at fully amorphous and crystallized states. We
found that the addition of 20 wt% of MFC in PLA signicantly im-
proved the storage modulus of crystallized PLA at a high
temperature.
2. Materials and methods
2.1. Materials
PLA, trade name LACEA H-100, was supplied by Mitsui Chemi-
cals, Inc., Japan. The density was 1.26 g/cm
3
, glass transition tem-
perature (T
g
) was 58 C, and melting point (T
m
) was 164 C. The
melt ow rate (MFR) was 8 g/10 min (190 C, 2.16 kg).
Microbrillated cellulose, trade name Celish KY-100G, was sup-
plied by Daicel Chemical Industries, Ltd., Japan. This product con-
sists of a 10 wt% ber content slurry.
2.2. Preparation of composites
An adequate amount of MFC (refer to Table 1) was stirred in
300 g of acetone for 2 h, followed by centrifugation to remove
the liquid phase (water and acetone). The process was repeated
four times in order to completely replace water by acetone. Next,
the acetone was replaced by dichloromethane using the same pro-
cess and repeating it two times. After that, the precipitate of MFC
was suspended in dichloromethane at the ratio described in Table
1 and treated by a homogenizer (Ultra-Turrax T25 Basic, IKA-Wer-
ke GmbH & Co. KG, Germany) for 2 min at 3000 rpm to obtain a
well-dispersed suspension. The suspension was stirred in a beaker,
and PLA was gradually added to the suspension at room tempera-
ture. After complete dissolution of PLA, the stirring was continued
for 2 h. The MFCdichloromethanePLA mixture was spread in
trays, and the solvent was evaporated in a fume hood at room tem-
perature for 12 h followed by vacuum-drying at 50 C for 8 h. The
mixture from which the solvent had been removed was kneaded
by a twin rotary roller mixer (Labo Plastomill, Toyo Seiki Sei-
saku-sho, Ltd., Japan) at 160 C, 40 rpm for 15 min. The compound
was crushed into small pieces and hot pressed into sheets at 180 C
in three steps: pre-heating for 5 min, 0.5 MPa for 5 min, and 1 MPa
for another 5 min. After hot pressing, the melted samples were
either immediately quenched in liquid N
2
to obtain a fully amor-
phous state (amorphous samples) or melt-crystallized in the hot
press at 100 C for 1 h to obtain highly crystallized solid structure
(crystallized samples).
2.3. X-ray diffraction
The diffraction patterns were obtained from radiation gener-
ated by the copper target of a Rigaku UltraX 18HF (Rigaku Corpo-
ration, Japan) set at 20 kV and 100 mA with the detector placed
on a goniometer scanning from 5 to 45.
2.4. Differential scanning calorimetry
DSC measurements were performed on a DSC6100 (SII Nano-
Technology Inc., Japan) in the temperature range of 25200 C
using approximately 8 mg of amorphous neat PLA and PLA/MFC
10 wt% composite. The samples were tested at non-isothermal
conditions at the same heating and cooling rates of 5 C/min in
three scans: heating, cooling, and heating. The thermal properties
such as glass transition temperature (T
g
), cold crystallization tem-
perature (T
cc
), melt temperature (T
m
), and enthalpy of melting
(DH
m
) were determined from the heating scan and the melt crys-
tallization temperature (T
mc
) was determined from the cooling
scan of the samples. The percentage crystallinity of each sample
was calculated by following the procedures described in previous
papers [13,14].
2.5. Tensile test
The tensile properties of samples were measured using an In-
stron 3365 universal materials testing machine. The specimen gage
length of about 25 mm was measured with a caliper for each sam-
ple upon gripping and the crosshead speed was set at 1 mm/min.
The specimens were 40 mm long, 4 mm wide, and 0.5 mm thick.
All results presented are the average values of ve measurements.
2.6. Dynamic mechanical analysis
The dynamic mechanical analysis (DMA) was performed on a
Rheovibron, DDV-25FP (Orientec A&D, Co. Ltd., Japan), with a
chuck distance of 21 mm, preload of 5 g and a frequency of 1 Hz.
The test specimens were 40 mm long, 4 mm wide, and 0.5 mm
thick. Specimens were cooled to 10 C with liquid N
2
, maintained
at this temperature for 10 min, and then heated to 120 C at a heat-
ing rate of 2 C/min.
3. Results and discussion
3.1. Effect of liquid N
2
quenching on the crystallinity of PLA
To independently evaluate the effect of MFC reinforcement and
the effect of PLA crystallization on the thermal and mechanical
properties of PLA composites, the melted samples were quenched
by liquid N
2
upon hot pressing to obtain a fully amorphous state.
X-ray diffraction analysis was performed to evaluate the crystallin-
ity of the samples.
Fig. 1a shows the X-ray diffraction patterns of amorphous neat
PLA (quenched by liquid N
2
) and crystallized neat PLA (melt-crys-
tallized at 100 C for 1 h). The patterns of corresponding PLA com-
posites at 10 wt% MFC content are shown in Fig. 1b. The neat PLA
and composite quenched by liquid N
2
exhibit no peaks showing
an amorphous nature. In the crystallized state, the neat PLA and
composite show peaks at around 2h = 16 and 19, indicating a
crystalline structure. These peak positions are in accordance with
those reported by Mathew et al. [15].
Based on the X-ray patterns, we could conrm that the samples
of PLA and its composite quenched by liquid N
2
were fully amor-
phous. Since the crystallinity of PLA inuences its thermal and
mechanical properties, quenching the samples with liquid N
2
al-
lowed us to assess the MFC reinforcement effect without the inu-
ence of crystallization.
3.2. Effect of the presence of MFC on the crystallization of PLA
The crystallization behavior of PLA in the presence of MFC was
studied using liquid N
2
quenched samples (amorphous state).
Table 1
Composition of the samples.
Samples MFC (g) PLA (g) Dichloromethane (g)
Neat PLA 0 60.0 300
PLA/MFC 3 wt% 1.8 58.2 390
PLA/MFC 5 wt% 3.0 57.0 450
PLA/MFC 10 wt% 6.0 54.0 600
PLA/MFC 20 wt% 12.0 48.0 900
1188 L. Suryanegara et al. / Composites Science and Technology 69 (2009) 11871192
Fig. 2a shows the thermograms of DSC for neat PLA and Fig. 2b pre-
sents the thermograms of PLA/MFC 10 wt% composite, obtained by
successive heating and cooling scans.
In Fig. 2a, in the rst heating scan, the thermogram of neat PLA
shows glass transition temperature (T
g
), cold crystallization tem-
perature (T
cc
), and melt temperature (T
m
), which are typical of
semi-crystalline polymer. During the cooling scan, the thermogram
shows a peak of melt crystallization temperature (T
mc
), but the
peak is lower than the peak of T
cc
in the rst heating. In the second
heating, the thermogram still exhibits the glass transition and cold
crystallization peaks, indicating that crystallization during the
cooling scan was not complete.
The thermogram of the composite (Fig. 2b) is different than the
thermogram of neat PLA. In the rst heating scan, the composite
shows the same trend as neat PLA, but the T
cc
of composite
(91.5 C) is lower than that of neat PLA (93.4 C). During the cooling
scan, the T
mc
peak of composite occurs at a higher temperature
(102.5 C) than that of neat PLA (95.6 C). It shows that PLA com-
posite started to crystallize sooner than neat PLA, which is an indi-
cation that the presence of MFC can act as nucleating agent on the
crystallization of PLA. In the second heating, the thermogramof the
composite exhibits only T
mc
while T
g
and T
cc
are almost absent,
indicating that the material was highly crystalline after the cooling
scan.
After the cooling scan, the degree of crystallinity of composite
(39.4%) was higher than that of neat PLA (17.2%). The higher crys-
tallinity of the composite compared to neat PLA, showed that the
presence of MFC accelerates the crystallization of PLA. Similar phe-
nomena were observed in microcrystalline cellulose (MCC) and
pulp ber reinforcements [16].
3.3. Effect of MFC reinforcement on the mechanical properties of PLA
To evaluate the effect of MFC reinforcement in the PLA matrix,
we studied the mechanical properties of the composite in the
amorphous state. Afterward, the samples were crystallized, allow-
ing us to evaluate the effect of crystallization of PLA on the
mechanical properties of MFC/PLA composite.
Fig. 3a shows the typical stressstrain curves of amorphous PLA
and the composites. The mean values and standard deviations of the
mechanical properties of PLA and composites in amorphous state
are summarized in Table 2. The gure clearly shows that both tensile
modulus and strength of PLA improved with an increase of MFC con-
tent, while the strainat break was decreased. The additionof MFCat a
ber content of 20 wt% improved the modulus of amorphous PLA
from 3.3 GPa to 5.2 GPa and the tensile strength (maximum stress)
fromaround58 MPa to70 MPa, while the strainat breakwas reduced
fromaround 7 to 2%. The Youngs modulus of neat PLA was increased
by adding stiff bers, whereas the increase of strength indicates good
interfacial adhesion between the matrix and the bers. These effects
of MFC reinforcement are in accordance with the results obtained in
the previous study for fully amorphous grade PLA [12].
Fig. 3b shows the tensile properties of crystallized PLA and com-
posites. Compared to the amorphous state (Fig. 3a), crystallization
of PLA increases the tensile modulus and strength of neat PLA and
the strain at break is reduced. As shown in Table 2, the modulus of
neat PLA is increased from 3.3 GPa to 4.0 GPa and the strength
from 50.2 MPa to 60.9 MPa, while the strain at break is reduced
from around 7 to 3%. In general, crystallization of semi-crystalline
polymer results in embrittlement of the polymer and hence de-
creases the strain at break [17].
Fig. 1. X-ray patterns of (a) PLA and (b) PLA/MFC 10 wt%, quenched by liquid N
2
and crystallized at 100 C for 1 h.
Fig. 2. DSC thermogram of (a) neat PLA and (b) PLA/MFC 10 wt% composite.
L. Suryanegara et al. / Composites Science and Technology 69 (2009) 11871192 1189
The crystallization increased the modulus of PLA/MFC compos-
ites without signicant reduction in the strain at break. The mod-
ulus of the composite at 20 wt% MFC content was improved from
5.2 GPa to 5.7 GPa. It should be emphasized that although the crys-
tallization of PLA makes neat PLA brittle, the yield strain of the
composites was not reduced due to the crystallization, and conse-
quently the MFC reinforcing effect obtained in the strength of PLA/
MFC composite in amorphous state was preserved.
The relationship between mechanical properties and MFC con-
tent in the crystallized state is plotted in Fig. 4. These results reveal
that the modulus and strength increase linearly with an increase in
MFC content up to 20 wt%. Mathew et al. [16] reported that the
addition of microcrystalline cellulose (MCC) increases the modulus
of PLA, while the strength decreases signicantly. In contrast, this
study demonstrates that MFC reinforcement improves both the
modulus and strength of crystallized PLA.
3.4. Effect of MFC reinforcement on the thermo-mechanical properties
of PLA
Compared to PLA in amorphous state, PLA in crystallized state
has better heat resistance at high temperature. To study the effect
of MFC reinforcement on the thermo-mechanical properties of PLA,
we performed DMA tests on samples in both amorphous and crys-
tallized states.
Fig. 5a shows the temperature dependence of the storage mod-
ulus (E
0
) of PLA and composites measured from samples in the
amorphous state. The gure shows that the storage modulus of
neat PLA below T
g
is almost constant at around 3 GPa. Above T
g
,
the modulus drops to 4 MPa at 80 C, and then increases to
200 MPa at 100 C. The increase of modulus refers to the cold crys-
tallization of the amorphous PLA.
In amorphous state, the addition of MFC improved the storage
modulus of neat PLA. The addition of MFC at a ber content of
20 wt% improved the modulus of neat PLA by 50% from
3508 MPa to 5223 MPa at 20 C (glassy state), and interestingly,
by 27 times from 4 MPa to 109 MPa at 80 C (rubbery state). This
immense difference in modulus improvement at 80 C compared
to that at roomtemperature is explained by the fact that the matrix
becomes extremely soft in the rubbery state and the reinforcement
becomes much more noticeable at high temperatures.
The effect of MFC content on the temperature dependency of
the storage modulus in PLA in the crystallized sate is shown in
Fig. 5b. The gure shows that the storage modulus of samples
is almost constant in the glassy state. Around T
g
, 60 C, the mod-
ulus of crystallized PLA decreased gradually with increase in tem-
perature, even though the PLA had been previously crystallized,
indicating that the material (PLA) was not 100% crystalline. The
decrease of modulus around T
g
is due to a-relaxation of amor-
phous structure.
The crystallization of PLA slightly improved the storage modu-
lus of amorphous neat PLA at 20 C from 3508 MPa to 3896 MPa,
whereas at 80 C the modulus increased by 126 times from
4 MPa to 505 MPa. The addition of MFC at a ber content of
20 wt%, as shown in Fig. 5b, further improved the storage modulus
of crystallized neat PLA from 505 MPa to 1616 MPa at 80 C and
from 293 MPa to 1034 MPa at 120 C. It should be emphasized that
the crystallized composite with MFC content of 20 wt% shows a
storage modulus around 1 GPa at 120 C. Such a high storage mod-
ulus under high temperature could not be achieved by amorphous
PLA [12] or semi-crystalline PLA without MFC reinforcement
(Fig. 5b). These results conrm that the combination of MFC rein-
forcement and crystallization of PLA in the matrix contributes to
better heat resistance of the composite.
To better visualize how the MFC reinforcement depends on the
rigidity of PLA, which is determined by temperature, we plotted
the storage modulus of composites divided by the storage modulus
of neat PLA (here referred to as the reinforcement efciency) as a
function of the modulus of neat PLA. Fig. 6a and b show the PLA
reinforcement efciency in the amorphous and crystallized states,
respectively. Both graphs show that the reinforcing efciency is
markedly high at a very low PLA modulus corresponding to high
temperatures. This means that MFC reinforcement is more effec-
tive at high temperatures, contributing to the heat resistance of
PLA composites.
Fig. 3. Typical stressstrain curves of neat PLA and PLA/MFC composites under (a)
amorphous and (b) crystallized states.
Table 2
Mechanical properties of amorphous and crystallized PLA and PLA/MFC composites at
room temperature.
Samples Condition Tensile modulus
(GPa)
Tensile strength
a
(MPa)
Strain at
break (%)
Neat PLA Amorphous 3.3 0.2 57.7 1.5 6.8 2.1
Crystallized 4.0 0.1 60.9 1.6 3.1 0.4
PLA/MFC 3 wt% Amorphous 3.8 0.1 61.4 1.6 2.7 0.2
Crystallized 4.4 0.1 63.6 0.7 2.0 0.1
PLA/MFC 5 wt% Amorphous 3.9 0.1 63.4 1.1 2.5 0.2
Crystallized 4.6 0.1 64.4 0.8 2.0 0.1
PLA/MFC 10 wt% Amorphous 4.5 0.1 65.4 1.9 2.2 0.1
Crystallized 4.7 0.1 66.2 3.5 2.0 0.2
PLA/MFC 20 wt% Amorphous 5.2 0.2 70.2 3.7 1.9 0.1
Crystallized 5.7 0.1 69.4 1.4 1.7 0.1
a
Maximum stress.
1190 L. Suryanegara et al. / Composites Science and Technology 69 (2009) 11871192
4. Conclusions
The thermal and mechanical properties of PLA/MFC nanocom-
posites were studied by using a semi-crystalline grade of PLA.
DSC measurements revealed that the presence of MFC can act as
nucleating agent on the crystallization of PLA. The addition of
MFC increased the tensile modulus of crystallized neat PLA by
42% and the strength by 14% at an MFC content of 20 wt%. DMA
measurements showed that the storage modulus of the crystallized
composite with an MFC content of 20 wt% was around 1 GPa at a
high temperature (120 C). These results demonstrated that MFC
could extend the application of PLA, particularly for products ex-
posed to high temperature.
Acknowledgements
The authors would like to thank Mitsui Chemicals, Inc., Japan
for supplying the PLA and Dr. K. Umemura, Dr. K. Abe, Dr. M. Nogi,
Fig. 5. Effect of MFC contents (wt%) on the temperature dependency of the storage
modulus under (a) amorphous and (b) crystallized states.
Fig. 6. The efciency of MFC reinforcement (E
0
composite divided by E
0
PLA at each
temperature) against rigidity (E
0
PLA) under (a) amorphous and (b) crystallized
states.
Fig. 4. MFC content dependence of (a) tensile modulus and (b) tensile strength, in the crystallized states.
L. Suryanegara et al. / Composites Science and Technology 69 (2009) 11871192 1191
and Dr. T. Nakatani, Research Institute for Sustainable Humano-
sphere, Kyoto University for their valuable suggestions and
discussions.
L. Suryanegara is indebted to the scholarship for doctoral
course from the Ministry of Education, Culture, Sports, Science
and Technology-Japan. This research was supported by a
Grant-in-Aid for Scientic Research (B) (No. 1538012, 2003.4
2007.3) from the Ministry of Education, Culture, Sports, Science
and Technology-Japan, and a Grant-in-Aid for Research and
Development for regional innovation consortium (No. 17S5018,
2006.9 2007.3) from the Ministry of Economy, Trade and
Industry, Japan.
References
[1] Saheb DN, Jog JP. Natural ber polymer composites: a review. Adv Polym
Technol 1999;18(4):35163.
[2] Garlota D. A literature review of poly (lactic acid). J Polym Environ
2001;9(2):6384.
[3] Auras R, Harte B, Selke S. An overview of polylactides as packaging materials.
Macromol Biosci 2004;4:83564.
[4] Mohanty AK, Misra M, Drzal LT. Sustainable bio-composites from renewable
resources: opportunities and challenges in the green materials world. J Polym
Environ 2002;9(2):1926.
[5] Samir MASA, Alloin F, Dufresne A. Review of recent research into cellulosic
whiskers, their properties and their application in nanocomposite eld.
Biomacromolecules 2005;6:61226.
[6] Oksman K, Skrifvars M, Selin JF. Natural bers as reinforcement in polylactic
acid (PLA) composites. Compos Sci Technol 2003;63(9):131724.
[7] Huda MS, Drzal LT, Misra M, Mohanty AK, Williams K, Mielewski DF. A study
on biocomposites from recycled newspaper ber and poly(lactic acid). Ind Eng
Chem Res 2005;44(15):5593601.
[8] Zimmermann T, Pohler E, Geiger T. Cellulose brils for polymer reinforcement.
Adv Eng Mater 2004;6(9):75461.
[9] Nakagaito AN, Yano H. Novel high-strength biocomposites based on
microbrillated cellulose having nano-order-unit web-like network
structure. Appl Phys A 2005;80(1):1559.
[10] Svagan AJ, Samir MASA, Berglund LA. Biomimetic polysaccharide
nanocomposites of high cellulose content and high toughness. J Polym
Environ 2007;8:255663.
[11] Seydibeyoglu M, Oksman K. Novel nanocomposites based on polyurethane
and micro brillated cellulose. Compos Sci Technol 2008;68:90814.
[12] Iwatake A, Nogi M, Yano H. Cellulose nanober reinforced polylactic acid.
Compos Sci Technol 2008;68:210316.
[13] Nam PH, Maiti P, Okamoto M, Kotaka T, Hasegawa N, Usuki A. A hierarchical
structure and properties of intercalated polypropylene/clay composites.
Polymer 2001;42:9633.
[14] Nam JY, Ray SS, Okamoto M. Crystallization behavior and morphology of
biodegradable polylactide/layered silicate nanocomposites. Macromolecules
2003;36:7126.
[15] Mathew AP, Oksman K, Sain M. The effect of morphology and chemical
characteristics of cellulose reinforcements on the crystallinity of polylactic
acid. J Appl Polym Sci 2006;101(1):30010.
[16] Mathew AP, Oksman K, Sain M. Mechanical properties of biodegradable
composites from poly lactic acid (PLA) and microcrystalline cellulose (MCC). J
Appl Polym Sci 2005;97(5):201425.
[17] Park SD, Todo M, Arakawa K. Effect of annealing on the fracture toughness of
poly(lactic acid). J Mater Sci 2004;39:11136.
1192 L. Suryanegara et al. / Composites Science and Technology 69 (2009) 11871192

Você também pode gostar