Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Composites: Part A 41 (2010) 11231129

Contents lists available at ScienceDirect

Composites: Part A
journal homepage: www.elsevier.com/locate/compositesa

Cardanolformaldehyde thermoset composites reinforced with buriti bers:


Preparation and characterization
Reginaldo da Silva Santos a, Alexandre Arajo de Souza a, Marco-Aurelio De Paoli b,*,
Cleide Maria Leite de Souza a
a
b

Departamento de Qumica, Universidade Federal do Piau, Campus Ininga, 64049-550 Teresina, PI, Brazil
Instituto de Qumica, Unicamp, C. Postal 6154, 13084-970 Campinas, SP, Brazil

a r t i c l e

i n f o

Article history:
Received 9 October 2009
Received in revised form 26 March 2010
Accepted 13 April 2010

Keywords:
A. Discontinuous reinforcement
B. Fiber/matrix bond
B. Mechanical properties
D. Mechanical testing

a b s t r a c t
Buriti (Mauritia exuosa) is a robust palm tree, abundant in the northeast of Brazil. Buriti bers are rich in
cellulose and have potential applications in polymer reinforcement. In this work, buriti bers were used:
as received, mercerized and in a silanized form. They were combined with cardanolformaldehyde thermoset resin and hot-pressed to form composites. Scanning electron micrograph showed that the mercerization process induced a roughness onto the ber surface. Dynamic mechanical analyses showed that
increasing the ber content provides an improvement in the rigidity of the composites. The mercerization
treatment showed the best results for ber/matrix adhesion.
2010 Elsevier Ltd. All rights reserved.

1. Introduction
At present there is a signicant trend to use renewable
resources to produce materials with appropriate mechanical properties for several applications. Composites using natural bers are
the best candidates for substituting metals and short berglass
reinforced thermosets. The mixture of materials in a composite
aims to improve the thermal and mechanical properties, giving
them more versatility. Composite materials have a broad range of
industrial applications, particularly in the automotive industry
[13]. The mixture of different materials allows the utilization of
synthetic and natural materials or the utilization of chemically
modied natural compounds.
The investigation of plant bers, such as cotton, curaua, coconut
and banana, dispersed in a polymeric matrix is a growing eld of
research, because of their low abrasiveness to the processing
equipment, reducing energy consumption and consequently lowering costs in the productive sector. Plant bers are biodegradable
and their composites with polymers can be thermally recycled,
yielding carbon credits. Besides, natural bers are very efcient
for sound absorption and fracture resistance when compared to
glass bers [4]. However, the natural bers have disadvantages
such low mechanical resistance, when compared to conventional
bers (asbestos, Kevlar, carbon and glass bers) and present the
* Corresponding author.
E-mail address: mdepaoli@iqm.unicamp.br (M.-A. De Paoli).
1359-835X/$ - see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.compositesa.2010.04.010

impossibility of high temperature processing because they start


to degrade at 200220 C [58].
Modications of the plant ber surfaces have been used to improve the chemical interaction of the components of composite
polymer materials [911]. Mercerization is currently used in the
rening process of cellulose bers. The process consists in treating
the material with concentrated sodium hydroxide solutions. The
aim of mercerization is to increase the number of hydroxyl groups
at the ber surface to strengthen the interactions with a polar
polymer matrix [12]. Another kind of modication of the ber surface is silanization, to increase the interaction with a non-polar
polymer matrix. In this process, a silanic coupling agent replaces
the hydroxyl groups, forming non-polar radicals, which avoids
the absorption of water molecules in the bulk of the matrix and
presents a good afnity for the polymer matrix [13].
Buriti (Mauritia exuosa) is a robust palm tree, with big leaves
arranged in the form of a fan. Different Buriti species are found
in the damp and swampy regions of Brazil, between 17 latitude
south and 3 latitude north and to the west of 40 longitude west.
Its fruit is used for nutrition [14,15] and the bers obtained from
the leaves are used for handicrafts. Investigation of the use of these
bers as reinforcement for polymers is an excellent alternative to
improve the development of these regions. Additionally, thermoset
polymers used as matrix in composites can be also obtained from
natural sources. Among these, the thermoset prepared form cashew nut shell liquid (CNSL) is an important candidate, because cashew nuts are also produced in the same geographic region.

1124

R. da Silva Santos et al. / Composites: Part A 41 (2010) 11231129

OH

C15H 31-2n
Cardanol
Fig. 1. Cardanol, a natural phenolic compound obtained from the cashew nut shell
liquid (CNSL).

The cashew tree (Anacardium occidentale) is found in the northeast region of Brazil, mainly in the Piau and Cear states. During
industrial processing, the cashew nut liberates the CNSL. The major
component in industrial CNSL is cardanol, Fig. 1, a phenolic
compound with a long unsaturated hydrocarbon chain (C15) substituted in the meta position [1618]. This hydrocarbon chain acts as
an internal plasticizer, giving exibility to the phenolformaldehyde resins that can be formed with it. Cardanol condenses with
formaldehyde producing a polymer network similar to phenol
formaldehyde resins. Misra and Pandey investigated in detail the
reaction mechanism of cardanolformaldehyde resin formation
[19].
In this work we describe in detail the preparation of the cardanolformaldehyde thermoset resin, chemical treatment of the
buriti bers and the composite preparation. The natural bers used
in the preparation of the composite were obtained from the straw
of buriti leaves. The aims of this work are: (1) to prepare composite
materials from thermoset cardanolformaldehyde resin using
buriti bers as reinforcement in different proportions; (2) to investigate the morphologies of the bers before and after treatment
and of the composite materials; (3) to study the thermal and dynamic mechanical properties of the composite materials using
thermogravimetry, differential scanning calorimetry and dynamic
mechanical analysis.

at break 3.3 0.7% and modulus 10.2 2.2 GPa. The density, measured using a helium picnometer, was 1.312 0.009 g cm3.
For mercerization, the buriti bers were treated with a 20% (w/
v) aqueous NaOH solution for 60 min at 25 C, washed with distilled water until neutral pH and dried in an oven at 60 C for
24 h, yielding mercerized bers (MBF).
For silanization the buriti bers were immersed in a 2% (v/v) 3aminopropyltrimethoxysilane solution in ethanol (95% purity) for
5 min. The pH was adjusted to 5.5 using acetic acid. To hydrolyze
the coupling agent the bers were exposed to air for 30 min. The
reaction between the hydroxyl groups from the ber and the silane
was induced by heating in an oven at 60 C for 2 h [20,21], yielding
silanized bers (SBF).
2.3. Preparation of the composites

2. Experimental procedure

One milliliter a 2.0 mol/L NaOH solution was added to 11.5 g of


cardanol, under mechanical stirring, in a water bath at 90 C. After
dissolution, 2.3 g of formaldehyde were added to the mixture and
it was stirred further for 60 min. Ground and sieved buriti bers
were added 20 min before the end of the stirring time. The mixture
was transferred to a stainless steel mold, previously coated with
aluminum foil to facilitate the removal of the material, and left
at room temperature for 12 h. During this period, air bubbles
formed in the interior of the resin were released. Finally, the mixture was pressed under 6.0 ton for 1.5 h, at 150 C, to obtain the
cured composite material.
The composite materials were prepared using pristine bers
and the chemically treated bers (mercerized or silanized). The ber content in the composites was varied, using: 5, 10 and 15 wt.%.
Composites with silanized bers were denominated SC-5, SC-10
and SC-15, respectively. Composites prepared with mercerized bers were denominated MC-5, MC-10 and MC-15, respectively.
Composites with higher concentration of bers could not be prepared because the low density of the bers precluded its mixture
with the cardanolformaldehyde resin.
To compare the effect of the chemical treatment of the bers, a
composite with 5 wt.% of pristine (untreated) ber was prepared
and denominated UTC-5. The pure thermoset resin was denominated CFR.

2.1. Materials

2.4. Composite characterizations

Green leaves of the native specimen M. exuosa were collected


in Teresina (Brazil, 550 latitude S and 42480 longitude W), during
September 2005. These were dried for 2 weeks using outdoor
exposure, until the leaves attained the characteristic color of dry
straw. The straw was manually separated from the tallows and
transversally cut in strips of 5.010 mm length, which were subsequently dried in oven at 60 C for 24 h. The strips were ground and
sieved to 0.250 and 0.300 mm. The plant bers retained at
0.250 mm were chemically modied and used as discontinuous
phase for composites preparation. Cardanol was supplied by the
Parque de Desenvolvimento Tecnolgico (PADETEC, Universidade
Federal do Cear). Other reagents were: formaldehyde (95% purity,
from Synth), 2.0 mol/L NaOH (98% purity, Vetec) and 3-aminopropyltrimethoxysilane (97% purity, Aldrich).

A DSC (TA Instrument DSC-2920) was used for thermal analysis


of the resin, bers, and composites. Samples of 1.8 mg were placed
in aluminum pans. A constant nitrogen ow of 50 mL/min was
used to purge the instrument. The samples were heated from 30
to 500 C at a rate of 10 C/min. Surfaces of cryogenically fractured
samples were examined with a Zeiss DSM-940A scanning electron
microscope (SEM), to investigate the effects of the chemical treatments on the bers and on the adhesion at the bers/resin interface. For this purpose, gold vapor deposition was used to have a
conductive layer.
The DMA curves were obtained from a TA Instruments Q800
Dynamic Mechanical Analyzer. The measurements were done
using the dual cantilever accessory. During the experiments with
dynamic temperature, the samples were submitted to exural
modes. The experimental conditions were: 20 lm oscillation
amplitude, 1 Hz frequency, 3 C/min heating rate and temperature
range from 35 to 300 C. The temperature of the equipment was
stabilized at 35 C for 5 min, before starting the experiments. The
dimensions of the samples were 60  13  3.2 mm. The stress
strain responses of the material were obtained in the exural mode
using the 3-point bend clamp. A force of 0.500 N/min was applied
up to rupture of the sample or end of the experiment (the limiting
force applied was 18 N). The equations used for calculation of the

2.2. Fiber characterization and treatments


Untreated bers (UTBF) with an average diameter of 182
12 lm and 50 mm length were characterized by their mechanical
tension properties in stressstrain measurements in an EMIC
DL-2000 universal testing machine using 20 specimens. Prior to
the measurements, the bers were conditioned for 24 h at 25 C
and 50% humidity. Results were: yield stress = 271 72 MPa, strain

1125

R. da Silva Santos et al. / Composites: Part A 41 (2010) 11231129

exural stress and strain were rf = 3PL/2bh2 and ef = 6Dh/L2 respectively, where P is the load (N), L is the support span (mm), D is the
maximum deection of the center of the sample (mm), b and h are
the width and dept (mm) of the samples, respectively [22].
3. Results and discussion
3.1. Thermogravimetric analysis
The TG curves for UTBF, SBF and MBF are shown in Fig. 2a and
present two main mass loss processes. The rst occurs between 60
and 100 C with a mass loss of ca. 8%, which can be attributed to
elimination of water from the silanized and untreated bers. This
mass loss is due to the hydrophilicity of the plant bers [2325].
The second process of mass loss starts at ca. 150 C for the UTBF
and SBF samples. This process can be attributed to thermal decomposition of the constituents of plant bers (pectin, lignin and hemicellulose). For MBF this process showed a signicant shift to higher
temperatures (230 C) suggesting an increased thermal stability
in comparison to UTBF and SBF. In addition, the MBF showed only
3.0% of mass loss attributed to elimination of water. Residues of
36%, 34% and 25%, at 800 C, for SBF, MBF and UTBF, respectively,
are probably due to the non-oxidizing atmosphere used in the
experiment.
The comparison between the TG curves for MBF, MC-10 and
CFR, Fig. 2b, indicate that the composite presents an intermediate
thermal behavior in relation to the ber and the matrix. The TG
curve for CFR and MC-10 showed a low mass loss starting at ca.
70 C, associated to the elimination of water obtained as a product
during the thermoset cure reaction. The composite curve shows an
inection at ca. 350 C, due to the thermal degradation of the ber
and at ca. 400 C, due to the thermal degradation of the thermoset
resin, as compared to the corresponding mass loss temperature of
the pure thermoset.
3.2. Differential scanning calorimetry
The DSC curves for the composite (with 10 wt.% of mercerized
ber), pure resin, untreated and treated bers are shown in
Fig. 3. The curves in Fig. 3a, for UTBF, MBF and SBF, show endothermic peaks between 60 and 80 C assigned to water loss. The water
loss for MBF occurred at temperatures higher than for the UTBF
and SBF samples. This probably occurs due to the higher number
of hydroxyl groups at the surface of the mercerized bers, which
allows for stronger hydrogen bonding with the water. The large
exothermic events that occur between 200 and 370 C are assigned

(a)

UTBF
SBF
MBF

The SEM micrographs of the UTBF, MBF and SBF samples, Fig. 4,
show the morphological changes on the bers surfaces due to the
different chemical treatments applied. The SEM image for mercerized ber (Fig. 4b) shows a rough aspect, probably due to the removal of low molar mass compounds, which occurs during
mercerization, leaving cavities at the surface. The high roughness
of the ber surface provides a larger number of anchorage points
for the polymeric matrix, increasing the interaction between the
components of the composite. The silanized ber, shown in
Fig. 4b, presents a more regular surface. This suggests that silanization produces a lm coating the entire ber surface.

(b)

MBF
MC-10
CFR

100

80

80

60

40

20

3.3. Scanning electron microscopy

Weight (%)

Weight (%)

100

to the decomposition process of the major components of the plant


bers, as discussed above [26].
For the mercerized bers, the weak DSC shoulders, assigned to
the decomposition of the constituents of the ber, are overlapped,
suggesting partial removal of these constituents during mercerization. Residues present in the mercerized bers decompose at
slightly higher temperatures, when compared to the untreated
and silanized bers. The DSC peaks, assigned to the decomposition
of lignin, appear at ca. 340 C in the curves for UTBF, MBF and SBF.
Studies developed by rfo et al. [27] revealed that the thermal
decomposition of lignocellulosic bers can be modeled using three
independent rst order reactions of three components, the rst
and second corresponding to cellulose and hemicellulose and the
third to lignin. All DSC curves obtained for untreated and treated
buriti bers (Fig. 3a) show large exothermic peaks between 200
and 370 C. Considering this model the low temperature exothermic process corresponds to cellulose and hemicellulose degradation and the high temperature part corresponds to lignin
degradation. Bilba and Ouensanga [28] investigated the thermal
behavior for sugarcane bagasse using FTIR analysis. They suggested
that, near 380 C a decrease of intensity of bands assigned to unsaturations and a formation of bands assigned to alkyl bonds occur.
This will correspond to lignin degradation, with liberation of water,
CO and CO2.
The cardanolformaldehyde resin and the composite with
10 wt.% of mercerized bers (MC-10) showed a weak endothermic
event in the DSC curve, Fig. 3b, assigned to water liberation during
formation of the resin. The intense exothermic peak that appears
between 260 and 370 C in the DSC curves for the composite
MC-10 corresponds to the thermal decomposition of the CFR component in this sample. These results also corroborate with the thermogravimetric measurements discussed above.

60

40

20

100 200 300 400 500 600 700 800 900

Temperature (C)

100 200 300 400 500 600 700 800 900

Temperature (C)

Fig. 2. Thermogravimetric curves under inert atmosphere for: (a) mercerized (MBF), silanized (SBF) and untreated (UTBF) buriti bers and (b) composite with 10 wt.% MBF
(MC-10), mercerized ber and cardanolformaldehyde resin (CFR).

1126

R. da Silva Santos et al. / Composites: Part A 41 (2010) 11231129

(a)

(b)

Exo

MBF

SBF
UTBF

50

100

150

200

250

MC-10

Heat flow (a. u.)

Heat flow (a.u)

Exo

300

350

400

CFR
MBF

50

Temperature (C)

100

150

200

250

300

350

400

Temperature (C)

Fig. 3. DSC curves for: (a) untreated, mercerized and silanized buriti bers and (b) mercerized ber, cardanolformaldehyde thermoset resin and for the composite
containing 10 wt.% of mercerized buriti ber. Symbols in the curves were used only to identify the samples, they do not correspond to single experimental results.

90 m

90 m

400 m

(a)

400 m

(b)

(c)

400 m

(d)

90 m

400 m

Fig. 4. SEM micrographs: (a) untreated buriti ber; (b) mercerized buriti ber; (c) silanized buriti ber (inserts show details with higher magnication) and (d) composite
MC-10 sample (arrows indicate the bers).

The SEM micrograph for the MC-10 composite sample, Fig. 4d,
shows a lamellar like morphology for the polymeric matrix. This
may confer mechanical fragility to this material. As can be observed in Fig. 4d, there is an excellent adhesion between the matrix
and the mercerized bers (indicated by the arrows in Fig. 4d). In
the case of this composite, the good coverage of the ber by the resin may prevent its exposure to air, improving its stability to
oxidation.
Phenolic thermoset resins usually present micropores that provide low thermal resistance [29,30]. In the case of CFR and the

composite samples, the micropores are formed during the process


of water elimination, resulting from the condensation reactions
[31]. These micropores can act as stress concentrators, lowering
the charge transportation capacity and energy absorption by the
composite.
3.4. Dynamic mechanical analysis
Fig. 5a presents the dynamic mechanical analysis, DMA, curves
for the thermoset and the composites prepared with 5 wt.% of

1127

R. da Silva Santos et al. / Composites: Part A 41 (2010) 11231129

400
350

tan

300

CFR
MC-5
MC-10
MC-15

tan

0.20

600

150

450

0.12

300

0.08

0.08
150

0.04

0.04
50

E'

E'
0

50

100

150

200

250

300

0.00

50

100

Temperature (C)

(d)

0.28

200

250

300

0.00

21

0.20

18

0.16
0.12

100

tan

200

24

0.24

E" (MPa)

CFR
CS-5
CS-10
CS-15

tan

150

Temperature (C)

400

300

E' (MPa)

0.16

tan

0.12

200

100

(c)

0.20

0.16

250

tan

E' (MPa)

(b) 750

0.24

CFR
UTC-5
MC-5
SC-5

E' (MPa)

(a)

RCF
CM-5
CS-5
CUT-5

15
12

0.08

0.04

0.00

E'

0
0

50

100

150

200

250

300

Temperature (C)

50

100

150

200

250

300

Temperature (C)

Fig. 5. Variation of E0 modulus and tan d as a function of temperature for the composites with: (a) 5 wt.% mercerized, silanized and natural ber; (b) with 0, 5, 10 and 15 wt.%
of mercerized bers; (c) with 0, 5, 10 and 15 wt.% of silanized bers and (d) variation of E00 modulus for the composites with 5 wt.% of mercerized, silanized and natural ber,
in function of temperature.

untreated and treated bers, to investigate the inuence of the


chemical treatment on the mechanical characteristics of the composites. It was observed that, as the temperature increases to near
150 C, a lowering in the values of the E0 modulus occurs for all the
materials, indicating lowering of the rigidity with the increase of
temperature [32]. At temperatures above 150 C, in the CFR resin,
an increase in the number of crosslinking bonds, due to the curing
process, probably occurs.
The curves of E0 modulus for CFR and the SC-5 sample (Fig. 5a)
shows the lowest values for this modulus, when compared to the
curve obtained for MC-5 and UTC-5 samples, suggesting that the
mercerization and the insertion of the bers in the polymeric matrix
provide an increase of the rigidity of the material due to the reinforcement effect of the bers that allows a higher degree of stress
transfer at the ber/matrix interface. The low value of the E0 modulus
observed for the SC-5 sample suggests that silanization does not
produce an effective improvement on the adhesion of the ber to
the matrix.
The tan d value can be assumed as an indication of the intensity of
the ber/matrix interface interaction. When the interaction is strong
the mobility of the polymeric segments is low at the interface, and
the tan d values are lowered [3335]. The DMA curves for the composites UTC-5 and MC-5 samples (Fig. 5a) shows a reduction of the
tan d values, when compared to the RCF sample. The DMA curve
for the SC-5 composite sample shows higher tan d values, making
it evident that, when the composite material presents a weak interfacial interaction between the ber and the matrix, it tends to dissipate more energy, exhibiting higher tan d values, in comparison to
the materials that present stronger interfacial interactions.

Fig. 5b shows the curves of E0 modulus and tan d for the composites with different concentrations of mercerized bers. The increase
in the mercerized ber/resin ratio provides an increase in the E0 modulus, and a lowering in tan d. This indicates an increased rigidity and
lower energy dissipation for the composites richer in bers. Additionally, a distinct behavior was observed for composites obtained
with silanized bers, as observed in Fig. 5c, showing that increasing
the bers content, up to 15 wt.%, provides a lowering of the E0 modulus and an increase of tan d. This suggests a lower ber/matrix
interaction, reducing the stiffness of the composite material.
Fig. 5d shows the variation of loss modulus (E00 ) versus temperature for the thermoset and the composites with untreated and
treated bers. The curves show two badly dened peaks, the rst
at ca. 80 C, and the second at ca. 175 C. The rst peak suggests
the occurrence of chain segment movements, localized at the
crosslinking bonds, i.e. similar to a glass transition (Tg). The second
peak is assigned to the curing of the resin. It is also observed that a
relative decrease in the amount of resin in the material provides a
lowering in the intensity of the Tg peak. This fact suggests a reduction of the free volume in the composite.
The stressstrain behavior of composites with different buriti ber treatments and contents are shown in Fig. 6. The stressstrain
curves for the CFR and the composites with 5 wt.% of ber, shown
in Fig. 6a, show that the pure thermoset is more fragile than the
composites, with lower yield stress and strain at break. Although
the module of the pure thermoset and the composites are very
similar, the composite with mercerized bers showed the higher
stress at yield. This corroborates again with the better ber to matrix adhesion in this composite.

1128

R. da Silva Santos et al. / Composites: Part A 41 (2010) 11231129

(a)

(b)

MC-5

MC-10
UTC-5
2

CFR

MC-5

MC-15

SC-5

Stress (MPa)

Stress (MPa)

CFR
1

10

12

Strain (%)

10

12

Strain (%)

Fig. 6. Stressstrain curves for: (a) thermoset resin compared to the composites with 5 wt.% of chemically treated bers and (b) composites containing 0, 5, 10 and 15 wt.% of
mercerized bers.

The behavior in Fig. 6b shows that a higher mercerized ber


content in the composite contributes to an increase in the tensile
modulus and yield stress. However, for the composites with 10
and 15 wt.% of mercerized bers the stress is beyond the detection
limit of the equipment and the strain at break could not be measured. The increase in modulus, by itself, is a strong evidence for
the reinforcing effect exerted by the bers.
4. Conclusions
Composites prepared from cardanolformaldehyde thermoset
resin and buriti bers were prepared. Thermogravimetric and
DSC curves showed that the composites have an intermediate thermal behavior compared to those presented by the bers and the
pure matrix. The DSC curves showed that the liberation of water
for MBF samples occurred at higher temperatures than for UTBF
and SBF samples, suggesting a higher number of hydroxyl groups
at the surface of the mercerized bers.
The SEM micrographs showed that removal of substances from
the ber surface occurs during mercerization. The cavities formed
allow better contact with the polymeric matrix, improving the ber/matrix adhesion. This reects in the mechanical properties of
the materials. DMA data conrm this behavior through the increase of the E0 modulus and the decrease of the tan d values.
The lamellar morphology presented by the resin probably is
responsible for the low mechanical resistance observed in the resin
and in the composites.
The DMA curves also show that, increasing the ber content in
the material provides a higher rigidity. The mercerization treatment showed the best results for ber/matrix adhesion.
Acknowledgements
RSS thanks CAPES for a fellowship. The authors thank LIEC of
UFSCar for the SEM analyses, PADETEC at the Universidade Federal
do Cear and Lapetro/UFPI for the utilization of the DSC, TG and
DMA equipments. We also thank FAPESP (04/15084-6) for nancial
support to MAP.
References
[1] Netravali AN, Chabba S. Composites get greener. Mater Today 2003;6(4):229.
[2] Huang X, Netravali A. Characterization of ax ber reinforced soy protein resin
based green composites modied with nano-clay particles. Compos Sci
Technol 2007;67(10):200514.

[3] Bodros E, Pillin I, Montrelay N, Baley C. Could biopolymers reinforced by


randomly scattered ax bre be used in structural applications? Compos Sci
Technol 2007;67(3):46270.
[4] Trindade WG, Hoareau W, Megiatto JD, Razera IAT, Castellan A, Frollini E.
Thermoset phenolic matrices reinforced with unmodied and surface-grafted
furfuryl alcohol sugar cane bagasse and curaua bers: properties of bers and
composites. Biomacromolecules 2005;6(5):248596.
[5] Geethamma VG, Kalaprasad G, Groeninckx G, Thomas S. Dynamic mechanical
behavior of short coir ber reinforced natural rubber composites. Composites:
Part A 2005;36(11):1499506.
[6] ODonnell A, Dweib MA, Wool RP. Natural ber composites with vegetal oilbased resin. Compos Sci Technol 2004;64(9):113545.
[7] Wielage B, Lampke T, Utschick H, Soergel F. Processing of natural-bre
reinforced polymers and the resulting dynamic-mechanical properties. J Mater
Process Technol 2003;139(1):1406.
[8] Ray D, Sarkar BK, Das S, Rana AK. Dynamic mechanical and thermal analysis of
vinylester-resin-matrix composites reinforced with untreated and alkalitreated jute bres. Compos Sci Technol 2002;62(7):9117.
[9] Beckermann GW, Pickering KL. Engineering and evaluation of hemp bre
reinforced polypropylene composites: bre treatment and matrix
modication. Composites: Part A 2008;39(6):97988.
[10] Arajo JR, Waldman WR, De Paoli MA. Thermal properties of high density
polyethylene composites with natural bres: coupling agent effect. Polym
Degrad Stabil 2008;93(10):17705.
[11] Li Y, Mai YW, Ye L. Sisal bre and its composites: a review of recent
developments. Compos Sci Technol 2000;60(11):203755.
[12] Herrera-Franco PJ, Valadez-Gonzlez A. A study of the mechanical properties of
short natural-ber reinforced composites. Composites: Part B 2005;36(8):
597608.
[13] Salon MCB, Abdlmouleb M, Bou S, Belgacem MN, Gardini A. Silane adsorption
onto cellulose bers: hydrolysis and condensation reactions. J Colloid Interf Sci
2005;289(1):24961.
[14] Frana LF, Reber G, Meireles MAA, Machado NT, Brunner GJ. Supercritical
extraction of carotenoids and lipids from buriti (Mauritia exuosa), a fruit from
the Amazon region. J Supercrit Fluids 1999;14(3):24756.
[15] Barbosa RI, Fearnside PM. Wood density of trees in open savannas of the
Brazilian Amazon forest. Ecol Manage 2004;199(1):11523.
[16] Chuayjuljit S, Rattanametangkool P, Potiyaraj P. Preparation of cardanol
formaldehyde resins from cashew nut shell liquid for the reinforcement of
natural rubber. J Appl Polym Sci 2007;104(3):19972007.
[17] Mwaikambo LY, Ansell MP. Hemp bre reinforced cashew nut shell liquid
composites. Compos Sci Technol 2003;63(9):1297305.
[18] Mwaikambo LY, Ansell MP. Cure characteristics of alkali catalysed cashew nut
shell liquidformaldehyde resin. J Mater Sci 2001;36(15):36938.
[19] Misra AK, Pandey GN. Kinetics of alkaline-catalyzed cardanol formaldehyde
reaction II. Mechanism of the reaction. J Appl Polym Sci 1985;30(3):96977.
[20] Bledzki AK, Glassan J. Composites reinforced with cellulose based bres. Prog
Polym Sci 1999;24(2):22174.
[21] Paul A, Joseph K, Thomas S. Effect of surface treatments on the electrical
properties of low-density polyethylene composites reinforced with short sisal
bers. Compos Sci Technol 1997;57(1):6779.
[22] Munoz JC, Ku H, Cardona F, Rogers D. Effect of catalyst and post-curing
conditions in the polymer network of epoxy and phenolic resins: preliminary
results. J Mater Process Technol 2008;202(1):48692.
[23] Wielage B, Lampke Th, Marx G, Nestler K, Starke D. Thermogravimetric and
differential scanning calorimetric analysis of natural bres and polypropylene.
Thermochim Acta 1999;337(1):16977.
[24] Scheirs J, Camino G, Tumiatti W. Overview of water evolution during the
thermal degradation of cellulose. Eur Polym J 2001;37(5):93342.

R. da Silva Santos et al. / Composites: Part A 41 (2010) 11231129


[25] Garcia-Perez M, Chaala A, Yang J, Roy C. Co-pyrolysis of sugarcane bagasse with
petroleum residue. Part I: thermogravimetric analysis. Fuel 2001;80(9):
124558.
[26] Hatakeyama T, Nakamura K, Hatakeyama H. Vaporization of bound water
associated with cellulose bres. Thermochim Acta 2000;352353:2339.
[27] rfo JJM, Antunes FJA, Figureueredo JL. Pyrolysis kinetics of lignocellulosic
materials three independent reactions model. Fuel 1999;78(3):34958.
[28] Bilba K, Ouensanga A. Fourier transform infrared spectroscopic study of thermal
degradation of sugar cane bagasse. J Anal Appl Pyrolysis 1996;38(1):6173.
[29] Wang S, Adanur S, Jang BZ. Mechanical and thermo-mechanical failure
mechanism analysis of ber/ller reinforced phenolic matrix composites.
Composite: Part B 1997;28(3):21531.
[30] Paiva JMF, Frollini E. Sugarcane bagasse reinforced phenolic and lignophenolic
composites. J Appl Polym Sci 2002;83(4):8808.

1129

[31] Mafezzoli A, Cal E, Zurlo S, Mele G, Tarzia A, Stifani C. Cardanol based matrix
biocomposites reinforced with natural bres. Compos Sci Technol
2004;64(6):83945.
[32] Mohanty S, Verma SK, Nayak SK. Dynamic mechanical and thermal properties of
MAPE treated jute/HDPE composites. Compos Sci Technol 2006;66(3):53847.
[33] Menard PK. Dynamic mechanical analysis: a practical introduction. New
York: RCR Press LLC; 1999.
[34] Pothan LA, Thomas S. Polarity parameters and dynamic mechanical behaviour
of chemically modied banana ber reinforced polyester composites. Compos
Sci Technol 2003;63(9):123140.
[35] Keusch S, Haessler R. Inuence of surface treatment of glass bres on the
dynamic mechanical properties of epoxy resin composites. Composites: Part A
1999;30(8):9971002.

You might also like