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

Materials and Design 47 (2013) 9–15

Contents lists available at SciVerse ScienceDirect

Materials and Design


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

Influence of fiber content on the mechanical and dynamic mechanical properties


of glass/ramie polymer composites
Daiane Romanzini a, Alessandra Lavoratti a, Heitor L. Ornaghi Jr. b, Sandro C. Amico b, Ademir J. Zattera a,⇑
a
Program of Postgraduate in Process and Technology Engineering (PGEPROTEC), University of Caxias do Sul, Rua Francisco Getúlio Vargas, 1130, 95070-560 Caxias do Sul, RS, Brazil
b
Program of Postgraduate Studies in Mining, Metals and Materials Engineering (PPGEM), Federal University of Rio Grande do Sul, Av. Bento Gonçalves, 9500, 91501-970 Porto Alegre,
RS, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: The combination of glass and ramie fibers with a polyester matrix can produce a hybrid material that is
Received 19 October 2012 competitive to all glass composites (e.g. those used in the automobile industry). In this work, glass and
Accepted 12 December 2012 ramie fibers cut to 45 mm in length were used to produce hybrid polymer composites by resin transfer
Available online 22 December 2012
molding (RTM), aiming to evaluate their physical, mechanical and dynamic mechanical properties as a
function of the relative glass–ramie volume fractions and the overall fiber content (10, 21 and
Keywords: 31 vol.%). Higher fiber content and higher ramie fiber fraction in the hybrid composites yielded lower
Hybrid composites
weight composites, but higher water absorption in the composite. The mechanical properties (impact
Vegetable fiber
Dynamic mechanical analysis
and interlaminar shear strength) of the composites were improved by using higher fiber content, and
the composite with 31 vol.% of reinforcement yielded the lowest value for the reinforcement effective-
ness coefficient C, as expected. Although the mechanical properties were improved for higher fiber con-
tent, the glass transition temperature did not vary significantly. Additionally, as found by analyzing the
adhesion factor A, improved adhesion tended to occur for the composites with lower fiber content (10%)
and higher ramie fiber fraction (0:100) and the results for the adhesion factor A did not correspond to
those found by the analysis of the tan delta peak height.
Ó 2012 Elsevier Ltd. All rights reserved.

1. Introduction fiber with length ranges within 150–200 mm [5]. The high mois-
ture absorption of natural fibers is one of their deleterious charac-
Hybrid composites are one of the emerging fields in polymer teristics that can benefit from hybridization, in order to avoid the
science that are gaining attention for application in various sectors, reduction in mechanical properties of their composites [1,6].
such as building, aeronautical and automotive [1]. The concept of Dynamic mechanical analysis (DMA) has been used for the
hybridization provides flexibility to the design engineer to tailor characterization of polymers to estimate their dynamic modulus
the material properties according to particular requirements [2], (E) and damping behavior (tan delta) [7]. DMA analysis can deter-
and the behavior of the hybrid composites appears to be, in gen- mine the primary relaxations, such as the glass transition (Tg) [8]
eral, a weighted sum of the contributions of the individual compo- and other parameters, such as dynamic fragility [9], crosslinking
nents in which there is a balance between their advantages and density [10] and the non-Arrhenius variation of relation times with
disadvantages [3]. temperature [11].
Among the vegetable fibers used for composites, the bast fibers, The present study focuses on the evaluation of the physical,
extracted from the stem of plants, e.g. kenaf, flax, ramie and hemp, mechanical and dynamic mechanical properties of glass–ramie
are usually accepted to show good mechanical properties com- hybrid polymer composites as a function of the overall fiber
pared with vegetable fibers extracted from the leaf (e.g., sisal and content (10, 21 and 31 vol.%) and the glass/ramie volume fractions
banana) or from the seed (e.g., coconut, cotton and palm oil) [4]. (0:100/25:75/50:50/75:25) in the composite.
China, Philippines and Brazil are the largest producers of ramie
fiber, which is a plant that belongs to the Urticaceae family derived 2. Materials and methods
from the bast of Boehmeria nivea and Boehmeria tenacissima, whose
2.1. Materials
⇑ Corresponding author. Tel.: +55 (54) 3218 2371; fax: +55 (54) 3218 2253.
E-mail addresses: dairomanzini@ibest.com.br (D. Romanzini), alelvt@gmail.com
Ramie roving was purchased from Sisalsul Fibras Naturais (São
(A. Lavoratti), ornaghijr.heitor@yahoo.com (H.L. Ornaghi Jr.), amico@ufrgs.br (S.C. Paulo/SP, Brazil) and glass fiber roving (EC 2400 P207) from CPIC
Amico), ajzattera@ucs.com.br (A.J. Zattera). Fiberglass do Brasil (Capivari/SP, Brazil). Unsaturated polyester

0261-3069/$ - see front matter Ó 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.matdes.2012.12.029
10 D. Romanzini et al. / Materials and Design 47 (2013) 9–15

resin UCEFLEX UC 5530-M was supplied by Elekeiroz S.A (Varzea According to the literature [14], the adhesion efficiency can be
Paulista/SP, Brazil). Mold-releasing agent poly(vinyl alcohol) estimated by the adhesion factor A, which is calculated using the
(PVA), the curing agent methyl ethyl ketone peroxide in di-isobutyl following equation:
phthalate (BUTANOX LPT) and the catalyst dimethylaniline (DMA)
were purchased from Disfibra (Caxias do Sul/RS, Brazil). 1 tan dc
A¼ 1 ð2Þ
1  V f tan dp

2.2. Mat manufacturing and composite preparation where tan dc and tan dp are the relative damping of the composite
and the polymer, respectively, at the glass transition temperature
Glass and ramie fibers were first cut to 45 mm length. The nat- (obtained from the maximum tan delta peak), and Vf is the fiber vol-
ural fiber was immersed in distilled water for 10 min and then un- ume fraction.
twisted and washed (with distilled water at 20–25 °C) for 50 min.
The fiber was oven-dried at 105 °C with air circulation for 60 min.
Then, the fibers were manually mixed and arranged in a pre-mold 3. Results and discussion
of the same shape of the mold to produce a hybrid mat. Three over-
all fiber content of the hybrid mats were studied, 10, 21 and 3.1. Physical characterization
31 vol.%, which was the maximum possible fiber load due to the
design of the mold. The influence of fiber length and hybridization The density (g/cm3) and water absorption (%) values for the
has already been reported in a previous study of the group [5]. The composites are shown in Table 1. Density increased with increas-
relative volume fraction between glass fiber (GF) and ramie fiber ing glass fiber content due to the higher density of the glass fiber
(RF) were as follow: 0:100, 25:75, 50:50, and 75:25 (GF:RF). It (2.5 g/cm3 for glass and 1.49 g/cm3 for ramie [5]), but water
was not possible to obtain pure glass composites due to severe absorption decreased. The reduction in water absorption occurs
fiber washing during molding. due to the reduced water absorption of the glass fibers in compar-
Before molding, the mats were pressed under the following con- ison with the natural fibers, in which the many OH groups of the
ditions: 10 min, 49 kN and 80 °C. Additionally, a mold-releasing cellulose are more exposed, thereby increasing the hydrophilicity
agent (PVA) was applied to the mold. The polyester resin was of the system [15].
mixed with 0.5 wt% of Butanox LPT (curing agent) and 0.1 wt% of Overall, a higher fiber content and glass relative volume fraction
dimethylaniline (catalyst). The process parameters used in the (compared to the same overall volume fraction) resulted in higher
RTM were mold temperature between 20 and 25 °C and positive density composites and lower water absorption. Similar behavior
pressure of 0.5 bar. The composites were cured at 25 °C for 24 h, was reported by Athijayamani et al. [6], studying randomly ori-
followed by a first post-curing at 80 °C for 6 h and a second post- ented natural fibers (roselle and sisal)/polyester hybrid composites
curing at 120 °C for 2 h. The first post-curing ensures that no fur- in different fiber contents (10–30 wt%), who reported that mois-
ther chemical modification occurs, and the second post-curing ture absorption increased with fiber incorporation due to greater
eliminates mechanical residual stress [12]. micro-void formation in the matrix.

3.2. Mechanical characterization


2.3. Characterization

Fig. 1 shows the impact strength results of the composites. The


Density and water absorption were evaluated according to
ASTM: D792-08 and ASTM: D570-10, respectively. Izod impact test composites showed higher impact strength than the pure resin,
and when glass fiber was added to the composite, impact strength
was performed using a CEAST impact machine in accordance with
ASTM: D256-10. Unnotched specimens (dimensions: significantly increased for all glass–ramie fractions. Additionally,
impact strength increases with increasing overall fiber content ex-
63.5  12.7  4 mm) were prepared and an average value from
ten samples for each family is reported. Interlaminar shear cept for the pure ramie composite, possibly due to poor overall
quality of the composite and to large fiber-to-fiber contact in
strength (ILSS) was carried out in a universal testing machine EMIC
DL-3000 in accordance with ASTM: D2344-06. A span-to-depth ra- which case matrix breakage becomes the predominant failure
tio of 4:1 and a bending rate of 1.0 mm/min were used and 10 mechanism [2].
specimens were tested in each case. All tests were conducted at
23 ± 2 °C and with relative humidity of 50 ± 5%. Table 1
The cross-section surface of the hybrid composites (after cryo- Density (g/cm3) and water absorption (%) of the composites.
genic fracture) was examined using a scanning electron micro-
Fiber (GF:RF) Density Water Water absorption at
scope (SEM-JEOL JSM-6060). All specimens were sputtered with a content (g/cm3) absorption at saturation 672 h (%)
gold layer prior to SEM observations. (vol.%) 24 h (%)
Dynamic mechanical properties were assessed using a Dynamic 10% 0:100 1.19 (±0.01) 0.58 ± 0.11 1.78 ± 0.16
Mechanical Analyzer DMA 2980. Analysis of the specimens 25:75 1.23 (±0.01) 0.56 ± 0.08 1.67 ± 0.21
(dimensions: 60  10  4 mm) were performed in dual cantilever 50:50 1.25 (±0.02) 0.42 ± 0.05 1.18 ± 0.02
mode (oscillation amplitude: 15 lm) at 1 Hz frequency, from room 75:25 1.29 (±0.02) 0.30 ± 0.04 0.93 ± 0.04

temperature to 180 °C (heating rate of 3 °C/min). The effectiveness 21% 0:100 1.21 (±0.01) 1.38 ± 0.47 3.46 ± 0.36
coefficient C of the reinforcement [13] was calculated using the fol- 25:75 1.28 (±0.01) 1.08 ± 0.39 2.87 ± 0.89
50:50 1.33 (±0.03) 0.68 ± 0.28 2.04 ± 0.58
lowing equation: 75:25 1.41 (±0.02) 0.55 ± 0.26 1.41 ± 0.51
31% 0:100 1.21 (±0.01) 2.20 ± 0.43 5.52 ± 0.67
E0g =E0r ðcompositeÞ 25:75 1.30 (±0.02) 1.53 ± 0.06 3.81 ± 0.11
C¼ ð1Þ
E0g =E0r ðresinÞ 50:50 1.38 (±0.03) 1.45 ± 0.15 3.52 ± 0.52
75:25 1.52 (±0.03) 1.05 ± 0.29 2.51 ± 0.76

where E0g and E0r are the storage modulus values in the glassy (40 °C) Resin data: density = 1.20 (±0.01) g/cm3, water absorption (24 h) = 0.19 ± 0.03%, and
and rubbery (160 °C) regions, respectively. water absorption at saturation (672 h) = 0.64 ± 0.06%.
D. Romanzini et al. / Materials and Design 47 (2013) 9–15 11

Fig. 1. Effect of overall fiber content and hybridization on impact strength of the
Fig. 4. Interlaminar shear strength of the studied composites.
composites.

In the short beam shear test, the maximum shear stress occurs
at the neutral plane where the normal stresses are zero. This re-
sults in combination of failure modes, such as fiber rupture, micro
buckling and interlaminar shear cracking [16]. The obtained inter-
laminar shear strength value was taken into account only for the
samples that exhibited pure interlaminar shear failure, as de-
scribed by Fan et al. [17] and Selmy et al. [18]. According to these
authors, typical load–displacement curves exhibit a nearly linear
elastic trend during the early loading stage until an apparent elas-
tic limit is reached and from them on, load starts decreasing. More-
over, the failed specimens were examined to assess the type of
damage induced in the composite. Fig. 2 shows examples of the
curves that were selected in this study.
Fig. 3 illustrates different types of failure modes identified in
the samples after short beam testing. Fig. 3a shows whitish and
opaque areas indicating localized damage caused by the large
stress concentration just beneath the loading nose, resulting in
crushing of the matrix and fiber damage [18]. Fig. 3b shows inter-
laminar shear fracture, which can be better appreciated using SEM
Fig. 2. Load versus displacement curves of the 50:50 composite with variable fiber micrographs, mostly transverse tensile cracks.
content. Fig. 4 shows interlaminar shear strength of the studied compos-
ites. The constituent materials (matrix and fiber), the adhesion be-
tween fiber and matrix, the fiber volume fraction and fiber
orientation and also void content govern the interlaminar shear
strength of the composites [18,19].
Regarding hybridization, an increase in interlaminar shear
strength was observed for higher glass fiber fraction. Ahmed and
Vijayarangan [16] studied woven jute and jute–glass fabric rein-
forced polyester composites and reported that jute laminates
(36 vol.%) exhibited an ILSS of 13.9 MPa which increased by 9.4%
and 21.1% with the addition of glass fiber (20% and 40% of the total
fiber weight, respectively). According to these authors, the inter-
laminar shear strength depends primarily on the matrix properties
and the fiber–matrix interfacial strength rather than the fiber
properties. Nevertheless, the glass fiber composites exhibited im-
proved ILSS compared to ramie composites, but the difference is
small at low glass fiber fraction (25:75). Maximum ILSS occurred
for the 75:25 composite at 31 vol.% (18 ± 2 MPa), thus, for 31% fiber
incorporation, the reinforcement is still effective in enhancing the
fiber/matrix stress transfer.

3.3. Dynamic mechanical analysis

Fig. 5 shows the storage modulus (E0 ) of all composites. Accord-


Fig. 3. Typical failure mode identified in 75:25 composites after short beam testing ing to this figure, an enhancement in storage modulus with fiber
indicating regions of (a) localized damage and (b) transverse tensile cracks. content was observed over the entire temperature range for all
12 D. Romanzini et al. / Materials and Design 47 (2013) 9–15

Fig. 5. Storage modulus (E0 ) curves for the composites 0:100 (a) and 75:25 (b).

recoverable viscoelastic deformation. Thus, stiffness of the com-


posites increased for higher fiber content.
Table 2
Effectiveness coefficient C (40–160 °C) for the studied composites. Table 2 shows the calculated effectiveness coefficient C
(40–160 °C) for the composites. The lowest C values were observed
Fiber content (vol.%) Effectiveness coefficient C
for composites with 31% of fiber content.
0:100 25:75 50:50 75:25 Idicula et al. [22] studied the reinforcement effectiveness of ba-
10 0.12 0.14 0.11 0.13 nana/sisal (1:1) polyester composites and reported a minimum C
21 0.09 0.09 0.10 0.11 value at 40% fiber volume fraction which was justified base on
31 0.08 0.08 0.07 0.08
the severe fiber-to-fiber contact that occurs for higher fiber load-
ing, decreasing the effective stress-transfer between fibers and ma-
trix. In the present work, the lowest value was found for 31 vol.%
glass–ramie fiber fractions studied. According to Ornaghi et al. and the existence of a minimum could not be verified because
[20], this enhancement is due to the higher restriction imposed the volume fraction could not be raise above 31% due to molding
by the former fibers on the matrix, which would increase the stress constraints.
transfer at the fiber interface. Another factor that could improve The loss modulus (E00 ) curves for the 0:100 and 75:25 compos-
the E0 is the interference of neighboring chains, since greater ites are presented in Fig. 6. The loss modulus increased with fiber
molecular cooperation is required to allow the relaxation process content over the entire temperature range analyzed and for all
to occur. glass–ramie ratios studied. All loss modulus curves reach a maxi-
Margem et al. [21] studying ramie composites verified that mum (for maximum dissipation of mechanical energy) and de-
increasing the fiber content resulted in an enhancement in the creases for higher temperatures, as a result of the free movement
ability of the resin to allow the mechanical constraints with of the polymer chains.

Fig. 6. Loss modulus (E00 ) curves for the composites 0:100 (a) and 75:25 (b).
D. Romanzini et al. / Materials and Design 47 (2013) 9–15 13

Table 3 pure resin (115 °C). Thus, the introduction of fibers showed more
Peak height and glass transition temperature (from E00 curve) obtained for the influence on the tan delta peak than in the measured glass transi-
composites.
tion temperature. According to Pothan et al. [13], composites with
GF:RF Peak height (MPa) Tg (°C) a poor fiber–matrix interface tend to dissipate more energy than
10 vol.% 21 vol.% 31 vol.% 10 vol.% 21 vol.% 31 vol.% the composites with good interface bonding, i.e. greater damping
0:100 257 299 353 98 102 103
indicates weaker interfacial adhesion. In Fig. 7, the increase in fiber
25:75 290 325 377 98 100 102 content showed a decrease in the tan delta peak because the over-
50:50 291 340 454 100 99 102 all interface area within the composite increases.
75:25 377 434 553 101 101 101 From the tan delta peak height values, the interfacial adhesion
Data for the pure resin: peak height of 243 MPa and Tg of 94 °C. was determined using the methodology described by Correa
et al. [14]. According to these authors, for strong interface adhe-
sion, molecular mobility around the reinforcement is reduced,
The increase in loss modulus peak height (Table 3) with glass fi- and low A values suggest improved interactions at the matrix-fiber
ber content is likely to have occurred due to the inhibition of the interface.
relaxation process within the composite due to the enhancement Table 4 shows the calculated A values in the present work. Low-
in the number of chain segments and in the free volume upon fiber er A values (i.e. stronger adhesion) were found for lower overall fi-
addition [23]. Additionally, the peak height exhibited the same ber content and for the composites with higher proportion of ramie
trend found for the impact strength results. fiber.
According to Almeida et al. [24], systems containing more Analyzing the SEM micrographs (Fig. 8a and b), with higher fi-
restrictions and a higher degree of reinforcement tend to exhibit ber content, an increase in fiber–fiber contact was observed, that
higher Tg. Comparing the resin to the composites (Table 3), the Tg could decrease the interaction between fiber and matrix. For the
increased from 94 °C to 98–103 °C, but no significant change in glass–polyester interface and the ramie–polyester interface
Tg could be identified among the composites. (Fig. 8c), at higher glass fiber fraction, agglomeration occurs, there-
The tan delta curves for all composites are shown in Fig. 7. With by decreasing the effective stress transfer between the glass fibers
the incorporation of fibers, the tan delta peak is lowered due to and the matrix [25]. Thus, strong interaction occurs at higher ramie
greater restriction in the movement of the polymer molecules fiber fraction, which corresponds to the observed adhesion factor A
caused by the presence of stiff fibers [13,20]. However, when the values.
fiber content is low, there will be regions with high concentration It is known that the hybrid composites are a more complex sys-
of resin (matrix) in the composite which will not be affected by the tem than the composites analyzed by Correa et al. [14], in which
presence of fibers. the adhesion factor was calculated for different coupling agents
The glass transition temperature obtained from the tan delta in polypropylene–wood composites. Analyzing the different fiber
curve did not exhibit a clear trend. All Tg values for the composites content of the present work, the results for the adhesion factor A
were within 112–116 °C in fact, in the same range of the Tg for the did not correspond to those found for the tan delta peak height.

Fig. 7. Effect of fiber content on tan delta curve for the composites 25:75 (a) 50:50 (b) and 75:25 (c).
14 D. Romanzini et al. / Materials and Design 47 (2013) 9–15

Table 4
Adhesion factor A (at Tg) for the composites.

Adhesion factor A
Fiber content (vol.%) 0:100 25:75 50:50 75:25
10 0.316 0.295 0.310 0.246
21 0.306 0.268 0.245 0.170
31 0.250 0.214 0.200 0.184

Fig. 9. Adhesion factor A for the composites at 31 vol.%.

Fig. 10. Cole–Cole plot for the resin and for the 50:50 composite for different fiber
content values.

In crosslinked polymers, the structural changes that occur after


the incorporation of fibers to a polymeric matrix can be studied
using the Cole–Cole method and the nature of the Cole–Cole plot
is indicative of the nature of the system and a homogeneous poly-
meric system is reported to exhibit a semi-circle diagram [13].
Some authors [13,22,26] obtained imperfect semi-circles for poly-
mer composites, which are indicative of system heterogeneity
associated with greater differences in the relaxation process as
more fiber are incorporated. The Cole–Cole plot of the pure resin
and the 50:50 composite are plotted in Fig. 10. Ratifying the results
of the cited references, the Cole–Cole diagrams showed imperfect
semi-circles, characteristic of heterogeneous systems.

4. Conclusions

Fig. 8. SEM micrographs of the composites 25:75/10 (a) 25:75/31 (b and c).
Hybridization enables the production of lower weight compos-
ites with intermediate water absorption values. The mechanical
An interesting behavior was observed regarding the variation of properties (impact and interlaminar shear strength) of the com-
the A factor with the temperature (Fig. 9). The ‘‘A’’ values of all posites improved for higher fiber content, and the reinforcement
composites decreased with the temperature following a similar effectiveness at 31 vol.% was confirmed by lower C values deter-
pattern. At first, the composites with higher glass fiber fraction mined by the storage modulus analysis. In addition, the loss mod-
exhibited lower ‘‘A’’ values. However, at a certain temperature (in ulus peak height exhibited the same trend found for the impact
the 80–90 °C range), this behavior was reversed, and the ramie fi- strength results. Although the mechanical properties improved
bers yielded higher A values. for higher fiber content, Tg did not change significantly.
D. Romanzini et al. / Materials and Design 47 (2013) 9–15 15

Analyzing the calculated adhesion factor A, the trend of im- [10] Pistor V, Ornaghi FG, Ornaghi Jr HL, Zattera AJ. Dynamic mechanical
characterization of epoxy/epoxycyclohexyl-POSS nanocompósitos. Mater Sci
proved adhesion occurs for the composites with a lower fiber con-
Eng A-Struct 2012;532:339–45.
tent (10 vol.%) and a higher natural fibers fraction (0:100). [11] Qazvini NT, Mohammadi N. Dynamic mechanical analysis of segmental
Analyzing the SEM micrographs, at higher fiber content, an in- relaxation in unsaturated polyester resin networks: effect of styrene content.
crease in fiber–fiber contact was observed, which could decrease Polymer 2005;46:9088–96.
[12] Bureau E, Chebli K, Cabot C, Saiter JM, Dreux F, Marais S, et al. Fragility of
the interaction between fibers and the matrix. At higher glass fiber unsaturated polyester resins cured with styrene: influence of the styrene
fractions, agglomeration occurs, resulting in an incompatibility concentration. Eur Polym J 2001;37:2169–76.
between the glass fiber and the matrix. Thus, a strong interaction [13] Pothan LA, Oommen Z, Thomas S. Dynamic mechanical analysis of banana
fiber reinforced polyester composites. Compos Sci Technol 2003;63:283–93.
between the fibers and the matrix occurs at higher ramie fiber frac- [14] Correa CA, Razzino CA, Hage Jr E. Role of maleated coupling agents on the
tions, which corresponds with the adhesion factor A values. There- interface adhesion of polypropylene–wood composites. J Thermoplast Compos
fore, analyzing composites with different fiber content values, the 2007;20:323–39.
[15] Jawaid M, Abdul Khalil HPS, Abu Bakar A, Khanam NP. Chemical resistence,
results obtained for the adhesion factor A did not correspond to void content and tensile properties of oil palm/jute fibre reinforced polymer
those obtained for the tan delta peak height. Finally, the Cole–Cole hybrid composites. Mater Des 2011;32:1014–9.
diagram showed imperfect semi-circles, which are characteristic of [16] Ahmed KS, Vijayarangan S. Tensile, flexural and interlaminar shear properties
of woven jute and jute–glass fabric reinforced polyester composites. J Mater
heterogeneous systems. Process Tech 2008;207:330–5.
[17] Fan Z, Santare MH, Advani SG. Interlaminar shear strength of glass fiber
Acknowledgements reinforced epoxy composites enhanced with multi-walled carbon nanotubes.
Compos Part A – Appl Sci 2008;39:540–54.
[18] Selmy AL, Elsesi AR, Azab NA, El-baky A. Interlaminar shear behavior of
The authors thank CNPq and CAPES for their financial support unidirectional glass fiber (U)/random glass fiber (R)/epoxy hybrid and non-
and Elekeiroz S.A for providing the polyester resin. hybrid composite laminates. Compos Part B 2012;43:1714–9.
[19] John NA, Brown JR. Flexural and interlaminar shear properties of glass–
reinforced phenolic composites. Compos Part A – Appl Sci 1998;29A:939–946,.
References [20] Ornaghi Jr HL, Bolner AS, Fiorio R, Zattera AJ, Amico SC. Mechanical and
dynamic mechanical analysis of hybrid composites molded by resin transfer
[1] Jawaid M, Abdul Khalil HPS. Cellulosic/synthetic fiber reinforced polymer molding. J Appl Polym Sci 2010;118:887–96.
hybrid composites: a review. Carbohyd Polym 2011;86(1):1–18. [21] Margem FM, Monteiro SN, Neto JB, Rodriguez RJS, Soares BG. The dynamic-
[2] Velmurugan R, Manikandan V. Mechanical properties of palmyra/glass fiber mechanical behavior of epoxy matrix composites reinforced with ramie fibers.
hybrid composites. Compos Part A – Appl Sci 2007;38:2216–26. Rev Mat 2010;15:167–75.
[3] Mishra S, Mohanty KA, Drzal LT, Misra M, Parija S, Nayak SK, et al. Studies on [22] Idicula M, Malhotra SK, Joseph K, Thomas S. Dynamic mechanical analysis of
mechanical performance of biofibre/glass reinforced polyester hybrid randomly oriented intimately mixed short banana/sisal hybrid fiber reinforced
composites. Compos Sci Technol 2003;63:1377–85. polyester composites. Compos Sci Technol 2005;65:1077–87.
[4] Cicala G, Recca G, Ziegmann G, Ziegmann G, El-Sabbagh A, Dickert M. [23] Hameed N, Sreekumar PA, Francis B, Yang W, Thomas S. Morphology, dynamic
Properties and performances of various hybrid glass/natural fibre composites mechanical and thermal studies on poly (styrene-co-acrylonitrile) modified
for curved pipes. Mater Des 2009;30:2538–43. epoxy resin/glass fibre composites. Compos Part A – Appl Sci
[5] Romanzini D, Ornaghi Jr HL, Amico SC, Zattera AJ. Preparation and 2007;38:2422–32.
characterization of ramie–glass fiber reinforced polymer matrix hybrid [24] Almeida Jr HS, Ornaghi Jr HL, Amico SC, Amado FDR. Study of hybrid
composites. Mater Res 2012;15(3):415–20. intralaminate curaua/glass composites. Mater Des 2012;42:111–7.
[6] Athijayamani A, Thiruchitrambalam M, Natarajan U, Pazhanivel B. Effect of [25] Devi LU, Bhagawan SS, Thomas S. Dynamic mechanical analysis of pineapple
moisture absorption on the mechanical properties of randomly oriented leaf/glass hybrid fiber reinforced polyester composites. Polym Compos
natural fibers/polyester hybrid composite. Mater Sci Eng A 2009;517:344–53. 2010;31:956–65.
[7] Menard KP. Dynamic mechanical analysis: a practical introduction. CRC Press [26] Ornaghi Jr HL, Silva HSP, Zattera AJ, Amico SC. Hybridization effect on the
LLC; 1999. p. 13–29. mechanical and dynamic mechanical properties of curaua. Compos Mater Sci
[8] Cassu SN, Felisberti MI. Comportamento dinâmico-mecânico e relaxações em Eng A 2011;528:7285–9.
polímeros e blendas poliméricas. Quim Nova 2005;28(2):255–63.
[9] Ornaghi Jr HL, Pistor V, Zattera A. Effect of the epoxycyclohexyl polyhedral
oligomeric silsesquioxane content on the dynamic fragility of an epoxy resin. J
Non-Cryst Solids 2012;358:427–32.

You might also like