Вы находитесь на странице: 1из 12

Polymer Testing 67 (2018) 533–544

Contents lists available at ScienceDirect

Polymer Testing
journal homepage: www.elsevier.com/locate/polytest

Material Properties

Effects of sodium carbonate on the performance of epoxy and polyester coir- T


reinforced composites
Júlio Cesar dos Santosa,b, Renato Luiz Siqueirac, Luciano Machado Gomes Vieirad,
Rodrigo Teixeira Santos Freirea,b, Valdir Manob, Túlio Hallak Panzeraa,∗
a
Centre for Innovation and Technology in Composite Materials, Department of Mechanical Engineering, Federal University of São João Del Rei, Brazil
b
Department of Natural Sciences, Federal University of São João Del Rei, Brazil
c
Department of Materials Engineering, Center of Technological Education of Minas Gerais, Brazil
d
Department of Production Engineering, Federal University of Minas Gerais, Brazil

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

Keywords: Surface modification induced by sodium carbonate on coir fibres was studied in view of its application as a
Composites reinforcing agent in polyester and epoxy polymer fibre-reinforced composites. Coir fibres were treated with a
Coir fibre 10 wt% sodium carbonate solution for different periods of time (24, 96 and 168 h) at room temperature. The
Chemical treatment surface treatment was evaluated by scanning electron microscopy, X-ray diffraction and thermogravimetric
Mechanical properties
analysis. Tensile, flexural and impact properties of treated and untreated coir fibre-reinforced composites were
TGA
compared. Coir-reinforced epoxy and polyester composites were manufactured by the uniform dispersion of
XRD
SEM randomly oriented coir fibres within the polymeric matrix. Tensile and flexural modulus of coir-reinforced
polyester composites increased nearly 28% and 25%, respectively after 96 h of coir treatment. An analogous
increase of 31% and 17% was obtained for coir-reinforced epoxy composites. In addition, coir-polyester com-
posites achieved superior tensile (∼17%) and flexural (∼5%) modulus and impact strength (∼193%) compared
with coir-epoxy-derived composites. In contrast, coir-epoxy composites led to superior tensile and flexural
strength. The experimental results revealed that sodium carbonate solution effectively removes hemicellulosic
compounds, promotes swelling and increases superficial roughness of the fibres, improving the mechanical
properties (modulus and strength) of the coir-reinforced composites studied.

1. Introduction fact, thousands of different natural fibres in the world and only few of
these have been studied, which opens a vast field of research for fibre-
Natural fibres have been appraised as environmentally correct ma- reinforced composites [1]. Natural fibres, however, present low re-
terials due to their renewability, biodegradability and low processing sistance to microbial attack and, owing to their hydrophilicity, are
energy demands [1,2]. Lignocellulosic fibres, in particular, are neutral susceptible to moisture absorption and incompatible with hydrophobic
with respect to the emission of CO2, which is a very important char- macromolecular matrices, which impairs fibre-matrix adhesion. In spite
acteristic in view of the Kyoto protocol [3]. of these drawbacks, these materials are currently used in many appli-
Natural fibres consist of thick-walled cells, referred to as elementary cations due to their resistance to corrosion and abrasion, low density,
fibres, which make the structure stiff and strong. The elementary fibre low cost and high specific mechanical properties. The selection of the
is composed of multilayered cellulose-lignin/hemicellulose walls of composite matrix phase is, however, limited by the temperature at
variable thickness, built on an array of cellulose-based crystalline which natural fibres degrade. Most natural fibres are thermally unstable
structures called microfibrils, arranged at a particular angle (the mi- above 200 °C, and thus thermoset polymers that can be cured below this
crofibrillar angle – MFA). Cell walls are composed of a primary and a temperature are used, especially unsaturated polyester and epoxy types
secondary wall layers, whose chemical composition (i.e., the content of [6]. Epoxy polymers are currently the dominant choice for low and
lignin, cellulose and hemicellulose) varies from fibre to fibre. The moderate temperatures (up to 135 °C) and the most common material
physical and mechanical properties of natural fibres are determined by used in high-performance engineering composites. Despite their lower
their chemical composition, cell wall/lumen ratio and cellulose mi- cost and extensive use in commercial applications, polyesters, which
crofibrillar angle in the dominant cell wall layers [2,4,5]. There are, in can be used at approximately the same temperatures as epoxy


Corresponding author.
E-mail address: panzera@ufsj.edu.br (T.H. Panzera).

https://doi.org/10.1016/j.polymertesting.2018.03.043
Received 30 January 2018; Received in revised form 22 March 2018; Accepted 24 March 2018
Available online 27 March 2018
0142-9418/ © 2018 Elsevier Ltd. All rights reserved.
J.C. dos Santos et al. Polymer Testing 67 (2018) 533–544

Fig. 1. Treatment process: a) portions of coir fibres b) Na2CO3 bath treatment c) post treatment washing.

polymers, are rarely chosen for high-performance composites [7]. lignocellulosic materials [22]. The effectiveness of this salt is compar-
Coir, in particular, is a lightweight, strong type of fibre that has able to that of regular alkaline reagents, e.g., NaOH. Sodium carbonate
lately attracted scientific and commercial interest owing to its specific can also potentially overcome the major operational problems with
properties and availability [5,8–11]. Compared with other typical conventional alkaline pretreatments such as corrosion and environ-
natural fibres, coir has higher lignin and lower cellulose and hemi- mental hazards [22], as well as serve as an effective pretreatment of
cellulose content which, along with its high microfibrillar angle, pro- natural fibres at reduced exposure times compared to sodium bicarbo-
vides several valuable properties, such as resiliency, toughness, as well nate. Such use, to the best of these authors' knowledge, has received
as damping, abrasion, weather resistance and high enlongation at break much less attention relative to other methods. This work, therefore,
[11]. Hydrophilicity, however, limits the use of coir and other natural aims to investigate the effects of the alternative, eco-friendly sodium
fibres as reinforcement in hydrophobic polymers [12–14]. Chemical carbonate pretreatment on the properties of epoxy and polyester coir-
modifications of the fibres have been proposed to improve the com- reinforced composites. Morphological and chemical characteristics of
posite mechanical performance and durability [6,12,13,15]. Several treated and untreated coir were investigated through scanning electron
studies indicate that, in order to improve physical and mechanical microscopy (SEM), X-Ray Diffraction (XRD), thermogravimetric ana-
properties, the surface of coir fibres can be modified by alkaline lysis (TGA) and density. In order to investigate the effect of the pro-
treatment to enhance fibre-matrix compatibility. Karthikayen & Bala- posed treatment on the mechanical properties of coir-epoxy/polyester
murugan [16] reported that the alkaline treatment of coir-reinforced composites, quasi-static tensile and flexural tests as well as flatwise
epoxy composites enhanced the impact resistance (∼78%), as well as charpy impact tests were carried out.
the tensile (∼18%) and flexural (∼17%) strength relative to untreated
composites. Similar behaviour is observed with other fibres. For in-
2. Material and methods
stance, Kumar et al. [14] have used NaOH to treat coconut-sheath fibres
in reinforced epoxy composites, achieving higher tensile (∼21%),
2.1. Alkali treatment of coir fibres
flexural (∼19%) and impact (∼24%) strength relative to non-treated
composites. Rout et al. [17] also reported the improvement of the fibre-
Coir fibres in entangled mats were supplied by Deflor Bioengenharia
matrix interfacial adhesion in coir-polyester reinforced composites after
(Belo Horizonte – Brazil). Portions of manually removed coir fibres
alkali treatment of the fibres. The higher tensile (∼26%) strength was
(Fig. 1a) were soaked in 10 wt% Na2CO3 solution (Fig. 1b), with mea-
obtained after treatment with a 2 wt% NaOH solution, while flexural
sured pH nearly 12 for 24, 96 or 168 h at room temperature. The coir
strength and impact resistance were higher (∼17 and ∼41%, respec-
samples were washed in water and subsequently immersed in fresh
tively) for treatments with a 5% alkali NaOH solution.
water for 30 min (post-treatment - Fig. 1c) and oven-dried at 50 C for
Jayabal et al. [18] investigated woven coir fibre-reinforced polye-
24 h. The average length of the fibres was 16 ± 5 cm.
ster composites and reported a 40% (42%) increase in tensile (flexural)
strength and 20% increase in impact resistance in comparison to un-
treated fibre composites. Enhanced mechanical properties were ob- 2.2. Fabrication of the composites
tained for the following NaOH concentration/soaking time combina-
tions: (2 wt%, 96 h), (5 wt%, 72 h) and (8 wt%, 24 h). However, the use The thermosetting polymers, epoxy and polyester, were supplied by
of chemical reagents such as sodium hydroxide have become less at- Hunstman-Brazil. The epoxy polymer consisted of Renlam M resin and
tractive owing to their cost as well as operational and environmental Aradur HY 956 hardener at a proportion of 5:1, respectively. An un-
hazards, which affect the environmental acceptability and commercial satured orthophtalic polyester resin was mixed with 2 wt% of methyl-
viability of biocomposites manufacturing [19,20]. ethyl-ketone as a catalyst. Composite samples were manufactured by
New eco-friendly and cost-effective treatments based on sodium hand lay-up followed by cold uniaxial compaction. Randomly oriented
bicarbonate have been proposed for kenaf, sisal and flax fibre-re- coir fibres were placed on an aluminium plate (0.5 mm thickness) in a
inforced epoxy composites [1,20,21]. Such treatments enhance fibre square steel mould frame, of dimensions 310 mm × 310 mm × 3 mm
stiffness as well as fibre-matrix adhesion by removing hemicellulose (Fig. 2a). Based on preliminary experiments [23], the fibre grammage
and superficial impurities of the fibres. However, the mildly alkaline was set at 900 g/m2. From the mean apparent density of coir fibres
character of sodium bicarbonate demanded longer treatment periods obtained in this work, 0.86 ± 0.02 g/cm3, the fibre volume fraction of
(about 120 h) compared to highly alkaline NaOH solution [19]. approximately 30% was determined.
Sodium bicarbonate, NaHCO3, also known as sodium acid carbo- The polymer mixture was then poured into the mould and, once the
nate, is frequently used as an antacid for medical purposes, as a lea- fibres were completely wet (Fig. 2b), a similar aluminium plate was
vening agent to bake cakes and bread and in the manufacture of sodium placed on the surface and the mould was closed with a steel lid
carbonate (Na2CO3). Sodium carbonate, Na2CO3, a comparatively (Fig. 2c). Samples were cured for 60 h at room temperature under a
stronger alkali, is used in laundry detergents as a softening agent. constant uniaxial pressure at 645 kPa. This pressure level was set based
Sodium carbonate acts as a relatively strong alkali in aqueous so- on several preliminary studies. The Tec Glaze-N mould release agent
lution and has been previously studied for the pretreatment of was used on the surface of aluminium plates to ensure better surface
finish of the composite samples. Subsequently, the aluminium plates

534
J.C. dos Santos et al. Polymer Testing 67 (2018) 533–544

2.4.4. Thermogravimetric analysis (TG/DTG)


The thermal stability of the raw and treated fibres was evaluated by
thermogravimetry using a Shimadzu DTG-60H differential thermal
analyzer. The analyses were performed under nitrogen atmosphere at a
constant flow (∼100 ml/min), with a heating rate of 10 °C/min and
temperature ranging from 30 to 800 °C.

2.5. Mechanical tests

The uniaxial tensile and flexural tests were conducted on a


Shimadzu AGX-Plus Universal Testing Machine equipped with a 100 kN
load cell according to ASTM D3039 [30] and ASTM D790 [31], re-
Fig. 2. Hand Lay-up manufacturing process. spectively, at a crosshead speed of 2 mm/min. Samples of dimensions
250 mm × 25 mm were used for the tensile test and a video-ex-
tensometer was used for the elongation measurements. For each ex-
were removed, and the composite panel with nearly 3 mm in thickness
perimental condition, six (6) specimens were fabricated with two re-
was post-cured at room temperature for 2 weeks (according to the
plicates running a total of 96 specimens. A second composite plate was
manufacturer's recommendation).
manufactured to obtain a new set of samples (Replicate 2).
The flexural test was performed on samples whose length and width
2.3. Design of experiment were determined in accordance to ASTM D790, so that the length of the
support span was 16 times the specimen thickness. For each experi-
The Design of Experiment (DoE) and Analysis of Variance (ANOVA) mental condition, ten (10) specimens were fabricated with two re-
techniques involve arrays to organize the factors affecting the me- plicates to total of 160 specimens. The impact test was carried out using
chanical and physical properties studied and the levels at which they a Charpy Impact Tester (XJJ Series) equipped with an 8.04 Nm hammer
should be set. A full factorial design (21∙31 = 6) with six experimental of 15 J of energy with samples of 80 mm × 10 mm of dimensions, ac-
conditions (E.C.) was established to investigate the effect of the factors cording to ISO 179-1 [32]. For each condition, ten (10) specimens were
(levels) Treatment Time, TT (24, 96, 168 h) and Matrix Phase, MP fabricated with two replicates to a total of 160 specimens. Tests were
(epoxy and polyester), on the flexural and tensile modulus and strength, conducted at controlled temperature (23 °C) and humidity (55%).
charpy impact resistance and apparent density of the composites. The
treated (untreated) composites are identified with the sample codes E- 3. Results
24, E-96 and E-168 (E-raw) for epoxy composites and P-24, P-96 and P-
168 (P-raw) for polyester composites. A randomization procedure was 3.1. Morphological analysis
also adopted during manufacturing and experimental tests, making the
whole numerical process more robust against non-controllable vari- In addition to the removal of lignin from the fibre surface, sodium
ables [25,26]. carbonate also removes impurities, wax and fatty substances adsorbed
in the surface [19]. The mechanism of action of alkalis is the disruption
2.4. Morphological and chemical characterization of fibres of hydrogen bonds on the fibre surface, thereby increasing its roughness
[5,16,28]. The native surface of the fibre (Fig. 3a) is smooth due to a
2.4.1. Scanning electron microscopy (SEM) waxy layer [33]. A rougher fibre surface is achieved after 24 h of
Raw and treated fibres were analysed by scanning electron micro- treatment and roughness progressively increases up to 96 h, owing to
scopy using a Philips XL-30 FEG microscope and Hitachi TM 3000 the removal of waxes, cuticles and globular particles, creating pits
microscope model coupled to an energy dispersive X-ray spectrometer (tyloses) on the fibre surface (Fig. 3b and c), as reported by Nam et al.
(EDS). [34]. It is noteworthy, however, that longer periods of treatment pro-
duces a smoother surface owing to the complete removal of cuticles and
2.4.2. Fibre apparent density globular particles, as shown in Fig. 3d, for 168 h. Longer periods of
The apparent density of coir fibres was measured by the density- treatment also leads to degradation of the fibres [1], also observable in
gradient column method for 5 samples composed of a bunch of fibres Fig. 3d. This effect may account for the decrement of the tensile and
immersed in pure ethanol (ASTM D276 [24]). flexural strength obtained from the composite mechanical character-
ization, as commented in section 3.3.

2.4.3. X-ray diffraction - crystallinity index (Ic) 3.2. Physical, chemical and thermal characterization
X-ray diffraction was used to evaluate the removal of superficial
compouds and exposure of cellulose molecules. The fibre bundles were The physical, chemical and thermal properties of the fibres studied
manually aligned and scissor-lacerated to micro-segments to be ana- are summarized in Table 1. It is possible to observe that raw coir fibres
lysed by X-ray diffraction using a Rigaku Ultima IV X-ray diffractometer have lower density and relatively higher thermal stability.
operating with CuKα radiation (λ = 0.15418 nm). The diffraction pat-
terns were obtained for values of 2θ ranging from 5 to 40° in continuous 3.2.1. Apparent density and structure
scan mode at 1°/min. The crystallinity index (Ic) of the samples was The apparent density of coir fibres varied from 0.86 to 0.97 g/cm3,
calculated using the formula: higher values being associated with the sodium carbonate treatment. As
I − Iam ⎞ shown in Fig. 4a, the morphological analysis of the cross section of a
Ic = ⎛ 002
⎜ × 100,

single untreated coir fibre reveals numerous elementary fibres, each
⎝ I002 ⎠ (1)
with its own lumen and pores (Fig. 4a). The apparent density was de-
where I002 is the maximum intensity of diffraction on the (002) crys- termined considering, as mentioned above, the whole volume occupied
tallographic plane at a 2θ angle between 22° and 23°, and Iam is the by the fibre, including the enclosed air volume. The fibre density is
intensity of diffraction of the amorphous material taken at a 2θ angle affected by fibre length and porosity [2], since pores reduce the density.
between 18° and 19° where the intensity is at a minimum [23–25]. Sodium carbonate increased the apparent density (see Table 1), owing

535
J.C. dos Santos et al. Polymer Testing 67 (2018) 533–544

Fig. 3. SEM surface analysis of coir fibres (a) raw (b) 24 h treated (c) 96 h treated (d) 168 h treated.

to cellulose swelling, which promotes the densification of cell walls and Based on the Na2CO3 dissolution, Na+ ions released into the aqu-
reduces the lumen of the elementary fibres (compare Fig. 4b and c). The eous medium in contact with cellulose form I disrupt hydroxyl bonds
low packed crystalline regions contributes to the diffusion of the Na+ and, having a favorable diameter to penetrate the interlattice space
ions through the structure [35]. Due to low cellulose content (micro- between adjacent planes, diffuse into these planes and form a new Na-
fibrils) (∼45% of fibre structure [4]) the Na+ ions can easily diffuse cellulose I [41] lattice with relatively large distances between the cel-
through the structure of raw coir fibres, since the amorphous regions lulose molecules which play a crucial role in widening even the smallest
are more easily accessed by solvent molecules. The literature reports pores. Considering that most of the hydroxyl groups (OH) present in the
that, after washing with water, the Na+ ions tend to be excluded from fibres correspond to alcoholic hydroxyls, an interaction analogous to
the crystalline lattice [34,36,37]. However, EDS analysis reveals that that which occurs during a traditional mercerization treatment may be
the post-treatment applied was not enough to remove all sodium ions proposed [1,27,34,42].
from inter- and intramolecular cellulose fibrils (Fig. 4c). Higher con-
Coir-OH + Na+ + OH− → Coir-O-Na+ + H2O (2)
centrations of alkali, however, may cause depolymerization of cellulose
molecules, damaging the cell wall and reducing the density [38]. Such The OH groups of cellulose are converted into ONa groups [18] by
effect was not observed, which implies that the alkaline concentration expanding the cellulose molecule dimensions, thereby increasing the
used did not cause cell wall damage. pore size. Cell wall swelling enables large molecules to penetrate
Cellulose swelling is a complex phenomenon which affects both crystalline regions and exert very large forces on the weaker Na-cellu-
molecular and supramolecular structures [1,36,39,40]. Cellulose mo- lose I lattice; thus, the lumen collapses resulting in a more compact
lecules, which consist of a series of glucopyranose rings linked together structure, leading to a closer packing of the chains of cellulose fibrils.
by valence bonds, are arranged parallel to the fibre axis in structures The swollen cell wall does not revert to the original shape after drying;
called cellulosic chains or microfibrils linked by hydroxyls bonds. In cell walls remain swollen and the lumen is permanently collapsed.
aqueous media, sodium carbonate dissociates into Na+ and CO32− ions.
Na+ ions do not participate in the hydrolysis reaction but, on the other
hand, the CO32− ions interact with H+ of the water releasing OH−.

Table 1
Physical, chemical and thermal characterization of coir fibre.
Fibre treatment Apparent Density Crystallinity Index Weight loss Cellulose Thermal Stability

3
g/cm % Range (°C) % (°C)

None 0.86 ± 0.02 44.79 30–124 7.23 –


223–395 48.98 357
395–622 42.58 –
Na2CO3 24 h 0.91 ± 0.04 47.73 30–137 9.61 –
210–403 41.39 314
403–597 45.06 –
96 h 0.96 ± 0.02 47.84 30–140 13.59 –
208–421 38.83 312
421–600 44.51 –
168 h 0.97 ± 0.03 47.85 30–139 9.38 –
215–410 42.43 318
410–644 45.12 –

536
J.C. dos Santos et al. Polymer Testing 67 (2018) 533–544

Fig. 4. SEM cross-sectional area analysis of coir fibres: a and c) raw b) 96 h treatment and d) 96 h treatment (EDS analysis). The EDS spectra of raw and treated fibres
are also presented.

3.2.2. X-ray diffraction - crystallinity index (Ic) of raw and treated coir have considerable effects on the regeneration of cellulose I to cellulose
fibres II at room temperature, while lower concentrations result only in a
The X-ray diffractograms of the untreated and Na2CO3-treated fibres slight increase of crystallinity and a negligible amount of cellulose II.
for different times are shown in Fig. 5. They reveal that both untreated This finding helps to explain the present results, since the pH resulting
and treated coir exhibit the same diffraction pattern of cellulose I, in from a 10 wt% Na2CO3 solution is approximately 12, whereas that of a
which the main intensity peaks are located at 2θ between 22 and 23° NaOH solution at the same concentration is much higher. Therefore, it
(between 18 and 22° for cellulose II, a polymorphic form) [27,29,36]. is reasonable to state that the treatment proposed here constitutes a
This reveals that the treatment of coir fibre with 10 wt% Na2CO3 does milder and unfavorable condition for the complete formation of cellu-
not completely transform cellulose I into cellulose II, but it is sufficient lose II.
to reduce the amorphous phase (including cementing components: When fibres contain large amounts of amorphous material (i.e
hemicellulose, lignin [43]), consequently, leading to a slight increase of lignin, hemicellulose, pectins and amorphous cellulose), the intensity of
the crystallinity index of treated samples (see Table 1). The increase of diffraction of the amorphous material seem to be smeared out and
crystallinity may also be attributed to the enhanced packing of cellulose appear as a single broad peak. However, when the crystalline content is
chains (microfibrils) after swelling [44]. high these peaks are more pronounced [40]. Fig. 5 reveals that, for
Liu and Hu [45] report that NaOH concentrations above 11 wt% treated fibres, an inflection point begins to appear, and the broad peak

537
J.C. dos Santos et al. Polymer Testing 67 (2018) 533–544

(96 h) and 29.73% (168 h). These results indicate enhanced interaction
between water molecules and coir fibre components (specially hydroxyl
groups of cellulose chains) after the removal of the superficial cuticle
layer and other chemical modifications produced by the Na2CO3
treatment [47]. It is noteworthy that moisture absorption depends on
the extent of the alkali treatment, rising to 88% for 96 h with sub-
sequent decrease to 30%. Similar behaviour was reported by Silva. et al.
[47], who also observed an increase in the moisture absorption of 5 wt
% NaOH-treated coir fibres in the range of 50%, 100% and 50% for 24,
48 and 72 h, respectively. To the best of these authors' knowledge, no
explanation for this behaviour has been found in the literature. How-
ever, SEM images (Fig. 3) imply that tyloses are more exposed at 96 h
than at 24 h, indicating the progressive cuticle removal by sodium
carbonate, which may explain the initial increment in moisture ab-
sorption. On the other hand, Medronho et al. [35] note that cellulose,
consisting of glucopyranose monomers, bears an amphiphilic character,
being hydrophobic in the equatorial plane of the glucopyranose rings
and hydrophobic in the axial plane. Gassan and Bledzki [44] report
that, on alkali treatment and the subsequent removal of lignin and
hemicellulose, cellulose molecules are more susceptible to spatial re-
Fig. 5. X-ray diffractograms of raw and Na2CO3 treated coir fibres.
orientation. This mechanism could account for the progressively more
hydrophobic character of the fibres through the incremented exposure
tends to resolve into two individual peaks. The peak at 2θ = 16° is al- of the axial plane of the glucopyranoses.
ready very pronounced and this phenomenon implies an increase in For the untreated fibres, the second stage of mass loss, which ap-
crystalline cellulose, whereas the amorphous phase is reduced relative pears as a shoulder centred at 307 °C in DTG curves, can be associated
to untreated fibres The increase of crystallinity induced by Na2CO3 mainly with hemicellulose degradation [1]. This stage of mass loss was
solution may be thus attributed to two effects: 1) amorphous material not observed for all treated fibres, demonstrating that the treatments of
removal and exposure of surface crystallites [5], and 2) structural re- 24, 96 and 168 h with 10 wt% Na2CO3 solution were efficient in re-
arrangement of crystalline regions [27,36,44–46]. moving a considerable amount of their hemicellulose. These results
During swelling, interfibrillar regions become less dense and rigid, corroborate the X-ray analysis presented in Fig. 5, which revealed a
thus more prone to structural rearrangements [44]. After washing with reduction in amorphous compounds (including hemicellulose) from
water, some authors [36,41] report that Na-cellulose I undergoes coir fibres promoted by Na2CO3.
structural changes with complete removal of Na+ ions. Such removal The most intense mass loss occurred in the range of 320–400 °C for
did not completely occur here, as shown in Fig. 4d. Na-cellulose I untreated fibres, in which the maximum mass flow occurred from the
contains in fact two coexisting phases - Na-cellulose II, which re- solid to the vapor phase at 358 °C, corresponding to the thermal de-
generates to cellulose I, and Na-cellulose III, also called cellulose IV, gradation of cellulose [14,15]. This temperature (358 °C) of maximum
which regenerates to cellulose II [36,37,45]. degradation rate decreased to approximately 310 °C for treated fibres,
observed in DTG curves, showing that the alkaline treatment also de-
3.2.3. Thermogravimetric analysis (TGA) creased the thermal stability of the cellulose. Previously studies have
The TG and DTG curves of raw and treated coir fibres are shown in reported a similar effect for sisal [1] and coir [47] fibres treated with
Fig. 6 and Table 1 (weight loss and thermal stability columns). NaHCO3 and NaOH solutions at room temperature, respectively.
Previous studies report that the thermal decomposition profile of Some authors have, however, reported an improvement of the
coir fibres is characterized by three peaks [14,15,47]. The first peak is thermal stability after alkaline treatment of fibres (sisal, munguba and
associated with moisture loss, which occurs up to approximately 100 °C. coir) owing to cellulose II [42,48,49] – which was not observed in this
Sodium carbonate induces a significant increase in moisture absorption study with the proposed alkaline treatment (see Fig. 6). It is important
relative to untreated coir fibres in the range of 32.92% (24 h), 87.97% to note that, although the transformation of cellulose I to II has not

Fig. 6. Thermal analysis curves (a) TG and (b) DTG for coir fibres.

538
J.C. dos Santos et al. Polymer Testing 67 (2018) 533–544

occurred, the alkaline treatment was sufficient to modify the structure matrix (Fig. 8b). Such behaviour was also observed in the flexural test
of the cellulose I, converting some OH-groups into ONa-groups. and will be discussed in the following section.
Cellulose II presents more intermolecular hydrogen bonds than It is well established by the theory of composites that the tensile
cellulose I which increases the thermal stability of fibres [40,50–52]. It modulus of fibre-reinforced composites depends on the tensile modulus
is worth noting that Liu and Hu [45] have observed that the poly- of the fibres and matrix, fibre-matrix volume fraction and fibre or-
morphic transition from cellulose I to cellulose II is thermodynamically ientation [34]. Coir fibres are linear elastic at low stress, and then show
favorable depending on the interaction of two factors: NaOH con- plastic behaviour until fibre failure at a very high strain [2]. As pre-
centration and treatment temperature [45]. These authors have in fact viously reported by other authors, alkali treatment significantly en-
observed a negligible amount of cellulose II at room temperature after hances the tensile modulus and strength of coir fibres [13,14,34],
treatment with 10 wt% NaOH. However, in this work the remaining owing to the denser packing of the chains of cellulose fibrils promoted
Na+ ions might have contributed to reduce the new intermolecular by swelling. This fact is corroborated by the experimental results de-
hydrogen bonds, consequently affecting the thermal stability of the scribed in Section 3.2.1. The increase in the tensile modulus of the
cellulose. natural fibre is due to interacting factors, [54]. Alkali treatments pro-
The thermal events above 400 °C may be attributed to (1) lignin mote the rupture of sensitive bonds between cellulose and hemi-
degradation, which starts at room temperature and proceeds up to cellulose. The removal of hemicellulose induces rearrangements of the
700 °C and beyond [1,15,43] or (2) aromatization involving dehy- cellulose chains with the formation of new hydrogen bonds. Hence, the
drogenation reactions [53]. fibres tend to get closely packed [1,54,55]. As a result, the fibre swells
making the fibre more homogeneous and contributing to stress transfer
3.3. Mechanical tests between interfibrillar regions [1,55].
The predominant loading mode in random fibre networks is the
Table 2 depicts the average values obtained for all mechanical tests transfer of axial stress directly from one fibre to another at their in-
with the respective standard deviation. tersection [56]. The increase in roughness provided by Na2CO3 (up to
Table 3 presents the mechanical properties of the epoxy and 96 h) can, therefore, improve the axial stress transfer, and hence the
polyester resins in pristine conditions. It can be observed that the epoxy composite strength (16.8 and 14.8%, for epoxy and polyester, respec-
matrix is better in strength while polyester is better in modulus and tively). However, even after alkali treatment, fibre-reinforced compo-
impact resistance. sites presented much lower tensile strength than their neat polymer
The analysis of variance for tensile, flexural and impact tests is counterparts (56.12 and 63.9% for epoxy and polyester, respectively)
described in Table 4. The underlined p-values represent the factors or (Table 3). In addition, longer periods of treatment (above 96 h) also
interaction of factors that were statistically significant to the responses. reduced the tensile strength owing to fibre degradation. The tensile
The Anderson-Darling normality test presented pAD-values greater than strength, for a given fibre content, depends on the fibre orientation and
0.05, and R2 was greater than 91.25%, indicating that the data is well the fibre-matrix interfacial adhesion [1,5,28]. Fibres are randomly
described by the underlying statistical model used for ANOVA. distributed in the matrix and only those oriented perpendicular to the
applied load offer reinforcement to the composite. Fibres aligned
3.3.1. Tensile test transversely to the load direction virtually do not offer any reinforce-
Fig. 7a and b presents the typical tensile stress/strain curves of ment and act as barriers that impair the homogeneity of the stress
polyester and epoxy coir-reinforced composites, respectively. As re- distribution throughout the matrix. Such high concentration regions
vealed by the statistical analysis presented below, it is clear from these result in fracture that usually occurs at very low stress [5,57], which
figures that sodium carbonate enhances the tensile modulus of the coir- reduces the tensile strength of the composite material, as also observed
reinforced composites. The tensile strength is also significantly en- by Yan et al. [5].
hanced for 96 h of alkaline treatment, especially for epoxy composites.
Table 2 displays the average values obtained for tensile modulus and 3.3.2. Three-point bending test
strength with the respective standard deviation. The tensile modulus Fig. 9a and b shows the typical flexural load/displacement curves of
varied from 2.32 to 3.13 GPa. Matrix Phase (MP) as well as Treatment the coir-polyester and coir-epoxy polyester derived composites, re-
Time (TT) significantly affected the tensile modulus (p-values < 0.05) spectively. These figures indicate that the flexural maximum load is
and their effect was analysed via the interaction plot presented in significantly enhanced for 96 h of alkaline treatment, especially for
Fig. 8a, revealing that higher tensile modulus was achieved when epoxy composites. Such observations are corroborated by the statistical
polyester was used. analysis presented below. The flexural modulus of the composites
The tensile strength varied from 12.57 to 20.75 MPa. Only the main varied from 2.27 to 2.87 GPa (Table 2). Both factors – Matrix Phase and
factors contributed significantly to the tensile strength. Tensile modulus Treatment Time – significantly contributed to the flexural modulus,
and strength both increased with treatment time up to 96 h. The tensile which was increased by the coir fibre inclusions. The main effect plots
strength, however, presents the opposite behaviour, and higher strength are shown in Fig. 10a. The sodium carbonate treatment also enhanced
was achieved when coir fibres were embedded in the epoxy polymer the flexural modulus, which was higher for 96 h. As reported in the
literature, the modulus of raw coir fibres varies from 4.9 to 6.0 GPa [2],
Table 4 and increases with alkali treatments [13,54,58]. Such enhancement is
Analysis of variance for – ANOVA (p ≤ 0.05). due to the incorporation of such rigid fibres. Since the flexural modulus
Experimental factors Tensile Flexural Impact
is measured at small strains, its dependence on the interfacial adhesion
is small. Polyester, which presents a slightly higher flexural modulus,
Modulus Strength Modulus Strength Flatwise provided also higher flexural modulus compared to epoxy composites.
The flexural strength, also affected by Matrix Phase and Treatment
Main MP 0.000 0.000 0.000 0.000 0.000
TT 0.000 0.007 0.000 0.003 0.031
Time, varied from 23.27 to 42.94 MPa (Table 2). Epoxy composites
presented higher flexural strength compared to polyester composites
Interaction 0.000 0.507 0.157 0.462 0.056 (Fig. 10b). This behaviour is the same reported for tensile tests and
(MP x TT) deserves further analysis.
Weak adhesion between coir fibres (hydrophilic) and thermosetting
R2 (Adj.,%) 99.57 93.90 96.94 91.25 99.59
pAD (pAD ≥ 0.05) 0.659 0.319 0.654 0.982 0.729 polymers (hydrophobic) affects the flexural strength, and better fibre-
matrix interaction occurs if the fibre is somehow more hydrophobic

539
J.C. dos Santos et al. Polymer Testing 67 (2018) 533–544

Fig. 7. Typical tensile stress/strain curves of the a) coir-derived polyester and b) coir-derived epoxy composites.

[14,46,59]. The adhesion between two solids is directly related to their composites are presented in Fig. 11a. Impact strength values varied
surface energies and wetting is a requisite to get a good interfacial from 6.13 to 18.59 kJ/m2. Polyester composites presented higher im-
adhesion [60]. Chemical treatments can in fact, to some extent, change pact strength, which was 193% higher compared to epoxy polymer
the nature of the fibre surface from hydrophilic to hydrophobic due to composites (Fig. 11b). These results were expected, since impact
condensation reactions which expose the fibre surface to compatible strength is deeply affected by fibre-matrix adhesion and, in general,
chemical groups of the matrix, promoting better wetting by hydro- very strong interfaces have a detrimental effect on impact properties.
phobic matrices such as epoxy polymer [1,5] and polyester [3,58,61]. Fracture is usually associated with the formation of macro-cracks in the
As already discussed above (section 3.2), Na2CO3 deeply removes matrix, followed by fibre failure. Most of the energy absorbed during
hemicellulose exposing lignin. Lignin being a phenolic natural polymer impact is related to fibre pull-out [57].
[62], should be chemically compatible with polyester [13]. However, Crack formation takes place in the matrix, which is usually less
when a certain amount of lignin is removed, an increased amount of resistant than fibres. Behind the crack front, bridging fibres stretch
hydroxyl groups present in the cellulose becomes available [16,46,63]. freely along the separating crack faces and, in analogy to Hook's law,
Epoxy polymers are characterized by the presence of more than two absorb energy that will otherwise be available at the crack tip. Fibres
epoxide groups per molecule, which react with the hydroxyl groups of also stretch along the debond length, where stretching is restricted by
cellulose at the interface increasing chemical bonding [14,64]. As a friction, which dissipates additional (thermal) energy once sliding takes
result, epoxy composites presented superior flexural (and tensile) place along the fibre-matrix interface. As the stress increases, if the fibre
strength relatively to polyester composites. failure occurs along the debond length, the fibre will pull-out and
It is also noteworthy that the fibre-matrix adhesion is also enhanced thermal energy will be dissipated due to the development of shear
by the alkali treatment owing to increased mechanical anchoring and frictional forces during sliding. Considering the abundance of fibres
wetting, resulting from the superficial roughening of fibres. dispersed in the coir fibre composites and the vast surface area avail-
able for fibre sliding in various directions, pull-out is considered a
3.3.3. Impact test powerful energy dissipation mechanism. Fibre pull-out is evident in
Charpy impact data obtained for raw and treated coir reinforced polyester composites, as shown in Fig. 12a.

Table 2
Summary of the mechanical tests.
E.C. Tensile Flexural Impact

Modulus Strength Modulus (GPa) Strength (MPa) kJ/m2


(GPa) (MPa)

r1 r2 r1 r2 r1 r2 r1 r2 r1 r2

E-Raw 2.32 2.33 17.31 18.12 2.26 2.28 32.64 37.15 6.41 4.25
( ± 0.15) ( ± 0.14) ( ± 0.54) ( ± 0.6) ( ± 0.11) ( ± 0.08) ( ± 4.47) ( ± 4.98) ( ± 0.42) ( ± 0.17)
E-24 h 2.56 2.53 17.90 17.60 2.59 2.60 40.09 38.65 6.77 6.07
( ± 0.23) ( ± 0.19) ( ± 0.9) ( ± 1.08) ( ± 0.13) ( ± 0.10) ( ± 3.85) ( ± 4.56) ( ± 0.67) ( ± 0.84)
E-96 h 2.59 2.62 20.91 20.63 2.72 2.64 43.05 42.79 6.29 5.97
( ± 0.10) ( ± 0.20) ( ± 0.66) ( ± 1.86) ( ± 0.21) ( ± 0.08) ( ± 4.08) ( ± 4.40) ( ± 0.69) ( ± 0.30)
E-168 h 2.60 2.59 18.68 19.88 2.67 2.68 40.07 34.49 6.42 6.82
( ± 0.10) ( ± 0.08) ( ± 3.29) ( ± 1.75) ( ± 0.14) ( ± 0.18) ( ± 5.10) ( ± 4.35) ( ± 0.79) ( ± 1.17)
P-Raw 2.41 2.52 12.83 12.17 2.40 2.36 23.58 25.87 17.41 18.89
( ± 0.04) ( ± 0.05) ( ± 0.86) ( ± 1.25) ( ± 0.14) ( ± 0.12) ( ± 3.41) ( ± 3.60) ( ± 3.18) ( ± 2.46)
P-24 h 2.96 2.87 11.68 14.26 2.68 2.62 29.24 29.97 17.05 17.64
( ± 0.18) ( ± 0.31) ( ± 0.36) ( ± 0.67) ( ± 0.17) ( ± 0.08) ( ± 3.66) ( ± 3.68) ( ± 0.79) ( ± 1.90)
P-96 h 3.12 3.13 14.40 14.36 2.82 2.90 30.42 31.20 18.13 17.99
( ± 0.09) ( ± 2.95) ( ± 0.05) ( ± 2.27) ( ± 0.09) ( ± 3.37) ( ± 0.14) ( ± 5.28) ( ± 0.77) ( ± 2.61)
P-168 h 3.12 3.12 13.32 13.59 2.81 2.80 24.51 21.86 16.93 17.02
( ± 0.24) ( ± 0.31) ( ± 0.38) ( ± 1.01) ( ± 0.10) ( ± 0.11) ( ± 2.99) ( ± 4.34) ( ± 2.71) ( ± 2.24)

540
J.C. dos Santos et al. Polymer Testing 67 (2018) 533–544

Fig. 8. Tensile testing: a) interaction plot for tensile modulus and b) main effect plots for tensile strength.

Table 3 surface, rendering them more prone to premature rupture.


Mechanical properties of the matrices.
Property (Unit) Type of Matrix
4. Conclusions
Epoxy Polyester
Sodium carbonate treatment of coir fibres efficiently removed ex-
Tensile Strength (MPa) 47.29 ± 2.02 39.80 ± 3.05
ternal amorphous compounds and increased the exposure of crystalline
Modulus (GPa) 2.24 ± 0.11 2.44 ± 0.14
cellulosic fibres. Moreover, the swelling of the internal structure led to a
Flexural Strength (MPa) 69.26 ± 3.96 55.85 ± 2.91 substantial densification of the fibres. These effects promoted a sa-
Modulus (GPa) 2.14 ± 0.04 2.19 ± 0.08 tisfactory increase of the tensile and flexural modulus of the fibre re-
inforced composites for both polymeric matrices-epoxy and polyester-
Impact Resistance (kJ/m2) 6.33 ± 0.41 5.29 ± 0.31
studied.
The alkali treatment also increased the superficial roughness of coir
However, if the fibre-matrix interface is too strong, the mechanisms fibres, which increased the tensile and flexural strength relative to
of energy dissipation described above - fibre-matrix debonding, brid- untreated coir composites.
ging, sliding and pull-out - are inhibited and the impact resistance drops The alkali treatment resulted more effective for 96 h, period after
as the impact energy is canalized to matrix cracking [56,65,66]. This is which the tensile and flexural properties tend to decrease.
the case for epoxy polymer composites in which enhanced fibre-matrix However, the tensile and flexural strength of the fibre reinforced
interaction results in fracture of fibres at the crack plane with little fibre composites were in fact lower in comparison to the neat polyester and
pull-out, as shown in Fig. 12b. It is worth noting that Na2CO3 also led to epoxy matrices. As discussed, such behaviour is due to the random
a significant reduction, albeit small, of impact strength (see Fig. 11b). orientation of the fibres in the matrix. Only those oriented perpendi-
This behaviour may be explained by the enhancement of mechanical cular to the applied load offer reinforcement to the composite. Fibres
interlocking produced by the alkali treatment, owing to the superficial aligned transversely to the load direction act as stress concentration
roughening of the fibres (as discussed in Section 3.3.2), which reduces centres, which reduce the tensile strength of the composite material.
fibre pull-out (Fig. 12c) and impact energy dissipation. In addition, It is noteworthy that epoxy composites presented superior flexural
Dassios [66] has observed that, in the event of fibre sliding, the in- (and tensile) strength relative to polyester composites, since epoxy
creased superficial roughness also causes the tearing of interlocking polymers are characterized by the presence of more than two epoxide
fibre-matrix blocks, which magnifies the effects of failures on the fibre groups per molecule, which react with the hydroxyl groups of cellulose
at the interface, increasing chemical bonding.

Fig. 9. Typical flexural force/displacement curves of the a) coir-derived polyester and b) coir-derived epoxy composites.

541
J.C. dos Santos et al. Polymer Testing 67 (2018) 533–544

Fig. 10. Flexural tests: a) main effect plots for modulus and b) strength.

Fig. 11. Impact properties of coir fibre composites (a) strength variation of impact strength of epoxy and polyester composites and (b) main effect plots for impact
strength.

Fig. 12. Charpy impact test: fracture analysis of untreated (a) polyester, (b) epoxy; (c) 96 h treated polyester composites.

However, owing to this superior fibre-matrix interaction, the impact FAPEMIG (CEX - PPM-00075-17), CAPES (PhD scholarship) and CNPq
strength was lower for epoxy fibre reinforced composites since pull-out, (306767/2016-3) for the financial support provided.
the main mechanism for impact energy dissipation, is dramatically re-
duced. The same behaviour is observed, to a lesser extent, for alkaly Appendix A. Supplementary data
treated fibre reinforced polyester composites, as sodium carbonate in-
creases the superficial roughness of the fibres, and thus fibre-matrix Supplementary data related to this article can be found at http://dx.
interlocking. doi.org/10.1016/j.polymertesting.2018.03.043.
In summary, the sodium carbonate treatment of coir fibres for 96 h
provides significant enhancement of tensile (and flexural) modulus and References
strength of coir fibre reinforced epoxy and polyester composites and
overcomes the major operational problems with conventional alkaline [1] V. Fiore, T. Scalici, F. Nicoletti, G. Vitale, M. Prestipino, A. Valenza, A new eco-
treatments such as corrosion, extensive neutralization and environ- friendly chemical treatment of natural fibres: effect of sodium bicarbonate on
properties of sisal fibre and its epoxy composites, Compos. B Eng. 85 (2016)
mental hazards. 150–160, http://dx.doi.org/10.1016/j.compositesb.2015.09.028.
[2] L.Q.N. Tran, T.N. Minh, C.A. Fuentes, T.T. Chi, A.W. Van Vuure, I. Verpoest,
Investigation of microstructure and tensile properties of porous natural coir fibre for
Acknowledgements use in composite materials, Ind. Crop. Prod. 65 (2015) 437–445, http://dx.doi.org/
10.1016/j.indcrop.2014.10.064.
[3] S.N. Monteiro, L.A.H. Terrones, J.R.M. D'Almeida, Mechanical performance of coir
The authors would like to thank the Brazilian Research Agencies

542
J.C. dos Santos et al. Polymer Testing 67 (2018) 533–544

fiber/polyester composites, Polym. Test. 27 (2008) 591–595, http://dx.doi.org/10. Properties of Polymer Matrix Composite Materials vol. 1, (2014), pp. 1–13, http://
1016/j.polymertesting.2008.03.003. dx.doi.org/10.1520/D3039.
[4] O. Faruk, A.K. Bledzki, H.P. Fink, M. Sain, Biocomposites reinforced with natural [31] ASTM D790-15, Standard Test Methods for Flexural Properties of Unreinforced and
fibers: 2000-2010, Prog. Polym. Sci. 37 (2012) 1552–1596, http://dx.doi.org/10. Reinforced Plastics and Electrical Insulating Materials D790 (2015), pp. 1–12,
1016/j.progpolymsci.2012.04.003. http://dx.doi.org/10.1520/D0790-15E02.
[5] L. Yan, N. Chouw, L. Huang, B. Kasal, Effect of alkali treatment on microstructure [32] ISO Standard 179-1, International Organization for Standardization, Plastics–
and mechanical properties of coir fibres, coir fibre reinforced-polymer composites Determination of Charpy Impact Properties, (2010).
and reinforced-cementitious composites, Construct. Build. Mater. 112 (112) (2016) [33] M. Brahmakumar, C. Pavithran, R.M. Pillai, Coconut fibre reinforced polyethylene
168–182, http://dx.doi.org/10.1007/s10570-016-1116-6. composites: effect of natural waxy surface layer of the fibre on fibre/matrix inter-
[6] K.L. Pickering, M.G.A. Efendy, T.M. Le, A review of recent developments in natural facial bonding and strength of composites, Compos. Sci. Technol. 65 (2005)
fibre composites and their mechanical performance, Compos. Appl. Sci. Manuf. 83 563–569, http://dx.doi.org/10.1016/j.compscitech.2004.09.020.
(2016) 98–112, http://dx.doi.org/10.1016/j.compositesa.2015.08.038. [34] T.H. Nam, S. Ogihara, N.H. Tung, S. Kobayashi, Effect of alkali treatment on in-
[7] F.C. Campbell, Matrix resin systems, Structural Composite Materials, 2010, pp. terfacial and mechanical properties of coir fiber reinforced poly(butylene succinate)
63–100. biodegradable composites, Compos. B Eng. 42 (2011) 1648–1656, http://dx.doi.
[8] L. Yan, S. Su, N. Chouw, Microstructure, flexural properties and durability of coir org/10.1016/j.compositesb.2011.04.001.
fibre reinforced concrete beams externally strengthened with flax FRP composites, [35] B. Medronho, B. Lindman, Competing forces during cellulose dissolution: from
Compos. B Eng. 80 (2015) 343–354, http://dx.doi.org/10.1016/j.compositesb. solvents to mechanisms, Curr. Opin. Colloid Interface Sci. 19 (2014) 32–40, http://
2015.06.011. dx.doi.org/10.1016/j.cocis.2013.12.001.
[9] T. Lecompte, A. Perrot, A. Subrianto, A. Le Duigou, G. Ausias, A novel pull-out [36] M. Cai, H. Takagi, A.N. Nakagaito, K. Kusaka, M. Katoh, Y. Li, Influence of alkali
device used to study the influence of pressure during processing of cement-based treatment on internal microstructure and tensile properties of abaca fibers, Adv.
material reinforced with coir, Construct. Build. Mater. 78 (2015) 224–233, http:// Mater. Res. 1110 (2015) 302–305, http://dx.doi.org/10.4028/www.scientific.net/
dx.doi.org/10.1016/j.conbuildmat.2014.12.119. AMR.1110.302.
[10] M. Ramesh, K. Palanikumar, K.H. Reddy, Plant fibre based bio-composites: sus- [37] B. Medronho, B. Lindman, Brief overview on cellulose dissolution/regeneration
tainable and renewable green materials, Renew. Sustain. Energy Rev. 79 (2017) interactions and mechanisms, Adv. Colloid Interface Sci. 222 (2015) 502–508,
558–584, http://dx.doi.org/10.1016/j.rser.2017.05.094. http://dx.doi.org/10.1016/j.cis.2014.05.004.
[11] L. Zhang, Y. Hu, Novel lignocellulosic hybrid particleboard composites made from [38] S.H. Aziz, M.P. Ansell, The effect of alkalization and fibre alignment on the me-
rice straws and coir fibers, Mater. Des. 55 (2014) 19–26, http://dx.doi.org/10. chanical and thermal properties of kenaf and hemp bast fibre composites: Part 1-
1016/j.matdes.2013.09.066. polyester resin matrix, Compos. Sci. Technol. 64 (2004) 1219–1230, http://dx.doi.
[12] S. Haghdan, G.D. Smith, Natural fiber reinforced polyester composites: a literature org/10.1016/j.compscitech.2003.10.001.
review, J. Reinforc. Plast. Compos. 34 (2015) 1179–1190, http://dx.doi.org/10. [39] J.G. Gwon, S.Y. Lee, G.H. Doh, J.H. Kim, Characterization of chemically modified
1177/0731684415588938. wood fibers using FTIR spectroscopy for biocomposites, J. Appl. Polym. Sci. 116
[13] S.V. Prasad, C. Pavithran, P.K. Rohatgi, Alkali treatment of coir fibres for coir- (2010) 3212–3219.
polyester composites, J. Mater. Sci. 18 (1983) 1443–1454, http://dx.doi.org/10. [40] A. El Oudiani, Y. Chaabouni, S. Msahli, F. Sakli, Crystal transition from cellulose i to
1007/BF01111964. cellulose II in NaOH treated Agave americana L. fibre, Carbohydr. Polym. 86 (2011)
[14] S.M.S. Kumar, D. Duraibabu, K. Subramanian, Studies on mechanical, thermal and 1221–1229, http://dx.doi.org/10.1016/j.carbpol.2011.06.037.
dynamic mechanical properties of untreated (raw) and treated coconut sheath fiber [41] K. Kobayashi, S. Kimura, E. Togawa, M. Wada, Crystal transition from Na-cellulose
reinforced epoxy composites, Mater. Des. 59 (2014) 63–69, http://dx.doi.org/10. IV to cellulose II monitored using synchrotron X-ray diffraction, Carbohydr. Polym.
1016/j.matdes.2014.02.013. 83 (2011) 483–488, http://dx.doi.org/10.1016/j.carbpol.2010.08.006.
[15] M.F. Rosa, B. sen Chiou, E.S. Medeiros, D.F. Wood, T.G. Williams, L.H.C. Mattoso, [42] a. C.H. Barreto, D.S. Rosa, P.B. a. Fechine, S.E. Mazzetto, Properties of sisal fibers
W.J. Orts, S.H. Imam, Effect of fiber treatments on tensile and thermal properties of treated by alkali solution and their application into cardanol-based biocomposites,
starch/ethylene vinyl alcohol copolymers/coir biocomposites, Bioresour. Technol. Compos. Part A Appl. Sci. Manuf 42 (2011) 492–500, http://dx.doi.org/10.1016/j.
100 (2009) 5196–5202, http://dx.doi.org/10.1016/j.biortech.2009.03.085. compositesa.2011.01.008.
[16] A. Karthikayen, K. Balamurugan, Effect of alkali treatment and fiber length on [43] M.F. Rosa, E.S. Medeiros, J.A. Malmonge, K.S. Gregorski, D.F. Wood,
impact behavior of coir fiber reinforced epoxy composites, J. Sci. Ind. Res. 17 L.H.C. Mattoso, G. Glenn, W.J. Orts, S.H. Imam, Cellulose nanowhiskers from co-
(2012) 1–5, http://dx.doi.org/10.1007/s12221-011-0073-9. conut husk fibers: effect of preparation conditions on their thermal and morpho-
[17] J. Rout, M. Misra, S.S. Tripathy, S.K. Nayak, A.K. Mohanty, The influence of fibre logical behavior, Carbohydr. Polym. 81 (2010) 83–92, http://dx.doi.org/10.1016/j.
treatment on the performance of coir - polyester composites, Compos. Sci. Technol. carbpol.2010.01.059.
61 (2001) 1303–1310, http://dx.doi.org/10.1016/S0266. [44] J. Gassan, A.K. Bledzki, Alkali treatment of jute Fibers : relationship between
[18] S. Jayabal, S. Velumani, P. Navaneethakrishnan, K. Palanikumar, Mechanical and structure and mechanical properties, J. Appl. Polym. Sci. (1998) 623–629, http://
machinability behaviors of woven coir fiber-reinforced polyester composite, Fibers dx.doi.org/10.1002/(SICI)1097-4628(19990124)71:4<623::AID-APP14>3.0.
Polym. 14 (2013) 1505–1514, http://dx.doi.org/10.1007/s12221-013-1505-5. CO;2-K.
[19] S. Chaitanya, I. Singh, Sisal fiber-reinforced Green composites: effect of ecofriendly [45] Y. Liu, H. Hu, X-ray diffraction study of bamboo fibers treated with NaOH, Fibers
fiber treatment, Polym. Compos. (2017), http://dx.doi.org/10.1002/pc.24511. Polym. 9 (2008) 735–739, http://dx.doi.org/10.1007/s12221-008-0115-0.
[20] V. Fiore, T. Scalici, A. Valenza, Effect of sodium bicarbonate treatment on me- [46] S.H.P. Bettini, A.C. Biteli, B.C. Bonse, Andreia de A. Morandim-Giannetti,
chanical properties of flax-reinforced epoxy composite materials, J. Compos. Mater. Polypropylene Composites Reinforced With Untreated and Chemically Treated Coir:
(2017), http://dx.doi.org/10.1177/0021998317720009. effect of the Presence of Compatibilizer, Polym. Eng. Sci. (2015) 2050–2057,
[21] V. Fiore, G. Di Bella, A. Valenza, The effect of alkaline treatment on mechanical http://dx.doi.org/10.1002/pen.
properties of kenaf fibers and their epoxy composites, Compos. B Eng. 68 (2015) [47] G.G. Silva, D.A. De Souza, J.C. Machado, D.J. Hourston, Mechanical and thermal
14–21, http://dx.doi.org/10.1016/j.compositesb.2014.08.025. characterization of native brazilian coir fiber, J. Appl. Polym. Sci. 76 (2000)
[22] S. Mirmohamadsadeghi, Z. Chen, C. Wan, Reducing biomass recalcitrance via mild 1197–1206, http://dx.doi.org/10.1002/(SICI)1097-4628(20000516)
sodium carbonate pretreatment, Bioresour. Technol. 209 (2016) 386–390, http:// 76:7<1197::AID-APP23>3.0.CO;2-G.
dx.doi.org/10.1016/j.biortech.2016.02.096. [48] I.F. Pinheiro, A.R. Morales, L.H. Mei, Polymeric biocomposites of poly (butylene
[23] L. Oliveira, J.C. Santos, T.H. Panzera, R.T.S. Freire, L.M.G. Vieira, F. Scarpa, adipate-co-terephthalate) reinforced with natural Munguba fibers, Cellulose 21
Evaluation of hybrid-short-coir-fibre-reinforced composites via full factorial design, (2014) 4381–4391, http://dx.doi.org/10.1007/s10570-014-0387-z.
Compos. Struct. (2018), http://dx.doi.org/10.1016/j.compstruct.2018.01.088. [49] I.K. Varma, D.S. Varma, M. Varma, Thermal behaviour of coir fibres, Thermochim.
[24] ASTM D276-12, Standard test methods for identification of fibers in textiles, Astm Acta 108 (1986) 199–210.
(2012) 1–14, http://dx.doi.org/10.1520/D0276-00AR08.2. [50] A. Rencurosi, J. Röhrling, J. Pauli, A. Potthast, C. Jäger, S. Pérez, P. Kosma,
[25] L. Silva, T. Panzera, V. Velloso, J. Rubio, a. Christoforo, F. Scarpa, Statistical design A. Imberty, Polymorphism in the crystal structure of the cellulose fragment ana-
of polymeric composites reinforced with banana fibres and silica microparticles, J. logue methyl 4-O-methyl-β-D-glucopyranosyl-(1-4)-β-D-glucopyranoside, Angew.
Compos. Mater. 47 (2012) 1199–1210, http://dx.doi.org/10.1177/ Chem. Int. Ed. 41 (2002) 4277–4281, http://dx.doi.org/10.1002/1521-
0021998312446499. 3773(20021115)41:22<4277::AID-ANIE4277>3.0.CO;2-I.
[26] J.C. Santos, L.M.G. Vieira, T.H. Panzera, M. a. Schiavon, A.L. Christoforo, F. Scarpa, [51] M. Zuber, K.M. Zia, I.A. Bhatti, Z. Ali, M.U. Arshad, M.J. Saif, Modification of
Hybrid glass fibre reinforced composites with micro and poly-diallyldimethy- cellulosic fibers by UV-irradiation. Part II: after treatments effects, Int. J. Biol.
lammonium chloride (PDDA) functionalized nano silica inclusions, Mater. Des. 65 Macromol. 51 (2012) 743–748, http://dx.doi.org/10.1016/j.ijbiomac.2012.07.001.
(2014) 543–549, http://dx.doi.org/10.1016/j.matdes.2014.09.052. [52] M.G. Northolt, H. Boerstoel, H. Maatman, R. Huisman, J. Veurink, H. Elzerman, The
[27] A.E.O. Ben Sghaier, Y. Chaabouni, S. Msahli, F. Sakli, Morphological and crystalline structure and properties of cellulose fibres spun from an anisotropic phosphoric acid
characterization of NaOH and NaOCl treated Agave americana L. fiber, Ind. Crop. solution, Polymer (Guildf) 42 (2001) 8249–8264, http://dx.doi.org/10.1016/
Prod. 36 (2012) 257–266, http://dx.doi.org/10.1016/j.indcrop.2011.09.012. S0032-3861(01)00211-7.
[28] P. Muensri, T. Kunanopparat, P. Menut, S. Siriwattanayotin, Effect of lignin removal [53] I.K. Varma, A. Krishnan, S. Krishnamoorthy, Effect of chemical treatment on
on the properties of coconut coir fiber/wheat gluten biocomposite, Compos. Part A thermal behavior of jute fibers, Textil. Res. J. (1988) 486–494.
Appl. Sci. Manuf 42 (2011) 173–179, http://dx.doi.org/10.1016/j.compositesa. [54] D.S. Varma, M. Varma, I.K. Varma, Coir fibers, Textil. Res. J. 54 (1984) 827–832,
2010.11.002. http://dx.doi.org/10.1177/004051758405401206.
[29] D.R. Mulinari, C.A.R.P. Baptista, J.V.C. Souza, H.J.C. Voorwald, Mechanical prop- [55] K. Obi Reddy, C. Uma Maheswari, M. Shukla, J.I. Song, A. Varada Rajulu, Tensile
erties of coconut fibers reinforced polyester composites, Procedia Eng 10 (2011) and structural characterization of alkali treated Borassus fruit fine fibers, Compos. B
2074–2079, http://dx.doi.org/10.1016/j.proeng.2011.04.343. Eng. 44 (2013) 433–438, http://dx.doi.org/10.1016/j.compositesb.2012.04.075.
[30] ASTM D3039/D3039M − 14, Standard, Standard Test Method for Tensile [56] W. Hamad, Fibrous structures: networks and composites, Cellulosic Materials:

543
J.C. dos Santos et al. Polymer Testing 67 (2018) 533–544

Fibres, Networks and Composites, 2002, p. 171. alkali treatment on the fine structure of coir fibres (Cocos Nucifera), J. Mater. Sci.
[57] A.C. De Albuquerque, K. Joseph, L. Hecker De Carvalho, J.R.M. D'Almeida, Effect of 31 (1996) 721–726, http://dx.doi.org/10.1007/BF00367891.
wettability and ageing conditions on the physical and mechanical properties of [62] J. Xu, J. Jiang, W. Dai, Y. Xu, Liquefaction of sawdust in hot compressed ethanol for
uniaxially oriented jute-roving-reinforced polyester composites, Compos. Sci. the production of bio-oils, Process Saf. Environ. Protect. 90 (2012) 333–338, http://
Technol. 60 (2000) 833–844, http://dx.doi.org/10.1016/S0266-3538(99)00188-8. dx.doi.org/10.1016/j.psep.2012.01.001.
[58] S. Kuriakose, D. Varma, V.G. Vaisakh, Mechanical Behaviour of Coir Reinforced [63] P.N. Khanam, H.A. Khalil, Tensile, flexural and chemical resistance properties of
Polyster Composites: an Experimental Investigation vol. 2, (2012). sisal fibre reinforced polymer composites: effect of fibre surface treatment, J. Polym
[59] C.A.S. Hill, H.P.S. Abdul Khalil, Effect of fiber treatments on mechanical properties (2011) 115–119, http://dx.doi.org/10.1007/s10924-010-0219-7.
of coir or oil palm fiber reinforced polyester composites, J. Appl. Polym. Sci. 78 [64] S. Ebnesajjad, Theories of adhesion, Surface Treatment of Materials for Adhesive
(2000) 1685–1697, http://dx.doi.org/10.1002/1097-4628(20001128) Bonding Surface Treatment of Materials for Adhesive Bonding, 2014.
78:9<1685::AID-APP150>3.0.CO;2-U. [65] C. Wang, X. Ji, A. Roy, V.V. Silberschmidt, Z. Chen, Shear strength and fracture
[60] L.Q.N. Tran, C.A. Fuentes, C. Dupont-Gillain, A.W. Van Vuure, I. Verpoest, Wetting toughness of carbon fibre/epoxy interface: effect of surface treatment, Mater. Des.
analysis and surface characterisation of coir fibres used as reinforcement for com- 85 (2015) 800–807, http://dx.doi.org/10.1016/j.matdes.2015.07.104.
posites, Colloid. Surface. Physicochem. Eng. Aspect 377 (2011) 251–260, http://dx. [66] K.G. Dassios, A review of the pull-out mechanism in the fracture of brittle- matrix
doi.org/10.1016/j.colsurfa.2011.01.023. fibre-reinforced composites, Adv. Compos. Lett. (2007) 17–24.
[61] S. Sreenivasan, P. Bhama Iyer, K.R. Krishna Iyer, Influence of delignification and

544

Вам также может понравиться