Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Journal of Natural Fibers

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/wjnf20

Manufacturing and Characterization of New


Composite Based on Epoxy Resin and Lygeum
Spartum L. Plant

Zouheyr Belouadah , Mansour Rokbi & Abdelaziz Ati

To cite this article: Zouheyr Belouadah , Mansour Rokbi & Abdelaziz Ati (2020): Manufacturing
and Characterization of New Composite Based on Epoxy Resin and Lygeum Spartum L. Plant,
Journal of Natural Fibers, DOI: 10.1080/15440478.2020.1856273

To link to this article: https://doi.org/10.1080/15440478.2020.1856273

Published online: 28 Dec 2020.

Submit your article to this journal

Article views: 16

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=wjnf20
JOURNAL OF NATURAL FIBERS
https://doi.org/10.1080/15440478.2020.1856273

ARTICLE

Manufacturing and Characterization of New Composite Based on


Epoxy Resin and Lygeum Spartum L. Plant
a
Zouheyr Belouadah , Mansour Rokbib, and Abdelaziz Atic
a
Laboratoire des Sciences et Techniques de l'Environnement, Ecole Nationale Polytechnique, , El-
Harrach, Alger, Algéria; bDépartement de Génie Mécanique, Université Mohamed Boudiaf - M'sila, , Algéria;
c
Laboratoire De Technologie Des Matériaux Et Génie Des Procédés, Faculté De Technologie, Université Bejaia 06000,
Algéria
ABSTRACT
In the present study, a new composite based on epoxy resin reinforced with Lygeum spartum L. fibers was
prepared using compression molding technique. Chemical, mechanical as well as morphological analyses were
carried out on the Epoxy/Lygeum spartum composite (Epoxy/LS) using ATR-FTIR, tensile and flexural tests, and
SEM spectroscopy. An increase in the intensity of several bands associated to the main complements of
vegetable fibers (cellulose, hemicellulose, and lignin) was showan using ATR-FTIR analysis. An improvement
in the mechanical properties of this new material was observed, with respect to the neat epoxy resin. This was of
the order of 117 % and 85 % for tensile and flexural modulus respectively, and of the order of 40 % and 9 % for
tensile and flexural strength, correspondingly. A good fiber-matrix adhesion was also observed from tensile
fracture surface of composite using scanning electron microscope (SEM). All the obtained results demonstrated
the effectiveness of the Lygeum spartum fibers as reinforcement of composite materials.
KEYWORDS Natural fiber; composite; epoxy; Lygeum spartum L.; mechanical properties; atr-FTIR

Introduction
Research development on the use of composites reinforced with natural fibers began in 1900
(Westman et al. 2010). For many decades, it has been a topic of great interest. In 1930s, Henry Ford
developed a soy matrix reinforced with natural fibers to form natural fiber composites used in car body
(Salit 2014). Since the 1970s, kenaf, hemp, sisal, flax, and jute have extensively been investigated for
composite applications (Sarasini and Fiore 2018), and in the last few years, interest has increased and
research on natural fibers has intensified. This is explained by the variety and multiplicity of their
characteristics such as biodegradability, low density, non-toxicity, low cost, and recyclability
(Belouadah, Ati, and Rokbi 2015; Newman et al. 2010; Sreenivasan et al. 2011). These characteristics
also justify the replacement of synthetic fibers (glass, Kevlar, carbon . . .) by natural ones. Vegetable
fiber can be obtained from: stems (Alfa, Flax, Hemp, Jute, and Kenaf), fruits (Cotton, Oil palm empty
fruit bunch and Coir) and leaves (Agave, Abaca, Henequen, Banana, and Sisal). Long or short fibers,
particulates, or whisker are the most common forms of reinforcement materials (Dandekar and Shin
2012). Long fibers are preferred over the short ones (Rokbi et al. 2018; Vaikhanski and Nutt 2003).
Joseph et al. (Joseph 2002) have reported that a composite with fiber length less than five times the
critical length of the fiber have a strength significantly lower than that of a continuous fiber composite
with the same fiber volume fraction. In another study, Fiore, Di Bella, and Valenza (2015) reported
that the Kenaf fibers, arranged along the load direction in the UD composites, give higher stiffness and
strength, and lower void contents compared to the composites reinforced with short fibers randomly
oriented (MAT). Composites reinforced with natural fibers environmentally friendly, are attracting
great interest in many industrial sectors: transportation (aeronautics, automobiles), construction and
building industry (partition boards, ceiling paneling), military applications, consumer products and
packaging, biomedical applications, and 3D printing (Mohammadinejad et al. 2019; Sahu 2013; Ullah

CONTACT Zouheyr Belouadah zouheyr.belouadah@g.enp.edu.dz Laboratoire Des Sciences Et Techniques De


L’environnement LSTE, El-Harrach, Alger, Algéria
© 2020 Taylor & Francis
2 Z. BELOUADAH ET AL.

and Chen 2020). They are also used in thermal insulation of buildings and in the manufacture of paper
(Abdelaziz et al. 2016; Salit 2014). This fact is mainly explained by the interesting properties (physico-
chemical, thermal, and mechanical properties) exhibited by the selected composites. Although these
vegetable fibers have several benefits, they also have some drawbacks such as their anisotropic
structure, their high rate of moisture absorption, and their hydrophilic nature, which prevents
a good adhesion to the polymeric matrix. Recently, several authors have investigated the potential
use of some natural fiber as new reinforcement agents of composite materials in several stages, starting
by studying the fibers individually, then treating and testing them in composites. For example, Indran
and Raj (2015) have first characterized, the properties of new natural fibers extracted from Cissus
quadrangularis stem as reinforcement of composite materials. As a second step, they investigated
(Indran et al. 2015), the properties of composites based on unsaturated polyester matrix reinforced by
these fibers. In a similar work, novel natural cellulose fabric from Manicaria saccifera palm was
characterized as a possible reinforcement of composite materials by Porras, Maranon, and Ashcroft
(2015), while in another work, these fibers were used to reinforce the PolyLactic Acid (PLA) as novel
green composite lamina (Porras, Maranon, and Ashcroft 2016). The previous authors found that the
addition of Manicaria fabric considerably improves the tensile properties of neat PLA by 51% and 26%
of Young’s modulus and the tensile strength, respectively. The present work focuses on bringing out
the contribution of Lygeum spartum fibers to improve the properties of composites materials.
Amiralian and Martin (2017) reported that the Lygeum spartum plant can be used in the development
of composite material, comprising an elastomer and nanocellulose, which can be used in a wide variety
of potential uses, including: industrial seals, wear liners in mining applications, tyres, conveyor belts,
balloon manufacture, etc. Arenas et al.(2020) examined the sound-absorption properties of Lygeum
spartum and Stipa tenacessima fibers as alternatives to synthetic ones in the manufacturing of sound-
absorbing materials, and found that practical and sustainable sound-absorbing panels can be con­
structed based on these two kinds of fibers. Salem et al. (2020) evaluated the fiber composition of 21
xero-halophytic species from Tunisian sabkhas, and found that some xero-halophytes like Aristida
pungens, Lygeum spartum, Atriplex halimus, etc., can be used in the production of high-quality
lightweight paper. In the present work, we have developed unidirectional composites based on
epoxy as a polymer matrix reinforced with the Lygeum Spartum L. fibers. The chemical analysis of
composites was studied with attenuated total reflectance-Fourier transform infrared spectroscopy
(ATR-FTIR). The tensile and flexural properties of composites are investigated and compared to
others composite materials reported in literature. The fracture surface of tensile specimens was
analyzed by using scanning electron microscopy (SEM)

Materials and experimental procedure


Materials
The used epoxy resin has a trade name of “MEDAPOXY STR”. It is supplied by the company
GRANITEX-Algeria. It comes in the form of a pre-metered kit of 2 components: element A (resin)
and element B (hardener). The physical and mechanical properties of this type of resin are shown in
Table 1. In this table, the tensile properties were measured according to ISO 527B.

Table 1. Some properties of MEDAPOXY STR resin.


Properties Epoxy resin (MEDAPOXY STR)
Density (ISO 758) (g cm−1) 1.1 ± 0.05
Viscosity (NF T76-102) mPaS (25°C) 11,000
Hardening time at 20°C and 65% HR Tack-free 6 h
Hard 16 h
Tensile strength (MPa) 63.39 ±1.42
Tensile modulus (GPa) 3.30 ± 0.32
Flexural strain (%) 5.54 ± 0.2
Flexural strength(MPa) 109.15 ± 7.4
Flexural modulus (GPa) 3.24 ± 0.2
JOURNAL OF NATURAL FIBERS 3

The fiber reinforcement, Lygeum spartum L. plant was harvested in Algeria. An eco-friendly
technique has been adopted to extract the fibers. The protocol of fiber extraction is well detailed in
our previous work (Belouadah, Ati, and Rokbi 2015).

Manufacturing of unidirectional composites


The choice of the manufacturing technique of composite materials depends on the type of matrix
(Thermoplastic, thermosetting), and the quality of the reinforcement (Short, long). In general, the
techniques used for the manufacturing of composites reinforced with plant fibers are the same as
those used in the manufacturing of composites reinforced with synthetic fibers. In this study, it is
the compression molding technique, which is used to manufacture the composites because it has
a significant effect on the final quality of the composite. The L. spartum fibers are unidirectionally
oriented in the composite. The fibers were cut to the length of 100 mm and were placed in an oven
at 70°C for 2 hours. Fibers were then weighed to ensure the same fiber weight fraction in each
composite plate specimen, and thereafter, they were placed in a metal mold. The assembly mold-
fibers was placed in a glass desiccator connected to a vacuum pump for degassing the fibers during
14 min.
For the preparation of the matrix, the epoxy resin was mixed with the hardener at a ratio by weight
of (2: 1). This mixture was carried out using a stirrer at low speed for 3 min to obtain good
homogeneity. Subsequently, the resulting mixture has undergone a degassing for a period of 13 min­
utes to remove the air bubbles using a vacuum pump and a glass desiccator. The degassed mixture was
then poured onto the fibers contained in the mold. A counter-mold was put in place to ensure
compression by means of a tightening of the bolts that make it possible both to eliminate the air-
bubbles and to maintain an uniform thickness of about 2 mm of composite plate. In order to avoid
adhesion of the composite plate to the mold, the inner surface of the latter was covered by an adhesive
tape. The mold was kept in an ambient temperature and the composite plates were demolded after
18–20 hours. The obtained unidirectional composites specimens were kept for 20 days followed by
drying in an oven at 70°C for 4 hours.
The numerous tests carried out to finalize this protocol of elaboration allowed us to show the great
importance of the two stages of drying and degassing on the final quality of composite plates
manufactured. This is well demonstrated in Figure 1. Indeed, in addition to the initial protocol of
drying and degassing, this step allows to significantly eliminate the air bubbles.

Figure 1. Micrograph of an area of a composite specimen: a) elaboration without drying of the fibers and without degassing of the
fibers and resin, b) elaboration with drying of the fibers and with degassing of the fibers and resin.
4 Z. BELOUADAH ET AL.

Characterization methods
ATR–FTIR analysis
The attenuated total reflectance–Fourier transform infrared spectroscopy (ATR–FTIR) technique
was used to analyze the chemical composition of composite plate Epoxy/LS. In this study, ATR-
FTIR analyses were performed using the Perkin–Elmer instrument at room temperature. This
machine is driven by computer software (Perkin Elmer Spectrum) with spectral resolution of
4 cm−1 and recorded in the region 4000–400 cm−1. Concerning the preparation of samples, pieces
of neat resin and of elaborate composite are placed on the ATR crystal and pushed gently with
a clamp.

Composite tensile test


The tensile tests on the unidirectional composite plates were carried out on five specimens using
a Zwick/Roell tensile test machine (Z250) equipped with a 2.5 kN load cell. During each tensile test, we
adopted an extensometer set to a base length of 30 mm. The tests were carried out in ambient
temperature at a crosshead speed of 1 mm/min. The unidirectional composite specimens have the
dimensions: 2x6x100 mm3. Composite tabs Epoxy/glass fibers (± 45 °) have been glued on each side of
the tensile specimen to avoid the rupture at the ends and to allow the transfer of stress to the useful
part of the specimen (Figure 2).
As regards to the fiber volume fraction in composites, many tests on composites elaboration
have been carried out using various fiber proportions to achieve the optimum percentage of
reinforcement in composites of 25%. The research works of Sathishkumar, Navaneethakrishnan,
and Shankar (2012) on the tensile and flexural properties of composites with isophthalic polyester
matrix reinforced with Snake fibers showed that composites reinforced with 25% of the fibers
offer the best mechanical properties. For Napier fiber-reinforced composites, Haameem et al.
(2016) also show that it is the rate of 25% of fibers, which gives the composite the best tensile and
flexural strength.
For this experiment, fibers were weighed beforehand to maintain the same fiber content by weight
(25%) in all the composite plates. However, to achieve this, it is necessary to calculate the volume ratio
of fibers which was performed using the image analysis software (ImageJ) (Schneider, Rasband, and
Eliceiri 2012). This software has already been used by several researchers in the same context
(Abdelaziz et al. 2016; Coroller et al. 2013; De Andrade Silva, Chawla, and de Toledo Filho 2008;
Martin, Davies, and Baley 2014). Martin, Davies, and Baley (2014) compared two techniques to
calculate the volume fraction of flax fibers reinforced epoxy composite using the image analysis
(imageJ) and the rule of mixtures of composite. They reported a similar volume fraction values for
the two techniques. Figure 3 illustrates the different steps for calculating the fiber ratio using the
ImageJ software.
Several micrographs on cross sections of two composite plates across several zones were taken
using a scanning electron microscope (SEM) to confirm the desired ratio. Each micrograph was
then processed by performing a binary thresholding black and white to distinguish the fibers (in
black) from the matrix (in white). The software calculates the area of each element and then

Figure 2. Scheme of tensile specimen of the unidirectional composite Epoxy/Vegetable fibers with composite heel Epoxy/glass fibers.
JOURNAL OF NATURAL FIBERS 5

Figure 3. Different steps of micrograph processing by imagej software. a) raw SEM image, b) selection of the cross-section, c) black
and white thresholding, d) particle analysis and surface detection.

generates a report representing the number of detected elements, the surface area of each element,
the total area of the elements (fibers), and the proportion of black area (fibers) relative to the surface
of the black element (matrix). The volume fraction of fibers measured using this technique was
23.41 ± 0.14%.

Composite flexural test


The three-point bending tests on the different test pieces were carried out according to ASTM D790-
10 using Zwick/Roell testing machine. The latter is equipped with 2.5 kN load cell and crosshead speed
of 2 mm/min. The ratio of the distance between the supports “L” and the thickness “h” of the specimen
(L/h) was fixed at 16 mm (Figure 4).
The flexural stress is evaluated using the following equation:
3FL
σf ¼ (1)
2bd2
6 Z. BELOUADAH ET AL.

Figure 4. Principle of a three-point bending test. (A) Three point bending system adaptable in Zwick/Roell machine.(B) Main
dimensions required for the three-point bending test: “L” length between the two fixed supports, the width “b” and the thickness “h”
of the specimen.

Where: F is the applied load expressed in newtons (N), L is the support span (mm), b is the width of
the specimen in millimeters (mm), and d is the Thickness of the specimen (mm).
The flexural modulus was calculated by the following formula:

mL3
Ef ¼ (2)
4bd3

Where, m: is the slope of tangent to the initial straight-line portion of the load-deflection curve (N/mm)

Fracture surface analysis


Microscopic analysis was performed to observe the detailed microstructure of the tensile fracture
surface of Epoxy/LS composite and thus to determine the fiber content in this composite. This analysis
was carried out with a PHILIPS XL30 Scanning Electron Microscope (SEM) at an electron beam-
accelerating potential of 20 kV.
JOURNAL OF NATURAL FIBERS 7

Results and discussion


ATR-FTIR Analysis
Figure 5 shows the ATR-FTIR spectra of Epoxy/LS composite and that of pure epoxy resin. The
identification of some of the most important peaks was achieved according to the literature
(Dharmalingam et al. 2020; Fiore, Scalici, and Valenza 2014; González, Baselga, and Cabanelas
2012; Liu et al. 2019; Saravanakumar et al. 2013; Sathishkumar et al. 2013; Sreekumar et al. 2008).
The large band centered at 3321 cm −1 is due to tretching of hydroxyl group H–O bonds of cellulose.
The bands peaking at 2925 cm −1 and 2851 cm −1 are attributed to C-H stretching vibration of CH and
CH2 groups in cellulose and hemicellulose and that around 1737 cm −1 to C = O stretching vibration of
carboxylic acid in lignin or ester components in hemicellulose. The wave number around 1610 cm −1 is
associated with the presence of water in the fibers. The intense peak at 1510 cm −1 is associated to
C = C stretching of aromatic ring of the lignin in Lygeum spartum fiber, whereas that located at around
1456 cm−1 corresponds to the CH2 symmetric bending band which represents the hemicellulose
xylene, wax, and impurities. The smaller band centered at 1363 cm−1 is associated to the bending
vibration of C-H and C-O groups of the aromatic ring in polysaccharides. The intense peak at 1242 cm
−1
corresponds to the C-O stretching vibration of the acetyl group in lignin and hemicellulose.
Furthermore, the presence of peak around 1180 cm −1 corresponds to the C-O-C stretching vibration
of the pyranose ring in polysaccharides. The vibratory activity around 1106 cm−1 corresponds to C–
O–C stretching and C-O stretching of cyclic alcohols of cellulose and glucoside linkage of lignin. The
other more intense peak at 1034 cm −1 is associated to the (C-O) groups of cellulose, while that present
at 826 cm −1 is attributed to the presence of β-glycosidic linkages in cellulose and hemicellulose. The
comparison of these two spectra (neat resin and Epoxy/LS) shows that the reinforcement of the epoxy
resin by Lygeum spartum fibers gives rise to an enhancement in the intensity of several bands such as:
3321 cm −1, 2925 cm −1, 2851 cm −1, 1737 cm −1, 1180 cm −1, 1106 cm −1, and 1034 cm −1. These bands
are associated with the cellulose, hemicellulose, and lignin, which are the main constituents of plant
fibers.

Figure 5. Superposition of the ATR-FTIR spectra of pure epoxy resin and Epoxy/LS composite.
8 Z. BELOUADAH ET AL.

Composite tensile properties


The mechanical behavior of a composite material depends mainly on the nature of the matrix and of
the fibers, the fiber content and the quality of the interfacial adhesion Fibers/Matrix. In order to
highlight the main conclusions of the composite tensile behavior and the beneficial effect of
L. spartum fibers as reinforcement in an epoxy resin matrix, we have assembled in histograms,
the results of the mechanical tensile properties of the Epoxy/LS composite and neat epoxy resin
(Figure 6). The beneficial effect of L. spartum fibers, as reinforcements of composite materials with
an epoxy resin base, is clearly established in view of the higher obtained values of tensile strength
and Young’s modulus of composite compared to the neat epoxy resin. This superiority is due
essentially to the high rigidity of L. spartum fibers (13 GPa of Young’s modulus and 280 MPa of
tensile strength, (Belouadah, Ati, and Rokbi 2015)) and to the high interfacial adhesion of fibers with
the epoxy matrix. The improvement obtained by this type of composite with respect to the neat
epoxy resin is of the order of 117% for Young’s modulus and 40% for tensile strength. Le and
Pickering (2015) reported an enhancement of the tensile properties of the neat epoxy resin by the
addition of aligned long harakeke fiber.
Table 2 summarizes the tensile properties of Epoxy/LS composite and other unidirectional fiber-
reinforced epoxy composites taken from previously published works. One may infer from this table
that the obtained tensile modulus values in Epoxy/LS composite with 23% of fibers content are higher
that than obtained in Epoxy/Sisal (Padmavathi, Naidu, and Rao 2012), Epoxy/Piassava (Nascimento
et al. 2012), and Epoxy/Agave American (Mylsamy and Rajendran 2011) with fibers content of 22.8,
20, and 35%, respectively. In the same context, the values obtained of the tensile strength are higher
than the ones found in Epoxy/Kenaf (Mahjoub et al. 2014), Epoxy/Piassava (Nascimento et al. 2012),
and Epoxy/Agave American (Mylsamy and Rajendran 2011), with fibers content of 10, 20, and 35%,
respectively. Compared to the Epoxy/LS composite, large values of tensile strength and Young’s
modulus have been recorded when the epoxy was reinforced by the Jute fibers (Hossain et al. 2013),
Scutched Flax fibers, and Flax tows fibers (Martin, Davies, and Baley 2014). This is mainly due to the
higher specific stiffness values of individual fibers, which are 55.66 ± 2.11 GPa, 47.0 ± 15.7 GPa, and
50.8 ± 15.7 GPa, respectively.

Figure 6. Tensile and Flexural properties of the Epoxy/LS composite and neat epoxy resin.
JOURNAL OF NATURAL FIBERS 9

Table 2. Comparison of tensile properties of Epoxy/LS with other unidirectional fiber-reinforced composites resulting from the
literature.
Composite consti­
tuents tensile
Composite modulus (GPa) Composite tensile properties
Matrix/Fibers (content Tensile strength Young’s Strain to fail­
%) Matrix Fiber (MPa) Modulus (GPa) ure (%) References
Epoxy/Lygeum. 3.30 ± 4.47– 88.38 ± 11.24 7.17 ± 0.34 1.41 ± 0.11 Current work
S (23.41 ± 0.14) 0.32 13.27
Epoxy/Kenaf (10) 2.131 40 58 6.8 - (Mahjoub et al. 2014)
Epoxy/Kenaf (30) 2.131 40 124 14.4 - (Mahjoub et al. 2014)
Epoxy/Sisal (22.8) 3.5 - 172.36 6.02 (Padmavathi, Naidu, and Rao
2012)
Epoxy/Jute (25) 3.89 ± 55.66 ± 112.69 ± 18.31 14.59 ± 2.28 0.82 ±0.17 (Hossain et al. 2013)
0.53 2.11
Epoxy/Phormium (51) - - 211 14.7 ± 0.8 - (Newman et al. 2010)
Epoxy/UT. Kenaf (48.6) 2.17 40.32 ± 95.4 10.34 - (Fiore, Di Bella, and Valenza
9.27 2015)
Epoxy/T. Kenaf (48.6) 2.17 37.82 ± 106.1 10.7 -
6.78
Epoxy/Sisal (32) 3.1– 3.2 24 132 15 1.2 (Oksman, Wallström, and
Berglund 2002)
Epoxy/Scut Flax (26 ± 2.7 ± 47.0 ± 234 ± 12 16.0 ± 1.0 1.7 ± 0.1 (Martin, Davies, and Baley
0.4) 0.1 15.7 2014)
Epoxy/Flax tows (27 ± 2.7 ± 50.8 ± 203 ± 25 14.6 ± 2.7 1.4 ± 0.1
3.0) 0.1 15.7
Epoxy/Harakeke (55) 3.91 32.09 223 (14) 16.8 (0.62) 1.44 (Le and Pickering 2015)
Epoxy/Piassava (20) 2.35 ± - 71.45 ± 15.68 2.65 ± 0.62 - (Nascimento et al. 2012)
0.31
Epoxy/NT. Agave. - - - 0.263 - (Mylsamy and Rajendran
A (35) 2011)
Epoxy/T. Agave. A (35) - - 41.2 0.270 -

Composite flexural properties


The histogram in Figure 6 shows the results of the mechanical flexural properties of the Epoxy/LS
composite and neat Epoxy resin. It is noted that for the strength properties, the Lygeum spartum fibers
used as reinforcement in the epoxy matrix confer a significant gain of 85% in flexural modulus and of
9% in flexural strength. However, there is a strong decrease in the flexural strain to failure with respect
to the neat resin of the order of 53%. As for the case of tensile properties, the notable improvement of
the flexural strength and flexural modulus values as well as the strong decrease on the strain to failure
rate of composite is due essentially to: (i) the high stiffness of L. Spartum fibers and to, (ii) the high
interfacial adhesion of the fibers with the epoxy matrix.
Table 3 summarizes the flexural strength, the flexural modulus and the strain to failure rate values
of the Epoxy/LS composite and neat resin. In assessing the flexural mechanical characteristics, we have
added to this table the results of other types of unidirectional fiber-reinforced epoxy composites cited
in the literature.
The table shows that the obtained flexural modulus values in Epoxy/LS composite with 23% of
fibers content are similar with those obtained in Epoxy/Hemp composite (Yousif et al. 2012) and in
Epoxy/treated Sugar Palm composite (Sapuan, Bachtiar, and Hamdan 2010), with fibers content of
38–41% and 10%, respectively. It is also noted that this value is higher than that one found in Epoxy/
Piassava (Nascimento et al. 2012) with fibers content of 20% and lower than that found in Epoxy/Fique
(Hoyos and Vázquez 2012) and in Epoxy/Harakeke (Le and Pickering 2015) with fibers content of 40%
and 49%, respectively. With regard to the flexural strength, the obtained values in Epoxy/LS composite
are higher than that found in Epoxy/Piassava (Nascimento et al. 2012) and Epoxy/Sugar Palm
(Sapuan, Bachtiar, and Hamdan 2010) with fibers content of 20% and of 10%, respectively, and
10 Z. BELOUADAH ET AL.

Table 3. Comparison of flexural properties of Epoxy/LS composite with other with other unidirectional fiber-reinforced epoxy
composites resulting from the literature.
Composite Flexural strength Young’s Modulus Strain to failure
Matrix/Fibers (fiber content %) (MPa) (GPa) (%) References
Epoxy 109.15 ± 7.4 3.24 ± 0.2 5.54 ± 0.2 Current work
Epoxy/Lygeum spartum 118.99 ± 11.8 6.02 ± 0.3 2.57 ± 0.3
(23.41 ± 0.14)
Epoxy/Harakeke (49) 223 13.7 - (Le and Pickering 2015)
Epoxy/Piassava (20) 60.15 ± 29.45 2.96 ± 0.90 (Nascimento et al. 2012)
Epoxy/NT. Hemp (38–41) 235.13 ± 27 5.572 ± 0.8 - (Yousif et al. 2012)
Epoxy/T Hemp (38–41) 301.64 ± 72 6.743 ± 1.8 - (Yousif et al. 2012)
Epoxy/Jute (25) 138.94 ± 18.62 10.31 ±3.95 1.94 ±0.70 (Hossain et al. 2013)
Epoxy/Fique (40) 168 ± 11 8.0 ± 0.7 2.7 ± 0.4 (Hoyos and Vázquez 2012)
Epoxy/NT. Sugar. P (10) 77.73 ± 9.39 2.805 ± 0.927 - (Sapuan, Bachtiar, and Hamdan
Epoxy/T Sugar. P (10) 58.17 ± 8.97 6.947 ± 1.389 - 2010)
Epoxy/T Bamboo (48) 279 ± 10 23 ± 0.6 2.2 (Osorio et al. 2011)

lower than that found in Epoxy/Harakeke (Le and Pickering 2015), Epoxy/Hemp (Yousif et al. 2012),
Epoxy/Fique (Hoyos and Vázquez 2012), and in Epoxy/treated Bamboo (Osorio et al. 2011) with fibers
content of 49%, 38–41%, 40%, and 48%, respectively.

Fracture surface analysis


Figure 7 shows a typical example of the tensile fracture surface of Epoxy/LS composite observed by
scanning electron microscope (SEM). The micrograph (a) gives a global overview at low magnification
of the observed fracture surface of one end of the tensile specimen. There is a smooth-fractured surface
of the epoxy matrix. Moreover, no decohesion at the interfaces Fibers/Matrix is visible around all
fibers. We zoomed on a part in the central area of the micrograph (a) which indicated by yellow
square. This zoom of a magnification (300X) is presented in the micrograph (b). It shows the detailed
morphology of a broken fiber composed of fiber cells (elementary fibers) and vascular bundles. Visible
holes on the fiber center are the lumens of the fiber cells. On the one hand, no presence of epoxy resin
has been visible within these lumens. Oksman, Wallström, and Berglund (2002) have observed the
same phenomenon on fractured cross sections of the sisal–epoxy composites. On the other hand, it is
clearly observed that the rupture of the fibers took place in the same plane of rupture of the matrix.
This could be due to the good fiber/matrix adhesion that justifies the good tensile and flexural
properties (Young’s modulus and strength).

Figure 7. SEM micrographs of tensile fracture surface of Epoxy/LS composite. (a) Global view revealing a smooth facies of the epoxy
matrix and the good adhesion Fibers/matrix.(b) Zoom of part of the micrograph (a).
JOURNAL OF NATURAL FIBERS 11

Conclusion
This paper constitutes a contribution to the major research axis of the composite materials develop­
ment reinforced with vegetable fibers, which have become an important issue in view of their use in
various industrial sectors. Lygeum spartum fibers, available in Algeria, are selected for manufacturing
unidirectional composite materials with epoxy matrix.
The chemical, mechanical (tensile and flexural three-point bending), and morphological character­
ization, of Epoxy/LS composite made it possible to draw the following conclusions:

● The ATR-FTIR analysis shows an increase in the intensity of several bands, which associated to
main constituents of plant fibers (cellulose, hemicellulose, and lignin), after the reinforcement of
the epoxy resin by Lygeum spartum fibers.
● The beneficial effect of L. spartum fibers, as reinforcements of composite materials with an epoxy
resin base, is clearly established in view of the higher obtained values of tensile and flexural
properties of composite compared to the neat epoxy resin. This improvement is due essentially
to: (i) the high stiffness of L. Spartum fibers and to, (ii) the high interfacial adhesion of the fibers
with the epoxy matrix.
● The tensile fracture surface of Epoxy/LS composite observed by scanning electron microscope
(SEM) shows that the no decohesion at the interface Fibers/Matrix and the rupture of the fibers at
the same plane of the rupture of matrix are the most failure mechanisms under tensile test. This is
due to the very strong bond between fibers and matrix that justifies the good tensile and flexural
properties.

Taking into account the mechanical properties obtained in tensile and flexural tests (Young’s modulus
and strength), we conclude that the Lygeum spartum fibers have proved their effectiveness as
reinforcement of composite materials as this study constitutes a successful experiment based on the
available literature.

Acknowledgments
Authors would like to thank DGRSDT (Direction Générale de la Recherche Scientifique et du Développement
Technologique) for their support, URMPE of boumerdes university, Pr. Chabira from Laghouat University for their
technical assistance. Mme BOUAYED Maliha from National Polytechnic School of Algiers, for her language consulting.

ORCID
Zouheyr Belouadah http://orcid.org/0000-0003-4640-7693
Mansour Rokbi http://orcid.org/0000-0001-5856-662X
Abdelaziz Ati http://orcid.org/0000-0002-2854-8815

References
Abdelaziz, S., S. Guessasma, A. Bouaziz, R. Hamzaoui, J. Beaugrand, and A. A. Souid. 2016. Date palm spikelet in mortar:
Testing and modelling to reveal the mechanical performance. Construction and Building Materials 124:228–36. doi:
doi:doi:10.1016/j.conbuildmat.2016.07.039
Amiralian, N. and D. J. Martin. 2017. Nanocomposite Elastomers. US Patent 2017/0333602 A1, filed december 8, 2014,
and issued november 5, 2017
Arenas, J. P., R. Del Rey, J. Alba, and R. Oltra. 2020. Sound-absorption properties of materials made of esparto grass
fibers. Sustainability 2020 (12):5533. doi:10.3390/su12145533.
Belouadah, Z., A. Ati, and M. Rokbi. 2015. Characterization of new natural cellulosic fiber from Lygeum spartum L.
Carbohydrate Polymers 134:429–37. doi:doi:doi:10.1016/j.carbpol.2015.08.024
Coroller, G., A. Lefeuvre, A. Le Duigou, A. Bourmaud, G. Ausias, T. Gaudry, and C. Baley. 2013. Effect of flax fibres
individualisation on tensile failure of flax/epoxy unidirectional composite. Composites. Part A, Applied Science and
Manufacturing 51:62–70. doi:10.1016/j.compositesa.2013.03.018.
12 Z. BELOUADAH ET AL.

Dandekar, C. R., and Y. C. Shin. 2012. Modeling of machining of composite materials: A review. International Journal of
Machine Tools & Manufacture 57:102–21. doi:doi:doi:10.1016/j.ijmachtools.2012.01.006
De Andrade Silva, F., N. Chawla, and R. D. de Toledo Filho. 2008. Tensile behavior of high performance natural (sisal)
fibers. Composites Science and Technology 68:3438–43. doi:10.1016/j.compscitech.2008.10.001.
Dharmalingam, S., O. Meenakshisundaram, V. Elumalai, and R. S. Boopathy. 2020. An investigation on the interfacial
adhesion between amine functionalized luffa fiber and epoxy resin and its effect on thermal and mechanical proper­
ties of their composites. Journal of Natural Fibers. doi:10.1080/15440478.2020.1726238.
Fiore, V., G. Di Bella, and A. Valenza. 2015. The effect of alkaline treatment on mechanical properties of kenaf fibers and
their epoxy composites. Composites Part B: Engineering 68:14–21. doi:10.1016/j.compositesb.2014.08.025.
Fiore, V., T. Scalici, and A. Valenza. 2014. Characterization of a new natural fiber from Arundo donax L. as potential
reinforcement of polymer composites. Carbohydrate Polymers 106:77–83. doi:10.1016/j.carbpol.2014.02.016.
González, M. G., J. Baselga, and J. C. Cabanelas. 2012. Applications of FTIR on epoxy resins-identification, monitoring
the curing process, phase separation and water uptake. Infrared Spectroscopy - Materials Science, Engineering and
Technology, InTech 13: 261–84.
Haameem, M. J. A., M. S. Abdul Majid, M. Afendi, H. F. A. Marzuki, I. Fahmi, and A. Gibson. 2016. Mechanical
properties of Napier grass fibre/polyester composites. Composite Structures 136:1–10. doi:10.1016/j.
compstruct.2015.09.051.
Hossain, M. R., M. A. Islam, A. Van Vuurea, and I. Verpoest. 2013. Tensile behavior of environment friendly jute epoxy
laminated composite. Procedia Engineering 56:782–88. doi:10.1016/j.proeng.2013.03.196.
Hoyos, C. G., and A. Vázquez. 2012. Flexural properties loss of unidirectional epoxy/fique composites immersed in water
and alkaline medium for construction application. Composites Part B: Engineering 43 (8):3120–30. doi:10.1016/j.
compositesb.2012.04.027.
Indran, S., and R. E. Raj. 2015. Characterization of new natural cellulosic fiber from Cissus quadrangularis stem.
Carbohydrate Polymers 117:392–99. doi:10.1016/j.carbpol.2014.09.072.
Indran, S., R. E. Raj, B. Daniel, and S. Saravanakumar. 2015. Cellulose powder treatment on Cissus quadrangularis stem
fiber-reinforcement in unsaturated polyester matrix composites. Journal of Reinforced Plastics and Composites 35
(3):212–27. doi:10.1177/0731684415611756.
Joseph, S. 2002. A comparison of the mechanical properties of phenol formaldehyde composites reinforced with banana
fibres and glass fibres. Composites Science and Technology 62 (14):1857–68. doi:10.1016/S0266-3538(02)00098-2.
Le, T. M., and K. L. Pickering. 2015. The potential of harakeke fibre as reinforcement in polymer matrix composites
including modelling of long harakeke fibre composite strength. Composites. Part A, Applied Science and
Manufacturing 76:44–53. doi:10.1016/j.compositesa.2015.05.005.
Liu, Y., X. Lv, J. Bao, J. Xie, X. Tang, J. Che, Y. Ma, and J. Tong. 2019. Characterization of silane treated and untreated
natural cellulosic fibre from corn stalk waste as potential reinforcement in polymer composites. Carbohydrate
Polymers 218:179–87. doi:10.1016/j.carbpol.2019.04.088.
Mahjoub, R., J. M. Yatim, A. R. M. Sam, and M. Raftari. 2014. Characteristics of continuous unidirectional kenaf fiber
reinforced epoxy composites. Materials & Design 64:640–49. doi:10.1016/j.matdes.2014.08.010.
Martin, N., P. Davies, and C. Baley. 2014. Comparison of the properties of scutched flax and flax tow for composite
material reinforcement. Industrial Crops and Products 61:284–92. doi:10.1016/j.indcrop.2014.07.015.
Mohammadinejad, R., H. Maleki, E. Larrañeta, A. R. Fajardo, A. B. Nik, A. Shavandi, A. Sheikhi, M. Ghorbanpour,
M. Farokhi, and P. Govindh. 2019. Status and future scope of plant-based green hydrogels in biomedical engineering.
Applied Materials Today 16:213–46. doi:10.1016/j.apmt.2019.04.010.
Mylsamy, K., and I. Rajendran. 2011. The mechanical properties, deformation and thermomechanical properties of
alkali treated and untreated Agave continuous fibre reinforced epoxy composites. Materials & Design 32 (5):3076–84.
doi:10.1016/j.matdes.2010.12.051.
Nascimento, D. C. O., A. S. Ferreira, S. N. Monteiro, R. C. M. Aquino, and S. G. Kestur. 2012. Studies on the
characterization of piassava fibers and their epoxy composites. Composites. Part A, Applied Science and
Manufacturing 43 (3):353–62. doi:10.1016/j.compositesa.2011.12.004.
Newman, R. H., M. J. Le Guen, M. A. Battley, and J. E. Carpenter. 2010. Failure mechanisms in composites reinforced
with unidirectional phormium leaf fibre. Composites. Part A, Applied Science and Manufacturing 41 (3):353–59.
doi:10.1016/j.compositesa.2009.11.001.
Oksman, K., L. Wallström, and L. Berglund. 2002. Morphology and mechanical properties of unidirectional sisal–epoxy
composites. Journal of Applied Polymer Science 84 (13):2358–65. doi:10.1002/app.10475.
Osorio, L., E. Trujillo, A. Van Vuure, and I. Verpoest. 2011. Morphological aspects and mechanical properties of single
bamboo fibres and flexural characterization of bamboo/epoxy composites. Journal of Reinforced Plastics and
Composites 30 (5):396–408. doi:10.1177/0731684410397683.
Padmavathi, T., S. V. Naidu, and R. Rao. 2012. Studies on mechanical behavior of surface modified sisal fiber-epoxy
composites. Journal of Reinforced Plastics and Composites 31 (8):519–32. doi:10.1177/0731684412438954.
Porras, A., A. Maranon, and I. Ashcroft. 2015. Characterization of a novel natural cellulose fabric from manicaria
saccifera palm as possible reinforcement of composite materials. Composites Part B: Engineering 74:66–73.
doi:10.1016/j.compositesb.2014.12.033.
JOURNAL OF NATURAL FIBERS 13

Porras, A., A. Maranon, and I. Ashcroft. 2016. Thermo-mechanical characterization of manicaria saccifera natural fabric
reinforced poly-lactic acid composite lamina. Composites. Part A, Applied Science and Manufacturing 81:105–10.
doi:10.1016/j.compositesa.2015.11.008.
Rokbi, M., A. Imad, C. Herbelot, and Z. Belouadah. 2018. Fracture toughness of random short natural fibers polyester
composites. Diffusion Foundations 18: 94–105. doi:10.4028/www.scientific.net/DF.18.94
Sahu, J. 2013. STUDY OF TENSILE AND FLEXURAL PROPERTIES OF LUFFA FIBER REINFORCED EPOXY
COMPOSITES. B.Tech. Thesis. National Institute of Technology Rourkela, Deemed University, India.
Salem, S., S. Nasri, S. Abidi, A. Smaoui, N. Nasri, P. Mutjé, and K. B. Hamed. 2020. Lignocellulosic biomass from sabkha
native vegetation: A new potential source for fiber-based bioenergy and bio-materials. Sabkha Ecosystems: Springer
49: 407–12. doi:10.1007/978-3-030-04417-6_25
Salit, M. S. 2014. Tropical natural fibre composites: Properties, Manufacture and Applications. Singapore: Springer.
doi:10.1007/978-981-287-155-8
Sapuan, S., D. Bachtiar, and M. Hamdan. 2010. Flexural properties of alkaline treated sugar palm fibre reinforced epoxy
composites. International Journal of Automotive and Mechanical Engineering (IJAME) 1:79–90. doi:10.15282/
ijame.1.2010.7.0007.
Sarasini, F., and V. Fiore. 2018. A systematic literature review on less common natural fibres and their biocomposites.
Journal of Cleaner Production 195:240–67. doi:10.1016/j.jclepro.2018.05.197.
Saravanakumar, S. S., A. Kumaravel, T. Nagarajan, P. Sudhakar, and R. Baskaran. 2013. Characterization of a novel
natural cellulosic fiber from prosopis juliflora bark. Carbohydrate Polymers 92 (2):1928–33. doi:10.1016/j.
carbpol.2012.11.064.
Sathishkumar, T., P. Navaneethakrishnan, and S. Shankar. 2012. Tensile and flexural properties of snake grass natural
fiber reinforced isophthallic polyester composites. Composites Science and Technology 72 (10):1183–90. doi:10.1016/j.
compscitech.2012.04.001.
Sathishkumar, T., P. Navaneethakrishnan, S. Shankar, and R. Rajasekar. 2013. Characterization of new cellulose
sansevieria ehrenbergii fibers for polymer composites. Composite Interfaces 20 (8):575–93. doi:10.1080/
15685543.2013.816652.
Schneider, C. A., W. S. Rasband, and K. W. Eliceiri. 2012. NIH Image to ImageJ: 25 years of image analysis. Nature
Methods 9 (7):671–75. doi:10.1038/nmeth.2089.
Sreekumar, P., R. Saiah, J. M. Saiter, N. Leblanc, K. Joseph, G. Unnikrishnan, and S. Thomas. 2008. Effect of chemical
treatment on dynamic mechanical properties of sisal fiber-reinforced polyester composites fabricated by resin transfer
molding. Composite Interfaces 15 (2–3):263–79. doi:10.1163/156855408783810858.
Sreenivasan, V., S. Somasundaram, D. Ravindran, V. Manikandan, and R. Narayanasamy. 2011. Microstructural,
physico-chemical and mechanical characterisation of Sansevieria cylindrica fibres–An exploratory investigation.
Materials & Design 32 (1):453–61. doi:10.1016/j.matdes.2010.06.004.
Ullah, S., and X. Chen. 2020. Fabrication, applications and challenges of natural biomaterials in tissue engineering.
Applied Materials Today 20:100656. doi:10.1016/j.apmt.2020.100656.
Vaikhanski, L., and S. R. Nutt. 2003. Fiber-reinforced composite foam from expandable PVC microspheres. Composites.
Part A, Applied Science and Manufacturing 34 (12):1245–53. doi:10.1016/S1359-835X(03)00255-0.
Westman, M. P., L. S. Fifield, K. L. Simmons, S. Laddha, and T. A. Kafentzis. 2010. Natural fiber composites: A review.
Pacific Northwest National Laboratory, US Department of Energy. doi:10.2172/989448
Yousif, B., A. Shalwan, C. Chin, and K. Ming. 2012. Flexural properties of treated and untreated kenaf/epoxy composites.
Materials & Design 40:378–85. doi:10.1016/j.matdes.2012.04.017.

You might also like