You are on page 1of 18

Journal of Natural Fibers

ISSN: 1544-0478 (Print) 1544-046X (Online) Journal homepage: https://www.tandfonline.com/loi/wjnf20

Physical and Thermal Properties of Novel Native


Andean Natural Fibers

Sandra Mori, Samuel Charca, Elena Flores & Holmer Savastano Jr.

To cite this article: Sandra Mori, Samuel Charca, Elena Flores & Holmer Savastano Jr. (2019):
Physical and Thermal Properties of Novel Native Andean Natural Fibers, Journal of Natural Fibers,
DOI: 10.1080/15440478.2019.1629150

To link to this article: https://doi.org/10.1080/15440478.2019.1629150

Published online: 05 Jul 2019.

Submit your article to this journal

Article views: 18

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/10.1080/15440478.2019.1629150

ARTICLE

Physical and Thermal Properties of Novel Native Andean Natural


Fibers
Sandra Moria, Samuel Charcaa, Elena Floresa, and Holmer Savastano Jr.b
a
Departemt of Mechanical Engineering, Universidad de Ingeniería y Tecnología, UTEC, Lima-Peru, Perú; bDepartment
of Biosystems Engineering, FZEA, Pirassununga, Brazil

ABSTRACT KEYWORDS
Natural fibers are a renewable resource that has become an economic Natural fibers; ichu; alkali
alternative to replace synthetic fibers as reinforcement in composite mate- treatment; amorphous silica
rials. In this study, the aim was to characterize physical and thermal proper- particles; composite
materials
ties of the Andean Stipa obtusa and Jarava ichu leaves (grasses known as
ichu) before and after the alkaline treatment. Energy dispersive spectrum 关键词
identified external phytoliths on the surface of nontreated raw leaves. With 天然纤维; 碱处理; 无定形
the alkaline treatment, improvement on the thermal stability and increment 二氧化硅颗粒; 复合材料
in the crystalline index and size was observed. Additionally, raw leaves
present a hydrophobic behavior; however, with the treatment this hydro-
phobicity was lost for Stipa obtusasignificantly but not for the Jarava ichu.
Furthermore, low surface energy was observed for the nontreated raw
leaves and increasing consistently for the treated fibers. According with
these results, these grasses are quite interesting materials from which fibers
can be extracted for engineering applications such as composite materials.

摘要
天然纤维是一种可再生资源,已成为替代合成纤维作为复合材料增强材
料的经济替代品. 本研究旨在探讨安第斯山针茅和jarava ichu叶(禾本科
称为ichu)在碱处理前后的物理和热特性. 能量色散谱分析了未处理生叶
表面的外来植物体. 经碱处理后,热稳定性得到改善,结晶指数和晶粒尺
寸增大. 此外,生叶具有疏水性;然而,经过处理后,这种疏水性对钝刺
针茅明显丧失,而对jarava ichu则没有. 此外,未处理的生叶表面能较
低,处理后的纤维表面能持续增加. 根据这些结果,这些草是非常有趣的
材料,从中提取纤维可用于工程应用,如复合材料.

Introduction
In the last decades, lignocellulosic fibers from plants, also called “natural fibers (NFs)” are being
widely used by industries like textile, construction, packaging, automobile, aerospace, medical, and
sporting facilities (Cherian et al. 2011; Cheung et al. 2009; Sanjay et al. 2016; Savastano et al. 2016). It
is estimated that the NFs demand will be increasingly important during the next years, due to
legislations that pressure to the industries to reduce environmental impact of products and processes
(Adekomaya et al. 2017; COMMISSION, EUROPEAN 2000; Ferris 2016; Son et al. 2011). In this
framework, attentions on the use of new natural resources as source of NFs are increasing in the last
years, especially if the fibers obtained have outstanding price-performance ratio with low density,

CONTACT Samuel Charca samuel.charca@upr.edu Departemt of Mechanical Engineering, Universidad de Ingeniería


y Tecnología, UTEC, Lima-Peru
This article has been republished with minor changes. These changes do not impact the academic content of the article.
Color versions of one or more of the figures in the article can be found online at www.tandfonline.com/wjnf.
© 2019 Taylor & Francis
2 S. MORI ET AL.

that they can be applied in several industries with particular interest in polymeric composites
(Bismarck, Baltazar-Y-Jimenez, and Sarikakis 2006; Brown et al. 2014).
There are different methods to extract and treat NFs, from which alkaline treatment is the most
common method, especially using NaOH (Bachtiar, Sapuan, and Hamdan 2009; Boopathi, Sampath,
and Mylsamy 2012; Das and Chakraborty 2008; Le Troedec et al. 2008; Lu et al. 2013; Saha et al.
2010). Different researches have reported improvements in mechanical properties of alkaline-treated
NFs, which is mainly due to the enhancement on their dimensional stability and crystalline region
(Boopathi, Sampath, and Mylsamy 2012; Borchani, Carrot, and Jaziri 2015; Das and Chakraborty
2008, 2009; El et al. 2011; John and Anandjiwala 2008; Liu and Hong 2008; Lu et al. 2013; Park et al.
2010; Saha et al. 2010; Sawpan, Pickering, and Fernyhough 2011; Yanjun 2010). Moreover, alkaline
treatment improves its thermal stability, which allows their adequate behavior along life cycle in any
application (Kumar et al. 2008; Le Troedec et al. 2008; Monteiro et al. 2012; Norul Izani et al. 2013;
Reddy et al. 2013). On the other hand, NFs in their natural state are hydrophilic due to the presence
of hydroxyl groups (polysaccharides), which gives to the NFs polar characteristic, obtaining a poor
(fiber/matrix) interface when it interacts with nonpolar resins (matrix) (Doan, Brodowsky, and Edith
2012; Heng et al. 2007; Kevin and John 2007; Tran et al. 2011). However, with alkali treatment
hydrophilic can change to hydrophobic character and with that their compatibility to hydrophobic
polymer matrix can be improved, allowing proper stresses transfer from the matrix to the reinforce-
ment (fibers) (George, Sreekala, and Thomas 2001; Tran et al. 2013).
The grass plant commonly called ichu had been extensively used until 50 years ago along the
Andean region as a local construction material in roofs and as a rope material. However, with the
development of the new construction materials, its use has been gradually reduced. In the recent
years, ichu had been studied to access its thermal insulation properties, which show promising
results (Charca et al. 2015). Furthermore, fibers extraction methodology from ichu was studied and
some mechanical properties were determined, showing equivalent result to the commercial NFs
(Mori et al. 2018). Other advantages of this material are the great availability and low cost of
harvesting (~0.15 USD/kg) (Charca et al. 2015). The ichu plant is endemic, so it does not need to be
cultivated, and it does not compete with the production of any other agricultural products in the use
of land. Moreover, it is estimated that the annual production of ichu in Peru (South America) is
about 76 kt. The principal aim of this research was to characterize physical and thermal properties of
novel NFs, obtained from Stipa obtusa and Jarava ichu (species of ichu) before and after alkaline
treatment. Properties like morphology, crystallinity index, thermal degradation, contact angle, and
surface energy were studied and analyzed.

Materials and methodology


Materials and alkaline treatment
Raw Jarava ichu and Stipa obtusa were harvested from Cusco (Peru) at 3600 m.a.s.l (Figure 1). Since
these grasses have parts with different sensibility to defibrillation (leaves, stems, sheath, flowers, and
seeds), leaves were selected to perform this study and were named “raw leaves.” Raw leaves have
length between 300 and 500 mm with cross-section area between 0.13 and 4.52 mm2 and chemical
composition (according to TAPPI standard) is similar to conventional natural fibers, as shown in
Table 1 (Tenazoa et al. 2017; Meng et al. 2016; Yan et al. 2016; Abdul Khalil et al. 2010, Ali et al.
2016). Selected raw leaves were soaked in alkaline solution with different NaOH concentrations and
treatment time established in Table 2, which were determined according to the maximum properties
reached in polymeric composites materials under flexural test (fiber and interface quality) (Mori,
Flores, and Charca 2017; Mori et al. 2018). The main purpose of the alkaline treatment was to clean
the surface and remove certain amount of noncellulosic components from the surface and cause
certain level of defibrillation and stabilization. Densities of these fibers were measured using
Ultrapyc 1200-e pycnometer and the results are depicted in Table 3.
JOURNAL OF NATURAL FIBERS 3

Figure 1. Picture of (a) Stipa obtusa and (b) Jarava ichu plants in Tinta, Cusco, Peru.

Table 1. Chemical composition of ichu leaves (wt%) and of conventional fibers [40–44], according to TAPPI standard.
Composition Moisture Cellulose Hemicellulose Lignin Extractives
TAPPI standard T 421 om-02 T 203 cm-99 Herbst technique T 222 om-02 T 204 cm-97
Jarava ichu 8.6 ± 0.58 42.81 ± 0.51 28.71 ± 1.38 12.99 ± 0.26 16.48 ± 0.14
Stipa obtusa 9.33 ± 0.12 38.07 ± 0.91 26.52 ± 1.21 15.56 ± 0.72 14.57 ± 0.06
Bamboo (P. pubescens) – 41.54 ± 0.70 25.51 ± 1.1 22.06 ± 0.88 6.27 ± 1.53
Coir – 39.5 ± 3.5 0.2 43 ± 2 1.8
Kenaf – 53.8 33.9 21.2 6.4

Table 2. Alkaline treatment parameters.


Parameters Stipa obtusa Jarava ichu
Concentration (NaOH molar) 0.5 1.5
Temperature (°C) 70 70
Time (h) 3.5 1
Afterward fibers were washed with distilled water and dried at 60°C for 12 h

Table 3. Density of ichu fibers.


Fibers Condition Density (g/cm3)
Stipa obtusa Raw leaves 1.40
Alkaline treated 1.47
Jarava ichu Raw leaves 1.30
Alkaline treated 1.55

Morphology
The morphology of the materials was characterized using a scanning electron microscope (SEM)
TM300 Hitachi operated at 15 kV. Before the measurements, samples around 10 mm length were
dried at 60°C for 24 h. In order to identify chemical elements on the fiber surface, energy dispersive
4 S. MORI ET AL.

spectrum (EDS) technique was used. The roughness of fibers surface was measured through image
analysis in terms of root mean square (RMS) of roughness profile using Gwyddion software. Ten
measurements were carried out for each scanned area (1, 5, 10, 15 y 20 µm2).

X-ray diffraction (XRD)


The crystallinity characteristics of the cellulose were examined by XRD equipment (Miniflex 600,
Rigaku) using CuK radiation generated at 40 kV and 15 mA, measured from 5° to 30° with scan
steps of 0.02° at 2°/min. Samples were previously cut and sieved (28–30 mesh). Crystallinity index
(CI) was determined using Segal (Eq. 1) and Hermans (Eq. 2) method (Popescu et al. 2011; Segal
et al. 1959):
I002  Iam
C:I: ¼  100 (1)
I002
where I002 is the maximum peak intensity corresponding to the 002 plane and Iam is the maximum
intensity of the noncrystalline part of the material;
Acryst
C:I: ¼ (2)
Atotal
where Acryst is the area under the crystalline curves and Atotal is the total area under the XRD spectra.
In order to use adequately those methods, noncrystalline standard lignin spectra was obtained as
well.
The apparent crystallite size (L) was determined according to Scherrer equation (Eq. 3)
(Ahvenainen, Kontro, and Svedström 2016).
Kλ
L¼ (3)
β  cosðθÞ
where K is the Scherrer constant (0.94), λ is the X-ray radiation wavelength (0.1506), β is the half
height width of the analyzed band (plane), and θ is the Bragg angle corresponding to the analyzed
plane. In order to identify the peak, crystalline bands and β parameter, spectra subtraction was
performed from the original materials (raw leaves and alkaline treated) by the noncrystalline
material (standard lignin); afterward, the final spectra was deconvoluted using Gaussian fitting
curve.

Thermogravimetric analysis
The thermal degradation characterization was carried out using STA 449F3 Jupiter NETZH, with
nitrogen as purge gas at a flow rate of 50 mL/min at temperatures range from (26 ± 2)°C to 700°C
with heating rate of 10°C/min. Samples were previously weighted at approximately 5 mg and
afterward milled and sieved between 28 and 30 meshes.

Advancing contact angle and surface energy


The advancing contact angle (CA) was determined by tensiometry technique according to Wilhelmy
method (Eq. 4) using a Data-Physics Instruments GmbH tensiometry (model DCAT21) and SCAT
11 software.
FðhÞ ¼ p  γ  cosðθÞ (4)
where, p is the wet perimeter of the fiber and γ is the surface tension of the liquid (Mahesh 2014).
For the measurement, deionized water (W), ethylene glycol (EG), glycerol (G), and n-hexane
liquid were used, whose properties are shown in Table 4. For the practical purpose, the theoretical
JOURNAL OF NATURAL FIBERS 5

Table 4. Properties of liquid considered for the contact angle measurement.


Deionized Ethylene
Properties water (DW) glycol (EG) Glycerol (G) n-Hexane
Density (g/cm3) 1.00 1.113 1.2613 0.6603
Surface tension (mN/m) 72.75 48.45 63.4 18.43
Viscosity (mPa.s) 1.00 21.81 1412 0.308
Temperature (°C) 20.00 20.00 20.00 20.00

static contact angle for the n-hexane, θo = 0 was used for all fibers (Pucci et al. 2017). Samples were
prepared in 20 mm of length without any twist to make sure that they enter vertically into the liquid.
Since the cross-section area of the fibers is irregular and it is not uniform along the fiber length,
numbers of uniform fibers were selected randomly with similar thickness. Selected fibers were
embedded into the resin and polished afterward and the cross-sectional perimeter was measured
using optical microscope; results are presented in Table 5. Those results were used as a perimeter to
determine the advancing contact angle. Before the measurement, fibers were dried in an oven at 60°
C for 12 h in order to eliminate the moisture and after that they were stored in a desiccator.
Advancing contact angle measurement was performed in the raw leaves and alkaline treated samples
with minimum 10 replicates. To determine static contact angle (θo) the advancing contact angles were
measured considering six advancing velocities (0.002, 0.01, 0.05, 0.1, 0.5, and 1 mm/s), afterward fitted to the
molecular kinetic theory (MKT) equation (Eq. 5) (Blake 2006, 1993; Tran et al. 2011).
 2 
γL λ
υ ¼ 2Ko λ  sinh ðcosðθo Þ  cosðθÞÞ (5)
2kT

where Ko is the equilibrium displacement frequency, λ is the average length of each molecular
displacement, γL is the surface energy of the liquid, k is the Boltzmann constant, T is the absolute
temperature, and θ is the corresponding advancing contact angle to the velocity measured.
Finally, surface energy was estimated using Owens and Wendt method (Fuentes, Verpoest, and
Van Vuure 2015; Owens and Wendt 1969; Tran et al. 2011) where the surface energy is divided in
p p
two components γL and γdL for the liquid and γS and γdS for the solid. Equation 6 expresses the
relation between them.
0qffiffiffiffiffi1
qffiffiffiffiffi p
γL C qffiffiffiffiffi
γL ð1 þ cosðθo ÞÞ pB
qffiffiffiffiffi ¼ γS @qffiffiffiffiffiA þ γdS (6)
2 γLd
γLd

p p
where γL and γS are the polar component for the liquid and solid, respectively, γdL and γdS are the
dispersive component for
 qffiffiffiffi liquid and the solid as well and θo is the static contact angle. Plotting the
ffi qffiffiffiffiffi qffiffiffiffiffi
p
γL ð1 þ cosðθo ÞÞ= 2 γdL in the y axis and γL = γdL in the x axis, linear fit can be applied giving
qffiffiffiffiffi qffiffiffiffiffi
p
as a slope the γS and an intercept of the y axis the γdS .

Table 5. Cross-sectional perimeters (average and the standard deviation) for


Jarava ichu and Stipa obtusa raw leaves and alkali treated.
Perimeter (p, mm)
Fiber Condition Average SD
Jarava ichu Raw leaves 3.15 0.471
Alkaline treated 2.77 0.303
Stipa obtusa Raw leaves 4.73 0.407
Alkaline treated 0.45 0.039
6 S. MORI ET AL.

Figure 2. SEM images of surfaces on raw leaf of Stipa obtusa (a) and (b), alkaline-treated Stipa obtusa (c), raw leaf of Jarava ichu,
(d) and alkaline-treated Jarava ichu (e) and (f).

Results and analysis


Morphology
SEM micrographs of the fibers surface are shown in Figure 2. According to these images, rough
compact surface can be observed on the raw leaves for Stipa obtusa and Jarava ichu. However, after
alkaline treatment it surfaces has changed; where bundled of micro fiber cells are appreciated after
the treatment, which is similar to conventional NFs (Boopathi, Sampath, and Mylsamy 2012;
Borchani, Carrot, and Jaziri 2015; Das and Chakraborty 2008, 2009; El et al. 2011; John and
Anandjiwala 2008; Liu and Hong 2008; Lu et al. 2013; Park et al. 2010; Saha et al. 2010; Sawpan,
Pickering, and Fernyhough 2011; Yanjun 2010). On the other hand, at the surface of the raw leaves
(Stipa obtusa and Jarava ichu) extracellular phytoliths (rigid silica microstructures) are observed,
which agrees with silicon detected in that area using EDS analysis (Figure 3). Phytoliths hair shapes
for both species are quite similar; differences are mainly in their sizes (length), observing greater
values for the Jarava ichu than Stipa obtusa (42 ± 5.8) µm and (35 ± 4.8) µm, respectively. Moreover,
for the Jarava ichu dumbbell shapes were observed at the surface with average area of (66 ± 10) µm2;
for the Stipa obtusa ovals shapes was observed as well with average area of (50.1 ± 13) µm2 Dumbbell
shapes phytoliths observed are quite similar to the ones found in rice straw, but the latter are
intracellular, and their average area is greater (80 µm2) (Roselló et al. 2017). The presence of
phytoliths is an indicator of a presence of amorphous silica; therefore, these raw leaves can be
potential to use as a reinforcement or additive (ash) for cement composites (Roselló et al. 2017).
After the fibers were subjected to alkaline treatment, phytoliths were removed as shows Figure 2c
and f. Even more, in these images, surface of the treated fibers present greater defibrillation in
comparison to raw leaves surface (not considering the phytoliths), which agrees to the average
roughness obtained for both species in Figure 4. Based on this alkaline treatment duplicate, the
roughness of raw leaves surfaces and of which Stipa obtusa after the treatment obtained greater
defibrillation degree (RMS of 28.78 ± 6 nm), which also agrees to the SEM images (Figure 2). Hence,
Stipa obtusa leaves are very sensible to the alkaline treatment; therefore, lower NaOH concentration
JOURNAL OF NATURAL FIBERS 7

Figure 3. EDS result of Stipa obtusa and Jarava ichu.

(0.5 M) is enough to remove significant amount of noncellulosic components. RMS values reveal
that treated Stipa obtusa has similar values of RMS for the treated flax fibers (RMS = 32 ± 8 nm,
treated using 0.5 M of NaOH) (Balnois et al. 2007). These roughness were obtained by establishing
the scanned area at 10 μm2, based on the analysis at different areas (5, 10, 15 y 20 μm2), where at this
area (10 µm2) the roughness finishes to increase and begins to stabilize, as shown in Figure 4. This
behavior is due to the probability to find slits between the cell fiber, similar to flax fiber results
(Balnois et al. 2007).
Figure 5a–b show a cross section of the micrograph, where the images present a complex and
unique morphology. Sole macrofibers are composed of a large number of unitary cells with different
shapes, sizes, and thicknesses, each type has their specific function in the plant (Figure 6) (Evert,
8 S. MORI ET AL.

Figure 4. Interval plot of roughness (RMS) as a function of the scan area for the surfaces of raw leaves and alkaline treated for
Jarava ichu (JI) and Stipa obtusa (SO) fibers.

Figure 5. SEM cross section of raw leaf of Stipa obtusa (a), alkaline-treated Stipa obtusa (b), raw leaf of Jarava ichu (c), and alkaline-
treated Jarava ichu (d).

Esau, and Esau 2006). In detailed view, it is possible to note a subset of unitary cell form a pattern,
which is generally called “fiber bundle” according to Rudall et al. (Rudall and Kew 2007). On the
other hand, morphology of cells (fibers) are different in the epidermis, which according to Table 6,
the cell walls are thicker and their lumen are smaller, especially in Jarava ichu (2 µm2 and 2.2 µm,
respectively). These features allow to the plant avoid water loss in drought times (Rudall and Kew
JOURNAL OF NATURAL FIBERS 9

Figure 6. Vascular bundle of: (a) Stipa obtusa and (b) Jarava ichu.

Table 6. Morphology of ichu leaves in epidermis and cortex area.


Epidermis Cortex
Raw leaves Lumen area (µm2) Cell wall thickness (µm) Lumen area (µm2) Cell wall thickness (µm)
Stipa obtusa 20.9 ± 10.3 2.7 ± 1.3 28.7 ± 15.9 0.8 ± 0.2
Jarava ichu 2.2 ± 1.0 2.2 ± 0.5 37.2 ± 41.3 0.8 ± 0.2

2007). However, during the treatment process, this thick area makes difficult the entry of alkaline
solution into the internal structure; therefore, the effectiveness of the treatment is reduced. That is
the reason why Jarava ichu treated with 1.5 M NaOH is still maintaining a column of joined fibers;
since it is more difficult for the alkaline solution to enter through those compact walls (Figure 5b).
Contrarily, Stipa obtusa raw leaves after treatment suffer a break to form small fibers bundles and the
cellular structure suffer a collapse (reduction of lumen area), which leads to a better packing of
cellulose chains according to Vasquez et al. (Alvarez and Analía 2006).

Crystallinity index
The diffraction pattern of raw leaves and alkali-treated (Stipa obtusa and Jarava ichu) are shown in
Figure 7. These graphs confirm the two well-defined peaks at 2θ between 16.24– 16.48° and 21.50–
22.05°, which correspond to the (110) and (002) crystallographic planes, respectively (French 2014).
According to Segal method, the CI increased with the alkaline treatment, in ~7% and ~51% for Stipa
10 S. MORI ET AL.

Figure 7. X-ray diffraction patterns of untreated (black) and alkali-treated (red): (a) Stipa obtusa (SO) and (b) Jarava ichu (JI).

Table 7. Crystallinity characteristics of ichu fibers using Segal and Herman’s method for crystallinity index (CI) and Scherrer
equation for apparent crystallite size.
Fibers Condition Segal CI (%) Hermans CI (%) 2θ110 (°) L110 (nm) 2θ002 (°) L002 (nm)
Stipa obtusa Raw leaves 64.31 44.20 16.35 1.55 21.96 2.51
Alkaline treated 69.35 47.07 16.40 1.77 21.92 2.66
Jarava ichu Raw leaves 43.45 21.11 16.24 1.44 21.50 2.75
Alkaline treated 65.97 42.25 16.48 1.63 22.05 2.80

obtusa and Jarava ichu, respectively (Table 7). Analogous increment was observed using the
Herman’s method (~6%) for Stipa obtusa; however, pronounced increment was registered for
Jarava ichu (~100%). Similar behavior was observed on natural fibers from alfa, saharan aloe vera,
and hemp (Balaji and Nagarajan 2017; Benyahia et al. 2013; Zhang, Dhakal, and MacMullen 2015).
The increase of CI can be due to the removal of the noncellulosic components and the amorphous
cellulose as well (Poletto, Ornaghi, and Zattera 2014). Peak curve fitting (deconvolution) was
performed to the subtracted spectra using the Gaussian profile, results showed r2 over 0.97 in all
the cases. From these results, only two peaks were identified and their locations are detailed in Table
7. Apparent crystalline size was estimated and the results are shown in Table 7 as well. Results
showed an increment in the crystal size with the alkali treatment for (110) and (002) planes for Stipa
obtusa and Jarava ichu. This increment may be associated to the reduction of the noncrystalline
constituents (Zhang, Dhakal, and MacMullen 2015). Despite of this minimal increase, it can be
JOURNAL OF NATURAL FIBERS 11

related to a better alignment of cellulose chains after removal of pectin and waxes, according to
Poleto et al. (Pickering et al. 2007). However, intense or prolonged treatment can degrade the
crystalline part of cellulose (Zhang, Dhakal, and MacMullen 2015). Estimated crystal sizes for the
studied fibers are similar to the commercial natural fibers reported (Poletto et al. 2012) (Poletto,
Ornaghi, and Zattera 2014).

Thermogravimetry analysis
The measurement data obtained from thermogravimetric decomposition process are shown in
Figure 8 and summarized in Table 8. From the results, the curve decreases around 1.5– 2.5% (by
weight) between 40°C and 100°C, this can be associated to the dehydration process, where the water
contained into the fiber is removed. For raw leaves fibers, the initial degradation starts at 200°C for
Stipa obtusa and 217°C for Jarava ichu, due to the thermal depolymerization of hemicellulose and
the glycosidic linkages of cellulose mainly (Reddy et al. 2013). After treatment, this process cannot be
visualized in derived thermogravimetric (DTG) diagram, due to its low content of hemicellulose,
since alkaline treatment reduces it (Manimaran et al. 2018). Furthermore, alkaline treatment
improves cellulose thermal stability from 220 to 225°C, as shown in Figure 8 (Nicoleta, Ibbett, and
Schuster 2011). This can be due to the increment in the CI after alkaline treatment, since the
degradation takes place initially in the amorphous regions of the cellulose in agreement with Poleto

Figure 8. Thermogravimetric decomposition and its derivate of raw leaves (black) and alkaline-treated (red) ichu at 10°C/min
under nitrogen: (a) Stipa obtusa (SO) and (b) Jarava ichu (JI).
12 S. MORI ET AL.

Table 8. Characteristic of thermal decomposition of Jarava ichu and Stipa obtusa at 10°C/min under nitrogen environment.
Fiber Condition Transition temperature range (°C) Weight loss (%) Weight loss (%)
Stipa obtusa Raw leaves 40–145 90.44 1.42
200–315 289.41 18.00
315–400 351.68 48.98
Alkaline treated 40–160 98.77 2.28
225–400 346.69 50.96
Jarava ichu Raw leaves 40–148 91.76 2.08
217–300 282.49 18.68
300–400 337.98 42.45
Alkaline treated 40–150 94.98 1.33
220–400 349.39 46.74

et al. (Poletto et al. 2012). Finally, lignin degradation is more complex, due to the fact that its
degradation can start around 160°C and extend until 900°C at low decomposition rate, and thus is
not detectable with TGA and DTG analyses (Park et al. 2010).

Advancing contact angle


Figures 9 and 10 show the contact angles for the fibers as function of advancing velocity measured in
deionized water. Results show high degree of variability; however, the average values have a clear
trend in all the cases. Raw leaves in deionized water shows a hydrophobic behavior (θ > 90), in
contrast to conventional natural fibers that are hydrophilic (Cao et al. 2012; Gañán et al. 2004; Sever
et al. 2012; Tran et al. 2011), increasing even more their contact angle for higher speeds. Since
external surface of the raw leafs is in contact with environment, these develop an unique surface
(hydrophobic due to waxes and extractives); while fibers from sisal, flax, and hemp are inside the
pulping material or covered by the external shells, which is removed during the fiber extraction
processes (Ahmad et al. 2017). However, after alkali treatment Stipa obtusa becomes hydrophilic as
shown in Figure 9 at low speed (~51°), this may be due to the effectiveness to remove the
noncellulosic components from the raw leaves surfaces. Nevertheless, for high advancing velocities
their behavior is still hydrophobic. On the other hand, Jarava ichu performs an increment in their
advancing contact angles (Figure 10). Similar trends were observed when the fibers were evaluated in
ethylene glycol and glycerol liquids.

Figure 9. Advancing contact angle as function of velocity for Stipa obtusa (SO) measured in deionized water (raw leaves and
alkaline treated).
JOURNAL OF NATURAL FIBERS 13

Figure 10. Advancing contact angle as function of velocity for Jarava ichu (JI) measured in deionized water (raw leaves and
alkaline treated).

Table 9. Static contact angle obtained using fitting the MKT equation (Eq. 5) to the experimental
results (DW: Deionized water, EG: Ethylene glycol, G: Glycerol).
Static contact angle (θo, °)
Fiber Condition DW EG G
Jarava ichu Raw leaf 98.81 91.27 87.72
Alkaline treated 78.42 83.33 86.14
Stipa obtusa Raw leaf 91.94 89.69 94.71
Alkaline treated 48.66 33.28 78.57

Static contact angles obtained from fitting the MKT equation (Eq. 6) to the experimental data are
shown in Table 9, with r2 over 0.85. Estimated static contact angle decreases with the alkaline
treatment for the Stipa obtusa and for Jarava ichu. Furthermore, from the fitting, the average length
of each molecular displacement (λ) shows a consistent value from 0.86 to 2.05 nm for deionized
water, from 1.43 to 5.88 nm to ethylene glycol, and from 1.94 to 3.75 nm for glycerol, which are
similar values to that reported for PET and coir fibers (Blake 1993; Tran et al. 2011). However, the
equilibrium displacement frequency (Ko) shows a significant variation, which can go from 0.74 to
5.02 × 106 s−1. This result is not new; Blake showed that the viscosities of the fluid affect widely the
Ko (Blake 2006, 1993). Furthermore, as the cross-sectional shapes of the fiber are irregular (Figure 5),
the micropores can affect significantly the viscous bending phenomena, even though the viscosity of
the fluid is very low, causing an irregular contact line velocity between the fluid and the solid during
the immersion. Similar effect can cause the surface roughness, which is considerable as shown results
in Figure 4.
Applying Owens and Wendt method (Eq.6) and using the linear fit, the slope and the intercept to the y
axis were determined with an r2 > 0.80. Thereafter, the surface energies were determined (Table 10).
Since raw leaves fibers have high static contact angles (θo > 80°) in all cases, their surface energies are
lower (<18 mN/m for Stipa obtusa and Jarava ichu). This is mainly due to the presence of waxes and
other extractive components that makes the fibers hydrophobic and limiting to the water flow along the
raw leaves surfaces. However, after alkaline treatment the surface energy of Stipa obtusa fibers suffers a
significant increment (44.45 mN/m), being the polar component, which shows a significant increment,
as shown in Table 10. High polar component value was reported for flax fibers as well (Belgacem and
Gandini 2009). On the other hand, surface energy for Jarava ichu suffers a slight increment. Low surface
14 S. MORI ET AL.

Table 10. Surface energy (polar and dispersive) for Stipa obtusa and Jarava ichu.
Surface energy (γs, mN/m)
Fiber Condition Polar (γps) Dispersive (γds) γs
Stipa obtusa Raw leaves 2.31 15.42 17.73
Alkaline treated 11.14 11.73 22.87
Jarava ichu Raw leaves 4.57 12.74 17.32
Alkaline treated 32.26 12.19 44.45

energy of the fibers, not necessarily is an effective indicator to argue that the fibers are compatible to the
resins. Components like waxes and other extractives at the surface also have a key role, which should be
considered during the process of fibers selections.

Conclusions
Stipa obtusa and Jarava ichu were evaluated in their physical and thermal stability properties. At the
surface of raw leaves (Stipa obtusa and Jarava ichu) phytoliths were identified which could be very
attractive to the cement industries as fiber or pozzolanic supplementary material. With the alkaline
treatment, high degree of defibrillation of technical fibers was observed on Stipa obtusa, according to the
morphological analysis but not as well as for the Jarava ichu. On the other hand, thermal stability and
crystallinity index of these fibers achieves similar result to other commercial natural fibers. Furthermore,
crystalline index and crystal sizes increases with the alkaline treatment. Finally, static contact angle
decreases with the alkaline treatment for the Stipa obtusa and the Jarava ichu; however, surface energy
increases with the alkaline treatment for Stipa obtusa and Jarava ichu, with a particular increment on the
polar component for the Stipa obtusa. Therefore, Andean raw Stipa obtusa and .arava ichu could be
technically feasible material to use as reinforcement in composites materials.

Acknowledgments
This paper was written in the context of the project: “Estudio comparativo del potencial de fibras naturales endémicas
del Perú para su uso como refuerzo en materiales compuestos laminados” founded by the Consejo Nacional de
Ciencia, Tecnología e Innovación Tecnológica (CONCYTEC) under the contract number N° 117-2016-FONDECYT.
The authors of this paper appreciate the financial support from the Peruvian Government.

Funding
This work was supported by the Cienciactiva - FONDECYT, Perú [117-2016].

Notes on contributors
Sandra Mori Graduate at Industrial Engineering ta Universidad de Ingeniería y Tecnoligía – UTEC (2018), currently
develop projects oriented to industrial process optimization and maintenance planning in the ceramic industries.

Samuel Charca Graduate at Mechanical Engineering at Universidad Nacional de San Antonio Abad del Cusco
(UNSAAC, 2001), doctor degree at Civil Engineering from University of Puerto Rico at Mayaguez Campus (2009),
and post-doctoral fellowship at Aalborg University. Full professor of the Universidad de Ingeniería y Tecnología –
UTEC in the Mechanical Engineering Department, he is responsible of the experimental mechanics research group.

Elena Flores Graduate at Chemistry at Pontificia Universidad Católica del Perú (PUCP), Master degree at Chemistry
from PUCP. Full professor of the Universidad de Ingeniería y Tecnología – UTEC in the Chemistry Engineering
Department.

Holmer Savastano Jr. Doctor degree in Civil Construction at Universidade de Sao Paulo (Brazil). Full professor of the
Universidade de São Paulo, Biosystems Engineering Department, Pirassununga, SP, Brazil. Co-Editor in Chief of the
Brazilian Journal Ambiente Construído (Antac, 2010–2016). He is the coordinator of the Research Nucleus for Biosystems
JOURNAL OF NATURAL FIBERS 15

Materials at USP since Sep/2012. Member of the coordination committee at FAPESP in the Engineering area. More than
170 full papers published in peer reviewed journals. Co-PI of the INCT Project – Advanced eco-efficient technologies for
cement-based materials (aeCEM, since Dec/2016). Responsible for research projects in collaboration with industries for the
development of innovative construction components based on fiber-cement composites. Dr. Holmer has experience in
Materials and Components for Construction, working on the following subjects: vegetable fiber, animal welfare, rural
construction, civil construction and sustainability. (Source: Lattes Curriculum).

References
Abdul Khalil, H. P. S., A. F. Ireana Yusra, A. H. Bhat, and M. Jawaid. 2010. Cell wall ultrastructure, anatomy, lignin
distribution, and chemical composition of Malaysian cultivated kenaf fiber. Industrial Crops and Products 31
(1):113–21. doi:10.1016/j.indcrop.2009.09.008.
Adekomaya, O., T. Jamiru, R. Sadiku, and Z. Huan. 2017. Negative impact from the application of natural fibers.
Journal of Cleaner Production 143 (Supplement C):843–46. doi:10.1016/j.jclepro.2016.12.037.
Ahmad, T., H. S. Mahmood, Z. Ali, M. A. Khan, and S. Zia. 2017. Design and development of a portable sisal
decorticator. Pakistan Journal of Agricultural Research 30 (3):9. doi:10.17582/journal.pjar/2017.30.3.209.217.
Ahvenainen, P., I. Kontro, and S. Kirsi. 2016. Comparison of sample crystallinity determination methods by X-ray
diffraction for challenging cellulose I materials. Cellulose 23 (2):1073–86. doi:10.1007/s10570-016-0881-6.
Ali, A., K. Shaker, Y. Nawab, M. Jabbar, T. Hussain, J. Militky, and V. Baheti. 2016. Hydrophobic treatment of natural
fibers and their composites—A review. Journal of Industrial Textiles. 1528083716654468.
Alvarez, V. A., and V. Analía. 2006. Influence of fiber chemical modification procedure on the mechanical properties
and water absorption of MaterBi-Y/sisal fiber composites. Composites Part A: Applied Science and Manufacturing 37
(10):1672–80. doi:10.1016/j.compositesa.2005.10.005.
Bachtiar, D., S. M. Sapuan, and M. M. Hamdan. 2009. The influence of alkaline surface fibre treatment on the impact
properties of sugar palm fibre-reinforced epoxy composites. Polymer-Plastics Technology and Engineering 48
(4):379–83. doi:10.1080/03602550902725373.
Balaji, A. N., and K. J. Nagarajan. 2017. Characterization of alkali treated and untreated new cellulosic fiber from
Saharan aloe vera cactus leaves. Carbohydrate Polymers 174:200–08. doi:10.1016/j.carbpol.2017.06.065.
Balnois, E., F. Busnel, C. Baley, and Y. Grohens. 2007. An AFM study of the effect of chemical treatments on the
surface microstructure and adhesion properties of flax fibres. Composite Interfaces 14 (7–9):715–31. doi:10.1163/
156855407782106537.
Belgacem, M. N., and A. Gandini. 2009. Chapter 2: Natural fibre-surface modification and characterization. Natural
Fibre Reinforced Polymer Composites: from Macro to Nanoscale 14–46.
Benyahia, A., A. Merrouche, M. Rokbi, and Z. Kouadri. 2013. Study the effect of alkali treatment of natural fibers on
the mechanical behavior of the composite unsaturated Polyester-fiber Alfa. composites 2:3.
Bismarck, A., A. Baltazar-Y-Jimenez, and K. Sarikakis. 2006. Green composites as panacea? Socio-economic aspects of
green materials. Environment, Development and Sustainability 8 (3):445–63. doi:10.1007/s10668-005-8506-5.
Blake, T. D. 1993. Dynamic contact angles and wetting kinetics. Wettability 49:251.
Blake, T. D. 2006. The physics of moving wetting lines. Journal of Colloid and Interface Science 299 (1):1–13.
doi:10.1016/j.jcis.2006.03.051.
Boopathi, L., P. S. Sampath, and K. Mylsamy. 2012. Investigation of physical, chemical and mechanical properties of
raw and alkali treated borassus fruit fiber. Composites Part B: Engineering 43 (8):3044–52. doi:10.1016/j.
compositesb.2012.05.002.
Borchani, K., C. Carrot, and M. Jaziri. 2015. Untreated and alkali treated fibers from Alfa stem: Effect of alkali
treatment on structural, morphological and thermal features 22(3):1577–89.
Brown, D., P. V. Colin Polsky, S. D. Bolstad, D. H. Brody, R. Kroh, T. Loveland, and A. M. Thomson. 2014. Land use
and land cover change. In US global change research program, ed. J. M. Melillo, T. C. Richmond, and G. Q. Yohe,
318–32. Washington DC, United States(US): Pacific Northwest National Laboratory (PNNL), Richland, WA (US).
Cao, Y., F. Chan, Y.-H. Chui, and H. Xiao. 2012. Characterization of Flax fibres modified by Alkaline, Enzyme, and
Steam-heat treatments. 2012. Flax Fibres; Surface Treatments; Analysis; Properties7(3):4109–4121
Charca, S., J. Noel, D. Andia, J. Flores, A. Guzman, C. Renteros, and J. Tumialan. 2015. Assessment of ichu fibers as
non-expensive thermal insulation system for the Andean regions. Energy and Buildings 108:55–60. doi:10.1016/j.
enbuild.2015.08.053.
Cherian, B. M., A. L. Leão, S. F. de Souza, L. M. M. Costa, G. M. de Olyveira, M. Kottaisamy, E. R. Nagarajan, and
S. Thomas. 2011. Cellulose nanocomposites with nanofibres isolated from pineapple leaf fibers for medical
applications. Carbohydrate Polymers 86 (4):1790–98. doi:10.1016/j.carbpol.2011.07.009.
Cheung, H.-Y., H. Mei-po, K.-T. Lau, F. Cardona, and D. Hui. 2009. Natural fibre-reinforced composites for
bioengineering and environmental engineering applications. Composites Part B: Engineering 40 (7):655–63.
doi:10.1016/j.compositesb.2009.04.014.
16 S. MORI ET AL.

COMMISSION, EUROPEAN. 2000. Directive 2000-53/EC of the European parliament and of the council of september
18, 2000 on end-of-life vehicles. Spain: Union OJotE.
Das, M., and D. Chakraborty. 2008. Evaluation of improvement of physical and mechanical properties of bamboo
fibers due to alkali treatment. Journal of Applied Polymer Science 107 (1):522–27. doi:10.1002/(ISSN)1097-4628.
Das, M., and D. Chakraborty. 2009. Effects of alkalization and fiber loading on the mechanical properties and
morphology of bamboo fiber composites. II. Resol matrix. Journal of Applied Polymer Science 112 (1):447–53.
doi:10.1002/app.v112:1.
Doan, T.-T.-L., H. Brodowsky, and M. Edith. 2012. Jute fibre/epoxy composites: Surface properties and interfacial
adhesion. Composites Science and Technology 72 (10):1160–66. doi:10.1016/j.compscitech.2012.03.025.
El, O. A., Y. Chaabouni, S. Msahli, and F. Sakli. 2011. Crystal transition from cellulose I to cellulose II in NaOH treated
Agave americana L. fibre. Carbohydrate Polymers 86 (3):1221–29. doi:10.1016/j.carbpol.2011.06.037.
Evert, R. F., K. Esau, and K. Esau. 2006. Esau’s Plant anatomy: Meristems, cells, and tissues of the plant body: Their
structure, function, and development. New Jersey: John Wiley & Sons, Inc.
Ferris, R. 2016. China auto regulatory trends 2016: New-energy, materials restrictions, recalls and emissions figure
prominently. The National Law.
French, A. D. 2014. Idealized powder diffraction patterns for cellulose polymorphs. Cellulose 21 (2):885–96.
doi:10.1007/s10570-013-0030-4.
Fuentes, C. A., I. Verpoest, and A. W. Van Vuure. 2015. Chapter 4: Interfacial compatibility and adhesion in natural
fiber composites. In Natural fiber composites, 139–68. CRC Press.
Gañán, P., J. Cruz, S. Garbizu, A. Arbelaiz, and I. Mondragon. 2004. Stem and bunch banana fibers from cultivation
wastes: Effect of treatments on physico-chemical behavior. Journal of Applied Polymer Science 94 (4):1489–95.
doi:10.1002/app.v94:4.
George, J. M., S. Sreekala, and S. Thomas. 2001. A review on interface modification and characterization of natural
fiber reinforced plastic composites. Polymer Engineering & Science 41 (9):1471–85. doi:10.1002/pen.10846.
Heng, J. Y. Y., F. P. Duncan, F. Thielmann, T. Lampke, and A. Bismarck. 2007. Methods to determine surface energies
of natural fibres: A review. Composite Interfaces 14 (7–9):581–604. doi:10.1163/156855407782106492.
John, M. J., and R. D. Anandjiwala. 2008. Recent developments in chemical modification and characterization of
natural fiber-reinforced composites. Polymer Composites 29 (2):187–207. doi:10.1002/pc.20461.
Kevin, H., and B. John. 2007. Dynamic wettability properties of single wood pulp fibers and their relationship to
absorbency. Wood and Fiber Science 20 (1):15.
Kumar, R., V. Choudhary, S. Mishra, and I. Varma. 2008. Banana fiber-reinforced biodegradable soy protein
composites. Frontiers of Chemistry in China 3 (3):243–50. doi:10.1007/s11458-008-0069-1.
Le Troedec, M., D. Sedan, C. Peyratout, J. P. Bonnet, A. Smith, R. Guinebretiere, V. Gloaguen, and P. Krausz. 2008.
Influence of various chemical treatments on the composition and structure of hemp fibres. Composites Part A:
Applied Science and Manufacturing 39 (3):514–22. doi:10.1016/j.compositesa.2007.12.001.
Liu, Y., and H. Hong. 2008. X-ray diffraction study of bamboo fibers treated with NaOH. Fibers and Polymers 9
(6):735–39. doi:10.1007/s12221-008-0115-0.
Lu, T., M. Jiang, Z. Jiang, D. Hui, Z. Wang, and Z. Zhou. 2013. Effect of surface modification of bamboo cellulose
fibers on mechanical properties of cellulose/epoxy composites. Composites Part B: Engineering 51:28–34.
doi:10.1016/j.compositesb.2013.02.031.
Mahesh, T. N. 2014. Influence of surface treatments on glass fibers and their wetting behavior. In Interfacial
phenomena in glass fiber composites., ed. North Carolina State University, 65–101. EEUU. PhD Thesis, North
Carolina.
Manimaran, P., P. Senthamaraikannan, M. R. Sanjay, M. K. Marichelvam, and M. Jawaid. 2018. Study on character-
ization of Furcraea foetida new natural fiber as composite reinforcement for lightweight applications. Carbohydrate
Polymers 181:650–58. doi:10.1016/j.carbpol.2017.11.099.
Meng, F.-D., Y. Yang-lun, Y.-M. Zhang, Y. Wen-ji, and J.-M. Gao. 2016. Surface chemical composition analysis of
heat-treated bamboo. Applied Surface Science 371:383–90. doi:10.1016/j.apsusc.2016.03.015.
Monteiro, S. N., V. Calado, R. Rubén Jesus S, and F. M. Margem. 2012. Thermogravimetric behavior of natural fibers
reinforced polymer composites—An overview. Materials Science and Engineering: A 557:17–28. doi:10.1016/j.
msea.2012.05.109.
Mori, S., C. Tenazoa, S. Candiotti, E. Flores, and S. Charca. 2018. Assessment of ichu fibers extraction and their use as
reinforcement in composite materials. Journal of Natural Fibers 1–16. doi:10.1080/15440478.2018.1527271.
Mori, S., E. Flores, and S. Charca. 2017. Ichu: New natural fibers for composites and its extraction methodology. 17th
International Conference on Non-Convetional Materials and Technologies Merida, Mexico.
Nicoleta, T., R. Ibbett, and K. C. Schuster. 2011. Overview on native cellulose and microcrystalline cellulose
I structure studied by X-ray diffraction (WAXD): Comparison between measurement techniques. Lenzinger
Berichte 89:14.
Norul Izani, M. A., M. T. Paridah, U. M. K. Anwar, M. Y. Mohd Nor, and P. S. H’ng. 2013. Effects of fiber treatment
on morphology, tensile and thermogravimetric analysis of oil palm empty fruit bunches fibers. Composites Part B:
Engineering 45 (1):1251–57. doi:10.1016/j.compositesb.2012.07.027.
JOURNAL OF NATURAL FIBERS 17

Owens, D. K., and R. C. Wendt. 1969. Estimation of the surface free energy of polymers. Journal of Applied Polymer
Science 13 (8):1741–47. doi:10.1002/app.1969.070130815.
Park, S., J. O. Baker, M. E. Himmel, P. A. Parilla, and D. K. Johnson. 2010. Cellulose crystallinity index: Measurement
techniques and their impact on interpreting cellulase performance. Biotechnology for Biofuels 3 (1):10. doi:10.1186/
1754-6834-3-20.
Pickering, K. L., G. W. Beckermann, S. N. Alam, and N. J. Foreman. 2007. Optimising industrial hemp fibre for
composites. Composites Part A: Applied Science and Manufacturing 38 (2):461–68. doi:10.1016/j.
compositesa.2006.02.020.
Poletto, M., A. J. Zattera, M. M. C. Forte, and R. M. C. Santana. 2012. Thermal decomposition of wood: Influence of
wood components and cellulose crystallite size. Bioresource Technology 109:148–53. doi:10.1016/j.
biortech.2011.11.122.
Poletto, M., H. Ornaghi, and A. Zattera. 2014. Native cellulose: Structure, characterization and thermal properties.
Materials 7 (9):6105. doi:10.3390/ma7096105.
Popescu, M.-C., C.-M. Popescu, G. Lisa, and Y. Sakata. 2011. Evaluation of morphological and chemical aspects of
different wood species by spectroscopy and thermal methods. Journal of Molecular Structure 988 (1–3):65–72.
doi:10.1016/j.molstruc.2010.12.004.
Pucci, M. F., M. C. Seghini, P.-J. Liotier, F. Sarasini, J. Tirilló, and S. Drapier. 2017. Surface characterisation and
wetting properties of single basalt fibres. Composites Part B: Engineering 109:72–81. doi:10.1016/j.
compositesb.2016.09.065.
Reddy, K., K. Obi, R. N. Reddy, J. Zhang, J. Zhang, and A. V. Rajulu. 2013. Effect of alkali treatment on the properties
of century fiber. Journal of Natural Fibers 10 (3):282–96. doi:10.1080/15440478.2013.800812.
Roselló, J., M. Lourdes Soriano, J. L. Pilar Santamarina, J. M. Akasaki, and P. Jordi. 2017. Rice straw ash: A potential
pozzolanic supplementary material for cementing systems. Industrial Crops and Products 103:39–50. doi:10.1016/j.
indcrop.2017.03.030.
Rudall, P. J., and R. B. G. Kew. 2007. Anatomy of flowering plants an introduction to structure and development. 3rd ed.
Cambridge UK: Cambridge University Press.
Saha, P., S. Manna, S. R. Chowdhury, R. Sen, D. Roy, and B. Adhikari. 2010. Enhancement of tensile strength of
lignocellulosic jute fibers by alkali-steam treatment. Bioresource Technology 101 (9):3182–87. doi:10.1016/j.
biortech.2009.12.010.
Sanjay, M., G. Arpitha, L. Naik, K. Gopalakrishna, and B. Yogesha. 2016. Applications of natural fibers and its
composites: An overview. Natural Resources 7 (3):6. doi:10.4236/nr.2016.73011.
Savastano, H., Jr, S. F. Santos, J. Fiorelli, and V. Agopyan. 2016. 19 - Sustainable use of vegetable fibres and particles in
civil construction A2 - Khatib. In Sustainability of Construction Materials, ed. Jamal M, 477–520. 2nd ed. Boston,
MA: Woodhead Publishing.
Sawpan, M. A., K. L. Pickering, and A. Fernyhough. 2011. Effect of various chemical treatments on the fibre structure
and tensile properties of industrial hemp fibres. Composites Part A: Applied Science and Manufacturing 42
(8):888–95. doi:10.1016/j.compositesa.2011.03.008.
Segal, L., J. J. Creely, A. E. Martin, and C. M. Conrad. 1959. An empirical method for estimating the degree of
crystallinity of native cellulose using the X-ray diffractometer. Textile Research Journal 29 (10):786–94. doi:10.1177/
004051755902901003.
Sever, K., M. Sarikanat, Y. Seki, G. Erkan, Ü. H. Erdoğan, and S. Erden. 2012. Surface treatments of jute fabric: The
influence of surface characteristics on jute fabrics and mechanical properties of jute/polyester composites. Industrial
Crops and Products 35 (1):22–30. doi:10.1016/j.indcrop.2011.05.020.
Son, H., C. Kim, W. K. Chong, and J.-S. Chou. 2011. Implementing sustainable development in the construction
industry: Constructors’ perspectives in the US and Korea. Sustainable Development 19 (5):337–47. doi:10.1002/sd.442.
Tenazoa, C., S. Charca, M. Quintana, and E. Flores. 2017. Chemical characterization for the comparative study of
peruvian natural fibers. 17th International Conference on Non-Convetional Materials and Technologies Merida,
Mexico.
Tran, L. Q. N., C. A. Fuentes, C. Dupont-Gillain, A. W. Van Vuure, and I. Verpoest. 2011. Wetting analysis and
surface characterisation of coir fibres used as reinforcement for composites. Colloids and Surfaces A:
Physicochemical and Engineering Aspects 377 (1):251–60. doi:10.1016/j.colsurfa.2011.01.023.
Tran, L. Q. N., C. A. Fuentes, C. Dupont-Gillain, A. W. Van Vuure, and I. Verpoest. 2013. Understanding the
interfacial compatibility and adhesion of natural coir fibre thermoplastic composites. Composites Science and
Technology 80:23–30. doi:10.1016/j.compscitech.2013.03.004.
Yan, L., N. Chouw, L. Huang, and B. Kasal. 2016. Effect of alkali treatment on microstructure and mechanical
properties of coir fibres, coir fibre reinforced-polymer composites and reinforced-cementitious composites.
Construction and Building Materials 112:168–82. doi:10.1016/j.conbuildmat.2016.02.182.
Yanjun, X. 2010. Silane coupling agents used for natural fiber/polymer composites: A review. Composites Part A:
Applied Science and Manufacturing 41 (7):14.
Zhang, H. N., J. Dhakal, and Z. Y. MacMullen. 2015. Capter 6: Testing and characterization of natural fiber-reinforced
composites in natural. In Natural fiber composites, ed. R. D. S. G. Campilho, 175–98.

You might also like