You are on page 1of 19

materials

Article
Chemical Treatment of Waste Abaca for Natural
Fiber-Reinforced Geopolymer Composite
Roy Alvin J. Malenab 1 , Janne Pauline S. Ngo 1 and Michael Angelo B. Promentilla 1,2, *
1 Chemical Engineering Department, De La Salle University, Manila 0922, Philippines;
roy.malenab@dlsu.edu.ph (R.A.J.M.); janne_ngo@dlsu.edu.ph (J.P.S.N.)
2 National Research Council of the Philippines, Taguig, Metro Manila 1631, Philippines
* Correspondence: michael.promentilla@dlsu.edu.ph; Tel.: +63-2-536-0223

Academic Editor: Nicole Zander


Received: 29 March 2017; Accepted: 23 May 2017; Published: 25 May 2017

Abstract: The use of natural fibers in reinforced composites to produce eco-friendly materials is
gaining more attention due to their attractive features such as low cost, low density and good
mechanical properties, among others. This work thus investigates the potential of waste abaca (Manila
hemp) fiber as reinforcing agent in an inorganic aluminosilicate material known as geopolymer.
In this study, the waste fibers were subjected to different chemical treatments to modify the surface
characteristics and to improve the adhesion with the fly ash-based geopolymer matrix. Definitive
screening design of experiment was used to investigate the effect of successive chemical treatment of
the fiber on its tensile strength considering the following factors: (1) NaOH pretreatment; (2) soaking
time in aluminum salt solution; and (3) final pH of the slurry. The results show that the abaca fiber
without alkali pretreatment, soaked for 12 h in Al2 (SO4 )3 solution and adjusted to pH 6 exhibited the
highest tensile strength among the treated fibers. Test results confirmed that the chemical treatment
removes the lignin, pectin and hemicellulose, as well as makes the surface rougher with the deposition
of aluminum compounds. This improves the interfacial bonding between geopolymer matrix and the
abaca fiber, while the geopolymer protects the treated fiber from thermal degradation.

Keywords: waste utilization; abaca (Manila hemp) fiber; chemical treatment; fiber-reinforced
composite; geopolymer; tensile strength; definitive screening design of experiment

1. Introduction
The growing concerns for the environment and sustainability have served as a strong drive
for researches on developing eco-friendly composite materials [1]. Composite materials in general
are formed by combining two or more constituent materials, a continuous medium called matrix
and the dispersed phase/s, either fiber/s or particulate/s, in order to develop another material
with desired combination of properties [2]. In fiber-reinforced composite (FRC) materials, such
matrix is reinforced by fibers for their superior properties and load transfer characteristics. Recently,
natural fibers such as cotton, flax, jute, sisal and hemp are also becoming more attractive than
synthetic fibers because of their natural abundance, low cost, low density, good mechanical properties,
nontoxicity, etc. [1–4]. For example, abaca fibers have been widely studied as a reinforcement in
cement [5], polypropylene [6,7] and epoxy [8] matrices, among others.
Abaca fiber or Manila hemp, known for its commendable mechanical strength, has been produced
and widely exported by the Philippines. Commercial grade abaca has density 1.5 g/cm3 and tensile
strength of about 980 N/mm2 [9]. In 2015, the annual abaca fiber production was 70,400 metric
tons supplying around 87% of world requirements [10]. These fibers are extracted from a native
banana species and are harvested mostly for their use in making ropes, in textile, and more recently,

Materials 2017, 10, 579; doi:10.3390/ma10060579 www.mdpi.com/journal/materials


Materials 2017, 10, 579 2 of 19

in automotive [11]. The extraction processes from plant to usable fibers produce considerable amount
of waste or scrap fibers, which are still underutilized [12].
This study thus explores the utilization of scrap abaca fiber as reinforcement for fly ash-based
geopolymer matrix. Geopolymer is an inorganic polymer formed from the reaction of materials rich in
reactive alumina and silica with an alkaline solution [13,14]. This material has also attracted widespread
scientific and industrial interest for its potential to valorize waste or industrial by-product such as coal
ash while producing cementitious material with performance comparable to that of Portland cement
in many applications [15]. Furthermore, it has several other advantages such as superior heat and fire
resistance, high chemical resistance, and lower carbon footprint and embodied energy. Accordingly,
fiber-reinforced geopolymer composite could potentially provide better mechanical and thermal
properties over a wide temperature range as compared to Portland cement or organic polymer-based
matrix. The use of natural fibers such as bamboo, cornhusk, wool, cotton, or flax in geopolymer
matrix has been mentioned [16–18] and shown with promising results in the recent critical review
on fiber-reinforced geopolymers [19,20]. Another study also demonstrates the surface modification
of such fiber which improves its alkali resistance and results to an improved flexural strength of a
reinforced metakaolin-based geopolymer as compared to that of pristine geopolymer [21]. In contrast,
one recent exploratory study [22] reported no significant differences in the flexural strength between
that of pristine fly ash based-geopolymer and that of geopolymer reinforced with the untreated coconut
coir, cotton, or sisal fibers. This suggests that surface modification of the natural fibers is essential to
improve the performance of the reinforced geopolymer composites.
To our knowledge, no studies have been reported yet on a chemical treatment of abaca fiber
used for reinforcing fly ash-based geopolymer. Surface modification of abaca fibers could improve
the interfacial adhesion with such geopolymer matrix. In general, the matrix and fibers retain their
physical and chemical identities in FRCs, forming boundaries or interfaces [23]. Effectiveness of fiber
reinforcement in composites depends on the degree or mechanism of bonding at the fiber–matrix
interface and the ability to transfer stress from matrix to the fiber [24]. Stress transfer among the
components should be effective in order to attain desired strength. The developed interface between
the fiber and the matrix is thus a crucial factor of the composite’s static and dynamic stability [2].
This interface enables load transfer between the components. When the fiber is brought in contact
with the matrix, the chemical groups on the fiber surface can bond with those present on the matrix.
Electrostatic, hydrogen bonds and dipole-dipole interaction can develop depending on the nature and
compatibility of the fiber and matrix. This is the basis for the fiber’s adhesion to the matrix. Surface
treatment, which can modify or introduce chemical groups that could aid in bonding, is often applied
to render the fiber compatible with the matrix. Meanwhile, the presence of hydroxyl and carboxylic
groups in natural cellulosic fibers’ structure make them intrinsically hydrophilic and polar. When such
fibers are used as reinforcement in hydrophobic matrices, like polymers, the resulting composite is a
heterogeneous system that could have inferior properties due to lack of adhesion and chemical affinity.
Some of the organic components of natural fibers can inhibit its compatibility with the targeted
organic or inorganic matrix. Since the interface quality and bondage between matrix and fiber is
insufficient, surface modification of these fibers through chemical treatment and other techniques have
been pursued in research for years such as alkali, chemical coupling, oxidation, plasma, ultrasound,
and enzyme treatments [23,25]. The most common treatment for natural fibers discussed in literature
is alkali treatment, where the fiber is soaked in sodium hydroxide (NaOH) solution in order to
strip off some of the unwanted waxes, hemicellulose and lignin. Due to the hydrophilic nature of
some of the fiber components and thus, tendency for water absorption, natural fibers may exhibit
weak interfacial bonding with hydrophobic matrices [26,27]. Furthermore, “swelling” may occur
when a hydrophilic fiber within a hydrophobic matrix absorbs moisture within the matrix, making it
susceptible to weakening and deterioration [28]. This may occur in the event that water absorption
due to the fiber breaks hydrogen bonds between the fiber and matrix or among the fibers themselves.
In addition, alkali treatment induces rougher surfaces for better adhesion between the fiber and matrix.
Materials 2017, 10, 579 3 of 19

fiber and
Materials matrix.
2017, 10, 579 Prior
studies have also shown that alkali treatment of these fibers have improved3 of 19
the performance of reinforced composites in terms of thermal stability [29], water absorption and
tensile strength [30,31].
Prior studies have also shown that alkali treatment of these fibers have improved the performance of
Degradation of such natural fibers when used in cementitious or geopolymer matrix due to the
reinforced composites in terms of thermal stability [29], water absorption and tensile strength [30,31].
harsh environment is another concern. The decay of plant fibers when used as reinforcement in
Degradation of such natural fibers when used in cementitious or geopolymer matrix due to the
alkaline matrices such as that of concrete has been observed [32]. Therefore, plant-based fibers must
harsh environment is another concern. The decay of plant fibers when used as reinforcement in alkaline
be treated not only to be compatible to its matrix but also to avoid or delay its degradation ensuring
matrices such as that of concrete has been observed [32]. Therefore, plant-based fibers must be treated
that its reinforcing function is sustainable. Aluminum sulfate treatment of natural fibers is a potential
not only to be compatible to its matrix but also to avoid or delay its degradation ensuring that its
technique to address the possible deterioration of plant-based fibers exposed to cementitious matrix.
reinforcing function is sustainable. Aluminum sulfate treatment of natural fibers is a potential technique
This technique has been employed for pulped wood fibers to reinforce cementitious matrices [33].
to address the possible deterioration of plant-based fibers exposed to cementitious matrix. This technique
The treatment involves deposition of insoluble aluminum hydroxide on the fiber surface via
has been employed for pulped wood fibers to reinforce cementitious matrices [33]. The treatment
neutralization process. These ionic deposits are expected to have better affinity with the inorganic or
involves deposition of insoluble aluminum hydroxide on the fiber surface via neutralization process.
cementitious matrix when compared to the affinity of organic, hydrophilic nature of cellulosic fibers
These ionic deposits are expected to have better affinity with the inorganic or cementitious matrix
to such matrices.
when compared to the affinity of organic, hydrophilic nature of cellulosic fibers to such matrices.
In the present work, we investigate the effect of successive chemical treatment via NaOH
In the present work, we investigate the effect of successive chemical treatment via NaOH
pretreatment followed by an aluminum sulfate treatment of waste abaca fibers on their tensile
pretreatment followed by an aluminum sulfate treatment of waste abaca fibers on their tensile strength
strength and morphology. Fourier transform infrared spectroscopy (FTIR), scanning electron
and morphology. Fourier transform infrared spectroscopy (FTIR), scanning electron microscopy with
microscopy with energy-dispersive X-ray spectroscopy (SEM-EDX), X-ray diffraction (XRD), and
energy-dispersive X-ray spectroscopy (SEM-EDX), X-ray diffraction (XRD), and thermo-gravimetric
thermo-gravimetric analysis (TGA) were used to analyze the structural and chemical changes in the
analysis (TGA) were used to analyze the structural and chemical changes in the fibers after chemical
fibers after chemical treatment. Preliminary characterization of the fiber-reinforced geopolymer using
treatment. Preliminary characterization of the fiber-reinforced geopolymer using SEM-EDX and TGA
SEM-EDX and TGA combined with mechanical tests provides an impetus for future work on the
combined with mechanical tests provides an impetus for future work on the development of chemically
development of chemically treated abaca fiber-reinforced geopolymer composites.
treated abaca fiber-reinforced geopolymer composites.
2. Materials
2. Materials and
and Method
Method

2.1. Sources
2.1. Sources of
of Raw
Raw Materials
Materials
Abaca fibers
Abaca are extracted
fibers are extracted by by hand
hand stripping
stripping and
and are
are subsequently
subsequently segregated
segregated into
into standard
standard
grades based on color and texture. Residual grade abaca fibers shown in Figure
grades based on color and texture. Residual grade abaca fibers shown in Figure 1a were collected 1a were collected
from
from Dumaguete
Dumaguete City, Philippines.
City, Philippines. The average The diameter
average diameter of theabaca
of the collected collected
strandsabaca strands
is 162 is 162
micrometers.
micrometers.
The fibers were The
cutfibers were cut into 25
into approximately approximately
cm long and 25 cm long to
segregated and
100segregated
strands pertosample
100 strands
as shown per
sample
in Figure as1b.
shown
The inflyFigure
ash used 1b. to
The fly ashgeopolymer
produce used to produce geopolymer
matrix was obtained matrix
fromwas obtained
a coal fired from
power a
coal fired
plant power
situated plant
at the situated
central at the
region central Philippines.
of Luzon, region of Luzon,
X-rayPhilippines.
fluorescenceX-ray
(XRF,fluorescence (XRF,
Thermo Scientific,
Thermo Scientific,
Waltham, MA, USA) Waltham,
analysis ofMA, theUSA)
fly ashanalysis
used is of the fly ashinused
summarized Tableis1.summarized
Chemicals used in Table 1.
to treat
Chemicals
abaca used
fibers, to treat
sodium abaca fibers,
hydroxide sodium hydroxide
micropearls (99 wt %)micropearls
and aluminum (99 wt %) and
sulfate aluminum
powders (99 sulfate
wt %),
powders (99 wt %), and the water glass solution (modulus = 2.5; 34.1 wt % SiO , 14.7
and the water glass solution (modulus = 2.5; 34.1 wt % SiO2 , 14.7 wt % Na2 O) for geopolymer synthesis
2 wt % Na 2 O) for
geopolymer synthesis were supplied by a local company (Just-in-One
were supplied by a local company (Just-in-One Marketing) and were used as received. Marketing) and were used as
received.
Table 1. X-ray fluorescence (XRF) analysis of fly ash.
Table 1. X-ray fluorescence (XRF) analysis of fly ash.
Oxide SiO2 Fe2 O3 CaO Al2 O3 MgO K2 O Others
Oxide SiO2 Fe2O3 CaO Al2O3 MgO K 2O Others
Mass % 33.9
Mass % 33.9 26.5 26.5 13.6
13.6 13.5
13.5 7.9
7.9 1.2 1.2 3.4 3.4

Figure 1.
Figure 1. Waste
Waste abaca
abaca from
from Dumaguete,
Dumaguete, Philippines:
Philippines: (a)
(a) as
as received;
received; and
and (b)
(b) segregated
segregated into
into 100
100 strands.
strands.
Materials 2017, 10, 579 4 of 19

2.2. Chemical Treatment of Waste Abaca Fiber


Two types of chemical treatments of abaca fibers were used in this study: (1) alkali pretreatment;
and (2) aluminum sulfate treatment, which involves deposition of aluminum hydroxide on the fiber
surface via precipitation. Using SAS JMP 11.1.1 software, 10 experimental runs were prescribed
using Definitive Screening Design (DSD). DSD works for factor screening containing a combination of
continuous and two-level categorical factors. It can estimate unbiased main effects that are completely
independent of two-factor interactions and avoid confounding of any pair of quadratic effects [34,35].
Table 2 shows the factors, their levels and corresponding codes used in this design of experiment.
There are three factors considered, one is a categorical factor (NaOH pretreatment) and the other two
(soaking time and pH) are continuous.
For the alkali pretreatment, fibers were soaked in 6% by weight NaOH solution for 48 h in a 500 mL
beaker with constant agitation at room temperature. The said treatment can remove impurities of the
fiber resulting to rough fiber surface [36,37]. Fibers were then removed from the NaOH solution and
washed with distilled water several times. Alkali pretreated fibers were then air dried for 24 h. For the
Al2 (SO4 )3 treatment, untreated and alkali pretreated fibers were soaked in 10% by weight of Al2 (SO4 )3
solution in different beakers. The initial pH of Al2 (SO4 )3 solution is 3.5. Final pH of the abaca fiber and
aqueous Al2 (SO4 )3 mixture were varied based on the solubility and precipitation of Al(OH)x species
as function of pH [38]. Soaking period was also varied to three levels, 30 min, 6.25 h and 12 h, to allow
the penetration of ions into the fiber. Then, the pH was adjusted to a desired level (4.5, 6.0 or 7.5) by
gradually adding 2 M NaOH solution. Subsequently, the fibers were left soaked in the mixture for
another 24 h to allow further precipitation of Al(OH)3 on the fiber surface. Samples were then drawn
from the mixture, washed with distilled water and air-dried for at least 24 h before being stored in
sealed plastics. Figure 2 illustrates the flow of the chemical treatment procedure and characterization
performed for the treated fiber.

Table 2. Factors and levels used for the chemical treatment.

Factor/Level (Coded) −1 0 +1
NaOH pretreatment - No Yes
Soaking time in Al2 (SO4 )3 , hrs 0.5 6.25 12
Final pH 4.5 6.0 7.5

2.3. Preparation of Abaca Fiber-Reinforced Geopolymer


Fly ash and abaca fibers (99:1 by weight) were dry-mixed to ensure uniform fiber distribution.
For every 1 kg of fly ash–fiber mixture, 400 g of alkaline activator (well-mixed 80% water glass solution,
20% 12 M NaOH by weight) was added gradually while mixing using a laboratory motorized mixer.
The mass ratio was chosen to achieve a geopolymer mix with considerable compressive strength and
good workability [39]. Mixing was continued for about 15 min until the mixture appeared homogenous.
About 30 mL of distilled water was added to attain the desired consistency. The mixture was then
poured in 5 cm × 5 cm × 5 cm cube molds (for compressive strength test) and 5 cm × 5 cm × 18 cm
molds (for flexural strength test), and placed on a shaker for 1 min to remove trapped bubbles. After
24 h, samples were demolded and cured in the oven at 75 ◦ C for 24 h and then cured at ambient
conditions for 28 days. A specimen of pristine geopolymer, i.e., with no abaca fiber reinforcement
was also prepared to serve as the reference material. The specimens were then tested for compressive
strength and flexural strength.
Materials 2017, 10, 579 5 of 19
Materials 2017, 10, 579 5 of 19

Figure
Figure 2. Chemicaltreatment
2. Chemical treatmentand
and characterization
characterization ofof
waste abaca
waste fiber.
abaca fiber.

2.4. Tensile Strength Test of Abaca Fibers


2.4. Tensile Strength Test of Abaca Fibers
Average tensile strengths of the abaca fibers (untreated and treated) were determined by taking
Average
five strandstensile
from strengths
each bundleof of
the
100abaca fibers
strands (untreated
randomly. Eachandstrandtreated)
was cut were
to 5 determined
cm then both by taking
ends
five strands from each
were mounted bundle of
to cardboard (2 100
cm ×strands randomly.
2 cm) using quick-dry Each strand was cut to
cyanoacrylate-based 5 cm then
adhesive. both
Then, the ends
were prepared
mountedspecimen
to cardboard was (2 cm by
held × 2thecm)universal
using quick-dry cyanoacrylate-based
testing machine (Instron modeladhesive. Then, the
3324, Instron,
Norwood,
prepared specimen MA, wasUSA)held
withbygauge length of 3testing
the universal cm. Themachine
cardboard and themodel
(Instron hardened
3324,adhesive layer
Instron, in
Norwood,
contact with the fiber protect the fiber from possible deformation due to
MA, USA) with gauge length of 3 cm. The cardboard and the hardened adhesive layer in contact with clamping pressure. The
universal
the fiber protect testing machine
the fiber from(UTM) applies
possible tensile loaddue
deformation on the fiber strand
to clamping equivalent
pressure. Theto universal
an extension
testing
rate of 5 mm per min and recording the stress until it breaks. Prior to fracture, each strand was placed
machine (UTM) applies tensile load on the fiber strand equivalent to an extension rate of 5 mm per
under an optical microscope (Amscope microscope, AmScope, CA, USA) linked to a computer to
min and recording the stress until it breaks. Prior to fracture, each strand was placed under an optical
measure each of their diameters. Fiber strands were assumed to have approximately circular cross-
microscope
sectional(Amscope
areas andmicroscope,
were computed AmScope, CA, USA)
accordingly. linked
Tensile to a computer
strength to measure
of the fiber is definedeach of their
as the
diameters.
maximum Fiberstress
strands were
it can assumed
withstand, andtoishave approximately
computed by dividing circular
the loadcross-sectional
at failure by theareas and were
original
computed accordingly. Tensile strength of the fiber is defined as the maximum stress
cross-sectional area per fiber. The five readings were averaged to obtain the final value recorded. it can withstand,
and is computed by dividing the load at failure by the original cross-sectional area per fiber. The five
readings were averaged to obtain the final value recorded.
Materials 2017,
Materials 2017, 10,
10, 579
579 66 of
of 19
19

2.5. Scanning Electron Microscopy—Energy Dispersive Spectroscopy (SEM-EDS) Measurements


2.5. Scanning Electron Microscopy—Energy Dispersive Spectroscopy (SEM-EDS) Measurements
Surface morphology and elemental microanalyses of samples were performed using Field-emission
scanning electron
Surface microscope
morphology and (FESEM Dual microanalyses
elemental Beam Helios Nanolab 600i, FEI,
of samples were Hillsboro,
performed OR, using
USA)
equipped with electron-dispersive
Field-emission X-ray spectroscopy
scanning electron microscope (FESEM (EDS, FEI, Hillsboro,
Dual Beam OR, USA).
Helios Nanolab 600i,The accelerating
FEI, Hillsboro,
voltage
OR, USA)andequipped
beam current with for SEM analysis are X-ray
electron-dispersive 2.0 kVspectroscopy
and 43 pA, respectively. For EDS analysis,
(EDS, FEI, Hillsboro, OR, USA).the
accelerating
The voltage
accelerating and beam
voltage currentcurrent
and beam are 15.0forkVSEM
and analysis
0.69 nA, respectively.
are 2.0 kV and 43 pA, respectively.
For EDS analysis, the accelerating voltage and beam current are 15.0 kV and 0.69 nA, respectively.
2.6. X-ray Diffraction (XRD) Measurements
2.6. X-ray Diffraction (XRD) Measurements
XRD (Maxima XRD-7000 Shimadzu, Tokyo, Japan) analyses were carried out using X-ray beams
generated (Maxima
XRD from a Cu XRD-7000 Shimadzu,
Kα radiation source Tokyo, Japan)
(40 kV; analyses
30 mA) withwere carried out
wavelength usingÅX-ray
of 1.54 to scanbeams
the
generated from a Cu Kα radiation source (40 kV; 30 mA) with wavelength of
samples from 3.00° to 70.00° 2-theta angles. The resolution for this analysis was set at 0.020, with scan 1.54 Å to scan the
samples from 3.00◦atto2°70.00
speed maintained
◦ 2-theta angles. The resolution for this analysis was set at 0.020, with
per min. The fiber’s crystallinity index (CI) was calculated by XRD peak
scan speed maintained at 2 ◦ per min. The fiber’s crystallinity index (CI) was calculated by XRD
deconvolution method [40]. In this method, a curve-fitting program (e.g., PeakFit v4.12) was used to
peak
identifydeconvolution
and separate method [40]. In overlapping
the different this method,diffraction
a curve-fitting program
spectrum (e.g., the
of both PeakFit v4.12) was
crystalline and
used to identify
amorphous and separate
contributions, andthe different
then the areaoverlapping diffraction
was calculated under spectrum
each fittedofGaussian
both the peaks.
crystalline
It is
and
assumedamorphous
that thecontributions,
observed peak andbroadening
then the area wasspectra
in the calculated under each
is mainly fitted to
attributed Gaussian peaks.
the increased
It is assumedcontent
amorphous that theinobserved
the sample. peakForbroadening
example, in thecrystalline
four spectra is peaks
mainly(101,
attributed
101, 002 to and
the increased
040) and
amorphous content in the sample. For example, four crystalline peaks (101,
broad amorphous peaks between 18° to 22° were identified [41] and assumed as Gaussian functions 101, 002 and 040) and broad
amorphous peaks between 18 ◦ to 22◦ were identified [41] and assumed as Gaussian functions as shown
as shown in Figure 3. These Gaussian peaks were extracted from the XRD spectrum through the
in Figure 3. These
curve-fitting processGaussian peaks5000
with at least were extractedwhich
iterations, from corresponds
the XRD spectrum through of
to a coefficient thedetermination
curve-fitting
process 2 of 0.98.
(r ) of 0.98. Based from the results of curve fitting, CI is calculated as the ratio of the sum (r
2 with at least 5000 iterations, which corresponds to a coefficient of determination of )areas of
Based from the peaks
all crystalline resultstoof the
curve sumfitting, CI is calculated
of areas of all peaks as the ratio of the
(crystalline sum
and of areas of all
amorphous) as crystalline
shown in
peaks
Equation to the
(1) sum
[40]. of areas of all peaks (crystalline and amorphous) as shown in Equation (1) [40].

CI=∑ A
CI
∑ AA crystalline
crystalline
(1)
(1)
A
crystalline + AAamorphous
crystalline amorphous

Figure 3. X-ray diffraction (XRD) peak deconvolution method to measure crystallinity index.

2.7. Fourier
2.7. Fourier Transform Infrared (FTIR)
Transform Infrared (FTIR) Spectroscopy
Spectroscopy Measurements
Measurements
Attenuated total
Attenuated total reflectance
reflectance Fourier
Fourier transform
transform infrared
infrared (ATR-FTIR)
(ATR-FTIR) spectroscopy
spectroscopy (Perkin
(Perkin Elmer,
Elmer,
Waltham, MA, USA) was carried out to qualitatively identify the constituents
Waltham, MA, USA) was carried out to qualitatively identify the constituents of untreated and of untreated and
untreated abaca fibers. Test results were obtained using Perkin Elmer FTIR Spectrometer
untreated abaca fibers. Test results were obtained using Perkin Elmer FTIR Spectrometer Frontier with Frontier
with accessory
ATR ATR accessory in theof
in the range range of 4000–650
4000–650 cmATR
cm−1 . The −1. The ATR cell is equipped with a trapezoidal
cell is equipped with a trapezoidal diamond
diamond crystal as the internal reflection element. Baseline
crystal as the internal reflection element. Baseline correction was correction
appliedwas applied
to the to thetospectrum
spectrum improve
to improve its quality without distorting the band intensities in
its quality without distorting the band intensities in the final spectrum. the final spectrum.
Materials 2017, 10, 579 7 of 19
Materials 2017, 10, 579 7 of 19
2.8. Thermogravimetric Analysis (TGA) Measurements
Thermal analysis was performed using a thermogravimetric analyzer (TA Instruments TGA
2.8. Thermogravimetric Analysis (TGA) Measurements
Q50, New Castle, DE, USA). Samples were cut into small pieces (around 5 mm) of 20–25 mg and were
placedThermal analysis pan,
on a platinum was performed
then heatedusing
froma30
thermogravimetric
to 900 °C at a rateanalyzer
of 10 °C(TA
perInstruments TGA Q50,
min under Argon gas
New Castle, DE, USA). Samples were cut into small pieces (around 5 mm) of 20–25 mg
at 40 mL per min purge flow. Calculations of percentage mass loss and derivative plots were done and were
placed on TA
using the a platinum pan,
Universal then heated
Analysis from 30 to 900 ◦ C at a rate of 10 ◦ C per min under Argon gas at
software.
40 mL per min purge flow. Calculations of percentage mass loss and derivative plots were done using
2.9. TA
the Compressive
UniversalStrength and
Analysis Flexural Strength Test of Geopolymer Composites
software.
For compressive
2.9. Compressive strength
Strength testing,
and Flexural cured cubic
Strength Test of(5 cm × 5 cmComposites
Geopolymer × 5 cm) samples were subjected to a
constant stress rate of 0.25 MPa/s until failure. The compressive strength (CS) of a sample was
For compressive
calculated (2): testing, cured cubic (5 cm × 5 cm × 5 cm) samples were subjected
strength
using Equation
to a constant stress rate of 0.25 MPa/s until failure. The compressive strength (CS) of a sample was
P
calculated using Equation (2): CS  max (2)
A
Pmax
CS = (2)
where Pmax is the total load on the sample at failure andA A is the calculated area of the bearing surface
where Pmax is the total load on the sample at failure and A is the calculated area of the bearing surface
of the specimen.
of theFor
specimen.
flexural strengths or modulus of rupture (MOR) measurement, samples were subjected to a
For flexural
constant strengths
stress rate or modulus
of 0.25 MPa/s using aofthree-point
rupture (MOR)
flexuremeasurement,
loading until samples
failure aswere subjected
shown to 4.
in Figure a
constant
MOR was stress rate of 0.25
calculated basedMPa/s using
on the loadaatthree-point flexure
failure, cross loading
sectional until
area of failure as shown
the sample in Figureof
and distance 4.
MOR was calculated based
supports using Equation (3): on the load at failure, cross sectional area of the sample and distance of
supports using Equation (3):
3FL
MOR=
MOR  3FL2 (3)
(3)
22bd
bd 2
whereFFisisload
where at at
load failure, L isLthe
failure, is distance between
the distance support
between or bspan,
or span,
support is the bwidth
is theand d is the
width andthickness
d is the
of the specimen
thickness of the being tested.
specimen being tested.

Figure 4. Three-point flexural strength test.


Figure 4. Three-point flexural strength test.

3. Results and Discussion


3. Results and Discussion
3.1. Tensile
3.1. Tensile Strength
Strength of
of Chemically
Chemically Treated
TreatedAbaca
AbacaFiber
Fiber
Table33summarizes
Table summarizesthe the measured
measured tensile
tensile strength
strength ofchemically
of 10 10 chemicallytreatedtreated
wastewaste
abacaabaca
fibers fibers
using
using
the the treatment
treatment combinations
combinations according
according to the designed
to the designed experiment
experiment from DSD from (seeDSD (see Section
Section 2.2).
2.2). Tensile
Tensile strength of treated abaca fiber in this study was observed to range from
strength of treated abaca fiber in this study was observed to range from 210 to 450 MPa. Compared to 210 to 450 MPa.
Compared
the averageto the average
tensile strengthtensile strength
of untreated of untreated
abaca fiber (500abaca
MPa),fiber (500 MPa),
all treated fibersall treated
have fibers
tensile have
strength
tensile strength lower than that of the untreated. The sample with the lowest
lower than that of the untreated. The sample with the lowest tensile strength (210 MPa) was alkali tensile strength
(210 MPa) was
pre-treated, alkali in
soaking pre-treated, soaking in
Al(OH)3 solution forAl(OH)
12 h and3 solution for 12 h and partially neutralized up to
partially neutralized up to 4.5 pH while the
4.5 pH while the sample with the highest measured
sample with the highest measured tensile strength (450 MPa) tensile strength
did not(450 MPa)alkali
undergo did not undergo alkali
pretreatment but
Materials 2017, 10, 579 8 of 19

pretreatment but was directly soaked in Al(OH)3 solution also for 12 h with final pH of 6. Thus, the
Materials 2017, 10, 579 8 of 19
tensile strength of the fiber is observed to be significantly affected by the nature of chemical
treatment.
was directly soaked in Al(OH)3 solution also for 12 h with final pH of 6. Thus, the tensile strength of
Table
the fiber is observed 3. Measured
to be response
significantly (tensile
affected strength)
by the natureatof
different
chemicaltreatment conditions.
treatment.
NaOH Soaking Tensile
Table 3. Measured response (tensile strength) Final pH treatment conditions.
at different
Pretreatment Time (h) Strength (MPa)
NaOH Pretreatment
No 12
Soaking Time (h)
6.0
Final pH
450 ± 160
Tensile Strength (MPa)
No 0.5 7.5 430 ± 130
No 12 6.0 450 ± 160
No
No 0.5
6.25 6.0
7.5
420 ± 140
430 ± 130
No Yes 6.25 12 7.5
6.0 370
420± ±
82140
Yes No 12 0.5 4.5
7.5 340 ± 97
370 ± 82
No Yes 0.5 6.25 4.5
4.5 340340 ± 97
± 140
Yes 6.25 4.5 340 ± 140
Yes 6.25 6.0 300 ± 82
Yes 6.25 6.0 300 ± 82
Yes Yes 0.5 0.5 6.0
6.0 230 ± 82
230 ± 82
No No 6.25 6.25 7.5
7.5 220 ± 69
220 ± 69
Yes Yes 12 12 4.5
4.5 210210 ± 83
± 83

To quantify
To quantify the effect,
the effect, regression
regression analysis
analysis waswas conducted
conducted to model
to model the the response
response variable
variable (i.e.,(i.e.,
the tensile
the tensile strength)
strength) as aas a function
function of these
of these factors
factors namely
namely the alkali
the alkali pretreatment,
pretreatment, soaking
soaking timetime
andand
finalfinal
pH.pH. Results
Results fromfrom the statistical
the statistical analysis
analysis are summarized
are summarized in Figure
in Figure 5. 5%
5. At At 5% significance
significance level,
level,
the significant main-effect factors to tensile strength were the presence of NaOH pretreatment andand
the significant main-effect factors to tensile strength were the presence of NaOH pretreatment
the final
the final pH atpHwhich
at which
the Althe Al24(SO
2 (SO 4)3-treated
)3 -treated samples
samples werewere neutralized
neutralized (p-value
(p-value = 0.0117
= 0.0117 andand 0.0169,
0.0169,
respectively). NaOH pretreatment and higher final pH were generally observed to cause a decline in in
respectively). NaOH pretreatment and higher final pH were generally observed to cause a decline
tensile
tensile strength.
strength. The The soaking
soaking timetime of fiber
of fiber samples
samples in theinaluminum
the aluminum salt solution
salt solution was found
was found to haveto ahave
a relatively
relatively insignificant
insignificant effect effect
to the to the tensile
tensile strengthstrength
(p-value(p-value = 0.0806).
= 0.0806). Nonetheless,
Nonetheless, the interaction
the interaction of
of the soaking time with the other two factors was found to affect the
the soaking time with the other two factors was found to affect the tensile strength most significantly tensile strength most
significantly
(p-value = 0.0109).(p-value = 0.0109).
Furthermore, theFurthermore,
model was able the model was98.7%
to explain able toofexplain 98.7% of
the observed the observed
variation in
variation
tensile strengthin due
tensile strength
to these due to these factors.
factors.

Figure
Figure 5. Actual
5. Actual vs. Predicted
vs. Predicted Plot Plot
(- - - (-- 0.05
- - - 0.05 significance
significance curve;
curve; - - - --mean)
- - -mean) for regression
for the the regression model
model
and and the model
the model coefficients.
coefficients.

Figure
Figure 6 shows
6 shows a matrix
a matrix of interaction
of interaction plotsplots between
between the the factors
factors considered
considered in the
in the chemical
chemical
treatment. Each cell of the matrix shows the interaction of the row effect with the column effect.For
treatment. Each cell of the matrix shows the interaction of the row effect with the column effect.
For example, the left-most-lowest
example, the left-most-lowestcellcellrepresents
representsthetheinteraction
interaction of of
pHpH (row)
(row) andand
NaOHNaOH pretreatment
pretreatment
(column) effects. A line segment is plotted for each level of row effect using response
(column) effects. A line segment is plotted for each level of row effect using response values (tensile values (tensile
strength) predicted by the model. As observed in the interaction profile, abaca fiber samples that werethat
strength) predicted by the model. As observed in the interaction profile, abaca fiber samples
not were not pretreated
pretreated with NaOH with NaOH
showed showed
a much lessasignificant
much lessdecline
significant decline
in tensile in tensile
strength strength with
with increasing
final pH, whereas alkali-treated fibers steeply declined in strength with increasing final pH. This couldfinal
increasing final pH, whereas alkali-treated fibers steeply declined in strength with increasing
pH. This by
be explained could be explained
the extended by the
contact withextended contactused
NaOH solution withforNaOH solution used
neutralization, for neutralization,
resulting to further
stripping of cellulosic components. It was found that when the Al2 (SO4 )3 -treated abaca fibers were
Materials 2017, 10, 579 9 of 19

resulting to further stripping of cellulosic components. It was found that when the Al2(SO4)3-treated
adjusted
abaca to lower
fibers were final pH (4.5),
adjusted the tensile
to lower strength
final pH (4.5), improved
the tensilewith increased
strength soaking
improved time.
with When
increased
adjusted to a pH of 7.5, the tensile strength decreased with increasing soaking
soaking time. When adjusted to a pH of 7.5, the tensile strength decreased with increasing soakingtime in the aluminum
salt solution.
time Note that the
in the aluminum saltmain purpose
solution. Noteofthat
Al2 (SO
the4 )main
3 treatment
purpose is to
ofform deposits
Al2(SO that could protect
4)3 treatment is to form
the fiber from the harsh environment of geopolymer matrix. It is possible
deposits that could protect the fiber from the harsh environment of geopolymer matrix. that soaking in the aluminum
It is possible
salt solution inherently improves tensile strength but was counteracted by the
that soaking in the aluminum salt solution inherently improves tensile strength but was counteracted addition of NaOH
solution
by in adjusting
the addition of NaOHthe solution
final pH,inresulting
adjustinginstead
the finaltopH,
a decline
resultingin strength
instead towhen excessive
a decline alkali
in strength
solution is added to attain a high final pH.
when excessive alkali solution is added to attain a high final pH.

Figure 6. Interaction
Interaction plot
plot matrix
matrix of the factors for chemical treatment of the abaca fibers.

3.2.
3.2. Fourier
Fourier Transform Infrared Spectroscopy
Transform Infrared Spectroscopy (FTIR)
(FTIR) of
of Raw
Raw and
and Alkali-Treated
Alkali-TreatedAbaca
AbacaFiber
Fiber
FTIR
FTIR allows
allows the qualitative analysis
the qualitative analysis of of variations
variations in in surface
surface chemistry
chemistry of of natural
natural fibers after
fibers after
treatment. FTIR spectra of untreated, alkali-treated and NaOH-then-Al 2(SO4)3 treated abaca fibers are
treatment. FTIR spectra of untreated, alkali-treated and NaOH-then-Al2 (SO4 )3 treated abaca fibers
shown
are shown in Figure
in Figure7a–c. In In
7a–c. addition,
addition,Figure
Figure 7d7ddescribes
describesthe thespectra
spectraofofdried
driedprecipitate
precipitate or or residues
residues
from the spent solution after the Al 2(SO4)3 treatment of abaca fibers. Note that the spectra in these
from the spent solution after the Al2 (SO4 )3 treatment of abaca fibers. Note that the spectra in these
figures were offset
figures were offset vertically
vertically toto observe
observe thethe difference
difference more
more clearly.
clearly. Table
Table 44 shows
shows the the observed
observed bands
bands
in FTIR spectra and their assignments to functional groups. In general, untreated
in FTIR spectra and their assignments to functional groups. In general, untreated abaca fiber showed abaca fiber showed
loss
loss of
of functional
functional groups
groups based
based on the comparison
on the comparison of of its
its FTIR
FTIR spectrum
spectrum to to that
that of the NaOH-treated
of the NaOH-treated
fibers. Peaks that
fibers. Peaks that receded
recededin inthe
thetreated
treatedfibers
fibersover
overfrequencies
frequencies 1510
1510 cmcm−1 , ,1430
−1 1430cmcm −1
−1 andand1250
1250cm cm−−11
could
could entail
entail loss
loss of phenolic groups,
of phenolic groups, signifying
signifying removal
removal of of large amount of
large amount lignin and
of lignin and pectin
pectin whose
whose
structure is a phenolic polymer. Another visible difference was observed
structure is a phenolic polymer. Another visible difference was observed at 1740 cm due to C=O at 1740 cm −−1
1 due to C=O
stretching vibrationwhich
stretching vibration whichisisa characteristic
a characteristic band
band of hemicellulose
of hemicellulose [42].[42].
The The disappearance
disappearance of thisofband
this
band indicates the dissolution of hemicellulose after the alkali pretreatment.
indicates the dissolution of hemicellulose after the alkali pretreatment. However there is no evidence However there is no
evidence
indicatingindicating
complete complete
removal of removal of these impurities.
these impurities. The decreaseThe decrease
in tensileinstrength
tensile strength of the
of the samples
samples could be attributed to these extractions. In addition, characteristic peak
could be attributed to these extractions. In addition, characteristic peak of Al(OH)3 at 990 cm− 1 of Al(OH) 3 at 990 cm −1

corresponding
corresponding to to Al–O
Al–O bonds
bonds [43]
[43] was
was observed
observed in in the
the spectra
spectra of of NaOH-then-Al
NaOH-then-Al2(SO (SO4))3 treated abaca
2 4 3 treated abaca
fiber, and the
fiber, and the precipitate
precipitate or or residues
residues from
from thethe spent
spent solution.
solution. Thus,
Thus, the
the FTIR
FTIR results confirm the
results confirm the
possible formation of Al(OH) 3 during neutralization and its deposition on the abaca surface.
possible formation of Al(OH) during neutralization and its deposition on the abaca surface.
3
Materials 2017, 10, 579 10 of 19
Materials 2017, 10, 579 10 of 19

Figure 7. Fourier Transform Infrared Spectroscopy (FTIR) spectra of: (a) untreated abaca; (b) NaOH-
Figure 7. Fourier Transform Infrared Spectroscopy (FTIR) spectra of: (a) untreated abaca;
treated abaca; (c) NaOH + Al2(SO4)3 treated abaca; and (d) precipitate or residue from the spent
(b) NaOH-treated abaca; (c) NaOH + Al2 (SO4 )3 treated abaca; and (d) precipitate or residue from the
solution of Al2(SO4)3 treatment.
spent solution of Al2 (SO4 )3 treatment.
Table 4. Functional Groups of Observed Bands in FTIR spectra of abaca fiber [25,42,43].
Table 4. Functional Groups of Observed Bands in FTIR spectra of abaca fiber [25,42,43].
Wavenumber (cm−1) Vibration Source
3690
Wavenumber (cm )− 1 Free OH
Vibration moisture
Source
3350 O–H linked stretching Polysaccharide
3690 Free OH moisture
2920 C–H stretching Cellulose, Hemicellulose
3350 O–H linked stretching Polysaccharide
1740 C=O stretching Hemicellulose
2920 C–H stretching Cellulose, Hemicellulose
1630
1740 OH in H 2O, bending
C=O stretching Moisture
Hemicellulose
1510
1630 C=C aromatic symmetrical
OH in H2 O, bending stretching Lignin
Moisture
1430
1510 CH2 symmetric
C=C aromatic bending
symmetrical stretching Pectin, lignin
Lignin
1250
1430 CH2C–O aryl group
symmetric bending Lignin
Pectin, lignin
1250
980 C–O aryl
Al–O group
stretching Lignin3
Al(OH)
980
900 Al–O
Glycosidic bonds stretchingring stretching
symmetric Al(OH)3
Polysaccharide
900 Glycosidic bonds symmetric ring stretching Polysaccharide
3.3. XRD Analysis of Untreated and Alkali-Treated Abaca Fibers
3.3. XRD Figure
Analysis of Untreated and Alkali-Treated Abaca Fibers
8 shows the X-ray spectra or diffractograms of untreated and NaOH-treated abaca fibers.
Note
Figure that the first
8 shows two
the crystalline
X-ray spectrapeaks cannot be observed
or diffractograms separately
of untreated and because they overlapped.
NaOH-treated abaca fibers.
Note that the first two crystalline peaks cannot be observed separately because theywith
Moreover, due to the very broad XRD spectrum of the amorphous cellulose, it overlapped the
overlapped.
first three crystalline peaks (101, 101, and 002). In this figure, one can notice the broader peak around
Moreover, due to the very broad XRD spectrum of the amorphous cellulose, it overlapped with the
the peak 002 of untreated fiber as compared with that of treated fiber. Accordingly, fitted Gaussian
first three crystalline peaks (101, 101, and 002). In this figure, one can notice the broader peak around
crystalline peaks were obtained using the XRD peak deconvolution method described in Section 2.6
the peak 002 ofthe
to quantify untreated
individual fiber as compared
contributions with
of these that of treated
components fiber. Accordingly,
to the overall XRD spectrumfitted Gaussian
of the fiber
crystalline
sample. This was then used to compute the relative crystallinity values to interpret the changes in the2.6 to
peaks were obtained using the XRD peak deconvolution method described in Section
quantify the individual
cellulose-based contributions
structure of these
of the fiber after components
treatment. Results to
fromthethe
overall XRD
Gaussian peakspectrum of the fiber
fitting showed
sample.
the This wasof
presence then
the used to compute
four major the relative
crystalline peaks ofcrystallinity
cellulose (101,values to interpret
101, 002 and 040) the whichchanges
can bein the
observed at the
cellulose-based following
structure of 2-theta
the fiberangles,
after15.1°, 16.8°, 22.9°
treatment. and 34.7°,
Results from respectively.
the Gaussian The broad
peak peak showed
fitting for
the amorphous
the presence of thecellulose
four majorwas crystalline
observed at peaks
2-theta of
of 20.2°. Thus,(101,
cellulose the calculated
101, 002 and crystallinity
040) whichindicescan be
(CI) from the XRD peak deconvolution method
◦ using
◦ Equation
◦ (1)◦ for the untreated
observed at the following 2-theta angles, 15.1 , 16.8 , 22.9 and 34.7 , respectively. The broad peak for and NaOH-
treated abaca fibers were found to be 61.0% and 78.8%, respectively. It is anticipated that after
the amorphous cellulose was observed at 2-theta of 20.2◦ . Thus, the calculated crystallinity indices (CI)
sufficient alkali pretreatment, the fiber crystallinity index is bound to increase due to the extraction
from the XRD peak deconvolution method using Equation (1) for the untreated and NaOH-treated
or removal of its amorphous components including hemicelluloses, lignin and other non-cellulosic
abacacomponents
fibers werewhich found to bethe
allow 61.0% and fibers
cellulosic 78.8%,torespectively.
adopt a more It is anticipated
crystalline structurethat[31].
after
The sufficient
observedalkali
pretreatment, the fiber crystallinity index is bound to increase due to the
increase in crystallinity due to the removal of lignin and hemicellulose is also in agreement withextraction or removal
the of
its amorphous
results of FTIR components
analysis. including hemicelluloses, lignin and other non-cellulosic components
which allow the cellulosic fibers to adopt a more crystalline structure [31]. The observed increase in
crystallinity due to the removal of lignin and hemicellulose is also in agreement with the results of
FTIR analysis.
Materials 2017, 10, 579 11 of 19
Materials 2017, 10, 579 11 of 19

(a) (b)

Figure 8. XRD spectra of untreated (a) and NaOH-treated (b) abaca fibers.
Figure 8. XRD spectra of untreated (a) and NaOH-treated (b) abaca fibers.
3.4. Morphology and Elemental Microanalysis of Untreated and Treated Abaca Fiber Surface through SEM-EDS
3.4. Morphology and Elemental Microanalysis of Untreated and Treated Abaca Fiber Surface
Effect of alkali pretreatment to the morphology and chemistry of fiber surface is important in
through SEM-EDS
the development fiber–matrix interaction. SEM-EDS can also provide evidences of the surface
Effect of alkali
modification pretreatment
in terms to the morphology
of morphological and elemental andmicroanalyses.
chemistry of Based fiber surface is important
on the results of FTIRin the
analysis, there are fiber components from the untreated fiber that were
development fiber–matrix interaction. SEM-EDS can also provide evidences of the surface modification removed after its contact with
NaOH solution, thus it is expected to observe difference with
in terms of morphological and elemental microanalyses. Based on the results of FTIR analysis,their morphology. Figure 9 shows the there
SEMcomponents
are fiber images of raw andthe
from treated abaca fibers.
untreated fiber that Figure
were9a removed
shows impurities
after itson the textured
contact with NaOH surfacesolution,
of
untreated abaca fiber. After treatment with NaOH solution, impurities were almost removed
thus it is expected to observe difference with their morphology. Figure 9 shows the SEM images of raw
resulting to a more uniform but rougher or corrugated surface as shown in Figure 9b. This may be a
and treated abaca fibers. Figure 9a shows impurities on the textured surface of untreated abaca fiber.
result of stripping off the lignin and hemicellulose from fibers in alkali solution as indicated in the
AfterFTIR
treatment with NaOH solution, impurities were almost removed resulting to a more uniform
analysis. It is evident in the images that corrugation on the fiber surface became sharper after
but rougher or corrugated
treatment. This was observed surface as shown
among all theinfiveFigure
treated 9b.samples
This may thatbeunderwent
a result ofSEM stripping
analysis. off the
ligninCorrugation
and hemicellulose from fibers in alkali solution as indicated in the
on the fiber surface indicates a better fiber–matrix interface by virtue of mechanicalFTIR analysis. It is evident in
the images that corrugation
interlocking. Such corrugation on theinfiber
the surface became sharper
surface indicates after treatment.
higher frictional This was
bond between the observed
two
among components,
all the five which has previously
treated samples that shown to be effective
underwent SEMinanalysis.
resisting shear and bending
Corrugation on thestress [44].surface
fiber
indicates In Figure fiber–matrix
a better 9c, SEM imageinterfaceof a treated abaca fiber
by virtue (no alkali pretreatment,
of mechanical interlocking. 12Such
h soaking; final pHin the
corrugation
= 6) indicates the Al(OH) 3 deposits which almost covered the fiber surface. At higher magnification
surface indicates higher frictional bond between the two components, which has previously shown to
(5000×; Figure 9c1), the Al(OH)3 particles were observed to have porous structure. Figure 9d1 shows
be effective in resisting shear and bending stress [44].
the SEM image of another treated fiber (NaOH pretreated, 12 h soaking; final pH = 6). Similar to the
In Figure 9c, SEM image of a treated abaca fiber (no alkali pretreatment, 12 h soaking; final pH = 6)
previous image, it can be observed that the fiber is almost completely covered with Al(OH)3 particles.
indicates the Al(OH)
At higher 3 deposits
magnification which
(5000×; almost
Figure 9d2),covered
porous the fiber surface.
structure At higher
is still present with magnification
additional crystal (5000×;
Figure 9c1), the Al(OH)
like features. However, 3 particles
in Figure 9e,f, coarse and fine Al(OH)3 particles having porous and non- SEM
were observed to have porous structure. Figure 9d1 shows the
image of another
porous treated
structures werefiber (NaOH
observed pretreated,
on the treated abaca 12 fiber
h soaking;
covering final pH =the
partially 6).surface.
SimilarThis to the previous
implies
image,thatit can be observed
various forms of that Al(OH)the3fiber
wereisdeposited
almost completely
on the surface covered with Al(OH)
of abaca 3 particles.
fibers that were treated At higher
differently.(5000×; Figure 9d2), porous structure is still present with additional crystal like features.
magnification
However,EDS analysis
in Figure wascoarse
9e,f, also performed
and fine Al(OH)on the surface of alkali treated abaca surface as shown in
3 particles having porous and non-porous structures
were observed on the treated abaca fiber covering partiallyelemental
Figure 10. A region of interest was selected to measure the the surface. composition
This impliesin the saidvarious
that surface.forms
However, it should be noted that the quantification error could be large as sample volume for EDS is
of Al(OH)3 were deposited on the surface of abaca fibers that were treated differently.
small in an order of few micrometers in diameter. Nevertheless, the results of EDS is indicative of the
EDS analysis was also performed on the surface of alkali treated abaca surface as shown in
presence of compound with known chemical structure. For example, in Figure 10a, the measured
Figure 10. Atoregion
carbon oxygen of mass
interest ratiowas selected
(C/O) to measure
is around the elemental
1.3 whereas theoreticallycomposition
(based oninthe thechemical
said surface.
However,
formula) it should
the C/Obe of noted
cellulose that theand
is 0.9 quantification
C/O of ligninerror is 2.9could
[45,46].be large
This as sample
is indicative volume
that the ligninfor EDS
is small
andin an order of few
hemicellulose weremicrometers
almost removed in diameter.
from abaca Nevertheless,
fiber after alkali the results of EDS
treatment. is indicative
Moreover, the of
the presence
computedofO/Al compound
values inwith known
regions suchchemical
as that of structure.
Figure 10b For range example,
from 2.8in toFigure
3.2. This10a, the measured
is indicative
carbonof the presencemass
to oxygen of aluminum
ratio (C/O) compounds
is around in 1.3
the whereas
form of aluminum
theoretically hydroxide
(basedwhich
on thewas confirmed
chemical formula)
from of
the C/O thecellulose
FTIR results. Note
is 0.9 thatC/O
and the said O/Al calculation
of lignin is based
is 2.9 [45,46]. on the
This assumptionthat
is indicative that the
all oxygen
lignin and
atoms detected
hemicellulose wereare from removed
almost the organic components
from abaca fiber of abaca
after fiber
alkaliand aluminum
treatment. hydroxidethe
Moreover, deposits
computed
O/Al values in regions such as that of Figure 10b range from 2.8 to 3.2. This is indicative of the presence
of aluminum compounds in the form of aluminum hydroxide which was confirmed from the FTIR
results. Note that the said O/Al calculation is based on the assumption that all oxygen atoms detected
Materials 2017, 10, 579 12 of 19

Materials 2017, 10, 579 12 of 19

are from the organic components of abaca fiber and aluminum hydroxide deposits only, and that the
only, and that the carbon to oxygen mass ratio (C/O) of fiber is 1.3 to determine the oxygen mass from
carbon to oxygen mass ratio (C/O) of fiber is 1.3 to determine the oxygen mass from the organic.
the organic.

500x 100μm 500x 100μm 500x 100μm

impurities

a1 b1 c1
5000x 10μm 5000x 10μm 5000x 10μm Porous Al(OH)3

a2 b2 c2
500x 100μm 500x 100μm 500x 50μm

coarse particles

fine particles

d1 e1 f1
5000x 10μm fine particles

10,000x 1 μm 1 μm
10,000x
d2 e2 f2

Figure 9. Scanning electron microscopy (SEM) images at low and high magnification of raw
Figure 9. Scanning electron microscopy (SEM) images at low and high magnification of raw (untreated),
(untreated), NaOH-treated and Al2(SO4)3-treated abaca fibers: (a) raw (untreated); (b) NaOH treated;
NaOH-treated and Al2 (SO4 )3 -treated
(c) without NaOH pretreatment;
abaca fibers: (a) raw (untreated); (b) NaOH treated; (c) without
12 h soaking; final pH = 6; (d) with NaOH pretreatment; 12 h;
NaOH
pH =pretreatment;
6; (e) without12 h soaking;
NaOH final pH12= h;
pretreatment; 6; pH
(d) with
= 4.5;NaOH
and (f)pretreatment; 12 h; pH = 6;0.5
with NaOH pretreatment; (e)h,without
NaOH
pH =pretreatment;
6. Low and high12magnification
h; pH = 4.5;areand (f) with
labeled NaOH
as “1” pretreatment;
and “2”, respectively.0.5 h, pH = 6. Low and high
magnification are labeled as “1” and “2”, respectively.
Materials 2017, 10, 579 13 of 19
Materials 2017, 10, 579 13 of 19

Figure 10. Sample EDS spectra of: (a) NaOH-treated; and (b) Al2(SO4)3-treated abaca
Figure 10. Sample EDS spectra of: (a) NaOH-treated; and (b) Al2 (SO4 )3 -treated abaca.
3.4. Thermogravimetric Analysis of Raw and Treated Abaca Fiber
3.5. Thermogravimetric Analysis of Raw and Treated Abaca Fiber
Thermal stability of chemically treated abaca fibers was determined using thermogravimetric
analysis (TGA).
Thermal In of
stability TGA, thermal treated
chemically stability abaca
was studied in terms
fibers was of weight
determined losses
using under argon
thermogravimetric
atmosphere with respect to temperature ramping (10°/min) from 30
analysis (TGA). In TGA, thermal stability was studied in terms of weight losses °C to 900 °C. Note thatunder
naturalargon
fibers are mainly composed of hemicellulose, cellulose
◦ and lignin. Thermal

atmosphere with respect to temperature ramping (10 /min) from 30 C to 900 C. Note that degradation
◦ of these
natural
components was reported to take place in different temperature ranges [25,47]. The observed thermal
fibers are mainly composed of hemicellulose, cellulose and lignin. Thermal degradation of these
decompositions are summarized in Table 5. The value of Tmax represents the temperature at which
components was reported to take place in different temperature ranges [25,47]. The observed thermal
the maximum decomposition rate was observed based from the differential thermogravimetry (DTG)
decompositions
data.
are summarized in Table 5. The value of Tmax represents the temperature at which the
maximum decomposition rate was observed based from the differential thermogravimetry (DTG) data.
Table 5. Decomposition Temperature of Fiber Components [25,47]
Table 5. Decomposition Temperature of Fiber Components [25,47].
Components Temperature of Decomposition, °C Tmax Based on DTG, °C
Moisture 30–100 ◦ 80
Components Temperature of Decomposition, C Tmax Based on DTG, ◦ C
Hemicellulose (Xylan) 160–350 245 (side chain), 298 (backbone)
Moisture
Cellulose 30–100
240–365 33580
Hemicellulose (Xylan) 160–350 245 (side chain), 298 (backbone)
Lignin 300–500 337
Cellulose 240–365 335
Lignin 300–500 337
Thermograms of abaca samples treated differently in Figure 11 show the Tmax, percentage mass
loss and rate of decomposition at Tmax for each stage of thermal decomposition. For the four abaca
Thermograms
samples, the initial of abaca samples
mass loss ranging treated
from 7.8 differently in Figuremass
to 10.3 percentage 11 show the 50
loss from Tmax°C ,topercentage
around 120mass
°C is mainly due to the evaporation of free and absorbed moisture
loss and rate of decomposition at Tmax for each stage of thermal decomposition. For the four on the fiber. Considering theabaca
untreated abaca fiber, the main decomposition event was observed
samples, the initial mass loss ranging from 7.8 to 10.3 percentage mass loss from 50 C to aroundfrom 200 °C to 370 °C ◦
equivalent
to is
120 ◦ C percentage
mainly due masstoloss
theofevaporation
78.2%. At thisof temperature range, a shoulder
free and absorbed moisture before the fiber.
on the peak located at 300 the
Considering
°C and a peak was observed at 345 °C (DTG). Fastest rate of decomposition with respect to
untreated abaca fiber, the main decomposition event was observed from 200 ◦ C to 370 ◦ C equivalent to
temperature was recorded at 1.1 percentage mass loss per °C. From the onset of the main thermal ◦
percentage mass loss of 78.2%. At this temperature range, a shoulder before the peak located at 300 C
decomposition to the location of shoulder, the mass loss is mainly due to the decomposition of
and ahemicellulose.
peak was observed As the at 345 ◦ C (DTG).
temperature Fastest
increases, ratecellulose
more of decomposition
decomposed withandrespect
lignin to temperature
started its
was recorded at 1.1 percentage ◦
degradation. Beyond 370 °C,massslower loss per C. Fromofthe
decomposition theonset of thecomponents
remaining main thermal decomposition
proceeded up to to
the location of shoulder,
900 °C resulting themass
to a total massloss lossof is mainly
87.1% due tomass
or residual the decomposition
of 12.9%. On one of hemicellulose.
hand, thermogram of As the
temperature increases, more cellulose decomposed and lignin started its degradation. Beyond 370 ◦ C,
abaca fiber treated with Al 2 (SO 4) 3 only is shown in Figure 11b. The features of this plot are similar to
slowerthat of untreated abaca
decomposition except
of the for the less
remaining obvious shoulder
components proceededin DTGup plot and◦ C
to 900 lower percentage
resulting massmass
to a total
loss ofloss in the
87.1% main decomposition
or residual mass of 12.9%. event,
Onand one thus
hand,leaving more residues.
thermogram of abacaThese differences
fiber treated withcanAl2be(SO4 )3
associated to the partial removal of hemicellulose and lignin during the
only is shown in Figure 11b. The features of this plot are similar to that of untreated abaca except forneutralization process by
adding NaOH solution. Figure 11c,d show thermograms of abaca fiber treated with NaOH, and
the less obvious shoulder in DTG plot and lower percentage mass loss in the main decomposition
NaOH then with Al2(SO4)3, respectively. It can be observed that the shoulder feature is no longer
event,present
and thus leaving more residues. These differences can be associated to the partial removal of
in the DTG plots of these samples. Since these samples were exposed to a higher
hemicellulose
concentration andNaOHligninsolution
duringfor thea neutralization
longer period, more process by adding
hemicellulose NaOHwere
particles solution.
removed Figure
from11c,d
showthe thermograms of abaca fiber
fiber during the chemical treatment. treated with NaOH, and NaOH then with Al 2 (SO )
4 3 , respectively.
It can be observed that the shoulder feature is no longer present in the DTG plots of these samples.
Since these samples were exposed to a higher concentration NaOH solution for a longer period, more
hemicellulose particles were removed from the fiber during the chemical treatment.
Materials 2017, 10, 579 14 of 19
Materials 2017, 10, 579 14 of 19

Thermal
Thermal decompositionofofAl(OH)
decomposition Al(OH) asreported
3 3as reported from a previous
previousstudy
study[48] [48]starts
startsatatabout
about240 °C◦ C
240
upupto to about400
about ◦
400 °C
C overlapping
overlappingwith
withthethe
thermal decomposition
thermal of theoforganic
decomposition components
the organic of abaca of
components
fibers. Furthermore, due to the small amount of Al(OH) deposited on abaca
abaca fibers. Furthermore, due to the small amount of Al(OH)3 deposited on abaca surface,
3 surface, the effect ofthe
Al(OH)
effect of Al(OH)3 thermal decomposition to the overall thermograms of Al2 (SO4 )3 treated fibersnot
3 thermal decomposition to the overall thermograms of Al2 (SO 4 ) 3 treated fibers was was
notobvious.
obvious.

Figure Thermograms
11.11.
Figure Thermograms of:of:(a)
(a)untreated;
untreated;(b)
(b)treated
treatedsolely
solely with
with Al (SO44)3);3(c)
Al22(SO ; (c)alkali-pretreated;
alkali-pretreated; and
and
(d)(d)
alkali and Al2 (SO 4 )3 solution-treated abaca fiber samples.
alkali and Al2(SO4)3 solution-treated abaca fiber samples

3.6.3.4.
Preliminary Characterization
Preliminary CharacterizationofofFiber-Reinforced
Fiber-Reinforced Geopolymers
Geopolymers
Incorporatingabaca
Incorporating abacafiber
fiber(about
(about1% 1%byby weight
weight of fly ash)ash) in
in geopolymer
geopolymermatrix matriximproved
improved thethe
compressive
compressive strength
strength ofofuntreated
untreatedfiber-reinforced
fiber-reinforced (25.9
(25.9 MPa)
MPa) and
andtreated
treatedfiber-reinforced
fiber-reinforced composite
composite
(22.2
(22.2 MPa)
MPa) byby 20%20%and
and3%,3%,respectively,
respectively, compared
compared to to that
that of
of aa pristine
pristinegeopolymer
geopolymer(21.6 MPa).
(21.6 MPa).
Likewise, the flexural strength of the untreated fiber reinforced (5.5 MPa) and
Likewise, the flexural strength of the untreated fiber reinforced (5.5 MPa) and the fiber-reinforced the fiber-reinforced
composite
composite (7.3(7.3 MPa)
MPa) wasalso
was alsoimproved
improvedby by95%
95% and
and 161%,
161%, respectively,
respectively,compared
comparedtotothatthatofofa pristine
a pristine
geopolymer (2.8 MPa). Such results indicate that the fiber’s main contribution is to
geopolymer (2.8 MPa). Such results indicate that the fiber’s main contribution is to improve primarily improve primarily
thethe composite’s
composite’s flexuralstrength
flexural strengthandandnot
notmuch
much of of its
its compressive
compressive strength
strengthsince
sincethe
thefibers control
fibers control
the cracking that gives rise to a “graceful” fracture by bridging across the cracks. Figure 12 shows the
the cracking that gives rise to a “graceful” fracture by bridging across the cracks. Figure 12 shows
fracture surface of the pristine geopolymer and those of the abaca fiber-reinforced. Moreover, a
the fracture surface of the pristine geopolymer and those of the abaca fiber-reinforced. Moreover,
significant improvement in flexural strength of the composite reinforced with chemically treated
a significant improvement in flexural strength of the composite reinforced with chemically treated
abaca fiber suggests a better interfacial adhesion between the fiber and the matrix. However, the
abaca fiber suggests a better interfacial adhesion between the fiber and the matrix. However, the
difference of improvement between the mechanical properties of geopolymer reinforced with
difference
untreatedof fiber
improvement
and treatedbetween theneeds
fiber still mechanical
furtherproperties
verificationof in
geopolymer
future works.reinforced with untreated
fiber and treated fiber still needs further verification in future works.
Materials 2017, 10, 579 15 of 19
Materials 2017, 10, 579 15 of 19

Figure
Figure12.12.Fracture
Fracturesurface
surfaceofofpristine
pristine geopolymer, reinforcedwith
geopolymer, reinforced withuntreated
untreatedabaca
abacaand
and reinforced
reinforced
with treated abaca.
with treated abaca.

In this paper, we focus on the microstructure characterization of the abaca-fiber reinforced


In this paper, we focus on the microstructure characterization of the abaca-fiber reinforced
geopolymer composites through SEM and thermogravimetric analysis. Representative fractured
geopolymer composites through SEM and thermogravimetric analysis. Representative fractured
surfaces
surfaces ofofthethecomposites
compositeswere were analyzed
analyzed using SEM-EDS as
using SEM-EDS asshown
shownininFigure Figure 13.13.Heterogeneous
Heterogeneous
geopolymer of different morphology such as textured spheres, porous
geopolymer of different morphology such as textured spheres, porous and crystal-like structures and crystal-like structures
were
were observed for both composites samples. Results of EDS indicated
observed for both composites samples. Results of EDS indicated that Si and Al in the geopolymer that Si and Al in the
geopolymer have an overall atomic ratio of 2:1. Fibers were clearly
have an overall atomic ratio of 2:1. Fibers were clearly embedded on the geopolymer matrix as shownembedded on the geopolymer
matrix
fromas theshown from theFor
micrographs. micrographs.
compositesFor composites
reinforced withreinforced
untreatedwith abaca untreated abaca visible
(Figure 13a,c), (Figuregaps
13a,c),
visible gaps were observed between the geopolymer matrix and fiber
were observed between the geopolymer matrix and fiber indicating poor interfacial adhesion. This is indicating poor interfacial
adhesion.
likely due This is likely
to the due to theof
incompatibility incompatibility
the fiber and matrix. of the Circular-like
fiber and matrix. Circular-like
particles particles were
were also observed on
also
theobserved on theare
surface, which surface,
likely which are likely to
to be geopolymer be geopolymer
precursors or nuclei precursors
that adhered or nuclei
on thethatabaca adhered
surface.on
theOn the other
abaca surface. hand,
Onforthethe composites
other hand, for reinforced with treated
the composites abacawith
reinforced fiberstreated
(Figureabaca
13b,d), narrower
fibers (Figure
gaps were observed between the matrix and treated fiber. The zeolite-like
13b,d), narrower gaps were observed between the matrix and treated fiber. The zeolite-like particles particles on the fiber surface
onindicate
the fiberthat reaction
surface took place
indicate thatonreaction
the surface tookforplace
these on
structures to form.
the surface forClustering of abacato
these structures fibers
form.
was also observed in other region of the fractured surface. Although
Clustering of abaca fibers was also observed in other region of the fractured surface. Although geopolymer and zeolite deposits
were observed
geopolymer andwhichzeoliteindicates
depositsthe interaction
were observed between the geopolymer
which indicates matrix and fiber surface,
the interaction between pullthe
out sites were
geopolymer matrixobserved suggesting
and fiber surface,the pullformed interfacial
out sites adhesion was
were observed not enough
suggesting in some interfacial
the formed regions.
adhesion was not enough in some regions. Nevertheless, the pulling out of these fibers duringtothe
Nevertheless, the pulling out of these fibers during the loading process absorbs energy that results
the improvement of the flexural strength of composite.
loading process absorbs energy that results to the improvement of the flexural strength of composite.
After compressive strength testing, treated abaca fibers from the fractured composites were
After compressive strength testing, treated abaca fibers from the fractured composites were
removed from the composite and subjected to TG analysis using the same method of thermal stability
removed from the composite and subjected to TG analysis using the same method of thermal stability
analysis of free abaca fibers. Thermogram of abaca fibers collected from the composite is shown in
analysis of free abaca fibers. Thermogram of abaca fibers collected from◦ the composite is shown in
Figure 14b. Tmax of the main thermal decompositions was observed at 287 C and the total percentage
Figure 14b. Tmax of the main thermal decompositions was observed at 287 °C and the total percentage
mass loss was 43.3%. In comparison to free treated abaca fibers, the percentage mass loss for the main
mass loss was 43.3%.
decomposition eventIniscomparison to free
lower for fibers treated
from abaca fibers,
the composite thanthethepercentage
free fibermass
samples.loss for the main
Moreover,
decomposition event is lower for fibers from the composite than
◦ the free
the measured rate of decomposition at Tmax is lower at 0.2% per C versus the rate at Tmax of free abaca fiber samples. Moreover,
thefiber
measured
at averagerateofof1.2%
decomposition
per ◦ C. This at is T max is lower at 0.2% per °C versus the rate at Tmax of free abaca
likely due to sites of geopolymer growth formed on the surface
fiber at average of 1.2% per °C. This is
of the treated fibers. The abaca fibers were coveredlikely due to with
sites geopolymer
of geopolymer growth
particles andformed on the surface
thus protecting the
of fibers
the treated fibers. The
from thermal abaca fibers were covered with geopolymer particles and thus protecting
degradation.
the fibers from thermal degradation.
Materials 2017, 10, 579 16 of 19

Materials 2017, 10, 579 16 of 19

Figure
Figure13.13.SEM
SEM images of composite
images of compositefractured
fracturedsurfaces:
surfaces:
(a) (a) reinforced
reinforced with with untreated
untreated abaca;abaca;
(b)
(b)reinforced
reinforcedwith
withtreated
treated
abaca; (c) interface between untreated abaca fiber and geopolymer matrixmatrix
abaca; (c) interface between untreated abaca fiber and geopolymer and
and(d)
(d) interface
interface between
between treated
treated abaca
abaca fiber
fiber andand geopolymer
geopolymer matrix.
matrix.

Figure
Figure 14.14.
SEMSEM image(a)
image (a)and
andthermogram
thermogramof
of untreated
untreated abaca
abaca fibers
fibersand
andtreated
treatedabaca
abacafibers (b)(b)
fibers from
from
the geopolymer composite.
the geopolymer composite.

4. Conclusions
4. Conclusions
Chemical treatment of waste abaca fibers modifies their structure and chemical composition,
Chemical treatment of waste abaca fibers modifies their structure and chemical composition,
with the modification depending on the treatment conditions. The high tensile strength among the
with the modification
treated depending
fibers was achieved on the
without treatment
alkali conditions.
pretreatment (6% by The high tensile
wt NaOH solution strength among
for 48 h), soakedthe
treated
for 12fibers was achieved
h in aluminum without
sulfate alkali
solution andpretreatment
neutralized to(6% pHby6. wt NaOH
FTIR solution
results indicatefor 48the
that h), alkali
soaked
forpretreatment
12 h in aluminumsuccessfully dissolved some of the amorphous and hydrophobic components suchalkali
sulfate solution and neutralized to pH 6. FTIR results indicate that the as
pretreatment successfully dissolved
lignin, pectin and hemicellulose, some
and of theresults
the XRD amorphous
indicateand
an hydrophobic components
increase in fiber crystallinitysuch
that as
lignin,
may pectin and hemicellulose,
be attributed and the
to such dissolution. XRD results
Results indicate an increase
from thermogravimetric in fiber
analysis (TGA)crystallinity
confirm thethat
may be attributed
removal to pectin
of lignin, such dissolution. Resultsfrom
and hemicellulose from thethermogravimetric analysis
fibers, and also suggest that(TGA) confirm the
the geopolymer

removal of lignin,
itself coat pectin and
the treated hemicellulose
abaca fibers and from thethem
protect fibers,from
and also suggest
thermal that the geopolymer
degradation. The Al2(SOitself
4) 3
Materials 2017, 10, 579 17 of 19

coat the treated abaca fibers and protect them from thermal degradation. The Al2 (SO4 )3 treatment is
effective to form Al(OH)3 deposits that roughen the surface for better fiber- matrix interface and could
also protect the fiber from the harsh environment of the geopolymer matrix. Preliminary results on
abaca fiber-reinforced geopolymer indicate better adhesion for the treated fiber/matrix which results
in an improved flexural strength as compared to the pristine geopolymer. Future works will further
investigate the effect of chemical treatment on other properties of fiber such as swelling, as well as the
mechanical properties of such fiber-reinforced geopolymer composites.

Acknowledgments: The authors acknowledge the anonymous reviewers for their comments and suggestions
that help improve the content of the paper. The authors would like also to acknowledge the financial support for
the research project (P-015) from the National Research Council of the Philippines (NRCP), Philippine Council
for Industry, Energy and Emerging Technology Research and Development (PCIEERD) and Advanced Device
and Materials Testing Laboratory (ADMATEL). We would also to thank the Polymer Chemistry Laboratory and
Synthetic Organic Chemistry Research Laboratory, Institute of Chemistry, University of the Philippines Diliman.
Author Contributions: Michael Angelo Promentilla and Roy Alvin Malenab conceived the research and designed
the experiment including the write-up of the main narrative of the paper. Janne Pauline Ngo helped in writing the
manuscript, performing the experiments and analyzing the data.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Khalil, H.P.; Bhat, A.H.; Yusra, A.F. Green composites from sustainable cellulose nanofibrils: A review.
Carbohydr. Polym. 2012, 87, 963–979. [CrossRef]
2. Balasubramanian, M. Composite Materials and Processing; CRC Press Taylor & Francis Group: Boca Raton, FL,
USA, 2014; pp. 2–4.
3. Chandrasekar, M.; Ishak, M.R.; Sapuan, S.M.; Leman, Z.; Jawaid, M. A review on the characterisation
of natural fibres and their composites after alkali treatment and water absorption. Plast. Rubber Compos.
Process. Appl. 2017, 46, 119–136. [CrossRef]
4. Achour, A.; Ghomari, F.; Belayachi, N. Properties of cementitious mortars reinforced with natural fibers.
J. Adhes. Sci. Technol. 2017, 31, 1938–1962. [CrossRef]
5. Coutts, R.; Warden, P. Air-cured abaca reinforced cement composites. Int. J. Cem. Compos. 1987, 9, 69–73.
[CrossRef]
6. Punyamurthy, R.; Sampathkumar, D.; Ranganagowda, R.P.; Bennehalli, B.; Srinivasa, C.V. Mechanical
properties of abaca fiber reinforced polypropylene composites: Effect of chemical treatment by
benzenediazonium chloride. J. King Saud Univ. Eng. Sci. 2015. [CrossRef]
7. Bledzki, A.K.; Mamun, A.A.; Jaszkiewicz, A.; Erdmann, K. Polypropylene composites with enzyme modified
abaca fibre. Compos. Sci. Technol. 2010, 70, 854–860. [CrossRef]
8. Liu, K.; Zhang, X.; Takagi, H.; Yang, Z.; Wang, D. Effect of chemical treatments on transverse thermal
conductivity of unidirectional abaca fiber/epoxy composite. Compos. Part A Appl. Sci. Manuf. 2014, 66,
227–236. [CrossRef]
9. Bledzki, A.K.; Faruk, O.; Sperber, V.E. Cars from Bio-Fibres. Macromol. Mater. Eng. 2006, 291, 449–457.
[CrossRef]
10. Philippine Statistics Authority, Republic of the Philippines. Selected Statistics on Agriculture; PSA: Quezon
City, Philippines, 2016; pp. 13–18; ISSN 2012-0362. Available online: https://psa.gov.ph/sites/default/files/
Selected%20Statistics%20on%20Agriculture%202016.pdf (accessed on 15 March 2017).
11. Philippine Fiber Industry Development Authority, Department of Agriculture, Republic of the Phillipines.
Available online: http://fida.da.gov.ph/index.php/archived-articles/19-philippine-abaca-helps-in-global-
environment-conservation (accessed on 15 March 2017).
12. Dungani, R.; Karina, M.S.; Sulaeman, A.; Hermawan, D.; Hadiyane, A. Agricultural waste fibers towards
sustainability and advanced utilization: A review. Asian J. Plant Sci. 2016, 15, 42–55. [CrossRef]
13. Davidovits, J. Geopolymers and geopolymeric materials. J. Therm. Anal. 1989, 35, 429–441. [CrossRef]
14. Duxson, P.; Fernández-Jiménez, A.; Provis, J.L.; Lukey, G.C.; Palomo, A.; Deventer, J.S. Geopolymer
technology: The current state of the art. J. Mater. Sci. 2006, 42, 2917–2933. [CrossRef]
Materials 2017, 10, 579 18 of 19

15. Singh, B.; Ishwarya, G.; Gupta, M.; Bhattacharyya, S. Geopolymer concrete: A review of some recent
developments. Constr. Build. Mater. 2015, 85, 78–90. [CrossRef]
16. Ribeiro, R.A.; Ribeiro, M.G.; Sankar, K.; Kriven, W.M. Geopolymer-bamboo composite—A novel sustainable
construction material. Constr. Build. Mater. 2016, 123, 501–507. [CrossRef]
17. Musil, S.S.; Keane, P.F.; Kriven, W.M. Green Composite: Sodium-based Geopolymer Reinforced with
Chemically Extracted Corn Husk Fibers. In Development in Strategic Materials and Computational Design IV;
The American Ceramic Society; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2014; pp. 123–133. [CrossRef]
18. Alzeer, M.; Mackenzie, K.J.D. Synthesis and mechanical properties of novel composites of inorganic polymers
(geopolymers) with unidirectional natural flax fibres (phormiumtenax). Appl. Clay Sci. 2013, 75–76, 148–152.
[CrossRef]
19. Yan, L.; Kasal, B.; Huang, L. A review of recent research on the use of cellulosic fibres, their fibre fabric
reinforced cementitious, geo-polymer and polymer composites in civil engineering. Compos. Part B Eng.
2016, 92, 94–132. [CrossRef]
20. Davidovits, J. Reinforced Geopolymer Composites: A Critical Review. Available online:
http://www.materialstoday.com/polymers-soft-materials/features/reinforced-geopolymer-composites-
a-critical-review/ (accessed on 28 March 2017).
21. Alzeer, M.; Mackenzie, K.J. Synthesis and mechanical properties of new fibre-reinforced composites of
inorganic polymers with natural woolfibres. J. Mater. Sci. 2012, 47, 6958–6965. [CrossRef]
22. Korniejenko, K.; Fraczek,
˛ E.; Pytlak, E.; Adamski, M. Mechanical properties of geopolymer composites
reinforced with natural fibers. Procedia Eng. 2016, 151, 388–393. [CrossRef]
23. Mallick, P.K. Fiber-Reinforced Composites—Materials, Manufacturing and Design, 3rd ed.; CRC Press Taylor &
Francis Group: Boca Raton, FL, USA, 2008; pp. 19–126.
24. Pickering, K.L.; AruanEfendy, M.G.; Le, T.M. A review of recent developments in natural fibre composites
and their mechanical performance. Compos. Part A Appl. Sci. Manuf. 2016, 83, 98–112. [CrossRef]
25. Jabbar, A.; Militký, J.; Wiener, J.; Javaid, M.U.; Rwawiire, S. Tensile, surface and thermal characterization of
jute fibres after novel treatments. Indian J. Fibre Text. Res. 2016, 41, 249–254.
26. Ramadevi, P.; Sampathkumar, D.; Srinivasa, C.; Bennehalli, B. Effect alkali treatment on water absorption of
single cellulosic. BioResources 2012, 7, 3515–3524.
27. Onuaguluchi, O.; Banthia, N. Plant-based natural fibre reinforced cement composites: A review.
Cem. Concr. Compos. 2016, 68, 96–108. [CrossRef]
28. Kabir, M.M.; Aravinthan, H.; Wang, H.; Cardona, T.; Lau, K.T. Effects of natural fiber surface on composite
properties: A review. In Proceedings of the 1st International Postgraduate Conference on Engineering,
Designing and Developing the Built Environment for Sustainable Wellbeing, Darling Heights, Australia,
27–29 April 2011; pp. 94–99.
29. Izani, M.N.; Paridah, M.; Anwar, U.; Nor, M.M.; H’Ng, P. Effects of fiber treatment on morphology, tensile
and thermogravimetric analysis of oil palm empty fruit bunches fibers. Compos. Part B Eng. 2013, 45,
1251–1257. [CrossRef]
30. Ramadevi, P.; Sampathkumar, D.; Ranganagowda, R.; Bennehalli, B.; Badyankal, P.; Venkateshappa, S.C.
Surface modification of abaca fiber by benzene diazonium chloride treatment and its influence on tensile
properties of abaca fiber reinforced polypropylene composites. Ciênc. Tecnol. Mater. 2014, 26, 142–149.
[CrossRef]
31. Cai, M.; Takagi, H.; Nakagaito, A.N.; Li, Y.; Waterhouse, G.I. Effect of alkali treatment on interfacial bonding
in abaca fiber-reinforced composites. Compos. Part A Appl. Sci. Manuf. 2016, 90, 589–597. [CrossRef]
32. Lima, H.C.; Willrich, F.L.; Barbosa, N.P.; Rosa, M.A.; Cunha, B.S. Durability analysis of bamboo as concrete
reinforcement. Mater. Struct. 2007, 41, 981–989. [CrossRef]
33. Cook, J.T.; Schoggen, H.L.; Bell, R.I.; Smith, D.J.; Morton, G.H. Cementitious Material Reinforced with
Chemically Treated Cellulose Fiber. U.S. Patent No. US7357833, 2008.
34. Jones, B.; Nachtsheim, C.J. A class of screening designs robust to active second-order effects. J. Qual. Technol.
2010, 105–112. [CrossRef]
35. Jones, B.; Nachtsheim, C.J. Definitive screening designs with added two-level categorical factors.
J. Qual. Technol. 2013, 45, 121–129; ISSN: 00224065.
36. Mwaikambo, L.Y.; Ansell, M.P. Hemp fibre reinforced cashew nut shell liquid composites. Compos. Sci. Technol.
2003, 63, 1297–1305. [CrossRef]
Materials 2017, 10, 579 19 of 19

37. Sedan, D.; Pagnoux, A.; Smith, A.; Chotard, T. Mechanical properties of hemp fibre reinforced cement:
Influence of the fibre/matrix interaction. J. Eur. Ceram. Soc. 2008, 28, 183–192. [CrossRef]
38. Peden, M.E.; Amankwah, S.A.; Keller, B.J.; Krug, E.C.; Peden, J.M. Evaluation of Aluminum Speciation Using
Synthetics and Natural Samples: Final Report; SWS Contract Report 463; Illinois State Water Survey Division:
Champaign, IL, USA, 1989; pp. 13–22.
39. Kalaw, E.M.; Culaba, A.; Hinode, H.; Kurniawan, W.; Gallardo, S.; Promentilla, M.A. Optimizing and
Characterizing Geopolymers from Ternary Blend of Philippine Coal Fly Ash, Coal Bottom Ash and Rice
Hull Ash. Materials 2016, 9, 580. [CrossRef]
40. Park, S.; Baker, J.O.; Himmel, M.E.; Parilla, P.A.; Johnson, D.K. Cellulose crystallinity index: Measurement
techniques and their impact on interpreting cellulase performance. Biotechnol. Biofuels 2010, 3, 10. [CrossRef]
[PubMed]
41. Ouajai, S.; Shanks, R. Composition, structure and thermal degradation of hemp cellulose after chemical
treatments. Polym. Degrad. Stab. 2005, 89, 327–335. [CrossRef]
42. Le Troëdec, M.L.; Peyratout, C.S.; Smith, A.; Chotard, T. Influence of various chemical treatments on the
interactions between hemp fibres and a lime matrix. J. Eur. Ceram. Soc. 2009, 29, 1861–1868. [CrossRef]
43. Goudarzi, M.; Ghanbari, D.; Salavati-Niasari, M.; Ahmadi, A. Syntheis and characterization of Al(OH)3 ,
Al2 O3 Nanoparticles and Polymeric Nanocomposites. J. Clust. Sci. 2016, 27, 25–38. [CrossRef]
44. Ranjbar, N.; Mehrali, M.; Mehrali, M.; Alengaram, U.J.; Jumaat, M.Z. High tensile strength fly ash based
geopolymer composite using copper coated micro steel fiber. Constr. Build. Maters. 2016, 112, 629–638.
[CrossRef]
45. Sgriccia, N.; Hawley, M.C.; Misra, M. Characterization of natural fiber surfaces and natural fiber composites.
Compos. Part A Appl. Sci. Manuf. 2008, 39, 1632–1637. [CrossRef]
46. Brigida, A.I.S.; Calado, V.M.A.; Goncalves, L.R.B.; Coelho, M.A.Z. Effect of chemical treatments on properties
of green coconut fiber. Carbohydr. Polym. 2010, 79, 832–838. [CrossRef]
47. Zhang, J.; Feng, L.; Wang, D.; Zhang, R.; Liu, G.; Cheng, G. Thermogravimetric analysis of lignocellulosic
biomass with ionic liquid pretreatment. Bioresour. Technol. 2014, 153, 379–382. [CrossRef] [PubMed]
48. Hapuarachchi, T.D.; Peijs, T. Aluminum triohydroxide in combination with ammonium polyphosphate as
flame retardant for unsaturated polyester resin. Express Polym. Lett. 2009, 3, 743–751. [CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like