You are on page 1of 53

Accepted Manuscript

Title: Isolation and characterization of nanocrystalline


cellulose from sugar palm fibres (Arenga Pinnata)

Authors: R.A. Ilyas, S.M. Sapuan, M.R. Ishak

PII: S0144-8617(17)31331-0
DOI: https://doi.org/10.1016/j.carbpol.2017.11.045
Reference: CARP 12996

To appear in:

Received date: 8-6-2017


Revised date: 25-10-2017
Accepted date: 14-11-2017

Please cite this article as: Ilyas, RA., Sapuan, SM., & Ishak, M.R., Isolation
and characterization of nanocrystalline cellulose from sugar palm fibres (Arenga
Pinnata).Carbohydrate Polymers https://doi.org/10.1016/j.carbpol.2017.11.045

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Isolation and characterization of nanocrystalline cellulose from sugar

palm fibres (Arenga Pinnata)

R.A. Ilyas1*, S.M. Sapuan1, 2, M.R. Ishak3

PT
1
Laboratory of Biocomposite Technology, Institute of Tropical Forestry and Forest Products,

RI
Universiti Putra Malaysia, 43400 UPM Serdang, Selangor, Malaysia

SC
2
Department of Mechanical and Manufacturing Engineering, Universiti Putra Malaysia, 43400

UPM Serdang, Selangor, Malaysia

U
3
Department of Aerospace Engineering, Universiti Putra Malaysia, 43400 UPM Serdang,

Selangor, Malaysia N
A
* Corresponding author. Tel.: +603-89471788; Fax: +603-86567122
M

E-mail address: sapuan@upm.edu.my


D

ABSTARCT
TE

Cellulose was extracted from sugar palm fibres (arenga pinnata) by conducting delignification
EP

and mercerization treatments. Subsequently, sugar palm nanocrystalline celluloses (SPNCCs)

were isolated from the extracted cellulose with 60 wt% concentrated sulphuric acid. The
CC

chemical composition of sugar palm fibres were determined at different stages of treatment.

Structural analysis was carried out by Brunauer-Emmett-Teller (BET), X-ray diffraction (XRD)
A

and Fourier transform infrared spectroscopy (FT-IR). Morphological analysis of extracted

cellulose and isolated nanocrystalline cellulose (NCCs) was investigated by using field emission

scanning electron microscopy (FESEM), atomic force microscopy (AFM), and transmission

electron microscopy (TEM). The thermal stability of sugar palm fibres at different stages of
treatment was investigated by thermogravimetric analysis (TGA). The results showed that lignin

and hemicellulose were removed from the extracted cellulose through the delignification and

mercerization process, respectively. The isolated SPNCCs were found to have length and

diameters of 130 ± 30 nm and 9 ± 1.96 nm, respectively.

PT
Keywords: Sugar palm fibres; Cellulose; Sugar palm nanocrystalline cellulose; Nanofibre;

Thermal behavior; Acid hydrolysis.

RI
SC
1. Introduction

U
Nanocrystalline celluloses (NCCs) isolated from plant fibers attracted a tremendous interest in

N
material science due to its appealing intrinsic properties including nano-dimension, high surface

area (100 m2 g-1)(Islam et al., 2013; Savadekar & Mhaske, 2012; Silvério et al., 2013), high
A
M

aspect ratio of 100 (Rosa et al., 2010; Savadekar & Mhaske, 2012; Tee et al., 2013), high

crystallinity, low density, high mechanical strength, unique morphology along with availability,
D

renewability and biodegrability (Azizi Samir et al., 2005; Ilyas et al., 2017; H. Ng et al., 2015).
TE

Cellulose is the product of biosynthesis from bacteria and plants, whereas the general term

‘”nanocrystalline cellulose’’ refers to cellulosic isolation or extraction materials, with the


EP

outstanding feature of nano-scale structural dimension. The main component of plant fibres is
CC

cellulose, semicrystalline polymer, which composed of poly(1,4-β-D-anhydroglucopyranose)

units. These units are formed from strong hydrogen bond between hydroxyl groups. Other main
A

components that made up natural fibres structure are lignin and hemicellulose. Lignin is a highly

cross-linked phenolic polymer, whereas hemicellulose is a branched multiple polysaccharide

polymer composed of different types of sugars comprising xylose, glucose, arabinose, mannose

and galactose. However, both lignin and hemicellulose are amorphous polymers. Strong acid
hydrolysis is a well-known process for the removal of the amorphous region and isolation of

NCCs from natural fibers. Under controlled conditions, acid hydrolysis allows removal of the

amorphous regions of cellulose fibres whilst keeping the crystalline domains intact in the form of

crystalline nanoparticles. In fact, removing the amorphous region influences the structure,

PT
thermal stability, crystallinity as well as surface morphology of the fibres. Previous studies done

by Deepa et al., (2011a) and Li et al., (2009), have shown improvement in the thermal stability

RI
and crystallinity during the isolation of NCCs.

SC
In the past decades, many different resources have been used to prepare NCCs, such as hard- and

U
softwood fibres (Beck-Candanedo et al., 2005), wheat straw (Dufresne et al., 1997), sisal (Garcia

N
de Rodriguez et al., 2006; Morán et al., 2008), pineapple leaves (Cherian et al., 2010a), coconut
A
husk fibres (M. Rosa et al., 2010), mulberry (Li et al., 2009), bananas (Deepa et al., 2011a),
M

sugarcane bagasse (de Morais Teixeira et al., 2011), bamboo (Brito et al., 2012), mengkuang

leaves (Sheltami et al., 2012), jute (Cao et al., 2012), rice straw (Lu & Hsieh, 2012), eucalyptus
D

wood (Tonoli et al., 2012), soy hull (Flauzino Neto et al., 2013), cotton linter (Morais et al.,
TE

2013) and kenaf bast (Zaini et al., 2013). The purpose of the isolation of NCCs is as
EP

reinforcements in the field of nanocomposite that has gained tremendous attention since it was

first examined by Favier et al. (1995). However, no studies on the production, composition, or
CC

properties of natural nanocrystalline cellulose fibres from sugar palm fibres have been found in

the literature.
A

Sugar palm tree also known as Arenga Pinnata is a popular multipurpose tree dominantly found

in tropical regions. It belongs to the Palmae family which has about 181 genera with around
2600 known species (Ishak et al., 2013). The fruit can be eaten as sweet meal and the fibres can

be used for weaving hat and mats, making ropes, brooms, road construction, brushes, roof

materials, cushion and shelters for fish breeding. Besides that, its stem core can be used for

making sago flour and root as tea to cure bladder stones; insect repellent and posts for pepper,

PT
boards, tool handles, water pipes, musical instruments like drums (Adawiyah et al., 2013; Ishak

et al., 2013; Martini et al., 2012; Mogea et al., 1991). Up to the present time, the usage of sugar

RI
palm fibres has progressed to another successive level especially to numerous engineering

SC
applications. In example, it is being used for underground and underwater cables, substitution of

geo-textile fiberglass reinforcement in road construction for soil stabilization as well as a used as

U
reinforcement in polymer matrix composite in material engineering (Ishak et al., 2013). Several

N
studies have shown that sugar palm fibres have a huge potential to be used in various polymer
A
composite application (D. Bachtiar et al., 2008, 2010, 2011; Ishak et al., 2009; Sahari et al.,
M

2011, 2013a, 2013b). This study continues with sugar palm-derived cellulose reinforced with

starch polymer (Sanyang et al., 2016). To the best of our knowledge, no study on sugar palm
D

nanocrystalline cellulose (SPNCCs) has been found in the literature. Thus the aim of the current
TE

study was to extract and characterize nanocrystalline cellulose (NCCs) from sugar palm fibres.
EP

In this paper, cellulose and NCCs were extracted from sugar palm fibres by chemical and

mechanical methods. The effects of different chemical treatments on sugar palm fibres were
CC

investigated by determining their chemical composition (TAPPI standard), thermogravimetric

analysis (TGA), X-ray diffraction (XRD), surface area by Brunauer-Emmett-Teller (BET),


A

degree of polymerization (DP), Fourier transform infrared (FT-IR) spectroscopy and field

emission scanning electron microscopy (FESEM). The aspect ratio and dimensions of the
isolated NCCs were determined by using zeta potential nanoparticle sizer, atomic force

microscopy (AFM) and transmission electron microscopy (TEM).

2. Experimental

2.1. Materials

PT
Sugar palm fibres gathered in Bahau (Negeri Sembilan, Malaysia) were used in this study. The

RI
chemical reagents used were sodium chlorite, acetic acid, sodium hydroxide and sulphuric acid

SC
(purchased from Sigma–Aldrich).

2.2. Extraction of cellulose

U
Cellulose fibres can be extracted from sugar palm fibres (SPF) using two main processes, which
N
are delignification and mercerization (Ilyas et al., 2017; Sanyang et al., 2016; Tawakkal, Talib,
A
Abdan, & Ling, 2012; Y. B. Tee et al., 2013). The initial process was performed in accordance
M

with ASTM D1104-56 (1978) to prepare holocellulose through a chlorination or bleaching


D

process, primarily designed for the removal of lignin from SPF. 20 g of SPF was rinsed with tap
TE

water to remove foreign particle and dust. The clean SPF then soaked in a 1000 ml beaker

containing 650 ml of hot distilled water with temperature of 95oC, which then was transferred to
EP

a water bath, and the temperature was set at 70 ºC. 4 mL of acetic acid and 8g of sodium chlorite
CC

was added to the beaker every hour for 7 h, consecutively. The changing color of the SPF from

light brown to white indicated the level of delignification. The obtained celluloses are referred to
A

as holocellulose and were filtered, washed and rinsed with distilled water.

The holocellulose were further treated to produce α-cellulose according to ASTM D1103-60

(1977). The holocellulose were then soaked in 500 mL of 5% w/v NaOH solution for 2 h at 23±2

ºC. The cellulose produced were filtered and immerged in 500 mL of distilled water containing
approximately 7 mL of acetic acid to neutralize the cellulose. Then, the mixture was stirred for

approximately 30 s before it was allowed to settle for 5 min. It is then rinsed with water until the

cellulose residue was free from acid as indicated by a pH meter. Lastly, the cellulose denoted as

sugar palm cellulose (SPC) was dried in an oven at 103 ̊ C overnight.

PT
2.3. Isolation of Sugar palm nanocrystalline cellulose (SPNCCs)

RI
SPNCCs were prepared by acid hydrolysis of the cellulose obtained as described elsewhere

SC
(Sanyang et al., 2016). Acid hydrolysis was carried out using 60 wt% H2SO4 solution and kept

stirring with mechanical stirrer 1200 rpm for 45 minutes at 45 ºC. The time of hydrolysis in this

U
study was fixed at 45 min, which was found to be the optimum time (Bondeson, Mathew, &
N
Oksman, 2006a). The ratio of the obtained cellulose to liquor was 5:100 (wt %) (5 gram
A
cellulose: 100 gram H2SO4 solution). The hydrolyzed cellulose samples were washed four times
M

by centrifugation (6,000 rpm, 20 min, and 10 ̊ C) to eliminate the leftover sulphuric acid. The
D

suspension was then dialyzed against distilled water until a constant pH was reached (6.5 to 7).
TE

Then, the resultant suspension was sonicated for 30 min. Lastly, the suspension was freeze-dried

at -110 ºC using ethylene gas and stored at cool place prior to sample analysis (Bondeson et al.,
EP

2006a; Sheltami et al., 2012).


CC

2.4. Characterization

2.4.1. Determination of chemical composition


A

The chemical compositions of sugar palm fibres were determined at different stages of treatment

such as raw fibres, bleached fibres and alkali-treated fibres. The percentage of holocellulose was

determined according to the method described by Wise, Murphy, & D’Addieco, (1946).
Determination of lignin (acid insoluble) and α-cellulose and contents were determined according

to TAPPI standard methods T 222 (acid-insoluble lignin in wood and pulp) and T 203 (alpha-,

beta- and gamma-cellulose in pulp), respectively. It should be noted that the procedure used for

this standard has been modified by the Institute Of Tropical Forestry And Forest Products,

PT
Universiti Putra Malaysia (INTROP-UPM).

2.4.2. Field emission scanning electron microscopy (FESEM)

RI
Field emission scanning electron microscopy (FESEM) micrographs were taken using an FEI

SC
NOVA NanoSEM 230 machine with an accelerating voltage of 3 kV to observe the micro and

U
nanostructure surface of the longitudinal cross section of dried sugar palm fibres, as well as the

N
fibres after different stages of treatment and SPNCCs. Besides, to avoid over charging, all
A
samples were coated with gold (Hajalilou, Hashim, & Mohamed Kamari, 2014; Sheltami et al.,
M

2012).

2.4.3. Transmission electron microscopy (TEM)


D
TE

The transmission electron microscopy (TEM) machines are used to produce nanostructure

images of the SPNCCs. The Philips Technai 20 machine with acceleration voltage of 200 kV and
EP

standard inclined sample holder produces the TEM images was used. Firstly, a suspension of

SPNCCs were dispersed by ultrasonicating apatite particles in distilled water for 10 minutes.
CC

After that, one drop of a diluted suspension of SPNCCs were placed on a carbon coated copper

grid, then allowed to dry at room temperature. The TEM nanostructures images were enhanced
A

to improve the resolutions (Hofmeister & Platen, 1992).

2.4.4. Density
The density was measured using gas intrusion under a helium gas flow with an AccuPyc 1340

pycnometer. The samples are being oven dry at 105 ºC overnight to remove moisture content

within the fibres. After that the samples were place inside the desiccator to remove traces of

water from almost-dry sample before placed inside the pycnometer. Five measurements were

PT
conducted at 27°C and the average value was computed (R. Jumaidin et al., 2017).

2.4.5. Moisture content

RI
Five samples were prepared for the moisture content investigation. All samples were heated in an

SC
oven for 24 hours at 105 ºC. Weight of samples before, Mi and after, Mf the heating were

U
measured in order to calculate the moisture content (R. Jumaidin et al., 2017). Moisture content

was determined by using Eq. (1).


N
A
𝑀𝑖 −𝑀𝑓
Moisture content (%) = × 100 (1)
𝑀𝑖
M

2.4.6. Yield
D

Yield was calculated as percentage, % of initial weight of sugar palm cellulose (SPC) after
TE

hydrolysis. The suspension of the SPNCCs gained after dialysis treatment was freeze-dried and

compare to the initial weight of SPC. Weight of samples final of SPNCCs, Mf and initial of SPF,
EP

Mi were measured in order to calculate the yield (Bondeson et al., 2006a). Yield was determined
CC

by using Eq. (2).

𝑀
Yield (%) = 𝑀𝑓 × 100 (2)
A

2.4.7. Fourier transform infrared (FT-IR) spectroscopy


Fourier transform infrared (FT-IR) spectroscopy was used to detect possible changes in the

functional groups existing in sugar palm fibre at different stages of extraction. Spectra of the

material were obtained using an IR spectrometer (Nicolet 6700 AEM). FT-IR spectra of the

samples (10 × 10 × 3 mm) was collected in the range of 4000 to 500 cm−1. Ground samples

PT
were mixed with KBr, after which the mixture was pressed into thin transparent films that were

analysed (R. Jumaidin et al., 2017a, 2017b).

RI
2.4.8. X-ray diffraction (XRD)

SC
Rigaku D/max 2500 X-ray powder diffractometer (Rigaku, Tokyo, Japan) equipped with CuKα

radiation (λ = 0.1541 nm) in the 2θ range 10-40o was used to study X-ray diffraction patterns of

U
N
the raw, bleached, alkali-treated fibres and hydrolysis of sugar palm fibres . Then empirical
A
method (Segal, Creely, Martin, & Conrad, 1959) was used to obtain the crystallinity index of the

samples 𝑋𝑐 , as shown in Eq. (3):


M

𝐼002 −𝐼𝑎𝑚
𝑋𝑐 = × 100 (3)
D

𝐼002
TE

where 𝐼002 and 𝐼𝑎𝑚 are the peak intensities of crystalline and amorphous materials, respectively.
EP

2.4.9. Zeta Potential

The zeta potential measurements were performed using a Zetasizer Nano-ZS (Malvern
CC

Instruments, Worcestershire, UK) to determine the approximate size of SPNCCs and to

characterize the surface charge property of nanoparticles. A tenfold dilution of the samples in
A

pure water in a total volume of 1 mL was subjected to a particle size analyzer at 25°C. The

measurement was based on the electrophoretic mobility (μm/s) of the particles which was
converted to zeta potential by inbuilt software based on the Helmholtz–Smoluchowski equation

(Morais et al., 2013).

2.4.10. Thermogravimetric analysis (TGA)

The thermal degradation behaviour of composites was analyzed by TGA with respect to weight

PT
loss due to increase in temperature. TGA was performed with a Q series thermal analysis

RI
machine from TA Instruments (New Castle, DE, USA) to determine the thermal stability of

sugar palm fibres at different stages of extraction. The analysis was carried out in aluminum pans

SC
under a dynamic nitrogen atmosphere in temperature range 25–600 oC at a heating rate of 10 oC

/min-1 (R. Jumaidin et al., 2017a, 2017b).

U
2.4.11 Surface area and porosity measurements N
A
The surface areas, pore sizes and pore size distribution were measured by Brunauer-Emmett-
M

Teller (BET) method or N2 adsorption-desorption at 77 K by using a surface area and porosity


D

analyszer BELSorp Mini II (NIKKISO, Osaka, Japan). The samples were degassed at 105 oC
TE

under vacuum for 10 h. Specific surface areas were derived from the linear region of the

isotherms using the Brunauer-Emmett-Teller (BET) equation in a relative P/Po pressure range
EP

10-2 to 1, whereas pore size distribution were calculated from adsorption branch of the isotherms

by Barrett-Joyner-Halenda (BJH) method. The total pore volumes were estimated from the
CC

amount absorbed at a relative pressure of P/Po= 0.98.


A

2.4.12 Atomic force microscopy (AFM)

The AFM measurement were conducted using Dimension Edge with High-Performance AFM

(Bruker, Santa Barbara, CA, USA) equipment and Bruker Nanoscope analysis software (Version

1.7) operated using Peak/ Force tapping mode with one controller (Nanoscope V from Bruker)
for evaluating the evaluating the thickness of the SPNCCs. Initially, a drop of diluted SPNCCs

aqueous suspension was deposited on the surface of an optical glass slide and allowed to air dry.

The SPNCCs were scanned in air at room temperature of 25 oC and relative humidity in tapping

mode with OMCL-AC160TS standard Si probes (tip radius<10 nm, spring constant = 2.98 N/m,

PT
resonant frequency = ~310 kHz) under a 1 Hz scan rate.

2.4.13 Degree of polymerization (DP)

RI
The degree of polymerization (DP) of SPF, SPBF, SPC and SPNCCs fibres suspension was

SC
determined based on the intrinsic viscosity [𝜂]. Viscosity measurement for fibres suspension was

U
carried on according to TAPPI Standard Method T230 om-08 and ISO 5351-1 as reported by

N
Chauve et al. (2013). Fibres were diluted in solutions containing distilled water and copper (II)
A
ethylenediamine (CED) solution as dissolving agent with a ratio 0.01:1:1 (treated fibre: distilled
M

water: CED). The solution was shaken until all the fibres were completely dissolved. The

viscosity of this solution and the solvent was measured at 25 °C using Ubbelohde viscometer
D

tube (Type 231, PTA Laboratory Equipment, Vorchdorf, Austria). The experiment was
TE

performed for all the samples and repeated three times. The molecular weight of treated fibre

was calculated using the Mark-Houwink approach, which was using Eq. (4),
EP

[η]=KMα (4)
CC

where [η] the intrinsic viscosity and M is the molecular weight. The values of the constants were

taken as K=0.42 and α=1 for the CED solvent (Ilyas et al., 2017; Yasim-Anuar et al., 2017).
A

3. Results and Discussion


The chemical compositions of the sugar palm fibres were affected by the chemical extraction

process, NaClO2 followed by NaOH. From raw sugar palm fibres as feedstock, the initial stage

of the extraction process was bleaching treatment. This treatment was done to remove lignin

(Dufresne et al., 2000; Sheltami et al., 2012). The bleaching treatment effects are depending on

PT
the temperature and pH. Usually, the reaction happens at high temperature and low pH. The

second stage was alkali treatment that was done to remove and hydrolyze the hemicellulose,

RI
silica, soluble mineral salts, and ash (Deepa et al., 2011b; Ndazi et al., 2007). The morphology of

SC
the raw and treated sugar palm fibres were influenced by this alkali treatments.

3.1. Chemical Composition

U
N
The chemical compositions of the sugar palm fibres at different stages of treatment are shown in
A
Table 1. From the Table 1, it shows that the original sugar palm fibres consist of 43.88%
M

cellulose, 7.24% hemicellulose, 33.24 % lignin, 2.73 % extractive and 1.01% ash. After treating

the sugar palm fibres with NaClO2 solution, the cellulose content was increased by 12.79%,
D

whereas the lignin was reduced by about 32.97 %. Lignin was removed in large amount after 7th
TE

round of bleaching, with duration of each round is 1 hour. The usage of NaClO2 for the 1st stage

is an excellent choice for lignin removal as natural fibres can be considered as composite of
EP

hollow cellulose fibrils collected by lignin as binder in a hemicellulose matric. Hence, the
CC

removal of the hemicellulose after the lignin removal would be eased. The extensive reduction of

the hemicellulose content in fibres occurred in alkali treatment, which was reduced to 3.97%
A

from 19.8%. Lignin was solubilized by bleaching treatment, whereas dilute alkali treatment

solubilized the hemicellulose and remaining lignin in the fibres. This was affected by cleavage of

the ester-linked substance of hemicellulose, where this treatment increased the surface area of the

sugar palm fibres to make polysaccharides more vulnerable to hydrolysis. Besides that, it is clear
from Table 1 that the chemical treatment removed most of the lignin and hemicelluloses from the

sugar palm fibers. The lignin content was decreased from 33.24% to 0.06% and hemicellulose

from 7.24 to 3.97% whereas α-cellulose content was increased from 43.88% to 82.33%. The

changes of chemical composition of sugar palm fibres after all treatment resulted in a better

PT
crystalline degree of cellulose and therefore improved the strength and thermal properties of the

fibers (Alemdar & Sain, 2008).

RI
3.2. Morphological analysis of sugar palm fibres, treated fibres and SPNCCs

SC
3.2.1 Morphological analysis of sugar palm fibres

U
The bleaching and alkali treatments not only resulted in changes of chemical composition

N
of the treated fibres, they also affect the structure of the fibres surfaces. Fig. 1 shows the sugar
A
palm tree and its fibres at different stages of treatment. The color of the sugar palm fibres
M

changed from black (Fig. 1d) to light brown after bleaching treatment (Fig. 1e) and became

white after alkali treatment (Fig 1 f). FESEM pictures of sugar palm fibres were taken to
D

investigate the structure of sugar palm fibers to reveal their homogeneity and micrometric
TE

dimensions, and are also shown in Fig. 1. Microscopic examination of the longitudinal and cross
EP

section of sugar palm fibres was depicted in Fig. 1b,c,g. Fig. 1g indicates sugar palm fibres

(SPF) with approximate diameter size of 212.01 ± 2.17 µm, in the original form were bonded
CC

together by cement components known as middle lamella, which were partially diminished after

the bleaching treatment (Fig. 1h). As pointed out by Alemdar & Sain, (2008) and M. F. Rosa et
A

al., (2010) worked with wheat straw and soy hull, and coconut fibres, respectively, these images

shows the partial removal of impurities pectin, lignin and hemicellulose after chemical treatment,

which they are acting as cementing components surround fibre-bundles.


As seen in Fig. 1b,c, the view from the outer to the inner part showed that sugar palm fibres

consist of a middle lamella, a primary, a secondary and a tertiary cell wall, build up around an

opening, the lumen (Bledzki & Gassan, 1999). The middle lamella as seen in Fig. 1c, which is

surround the cell wall is mainly composed of pectins (macromolecules of galacturonic acid) that

PT
hold fibres together into a bundle, with a size around 1.98 ± 0.15 µm.

The interior of the sugar palm fibres consist of primary cell wall, secondary cell wall, and lumen.

RI
Primary cell wall is made of cellulose (a polymer baesd on glucose units) fibrils in an organic

SC
matrix of amorphous hemicelluloses and lignin, proteins, and low methylesterified pectins, with

the average diameter being around 10.38 ± 0.57 µm. The secondary cell wall consist of three

U
layers of cellulose fibrils with different axial orientation that are bound by lignin (Mohanty et al.,
N
2000). Primary cell wall and secondary cell wall also provide mechanical support to the plant.
A
Fig.1c shows a lumen with a thickness of around 3.72 ± 0.15 µm.
M

From Fig. 1g, it can be seen that the longitudinal section of SPF surface topography is rough,
D

with pore-like spots that appear in almost regular intervals, where these similar spots were also
TE

found at the surface of coir fibres (Ticoalu, Aravinthan, & Cardona, 2013). These pore-like

spots are known as tyloses, which their function is to cover the pits on the cell walls.
EP

Besides that, based on the observation on sugar palm fibres microstructure, there is strong
CC

evidence that the acid/ bleaching treatment (NaClO2) changed the physical surface appearance of

sugar palm fibres compared to raw sugar palm fibres (Fig. 1a, h). It showed that a drastic
A

physical changes with slight fibrillation was clearly observed on the outer surface of the fibres

where the outer surface became clear and this was due to the removal of the waxy layer on the

outer surface. The average diameter of bleached fibres was reduced after chemical treatment

from 212.01 ± 2.17 µm to 121.80 ± 10.57 µm. These were measured after the partial removal of
lignin and hemicelluloses lignin, in another words, because of the elimination of the primary

walls, which was further supported by chemical analysis data given in Table 1. Furthermore, the

bleaching treatment removed the extractives from the fibres indicated by the surface changing.

After alkali treatment (Fig. 1i), the fibre bundles where dispersed into individual fibres with

PT
diameters in the range 11.87 ± 1.04 µm. Compared with raw sugar palm fibres, bleached fibres

and alkali-treated fibres was almost double and eighteen time smaller than raw sugar palm fibres,

RI
respectively. The removal of lignin and hemicellulose through the process of delignification and

SC
mercerization of raw SPF had caused drastic reduction in the diameter of bleached and alkali-

treated fibres, besides the surface of SPC (Fig. 1i) had changed to smooth and groovy surface

U
with parallel arrangement along the cellulose. The diameter of SPC obtained is supported with

N
the average diameter of kenaf-derived cellulose (13 µm) and cellulose microfibres reported by
A
Sonia et al. (2013) and Tawakkal et al. (2012), respectively. Besides that, similar reports have
M

been documented on the surface appearance of many natural fibre-derived cellulose (Sanyang et

al., 2016; Tawakkal et al., 2012; Y. B. Tee et al., 2013). The chemical treatments used on the
D

fibres also affected the characterization of the separation of micro-sized fibres from the fibres
TE

bundle into individual micro-sized fibres.


EP

3.2.2 Isolation of SPNCCs


CC

Delignification and mercerization resulted in partial defibrillation and opening of fiber-bundles.

Removing the cementing components and defibrillation via bleaching-treatment and alkali
A

treatment are important steps towards producing more efficient SPNCCs hydrolysis. Each

microfibril bundle is composed of a bundle of NCCs interconnected along the microfibril by

amorphous region (defected structure). Under the hydrolysis process, these regions were

eliminated to yield highly NCCs, resulting in a diameter reduction and separation of fibres from
agglomerates micron-sized fibres to individual nanofibres (Azizi Samir et al., 2005; Bondeson et

al., 2006a). Usually these NCCs range in length and diameter from 100-250 nm and 5-80 nm,

respectively, for most cellulosic material, subsequently become more or less individualized thru

sonification (Bondeson et al., 2006a).

PT
The yield of SPNCCs was about 29 % after 45 min of hydrolysis treatment (of initial weight).

The obtained yield for SPNCCs is an agreement with the yield of sisal (30 %) and mengkuang

RI
leaves (28%) reported by Garcia de Rodriguez, Thielemans, & Dufresne, (2006b) and Sheltami

SC
et al., (2012), respectively. In a separate investigation, de Morais Teixeira et al.,(2011) also

reported that the yield of sugarcane bagasse NCCs is 58 %, which was twice higher than that the

U
yield SPNCCs. The differences of the yield gained is depending on the source sample,
N
pretreatment and hydrolysis condition (Eichhorn et al., 2010; H. M. Ng et al., 2015). It is well
A
known that the advantage of using sulphuric acid for hydrolysis treatment is, this chemical
M

contribute to the isolated crystalline particle and stable aqueous suspensions of NCCs, due to
D

negatively charged, hence do not tend to agglomerate. In more detail, these charged induce
TE

electrostatic repulsion forces between nanoparticles, thus lead to a stable suspension (Bondeson,

Mathew, & Oksman, 2006b; Samir, Alloin, Paillet, & Dufresne, 2004). This may be attributed to
EP

the process of esterification on the surface of hydroxyl groups from cellulose during hydrolysis

process, and as a consequence, the sulphate groups are introduced in the outer surface of NCCs
CC

(Habibi, Lucia, & Rojas, 2010). Moreover, the presence of sulphate group also has been verified
A

to decrease the material thermal stability (Roman & Winter, 2004a). Hence, all these properties

are the important key factor when NCCs are incorporated as nanoreinforcement within

nanocomposite.
The resultant suspension of isolated SPNCCs prepared from cellulose sugar palm fibre is shown

in Fig. 2a. The concentration of this suspension was 2 wt%. FESEM, TEM and AFM observation

and the distribution of length and diameter of SPNCCs are shown in Fig. 2b, c, d, e, f, g, h and i,

respectively. The SPNCCs were characterized and analyzed by particle length (L), diameter (D)

PT
and aspect ratio (L/D). The diameters and length of the fibers after hydrolysis treatments were

calculated by an image processing analysis program, Image J, using the FESEM, TEM and AFM

RI
images. The profile analysis was carried out with peak fitting program using Gaussian line shape

SC
to define the average diameter and length of SPNCCs. Sugar palm NCCs ranged in length from

45 to 238 nm, with an average value around 130 ± 30.23 nm. The diameter was in the range 3–

U
18.19 nm, with an average value around 9 ± 1.96 nm. The diameter measured were similar to the

N
nano-sized structures that were extracted from other agro-residue sources such as coconut husk
A
(5.5 ± 1.5 nm) (M. Rosa et al., 2010), rice straw (5.06 nm) (Thiripura Sundari & Ramesh, 2012),
M

soy hulls (4.43 ± 1.20 nm) (Flauzino Neto et al., 2013), sugarcane bagasse (4 ± 2 nm) (de Morais

Teixeira et al., 2011), banana residues (5nm) (Zuluaga et al., 2007) and smaller than microfibrils
D

from wheat straw (10–80nm) (Alemdar & Sain, 2008), sisal fibres (30.9 ± 12.5 nm) (Morán et
TE

al., 2008) and sugarcane bagasse (30nm) (Bhattacharya, Germinario, & Winter, 2008) as shown
EP

in Table 2. These resultant images revealed the efficiency used of the acid hydrolysis treatment

on sugar palm fibres and endorsed that the aqueous suspensions contained sugar palm NCCs
CC

residing mostly of individual crystal and some aggregates.


A

The optimum aspect ratios (length to diameter) of sugar palm NCCs was calculated to be 14. The

aspect ratios of sugar palm NCCs are not as much different from each other as their actual

dimensions, but quite low compared to the value of those reported on NCCs from other cellulosic

sources, such as sisal (60) (Garcia de Rodriguez et al., 2006a), cotton (15) (Qi, Cai, Zhang, &
Kuga, 2009), wood (35–38)(Roman & Gray, 2005), ramie (25–42)(Habibi et al., 2008), sisal (43)

(Siqueira, Bras, & Dufresne, 2010), and tunicate (251–493)(Iwamoto, Kai, Isogai, & Iwata,

2009).

Correspondingly to the TEM sample preparation, the AFM sample preparation can also cause the

PT
nanocrystalline cellulose to aggregate in their images. However the AFM analysis allows the

discernment of individual NCCs of agglomerated structure through transverse height profile.

RI
Moreover the detailed structure, topography, thickness and roughness of the NCCs also could

SC
gained from the AFM for the purpose of NCCs manufacturing process or functional application

developer (Flauzino Neto et al., 2013). The ultrafine fibres of SPNCCs were examined by an

U
AFM scanning probe microscope that used only moderate mechanical force and does not
N
required sputter coating. According to Aziz-Samir et al., (2005), the function of the sulphuric
A
acid hydrolysis is to cleave the amorphous region of microfibrils transversely, which resultant in
M

reduction of the fibres diameter from micro to nanometers. Fig 2e shows the AFM image of
D

sugar palm nanocrystalline cellulose (SPNCCs) prepared by the treatment of sugar palm
TE

cellulose fibres with sulphuric acid solution, signifying the dominant diameter ranged size of

SPNCCs was 10.7 nm± 2.34 nm, which is similar to those obtained from wood pulp, cotton and
EP

tunicin, probably because of the aggregation of monocrystals. These images also displayed
CC

needle-like nanoparticle. Besides, Fig. 2e also indicated the larger fibre width is due to the tip

coardening effect, which was common in AFM (Goetz et al., 2010; Lu & Hsieh, 2012). The
A

AFM phase image (Fig. 2c) showed a peak nanofibres height of 5.781 nm, comparable to the

average nanofibres diameter (10.7 ± 2.34 nm) determined from the AFM images. Fig. 2(g,i)

shows the histograms corresponding to these measurements. The AFM image also showed that

the length of the SPNCCs after the sulphuric acid hydrolysis, was around 130±16.87 nm.
Besides, these fibre height and width value gained from the analysis confirmed the essentially

cylindrical cross-sectional shape of SPNCCs self-assemble fibres. The average value of aspect

ratio obtained by AFM measurements was 13, which lies in the range of long nanocrystalline

cellulose which have great potential to be used as reinforcing agents in nanocomposites (M. Rosa

PT
et al., 2010).

3.3 Physical Properties

RI
One of the most important characteristics that must be considered for a new natural materials as

SC
potential filler, especially for polymer composite, is moisture content. Therefore, it is an

U
important parameter used to determine the properties and the end uses of the fibres. Low

N
moisture content is required, as high moisture content of the fibres could weaken the stability of
A
the polymer composites in terms of porosity formation, dimensions, and tensile strength (R.
M

Jumaidin et al., 2017; Razali et al., 2015). This can be explained with the some alkali labile

linkages (ester linkages and ether linkage) between lignin and polysaccharides that were broken
D

during the alkali treatment, which leaving a rigid structure (cellulose) that was durable to
TE

dissolve in the acidic solution. Therefore, as the percentage of the cellulose content in the fibres

increased, the moisture absorption rate also increased. This is due to the chemical structure of
EP

cellulose itself, which composed of hydroxyl groups that are accessible to water. Therefore, as
CC

related to the bleaching, alkali and hydrolysis treatment conducted on the fibres, the fibres

surface were induced to the abundant formation of hydrophilic ionic groups, which indirectly
A

stimulate the absorption of moisture. As a result, we can see the increment of the percentage of

moisture content from raw fibres (8.36 ± 0.0984 %) to SPNCCs (17.901 ± 0.0623 %).

Besides that, in material selection process, material weight is one of the crucial parameter that

have to be considered as it may affect the performance of the products. Moreover, the main
criteria that correlated to this material selection properties is density. The density value of raw

sugar palm fibre, bleached-treated, alkali-treated and hydrolysis fibres showed a decreasing

value throughout the treatment, which were 1.50, 1.30 ± 0.0023, 1.28 ± 0.0019 and 1.05 ± 0.0023

g/cm-3, respectively. This decreasing trend might be attributed to the removal of main component

PT
of the fibres such as lignin and hemicellulose from the fibres, the density of the fibres decreased,

as shown in Table 2. The removal of an amorphous non-cellulosic compounds component in

RI
bleaching and alkali treatment created voids in the structure of fibres, led swelling of fibres to

SC
occur. The constituent of fibres then became well separated. The increasing in volume with loss

in weight caused density value to decrease (Ray & Sarkar, 2001). Besides that, the density values

U
of SPNCCs were lower compared to conventional manmade fibres such as glass fibre (2.5

N
g/cm3), aramid (1.4 g/cm3) and carbon (1.7 g/cm3) (Mohanty et al., 2000).
A
Moreover, fibre density is also related to the surface area and the porosity of fibres. The density
M

of the fibre decrease as the pore volume of the fibre increase. The surface areas and porosities of
D

the macro fibres and self-assembled nanofibres were determined by N2 adsorption-desorption


TE

method 77K. Fibres from all treatment exhibited type IV isotherms accompanied by hysteresis

loop, which is associated with capillary condensation taking place in mesopores (2-50 nm). It is
EP

also represent the limiting uptake over a range of high P/Po. Moreover, these nanofibres shows

an increment in cumulative pore volumes, i.e., 0.061 cm3/g for SPF, 0.067 cm3/g for SPBF,
CC

0.195 cm3/g for SPC and 0.226 cm3/g for SPNCCs. The cumulative pore volume for
A

nanocrystalline cellulose (0.226 cm3/g) is 4 times larger than that for the raw sugar palm fibres

(0.061 cm3/g). The smaller cumulative pore volume of the raw sugar palm fibres was expected

with the rigid, closely aligned and tightly bound building elements via strongly hydrogen bonded

cellulose structures, thus leaving little interfacial spaces (Lu & Hsieh, 2012). The result of low
pore volume was confirmed by the FESEM, TEM and AFM observation (Fig. 1 and Fig. 2).

However, the pore volume of nanocrcystalline cellulose increased after loosely packed structures

took shape to create a huge amount of mesopores among the SPNCC. The similar trend was

observed with The Brunauer-Emmett-Teller (BET) surface area for SPF (7.578 m2/g), SPBF

PT
(10.351 m2/g), SPC (13.1815 m2/g) and SPNCCs (14.474 m2/g). The Brunauer-Emmett-Teller

(BET) surface area of nanosize fibres from SPNCCs is 14.474 m2/g, 2 times of those from SPF

RI
(7.578 m2/g). The higher specific surface area of the nanofibres assembled from SPNCCs is

SC
mainly due to the smaller fibres sizes than that of SPF.

Degree of polymerization (DP) is a crucial parameter for evaluating the length and branching of

U
cellulose chains, besides, to evaluate the influence of hydrolysis in the cellulose chains that are
N
present in nanocellulose. Besides that, it has been stated by Audrey et al. (2007), that DP and
A
molecular weight may affect the properties of cellulose such as spinnability, solubility, and the
M

mechanical properties of cellulose based materials. The degree of polymerization and viscosity-
D

average molecular weight of the treated fibres were determined using an intrinsic viscosity
TE

measurement. Table 2 gives the DP value for the various treated fibres prepared in this study.

The degree of polymerization of the SPBF, SPC and SPNCCs were 2963.33, 946.48 and 142.86,
EP

respectively, and the molecular weight were 480,513.39 g/mol, 153,458.51 g/mol, and 23,164.7

g/mol, respectively.
CC

From the same table , it can be seen that the DP obtained for the SPC was almost similar to the
A

DP of bamboo cellulose (891) (Wang et al. 2010), curaua cellulose (989) (Corrêa et al. 2010), oil

palm cellulose (967) (Yasim et al. 2017). According to Yasmin et al. (2017) the molecular

weight and degree of polymerization of bio-cellulose based material were ranged from

approximately 90,000 to 300,000 g/mol and 400 to 3000, respectively. The sugar palm cellulose
were in the range of the cellulose DP, thus in good agreement with those reported in the

literature.

Besides, it has been reported by Kumar et al. (2009), that delignification by the acid-chlorite

process has a significant effect on the cellulose chain, in which it causes a huge reduction in the

PT
average degree of polymerization of treated fibres. Extensive delignification of SPBF process

reduced the DP to 2963, and further mercerization and hydrolysis process caused the DP to

RI
decrease to 946.48 and 142.86 (reduction percentage of 95.18%), respectively. The decreasing

SC
trend observed for the DP of SPBF, SPC and SPNCCs was attributed by the removal of lignin,

hemicellulose, and amorphous region via the delignification, mercerization, and hydrolysis of the

U
raw SPF, where the SPF bundle, SPBF and SPC fibres was cleaved and nanocrystalline cellulose

N
were released (Talib et al. 2011). Moreover, the DP reduction of cellulose was also due to the
A
acid catalysed cleavage during acid-chlorite delignification (Hubbell and Ragauskas 2010).
M

Besides, mercerization treatment promotes more reactive sites on the cellulose surface (Corrêa et

al., 2010). Thus, this caused acid hydrolysis treatment to be more effective, which led to major
D

level of breaking the cellulose chains. According to Pääkkö et al. (2008) and Corrêa et al,
TE

(2010), the sulphate group of strong acid leads to an extensive hydrolysis of amorphous phase
EP

and catalyse the degradation of the cellulose chain, subsequently reduce DP. DP obtained for the

SPNCCs was almost similar to the DP of curaua nanofibre (101.6) (Corrêa et al., 2010) and
CC

sugar beet pulp (120) (Habibi & Vignon, 2008).


A
3.3. FTIR spectroscopy analysis

Different stages of treatment on the polysaccharides (hemicelluloses and cellulose) and lignin of

sugar palm fibres were examined using FT-IR spectroscopy. The fingerprints of the functional

PT
groups obtained for sugar palm fibres at different stages of treatment (bleaching, alkali-

treatments and hydrolysis) are labelled in Fig. 3. The band located at 1719 cm−1 in the spectrum

RI
of raw sugar palm is attributed to C=O stretching of the acetyl and uronic ester groups of

SC
hemicellulose or the ester linkage of carboxylic groups of ferulic and p-coumaric acids of lignin

U
and/or xylan in hemicellulose (Alemdar & Sain, 2008; Fabiyi & Ogunleye, 2015; X. F. Sun, Xu,

N
Sun, Fowler, & Baird, 2005). This band was still present in the FT-IR of fibres after bleaching
A
treatment and no longer present in the FT-IR spectra of fibres after alkali and hydrolysis
M

treatment. The disappearance of this band could have been caused by the removal of

hemicellulose and lignin from sugar palm fibres during the chemical extraction (Alemdar &
D

Sain, 2008; Jonoobi, Harun, Shakeri, Misra, & Oksmand, 2009; Sheltami et al., 2012). This was
TE

supported by the results obtained from determination of the chemical composition of the fibres

(Section 3.1 and Table 1), in which they indicated that lignin was almost completely removed
EP

after the bleaching treatment. However, hemicellulose was not completely removed after the
CC

bleaching treatment and it was still remained after the alkali-treatment. For this reason, the

disappearance of the C= O stretching band from the spectrum could be caused by cleavage of all
A

ester-linked substances of the hemicellulose by alkali treatment. Ether bonds between lignin and

hemicellulose were not affected by this treatment (Sheltami et al., 2012; Xiao, Sun, & Sun,

2001).
The peaks in the 1520-1510 cm−1 are determined as aromatic skeletal vibration of the functional

group of lignin and lignocellulose. The band observed at 1227 cm−1, 1507 cm−1 and 1593 cm−1

are corresponds to lignin (Faix, Lin, & Dence, 1992; N., Jane, & K., 2012). These peak ceased

after the bleaching treatment, suggesting that this treatment was effectively remove lignin from

PT
the fibres. These results were supported by determination of the chemical composition of the

fibres as shown in Table 1. Besides, the peaks at 1475-1600 cm-1 signify the structural polymer

RI
of stretching of the aromatic groups present in form of lignin (Himmelsbach, Khalili, & Akin,

SC
2002; Sahari, Sapuan, Ismarrubie, & Rahman, 2012).

Based on Fig. 3 also, the absorbance peaked in the region between 1630–1650 cm−1 and around

U
2900 cm−1 that reflected the stretching of the O–H and C–H groups, respectively. The peaks in
N
the 3700-3100 cm−1 region were assigned to the adsorbed water. The absorbance bands around
A
897, 1030, 1160, 1316, 1370 and 1424 cm−1 were associated with the C–H rocking vibrations,
M

C–O stretching, C-O-C asymmetric valence vibration, C-H2 rocking vibration and C-H2
D

deformation vibration , respectively, where these peaks are referred to cellulose form of the
TE

carbohydrates (Alemdar & Sain, 2008; Fan, Dai, & Huang, 2012; Sheltami et al., 2012). These

different bands can be seen in all spectra, regardless of the purification of the fibres. The intense
EP

peaks at 1656 cm-1 signified C=C cis stretching of unsaturated acids or sterols correspond to

tannin (R. C. Sun & Tomkinson, 2002).


CC

The intense peaks at 3200 – 3500 cm-1 indicated the present of O-H groups in untreated and
A

treated fibres, due to presence of hydroxyl group in cellulose, hemicellulose and lignin. Peaks at

1800-1600 cm-1 signified carbonyl groups (C=O) in lignin and hemicelluloses (Kazayawoko,

Balatinecz, & Woodhams, 1997). The high peaks at 1000-1300 cm-1 was determined to be
existing in all 4 types of fibres, which signify C-H stretching and C-O groups, respectively (Faix

et al., 1992).

3.4. X-ray diffraction measurements

PT
Various ordered crystalline arrangements were resulted from both intra- and intermolecular

RI
hydrogen bonding occur in cellulose hydroxyl groups. Fig. 4 shows the XRD patterns for sugar

SC
palm fibres at different stages of treatment. The characterization of fibres polymoph can be

determined via diffractograms of XRD, in which it displayed a mixture of polymorphs of

U
cellulose I (typical peaks at 2θ=15º and 22.6º) and cellulose II (peaks at 12.3◦ and 22.1◦) (Klemm

N
et al., 2005). Fig. 4. also shows the value of crystallinity index for sugar palm fibres, bleached
A
fibres, alkali-treated fibres and SPNCCs. Major intensity peak that was identified in the X-ray
M

diffractograms which located at 2θ value of around 22.6º shown the crystalline structure of

cellulose I for all samples, whereas the low intensity peak at a 2θ value of around 18º was
D

labelled as amorphous region (Segal et al., 1959).


TE

The peak at around 2θ =22.6º is sharper for chemically treated than untreated sugar palm fibres.
EP

Crystallinity degree in the structure of the fibers can be identified by observing at the sharpness

of the diffraction peak. The sharper the diffraction peak, the higher the crystallinity degree of the
CC

fibres (Alemdar & Sain, 2008). The crystallinity indices of raw sugar palm fibre, bleached fibres,

alkali-treated fibres and SPNCCs were found to be 55.8%, 65.9%, 76.5% and 85.9%,
A

respectively. These results clearly demonstrated that the crystallinity of the material

progressively increases. The higher crystallinity value of SPNCCs compared to sugar palm fibres

can be well understood by the reduction and removal of amorphous non-cellulosic compounds,
such as lignin and hemicellulose induced by the bleaching, alkali and hydrolysis treatments

conducted in the purification process. The crystallinity index of the sample after hydrolysis

treatment, SPNCCs was higher than pineapple leaves 54% (Cherian et al., 2010b), soy hulls

73.5% (Flauzino Neto et al., 2013) and sisal 78% (Brito et al., 2012). However, it was lower than

PT
curaua 87% (Brito et al., 2012), bamboo 87% (Brito et al., 2012), eucalyptus 89% (Brito et al.,

2012) and cotton linter 90% (Morais et al., 2013). Thus, the crystallinity value is depending on

RI
different types of plants and the duration of fibres purification and hydrolysis process. There are

SC
relation between the number of crystallinity degree region and the stiffness of cellulose, where

increment in number of crystallinity region will raise the stiffness of fibres. Higher crystallinity

U
in the chemical treated fibers is interrelated with higher tensile strength of the fibers. Therefore,

N
the mechanical properties of the nanocomposite material can be improved by using these treated
A
fibers as nanofiller (Bhatnagar, 2005; Rong et al., 2001).
M
D

3.5. Particle size measurement and zeta potential


TE

Zeta potential (estimated as surface charge property of nanomaterial) and zeta sizer (particle size
EP

measurement) are analyzed using Zetasizer Nano-ZS (Malvern Instruments, Worcestershire,

UK). The formation and the size of the SPNCCs obtained were also checked by TEM. The result
CC

gained shows that the SPNCCs possessed a large negative charged, with mean value of -61.50 ±

1.65 mV. The strong negative zeta potential value was measured, revealed the existence of
A

negatively-charged SPNCCs in the suspension. These large negatively-charged zeta potential

values were obtained due to the conversion of the hydroxyl groups in to sulfate ester groups

(e.g., conversion of cellulose-OH to cellulose-OSO3−H+), on SPNCCs surface during sulphuric

acid hydrolysis treatment (Börjesson & Westman, 2015). The suspension of NCCs were
considered to be stable as the absolute value were larger than -27 mV (de Mesquita, Donnici, &

Pereira, 2010). The particle size distribution resulted in only one main group with 100% of the

particles were around 138.93 ± 13.93 nm. This light scattering technique cannot measure the size

of particle accurately and precisely as it measured the length and diameter dimension of the

PT
particle thoroughly. Nevertheless, the size distribution are ranging from 5.6 – 1106 nm are

considered similar to the TEM measurement (Fig. 2). According to Braun et al., (2008), these

RI
techniques could substitute TEM for good estimation of NCCs dimensions, only if there is

SC
adoption of adequate mathematical method. Thus, for SPNCCs, the Helmholtz–Smoluchowski

equation of general purposes was effective for measuring the SPNCCs crystals’ dimensions.

U
3.6. Thermogravimetric analysis
N
A
M

It is crucial to study the thermal properties of the natural fibres in order to determine their

compatibility as nanofiller for biocomposite processing, since the thermoplastic processing


D

temperature would arise above 200ºC. The thermogravimetric analysis data results obtained from
TE

sugar palm fibers, as well as those obtained after bleaching, alkali and hydrolysis treatment are
EP

shown in Fig. 5. Besides that, the Fig. 5. also illustrated both raw TG curves (panel i) and

derived curves (DTG – panel ii) as a function of temperature. Table 3 shows the comparison of
CC

percentage weight loses for untreated and treated sugar palm fibres.

Generally, the thermal decomposition of raw sugar palm fibres was divided into four phases. The
A

initial phase was evaporation of moisture content in fibres (occurred from 45 to ~ 123 ºC), next

decomposition of lignocellulosic components of hemicelluloses (occurred from 220 ºC to ~ 315

ºC) , followed by cellulose (occurred from 315 to ~ 400 ºC), lignin (occurred from 165 ºC to ~
900 ºC) and finally their ash (1723 ºC) as reported elsewhere (Ishak et al.,, 2012; Yang et al.,

2007).

The initial weight loss started around 42 ºC was caused by water evaporation in these samples.

As fibres are heated, there was reduction in weight of the material initially. This was due to the

PT
loss of water and volatile extractive, which they tend to move to the surface of fibres (M. F. Rosa

et al., 2010). The movement of volatile extractives happened due to the migration of water from

RI
lower interior part of fibres to the higher potential of water at the surface of fibres, as the water

SC
on the surface of fibres was evaporated. Therefore, the migration of water was indirectly carried

the volatile extractive and left it on the fibres surface (Ishak et al., 2012). It can be seen that the

U
evaporation of moisture of (a) raw sugar palm fibres was completely evaporated at 196.56 ºC,
N
compared to the treated fibres (134-184 ºC) as shown in Fig. 5. This is because of higher
A
moisture content (MC) of sugar palm fibres, resulted in higher mass loss (10.38%) compared to
M

the other treated fibres as shown in Table 1. As the fibres were being treated, the weight loss
D

were decreased due to lowered moisture content. Moreover, low mass losses can be seen at low
TE

heating temperatures, which mainly linked with only loss of water in cell lumen and in cell wall,

and volatiles extractives in the fibers, whereas huge mass losses that estimated up to 70% losses
EP

were expected to be in form of basic constituent; starting from hemicellulose, cellulose and

lignin where they started to decompose at temperature of 100 ºC and above (Ishak et al., 2012;
CC

Yang et al., 2007).


A

The DTG curve of original sugar palm fibres displayed an earlier weigh loss started at 210.58 ºC,

which then reached the highest peak at 281 ºC, due to the low decomposition temperature of

hemicellulose and lignin (Morán et al., 2008). Hemicellulose was very easy to get rid of from

main steam and subsequently degrade into CO2, CO, and some hydrocarbon, etc. at low
temperatures between 220 to 315 ºC. This is due to the composition of hemicellulose which

composed of various saccharides such as xylose, mannose, glucose, galactose, etc., and it’s

amorphous structure which appeared in random and rich of branches (Yang et al., 2007). On the

other hand, the cellulose fibers obtained by bleaching treatment showed a higher decomposition

PT
temperature around 345 ºC, which was almost similar than those reported by Yang et al (355 ºC)

(Yang et al., 2007).

RI
Table 3 shows the comparison of degradation temperature on maximum weight-loss rate (TMax),

SC
weight loss (WL) and char yield for treated and untreated sugar palm fibres obtained from the TG

and DTG curves. These outcomes clearly shows that the thermal stability of the sugar palm

U
fibres decreased after bleaching treatment and further increased after alkali-treatment. The 1st
N
degradation temperature was found to be 281, 271.56 and 346.09 ºC for the raw sugar palm
A
fibres, bleached-treated, and alkali-treated fibres, respectively. This is due to the removal of the
M

fiber structure, which facilitated degradation of the remaining hemicellulose and lignin
D

components from the fibres. Bleaching treatment also led to fibres fibrillation, i.e. break down of
TE

the fibres bundle into individual fibers (Deepa et al., 2011b). These process enhanced the

removal of fiber constituents like hemicelluloses, lignin, and waxes from the fiber, resulting in
EP

elevated surface destruction with well-defined fibril aggregates. The reduction in the value of

temperature of bleached fibres compared to raw sugar palm fibres was expected to occur due to
CC

the low lignin percentage in chemical composition of bleached fibres as related in Table 1. In
A

facts, lignin was the most difficult to decompose compared to cellulose and hemicellulose, which

the temperature of decomposition started early as 160 ºC and extends as high as 900 ºC (Yang et

al., 2007). Besides that, lignin also was a very tough component and responsible for providing

stiffness to the cell wall as well as serving to bond individual cell together in the middle lamella
region (Ishak et al., 2012). As the lignin is removed via bleaching treatment, the chemical

component remaining in the fibers are mostly composed of hemicellulose and cellulose. Thus,

the degradation process was easy to occur in alkali-treated fibers due to temperature degraded of

hemicelluloses are around 220-315 ºC as hemicellulose has low temperature degraded compared

PT
to other macromolecular components i.e: cellulose and lignin (Shafizadeh & Chin, 1977).

Throughout the decomposition of the (a) raw sugar palm fibres and (b) bleached fibres, a small

RI
peak appeared at 281 and 271.56 ºC, respectively, nearby to the main peak. This peak was

SC
attributed to the decomposition of hemicellulose and lignin, then disappeared, as expected, in the

case of the alkali-treated fibres (Fortunati et al., 2013). Moreover, the weight loss is much more

U
progressive than that of chemically purified samples. In addition, a shoulder is observed in the
N
DTG curve at around 300ºC, but is no longer present after the alkali treatment, which likely
A
reveals the removal of a partial of the hemicellulose (Sheltami et al., 2012).
M

Moreover, (c) alkali-treated fibres illustrated an increase in the mass loss temperature
D

(WL=73.71%; TMax=346.09 for 1st degradation) compared to raw sugar palm fibres (WL=43.76%;
TE

TMax=345.45 for 2nd degradation) and bleached fibres (WL=52.39%; TMax=324.44 for 12nd

degradation). The degradation of cellulose in the fibres was started within 315~ 400 ºC (Yang et
EP

al., 2007). This proved the capability of the cellulose to sustain the heat deformation
CC

temperatures. The alkali-treatment process helped to remove the lignin and hemicellulose after

bleaching treatment process as displayed in Table 1. The mechanism of alkali-treatment involved


A

the oxidation of hemicellulose and lignin which led to dissolution of both component by

breaking down the polysaccharides to simple sugars and hence released the cellulose fibers

(Deepa et al., 2011b). The dissolution of these various component via the bleaching and alkali

process produced α-cellulose as a residue material, which has been described to be crystalline
(Beckermann & Pickering, 2008). The higher the crystalline structure, the higher the degradation

temperature (Cherian et al., 2008). From Table 3, we can see that there were increment in the

degradation temperature in (c) alkali- and (d) hydrolysis treated fibres (1st and 2nd thermal

degradation, respectively), due to the higher crystallinity structure of fibres. This would

PT
indirectly enabled the fiber to endure severe heat and processing conditions. The hydrolysis

treatment was proved to reduce long micro fibres chain to nano-dimention effectively, by

RI
hydrolyzing the amorphous region. Thus, it can be summarized from these results that the

SC
production of SPNCCs displayed an enhancement of thermal properties of fibres than untreated

fibres, which then making them suitable as reinforcing materials in bio-renewable composite

U
material preparation.

N
The degradation temperature of sugar palm fibres after bleaching and alkali treatments was
A
around 200 ºC, which was lower than that of raw fibres. This result was similar as reported by Li
M

et al., (2009) and Soares, Camino, and Levchik (1995) for mulberry barks cellulose and pure
D

cellulose, respectively. The degradation temperature become lower due to the remained
TE

hemicellulose component after the chemical treatment in the fibres. Hemicellulose was located

between cellulose fibril. Thus, the strong bonding between hemicellulose and cellulose fibril
EP

resulted to decrease the crystallinity of cellulose fibrils and thus assessable the beginning process

of thermal degradation (Deepa et al., 2011b; Petersson, Mathew, & Oksman, 2009).
CC

This lower degradation temperature were caused by the hemicellulose component, which
A

remained after the chemical treatments. Hemicellulose was located within and between the

cellulose fibrils. This strong association between hemicelluloses and cellulose fibrils was

believed to decrease the crystallinity of the cellulose fibrils and accelerate the beginning of the

thermal degradation process.


The result shows that the fibers that undergo hydrolysis treatment start to degrade earlier

compared to other fibres, around TOnset =185.78 ºC. This may be attributed to the residual of

sulphuric acid group within the fibre, thereby causing the degradation temperature of fibres to be

lower. This lower degradation temperature TOnset =185.78 ºC was also referring to the accessible

PT
degradation temperature of highly sulfated amorphous region which acting as the flame

retardants (Roman & Winter, 2004b), whereas the higher temperature phase correspond to the

RI
degradation of unsulfated crystal interior. Though the fibres have been neutralized to pH of 6-7

SC
during the process of dilution, however there are still lagging of sulphuric acid group present

within the SPNCCs. It has been reported elsewhere (Julien, Chornet, & Overend, 1993; Kim,

U
Nishiyama, Wada, & Kuga, 2001; Li et al., 2009) that the thermostability of NCCs could be

N
reduced a lot due to the treatment of sulphuric acid hydrolysis, as this treatment introduce
A
sulfated group into the NCCs.
M

The oxidation and breakdown of the charred residue is occurred at the final degradation phase
D

(DTG peak above 425 ºC), where these residues were converted to lower molecular weight
TE

gaseous products (Flauzino Neto et al., 2013). Besides that, after the lignin had been completely

decomposed, the residue left was the inorganic material which also assumed as ash content or
EP

char residue. This was due to the presence of inorganic material such as silica (silicon dioxide,

SiO2) in the fibres which can only be decomposed at a high temperature over 1723 ºC and above
CC

(Ishak et al., 2012). Moreover, less weight residue of bleached fibres and alkali-treated fibres
A

than that of raw sugar palm fibres was because of the hemicellulose and lignin were removed

from the both bleached fibres and alkali-treated fibres, (Nguyen, Zavarin, & Barrall, 1981),

where this non-cellulosic could encourage high formation of char residue. Whereas for SPNCCs,

the weight residue of SPNCCs was increased due to the sulfate groups acting as the flame
retardants (Roman & Winter, 2004b). These results were very consistent with results obtained

from the chemical composition, XRD and FTIR measurements.

4. Conclusion

PT
SPNCCs was successfully extracted and isolated from sugar palm fibers using the treatment of

RI
delignification (NaClO2), mercerization (NaOH) and hydrolysis (H2SO4). The chemical

SC
composition analysis showed an increase trending in the cellulose content after each chemical

treatment; it therefore changed from 43.88%% in the raw sugar palm fiber to 81.5% in the

U
bleached fibres and to 82.33% in the alkali-treated fibres. FTIR results also showed that the

N
chemical treatments of delignification (NaClO2) and mercerization (NaOH) was effectively used
A
to remove lignin and hemicellulose, respectively. Thermogravimetric analysis (TGA) shows that
M

the thermal stability increased in the materials having undergone chemical treatments. This was

due to the removal of hemicellulose and lignin as well as to an increase in crystallinity during
D

processing. XRD analyses pointed at a higher crystallinity for the chemically treated fibers, and
TE

the highest crystallinity was found for the SPNCCs. Overall, the crystallinity increased from
EP

55.8% in raw sugar palm fibres to 65.9% in the bleached fibres to 76.0% in the alkali-treated

fibres and to 85.9% in SPNCCs. TEM micrographs showed an adequate, random dispersion in
CC

solution and revealed that sugar palm fibres can be recommended as a preliminary material for

the preparation of SPNCCs. The SPNCCs presented a length (L) of around 130 ± 30 nm and
A

average diameter (D) 9 ± 1.96 nm. This conducted experiment had revealed that sugar palm

fibres can be used as a starting material for the preparation of nanocrystalline cellulose.

Acknowledgments
The authors would like to appreciate University Putra Malaysia for financial support through the

Graduate Research Fellowship (GRF) scholarship. The authors are thankful to Dr. Muhammad

Lamin Sanyang for guideline throughout the experiment. The authors also thanks the Forest

Research Institute Malaysia (FRIM) and Dr. Rushdan Ibrahim for their advice and fruitful

PT
discussions.

RI
SC
U
N
A
M
D
TE
EP
CC
A
References

Adawiyah, D. R., Sasaki, T., & Kohyama, K. (2013). Characterization of arenga starch in
comparison with sago starch. Carbohydrate Polymers, 92(2), 2306–2313.
https://doi.org/10.1016/j.carbpol.2012.12.014
Alemdar, A., & Sain, M. (2008). Isolation and characterization of nanofibers from agricultural
residues - Wheat straw and soy hulls. Bioresource Technology, 99(6), 1664–1671.
https://doi.org/10.1016/j.biortech.2007.04.029

PT
ASTM D1104-56 (1978). “Method of test for holocellulose in wood,” American Society for
Testing and Materials, USA.
ASTM D1103-60 (1977). “Method of test for alpha-cellulose in wood,” American Society for

RI
Testing and Materials, USA.
ASTM D1695-07(2012), Standard Terminology of Cellulose and Cellulose Derivatives, ASTM
International, West Conshohocken, PA, 2012

SC
Azizi Samir, M. A. S., Alloin, F., & Dufresne, A. (2005). Review of Recent Research into
Cellulosic Whiskers, Their Properties and Their Application in Nanocomposite Field.
Biomacromolecules, 6(2), 612–626. https://doi.org/10.1021/bm0493685

U
Bachtiar, D., Salit, M. S., Zainudin, E., Abdan, K., & Dahlan, K. Z. H. M. (2011). Effects of
alkaline treatment and a compatibilizing agent on tensile properties of sugar palm
N
fibrereinforced high impact polystyrene composites. BioResources, 6(4), 4815–4823.
Bachtiar, D., Sapuan, S. M., & Hamdan, M. M. (2008). The effect of alkaline treatment on
A
tensile properties of sugar palm fibre reinforced epoxy composites. Materials & Design,
29(7), 1285–1290. https://doi.org/10.1016/j.matdes.2007.09.006
M

Bachtiar, D., Sapuan, S. M., & Hamdan, M. M. (2010). Flexural properties of alkaline treated
sugar palm fibre reinforced epoxy composites. International Journal of Automotive and
Mechanical Engineering, 1(1), 79–90. https://doi.org/10.15282/ijame.1.2010.7.0007
D

Beck-Candanedo, S., Roman, M., & Gray, D. G. (2005). Effect of Reaction Conditions on the
Properties and Behavior of Wood Cellulose Nanocrystal Suspensions. Biomacromolecules,
TE

6(2), 1048–1054. https://doi.org/10.1021/bm049300p


Beckermann, G. W., & Pickering, K. L. (2008). Engineering and evaluation of hemp fibre
reinforced polypropylene composites: Fibre treatment and matrix modification. Composites
Part A: Applied Science and Manufacturing, 39(6), 979–988.
EP

https://doi.org/10.1016/j.compositesa.2008.03.010
Bhatnagar, A. (2005). Processing of Cellulose Nanofiber-reinforced Composites. Journal of
Reinforced Plastics and Composites, 24(12), 1259–1268.
CC

https://doi.org/10.1177/0731684405049864
Bhattacharya, D., Germinario, L. T., & Winter, W. T. (2008). Isolation, preparation and
characterization of cellulose microfibers obtained from bagasse. Carbohydrate Polymers,
A

73(3), 371–377. https://doi.org/10.1016/j.carbpol.2007.12.005


Bledzki, A. K., & Gassan, J. (1999). Composites reinforced with cellulose based fibers. Progress
in Polymer Science, 24(2), 221–274.
Bondeson, D., Mathew, A., & Oksman, K. (2006a). Optimization of the isolation of nanocrystals
from microcrystalline cellulose by acid hydrolysis. Cellulose, 13(2), 171–180.
https://doi.org/10.1007/s10570-006-9061-4
Bondeson, D., Mathew, A., & Oksman, K. (2006b). Optimization of the isolation of nanocrystals
from microcrystalline cellulose by acid hydrolysis, 171–180.
https://doi.org/10.1007/s10570-006-9061-4
Börjesson, M., & Westman, G. (2015). Crystalline Nanocellulose — Preparation, Modification,
and Properties. In Cellulose - Fundamental Aspects and Current Trends (pp. 159–191).
InTech. https://doi.org/10.5772/61899
Braun, B., Dorgan, J. R., & Chandler, J. P. (2008). Cellulosic Nanowhiskers. Theory and
Application of Light Scattering from Polydisperse Spheroids in the Rayleigh−Gans−Debye
Regime. Biomacromolecules, 9(4), 1255–1263. https://doi.org/10.1021/bm7013137
Brito, B. S. L., Pereira, F. V, Putaux, J.-L., & Jean, B. (2012). Preparation, morphology and

PT
structure of cellulose nanocrystals from bamboo fibers. Cellulose, 19(5), 1527–1536.
https://doi.org/10.1007/s10570-012-9738-9
Cao, X., Ding, B., Yu, J., & Al-Deyab, S. S. (2012). Cellulose nanowhiskers extracted from

RI
TEMPO-oxidized jute fibers. Carbohydrate Polymers, 90(2), 1075–1080.
https://doi.org/10.1016/j.carbpol.2012.06.046
Chauve, M., Barre, L., Tapin-lingua, S., Perez, S., Decottignies, D., Perez, S., & Ferreira, N. L.

SC
(2013). Evolution and impact of cellulose architecture during enzymatic hydrolysis by
fungal cellulases. Advances in Biosciences and Biotechnology’, 4(December), 1095–1109.
https://doi.org/10.4236/abb.2013.412146
Cherian, B. M., Leão, A. L., de Souza, S. F., Thomas, S., Pothan, L. A., & Kottaisamy, M.

U
(2010a). Isolation of nanocellulose from pineapple leaf fibres by steam explosion.
Carbohydrate Polymers, 81(3), 720–725. https://doi.org/10.1016/j.carbpol.2010.03.046
N
Cherian, B. M., Leão, A. L., de Souza, S. F., Thomas, S., Pothan, L. A., & Kottaisamy, M.
(2010b). Isolation of nanocellulose from pineapple leaf fibres by steam explosion.
A
Carbohydrate Polymers, 81(3), 720–725. https://doi.org/10.1016/j.carbpol.2010.03.046
Cherian, B. M., Pothan, L. A., Nguyen-Chung, T., Mennig, G., Kottaisamy, M., & Thomas, S.
M

(2008). A Novel Method for the Synthesis of Cellulose Nanofibril Whiskers from Banana
Fibers and Characterization. Journal of Agricultural and Food Chemistry, 56(14), 5617–
5627. https://doi.org/10.1021/jf8003674
D

Corrêa, A. C., de Morais Teixeira, E., Pessan, L. A., & Mattoso, L. H. C. (2010). Cellulose
nanofibers from curaua fibers. Cellulose, 17(6), 1183–1192. https://doi.org/10.1007/s10570-
TE

010-9453-3
de Mesquita, J. P., Donnici, C. L., & Pereira, F. V. (2010). Biobased Nanocomposites from
Layer-by-Layer Assembly of Cellulose Nanowhiskers with Chitosan. Biomacromolecules,
EP

11(2), 473–480. https://doi.org/10.1021/bm9011985


de Morais Teixeira, E., Bondancia TJ, Teodoro KBR, Correˆa AC, Marconcini JM, & Mattoso
LHC. (2011). Sugarcane bagasse whiskers : Extraction and characterizations. Ind Crops
CC

Prod, 33(63), 66. https://doi.org/10.1016/j.indcrop.2010.08.009


Deepa, B., Abraham, E., Cherian, B. M., Bismarck, A., Blaker, J. J., Pothan, L. A., …
Kottaisamy, M. (2011a). Structure, morphology and thermal characteristics of banana nano
fibers obtained by steam explosion. Bioresource Technology, 102(2), 1988–1997.
A

https://doi.org/10.1016/j.biortech.2010.09.030
Deepa, B., Abraham, E., Cherian, B. M., Bismarck, A., Blaker, J. J., Pothan, L. A., …
Kottaisamy, M. (2011b). Structure, morphology and thermal characteristics of banana nano
fibers obtained by steam explosion. Bioresource Technology, 102(2), 1988–1997.
https://doi.org/10.1016/j.biortech.2010.09.030
Dufresne, A., Cavaillé, J.-Y., & Helbert, W. (1997). Thermoplastic nanocomposites filled with
wheat straw cellulose whiskers. Part II: Effect of processing and modeling. Polymer
Composites, 18(2), 198–210. https://doi.org/10.1002/pc.10274
Dufresne, A., Dupeyre, D., & Vignon, M. R. (2000). Cellulose microfibrils from potato tuber
cells: Processing and characterization of starch-cellulose microfibril composites. Journal of
Applied Polymer Science, 76(14), 2080–2092. https://doi.org/10.1002/(SICI)1097-
4628(20000628)76:14<2080::AID-APP12>3.0.CO;2-U
Eichhorn, S. J., Dufresne, A., Aranguren, M., Marcovich, N. E., Capadona, J. R., Rowan, S. J.,
… Peijs, T. (2010). Review: current international research into cellulose nanofibres and
nanocomposites. Journal of Materials Science, 45(1), 1–33. https://doi.org/10.1007/s10853-

PT
009-3874-0
Fabiyi, J. S., & Ogunleye, B. M. (2015). Mid-infrared spectroscopy and dynamic mechanical
analysis of heat-treated obeche (Triplochiton scleroxylon)wood. Maderas. Ciencia Y

RI
Tecnología, 17(1), 5–16. https://doi.org/10.4067/S0718-221X2015005000001
Faix, O., Lin, S., & Dence, C. (1992). Fourier transform infrared spectroscopy. In Methods in
Lignin Chemistry. Springer-Verlag, 83–109.

SC
Fan, M., Dai, D., & Huang, B. (2012). Fourier Transform - Materials Analysis. (S. Salih, Ed.),
Fourier Transform - Materials Analysis. InTech. https://doi.org/10.5772/2659
Favier, V., Canova, G. R., Cavaillé, J. Y., Chanzy, H., Dufresne, A., & Gauthier, C. (1995).
Nanocomposite materials from latex and cellulose whiskers. Polymers for Advanced

U
Technologies, 6(5), 351–355. https://doi.org/10.1002/pat.1995.220060514
Flauzino Neto, W. P., Silvério, H. A., Dantas, N. O., & Pasquini, D. (2013). Extraction and
N
characterization of cellulose nanocrystals from agro-industrial residue – Soy hulls.
Industrial Crops and Products, 42, 480–488. https://doi.org/10.1016/j.indcrop.2012.06.041
A
Fortunati, E., Puglia, D., Monti, M., Peponi, L., Santulli, C., Kenny, J. M., & Torre, L. (2013).
Extraction of Cellulose Nanocrystals from Phormium tenax Fibres. Journal of Polymers and
M

the Environment, 21(2), 319–328. https://doi.org/10.1007/s10924-012-0543-1


Frenot, A., Henriksson, M. W., & Walkenström, P. (2007). Electrospinning of cellulose-based
nanofibers. Journal of Applied Polymer Science, 103(3), 1473–1482.
D

https://doi.org/10.1002/app.24912
Garcia de Rodriguez, N. L., Thielemans, W., & Dufresne, A. (2006a). Sisal cellulose whiskers
TE

reinforced polyvinyl acetate nanocomposites. Cellulose, 13(3), 261–270.


https://doi.org/10.1007/s10570-005-9039-7
Garcia de Rodriguez, N. L., Thielemans, W., & Dufresne, A. (2006b). Sisal cellulose whiskers
EP

reinforced polyvinyl acetate nanocomposites. Cellulose, 13(3), 261–270.


https://doi.org/10.1007/s10570-005-9039-7
Goetz, L., Foston, M., Mathew, A. P., Oksman, K., & Ragauskas, A. J. (2010). Poly(methyl
CC

vinyl ether- co -maleic acid)−Polyethylene Glycol Nanocomposites Cross-Linked In Situ


with Cellulose Nanowhiskers. Biomacromolecules, 11(10), 2660–2666.
https://doi.org/10.1021/bm1006695
Habibi, Y., Goffin, A.-L., Schiltz, N., Duquesne, E., Dubois, P., & Dufresne, A. (2008).
A

Bionanocomposites based on poly(ε-caprolactone)-grafted cellulose nanocrystals by ring-


opening polymerization. Journal of Materials Chemistry, 18(41), 5002.
https://doi.org/10.1039/b809212e
Habibi, Y., Lucia, L. A., & Rojas, O. J. (2010). Cellulose Nanocrystals: Chemistry, Self-
Assembly, and Applications. Chemical Reviews, 110(6), 3479–3500.
https://doi.org/10.1021/cr900339w
Habibi, Y., & Vignon, M. R. (2008). Optimization of cellouronic acid synthesis by TEMPO-
mediated oxidation of cellulose III from sugar beet pulp. Cellulose, 15(1), 177–185.
https://doi.org/10.1007/s10570-007-9179-z
Hajalilou, A., Hashim, M., & Mohamed Kamari, H. (2014). Effects of Additives and Sintering
Time on the Microstructure of Ni-Zn Ferrite and Its Electrical and Magnetic Properties.
Advances in Materials Science and Engineering, 2014, 1–6.
https://doi.org/10.1155/2014/138789
Himmelsbach, D. S., Khalili, S., & Akin, D. E. (2002). The use of FT-IR microspectroscopic
mapping to study the effects of enzymatic retting of flax (Linum usitatissimum L) stems.

PT
Journal of the Science of Food and Agriculture, 82(7), 685–696.
https://doi.org/10.1002/jsfa.1090
Hofmeister, W., & Platen, H. Von. (1992). Crystal Chemistry and Atomic Order in Brucite-

RI
related Double-layer Structures. Crystallography Reviews, 3(1), 3–26.
https://doi.org/10.1080/08893119208032964
Hubbell, C. A., & Ragauskas, A. J. (2010). Effect of acid-chlorite delignification on cellulose

SC
degree of polymerization. Bioresource Technology, 101(19), 7410–7415.
https://doi.org/10.1016/j.biortech.2010.04.029
Ilyas, R. A., Sapuan, S. M., Ishak, M. R., & Zainudin, E. S. (2017). Effect of Delignification on
the Physical, Thermal, Chemical, and Structural Properties of Sugar Palm Fibre.

U
BioResources, 12(4), 8734–8754. https://doi.org/10.15376/biores.12.4.8734-8754
Ilyas Rushdan, A., Sapuan Salit, M., Lamin Sanyang, M., & Ridzwan Ishak, M. (2017).
N
Nanocrystalline Cellulose As Reinforcement For Polymeric Matrix Nanocomposites And
Its Potential Applications: A Review. Current Analytical Chemistry, 13.
A
https://doi.org/10.2174/1573411013666171003155624
Ishak, M. R., Leman, Z., Sapuan, S. M., Salleh, M. Y., & Misri, S. (2009). the Effect of Sea
M

Water Treatment on the Impact and Flexural Strength of Sugar Palm Fibre Reinforced
Epoxy Composites. International Journal of Mechanical and Materials Engineering (
IJMME ), 4(3), 316–320.
D

Ishak, M. R., Sapuan, S. M., Leman, Z., Rahman, M. Z. A., & Anwar, U. M. K. (2012).
Characterization of sugar palm (Arenga pinnata) fibres. Journal of Thermal Analysis and
TE

Calorimetry, 109(2), 981–989. https://doi.org/10.1007/s10973-011-1785-1


Ishak, M. R., Sapuan, S. M., Leman, Z., Rahman, M. Z. A., Anwar, U. M. K., & Siregar, J. P.
(2013). Sugar palm (Arenga pinnata): Its fibres, polymers and composites. Carbohydrate
EP

Polymers, 91(2), 699–710. https://doi.org/10.1016/j.carbpol.2012.07.073


Islam, M. T., Alam, M. M., & Zoccola, M. (2013). Review on Modification of Nanocellulose for
Application in Composites. International Journal of Innovative Research in Science,
CC

Engineering and Technology, 2(10), 5444–5451.


ISO 5351-1 (2004). “Pulps - Determination of limiting viscosity number in cupri
ethylenediamine (CED) solution,” International Organization for Standardization, Geneva,
Switzerland
A

Iwamoto, S., Kai, W., Isogai, A., & Iwata, T. (2009). Elastic Modulus of Single Cellulose
Microfibrils from Tunicate Measured by Atomic Force Microscopy. Biomacromolecules,
10(9), 2571–2576. https://doi.org/10.1021/bm900520n
Jonoobi, M., Harun, J., Shakeri, A., Misra, M., & Oksmand, K. (2009). Chemical composition,
crystallinity, and thermal degradation of bleached and unbleached kenaf bast (Hibiscus
cannabinus) pulp and nanofibers. BioResources, 4(2), 626–639.
https://doi.org/10.15376/biores.4.2.626-639
Julien, S., Chornet, E., & Overend, R. P. (1993). Influence of acid pretreatment (H2SO4, HCl,
HNO3) on reaction selectivity in the vacuum pyrolysis of cellulose. Journal of Analytical
and Applied Pyrolysis, 27(1), 25–43. https://doi.org/10.1016/0165-2370(93)80020-Z
Jumaidin, R., Sapuan, S. M., Jawaid, M., Ishak, M. R., & Sahari, J. (2017). Characteristics of
Eucheuma cottonii waste from East Malaysia: Physical, Thermal, and Chemical
compositionle. European Journal of Phycology, 0(0), 1–8.
https://doi.org/10.1080/09670262.2016.1248498
Jumaidin, R., Sapuan, S. M., Jawaid, M., Ishak, M. R., & Sahari, J. (2017a). Effect of seaweed

PT
on mechanical, thermal, and biodegradation properties of thermoplastic sugar palm
starch/agar composites. International Journal of Biological Macromolecules, 99, 265–273.
https://doi.org/10.1016/j.ijbiomac.2017.02.092

RI
Jumaidin, R., Sapuan, S. M., Jawaid, M., Ishak, M. R., & Sahari, J. (2017b). Thermal,
mechanical, and physical properties of seaweed/sugar palm fibre reinforced thermoplastic
sugar palm Starch/Agar hybrid composites. International Journal of Biological

SC
Macromolecules, 97, 606–615. https://doi.org/10.1016/j.ijbiomac.2017.01.079
Kazayawoko, M., Balatinecz, J. J., & Woodhams, R. T. (1997). Diffuse reflectance Fourier
transform infrared spectra of wood fibers treated with maleated polypropylenes. Journal of
Applied Polymer Science, 66(6), 1163–1173. https://doi.org/10.1002/(SICI)1097-

U
4628(19971107)66:6<1163::AID-APP16>3.0.CO;2-2
Kim, D.-Y., Nishiyama, Y., Wada, M., & Kuga, S. (2001). High-yield Carbonization of
N
Cellulose by Sulfuric Acid Impregnation. Cellulose, 8(1), 29–33.
https://doi.org/10.1023/A:1016621103245
A
Kumar, R., Mago, G., Balan, V., & Wyman, C. E. (2009). Physical and chemical
characterizations of corn stover and poplar solids resulting from leading pretreatment
M

technologies. Bioresource Technology, 100(17), 3948–3962.


https://doi.org/10.1016/j.biortech.2009.01.075
Li, R., Fei, J., Cai, Y., Li, Y., Feng, J., & Yao, J. (2009). Cellulose whiskers extracted from
D

mulberry : A novel biomass production. Carbohydrate Polymers, 76(1), 94–99.


https://doi.org/10.1016/j.carbpol.2008.09.034
TE

Lu, P., & Hsieh, Y. (2012). Preparation and characterization of cellulose nanocrystals from rice
straw. Carbohydrate Polymers, 87(1), 564–573.
https://doi.org/10.1016/j.carbpol.2011.08.022
EP

Martini, E., Roshetko, J. M., van Noordwijk, M., Rahmanulloh, A., Mulyoutami, E., Joshi, L., &
Budidarsono, S. (2012). Sugar palm (Arenga pinnata (Wurmb) Merr.) for livelihoods and
biodiversity conservation in the orangutan habitat of Batang Toru, North Sumatra,
CC

Indonesia: Mixed prospects for domestication. Agroforestry Systems, 86(3), 401–417.


https://doi.org/10.1007/s10457-011-9441-0
Mogea, J., Seibert, B., & Smits, W. (1991). Multipurpose palms: the sugar palm (Arenga pinnata
(Wurmb) Merr.). Agroforestry Systems, 13(2), 111–129.
A

https://doi.org/10.1007/BF00140236
Mohanty, A. K., Misra, M., & Hinrichsen, G. (2000). Biofibres, biodegradable polymers and
biocomposites: An overview. Macromolecular Materials and Engineering, 276–277, 1–24.
https://doi.org/10.1002/(SICI)1439-2054(20000301)276:1<1::AID-MAME1>3.0.CO;2-W
Morais, J. P. S., Rosa, M. D. F., de Souza Filho, M. de sá M., Nascimento, L. D., do Nascimento,
D. M., & Cassales, A. R. (2013). Extraction and characterization of nanocellulose structures
from raw cotton linter. Carbohydrate Polymers, 91(1), 229–235.
https://doi.org/10.1016/j.carbpol.2012.08.010
Morán, J. I., Alvarez, V. A., Cyras, V. P., & Vázquez, A. (2008). Extraction of cellulose and
preparation of nanocellulose from sisal fibers. Cellulose, 15(1), 149–159.
https://doi.org/10.1007/s10570-007-9145-9
N., E., Jane, Z., & K., R. (2012). Image Watermarking in Higher-Order Gradient Domain. In
Advances in Wavelet Theory and Their Applications in Engineering, Physics and
Technology. InTech. https://doi.org/10.5772/35603
Ndazi, B. S., Nyahumwa, C., & Tesha, J. (2007). Chemical and thermal stability of rice husks

PT
against alkali treatment. BioResources, 3(4), 1267–1277.
https://doi.org/10.15376/biores.3.4.1267-1277
Ng, H. M., Sin, L. T., Tee, T. T., Bee, S. T., Hui, D., Low, C. Y., & Rahmat, A. R. (2015).

RI
Extraction of cellulose nanocrystals from plant sources for application as reinforcing agent
in polymers. Composites Part B: Engineering, 75, 176–200.
https://doi.org/10.1016/j.compositesb.2015.01.008

SC
Ng, H., Sin, L. T., Tee, T., Bee, S., Hui, D., Low, C., & Rahmat, A. R. (2015). Extraction of
cellulose nanocrystals from plant sources for application as reinforcing agent in polymers.
Composites Part B: Engineering, 75, 176–200.
https://doi.org/10.1016/j.compositesb.2015.01.008

U
Nguyen, T., Zavarin, E., & Barrall, E. M. (1981). Thermal Analysis of Lignocellulosic Materials.
Journal of Macromolecular Science, Part C, 20(1), 1–65.
https://doi.org/10.1080/00222358108080014 N
Pääkkö, M., Vapaavuori, J., Silvennoinen, R., Kosonen, H., Ankerfors, M., Lindström, T., …
A
Ikkala, O. (2008). Long and entangled native cellulose I nanofibers allow flexible aerogels
and hierarchically porous templates for functionalities. Soft Matter, 4(12), 2492.
M

https://doi.org/10.1039/b810371b
Petersson, L., Mathew, A. P., & Oksman, K. (2009). Dispersion and properties of cellulose
nanowhiskers and layered silicates in cellulose acetate butyrate nanocomposites. Journal of
D

Applied Polymer Science, 112(4), 2001–2009. https://doi.org/10.1002/app.29661


Qi, H., Cai, J., Zhang, L., & Kuga, S. (2009). Properties of Films Composed of Cellulose
TE

Nanowhiskers and a Cellulose Matrix Regenerated from Alkali/Urea Solution.


Biomacromolecules, 10(6), 1597–1602. https://doi.org/10.1021/bm9001975
Ray, D., & Sarkar, B. K. (2001). Characterization of alkali-treated jute fibers for physical and
EP

mechanical properties. Journal of Applied Polymer Science, 80(7), 1013–1020.


https://doi.org/10.1002/app.1184
Razali, N., Salit, M. S., Jawaid, M., Ishak, M. R., & Lazim, Y. (2015). A study on chemical
CC

composition, physical, tensile, morphological, and thermal properties of roselle fibre: Effect
of fibre maturity. BioResources, 10(1), 1803–1823.
Roman, M., & Gray, D. G. (2005). Parabolic Focal Conics in Self-Assembled Solid Films of
Cellulose Nanocrystals. Langmuir, 21(12), 5555–5561. https://doi.org/10.1021/la046797f
A

Roman, M., & Winter, W. T. (2004a). Effect of sulfate groups from sulfuric acid hydrolysis on
the thermal degradation behavior of bacterial cellulose. Biomacromolecules, 5(5), 1671–
1677. https://doi.org/10.1021/bm034519+
Roman, M., & Winter, W. T. (2004b). Effect of sulfate groups from sulfuric acid hydrolysis on
the thermal degradation behavior of bacterial cellulose. Biomacromolecules, 5(5), 1671–
1677. https://doi.org/10.1021/bm034519+
Rong, M. Z., Zhang, M. Q., Liu, Y., Yang, G. C., & Zeng, H. M. (2001). The effect of fiber
treatment on the mechanical properties of unidirectional sisal-reinforced epoxy composites.
Composites Science and Technology, 61(10), 1437–1447. https://doi.org/10.1016/S0266-
3538(01)00046-X
Rosa, M. F., Medeiros, E. S., Malmonge, J. A., Gregorski, K. S., Wood, D. F., Mattoso, L. H. C.,
… Imam, S. H. (2010). Cellulose nanowhiskers from coconut husk fibers: Effect of
preparation conditions on their thermal and morphological behavior. Carbohydrate
Polymers, 81(1), 83–92. https://doi.org/10.1016/j.carbpol.2010.01.059
Rosa, M., Malmonge, J., Delilah, W., & Imam, S. H. (2010). Cellulose nanowhiskers from

PT
coconut husk fibers : Effect of preparation conditions on their thermal and morphological
behavior. Carbohydrate Polymers, 81, 83–92. https://doi.org/10.1016/j.carbpol.2010.01.059
Sahari, J., Sapuan, S. M., Ismarrubie, Z. N., & Rahman, M. (2012). Physical and chemical

RI
properties of different morphological parts of sugar palm fibres. Fibres and Textiles in
Eastern Europe, 91(2), 21–24.
Sahari, J., Sapuan, S. M., Ismarrubie, Z. N., & Rahman, M. Z. a. (2011). Investigation on

SC
Bending Strength and Stiffness of Sugar Palm Fibre from Different Parts Reinforced
Unsaturated Polyester Composites. Key Engineering Materials, 471–472(October 2015),
502–506. https://doi.org/10.4028/www.scientific.net/KEM.471-472.502
Sahari, J., Sapuan, S. M., Zainudin, E. S., & Maleque, M. A. (2013a). Effect of Water

U
Absorption on Mechanical Properties of Sugar Palm Fibre Reinforced Sugar Palm Starch
(SPF/SPS) Biocomposites. Journal of Biobased Materials and Bioenergy, 7(1), 90–94.
https://doi.org/10.1166/jbmb.2013.1267 N
Sahari, J., Sapuan, S. M., Zainudin, E. S., & Maleque, M. A. (2013b). Mechanical and thermal
A
properties of environmentally friendly composites derived from sugar palm tree. Materials
& Design, 49(2), 285–289. https://doi.org/10.1016/j.matdes.2013.01.048
M

Samir, A., Alloin, F., Paillet, M., & Dufresne, A. (2004). Tangling Effect in Fibrillated Cellulose
Reinforced Nanocomposites. Macromolecules, 37, 4313–4316.
Sanyang, M. L., Sapuan, S. M., Jawaid, M., Ishak, M. R., & Sahari, J. (2016). Effect of sugar
D

palm-derived cellulose reinforcement on the mechanical and water barrier properties of


sugar palm starch biocomposite films. BioResources, 11(2), 4134–4145.
TE

https://doi.org/10.15376/biores.11.2.4134-4145
Savadekar, N. R., & Mhaske, S. T. (2012). Synthesis of nano cellulose fibers and effect on
thermoplastics starch based films. Carbohydrate Polymers, 89(1), 146–151.
EP

https://doi.org/10.1016/j.carbpol.2012.02.063
Segal, L., Creely, J. J., Martin, A. E., & Conrad, C. M. (1959). An Empirical Method for
Estimating the Degree of Crystallinity of Native Cellulose Using the X-Ray Diffractometer.
CC

Textile Research Journal, 29(10), 786–794. https://doi.org/10.1177/004051755902901003


Shafizadeh, F., & Chin, P. P. S. (1977). Thermal Deterioration of Wood. In Goldstein IS (Ed.),
Wood technology: chemical aspects (pp. 57–81). ACS Symposium Series 43.
https://doi.org/10.1021/bk-1977-0043.ch005
A

Sheltami, R. M., Abdullah, I., Ahmad, I., Dufresne, A., & Kargarzadeh, H. (2012). Extraction of
cellulose nanocrystals from mengkuang leaves ( Pandanus tectorius ). Carbohydrate
Polymers, 88(2), 772–779. https://doi.org/10.1016/j.carbpol.2012.01.062
Silvério, H. A., Flauzino Neto, W. P., Dantas, N. O., & Pasquini, D. (2013). Extraction and
characterization of cellulose nanocrystals from corncob for application as reinforcing agent
in nanocomposites. Industrial Crops and Products, 44, 427–436.
https://doi.org/10.1016/j.indcrop.2012.10.014
Siqueira, G., Bras, J., & Dufresne, A. (2010). New Process of Chemical Grafting of Cellulose
Nanoparticles with a Long Chain Isocyanate. Langmuir, 26(1), 402–411.
https://doi.org/10.1021/la9028595
Soares, S., Camino, G., & Levchik, S. (1995). Comparative study of the thermal decomposition
of pure cellulose and pulp paper. Polymer Degradation and Stability, 49(2), 275–283.
https://doi.org/10.1016/0141-3910(95)87009-1
Sonia, A., Priya Dasan, K., & Alex, R. (2013). Celluloses microfibres (CMF) reinforced poly
(ethylene-co-vinyl acetate) (EVA) composites: Dynamic mechanical, gamma and thermal

PT
ageing studies. Chemical Engineering Journal, 228, 1214–1222.
https://doi.org/10.1016/j.cej.2013.04.091
Sun, R. C., & Tomkinson, J. (2002). Comparative study of organic solvent-soluble and water-

RI
soluble lipophilic extractives from wheat straw 2: spectroscopic and thermal analysis.
Journal of Wood Science, 48(3), 222–226. https://doi.org/10.1007/BF00771371
Sun, X. F., Xu, F., Sun, R. C., Fowler, P., & Baird, M. S. (2005). Characteristics of degraded

SC
cellulose obtained from steam-exploded wheat straw. Carbohydrate Research, 340(1), 97–
106. https://doi.org/10.1016/j.carres.2004.10.022
TAPPI. (2006). “Acid-insoluble Lignin in Wood and Pulp”. TAPPI T 222, om-02. (1-14).
TAPPI (1999). “Alpha-, beta- and gamma-cellulose in pulp”. TAPPI 203, cm-99. (1-5).

U
Talib, R. A., Tawakkal, I. S. M. A., & Khalina, A. (2011). The Influence of Mercerised Kenaf
Fibres Reinforced Polylactic Acid Composites on Dynamic Mechanical Analysis. Key
Engineering Materials, 471–472, 815–820. N
https://doi.org/10.4028/www.scientific.net/KEM.471-472.815
A
Tawakkal, I. S. M. a. ., Talib, R. a. ., Abdan, K., & Ling, C. N. (2012). Mechanical and physical
properties of kenaf- derived cellulose (kdc)-filled polylactic acid (pla) composites.
M

BioResources, 7(2), 1643–1655.


Tee, T.-T., Sin, L. T., Gobinath, R., Bee, S.-T., Hui, D., Rahmat, A. R., … Fang, Q. (2013).
Investigation of nano-size montmorillonite on enhancing polyvinyl alcohol–starch blends
D

prepared via solution cast approach. Composites Part B: Engineering, 47, 238–247.
https://doi.org/10.1016/j.compositesb.2012.10.033
TE

Tee, Y. B., Talib, R. a., Abdan, K., Chin, N. L., Basha, R. K., & Md Yunos, K. F. (2013).
Thermally Grafting Aminosilane onto Kenaf-Derived Cellulose and Its Influence on the
Thermal Properties of Poly(Lactic Acid) Composites. BioResources, 8(3), 4468–4483.
EP

https://doi.org/10.15376/biores.8.3.4468-4483
Thiripura Sundari, M., & Ramesh, A. (2012). Isolation and characterization of cellulose
nanofibers from the aquatic weed water hyacinth—Eichhornia crassipes. Carbohydrate
CC

Polymers, 87(2), 1701–1705. https://doi.org/10.1016/j.carbpol.2011.09.076


Ticoalu, A., Aravinthan, T., & Cardona, F. (2013). A review on the characteristics of gomuti
fibre and its composites with thermoset resins. Journal of Reinforced Plastics and
Composites, 32(2), 124–136. https://doi.org/10.1177/0731684412463109
A

Tonoli, G. H. D., Teixeira, E. M., Corrêa, A. C., Marconcini, J. M., Caixeta, L. A., Pereira-da-
Silva, M. A., & Mattoso, L. H. C. (2012). Cellulose micro/nanofibres from Eucalyptus kraft
pulp: Preparation and properties. Carbohydrate Polymers, 89(1), 80–88.
https://doi.org/10.1016/j.carbpol.2012.02.052
Wang Yueping, Wang Ge, Cheng Haitao, Tian Genlin, Liu Zheng, Xiao Qun Feng, … Gao
Xushan. (2010). Structures of Bamboo Fiber for Textiles. Textile Research Journal, 80(4),
334–343. https://doi.org/10.1177/0040517509337633
Wise, L. E., Murphy, M., & D’Addieco, A. A. (1946). Chlorite, holocellulose, its fractionation
and bearing on summative wood analysis and on studies on the hemicellulose. Paper Trade
Journal, 122(2), 35–43.
Xiao, B., Sun, X. ., & Sun, R. (2001). Chemical, structural, and thermal characterizations of
alkali-soluble lignins and hemicelluloses, and cellulose from maize stems, rye straw, and
rice straw. Polymer Degradation and Stability, 74(2), 307–319.
https://doi.org/10.1016/S0141-3910(01)00163-X
Yang, H., Yan, R., Chen, H., Lee, D. H., & Zheng, C. (2007). Characteristics of hemicellulose,

PT
cellulose and lignin pyrolysis. Fuel, 86(12–13), 1781–1788.
https://doi.org/10.1016/j.fuel.2006.12.013
Yasim-Anuar, T. A. T., Ariffin, H., Norrrahim, M. N. F., & Hassan, M. A. (2017). Factors

RI
Affecting Spinnability of Oil Palm Mesocarp Fiber Cellulose Solution for the Production of
Microfiber. Bioresources, 12(1), 715–734.
Zaini, L. H., Jonoobi, M., Tahir, P. M., & Karimi, S. (2013). Isolation and Characterization of

SC
Cellulose Whiskers from Kenaf (Hibiscus cannabinus) Bast Fibers. Journal of Biomaterials
and Nanobiotechnology, 4(1), 37–44. https://doi.org/10.4236/jbnb.2013.41006
Zuluaga, R., Putaux, J.-L., Restrepo, A., Mondragon, I., & Gañán, P. (2007). Cellulose
microfibrils from banana farming residues: isolation and characterization. Cellulose, 14(6),

U
585–592. https://doi.org/10.1007/s10570-007-9118-z

N
A
M
D
TE
EP
CC
A
PT
RI
SC
U
N
A
M
D
TE
EP
CC
A

Fig. 1. Photographs of (a) the sugar palm tree, (d) raw sugar palm fibres, (e) bleached fibres and
(f) alkali-treated fibres; FESEM micrographs of sugar plant fibres: (g) cross section, (b)
longitudinal section, (c) primary, secondary cell wall and middle lamella, (h) alkali-treated
fibres, (i) and bleached fibres.
PT
RI
SC
U
N
A
M
D
TE
EP
CC
A

Fig. 2. (a) Aqueous suspension (2 wt%), (b) FESEM micrograph, (c) height measurement of NCCs

using atomic force microscopy (AFM), (d) transmission electron (TEM) micrograph, (e) atomic

force microscopy (AFM) and their length (f and g) and diameter (h and i) histograms of

nanocrystalline cellulose extracted from sugar palm fibres.


A
CC
EP
TE
D
M
A
N
U
SC
RI
PT
PT
RI
SC
U
N
A
M
D
TE

Fig. 3. FTIR spectra of (a) sugar palm fibres, (b) bleached fibres, (c) alkali-treated fibres, and (d)
EP

SPNCCs.
CC
A
1000

800 d

PT
a
600
Intensity

RI
b

SC
400
c

U
200
N
A
M

0
5 10 15 20 25 30 35 40
D

2(degree)
TE

Fig. 4. X-ray diffraction patterns of (a) sugar palm fibres, (b) bleaching fibres, (c) alkali-treated

fibres, and (d) SPNCCs.


EP
CC
A
100

PT
80

RI
Weight (%)

60

SC
U
40
N a
A
b
20
M

d
c
100 200 300 400 500 600
D

Temperature (oC)
TE

i
EP
CC
A
PT
Weight (%)

RI
a
b

SC
U
N d
A
c
M

100 200 300 400 500 600


D

Temperature (oC) ii
TE
EP

Fig. 5. (i) TG and (ii) DTG curves for (a) raw sugar palm fibres, (b) bleached fibres, (c) alkali-
CC

treated fibres and (d) SPNCCs.


A
Table 1

Chemical composition of sugar palm fibres at different stages of treatment.

Samples Cellulose (%) Hemicellulose (%) Holocellulose (%) Lignin (%) Extractive (%) Ash (%)

Sugar Palm fibres 43.88 7.24 51.12 33.24 2.73 1.01

PT
Bleached fibres 56.67 19.8 76.47 0.27 0.23 2.16

Alkali-treated fibres 82.33 3.97 86.3 0.06 - 0.72

RI
Table 2

SC
Physical properties of Sugar Palm fibres, bleached fibres, alkali-treated fibres, SPNCCs and

U
other materials.

Samples Diameter
Density
(g/cm-3)
Moisture
content (wt %)
Xc
(%)
NSurface
area
Pore
volume DP Mw (g/mol) Reference
A
(m2/g) (cm3/g)
Sugar Palm fibres 212.01 ± 1.50 8.36 ± 0.0984 55.8 7.58 0.0607 Current study
- -
2.17 µm
M
Bleached fibres 94.49 ± 1.30 ± 6.25 ± 0.0745 65.9 10.35 0.0678 Current study
2963.33 480,513.39
0.03 µm 0.0023
Alkali-treated 11.87 ± 1.28 ± 3.83 ± 0.1037 76.0 13.18 0.1950 Current study
946.48 153,458.51
fibres 1.04 µm 0.0019
Sugar Palm NCC 9 ± 1.96 nm 1.05 ± 17.901 ± 85.9 14.47 0.2260 Current study
D

142.86 23,164.7
0.0023 0.0623
Coconut Husk 5.5 ± 1.5 (M. Rosa et
- - - - - - -
NCC nm al., 2010)
TE

Rice Straw NCC (Thiripura


Sundari &
5.06 nm - - - - - - -
Ramesh,
2012)
Sisal Fibre NCC 30.9 ± 12.5 (Morán et al.,
EP

- - 75.0 - - - -
nm 2008)
Soy Hulls NCC (Flauzino
4.43 ± 1.20
- - 69.6 - - - - Neto et al.,
nm
2013)
CC

Sugarcane (de Morais


bagasse NCC 4 ± 2 nm - - 87.5 - - - - Teixeira et
al., 2011)
Results expressed as mean ± standard deviation
A
Table 3

Onset temperature (TOnset), degradation temperature on maximum weight-loss rate (TMax), weight

loss (WL) and char yield for raw sugar palm fibres, bleached fibres, alkali-treated fibres and

PT
SPNCCs obtained from the TG and DTG curves.

RI
Sample Water evaporation 1st thermal degradation 2nd thermal degrade Char yield
TOnset (◦C) TMax (◦C) WL (%) TOnset (◦C) TMax (◦C) WL (%) TOnset (◦C) TMax (◦C) WL (%) W (%)

SC
Raw sugar 41.73 10.38 210.58 281 15.13 308.05 345.45 43.76 30.73
106.78
palm fibres
Bleached 42.37 9.87 195.66 271.56 15.24 288.35 324.44 52.39 22.5
103.74
fibres
Alkali- 43.49 101.23 8.58 207.92 346.09 73.71 - - - 17.71

U
treated
fibres
SPNCCs 43.72 103.43 7.37 185.78 348.65 78.41 - - - 17.97

N
A
M
D
TE
EP
CC
A

You might also like