You are on page 1of 53

Accepted Manuscript

Pyrus pyrifolia fruit peel as sustainable source for spherical and


porous network based nanocellulose synthesis via one-pot
hydrolysis system

You Wei Chen, Muhammad Ariff Hasanulbasori, Phang Fung


Chiat, Hwei Voon Lee

PII: S0141-8130(18)34193-X
DOI: doi:10.1016/j.ijbiomac.2018.10.013
Reference: BIOMAC 10664
To appear in: International Journal of Biological Macromolecules
Received date: 10 August 2018
Revised date: 21 September 2018
Accepted date: 1 October 2018

Please cite this article as: You Wei Chen, Muhammad Ariff Hasanulbasori, Phang Fung
Chiat, Hwei Voon Lee , Pyrus pyrifolia fruit peel as sustainable source for spherical and
porous network based nanocellulose synthesis via one-pot hydrolysis system. Biomac
(2018), doi:10.1016/j.ijbiomac.2018.10.013

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.
ACCEPTED MANUSCRIPT

Pyrus pyrifolia fruit peel as sustainable source for spherical and porous network based
nanocellulose synthesis via one-pot hydrolysis system

You Wei Chen, Muhammad Ariff Hasanulbasori, Phang Fung Chiat, Hwei Voon Lee*

Nanotechnology & Catalysis Research Centre (NANOCAT), Institute of Postgraduate

PT
Studies, University of Malaya, 50603 Kuala Lumpur, Malaysia

*Corresponding author: leehweivoon@um.edu.my

RI
Tel: +603-7967 6954, Fax: +603-7957 6956

SC
Abstract
NU
In this paper, we have attempted to the revalorization of pear (Pyrus pyrifolia L.) peel
residue and converting them into high value-added products remains limited. A green and
facile one-pot isolation procedure was designed for simplifying the isolation of nanocellulose
MA

directly from pear peel residue. The one-pot approach employed in this work is interesting as
the reaction involved less harmful chemicals and multiple steps. The reaction was carried out
D

by adding hydrogen peroxide as an oxidant and chromium (III) nitrate as a catalyst in the
acidic medium under mild process conditions. FTIR spectroscopy proved the resultant
E

nanocellulose was the cellulose species without the presence of non-cellulosic layer. XRD
PT

study indicated that the isolated nanocellulose possessed of cellulose I polymorph with the
high crystallinity index of 85.7%. FESEM analysis clearly revealed the considerable size
CE

reduction during one-pot process. Remarkably, TEM analysis revealed that the isolated
nanocellulose consisted of networked nature and spherical shaped morphologies with the
AC

high aspect ratio of 24.6. TGA presented lower thermal stability for nanocellulose. This study
provides a cost-effective method and straightforward one-pot process for fabricating
nanocellulose from pear peel residue. This was the first investigation on the nanocellulose
extraction from pear fruit.

Keywords: Fruit residue; Oxidative depolymerization; One-pot isolation approach

1. Introduction

1
ACCEPTED MANUSCRIPT

Asian pear (Pyrus pyrifolia L.) is one of the most important, popular and widely
consumed fruit crops in the world. Asian pear is a pear tree species native to Asian and the
fruit is known by many names, including sand pear, Chinese pear, Korean pear and Japanese
pear. Cultivars derived from Pyrus pyrifolia are grown throughout East Asia and in some
other countries such as United State, Australia and New Zealand. In 2012, the global annual
production of fresh pears was around 23 million tons, according to the United Nations Food
and Agriculture Organization (FAO). In the same year, approximately 43646 tons of pears

PT
were imported by Malaysia. Pear is used mostly for fresh consumption or used in processed
food products including drinks, jams, candy, jellies, juices, puree and preserved fruits. Pear
has also been used as a traditional herbal medicine in Asia countries due to the presence of

RI
phytochemicals that exert preventive effects against various diseases in human [1]. Pear peel

SC
is normally not consumed with fruits due to the concerned about the pesticide and chemical
content on the surface and it does not provide the best taste and texture. Pear fruit is generally
NU
peeled for eating as it has a thick peel that contains plenty of pectin and stone cells with a
layer of wax. In recent years, the food processing industries have seen rapid growth
throughout the world, but also major losses and residue during processing. During the typical
MA

pulp extraction or pear juice manufacturing process, ca. 80‒85% of the fresh weight of the
fruits will be utilized while the remaining by-products lack commercial applications and
D

remained in large qualities as the primary byproducts. Therefore, it is one of the most
abundant, renewable and cheap fruit residues, while representing an average estimated
E

production of 100 kg of pears produced ca. 12 kg of solid residue [2]. Besides that, the pear
PT

peel residue is also contributed along the food chain (from agriculture to retail and household
consumption). Due to the high consumption and industrial processing of the edible parts of
CE

pear, pear pomaces (principally peel) are generated in large quantities from the industry
processing units. Therefore, increasing disposal of this residue is rapidly becoming more
AC

difficult to solve. At present, such enormous quality of residue is often dedicated to animal
feed or disposed of it via landfill and incineration. In fact, inappropriate management of
landfill resulted in the emission of harmful greenhouse gases (i.e. methane and carbon
dioxide) an incineration process involves the formation and release of pollutants and
secondary waste such as acid gases, dioxins, furans and particulates, which not only directly
contribute to huge environmental issues, also causes serious health risks [3]. For these
reasons, there is an urgent need of finding alternative valorization options and value-added
use for pear peel residue. In fact, nowadays the use of fruit residual is quickly becoming a
mainstream industrial approach for the production of advanced biodegradable materials, such

2
ACCEPTED MANUSCRIPT

as nanocellulose. The inexpensive and readily available use of pear peel residue is highly
cost-effective and minimizes environmental impact. For this reason, research for the
utilization of pear peel residue as sources of nanocellulose may be of considerable economic
benefits and has become increasingly attractive. From an environmental point of view,
recovery of valuable nanocellulose product from pear peel residue is an ever-growing
research area and considered one of the most effective solutions for the correct management
of fruit processing industry residues. Various efforts are being taken towards the waste to

PT
wealth concept all around the world. In this concept, the production of green, low-cost and
high-performance nanocellulose from renewable and sustainable resources is an expanding
research area.

RI
SC
The depletion of fossil resources together with the increase in environmental
awareness and health concerns are acting synergistically in providing the motivation for
NU
developing products and materials that satisfy the environmental concerns and independence
of petrochemical resources [4]. During the past two decades, more attention was being given
to the usage of bio-based materials that can prevent the widespread dependence on fossil
MA

fuels. Cellulose is one of the most abundant and ubiquitous renewable natural biopolymers on
Earth. It is considered as an inexhaustible source of feedstock for the increasing demand for
D

biocompatible and environmentally friendly materials and products because it is regularly


regenerated by nature within a short period of time with the annual production is estimated to
E

be more than 7.5 × 1010 tons [5]. It is considered to be water-insoluble compound and can
PT

play an important role in maintaining the structure of a plant cell wall. Since its discovery and
isolation by Anselme Payen in 1838, the structure and properties of cellulose have been
CE

largely studied and highlighted in the literature. To understand nanocellulose production it is


necessary to comprehend the cellulose biosynthesis in nature. It is composed of (1–4)-β-
AC

linked ᴅ-glucose chains with a syndiotactic configuration, assembled together by hydrogen


bonds and Van der Waals interaction into elementary microfibrils that further aggregate into
larger structures. Cellulose microfibrils involve a highly crystalline core surrounded by less
crystalline regions and interrupted by amorphous forms of cellulose. Compared with cellulose,
nanocellulose has attracted considerable attention in various fields due to its extraordinary
properties including low cost, low density, renewability, high surface area, high chemical
reactivity, high strength and modulus, high elasticity, high transparency, high tensile stiffness,
lightweight, low thermal expansion, recyclable and biodegradable natural advantages due to
its nano-dimensional properties [6, 7]. Nanocellulose is the nanoscale form of cellulose which

3
ACCEPTED MANUSCRIPT

can be produced in various morphological shapes such spherical, rod-like, ribbon-like,


needle-shaped, having a compact structure of ordered cellulose chains stabilized by inter and
intramolecular hydrogen bonding, which make very interesting solid crystalline nanoparticles
with unique characteristics [8].

Starting from pure cellulose materials, nanocellulose can be isolated via various
hydrolysis approaches including acid hydrolysis, mechanical disintegration, TEMPO-

PT
mediated oxidation and enzyme-assisted hydrolysis [9, 10]. Among these techniques, acid
hydrolysis process is the most popular and effective in nanocellulose production. During the
acid hydrolysis, the cellulose fibers are subjected to concentrated acid to remove the

RI
amorphous domains of the cellulose chains and leave the crystalline domains unaltered. In

SC
this context, the sulfuric acid (H2SO4) has been extensively used for nanocellulose extraction,
but hydrochloric (HCl), phosphoric (H3PO4) and hydrobromic (HBr) acids have also been
NU
reported for such purposes [11]. However, H3PO4 and HBr were less effective for acid
hydrolysis and required longer reaction time to complete the process. HCl generates low-
density surface charges on the nanocellulose with limited nanocrystal dispersibility, which
MA

tends to promote flocculation in aqueous suspensions. Therefore, sulfuric acid hydrolysis is a


simple process and it requires shorter reaction time than other processes [12]. Furthermore,
D

the nanocellulose produced through this process consists of functionalized surface, high
crystallinity and good colloidal stability in water. Unfortunately, this process has some severe
E

drawbacks for large-scale production, such as over-degradation of cellulose resulted in low


PT

yield (ca. 30%), pollution of environment, large amount of water consumption, equipment
corrosion and generation of huge amount of waste [13]. Thus, diluted or organic acids have
CE

been suggested for milder reactions, but such treatments have been found to be less effective
in reacting to the cellulose structure [14, 15]. In order to further improve the cellulose
AC

hydrolysis efficiency, Cr(NO3)3 as a co-catalyst has been added in the dilute sulfuric acid
hydrolysis process. Previous studies [16-18] proved that the rate of cellulose hydrolysis can
be significantly improved by dilute acid pretreatment of cellulose with the presence of metal
ion catalysts. In recent years, Cr(NO3)3 is chosen as a Lewis acid catalyst due to its superior
performance in complete solubilize of cellulose [19, 20], hemicellulose and lignocellulosic
[21] into water soluble chemicals and fuels. Thus, it is believed that controlled hydrolysis of
cellulose to produce solid state nanocellulose could be achieved by using [Cr(NO3)3/H2SO4]
hydrolyzing medium under less severe reaction conditions.

4
ACCEPTED MANUSCRIPT

At the industrial scale, one-pot cost-efficient methods for the isolation of


nanocellulose are beneficial. To prepare and isolate nanocellulose using existing methods, the
removal of the non-cellulosic components such as lignin, hemicellulose and waxes is a
prerequisite. This process involves a multiple stage chemical procedures such as alkalization
and bleaching process in order to obtain relatively pure cellulosic starting materials before the
acid hydrolysis. With this technique, tedious and time-consuming preparation steps are
required to obtain nanocellulose, and this impedes large-scale production and real-world

PT
applications of nanocellulose. Besides that, such multistep extraction procedures involved
many harmful chemicals, high volumes of water consumed and significant product losses
during multiple washing phases. These main drawbacks are potential obstacles to their use in

RI
the stage of industrial scaling. Fortunately, all these difficulties can be avoided because of the

SC
exciting discovery of the isolation of nanocellulose via an effective and green procedure
using hydrogen peroxide (H2O2). H2O2 is safe to use and creates no harmful by-product and
NU
decomposes inevitably to water and oxygen. According to the reference [22], the H2O2
decomposition mechanism can be implied as follows:
H2O2 → 2HO•
MA

H2O2 + HO• → HOO• + H2O


HO• + HOO• → O2 + H2O
D

Overall equation: 2 H2O2 → 2 H2O + O2


E

In general, H2O2 is able to increase the accessibility of the cellulose by detaching and
PT

loosen the lignocellulosic matrix. Several reactions may take places such as cleavage of alkyl
aryl ether linkages, displacement of side chains, electrophilic substitution or the oxidative
CE

cleavage of aromatic nuclei. H2O2 is a powerful oxidizing agent in degradation of lignin


components and capable of penetrating the amorphous domains as well as performing the
AC

decolorization of cellulose materials [23]. Furthermore, elimination of crosslinks contributed


to the separation of structural linkages between lignin and other lignocellulose components.
Therefore, removal of lignin and hemicellulose can be easily attained by this oxidizing agent.
Most importantly, H2O2 is considered one of the greenest and the most efficient oxidizing
agents because its oxidation by-product is only water and it has high active oxygen content
(about 47%) [24]. Indeed, H2O2 dissociates to form the hydroperoxide anion (HOO−) and this
anion reacts with undissociated H2O2 to form highly reactive hydroxyl (HO•) and superoxide
(O2‒•) radicals that cleaves ester link between hemicelluloses and lignin. HOO− is responsible
for lignin oxidation and decolorization while HO• and O2‒• bleach the material attacking the

5
ACCEPTED MANUSCRIPT

lignin side-chains and producing water soluble oxidation and low molecular weight products
[25]. The action of H2O2 causes less damage to carbohydrates and delignification is more
efficient [2]. According to a study conducted by Ching’s group [26], the results revealed that
the most of the lignin and hemicellulose had been simultaneously degraded during the
bleaching process and therefore, additional alkaline treatment did not require to hydrolyze
hemicellulose. Apart from that, the main disadvantage of NaOH treatment causes high water
use during neutralization, which is not required for producing cellulose I nanomaterial. In

PT
addition to that, bleaching process is proved to be effective in removing the extractives on the
surface of fiber [27]. In summary, the H2O2 bleaching agent is an excellent choice for the
removal of extractive, hemicellulose and lignin.

RI
SC
Previous studies on pear fruit have mostly focused on the extraction of valuable
bioactive constituents such as phenolic compounds, sugars derivatives, organic acids,
NU
vitamins and minerals [28, 29]. However, systematic studies on the revalorization of pear
peel residue into high-value added nanocellulose solid material have not yet been performed.
Concerning chemical composition, cell walls of pear fruit consist mostly of cellulose,
MA

hemicellulose and pectin, with some lignin in pears. Cellulose represents however more than
30% of dry matter in pear pomace [2]. Considering its large production and its cellulose
D

content, this biomaterial has been proposed within the biorefinery concept as a source of
broad range of high value-added products for the production of cellulose derivatives, such as
E

highly crystalline nanocellulose, which are considered, due to their outstanding properties, an
PT

innovative candidate to develop more sophisticated products demanded by the modern


society. In this context, pear peel residue could be an alternative raw material to produce
CE

nanocellulose that, to the best of our knowledge, has not been previously comprehensive
investigates for this purpose.
AC

Because of the trend to develop facile, versatile, safe and eco-efficient processes to
diminish harmful by-products, this project has focused on a one-pot fabrication system and
characterization of high purified cellulose-based nanoparticles through Cr(III)-catalyzed
oxidation of pear peel by H2O2. Based on the aforementioned, this driving force of this study
is to explore for the first time the revalorization of the cellulose-rich pear peel residue as a
new source for nanocellulose production. The resulting product was characterized in terms of
chemical composition and nanocellulose yield. Moreover, in order to complete the raw
material and as-obtained nanocellulose characterization, advanced techniques including

6
ACCEPTED MANUSCRIPT

Fourier transform infrared spectroscopy (FTIR), X-ray diffraction (XRD), thermogravimetric


analysis (TGA), field emission scanning electron microscopy (FESEM) and transmission
electron microscopy (TEM) were also performed to determine their physicochemical
properties. This study provides would not only introduce a new isolation approach for
production of spherical and networked-like nanocellulose from fruit residue but also help to
overcome environmental pollution issues. Moreover, the full utilization of fruits could lead
industries to a lower-waste agribusiness and thus increasing industrial profitability.

PT
2. Experimental
2.1 Materials and chemicals

RI
In this study, the raw material, pear (Pyrus pyrifolia L.) peel residue, was freshly

SC
collected from a local fruit market. In order to avoid microbial spoilage and contamination,
the peel fraction was filtered and washed thoroughly with deionized water and sonicated at
NU
40 °C for 15 min in an ultrasonic bath with the purpose of removing the pear juice remnants
and other water-soluble materials commonly found in the cellulosic fibers. The pear peel
residue was collected and dried in air for 2‒3 days. The air-dried pear peal were ground into
MA

fine powder form with a precision grinder and sieved through a ~80 μm mesh.
D

All the chemicals used in this study were of reagent grade and acquired from Merck
Malaysia and used as received without further treatment and purification. These chemicals
E

were obtained from commercial sources. The dialysis membrane used was purchased from
PT

Sigma-Aldrich with a molecular weight cut off between 12000 and 14000. Ultrathin carbon
film on holey carbon support copper TEM grids was purchased from Agar Scientific
CE

Malaysia. Deionized water (Millipore) was used throughout this work.


AC

2.2 Protocol for purification and isolation of nanocellulose


Fabrication of nanocellulose was carried out in the one-pot process: Dried pear peel
powder (10 g) was soaked in deionized water and vigorously dispersed by a magnetic stirrer
for 24 h, filtered to remove the fines and extra water from the pulp. After that, 100 mL of 30
wt% hydrogen peroxide (H2O2) solution was added to the swollen pulp in a round bottom
flask. The reaction process was performed at 90 °C for 5 h with mechanical agitation. The
temperature was well-controlled using a thermometer inside the reaction system. At the end
of the process, the reaction mixture was cooled down in an ice bath system to room
temperature. When the temperature reached to room temperature, 0.22 M of chromium (III)

7
ACCEPTED MANUSCRIPT

nitrate, Cr(NO3)3 metal salt catalyst was added into the same pot suspension together with 8
wt% H2SO4 and further heated up the mixture to 82 °C and held at that temperature for 1 h.
Afterward, the reaction mixture was diluted with ice cubes to quench the reaction and washed
with deionized water by successive centrifugations at 12000 rpm (25 °C) for 10 min for three
times until the hydrolyzing medium was completely removed (constant pH was obtained) in
order to collect the purified supernatant, a translucent suspension of pear nanocellulose. The
purposes of high-speed centrifugation are to concentrate the cellulose crystals, to remove the

PT
non-fibrillated cellulose fractions and to remove the excess reaction medium. The
supernatants consisted of chromium and sulfate ions as well as dissolved oligosaccharides
resulting from cellulose degradation process. The concentration of these ion residues must be

RI
kept below threshold values. This can be achieved by additional centrifugation or washing or

SC
by ion-exchange resins.
NU
The insoluble resultant nanocellulose was dialysis with using regenerated cellulose
dialysis membrane against the deionized water. In this dialysis process, cellulose membrane
acts as a removal of the solvent content of sulfur and metal ion residue of the nanocellulose.
MA

Dialysis fostered the diffusion of excess acid from the nanocellulose suspension toward
medium with lower concentration. The pH of the water was measured and replaced every 12
D

h until constant pH was noticed, reflecting the pH of the nanocellulose. To obtain well-
dispersed nanocellulose suspension and to minimize the agglomeration, the suspension was
E

then ultrasonicated repeatedly for 5 min (pulse 15 sec on and 5 sec off) in an ice water bath.
PT

To prevent localized excessive heating caused by cavitation, a double wall glass beaker
attached to a recirculating cooling system was used to ensure that the solution temperature
CE

was always kept at 20 °C. Finally, the resulting suspension was freeze-dried and stored in a
vacuum-tight desiccator for further physicochemical characterization. Freeze-drying
AC

(lyophilization) was carried out with a freeze dryer/lyophilizer (Labconco) at −40 °C and
drying by sublimation under reduced pressure. A freeze-drying method can be utilized to
minimize the hornification process.

The yield is defined as the ratio of the amount of nanocellulose in the supernatant to
the initial amount of nanocellulose before ultrasonication and centrifugation. Centrifugation
of a dilute nanocellulose suspension was shown to be an efficient means to separate the non-
fibrillated materials from those partially fibrillated fibers which sediment. The protocol was
carried out as follows: A dilute suspension of nanocellulose in water at ca. 0.1% (w/v) was

8
ACCEPTED MANUSCRIPT

centrifuged at 4500 rpm for 20 min to separate the nanofibrillated material (in the supernatant
fraction) from the non-fibrillated and partially fibrillated fibers remains in the precipitate at
the bottom (sediment portion). Then, the sediment fraction was dried to a constant weight at
105 °C. The results represented the average values of three replications. The nanocellulose
yield was calculated according to the relation [27]. At least three different replications of
each sample were considered to calculate the average.

Yield (%) = [1‒ (Wp/Ws)] × 100

PT
where Ws is the weight of nanocellulose before centrifugation and Wp is that of the dry

RI
sediment after centrifugation.

SC
2.3 Testing and characterization
NU
The raw material and isolated nanocellulose were characterized using various spectral
and analytical techniques.
MA

2.3.1 Determination of the composition


Since no literature is available on the composition of pear peel, the first aim was to
D

determine their chemical composition by ASTM standards. The meshed and powdered
sample was oven dried in a commercial hot air drying oven maintained at 100–105 °C for a
E

few hours until a constant weight was achieved. The following analysis was carried out on an
PT

oven-dry sample. Extractives (ASTM D1107-96), Klason lignin (ASTM D1106-56),


holocellulose (ASTM D1104-56) and α-cellulose (ASTM D1103-60) were analyzed. The
CE

difference between the values of holocellulose and α-cellulose gives the hemicellulose
content of the sample. All analytical determinations were performed in duplicate and the
AC

means were calculated.

Soxhlet method has been the most used extraction technique to remove the
compounds, which are not part of the biomass and may interfere with some analyses, by
means of a soxhletation method with the one-step process. Briefly, the raw material (~10 g)
in powder form was placed on a thimble holder, which was filled with the solvent mixture
(1/3 ethanol and 2/3 toluene by volume) from a distillation flask. When the liquid reached the
overflow level, the solution of the thimble holder was aspirated by a siphon and unloaded the
solution back into the distillation flask. This solution carries extracted solutes/extractives into

9
ACCEPTED MANUSCRIPT

the bulk liquid. This process ran repeatedly until the extraction was completed. According to
ASTM D1107-96, adjust the heaters in order to provide a boiling rate which will cycle the
specimens for not less than 24 extractions over a 4 to 5 hour period. After that, the material
was repeatedly washed with fresh solvent and dried in an oven for 1 hour at 105 °C and
cooled in a desiccator until constant weight was accomplished.

The lignin content of pear peel sample was determined in accordance with ASTM

PT
D1106-56. In this method, the isolated lignin was known as Klason lignin, which defined as
the component insoluble in a 72% H2SO4 solution. 1 g of extractive-free sample was treated
with 15 mL of 72% H2SO4 for 2 h at room temperature in order to hydrolyze and solubilize

RI
the carbohydrates components. The sample was then diluted with 560 mL of deionized water

SC
to reduce the acid concentration to 3 wt% and further boiled for 4 h. Next, lignin residue was
allowed to settle before being filtered through filter paper. The residue was repeatedly
NU
washed by hot water until a neutral pH was achieved. The dried insoluble residue represents
the lignin content.
MA

The holocellulose (cellulose and hemicellulose) of sample was determined by ASTM


D1104-56. During the treatment, 4 g of sample was reacted with 5 g of sodium chlorite and 2
D

mL of acetic acid at 70 °C for 4 h. The mixture was then cooled down naturally and the
residue was filtered and washed with deionized water and acetone. The residue was finally
E

dried at 105 °C in an oven until constant weight was achieved.


PT

α-cellulose is defined as the residue of holocellulose insoluble in a 17.5 wt% NaOH


CE

solution. In this procedure (ASTM D1103-60), 5 g of holocellulose (A) was added to a 100
mL of 17.5 wt% NaOH solution at room temperature with continuous stirring for a 30 min
AC

incubating period. Later, 100 mL of deionized water was added to dilute the NaOH, stirred
for 1 min and then left for 5 min. The residue was filtered and washed with deionized water.
Then, the addition of 15 mL of 10 wt% acetic acid served at hydrolyzing the degraded
cellulose and hemicellulose. The residue was filtered into a weighed glass crucible and
washed with hot water until neutralized before drying it to a constant weight (B). The amount
of α-cellulose was finally determined gravimetrically.

α-cellulose content (%) = [(A‒B)/A] × 100

10
ACCEPTED MANUSCRIPT

2.3.2 Fourier transform infrared spectroscopy (FTIR)


The functional groups of raw material and nanocellulose were analyzed by using
Perkin Elmer Infrared spectrometer (Perkin Elmer Spectrum Two). Effects of the treatments
on the chemical compositions can be tracked from the IR spectra. Prior to the experiment,
samples were carefully dried in an air circulating oven at 60 °C for 12 h. A small quantity of
each dried sample was mixed with previously dried anhydrous potassium bromide powder (1%
w/w) and compressed to form a disc. The samples were analyzed in the transmission mode in
the spectral range from 4000 to 400 cm-1 at intervals of 1 cm-1 with the spectral resolution

PT
was 4 cm-1 and an accumulation of 32 scans at room temperature.

RI
2.3.3 X-ray diffraction (XRD)

SC
Crystallinity and phase conformation (purity) of the samples were analyzed by XRD
analysis. Each sample in the form of milled powder was placed on the sample holder and
NU
leveled to obtain total and uniform X-ray exposure. XRD pattern was recorded on a Bruker
diffractometer D8 Advance using a scanning rate of 5s/step and a step size of 4°/min. All
assays were measured in the 5° < 2θ angles < 50° using CuKα radiation (λ = 1.5406 Å) as a
MA

target with a applied accelerator voltage of 40 kV and current of 30 mA at room temperature,


which allowed collecting information from a large area. The dried materials were pressed
D

manually into cylindrical wafers with a diameter of 2 cm and thickness of ca. 0.1 cm. A
silicon zero background plate was used to calibrate the effect of sample holder and to make
E

sure no signal was generated from the sample holder. The same sample holder position and
PT

sample holder (holder and silicon zero background plate) were used for all experiments.
Subtraction of background intensity from the measured intensity of nanocellulose samples
CE

was performed.
AC

The crystallinity index (CrI) was calculated through referring the diffraction intensity
of the crystalline and amorphous zones using the empirical method by the following
experimental equation.

CrI (%) = [(I002 ‒ Iam) / I002] × 100

where I200 is the integrated area of the intensity peak at the plane (200) (2θ = 22.6°) and Iam is
the subtending area for minimum intensity (110) (2θ = 16.5°) of the whole diffraction profile.

11
ACCEPTED MANUSCRIPT

This method is convenient and quick to determine the relative crystallinity of nanocellulose
which can be used for relative comparisons.

2.3.4 Thermogravimetric analysis (TGA)


The thermal stability and degradation temperature of all studied samples were
investigated using a TA Instruments Q500 thermogravimetric analyzer unit. For the analysis,
ca. 6 mg of sample was heated in a porcelain crucible with the heating ramp started from 25

PT
until 700 °C at a heating rate of 10 °C/min. All measurements were performed under the
nitrogen atmosphere with a gas flow rate of 40 mL/min in order to provide an inert
atmosphere for pyrolysis. The differential thermogravimetry (DTG) curves were obtained

RI
from the first derivatives of the weight loss rate. TGA thermal curve was displayed as weight

SC
percent (%) versus temperature.
NU
2.3.5 Field emission scanning electron microscopy (FESEM), energy-dispersive X-ray
spectroscopy (EDX)
The surface morphology of the raw material and isolated nanocellulose was studied
MA

by FESEM. The images were captured by a Fei Quanta 200F microscope at an accelerating
voltage of 10 kV. The samples were dried in an air oven at 60 °C for 12 h before the analysis.
D

A small amount of powdered sample was mounted on a metal stub with adhesive double
sided carbon tape followed by gold coated using sputter coating technique for 20 s using a
E

vacuum sputter coater for 5 min at 30 mA current and 0.4 millibars pressure to suppress the
PT

charging effect inside the FESEM.


CE

Identification of chemical compositions of samples and element mapping analysis


were further carried out by an energy-dispersive spectrometer. This analysis is considered a
AC

semi-quantitative analysis since some analytical variables, such as critical penetration depth
and detection limit might affect the results. EDX spectrometer coupled with FESEM unit was
used for identifying the elemental composition of nanocellulose. The EDX experiments were
conducted at an accelerated voltage of 10 kV. The tabulated results of the quantitative surface
analysis are expressed as both the weight percentages and atomic percent.

2.3.6 Transmission electron microscopy (TEM)


The morphology and dimension of the as-produced nanocellulose were determined
using high-resolution transmission electron microscope (JEOL JEMP-2100F brand). 1 g of

12
ACCEPTED MANUSCRIPT

nanocellulose powder was diluted with GC grade ethanol and the aqueous suspension was
mildly ultrasonicated in a water bath sonicator for 5 min. A drop of a diluted suspension (1
mL) was deposited on the surface of a glow-discharged carbon-coated copper grid and the
excess amount of liquid was removed with a filter paper. Excess solvent was carefully
wicked away from the edge of the grid with a filter paper by means of capillary forces. The
grid then placed in a partially covered glass petri dish to allow evaporation at ambient
temperature under windy condition before TEM analysis and the measurement was operated

PT
at 200 kV accelerating voltage. Compared to other methods widely used so far for the
preparation of nanocellulose sample for TEM analysis using dry particles, the dispersion of
the wet nanocellulose in ethanol medium had the advantage of reducing the preparation time

RI
providing results readily exploitable after dialysis. The use of ethanol as dispersing solution

SC
for nanocellulose was found to have a positive outcome on the preparation process. Since this
solvent is very volatile, it was possible to evaporate it rapidly at room temperature, leading to
NU
a drying period of less than 15 min at room temperature. The nanocellulose dimensions (i.e.
diameter and length distribution) were determined using digital image analyses (ImageJ
software) from the TEM images. A hundred nanocellulose fibers were randomly selected to
MA

determine the average diameter and length.


D

3. Results and discussion


E

3.1 Chemical composition study


The ASTM chemical composition analysis proved that the pear peel residue has
PT

relatively high cellulose content (38.5%) and low non-cellulosic components (23.6%
hemicellulose, 28.1% lignin and 9.8% extractive) levels. The pear peel residue is an attractive
CE

source of nanocellulose because it has more cellulose when comparing with obtained data
with available literature for other raw materials that commonly used. This information has
AC

abbreviated in Table 1. It is very important to select the raw material with high cellulose
content for the nanocellulose production. This is because the percentage of cellulose is a
strong indicator of the potential of this material as a nanocellulose source. In addition to that,
pear peel residue is also available in large amounts because it is a by-product of the food and
beverage industries. Upscaling of pear peel residue for commercial production of
nanocellulose requires a supply with little variation in the cellulose content and low
impurities content and other contaminants. Therefore, pear peel can be used as a potential
source for the fabrication of nanocellulose; furthermore, the pear peel can be effectively
utilized as the filler for the preparation of composites. In this study, the isolated nanocellulose

13
ACCEPTED MANUSCRIPT

rendered comparatively high α-cellulose percentage (97%) with a low amount of


hemicellulose (1.2%) and lignin (1.8%). This finding evidence that the one-pot process
demonstrates the effectiveness and efficient regarding the total delignification of biomass by
hydrolysis of β-ether bonds and also in the oxidation of the reducing groups of carbohydrates.
In addition, the one-pot process resulted in the opening of lignocellulosic structure of the
biomass, causing the hydrolysis of hemicellulose and cleavage of lignin-hemicellulose
linkage. This reflected that one-pot isolation process is highly effective in removal of most of

PT
the lignin, hemicellulose as well as other impurities available in the pear peel residue.

------Table 1------

RI
SC
Figure 1 shows the visual macroscopic evolution of the color changes of pear peel
before and after the treatment as well as the photographs illustrating the aqueous suspension
NU
stability for nanocellulose sample. Definitely, after the treatment, the color of the isolated
nanocellulose from pear peel raw material changed from dark brown to completely white
cellulose pulp was obtained as the end product. The H2O2 treatment was designed to
MA

solubilize lignin, hemicellulose residual and any other extracts in the raw material. The
solubilized components were washed out with water during successive centrifugation, leaving
D

the whiten cellulose pulp. The color change of treated fibers can be directly related to
chemical treatment efficiency, whereas coloring matter from oxidized carbohydrates and
E

residual lignin are completely removed. The delignification treatment with H2O2 was
PT

performed in order to break down, degrades and oxidizes the phenolic compounds or
molecules that consist of chromophoric groups (such as conjugated carbonyls, double bonds
CE

and their combination) present in lignin into non-chromophoric species. The highly reactive
species including hydroxyl radicals, superoxide anion radicals and hydroperoxide anions will
AC

eventually recombine into oxygen and water without leaving the toxic residues in the treated
matrix. A previous study [30] has suggested that the very white color of nanocellulose
corresponded to the absence of non-cellulosic components were probably eliminated within
the fibers. This indicated that the high purity nanocellulose material was obtained via the one-
pot process. Therefore, the disappearance of the dark brown color from the chemically treated
fibers can also be a clear indication that lignin and hemicellulose were removed and agreed
with the chemical composition data. Predominantly, the isolated white color of nanocellulose
is important for industry applications since the colored products could limit its use in the
polymeric composites applications that without color requirement. Color is an important

14
ACCEPTED MANUSCRIPT

property of nanocellulose fiber because it can affect consumer acceptance in many


applications. In this work, the yield of isolated nanocellulose was calculated at 23.7% with
respect to the initial amount of starting raw material. Table 2 tabulates the comparison of the
yield of nanocellulose obtained from various materials via different isolation routes.

------Table 2------

PT
As visually observed in Figure 1, the isolated nanocellulose was considered to be well
dispersed in water, since the nanoparticle/water mixture formed a uniform dispersion with no
obvious sedimentation and precipitation, whereas unstable dispersion state will lead the

RI
formation of aggregates and the two-phase system observed. According to the literature [31],

SC
the color of the suspension is directly related to the purity of nanocellulose, more specifically,
the suspension would appear as brown color if there was still a significant amount of lignin
NU
and hemicellulose within the fiber. The fact that the suspension of pear peel nanocellulose
was milky white in color indicates low lignin content, which can also be correlated to the
chemical analysis result (Table 1). The turbid phenomena of the suspension were highly
MA

compatible with previous literature [32, 33]. The nanocellulose suspension was well
dispersed in water and able to maintain stable without visible aggregation can be observed at
D

the bottom of the bottle due to its nanoscale particle size. Most importantly, the H2SO4 used
in hydrolysis resulted in the incorporation of sulfate groups on the nanocellulose surface
E

which created a strong electrostatic repulsion between the anionic sulfate ester groups at the
PT

surface resulting in a more stable final suspension [33]. A similar phenomenon has been
previously reported, where the stability of cellulose nanocrystal dispersion involves
CE

complicated phenomenon that is affected by two main parameters, namely the surface charge
of nanocellulose and the interaction between nanocellulose and the dispersion medium. The
AC

nanocellulose isolated by hydrolysis using sulfuric acid has a negatively charged surface
(OSO3−) and abundance of OH groups, which lead to stable aqueous dispersion of
nanocellulose. In this case, the interaction between nanocellulose and the medium occurs by
electrostatic repulsion of negative charge and hydrophilic affinity [34].

------Figure 1------

3.2 FTIR characterization

15
ACCEPTED MANUSCRIPT

FTIR spectroscopy has been extensively used to study the functional groups of
lignocellulosic biomass and the changes caused due to different treatments. The spectra offer
qualitative and semi-quantitative information suggesting the absence and presence of
lignocellulosic compounds, and whether the intensity of a peak has changed after treatment.
This technique is well known for its simplicity in sample preparation and speed of analysis
[35]. FTIR spectra of raw material and isolated nanocellulose are illustrated in Figure 2.
Generally, two main absorbance regions can be found at high wavenumber (4000‒2800 cm-1)
and at low wavenumber (1700–800 cm-1). A broad peak at 3400 cm-1 observed in all samples,

PT
which was associated with stretching vibration of inter- and intra-molecular hydrogen bonded
O‒H groups. As clearly observed, this absorption peak in the raw material was similar to that

RI
in isolated nanocellulose. This implied that most of the crystalline cellulose in the raw

SC
material was not disrupted during the nanocellulose isolation process [36]. Likewise, this
peak was attributed to the hydrogen bond of O‒H stretching and bending vibrations of the
NU
adsorbed water that mainly attributed from the environment, which is related to the
hydrophilic character of cellulosic materials [7]. The peak observed at 2900 cm-1 was due to
the aliphatic saturated symmetric C–H unit vibrations and CH2 stretching. The peak at 2850
MA

cm-1 was due to C-H deformation within the aromatic methoxyl group of lignin [37]. This
peak might be also attributed to asymmetric and symmetric methyl and methylene stretching
D

groups of the extractive content in the fibers as some compounds in organic extractives, such
as phenolic acid methyl esters and fatty acid methyl esters, contain these functional groups
E

[38]. The vanishing of this peak in nanocellulose spectra was indicative of the total removal
PT

of a high fraction of extractive and lignin during the treatment.


CE

The adsorption peak at 1730 cm-1 was detectable in raw material and has been
proposed to be associated with the alkyl ester of the acetyl group in hemicellulose component
AC

[36]. This band also was characteristics of the acetyl or uronic ester groups of hemicellulose.
Another possibility for this peak is that aldehyde or carboxyl absorption could have arisen
from the oxidation of the C–OH groups or opened terminal glycopyranose rings. In addition,
this band found at this region was assigned to the tension of the C–O bond at C=O of lactonic
ester and the tension of the C=C bond of vinyl or aromatic compounds. The ester linkage
C=O with an absorption peak at 1700 cm−1 was usually defined as the acetyl group in
hemicellulose structure as well as the linkage between lignin and hemicellulose [36].
Furthermore, the band at 1275 cm-1 was assigned to the syringyl ring breathing and C‒O
stretching in xylan. However, all these bands were completely absent in nanocellulose as a

16
ACCEPTED MANUSCRIPT

result of the successive removal of the hemicellulose from the pear peel during the treatment.
Similarly, the disappearance of the typical absorption bands of hemicellulose component, α-
glucan at 1040 cm-1, reflected the effectiveness of one-pot process in dissolution of
hemicellulose within the fiber to produce highly purified nanocellulose [35]. Moreover, the
diminished of the peak at 810 cm-1 in the nanocellulose spectra, which resulted from the in-
phase ring stretching of mannose residue, suggesting the degradation of glucomannan
(hemicellulose component) has occurred during the one-pot process.

PT
For all samples, the peak in the region of 1640 cm-1 was indicative of O–H bending
originated from absorbed water in the fiber. Even though all FTIR samples were subjected to

RI
a proper drying process, the complete removal of water molecules was very difficult due to

SC
the strong cellulose-water interaction [30]. The higher moisture content of nanocellulose may
be attributed to the absorption of moisture in the spaces left vacant after the removal of lignin
NU
and hemicellulose. This was due to open surfaces created in the chemically-treated fibers help
absorb moisture. Also, the higher moisture content for nanocellulose may probably due to the
higher cellulose content.
MA

Besides hemicellulose, lignin is also one of the main components in the


D

lignocellulosic biomass, thus the distribution of lignin-related bands also investigated. The
peaks at 1530, 1460, 1313 and 1247 cm-1 in the raw material spectrum were associated with
E

the characteristics of lignin. The small shoulder peak at 1530 cm-1 was attributed to the C=C
PT

unsaturated linkages in plane symmetrical stretching of aromatic rings present in lignin [39].
The bands at 1460 cm−1 have been assigned to absorption due to C-H deformation within the
CE

methoxyl groups of lignin. Furthermore, the adsorption peak due to aromatic hydroxyl groups
can be observed at 1313 cm-1 [35]. These adsorptions may be generated by the ether bonds
AC

within the lignin. This is because hemicellulose can form a linkage between lignin and
cellulose, and lignin-carbohydrate complex with lignin by ether bonds. Besides that, these
two peaks have been assigned to aromatic skeletal stretching. Furthermore, the C‒O groups
of the guaiacyl unit stretching vibration in lignin resulted in an adsorption peak at 1247 cm-1
[30]. The peaks detected at 856 cm-1 could be assigned to the C‒H out-of-plane deformation
vibration of lignin. Since all the adsorption of the peaks as mentioned above was disappeared
from the spectrum after the pear peel material had undergone the one-pot process. This
suggested that the complete removal of lignin in the nanocellulose sample has been achieved.

17
ACCEPTED MANUSCRIPT

This result is in good agreement with the chemical composition analysis, where the
nanocellulose sample exhibited the negligible lignin content of 1.8%.

The occurrence of a peak at 1436 cm-1 for all samples is due to CH2 asymmetric
bending motion in cellulose [30]. A previous study also suggests that this peak was owing to
the intermolecular hydrogen attraction at C6 group and the symmetrical angular deformations
of methylene groups (constituents of cellulose repeating unit). [40]. Some important spectral

PT
bands can be monitored to investigate the presence of cellulose parent chain were including
1375 cm-1 (asymmetric C–H deformation or C–H bending), 1322 cm-1 (CH2 wagging), 1164
cm-1 (C–O antisymmetric stretching), 1110 cm-1 (C–C and C–O stretching), 1060 cm-1 (C–O–

RI
C pyranose ring stretching vibration) and 890 cm-1 (C‒H rock vibration) [30]. In all samples,

SC
the observed peaks at 1110 and 890 cm-1 were associated with the presence of β-glucosidic
ether linkages correlated to the anomeric vibration modes of anhydro-glucopyranose ring
NU
skeleton, and to the β-glycosidic linkages between the anhydroglucose rings in the cellulose
chains [7]. According to Kawee et al. [41], the spectra of samples showed the characteristic
bands around 890 and 1436 cm-1, defining them as cellulose type I. It can be observed that
MA

the band positions in the FTIR spectra for all samples had no significantly different peaks and
no change in cellulose I type. It was clearly seen that the intensities of these two bands were
D

increased in the nanocellulose spectrum as compared to the raw material. This suggested that
the more cellulose content was exposed in the case of isolated nanocellulose due to the
E

successive degradation and removal of non-cellulosic contents (i.e. hemicellulose and lignin)
PT

in the raw material after the one-pot process. Likewise, the peak at 890 cm-1 remained
unchanged in the isolated nanocellulose proved that the β-ᴅ-glucopyranosyl structure in
CE

cellulose chains was not affected during the treatment [10]. Based on these findings, it is
evident that the nanocellulose was successfully isolated from pear peel and the end product is
AC

free of hemicellulose and lignin. The obtained results are in accordance with literature [14, 15,
40]. On the other hand, the peak at 890 cm-1, which determines of changes in the crystallinity
of material (the C1 group frequency or ring frequency is characteristic of β-glycosidic
linkages between the sugar units), became more prominent. The weak band of 890‒875 cm-1
was assigned to the absorption of the β-stereochemistry configuration of the ᴅ-glucopyranose
in cellulose. According to Prieto-García et al. [42], the nature of this absorption band in the
spectrum of cellulose and other polysaccharides have a complex character. Thus, this band
was weak in natural cellulose and its intensity depends on the subjected treatment. Therefore,
during the purification process, the intensity of the mentioned signal might be increased [42].

18
ACCEPTED MANUSCRIPT

Unfortunately, no 807 cm-1 band was observed (related to symmetric vibration of C‒O‒S
bonds associated with C‒O‒SO3− groups), which are derived from esterification occurred
during the hydrolysis reaction. Similar observation has been reported by [43], which the
FTIR spectra of H2SO4-hydrolyzed nanocellulose remained similar to HCl-treated
nanocellulose although the sulfur groups were introduced to the nanocellulose surfaces. The
absence of this peak could be due to only little amount of sulfur presented (determined by
EDX analysis, Section 3.5) on the isolated nanocellulose compared to other main components.

PT
Therefore, it is hard to be determined.

Comparable results have been obtained by Yahya et al., [10] who obtained similar

RI
results while fabricating the nanocellulose from various starting materials. Chen et al. [7] also

SC
reported the same result when isolating nanocellulose from Gelidium elegans using H2SO4
hydrolysis. The FTIR spectra confirm that the one-pot process performed to obtain
NU
nanocellulose from pear peel does not affect the chemical structure of the cellulosic
fragments. The surface morphologies may however be affected. In addition, there were no
new absorption peaks recorded in the nanocellulose spectra compared to the raw material
MA

curve. It can be concluded that no new functional groups were generated or metal salt ion
attached to the surface of fibers and indicated that the chemical structure of the nanocellulose
D

during the one-pot process.


E

------Figure 2------
PT
CE

3.3 XRD analysis


XRD studies of raw material and isolated nanocellulose were carried out to
AC

investigate the possible change in crystalline structure behavior and crystallinity index after
the chemical treatment. The crystallinity of nanocellulose is an important factor for
determining its thermal and mechanical properties. Native cellulose typically consists of two
crystalline forms in a single microfibril, namely Iα and Iβ. Iα shows triclinic unit cell, (100),
(010) and (110), containing one cellulose chain, whereas Iβ displays monoclinic unit cell, (1-
10), (110) and (200), which possess of two parallel cellulose chains. Figure 3 shows the
equatorial XRD pattern and their relative degrees of crystallinity of all the samples. XRD
pattern of all the samples display a well-defined polymorph of typical cellulose Iα structure
with an amorphous broad hump and crystalline peak. According to cellulose-ICCD

19
ACCEPTED MANUSCRIPT

(International Centre for Diffraction Data),the presence of cellulose type I can be observed by
the presence of characteristic peak at Bragg’s angle (2θ) around 16.2°, 22.5° and 34.6°, which
corresponding to the (110), (200) and (004) lattice planes, respectively. This finding is in
good agreement with the FTIR result. The values of crystallinity index (CrI) of all samples
provide a qualitative or semi-quantitative method to evaluate the amounts of amorphous and
crystalline cellulosic components in fibers. The crystallinity index is related to a large number
of secondary molecular bonds in the crystalline regions and to the level of compaction in

PT
these regions [44].

The crystallinity index of the raw material and isolated nanocellulose was found to be

RI
35.4 and 85.7% respectively, as determined from Segal’s equation. Low crystallinity was

SC
observed for the fibers given that crystalline domains were embedded in a matrix of
amorphous components, such as hemicelluloses, lignin and pectin. It is noteworthy that the
NU
diffraction peaks of the nanocellulose, especially positioned at 2θ value of 22.5° become
sharper and resolved. Compared with the raw material, the nanocellulose had broader peaks.
A good definition of diffractogram revealed that the sharpest peak at 2θ = 22.6 accounted for
MA

the superior (200) crystallographic peak, suggesting a significant increase of crystalline part
in the isolated nanocellulose. This finding suggested that the CrI of the samples increase
D

proportionally during the chemical treatments without altered the crystal structure of
nanocellulose. The increment of CrI from the raw material to nanocellulose was undoubtedly
E

ascribed to the priority on the removal of non-cellulosic compounds which exist in the
PT

amorphous nature (namely wax, extractive, lignin and hemicellulose) during one-pot process,
as confirmed by FTIR analysis (Figure 2) and enrichment in cellulose content in the treated
CE

fiber. Such phenomenon could be related, to the concentration reduction of most of the
residual hemicellulose and lignin, a tendency that was also reported by Fillat et al. [27]
AC

following the pretreatment of olive tree pruning residue. Most importantly, the increasing of
CrI value also indicated that the successive hydrolytic cleavage of glycosidic bonds between
anhydroglucose units in the cellulose amorphous parts. This action is counteracted by
rearrangement of the tangling chain ends, which is favored by release of internal strain,
leaving behind the high crystalline domains without any destruction. This is because the non-
crystalline region is more susceptible to the hydrolyzing agent than the crystalline domains as
the hydrolysis kinetics of amorphous regions is faster than the crystalline ones. Besides that,
the rearrangement of the glucose chains in cellulose molecules into more ordered and
crystalline structure after the dissolution of amorphous phase resulted in higher crystallinity

20
ACCEPTED MANUSCRIPT

[45]. Dissolution of the amorphous cellulose domain rendered it to become more susceptible
to hydrolysis than the crystalline parts. With progressing chemical treatments, the peak (200)
remained constant, which ensures the structural integrity of cellulose type I had been
maintained [46]. It indicates that the original crystalline structure of cellulose fibrils was well
maintained during one-pot process. This finding is highly consistent with the reports about
the crystallinity changes of nanocellulose caused by ionic liquid [47] or sulfuric acid [30, 48].
Further, the strong peak assigned to the (004) plane appeared for the nanocellulose suggested

PT
that one-pot reaction would provide relatively mild conditions to preserve the crystalline
domains of nanocellulose during the chemical treatment.

RI
Herein, the CrI of isolated nanocellulose prepared by one-pot process rendered is

SC
higher to that of nanocellulose extracted from degreasing cotton (62.7%; H2SO4 hydrolysis)
[32], spruce wood (63%; glycerol-methane sulfonic acid hydrolysis) [4], rice straw (63.4%;
NU
ultrasonication) [46], Colombian Fique leaves (65%; TEMPO-mediated oxidation) [25] and
empty fruit bunch (76%; ammonium persulfate oxidation) [9]. The high CrI for the isolated
nanocellulose also indicated the higher order and more hydrogen-bonding interactions in this
MA

material. Crystallinity index of nanocellulose is the prime importance to be considered during


the isolation process, as this is the main factor to determine the mechanical strength and
D

reinforcing capability of nanocellulose to be utilized in nanocomposite applications. The high


crystallinity nanocellulose is expected to be more effective in providing better reinforcement
E

potential for nanocomposite material due to its high stiffness and rigidity, resulting in a high
PT

Young’s modulus of the crystal region along the longitudinal direction [46]. Thus, the
absence of significant changes in the total crystallinity of high crystalline nanocellulose
CE

proves that the one-pot reaction has great potential in the production of nanocellulose
particles. In this sense, high crystallinity nanocellulose could more effectively provide better
AC

reinforcement to the flexible polymeric matrices due to its high elastic modulus of the crystal
regions while fewer crystalline domains are more flexible than crystalline zones, which can
provide improved ductility to stiff polymeric matrices.

------Figure 3------

3.4 TGA examination


The thermal degradation behavior of the raw material and isolated nanocellulose
samples was investigated using thermogravimetric analysis. TGA can be used to characterize

21
ACCEPTED MANUSCRIPT

material that exhibits a weight change upon heating and to detect the phase change due to the
decomposition process. The obtained TG/DTG curves of all samples are shown in Figure 4.
In general, the thermal decomposition of these fibers consists of several phases including
moisture evaporation, breakdown of lignocellulosic components of hemicellulose, followed
by cellulose, lignin and finally the ash. The first phase was the evaporation of the moisture
absorbed in fibers ranging from 45 to 125 °C. All samples presented the first weight loss
beginning at the temperature lower than 100 °C, which is attributed to the evaporation

PT
absorbed water, release of moisture and vaporization of low molecular mass compounds
(such as extractive) bonded on the surface and/or inside of the cellulosic fibers [49]. On the
other hand, less volatile extractives tended to migrate to the fiber surface together with the

RI
interlayer coordinated water molecules since the water on the fiber surface was evaporated

SC
[50]. Hence, these extractives compounds were tended to leave the fiber together with the
water. The second decomposition phase related to the degradation of hemicellulose. As the
NU
temperature is increased, there were chemical changes in their hemicelluloses components
due to the cellular breakdown [51]. Hemicellulose is made up of various saccharides
(galactose, glucose, mannose, xylose, etc.) that exist as amorphous, randomly-ordered and
MA

rich of branches. Therefore, it can be easily removed from the fiber and degraded into
volatiles compounds (i.e. CO, CO2, and some hydrocarbons) that can be evolving out at low
D

temperatures (around 220–315 °C). For this reason, hemicellulose is less thermally stable and
tended to be degraded to a greater extent compared to other lignocellulosic macromolecular
E

components [52]. Once the hemicellulose had completely decomposed, the degradation of
PT

cellulose will take place as the third decomposition phase. Cellulose composed of a number
of chains which are linked via extensive hydrogen bonding networks to form the microfibril
CE

which acts as the primary reinforcing agent in the cell wall. Due to the highly crystalline
nature of cellulose, it normally has relatively thermally stable and only started to be
AC

decomposed at the higher temperature (above 315 °C). The cellulose degradation step
involved several reactions, including depolymerization, dehydration, and decomposition of
anhydroglucose units, with the production of flammable volatiles, which is released through
dehydration, decarboxylation, hydrolysis, oxidation and transglycosylation [53]. The fourth
phase was related to the decomposition of lignin. Lignin is the most difficult to degrade
compared to cellulose and hemicellulose which may be due to highly stable aromatic
compounds. The decomposition of lignin started as early as 160 °C and extended to 900 °C to
complete its composition. This is due to the lignin is a tough component and responsible for
providing the rigidity to the plant materials as well as serving to bond individual cells

22
ACCEPTED MANUSCRIPT

together in the middle lamella region [51]. Finally, as soon as the lignin had completely
decomposed, the component left was referred to as inorganic material in the fibers, which can
be assumed as ash content. This may be due to the presence of inorganic materials such as
silica in the fiber which only decomposed at a very high temperature of above 1700 °C.

The onset of thermal decomposition (Tonset) indicates the beginning of the degradation
temperature and the temperature of maximum decomposition (Tmax) refers to the temperature

PT
of the maximum rate of degradation. Tmax is the temperature where the highest rate of thermal
decomposition occurs and can be used to compare the thermal stability of different materials.
These confront degradation at different temperatures from 100 to 700 °C due to differences in

RI
chemical structure. For raw material, the weight loss started at an earlier stage at Tonset value

SC
around 185.7 °C with the Tmax value of 363 °C. These results likely reflect the decomposition
of non-cellulosic organic compounds, such as wax, pectin, hemicellulose and lignin present
NU
in the raw material. These components reduce the crystallinity of material and lower the
thermal degradation. The temperature dislocation can be associated with the effectiveness of
removing the amorphous component by the one-pot process, as observed in the visual
MA

analysis and FTIR analysis, and further confirmed by the chemical composition results. If
depolymerization occurs more extensively than dehydration, volatilization of tar (high boiling
D

products) mostly composed of levoglucosan is observed. The high boiling products might be
further decomposed, with the production of light flammable gases, if it is not promptly
E

removed from the heating zone. On the other hand, a degradation peak was clearly observed
PT

at 224.1 °C in TGA curve. This finding was related to the presence of hemicellulose which
degraded at much lower temperature. Based on TG curve, the mass loss rate for both curves
CE

steadily showed in reaching the equilibrium state at temperature of about 400 °C and there
was no significant mass loss since almost lignocelluloses composition had been decomposed.
AC

As expected, the isolated nanocellulose exhibited lower Tmax (358.5 C) compared to


its raw material with Tonset value of 284.5 °C. The sulfuric acid hydrolysis decreases the
thermal stability of cellulose crystals due to the insertion of active sulfate groups (SO3‒)
during the hydrolysis and resulted in the replacement of OH by SO3‒ groups which can lower
the activation energy of the degradation of nanocellulose [10, 32]. The presence of sulfate
groups on the surface of the nanocellulose rendered the catalytic effect in the thermal
decomposition mechanism, resulted in the nanocellulose sample is less resistant to pyrolysis
[54]. It is well known that the H2SO4-treated nanocellulose showed two distinct stages of

23
ACCEPTED MANUSCRIPT

cellulose degradation, which the first decomposition process may associate with the
degradation of sulfated nanocellulose and the secondary degradation process may relate to the
breakdown of the interior unsulfured crystals [55]. However, the nanocellulose isolated in
this study was displayed one less stage of degradation. This can be explained by the fact that
fewer sulfate groups (less sulfate content) were introduced on the surface of nanocellulose
chains as only diluted H2SO4 (4 wt%) was involved in this study, with the expectation that
the catalytic effect was less pronounced in this case and therefore, the isolated nanocellulose

PT
would not degraded at lower temperature than typical concentrated H2SO4 hydrolysis. This is
not surprising as similar observation has been reported by [55], which their study found that
the nanocellulose hydrolyzed by lower sulfuric acid concentration (ca. 50 wt%) showed only

RI
one degradation peak in comparison to those isolated by high acid concentration (64 wt%)

SC
that possessed of two-stage degradation pattern. Therefore, it is widely accepted that the
nanocellulose suspension obtained with lower acid concentration displayed fewer degradation
NU
stages which closer to the raw material curves.

In addition, the decomposition temperature of nanocellulose was found to be


MA

somewhat lower than the raw material because of the massive reduction in molecular weight
of nanocellulose [47]. Some factors may result in the reduction of thermal stability of
D

nanocellulose. First, surface sulfation of nanocellulose due to adhesion of sulfate groups


caused the activation energy required for the thermal decomposition was lowered [56]. In fact,
E

the elimination of sulfuric acid in sulfated anhydroglucose units required less energy and
PT

being released at much lower temperature during heating. The released sulfuric acid would
further assist the depolymerization of cellulose by removing some of the hydroxyl groups
CE

either by esterification or direct catalysis mechanism. Second, thermal stability diminished


for nanocellulose due to the smaller fiber dimensions which lead to high surface areas
AC

exposing to heat in response to an increase in temperature [47]. Third, the lower


decomposition temperature might indicate faster heat transfer in nanocellulose. The
nanocellulose acts as the function as efficient pathways for phonons [48]. Fourth,
nanocellulose rendered more contact points compared to raw material and this could induce
thermal degradation of surrounding fiber when some nanocellulose fibers started to degrade.
Reduced thermal stability for nanocellulose isolated via different hydrolysis routes has been
reported, as presented in Table 3.

------Table 3------
24
ACCEPTED MANUSCRIPT

Regarding the reaction residuals, it was remarkable that production of carbonaceous


residues (charring phenomenon) from nanocellulose (10.88%) was almost 2.3 times
significantly lower than in raw material (24.79%). The char yield at 600 °C indicated
nonvolatile char material generated by pyrolysis. Ideally, pyrolysis of cellulose should
produce an ideal yield of 44.4% of carbonaceous residuals during the dehydration reaction
since all hydrogen and oxygen would be liberated in the form of H2O [25]. However, the

PT
cellulose pyrolysis contains other complex reactions including decarboxylation to CO2,
decarbonylation to CO and reactions to form other gaseous products (H2 and CH4) prevent
this from happening. In brief, the levoglucosan breaks down to provide a variety of low

RI
molecular weight fission products, including hydrocarbons and hydrogen as well as CO2, CO

SC
and H2O at a temperature above 600 °C. Secondary reactions may take place in the gas phase
as further decomposition of levoglucosan, in the solid phase as the condensation and
NU
crosslinking of intermediate chars to highly condensed polycyclic aromatic structures, or by
interaction of both phases as gasification of char by reaction with CO2 and H2O at high
temperatures to produce CO and H2. In this study, higher amount of residue in the raw
MA

material directly correlated to the presence of hemicellulose, lignin and other non-cellulosic
materials. During the chemical treatment, subsequent loss of non-cellulosic materials and
D

impurities in the matrix system occurred which resulted in lower residue percentage.
Therefore, the lowest char residue was obtained for isolated nanocellulose. Similar results
E

were found in the isolation of nanofibrils from the Helicteres isora plant [57] and moso
PT

bamboo culms [58]. Considering the thermal properties, the isolated nanocellulose in this
study has a potential for use as reinforcing filler in thermoplastic composites as the typical
CE

processing temperatures for polymeric composites (i.e. thermoplastic) is rise around 200 °C.
Therefore, the thermal resistance of isolated nanocellulose from pear peel residue via one-pot
AC

process remains sufficient to ensure that the nanocellulose can be processed at approximately
200 °C.

------Figure 4------

3.5 FESEM investigation


The effect of chemical treatment on the morphological structure of the surface and
bulk of pear peel was characterized using FESEM analysis. It is expected that the one-pot
process applied for nanocellulose production will induce morphological changes on the

25
ACCEPTED MANUSCRIPT

material. Thus, FESEM images demonstrate the morphological differences between raw
material and isolated nanocellulose. Morphological examination of the nanocellulose is
essential as the cellulose sources and hydrolysis route has a notable influence on the
properties and dimension of nanocellulose. Figure 5(a) shows the raw material had a smooth,
uniform and oriented surface with a bundle of fibers of an average of 6.22 ± 1.2 μm of
diameter. The original pear peel residue showed large fiber bundles and intact structures.
Similar to the literature [59], the micrographs for the raw material suggested the presence of

PT
globular wax particles, extractives, pectin as well as the amorphous constituents (i.e. lignin
and hemicellulose). The raw pear peel fiber had significantly bigger diameter with
presumably lower aspect ratio.

RI
SC
The effect of one-pot process was evident from the comparison of micrographs in
Figure 5(b). After one-pot process, some cell tissues had been destroyed and the nano-sized
NU
fibrils were obviously peeled off from the micro-sized fibers, conferring a rougher surface
with destroyed cell tissues. The increase in surface roughness can be contributed to the
successful removal of non-cellulosic layer (lignin and hemicellulose) as well as cellulose
MA

amorphous phase during the treatment. This fact was further confirmed by TEM analysis. It is
reasonable to conclude that the increase in surface roughness was produced by the chemical
D

treatment process since the hemicellulose coat the cellulose fibrils and lignin fill up the empty
spaces [34]. The removal of these non-cellulosic components is interesting because it
E

facilitates the diffusion of the hydrolyzing catalysts into the cellulose core during isolation of
PT

the nanocellulose. Therefore, the rougher surface after chemical treatments is one of the
indications of removal of the outer non-cellulosic layers composed of cementing materials
CE

including lignin, hemicellulose, wax, pectin and other impurities contained in the raw
material [60]. FESEM images also displayed that the isolated nanocellulose fibers were
AC

cylindrical in shape.

In addition to that, it can be clearly observed that there was a considerable reduction
in fiber size after one-pot process due to fiber defibrillation process. In fact, both pectin and
wax are known as the protective layer that surrounds the surface of the fiber. Besides that, it
is observed that the nanocellulose fiber bundles were separated into individual fibers and the
diameter of the fibrous material decreased to an average of around 1.24 ± 0.3 μm. This fiber
decrement indicated that under the strong chemical treatment conditions, almost all the non-
cellulosic components that bind the fibril structure together were degraded and removed, thus

26
ACCEPTED MANUSCRIPT

enabling the fibers to be separated into more individual form. This size reduction observed
was mainly attributed to the separation of the fibers primary cell wall due to the consecutive
removal of hemicellulose and lignin [60]. Most importantly, the size reduction will eventually
increase the aspect ratio (L/d, L being the length and d the diameter) of nanocellulose, which
should provide the higher reinforcing capability of the fibers for composite applications. This
gave rise to the presumably high aspect ratio of nanocellulose compared to the lower aspect
ratio of the raw material. This result was highly comparable to other reported works [47, 48].

PT
Considerable fiber size reduction, coupled with a loss of fibrous character occurs,
indicating the surface etching and erosion nature of the hydrolysis process [48]. For

RI
transportation, storage and generation of various structures, drying process of nanocellulose

SC
is important. During the freezing process, the nanocellulose fibers tended to gather around the
edge of ice crystals and massive hydrogen bonds are formed among the adjacent surfaces of
NU
nanocellulose. This behavior could be related to the greater interaction between adjacent
cellulose chains. As a result, the nanocellulose interacted and self-aggregated with each other
into long fibers through hydrogen binding in both longitudinal and lateral directions and
MA

eventually organized into larger-sized cellulose fibers or 2D sheet-like structures [61]. The
chemical treatment exposed the cellulose core (which consists of abundant in OH groups) due
D

to the removal of the hemicelluloses and lignin, which favors hydrogen bonding between
adjacent nanocellulose chains. This attraction of cellulose chains can lead to agglomeration of
E

nanocellulose fibers, as observed in Figure 5(b), which is a common phenomenon reported by


PT

many authors [10, 62]. According to Liu’s study [61], the nanocellulose fibers tended to be
aggregated into bundles during drying process. The drying process of nanocellulose
CE

suspension is important for its storage, transportation and generation of various desired
structures. Therefore, the real structures of nanocellulose are not likely to be understood by
AC

FESEM images since FESEM has two main drawbacks preventing observation of the native
fiber structure: the use of coatings, which increase the size of the nanomaterial, and self-
aggregation processes driven by solvent removal as result of nanocellulose drying.

The detailed explanation regarding the self-assembly mechanism of nanocellulose


during freeze-drying has been reported by Han et al. [63]. The lyophilization process consists
of two important steps, i.e. freezing (growing of the ice crystals) and drying (sublimation of
the ice molecules). Upon freezing, the water in the nanocellulose aqueous suspensions is
frozen into ice crystals and concentrating nanocellulose to closer contact with each other in

27
ACCEPTED MANUSCRIPT

both axial and lateral directions. After that, most of the nanocellulose particles above the
critical particle size are trapped by the moving water-ice front and confined into the
interstitial spaces between ice crystals. The final structure of nanocellulose foam is formed as
ice crystals are sublimated during the drying step. Sublimation of ice crystals leaves pores
among the fibers. Therefore, in addition to the freezing conditions, the particle size, surface
charge, suspension concentration and their relationship with the self-assembly behavior of
cellulose particles are theoretically related to the morphology and properties of the final foam.

PT
Both fibrillar morphologies and associated inter-fibrillar pore structures may be further
exploited to generate new fibrillar network structures and porous materials.

RI
FESEM technique has been widely utilized for morphology characterization of

SC
nanocellulose. Prior to imaging, coating the surface of the sample with gold or platinum is
needed. However, this resulted in broadening the nano-sized structure of fiber and probably
NU
reached millimeter-scale. Moreover, the nanocellulose fibers might to aggregate into thick
nanofibers or be dried into thin sheets during the drying process. Compared with traditional
FESEM, the resolutions of TEM are higher and coating the sample surfaces before
MA

observation is not necessary. The TEM analysis can more clearly evaluate the size of single
nanocellulose fiber and web-like twisted structure [61].
D

Direct information about elemental analysis and chemical composition of


E

nanocellulose was determined using EDX attached to the FESEM. The EDX profile of the
PT

nanocellulose is as shown in Figure 5. The characteristic peaks around 0.2, 0.5 and 2.2 keV
were corresponded to the binding energies of carbon, oxygen and sulfur, respectively. The
CE

results reveal the basic components of isolated nanocellulose with their respective mass and
atomic distribution within the sample are also summarized Figure 5. Based on the analysis, it
AC

manifested that the nanocellulose presence of 0.30 wt% elemental sulfur impurity along with
the main components (47.77% carbon and 51.93% oxygen). This amount of sulfur impurity is
most likely originated from the sulfate group that resulted from sulfuric acid hydrolyzing
agent and was remaining to some extent after dialysis of nanocellulose and unable to remove
completely during washing step. Moreover, the adsorption of sulfate group on the surface of
cellulose during esterification was believed to account for its presence. Most importantly, no
energy peaks were detected for elemental Cr, indicating that the obtained nanocellulose was
free of metal ion, which were originally from the metal salt catalyst employed. This also
proven that Cr3+ could be completely removed by centrifuge washing. Thus, the thermal

28
ACCEPTED MANUSCRIPT

stability of nanocellulose would not be affected by the adsorption of Cr3+ on the surface.
From the results, there was no evidence of metal ion elemental impurities were introduced
during the chemical treatment steps involved. Sulfated nanocellulose was prepared according
to the procedure shown in below:

Cellulose‒OH + SO3H‒OH → Nanocellulose‒OSO3H + H2O

PT
------Figure 5------

3.6 TEM observation

RI
The morphology and dimensions of isolated nanocellulose were carefully examined

SC
by TEM analysis, as presented in Figure 6. Imaging of raw material was not possible due to
the high dimensions of the particles. The size of nanocellulose is dependent on the hydrolysis
NU
conditions as well as its source. Figure 6(a‒b) clearly displayed that the isolated
nanocellulose had web-like network structure with long entangled cellulosic filaments
morphology, thus confirming their successful extraction from pear peel residue at the
MA

nanoscale. These results demonstrate the efficiency of the conditions used in the one-pot
process and confirm that the aqueous nanocellulose suspension contained individual
D

nanocrystals. Tibolla’s research group noticed the nanocellulose isolated from cassava root
bagasse and peelings with similar characteristics when observed by the TEM technique [44].
E

A tendency of agglomeration could be observed from TEM. This happens could be due to
PT

drying process of the suspension onto the carbon film covering the copper grids or if it
reflected the state of the suspension. The nanocellulose tended to be aggregated in parallel, an
CE

observation reported in previous studies [9, 14].


AC

The average diameter and length of isolated nanocellulose sample were examined by
analyzing the high-magnification TEM images using digital image analysis (Image J
software), which confirmed that the nanocellulose was extracted at nano-sized scale. From
the analysis, the average diameter and length of nanocellulose was 20.5 ± 6.3 nm and 503.8 ±
30.5 nm, respectively, with the aspect ratio of 24.6. The formation networked-like
nanocellulose can be attributed to several possible reasons: The hydrogen bonding interaction
between cellulose fibers making it possible to regulate the network-like structure of
nanocellulose fiber based porous material during freeze-drying process. Besides that, the
strong hydrogen bonding among nanocellulose overcomes the repulsion of surface negative

29
ACCEPTED MANUSCRIPT

charges, leading to the formation of self-assembled porous networks [64]. Another probable
explanation is that the network structured nanocellulose was actually formed by the over-
irradiation of electron beams during TEM observation [48]. Based on Chirayil’s study [57],
the network structure of nanocellulose was formed due to a number of branches of smaller
bundles or partly individualized nanofibrils were hooked up to larger aggregates. In other
words, the chains twisted or bundled together randomly to neighboring chain and networking
occurred when chains were split to join different bundles. Thus, large aggregates consisting

PT
of wire like cellulose fibrils with nanoscale widths are observed. It has been investigated that
the possibility of using this network-like structure material with nanosized pores has great
potential applications as high efficient sustainable building material, filter in food, air and

RI
medical fields.

SC
Interestingly, spherical-shaped nanocellulose was also being observed through TEM
NU
analysis, as clearly illustrated in Figure (c‒d). The spherical nanocellulose particles rendered
the mean size of 37.3 ± 4.1 nm. These two different forms of nanocellulose (namely,
network-structure and spherical-shaped) could not be separated by centrifugation or filtration
MA

and have not been reported by others. The formation of spherical nanocellulose could be due
to self-assembly of short cellulose rods via interfacial hydrogen bonds. Besides that, the
D

ultrasonic treatment might play an important role in forming the spherical nanocellulose [48].
The ultrasound in hydrolysis seems to play a role in keeping the reactants in good contact,
E

enhancing the hygroscopicity of the cellulose, facilitating the molecular penetration of acid
PT

into the cellulose and increasing the hydrolysis reaction rate. As reported by
Naduparambath’s study [48], it is believed that the network structured nanocellulose is
CE

actually embedded in the abundant spherical cellulose nanocrystals in the freeze-drying


process. Some scholars have previously studied the production of spherical cellulose
AC

nanoparticles. For instance, Meyabadi et al. [65] prepared nano-spherical cellulose structures
with the size less than 100 nm from waste cotton through enzymatic hydrolysis followed by
sonication treatment. Additionally, Azrina’s team [66] used the ultrasound assisted acid
hydrolysis to produce spherical cellulose nanoparticles with sizes ranging 30–40 nm from oil
palm empty fruit bunch pulp. Furthermore, Lu and Hsieh [64] have recently reported that a
great deal of the products from acid hydrolysis and freeze drying of the cotton cellulose were
10–100 nm spherical cellulose nanocrystals. Although spherical cellulose nanoparticles have
been produced by the hydrolysis of cellulose isolated from various biomasses, the exact
structure and formation mechanism are principally unknown. It is usually considered that

30
ACCEPTED MANUSCRIPT

cellulose in native plant cells is in the form of cellulose microfibrils bundles that are
imbedded between hemicellulose and lignin. At the end of chemical isolation and purification
processes, cellulose in the forms of microfibrils and other fragmented forms is separated.
During the collection and preparation processes, such as freeze-drying and deposition onto
other substrates for imaging in our case, it is possible that the nanocellulose particles were
created by a self-assembly process from spherical structure and their fragments. According to
the study published by Lu and Hsieh [67], during the drying process, more abundant nano-

PT
fragments associate around the fewer, but larger nano-rods via strong lateral hydrogen bonds
and with each other, possibly driven by a layer-by-layer and thermodynamically favored
process to the reduced specific surface.

RI
SC
It should be noted that the morphology of nanocellulose observed via TEM analysis
might not be observed under FESEM. This is mainly due to instrumental differences between
NU
these two imaging techniques, the observed dimensional difference between FESEM and
TEM points to potential influence of the different sample preparation method, i.e., FESEM
required freeze-dried powder sample whereas TEM needed nanocellulose suspension
MA

deposited on the copper grid. This is not surprising as similar observation has been reported
in the previous works [48, 64] where the nanocellulose observed in FESEM and TEM were
D

different.
E

Increasing the nanocellulose aspect ratio produced densely entangled networks


PT

stronger than the initial lignocellulosic material because of the creation of more and stronger
bonds, in a similar behavior as reported by Tibolla et al [44]. For comparison, the
CE

nanocellulose obtained in this study had higher aspect ratio than those cellulose
nanomaterials isolated from different sources in other studies, including cotton linter (5) [47],
AC

cotton fabric (9.96) [56], Eucalyptus kraft pulp (12‒14) [8], degreasing cotton (12), waste
cotton cloth (17) [32], Helicteres isora (15) [57] and mengkuang leaves (10‒20) [49].
Besides that, the aspect ratio measured is in close agreement with the values obtained for the
nanocellulose obtained from bacterial cellulose (20.1) [43] and coconut husk (22) [68]. The
aspect ratio of nanocellulose can be fundamental to their applications especially for
reinforcing fillers for polymer composites. According to the percolation theory [59], it is
possible to claim that longer nanocellulose fibers are able to provide better mechanical
reinforcement to the polymeric composite than those shorter ones at the same volume fraction
of particles. It is widely accepted that the raw material and the hydrolysis/pretreatment

31
ACCEPTED MANUSCRIPT

conditions are the most important factors that influenced the nanocellulose dimensions. The
aspect ratio of pear peel nanocellulose (24.6) does not overlap with any of those listed, and
this raw material is an option if nanocellulose with specific dimensions is required by
industry.

In general, nanocellulose with the aspect ratio greater than 10 promotes the formation
of an anisotropic phase within the polymer matrix and exhibits good mechanical properties
(tensile strength, bending strength and Young’s modulus) because a high aspect ratio may

PT
favor tension transfer in the nanofiber-matrix interface [59]. Moriana’s group [59] reported
that the aspect ratio value of nanocellulose above 10 was considered the minimum value for a

RI
good stress transfer from the matrix to the fibers during reinforcement. When the

SC
nanocellulose-based nanocomposites are subjected to external stresses, the energy is absorbed
and dissipated by the nanocellulose, thus improving the mechanical properties of the polymer
NU
matrix [34]. The nanocellulose can function as a barrier to oxygen during thermal degradation
and enhanced the thermal stability of the polymer. Moreover, the thermal stability of polymer
could be improved due to the good dispersion and interfacial interaction between polymer
MA

matrix and nanocellulose that can promote a tortuous path to the molecules of the oxidizing
gas that permeates through the nanocomposites materials. Herein, the high aspect ratio seen
D

here for isolated nanocellulose indicates the promise of this nanomaterial as a dispersed phase
for polymer nanocomposites materials with improved properties. Evaluating FTIR, chemical
E

composition, XRD, TGA and morphological results, it is possible to conclude that one-pot
PT

process is effective in removing non-cellulosic layer and cellulose amorphous content of pear
peel residue without damaging the cellulose fiber structure.
CE

------Figure 6------
AC

4. Conclusions
In summary, the one-pot process for isolating the nanocellulose from pear (Pyrus
pyrifolia) peel residue was designed by applying H2O2 as an oxidant and Cr(NO3)3 as metal
salt catalyst under acidic hydrolyzing medium. This simple and straightforward process
provided an opportunity for the effective revalorization of pear peel residue into value-added
nanocellulose that would otherwise be disposed of as a waste. The use of H2O2 has great
benefits over other chemicals due to its easily removing and rapidly decomposed into water
and oxygen after the reaction. The results of the physicochemical characterizations showed

32
ACCEPTED MANUSCRIPT

that the isolated nanocellulose had the high crystallinity of 85.7% with the average diameter
and length of 20.5 ± 6.3 nm and 503.8 ± 30.5 nm, respectively, giving rise to a high aspect
ratio of 24.6. The FTIR and XRD results showed that the original structure of cellulose fibrils
is preserved to a large extent during the isolation process. TEM analysis shows that the
isolated nanocellulose contained networked structures as well as spherical shaped particles.
Most importantly, the final yield of isolated nanocellulose was ca. 23.7% with respect to the
raw material. In this sense, the present work has proven that the one-pot isolation process is

PT
able to re-engineer the structure of pear peel residue into tailored nanocellulose with desired
characteristics for various applications. It can be concluded that the versatile one-pot process
is a better and more practical method for obtaining large-scale production of high-quality

RI
nanocellulose with less washing phases and purification steps involved from various fruit

SC
waste origins.
NU
Acknowledgement
The authors are grateful for the financial support by University of Malaya: SATU Joint
Research Scheme (ST015-2017) and RU Grant (RU007C-2017G).
MA

References
D

[1] K.H. Lee, J.-Y. Cho, H.J. Lee, Y.-K. Ma, J. Kwon, S.H. Park, S.-H. Lee, J.A. Cho, W.-S. Kim,
E

K.-H. Park, Hydroxycinnamoylmalic acids and their methyl esters from pear (Pyrus pyrifolia
PT

Nakai) fruit peel, Journal of agricultural and food chemistry 59(18) (2011) 10124-10128.

[2] H.N. Rabetafika, B. Bchir, C. Blecker, M. Paquot, B. Wathelet, Comparative study of alkaline
CE

extraction process of hemicelluloses from pear pomace, Biomass and Bioenergy 61 (2014) 254-
AC

264.

[3] N.A. Sagar, S. Pareek, S. Sharma, E.M. Yahia, M.G. Lobo, Fruit and Vegetable Waste:

Bioactive Compounds, Their Extraction, and Possible Utilization, Comprehensive Reviews in

Food Science and Food Safety 17(3) (2018) 512-531.

[4] M. Kunaver, A. Anžlovar, E. Žagar, The fast and effective isolation of nanocellulose from

selected cellulosic feedstocks, Carbohydrate Polymers 148 (2016) 251-258.

33
ACCEPTED MANUSCRIPT

[5] L. Brinchi, F. Cotana, E. Fortunati, J. Kenny, Production of nanocrystalline cellulose from

lignocellulosic biomass: technology and applications, Carbohydrate Polymers 94(1) (2013) 154-

169.

[6] Y.W. Chen, S.H. Binti Hassan, M. Yahya, H.V. Lee, Novel Superabsorbent Cellulose-Based

Hydrogels: Present Status, Synthesis, Characterization, and Application Prospects, Cellulose-

Based Superabsorbent Hydrogels (2018) 1-41.

PT
[7] Y.W. Chen, H.V. Lee, J.C. Juan, S.-M. Phang, Production of new cellulose nanomaterial from

RI
red algae marine biomass Gelidium elegans, Carbohydrate Polymers 151 (2016) 1210-1219.

SC
[8] G. Tonoli, E. Teixeira, A. Corrêa, J. Marconcini, L. Caixeta, M. Pereira-da-Silva, L. Mattoso,

Cellulose micro/nanofibres from Eucalyptus kraft pulp: preparation and properties, Carbohydrate
NU
polymers 89(1) (2012) 80-88.

[9] K.Y. Goh, Y.C. Ching, C.H. Chuah, L.C. Abdullah, N.-S. Liou, Individualization of
MA

microfibrillated celluloses from oil palm empty fruit bunch: comparative studies between acid

hydrolysis and ammonium persulfate oxidation, Cellulose 23(1) (2016) 379-390.


D

[10] M. Yahya, Y.W. Chen, H.V. Lee, W.H.W. Hassan, Reuse of Selected Lignocellulosic and
E

Processed Biomasses as Sustainable Sources for the Fabrication of Nanocellulose via Ni(II)-
PT

Catalyzed Hydrolysis Approach: A Comparative Study, Journal of Polymers and the Environment
CE

26(7) (2018) 2825-2844.

[11] Y. Habibi, L.A. Lucia, O.J. Rojas, Cellulose nanocrystals: chemistry, selfassembly, and
AC

applications, Chem Rev 110 (2010).

[12] D. Trache, M.H. Hussin, C.T. Hui Chuin, S. Sabar, M.R.N. Fazita, O.F.A. Taiwo, T.M.

Hassan, M.K.M. Haafiz, Microcrystalline cellulose: Isolation, characterization and bio-composites

application—A review, International Journal of Biological Macromolecules 93 (2016) 789-804.

[13] D. Trache, M.H. Hussin, M.M. Haafiz, V.K. Thakur, Recent progress in cellulose

nanocrystals: sources and production, Nanoscale 9(5) (2017) 1763-1786.

34
ACCEPTED MANUSCRIPT

[14] Y.W. Chen, H.V. Lee, S.B. Abd Hamid, Preparation and Characterization of Cellulose

Crystallites via Fe(III)-, Co(II)-and Ni(II)-assisted Dilute Sulfuric Acid Catalyzed Hydrolysis

Process, Journal of Nano Research 41 (2016) 96-109.

[15] Y.W. Chen, H.V. Lee, S.B.A. Hamid, Preparation of Nanostructured Cellulose via Cr(III)-

and Mn(II)-Transition Metal Salt Catalyzed Acid Hydrolysis Approach, BioResources 11(3) (2016)

7224-7241.

PT
[16] S.R. Kamireddy, J. Li, M. Tucker, J. Degenstein, Y. Ji, Effects and mechanism of metal

RI
chloride salts on pretreatment and enzymatic digestibility of corn stover, Industrial & Engineering

SC
Chemistry Research 52(5) (2013) 1775-1782.

[17] J. Li, H. Xiu, M. Zhang, H. Wang, Y. Ren, Y. Ji, Enhancement of cellulose acid hydrolysis
NU
selectivity using metal ion catalysts, Current Organic Chemistry 17(15) (2013) 1617-1623.

[18] J. Li, X. Zhang, M. Zhang, H. Xiu, H. He, Optimization of selective acid hydrolysis of
MA

cellulose for microcrystalline cellulose using FeCl3, BioResources 9(1) (2014) 1334-1345.

[19] S. Jing, X. Cao, L. Zhong, X. Peng, X. Zhang, S. Wang, R. Sun, In Situ Carbonic Acid from
D

CO2: A Green Acid for Highly Effective Conversion of Cellulose in the Presence of Lewis acid,
E

ACS Sustainable Chemistry & Engineering 4(8) (2016) 4146-4155.


PT

[20] W. Chen You, V. Lee Hwei, Recent progress in homogeneous Lewis acid catalysts for the
CE

transformation of hemicellulose and cellulose into valuable chemicals, fuels, and nanocellulose,

Reviews in Chemical Engineering, 2018.


AC

[21] L. Liu, J. Sun, C. Cai, S. Wang, H. Pei, J. Zhang, Corn stover pretreatment by inorganic salts

and its effects on hemicellulose and cellulose degradation, Bioresource Technology 100(23) (2009)

5865-5871.

[22] X. Jia, F. Zhang, Y. Xin, W. Xu, N. Shi, Z. Lv, A Simulation Investigation on the Free

Radical Decomposition Mechanism of Hydrogen Peroxide, Chemical Engineering Transactions 61

(2017) 1699-1704.

35
ACCEPTED MANUSCRIPT

[23] A.C. Leung, S. Hrapovic, E. Lam, Y. Liu, K.B. Male, K.A. Mahmoud, J.H. Luong,

Characteristics and Properties of Carboxylated Cellulose Nanocrystals Prepared from a Novel

One‐Step Procedure, Small 7(3) (2011) 302-305.

[24] H. Shi, J. Li, D. Shi, H. Shi, B. Feng, W. Li, Y. Bai, J. Zhao, A. He, Combined

Reduction/Precipitation, Chemical Oxidation, and Biological Aerated Filter Processes for

Treatment of Electroplating Wastewater, Separation Science and Technology 50(15) (2015) 2303-

PT
2310.

RI
[25] S. Ovalle-Serrano, F. Gómez, C. Blanco-Tirado, M. Combariza, Isolation and

SC
characterization of cellulose nanofibrils from Colombian Fique decortication by-products,

Carbohydrate polymers 189 (2018) 169-177.


NU
[26] Y.C. Ching, T.S. Ng, Effect of preparation conditions on cellulose from oil palm empty fruit

bunch fiber, BioResources 9(4) (2014) 6373-6385.


MA

[27] Ú. Fillat, B. Wicklein, R. Martin-Sampedro, D. Ibarra, E. Ruiz-Hitzky, C. Valencia, A.

Sarrión, E. Castro, M.E. Eugenio, Assessing cellulose nanofiber production from olive tree
D

pruning residue, Carbohydrate polymers 179 (2018) 252-261.


E

[28] G.-F. Deng, C. Shen, X.-R. Xu, R.-D. Kuang, Y.-J. Guo, L.-S. Zeng, L.-L. Gao, X. Lin, J.-F.
PT

Xie, E.-Q. Xia, Potential of fruit wastes as natural resources of bioactive compounds, International
CE

journal of molecular sciences 13(7) (2012) 8308-8323.

[29] T. Dias, R. Fragoso, E. Duarte, Anaerobic co-digestion of dairy cattle manure and pear waste,
AC

Bioresource technology 164 (2014) 420-423.

[30] M. Mohamed, W. Salleh, J. Jaafar, S. Asri, A. Ismail, Physicochemical properties of “green”

nanocrystalline cellulose isolated from recycled newspaper, RSC Advances 5(38) (2015) 29842-

29849.

[31] M.F. Rosa, E.S. Medeiros, J.A. Malmonge, K.S. Gregorski, D.F. Wood, L.H.C. Mattoso, G.

Glenn, W.J. Orts, S.H. Imam, Cellulose nanowhiskers from coconut husk fibers: Effect of

36
ACCEPTED MANUSCRIPT

preparation conditions on their thermal and morphological behavior, Carbohydrate Polymers 81(1)

(2010) 83-92.

[32] Z. Wang, Z. Yao, J. Zhou, Y. Zhang, Reuse of waste cotton cloth for the extraction of

cellulose nanocrystals, Carbohydrate Polymers 157 (2017) 945-952.

[33] R. Xiong, X. Zhang, D. Tian, Z. Zhou, C. Lu, Comparing microcrystalline with spherical

nanocrystalline cellulose from waste cotton fabrics, Cellulose 19(4) (2012) 1189-1198.

PT
[34] F. Ferreira, M. Mariano, S. Rabelo, R. Gouveia, L. Lona, Isolation and surface modification

RI
of cellulose nanocrystals from sugarcane bagasse waste: from a micro-to a nano-scale view,

SC
Applied Surface Science 436 (2018) 1113-1122.

[35] S.F. Sim, M. Mohamed, N.A.L.M.I. Lu, N.S.P. Sarman, S.N.S. Samsudin, Computer-assisted
NU
analysis of fourier transform infrared (FTIR) spectra for characterization of various treated and

untreated agriculture biomass, BioResources 7(4) (2012) 5367-5380.


MA

[36] T.-C. Hsu, G.-L. Guo, W.-H. Chen, W.-S. Hwang, Effect of dilute acid pretreatment of rice

straw on structural properties and enzymatic hydrolysis, Bioresource Technology 101(13) (2010)
D

4907-4913.
E

[37] R. Dhabhai, S.P. Chaurasia, A.K. Dalai, Effect of pretreatment conditions on structural
PT

characteristics of wheat straw, Chemical Engineering Communications 200(9) (2013) 1251-1259.


CE

[38] M. Poletto, H. Ornaghi , A. Zattera, Native Cellulose: Structure, Characterization and

Thermal Properties, Materials 7(9) (2014) 6105.


AC

[39] A. Mandal, D. Chakrabarty, Isolation of nanocellulose from waste sugarcane bagasse (SCB)

and its characterization, Carbohydrate Polymers 86(3) (2011) 1291-1299.

[40] M.K. Mohamad Haafiz, S.J. Eichhorn, A. Hassan, M. Jawaid, Isolation and characterization

of microcrystalline cellulose from oil palm biomass residue, Carbohydrate Polymers 93(2) (2013)

628-34.

37
ACCEPTED MANUSCRIPT

[41] N. Kawee, N.T. Lam, P. Sukyai, Homogenous isolation of individualized bacterial

nanofibrillated cellulose by high pressure homogenization, Carbohydrate polymers 179 (2018)

394-401.

[42] F. Prieto-García, E. Jiménez-Muñoz, O.A. Acevedo-Sandoval, R. Rodríguez-Laguna, R.A.

Canales-Flores, J. Prieto-Méndez, Obtaining and Optimization of Cellulose Pulp from Leaves of

Agave tequilana Weber Var. Blue. Preparation of Handmade Craft Paper, Waste and Biomass

PT
Valorization (2018) 1-17.

RI
[43] N.F. Vasconcelos, J.P.A. Feitosa, F.M.P. da Gama, J.P.S. Morais, F.K. Andrade, M.d.S.M. de

SC
Souza, M. de Freitas Rosa, Bacterial cellulose nanocrystals produced under different hydrolysis

conditions: Properties and morphological features, Carbohydrate Polymers 155 (2017) 425-431.
NU
[44] H. Tibolla, F.M. Pelissari, J.T. Martins, A. Vicente, F.C. Menegalli, Cellulose nanofibers

produced from banana peel by chemical and mechanical treatments: Characterization and
MA

cytotoxicity assessment, Food Hydrocolloids 75 (2018) 192-201.

[45] K. Rahbar Shamskar, H. Heidari, A. Rashidi, Preparation and evaluation of nanocrystalline


D

cellulose aerogels from raw cotton and cotton stalk, Industrial Crops and Products 93 (2016) 203-
E

211.
PT

[46] W. Chen, H. Yu, Y. Liu, Y. Hai, M. Zhang, P. Chen, Isolation and characterization of
CE

cellulose nanofibers from four plant cellulose fibers using a chemical-ultrasonic process, Cellulose

18(2) (2011) 433-442.


AC

[47] X.Y. Tan, S.B. Abd Hamid, C.W. Lai, Preparation of high crystallinity cellulose nanocrystals

(CNCs) by ionic liquid solvolysis, Biomass and Bioenergy 81 (2015) 584-591.

[48] S. Naduparambath, T. Jinitha, V. Shaniba, M. Sreejith, A.K. Balan, E. Purushothaman,

Isolation and characterisation of cellulose nanocrystals from sago seed shells, Carbohydrate

polymers 180 (2018) 13-20.

38
ACCEPTED MANUSCRIPT

[49] R.M. Sheltami, I. Abdullah, I. Ahmad, A. Dufresne, H. Kargarzadeh, Extraction of cellulose

nanocrystals from mengkuang leaves (Pandanus tectorius), Carbohydrate Polymers 88(2) (2012)

772-779.

[50] X. Feng, X. Meng, J. Zhao, M. Miao, L. Shi, S. Zhang, J. Fang, Extraction and preparation of

cellulose nanocrystals from dealginate kelp residue: structures and morphological characterization,

Cellulose 22(3) (2015) 1763-1772.

PT
[51] M. Ishak, S. Sapuan, Z. Leman, M. Rahman, U. Anwar, Characterization of sugar palm

RI
(Arenga pinnata) fibres: tensile and thermal properties, Journal of thermal analysis and calorimetry

SC
109(2) (2011) 981-989.

[52] H. Yang, R. Yan, H. Chen, D.H. Lee, C. Zheng, Characteristics of hemicellulose, cellulose
NU
and lignin pyrolysis, Fuel 86(12–13) (2007) 1781-1788.

[53] T.J. Bondancia, L.H.C. Mattoso, J.M. Marconcini, C.S. Farinas, A new approach to obtain
MA

cellulose nanocrystals and ethanol from eucalyptus cellulose pulp via the biochemical pathway,

Biotechnology progress 33(4) (2017) 1085-1095.


D

[54] F. Beltramino, M.B. Roncero, A.L. Torres, T. Vidal, C. Valls, Optimization of sulfuric acid
E

hydrolysis conditions for preparation of nanocrystalline cellulose from enzymatically pretreated


PT

fibers, Cellulose 23(3) (2016) 1777-1789.


CE

[55] K.C.C. de Carvalho Benini, H.J.C. Voorwald, M.O.H. Cioffi, M.C. Rezende, V. Arantes,

Preparation of nanocellulose from Imperata brasiliensis grass using Taguchi method,


AC

Carbohydrate polymers 192 (2018) 337-346.

[56] J. Han, C. Zhou, A.D. French, G. Han, Q. Wu, Characterization of cellulose II nanoparticles

regenerated from 1-butyl-3-methylimidazolium chloride, Carbohydrate polymers 94(2) (2013)

773-781.

[57] C.J. Chirayil, J. Joy, L. Mathew, M. Mozetic, J. Koetz, S. Thomas, Isolation and

characterization of cellulose nanofibrils from Helicteres isora plant, Industrial Crops and Products

59 (2014) 27-34.
39
ACCEPTED MANUSCRIPT

[58] J. Xie, C.-Y. Hse, F. Cornelis, T. Hu, J. Qi, T.F. Shupe, Isolation and characterization of

cellulose nanofibers from bamboo using microwave liquefaction combined with chemical

treatment and ultrasonication, Carbohydrate polymers 151 (2016) 725-734.

[59] R. Moriana, F. Vilaplana, M. Ek, Cellulose Nanocrystals from Forest Residues as

Reinforcing Agents for Composites: A Study from Macro- to Nano-Dimensions, Carbohydrate

Polymers 139 (2016) 139-149.

PT
[60] N. Johar, I. Ahmad, A. Dufresne, Extraction, preparation and characterization of cellulose

RI
fibres and nanocrystals from rice husk, Industrial Crops and Products 37(1) (2012) 93-99.

SC
[61] C. Liu, B. Li, H. Du, D. Lv, Y. Zhang, G. Yu, X. Mu, H. Peng, Properties of nanocellulose

isolated from corncob residue using sulfuric acid, formic acid, oxidative and mechanical methods,
NU
Carbohydrate Polymers 151 (2016) 716–724.

[62] H. Kargarzadeh, I. Ahmad, I. Abdullah, A. Dufresne, S.Y. Zainudin, R.M. Sheltami, Effects
MA

of hydrolysis conditions on the morphology, crystallinity, and thermal stability of cellulose

nanocrystals extracted from kenaf bast fibers, Cellulose 19(3) (2012) 855-866.
D

[63] J. Han, C. Zhou, Y. Wu, F. Liu, Q. Wu, Self-assembling behavior of cellulose nanoparticles
E

during freeze-drying: Effect of suspension concentration, particle size, crystal structure, and
PT

surface charge, Biomacromolecules 14(5) (2013) 1529-1540.


CE

[64] P. Lu, Y.-L. Hsieh, Preparation and properties of cellulose nanocrystals: rods, spheres, and

network, Carbohydrate Polymers 82(2) (2010) 329-336.


AC

[65] T.F. Meyabadi, F. Dadashian, G.M.M. Sadeghi, H.E.Z. Asl, Spherical cellulose nanoparticles

preparation from waste cotton using a green method, Powder Technology 261 (2014) 232-240.

[66] Z.Z. Azrina, M.D.H. Beg, M. Rosli, R. Ramli, N. Junadi, A.M. Alam, Spherical

nanocrystalline cellulose (NCC) from oil palm empty fruit bunch pulp via ultrasound assisted

hydrolysis, Carbohydrate polymers 162 (2017) 115-120.

[67] P. Lu, Y.-L. Hsieh, Cellulose isolation and core–shell nanostructures of cellulose nanocrystals

from chardonnay grape skins, Carbohydrate Polymers 87(4) (2012) 2546-2553.


40
ACCEPTED MANUSCRIPT

[68] D.M. Nascimento, J.S. Almeida, A.F. Dias, M.C.B. Figueirêdo, J.P.S. Morais, J.P.A. Feitosa,

M.d.F. Rosa, A novel green approach for the preparation of cellulose nanowhiskers from white

coir, Carbohydrate Polymers 110 (2014) 456-463.

PT
RI
SC
NU
MA
E D
PT
CE
AC

41
ACCEPTED MANUSCRIPT

Figure 1: Schematic representation of the overall steps of nanocellulose isolation from pear
peel residue and physical appearance of each obtained products by digital macroscopic
images.
Figure 2: FTIR spectra of (a) raw material and (b) isolated nanocellulose.
Figure 3: XRD pattern of (a) raw material and (b) isolated nanocellulose.
Figure 4: TG and DTG curves of (a) raw material and (b) isolated nanocellulose.

PT
Figure 5: FESEM micrographs of (a) raw material and (b) isolated nanocellulose. Scale bar
is 10 μm; (c) EDX diffraction spectrum for elemental analysis of nanocellulose.
Figure 6: TEM microscopic images of isolated nanocellulose in the form of (a‒b) porous

RI
network structure and (c‒d) spherical shaped.

SC
NU
MA
E D
PT
CE
AC

42
ACCEPTED MANUSCRIPT

Table 1: Comparison of the chemical composition of various raw materials reported in works
of literature.
Composition (%)
Raw material Reference
α-cellulose Hemicellulose Lignin Others
Corncob 34.1 42.5 12.8 10.6 [29]
Peanut shell 22.1 12.1 35.2 30.6
Rice straw 33 26 7 34

PT
Sugarcane 40 29 13 18
Sorghum leaf 31 30 11 28

RI
Sago seed shells 36.5 22.5 23.6 17.4
Wheat straw stem 39.8 34.2 19.8 ‒ [30]

SC
Empty fruit bunch 40 23 21 2
Sugar beet pulp 22 32 2 ‒
NU
Mengkuang leaves 37.3 34.4 24 2.5
Mulberry bark 37.4 25.3 10 ‒
MA

Coconut husk 32.5 ‒ 37 ‒


Rice husk 35 33 23 23
Apple 43.6 24.4 20.4 11.7 [31]
D

Cherry 18.4 10.7 69.4 1.51


E

Chokeberry 34.6 33.5 24.1 7.85


PT

Blackcurrant 12 25.3 59.3 2.73


Pear peel 38.5 23.6 28.1 9.8* Current study
CE

*Extractive content determined by ASTM D1107-96


AC

43
ACCEPTED MANUSCRIPT

Table 2: Assessment of the preparation of nanocellulose from different origins and isolation
routes
Origins Synthesis process Yield (%)* Reference
Pear peel residue One-pot treatment 23.7 This study
Woody chips Alkalization, bleaching and H2SO4 hydrolysis 20 [33]
Banana pseudostem Alkalization and H2SO4 hydrolysis 19.4 [34]
Woody branches Alkalization, bleaching and H2SO4 hydrolysis 16 [33]

PT
Pine needles Alkalization, bleaching and H2SO4 hydrolysis 13
Barley straw and husk Alkalization, bleaching and H2SO4 hydrolysis 12 [35]

RI
Giant cane (Arundo donax) Alkalization, bleaching and H2SO4 hydrolysis 11 [36]
Spruce bark Solvent extraction, bleaching, H2SO4 hydrolysis 11 [37]

SC
Rice straw Alkalization, bleaching and H2SO4 hydrolysis 6.6 [38]
Olive pomace Soxhlet extraction, alkalization, bleaching and 5.8 [39]
NU
H2SO4 hydrolysis
Licuri palm leaves Alkalization, bleaching and H2SO4 hydrolysis 5.7 [40]
MA

Garlic stalk Alkalization, bleaching and H2SO4 hydrolysis 4.6 [41]


* With respect to the initial amount of dried raw material.
E D
PT
CE
AC

44
ACCEPTED MANUSCRIPT

Table 3: Thermal properties of nanocellulose extracted by various treatments


Tmax value (°C)
Starting materials Hydrolysis method References
Before After
Pear peel residue 363 358.5 One-pot process Present study
Bamboo cellulose 355 324 Phosphoric acid + ball milling [68]
Cotton linter 350 240 Ionic liquid solvolysis [58]
Gluconacetobacter xylinus 313.6 298.2 High pressure homogenization [52]

PT
Colombian Fique 322 310 TEMPO-mediated oxidation [24]

RI
SC
NU
MA
E D
PT
CE
AC

45
ACCEPTED MANUSCRIPT

Highlights

 Nanocellulose has been successfully isolated from Asian pear (Pyrus pyrifolia) peel.
 A novel one-pot system was developed for the facile production of nanocellulose.
 Isolated nanocellulose rendered the high crystallinity (85.7%) with high aspect ratio (24.6).
 The nanocellulose yield was ca. 23.7% with respect to the raw material.

PT
RI
SC
NU
MA
E D
PT
CE
AC

46
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6

You might also like