ArticlePDF AvailableLiterature Review

Abstract and Figures

The high demand for plastic and polymeric materials which keeps rising every year makes them important industries, for which sustainability is a crucial aspect to be taken into account. Therefore, it becomes a requirement to makes it a clean and eco-friendly industry. Cellulose creates an excellent opportunity to minimize the effect of non-degradable materials by using it as a filler for either a synthesis matrix or a natural starch matrix. It is the primary substance in the walls of plant cells, helping plants to remain stiff and upright, and can be found in plant sources, agriculture waste, animals, and bacterial pellicle. In this review, we discussed the recent research development and studies in the field of biocomposites that focused on the techniques of extracting micro-and nanocellulose, treatment and modification of cellulose, classification, and applications of cellulose. In addition, this review paper looked inward on how the reinforcement of micro-and nanocellulose can yield a material with improved performance. This article featured the performances, limitations, and possible areas of improvement to fit into the broader range of engineering applications.
Content may be subject to copyright.
polymers
Review
Micro- and Nanocellulose in Polymer Composite Materials:
A Review
Abdoulhdi A. Borhana Omran 1, 2, * , Abdulrahman A. B. A. Mohammed 1, S. M. Sapuan 3, 4, *, R. A. Ilyas 5, 6, * ,
M. R. M. Asyraf 7, Seyed Saeid Rahimian Koloor 8and Michal Petr ˚u 8


Citation: Omran, A.A.B.;
Mohammed, A.A.B.A.; Sapuan, S.M.;
Ilyas, R.A.; Asyraf, M.R.M.; Rahimian
Koloor, S.S.; Petr˚u, M. Micro- and
Nanocellulose in Polymer Composite
Materials: A Review. Polymers 2021,
13, 231. https://doi.org/10.3390/
polym13020231
Received: 21 December 2020
Accepted: 6 January 2021
Published: 11 January 2021
Publisher’s Note: MDPI stays neu-
tral with regard to jurisdictional clai-
ms in published maps and institutio-
nal affiliations.
Copyright: © 2021 by the authors. Li-
censee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and con-
ditions of the Creative Commons At-
tribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
1Department of Mechanical Engineering, College of Engineering, Universiti Tenaga Nasional,
Jalan Ikram-Uniten, Kajang 43000, Selangor, Malaysia; rahman.aziz@uniten.edu.my
2Department of Mechanical Engineering, College of Engineering Science & Technology, Sebha University,
Sabha 00218, Libya
3Laboratory of Biocomposite Technology, Institute of Tropical Forestry and Forest Products (INTROP),
Universiti Putra Malaysia, Serdang 43400, Selangor, Malaysia
4
Advanced Engineering Materials and Composites Research Centre (AEMC), Department of Mechanical and
Manufacturing Engineering, Universiti Putra Malaysia, Serdang 43400, Selangor, Malaysia
5Sustainable Waste Management Research Group (SWAM), School of Chemical and Energy Engineering,
Faculty of Engineering, Universiti Teknologi Malaysia, Johor Bahru 81310, Johor, Malaysia
6
Centre for Advanced Composite Materials, Universiti Teknologi Malaysia, Johor Bahru 81310, Johor, Malaysia
7Department of Aerospace Engineering, Universiti Putra Malaysia, Serdang 43400, Selangor, Malaysia;
asyrafriz96@gmail.com
8Institute for Nanomaterials, Advanced Technologies and Innovation, Technical University of Liberec,
Studentská2, 461 17 Liberec, Czech Republic; seyed.rahimian@tul.cz (S.S.R.K.); michal.petru@tul.cz (M.P.)
*Correspondence: amhmad@uniten.edu.my (A.A.B.O.); sapuan@upm.edu.my (S.M.S.);
ahmadilyas@utm.my (R.A.I.)
Abstract:
The high demand for plastic and polymeric materials which keeps rising every year makes
them important industries, for which sustainability is a crucial aspect to be taken into account.
Therefore, it becomes a requirement to makes it a clean and eco-friendly industry. Cellulose creates
an excellent opportunity to minimize the effect of non-degradable materials by using it as a filler
for either a synthesis matrix or a natural starch matrix. It is the primary substance in the walls of
plant cells, helping plants to remain stiff and upright, and can be found in plant sources, agriculture
waste, animals, and bacterial pellicle. In this review, we discussed the recent research development
and studies in the field of biocomposites that focused on the techniques of extracting micro- and
nanocellulose, treatment and modification of cellulose, classification, and applications of cellulose.
In addition, this review paper looked inward on how the reinforcement of micro- and nanocellulose
can yield a material with improved performance. This article featured the performances, limitations,
and possible areas of improvement to fit into the broader range of engineering applications.
Keywords:
natural fiber; nanocellulose; microcellulose; biocomposite; nanocomposite; biopolymer;
synthetic polymer
1. Introduction
Petroleum-based synthesis polymers are non-degradable, with production, recycling,
and disposal releasing toxic emissions into the environment [
1
]. Cellulose offers excel-
lent properties to minimize this damage by utilization as a filler in the manufacturing of
either a synthesis matrix or a natural starch matrix. Cellulose is the main substance of
a plant’s cell walls, helping plants to remain stiff and upright, hence, it can be extracted
from plant sources, agriculture waste, animals, and bacterial pellicle [
2
,
3
]. It is composed
of polymer chains consisting of unbranched
β
(1,4) linked D glucopyranosyl units (an-
hydroglucose unit, AGU) [
4
,
5
]. Cellulose also possesses excellent mechanical properties,
such as tensile and flexural strengths, tensile and flexural moduli, and thermal resistance,
as well as low cost, due to its availability from different resources and abundance in nature,
Polymers 2021,13, 231. https://doi.org/10.3390/polym13020231 https://www.mdpi.com/journal/polymers
Polymers 2021,13, 231 2 of 35
and degradability which is not obtainable in synthetic fillers, that makes it an excellent
bio-filler for both synthesis or natural polymer matrixes [
6
]. Cellulose needs to be extracted
to be a useful substance. Cellulose extraction can be achieved via three approaches; me-
chanical, chemical, and bacterial techniques. Mechanical cellulose extraction comprises
of high-pressurized homogenization [
7
], grinding [
8
], crushing [
9
], and steam explosion
methods [
10
]. Chemical extraction methods include alkali treatment [
11
], acid retting,
chemical retting [12], and degumming [13].
Cellulose can be extracted in different sizes, depending on the intended application.
Micro- and nanocellulose are the common sizes of cellulose used in industrial applica-
tions. Nanocellulose is divided into three types, (1) nanofibrillated cellulose (NFC), also
known as nanofibrils or microfibrils or macrofibrillated cellulose or nanofibrillated cellu-
lose; (2) nanocrystalline cellulose (NCC), also known as crystallites, whiskers, or rod-like
cellulose microcrystals, and (3) bacterial nanocellulose (BNC), also known as microbial
cellulose or biocellulose [
14
,
15
]. The difference between microfibrillated cellulose and
nanocrystalline cellulose is the fiber size distributions that are wide in microfibrillated
cellulose and narrow or drastically shorter in nanocrystalline cellulose [
16
]. Figure 1
depicts the structural difference between nanofibrillated cellulose and nanocrystalline
cellulose. Similar to microfibrillated cellulose, bacterial cellulose also has a narrow size
distribution and high crystallinity, except for its source, which is bacteria. According to
Alain Dufresne [
17
] and Chirayil et al. [
18
], NCC and NFC are renowned not only for their
biodegradation, superb properties, unique structures, low density, excellent mechanical
performance, high surface area and aspect ratio, biocompatibility, and natural abundance,
but also for their possibility to modify their surfaces to enhance their nano-reinforcement
compatibility with other polymers due to the presence of abundant hydroxyl groups.
Nanocellulose-based materials, also known as a new ageless bionanomaterial, are non-
toxic, recyclable, sustainable, and carbon-neutral [
17
]. NCC and NFC have demonstrated
numerous advanced applications, including in the automotive industry, optically transpar-
ent materials, drug supply, coating films, tissue technology, biomimetic materials, aerogels,
sensors, three-dimensional (3D) printing, rheology modifiers, energy harvesters, filtration,
textiles, printed and flexible electronics, composites, paper and board, packaging, oil and
gas, medical and healthcare, and scaffolding [
19
,
20
]. In addition, macro and mesoporous
nanocellulose beads also are utilized in energy storage devices. The cellulose beads act as
electrodes that serve as complements to conventional supercapacitors and batteries [
21
],
and depend on the properties of the cellulose (e.g., origin, porosity, pore distribution,
pore-size distribution, and crystallinity) [
22
]. In consequence, the number of patents and
publications on nanocellulose over 20 years have increased significantly from 764 in 2000 to
18,418 in 2020. In addition, this increment of more than 2300% over 20 years indicates that
nanocellulose has become the advanced emerging material in the 21st century.
Figure 1.
Atomic force microscopy images show different structure between nanocrystalline cellulose (NCC) [
23
] and
nanofibrillated cellulose (NFC) [24]. (Reproduced with copyright permission from Ilyas et al. [23,24]).
Polymers 2021,13, 231 3 of 35
Applications of cellulose are vast and interfere with many fields concentrated on me-
chanical, medical, and industrial applications [
25
]. In industry, cellulose is used as a filler for
matrixes in the manufacturing of a degradable polymer. Cellulose is also used in packaging
applications, tissue engineering applications, electronic, optical, sensor, pharmaceutical ap-
plications, cosmetic applications, insulation, water filtration, hygienic applications, as well
as vascular graft applications [
26
28
]. For instance, in Li-ion battery application, cellulose
has been applied along with carbon nanotubes (CNT) as current collectors [
29
]. Previously,
the current collector in the battery used the conventional aluminum foil. From this point of
view, cellulose paper-CNTs-based electrodes showed ~17% improvement in areal capacity
compared to commercial aluminum-based electrodes. Another renowned application of
cellulose is the implementation of electrospun cellulose acetate nanofibers for antimicrobial
activity as mentioned by Kalwar and Shen [
30
]. Moreover, cellulose is highly efficient in an-
titumor drug delivery [
31
]. In this case, the application of carboxymethyl cellulose-grafted
graphene oxide drug delivery system has a huge potential in colon cancer therapy. The
cellulose can also be implemented in the oil and gas industry due to its large surface areas
and high volume concentrations along with unique mechanical, chemical, thermal, and
magnetic properties [
32
]. Cellulose can also be used as additive and reinforcement for cross
arm application in transmission towers in order to improve their mechanical properties
and electrical resistance performance [
33
,
34
]. To increase the base of potential applications,
cellulose’s properties need to be more flexible in terms of modification and improvement
to match the required properties of various applications [
35
]. In this paper, we focused
on the techniques of extracting micro- and nanocellulose, treatment and modification of
cellulose, classification, and applications of cellulose. Thus, the objective of this paper is
to demonstrate the recent state of development in the field of micro and nanocellulose,
explain the process of extracting and modifying different types of cellulose, and highlight
the properties improvement of cellulose through examples.
2. Classification of Cellulose
Cellulose can be classified into two types based on size, microcellulose and nanocel-
lulose, while nanocellulose can be classified in three types: (1) nano- or microfibrillated
cellulose (NFC)/(MFC), (2) nanocrystalline cellulose (NCC), and (3) bacterial nanocellulose
(BNC) [
36
,
37
]. The advantage of extracting or isolating cellulose is that the nanocellu-
lose can be obtained from microcellulose [
6
,
38
], producing different cellulose sizes in a
compatible procedure.
Nanocellulose can be categorized into the family in nanofibrillated cellulose (NFC),
nanocrystalline cellulose (NCC), and bacterial nanocellulose (BNC). The size of nanocel-
lulose ranges from 5 nm to 100 nm [
39
]. The difference between nanofibrillated cellulose
(NFC) and microfibrillated cellulose (MFC) is that NFC is usually produced using a chemi-
cal pretreatment followed by a high-pressurized homogenization, while MFC is commonly
yielded from chemical treatment [
40
]. The sources of NFC or MFC are wood, sugar beet,
potato tuber, hemp, and flax. The average diameter is 20–50 nm [
41
,
42
]. Meanwhile,
for nanocrystalline cellulose (NCC), the average range of NCC diameter and length are
5–70 nm and 100 nm, respectively [
43
]. NCC can be extracted from several sources like
plants (wood, cotton, hemp, flax, wheat straw, mulberry bark, ramie, avicel, and tunicin),
algae and bacteria, and animals (tunicates) [
44
]. Another type of nanocellulose that can be
produced from non-plant sources is bacterial nanocellulose (BNC). Using microorganisms
in the industry of biopolymers is vital because such microorganisms exhibit rapid growth,
allowing for high yields and year-round availability of the product [
45
]. There are two main
methods for producing BNC using microorganisms: static culture and stirred culture [
46
].
Static culture employs the accumulation of a thick, leather-like white BNC pellicle at the
air-liquid interface. The stirred culture synthesizes cellulose in a dispersed manner in the
culture medium, forming irregular pellets or suspended fibers [47].
It is better to produce bacterial cellulose by static culture because previous studies
have shown that bacterial cellulose produced from a static culture has higher mechanical
Polymers 2021,13, 231 4 of 35
strength and yields than those obtained from stirred culture. Moreover, stirred culture
has a higher probability of microorganism mutations, which might affect BNC produc-
tion. The disadvantage of a static culture is that it takes more time and a larger area of
cultivation [4851].
3. Microcellulose and Nanocellulose Extraction, Treatment, and Modification
Lately, natural fiber biopolymers have been significantly used as alternatives to syn-
thetic polymer which negatively affected the environment [
52
]. Green composites can be
enrolled in many applications, such as automobiles, packaging, construction, building ma-
terials, furniture industry, etc. [
53
58
]. Cellulose is the main component of several natural
fibers, such as sugarcane bagasse, cotton, cogon grass, flax, hemp, jute, and sisal [
59
64
],
and it can also be found in sea animals, bacteria, and fungi. Cellulose can be extracted in
microscale with an excessive amount of mineral acids, the crystalline phases at nanome-
ter range [
65
], with sizes of 10–200
µ
m [
66
], and the mean diameter of approximately
44.28 µm [67].
The structure of microcellulose can be divided into microfibrillated cel-
lulose or microcrystalline cellulose; microcrystalline cellulose has higher strength than
the microfibrillated cellulose [
4
]. Cellulose can also be extracted as nanocellulose size of
nanocellulose fiber, which generally contains less than 100 nm in diameter and several
micrometers in length [
68
]. Plant natural fiber consists of cellulose and non-cellulose mate-
rials such as lignin, hemicellulose, pectin, wax, and other extractives. Therefore, in order
to extract cellulose either as micro or nano, the non-cellulose materials must be removed.
There are two common methods to remove non-cellulosic materials that were used by
researchers, (I) acid chlorite treatment and (II) alkaline treatment [
69
,
70
]. Depending on
the conditions of extraction process and extraction technique, the crystalline region of the
cellulose can significantly vary in size and aspect ratio. This usually results in the types
of fibrils, crystalline, and particle sizes (micro- or nano-size). However, they are normally
anisometric.
3.1. Cellulose Extraction Techniques
There are several types of cellulose production techniques such as mechanical treat-
ment, chemical treatment, combination of chemi-mechanical process, as well as bacterial
production of cellulose.
3.1.1. Mechanical Extraction
High-pressurized homogenization is one of the mechanical extraction techniques.
High-pressurized homogenization is used for large-scale nanocellulose production by
forcing the material through a very narrow channel or orifice using a piston under high
pressure of 50–2000 MPa [
66
]. This is an environmentally friendly method for nanocellulose
isolation [
71
]. However, there is a possibility for the occurrence of mechanical damage to
the crystalline structure using this method [
72
]. Another mechanical technique is grinding.
Grinding is used to separate nanocellulose from fiber by applying shear stress on the fiber
by rotating grindstones at approximately 1500 rpm [
73
]. The heat produced by friction
during the fibrillation process leads to water evaporation, which improves the extraction
process [
74
]. In addition, crushing is also used to extract cellulose fiber. This method is used
to produce microcellulose in frozen places [
75
]. The size of the produced cellulose ranges
between 0.1 and 1
µ
m. This process can be used as a pretreatment prior to high-pressurized
homogenization to yield nanocellulose. Steam explosion is utilized for the extraction of
cellulose, which uses a low energy consumption method to extract the cellulose. Although
it does not completely remove lignin, it can be considered a pretreatment. After applying
this method, the obtained fiber needs mechanical modification.
3.1.2. Chemical Extraction
Chemical extraction procedures extract cellulose by using alkali retting, acid retting,
chemical retting, chemical assisted natural (CAN), or degumming to remove the lignin
Polymers 2021,13, 231 5 of 35
content in the fibers. These treatments also affect other components of the fiber microstruc-
ture, including pectin, hemicellulose, and other non-cellulosic materials [
76
79
]. One
of the examples using the chemical extraction method is alkali or acid retting. This ex-
traction method causes less fiber damage [
80
], while mechanical extraction is less costly.
It is performed by heating, cleaning, and soaking the fiber in alkali or acid solution [
81
].
This method has the ability to improve some properties of the fiber. Degumming, which is
one of the chemical extraction processes that is developed to hold the ramie fiber’s shape,
works by eliminating the gummy and pectin content [
82
]. Another chemical technique is
chemical retting. This procedure is used to reduce the lignin and water content in fibers.
Chemical retting is able to remove more lignin compared to alkali and acid retting but is
less effective in terms of eliminating moisture [
12
]. A combination of the chemical and
mechanical extraction methods can be applied to guarantee higher efficiency of lignin
removal, where the mechanical processes usually are done after chemical treatment [
83
].
Figure 2shows the extraction of nanocellulose from lignocellulosic biomass via mechanical
and chemical methods.
Figure 2.
Extraction of nanocellulose from lignocellulosic biomass (reproduced with copyright permission from
Sharma et al. [84]).
3.1.3. Bacterial Production of Cellulose
Bacterial cellulose is of similar molecular formula to plant origin cellulose, charac-
terized by a crystalline nanofibrillar structure which creates a large surface area that can
retain a large amount of liquid. There are many methods for bacterial cellulose preparation,
including static, agitated/shaking, and bioreactor cultures. The results of macroscopic
morphology, microstructure, mechanical properties of bacterial cellulose are different, de-
pending on the preparation method. The static culture method enhances the accumulation
of a gelatinous membrane of cellulose at the surface of the nutrition solution, whereas
the agitated/shaking culture affects the asterisk-like, sphere-like, pellet-like, or irregular
masses [
36
]. The required properties and the applications dictate the selection of the ap-
propriate preparation method. Producing cellulose-based bacterial resources gives higher
critical surface tension and higher thermal degradation temperature while the cellulose
extracted from plants via combination of the chemical and mechanical extraction methods
has a hierarchical organization and semi-crystalline nature.
Polymers 2021,13, 231 6 of 35
3.2. Cellulose Surface Treatment and Modification
Cellulose is the most abundant component that can be found almost exclusively in
plant cell walls; it can also be produced by some algae and bacteria [
85
]. The applications
of natural biopolymers have extended in last recent years due to the improvement in
the processes of surface treatment and modifications; these applications involve automo-
biles, construction, building materials like nano building blocks in composites, furniture
industry, and optical applications [
53
,
86
]. The cellulose fiber has two main drawbacks,
(1) high number of hydroxyl groups that makes the product’s structure gel-like and (2)
high hydrophilicity [
87
], which limit its uses in several applications. The purpose of the
modification is to improve these two drawbacks to enhance the cellulose’s properties and
broaden the applications of natural fiber [88].
To reduce energy consumption during cellulose manufacturing and to extract cellulose
in an effective way, pretreatment needs to be carried out. The pretreatment can be either
enzymatic pretreatment or TEMPO (2,2,6,6-tetramethylpiperidine-1-oxyl) pretreatment.
Enzymatic pretreatment can be divided into cellobiohydrolases and endoglucanases [
89
,
90
],
which show strong synergistic effects [
91
]. TEMPO-mediated oxidation pretreatment is a
treatment that must be performed in solution. TEMPO-mediated oxidation pretreatment
improves the reactivity of cellulose, and the C6 primary hydroxyl groups of cellulose are
converted to carboxylate groups via the C6 aldehyde groups [
92
,
93
]. The main cellulose
modifications are presented in the following subsections.
3.2.1. Molecule Chemical Grafting
The reaction mechanism in this technique is to improve the structure and properties of
cellulose, and only happens on the cellulose chains located on the surface of the cellulose.
The limitation on the extent of acetylation (ester bonds are formed between cellulose and
cyclodextrins) lies in the susceptibility and ease to maintain the surface; however, this
technique does not make the cellulose fully dissolved because of the complex network
formation [94].
3.2.2. Surface Adsorption on Cellulose
The adsorption on the surface of cellulose is usually done by using surfactants. There
are many types of surfactants, such as fluorosurfactant, e.g., perfluorooctadecanoic acid
used to coat cellulose [
95
], cationic surfactant [
96
,
97
], and polyelectrolyte solution [
9
,
98
].
Surfactants improve hydrophobic behavior; however, they might also change the physical
properties and produce some cracks that possibly make absorption of water and moisture
occur.
3.2.3. Direct Chemical Modification Methods
The properties of cellulose, such as its hydrophilic or hydrophobic character, elasticity,
water sorbency, adsorptive or ion exchange capability, resistance to microbiological attack,
and thermal resistance are usually modified by chemical treatments. The main methods of
cellulose chemical modification are esterification, etherification, halogenations, oxidation,
and alkali treatment [
99
]. The chemical modifications methods of cellulose are the best
methods to achieve adequate structural durability and an efficient adsorption capacity.
3.2.4. Cellulose Grafting
Grafting on cellulose by attaching or adding the particles of molecules covalently
to the cellulose can be done by either using coupling agents or activating the cellulose
substrates [
88
]. Vinyl monomers grafting on cellulose can be performed in homogeneous or
heterogeneous medium [
100
]. When polyhydroxybutyrate (PHB) is grafted with cellulose,
it improves the properties of the cellulose in terms of crystallinity, flexibility, and the
chemically linking of the fibers with the matrix [
101
]. Figure 3illustrates the general
mechanism of peroxide radical initiated grafting of PHB onto cellulose. Polyethylene glycol
Polymers 2021,13, 231 7 of 35
(PEG) and aminosilane are also used as grafting materials [
102
,
103
], where these materials
increase the cellulose polarity to have better compatibility with the polymer.
Figure 3.
The general mechanism of peroxide radical initiated grafting of polyhydroxybutyrate (PHB) onto cellulose
(reproduced with copyright permission from Wei et al. [101]).
4. Mechanical Properties of Microcellulose and Nanocellulose
Developing eco-friendly, sustainable, and easily available materials has gained great
attention in recent years, due to environmental issues and the depletion of petroleum.
Exploring alternatives for petroleum-based polymers by extracting cellulose from natural
resources like plants, animals, and bacteria provides us with an excellent opportunity to
save the environment. Nevertheless, this natural cellulosic polymer has some property
shortages. Improving its properties extends the abilities of using natural fiber in wider
applications. Usually, improving mechanical properties results in enhancements in the
other properties, such as thermal and chemical properties. This is due to the modifications
applied to the natural fiber improving the homogeneity between the particles, hence
creating a good chemical bonding [
55
]. The important mechanical tests to assess the
enhancements resulted from the modifications are tensile strength, tensile modulus, flexural
strength, flexural modulus, elongation, and stiffness [104,105].
4.1. Tensile Strength
It is essential to know the maximum load that the material can withstand before
failure prior to application in any field. Adding microcellulose to polymers has improved
the tensile strength of many polymers, e.g., adding microcrystalline cellulose (MCC) pre-
pared from cotton fabric waste to poly(vinyl chloride) film enhanced the tensile strength
with the increase of MCC content [
106
]. When MCC was added to poly(vinyl alcohol),
the tensile strength continued to increase with MCC addition. The maximum tensile
strength was achieved with 10% of MCC load; beyond that value, the tensile strength
started to drop gradually [
107
]. The composite of polypropylene (PP)/MCC revealed an
insignificant decrease in the tensile strength; this problem can be solved by using coupling
agents, such as aminopropyltriethoxysilane and maleic anhydride-grafted polypropylene
(MAPP). In the case of using coupling agents, the tensile strength was improved, giving a
chance to add more MCC [
108
]. The composite of hydroxypropyl starch/MCC showed
improvement in the tensile strength, where the maximum load was 6% of MCC, and
the tensile strength started to decrease when adding more MCC [
109
]. MCC-reinforced
polypropylene composites using maleic anhydride polypropylene (MAPP) as a coupling
agent improved the tensile strength by 27% higher than that of the composite containing
only MAPP [
110
]. Reinforcing nano-clay with MCC improved the tensile strength of the
Polymers 2021,13, 231 8 of 35
composite at 7.5 wt% nano-clay [
111
]. Incorporation of nano-cellulose into gelatin and
starch matrices showed that increasing nanocellulose composition to 10% led to increasing
the tensile strength [
112
]. In the composite of poly(vinyl alcohol)/NCC and nanosilica, the
addition of NCC and nanosilica to poly(vinyl alcohol) improved the tensile strength [
113
].
Nanocellulose reinforced unsaturated polyester (UPR) revealed significant increment of
tensile strength with the addition of 0.5–3 wt% of NCC with maximum tensile strength
obtained for 0.5 wt% [
114
]. In another study, the composite of chitosan reinforced with
NCC improved the tensile strength result until it reached 245 MPa [115].
4.2. Flexural Strength
Flexural strength is defined as the maximum bending stress that the material can
withstand before yield. Adding modified MCC to cement mortar was found to improve
the flexural strength by double [
116
]. The composite of MCC and cementitious improved
the cementitious by 19.2% with 1% of MCC addition [
117
]. Using nutshell with MCC to fill
high-density polyethylene also made remarkable improvements in the flexural strength of
the high-density polyethylene [
118
]. In the composite of polymethylmethacrylate (PMMA)
and nanocrystalline cellulose (NCC), the addition of NCC improved the flexural strength
to around 3 MPa [
119
]. The addition of NCC to cement paste showed 20–30% improvement
in flexural strength, with only 0.2% of NCC [
120
,
121
]. The composite of NCC to epoxy
nanocomposites exhibited a 20% enhancement in the flexure strength [122].
4.3. Elastic Modulus (Young’s Modulus)
To measure the stiffness of a material, we need to measure the tensile modulus.
The addition of MCC to commercial acrylic adhesive increased the modulus of the compo-
sition [
123
]. Solution casting of MCC and organophilic silica (R972) applied as a filler to
poly(lactic acid) (PLA) revealed an improvement in the tensile modulus [
124
]. Polycapro-
lactone/MCC/wood flour composites exhibited lower tensile modulus, and the maximum
tensile modulus was gained from polycaprolactone/wood flour without MCC [125].
When the NCC was extracted from MCC and added to polyamide 6, the tensile
modulus of the composite increased for almost 10 times that of the polyamide 6
0
s [
126
].
The montmorillonite (MMT)/NCC reinforcing polylactic acid (PLA) hybrid showed an
improvement in the tensile modulus by increasing the load of NCC [
127
]. When NCC
was coated on woven jute/green epoxy composite, enhancement in the tensile modulus
occurred [128].
4.4. Flexural Modulus
Like all materials, we need to test the resistance of bending of the newly developed
material. When microcrystalline cellulose (MCC) was used as a filler to nylon 6, the filler
loading improved the flexural modulus from 2.6 GPa with neat nylon 6 to 3.8 GPa with
20% of MCC loading [
129
]. MCC was also used as reinforcement material to cementitious
composites and revealed 106% enhancement in the flexural modulus as a result of adding
MCC [
130
]. The addition of epoxidized citric acid to polylactide/MCC showed the highest
result of flexural modulus (4.7 MPa) at 3% of MCC, which was 3.3 MPa before adding
MCC [
131
]. For nanocellulose used as reinforcement of epoxy composite, the addition of
NCC increased the flexural modulus until it reached 3.1 GPa at 0.75% of NCC; adding more
NCC resulted in the drop of the flexural strength [
132
]. Poly(lactic acid) (PLA) reinforced
with NCC exhibited 106% improvement in terms of flexural strength [133].
4.5. Elongation at Break
The deformation that occurs before a material eventually breaks needs to be measured
to know the ductility of the material. The addition of MCC extracted from waste-cotton
fabric to the hybrid of poly(lactic acid), poly(butylene succinate) revealed acceptable elon-
gation while the other properties were improved [
134
]. When regenerated MCC was added
to epoxidized natural rubber blend film, the elongation was improved by 39% at 20 wt%
Polymers 2021,13, 231 9 of 35
loading of MCC [
135
]. Adding NCC extracted from MCC to poly(lactic acid) for packaging
applications showed improvement in elongation, reaching 205% [
136
]. Phosphorylated
nanocellulose fibrils added to PVA nanocomposites showed a significant drop in elongation
at break encountered with general improvements in the other properties [137].
5. Cellulose Aging Resistance
It is known that polymer and plastic materials tend to age under the interaction
of oxygen and heat; this aging leads to scrapping [
138
]. Therefore, resistance to aging
is an essential feature of cellulose. The importance of aging resistance depends on the
applications that the materials will be used for. Usually, when the thermal and mechanical
properties are improved, the aging resistance will also be improved. Enrolling cellulose
in polymers gives excellent potential to enhance the polymer’s aging resistance [
139
,
140
].
Under normal conditions, the aging resistance can be assessed by tracking the changes in
their interfacial shear strength (IFSS) results.
It is clear that adding ash and other non-cellulosic components to the biopolymer
reduces the properties of the material, whether these properties are mechanical or thermal,
but it has been found that adding ash to the biopolymer improves the aging resistance of
the biopolymer. Using cellulose ash as contribution filler improved the aging resistance
of the asphalt mixtures from 45.3% to 48.6% [
141
]. Adding charcoal ash coconut shell to
bitumen also improved the aging resistance of the bitumen [142].
The composite of polylactic acid (PLA)/linear low-density polyethylene (LLDPE)/
microcrystalline cellulose (MCC) fiber revealed a good aging resistance [
143
]. The bacterial
cellulose used to make electrical insulating paper enhanced its aging resistance [
144
].
The addition of nanocrystalline cellulose/attapulgite (AT) to polypropylene (PP) improves
aging resistance, thermal stability, and gas barrier properties [
145
]. However, the addition
of NCC/AT to PP decreased the mechanical properties and the degradation degree of PP.
The composition of nanocrystalline cellulose and potato starch showed improvement in
the aging resistance after bleaching the pulps [
146
]. Comparing the aging resistance of
aramid nanofiber (ANF) and nanocrystalline cellulose revealed that ANF has better aging
resistance than NCC [
147
,
148
]. The NCC boosted the carbon in ionic liquid supercapacitors
and improved the aging resistance of the carbon/NCC. This improved the composite’s
stability even after three months [
149
]. Removing moisture and water content exhibited
enhancement in the aging resistance of the biopolymer [
150
]. The nanocrystalline cellulose
with copper-based dye-sensitized solar cells demonstrated great improvement in the
percentage of aging resistance of the composite [151].
There are three methods to improve aging resistance; these methods are (1) poly-
hedral oligomeric silsesquioxane, in which some materials can be used to improve the
fiber’s resistance, such as modified polyhedral oligomeric silsesquioxane [
152
]. When
poly(p-phenylene benzobisoxazole) nanocomposite fiber was modified by using polyhedral
oligomeric silsesquioxane, the aging resistance of poly(p-phenylene benzobisoxazole) was
improved [
153
]. (2) Polyhedral oligomeric silsesquioxanes (POSS) grafting, using 3 amino-
propyltrimethoxysilane (APTMS) as a bridging agent. Polyhedral oligomeric silsesquiox-
anes and silane agent used to graft ZnO nanowires (NWs) onto poly(p-phenylene ben-
zobisoxazole) (PBO) fibers enhanced the aging resistance of the PBO and improved the
other properties [
154
]. (3) Polyhedral oligomeric silsesquioxane (POSS) derivatives in
an ionic liquid 1-allyl-3-methylimidazolium chloride (AmimCl). The dispersion of POSS
(both aminophenyl or nitrophenyl groups (POSS-AN, NH
2
:NO
2
= 2:6)) in cellulose matrix,
POSS-AN nanoparticles were uniformly dispersed in cellulose at nanoscale, the POSS-AN
provided better compatibility in both the AmimCl and cellulose/POSS nanocomposite
films and increased the aging resistance [155].
6. Comparison between Plant-Derived Nanocellulose and Bacterial Nanocellulose
Fibers
There are differences between plant-based and bacteria-derived nanocellulose fibers.
Bacterial nanocellulose possesses higher critical surface tension and higher thermal degra-
Polymers 2021,13, 231 10 of 35
dation temperature while the cellulose extracted from plant has hierarchical organization
and semi-crystalline nature. The cellulose-based plant is available while the bacterial
cellulose is limited [
156
], and the bacterial cellulose has a higher purity and crystallinity
degree than the plant cellulose [
157
]. Cellulose from plants takes a longer time to be
harvested, depending on the plants’ growing duration. Other factors, e.g., plant and soil
types, nutrients, climate conditions, and susceptibility to insect pest infestation, also con-
tribute to variances between plant-originated cellulose fibers and bacteria-derived cellulose
fibers. The plant-derived cellulose compositions are cellulose, lignin, hemicellulose, and
ash, where the nanocellulose extraction process requires energy for harvesting and lignin
removal. The drawback of harvesting is that it might cause damage to the environment
rather than saving it. On the other hand, the cellulose extracted from bacteria usually needs
only days to grow. After cell removal, the cellulose can be extracted in a pure mode. This
method demonstrates energy consumption in the sterilization of nutrients and bacterial
cellulose and the cell removal processes, which negatively impacts the environment by
increasing the air pollution [
158
]. Figure 4illustrates the relationship between different
kinds of nanocelluloses.
Figure 4.
Relationship between different kinds of nanocelluloses [
159
]. (Reproduced with copyright
permission from Creative Commons Attribution License 3.0).
7. Applications of Nanocellulose
Applications of any material depend on the appropriateness between the properties
of the material and the application specifications and standards. Whenever the proper-
ties of any material are improved and are easy to modify, the material will cover more
Polymers 2021,13, 231 11 of 35
applications. In this case, the improvement in the properties of nanocellulosic materials
will extend their applications, where the nanocelluloses applications can be divided by
their types. Bacterial nanocellulose’s main applications in the medical field are wound,
burn, and ulcer dressings [
160
]. It is also used in packaging [
161
], tissue engineering,
electronic, optical, sensor, and catalysis applications [
162
]; whereas NFC and MFC are used
as additives in the paper-making process [
163
], coating [
164
], medical, pharmaceutical,
cosmetic, hygienic [
165
], barrier material [
166
], high-temperature thermal insulation [
167
],
and chronic wound healing applications [
168
]. Because of the small diameter of NCC, it is
useful in the medical field applications, especially in the vascular graft [
169
], electronics,
catalysis [
170
], packaging applications [
171
,
172
], synthetic plastic or polymers, fuel cells,
filtration, catalysis, tissue engineering, solar cells, and lithium-ion batteries [
173
]. The NCC
bionanocomposites can also be implemented in fire extinguishers [
35
] and automotive com-
ponents [
174
,
175
] due to their high thermal stability and tensile strength. The applicability
of any composites is decided by its controlled durability under the circumstances it is being
used. Nanocellulose has a great developing direction on flexible electronics, 3D printing
technologies, and smart materials, as well as the medical and energy field applications.
Moreover, microcrystalline cellulose is used in many applications, such as abrasives in
cosmetics, absorbent, anti-caking, bulking, and aqueous viscosity increasing agents, binder,
emulsion stabilizer, slip modifier, and texturizer [176,177].
Designing biocomposite-based cellulose has numerous challenges due to the large
variety of cellulosic fibers, polymers, and manufacturing processes, since there are a wide
variety of types of reinforcements, dissimilar fiber geometry, and many possibilities for the
orientation and fiber arrangement being used especially in large scale design [
172
]. The
sustainability of the environment directly affects the economic development. By increasing
the area of cellulose application, the agriculture of cellulose sources will improve, which im-
proves other industries such as agriculture machines industry. Moreover, the development
of a growing industry like biocomposites creates more jobs and sub-industries.
8. Cellulose Fiber for Injection Molding
Injecting molding is the processing technique (co-extrusion [
178
] and melt processing
techniques [
179
]) that is used for natural fiber, similar to the synthesis polymers injection
methods [
180
,
181
]. This method is able to produce parts with very precise dimensions at
very low cycle times [
182
]. The approach used with cellulose is usually the foam injection
mold, because it can be used with many different-sized parts and densities [
183
] to produce
a high strength and lightweight molded parts. Many applications require thicker walls of
mold than standard injection molding due to producing products effectively. However,
the structural foam process allows for a quicker process and lower cycle time on thicker
parts and is able to produce complex three-dimensional geometries. Polymeric foams
can be prepared by different techniques, such as extrusion, batch, and bead foaming, as
well as foam injection molding [
184
]. Microcellular injection molding technology, also
known as the (MuCell) process, is a type of foam injection molding technique that produces
parts with excellent dimensional stability using a lower injection pressure [
180
], which
means lower cost. This technique is typically used to produce nanocellulose [
185
]. Injecting
cellulose nanocrystals with biodegradable poly(lactic acid) foam improved the composite
properties at a low cost [
186
]. Microcellular injection molding technology can also be
used with cellulose nanofibers. This technique that inject-molded polypropylene foams by
introducing hydrophobic-modified cellulose nanofibers has improved the dispersion of
cellulose nanofibers in the composite [
187
]. There are three types of foaming processing
technologies.
8.1. Batch Foaming Processing
Batch-foaming is a discontinuous foaming process that shows good reproducibility
due to the precise process control. It is qualified to be employed to investigate the foaming
behavior of polymers and polymer systems and is also used in industrial applications [
188
].
Polymers 2021,13, 231 12 of 35
In this method, the polymer is placed in a high-pressure chamber saturated with inert gas,
then, the polymer sample is put under heat and pressure to lower the gas solubility of the
polymer [189].
8.2. Extrusion Foaming Processing
Foam extrusion is a continuous process of high industrial relevance. It allows produc-
ing semi-finished products with foam densities below 50 kg/m
3
[
188
]. The temperature
must be minimized in this method to avoid emissions during the extrusion foaming pro-
cessing. This method can be performed with either chemical foaming agents (chemical
blowing agents (CBAs)) or physical blowing agents (PBAs) [190,191].
8.3. Injection Foam Molding Process
Foam injection molding processing is a composite material produced when a polymer,
usually thermoset or thermoplastic, is combined with either an inert physical gas, such
as nitrogen, or a chemical blowing agent during the molding process [
192
]. This method
is used to fabricate three-dimensional shapes of polymers using either chemical foaming
agents (chemical blowing agents (CBAs)) or physical blowing agents (PBAs) [193,194].
9. Nanocellulose Reinforcing Polymer Composite
In general, nanocellulose can be reinforced with synthetic polymer with significantly
different chemical and or physical properties when reinforced. Later, the combination of
both materials would produce a material with characteristics different from the individual
components. The term is usually referred to nanocomposites, which are generally divided
into nanocrystalline cellulose (NCC) filled synthetic polymers and nanofabrillated cellulose
(NFC) filled synthetic polymers [
195
]. The addition of nanocellulose has significantly
improved synthetic polymer properties, such as tensile strength and thermal conductiv-
ity. Synthetizing metal-organic framework (MOF) powders on nanocellulose template
offered conventional with excellent mechanical flexibility and porosities as well as shifted
the balance of growth and nucleation for synthesizing MOF microcrystals. This low-cost
production pathway is capable of transforming MOF into flexible and shapeable form
and thus range their applications in more wide fields [
196
]. Table 1exhibits some exam-
ples of nanocrystalline cellulose-filled synthetic polymer and the effect of nanocellulose
reinforcement.
The issue of reinforcement of nanocrystalline cellulose with synthetic polymer has
received considerable critical attention. Roohani et al. [
197
] studied the influence of NCC
contents on the morphological, dynamic mechanical, and tensile properties of cotton
NCC reinforced with copolymers of polyvinyl alcohol and polyvinyl acetate. For dynamic
mechanical analysis results, glass transition temperature rose significantly with the addition
of cotton NCC due to the formation of a water layer at the interface, which caused its
matrix to become less plasticized by water. On top of that, the tensile strength and Young’s
modulus seemed to increase with the addition of cotton NCC. However, the humidity
elevation caused a significant decrease in the tensile modulus due to the glass transition
temperature changes towards lower values, below the room temperature. Cao et al. [
198
]
also looked at the effect of NCC content but with differences in NCC sources and its
polymer matrix. They studied the NCC contents of flax fillers and tested for tensile,
dynamic mechanical, and thermal gravimetric properties of flax NCC to reinforce nitrile
composites. The composites performed an increase of tensile strength and storage modulus
with the increase of filler content. Moreover, the rise in flax NCC also resulted in glass
transition temperature (T
g
) of the composites that was shifted from 10.8 to 17.2
C. In terms
of thermal gravimetric analysis, the degradation temperature corresponding to flax NCC
in nanocomposites was significantly higher than pure flax NCC.
Polymers 2021,13, 231 13 of 35
Table 1. Illustration of the effect of nanocrystalline cellulose on synthetic polymer matrix.
Source of
Nanocrystalline
Cellulose
Synthetic Polymer The Effect and Advantages of the Reinforcement Ref.
Cotton Copolymers of polyvinyl
alcohol and polyvinyl acetate
-
Improvement of dynamic mechanical, tensile strength,
and Young’s modulus as NCC content increased.
- Glass transition temperature rose significantly as the
addition of cotton NCC.
-
Elevation of humidity resulted in significant declining
of the tensile modulus.
[197]
Flax Nitrile rubber
-
Increase of tensile strength (7.7–15.8 MPa) and storage
modulus.
- Higher thermal degradation temperature.
- Tgof the nanocomposites was shifted 10.8 to 17.2 C
with flax NCC content increasing to 20 phr.
[198]
Oil palm Polypropylene/cyclic natural
rubber
- Improvement of interphase surface tension and
thermal stability.
- More homogenous than without the addition of the
nano-fillers.
[199]
Pineapple Polyvinyl alcohol
- The nanocomposites are used to develop green
pH/magnetic sensitive hydrogels.
- Improvement in terms of thermal stability, swelling
ability.
- Significant enhancement of naringin loading and
entrapment efficiency of the hydrogels.
[200]
Sisal leaf Rubber - Increase in tensile properties of the nanocomposites. [201]
Softwood Kraft pulp Poly Vinyl Alcohol
-
The gelatin and NCC fillers help to increase the tensile
strength and Young’s modulus.
- The moisture absorption of NCC/gelatin/PVA
nanocomposites tremendously increased as compared
to PVA films.
- Thermal properties such as thermal degradation and
glass transition temperature enhanced.
- Better crystallinity due to the existence of glycosidic
bonds in cellulose structure.
[202]
Tunicin Epoxy
- Improved tensile strength due to good adhesion with
epoxy matrix.
- Provided better dynamic mechanical properties in
their synthetic polymer.
[203]
In addition, Mahendra et al. [
199
] carried out a study on the effect of oil palm NCC
and TEMPO-oxidized nanocellulose on the compatibility of polypropylene/cyclic nat-
ural rubber (PP/CNR) blends. The result showed that the addition of NCC enhanced
the mechanical properties of the polymer nanocomposites compared to the neat polymer.
Moreover, the improvement of NCC nanocomposites was also observed as the result
of interphase surface tension and thermal stability. A recent study was carried out by
Dai et al. [
200
] on the potential of the green method to fabricate green pH/magnetic sensi-
Polymers 2021,13, 231 14 of 35
tive hydrogels based on pineapple peel crystalline nanocellulose (rPPNc) and polyvinyl
alcohol. The rPPNc improved the thermal stability, swelling ability, naringin loading, and
entrapment efficiency of the hydrogels. A study conducted by Jain and Pradhan [
201
]
stressed that the mechanical properties of sisal NCC-rubber composites stress strain graphs
displayed a ductile fracture behavior, where a peak yield stress occurred followed by
necking and cold drawing. The increase of sisal NCC fillers within 5–10% in the nanocom-
posites would increase about 0.365 to 0.360 MPa compared to pure rubber. The process of
extraction of NCC fillers from sisal leaves in this research is the acid hydrolysis method.
However, the tensile strength of sisal nanocomposite was less than the sisal fiber composite
due to the lack of homogeneity in mixing sisal NCC in latex and weak bonding between
cellulose and latex.
There have been several investigations into the causes of the effect of nanofibrillated
cellulose when reinforced with polymer. Karmaker et al. [
202
] implemented solution
casting to fabricate PVA-gelatin films with the addition of NCC fillers from softwood kraft
pulp. The addition of nanofiller showed drastic changes with its mechanical and thermal
properties. Those tensile moduli and strength were significantly increased as gelatin and
NCC fillers were added in the nanocomposite, which contributed to low elongation at
break. Moreover, the addition of NCC filler in the PVA-gelatin nanocomposites reduced the
moisture absorption as their thermal properties was improved. The surface morphology of
the nanocomposite permitted better crystallinity due to the existence of glycosidic bonds in
cellulose structure. Another study executed by Xu et al. [
203
] found that tunicin cellulose
had potent effects in terms of mechanical and thermal properties in epoxy nanocomposites.
The storage modulus and T
g
of the nanocomposites were significantly enhanced when the
increase of tunicin NCC filler. For instance, with 15 wt% of tunicin NCC fillers, the storage
modulus was increased by 100% relative to pure epoxy, while their T
g
increased to 75.5
C.
Moreover, the inclusion of NCC in epoxy nanocomposites tremendously increased the
tensile strength up to 60 MPa due to good surface adhesion NCC filler with epoxy matrix.
In addition to that, another well-known nanofiller is nanofibrillated cellulose (NFC). It is
commonly used in packaging and automotive applications due to its enhanced thermal,
mechanical, and crystallinity properties. Table 2summarizes the effect of NFC in synthetic
polymer matrix composites.
Table 2. Illustration of the effects of nanofibrillated cellulose on synthetic polymer matrix.
Source of Nanofibrillated
Cellulose Synthetic Polymer The Effect of the Reinforcement Ref.
Bagasse pulp Aluminium nitrite
- The effect of silane treatment along with NFC substrate
improved the thermal conductivity.
- The composite film is highly suitable for green electronic
devices applications.
[
204
]
Banana Epoxy
- Reduce the water uptake of nanocomposite film,
especially at 5 wt% of NFC.
- The mechanical and dynamic mechanical properties
significantly improved at 2–3 wt% of NFCs.
- Act as a catalytic curing agent.
[
205
]
Northern bleached softwood
kraft (NBSF) pulp Epoxy
- Improved the nanocomposite storage modulus as well as
their tan δ.
- Tensile and flexural properties significantly increased as
the inclusion of NFC fillers.
- Thermal stability and residual char of kenaf/epoxy
composites was well enhanced.
[
206
]
Polymers 2021,13, 231 15 of 35
Table 2. Cont.
Source of Nanofibrillated
Cellulose Synthetic Polymer The Effect of the Reinforcement Ref.
Western red cedar Epoxy
- Strong reinforcing effects displayed by the high residual
lignin containing NFCs on the mechanical, physical,
thermal properties of the nanocomposite.
- High residual lignin of NFC provided impermeable
medium for moisture in epoxy composites.
[
207
]
White bamboo Epoxy
- Improvement in tensile and flexural properties, fracture
toughness, as well as thermal property especially at 0.3
wt% of NFC.
- Better in dynamic mechanical properties in both tensile
and bending condition as the addition of NFCs fillers at
0.3 wt%.
[
208
]
Bamboo Starch/PVA
- Better homogeneity, cohesion, and more compact
structure, which promotes larger crystals in the
nanocomposite.
- Tensile strength and elongation at break improved at 24
and 15% as compared to the control blend.
[
209
]
As mentioned by Zhang et al. [
204
], bagasse pulp NFC was blended with
γ
-aminoprop-
yltriethoxysilane treated aluminum nitride nanosheets (TAlN). The results found that
dispersibility of the AlN nanosheets in the NFC substrate was enhanced because of the
silane treatment. This happened due to the treatment that lowered scattering between the
AlN and NFC interfaces, which induced better thermal conductivity. The new material has
shown enhancement of the mechanical properties of nanocellulose reinforced synthetic
polymer composites. Pandurangan and Kanny [
205
] evaluated morphological and curing
properties of banana NFC-filled epoxy composites. They stressed that the banana NFC
fillers acted as a catalytic curing agent by increasing the cross-link density during gelation of
epoxy. Moreover, good dispersion of the banana NFC particles in the matrix contributed to
10% increase in tensile strength and 26% increase in elongation at 3 wt% of NFC filled epoxy
nanocomposite. Along with the mechanical properties, nanocomposite film’s dynamic
mechanical properties were improved, especially at 2–3 wt% of NFC filler. Water uptake
results suggested that the water uptake of the NFC filled epoxy nanocomposites was
reduced as with higher concentration of NFC particles. The same study was also conducted
by Nair et al. [
207
] by using western red cedar NFC with high residual lignin. They found
a significant increase in terms of mechanical, thermal, and water barrier properties of the
high residual lignin NFCs nanocomposite.
Recently, studies on nanocellulose reinforced epoxy have been performed by re-
searchers in many fields such as aerospace, automotive, and marine construction [
205
208
].
Vu et al. [
208
] carried out a study on the influence of micro/nano white bamboo fibrils
on the physical characteristics of epoxy resin reinforced composites. They found that the
inclusion of the NFCs increased the flexural and tensile behaviors, fracture toughness, and
thermal properties of the nanocomposite. The presence of white bamboo NFCs enhanced
the tensile and flexural moduli, which exhibited improvement in the nanocomposite’s stiff-
ness. Another study by Junior et al. [
209
] also looked at the impact of the nanofibrillation
of bamboo pulp but with a slight difference in focus from Vu et al. [
208
]. They studied
the nanocomposite based on starch/PVA blend and tested for thermal, structural, and
mechanical properties. The results indicated that the higher NFC filler encouraged better
homogeneity, cohesion, and more compact structure, which promoted larger crystals in
Polymers 2021,13, 231 16 of 35
the nanocomposite. In addition, the tensile strength and elongation at break improved at
24 and 15% as compared to the control blend.
10. Nanocellulose Reinforcing Biopolymer Composite
As nanocellulose is used to reinforce synthesis polymer, it is also employed to reinforce
biopolymer from natural resources to improve the properties of the natural polymer, such as
polylactic acid (PLA), polyhydroxy acids (PHA), polyhydroxybutyrate (PHB), polybutylene
succinate (PBS), and starch biopolymer [210,211]. Tables 3and 4illustrate some examples
of NCC and NFC reinforced biopolymer and show the effects of NCC as well as NFC fillers.
Table 3. Illustration of the effect of nanocrystalline cellulose on biopolymer matrix.
Source of Nanocrystalline
Cellulose Source of Biopolymer The Effect of the Reinforcement Ref.
Maize starch Polylactic acid
- Addition of starch NCC filler showed the PLA
nanocomposite to have high potential to
improve the oxygen barrier and tensile
properties.
- Provided better filler dispersion and
interaction with the matrix.
[212]
Nata-de-coco Polylactic acid
- Enhancement in viscoelastic properties up to
175% in terms of storage modulus in bending.
- Addition of 2 wt% nanocellulose into PLA
resulted in moderate strength improvement.
[213]
Bamboo pulp Polylactic acid
- PLA-grafted NCC (PLA-g-NCC) films display
uniform dispersion of NCC due to the efficient
grafting, results in enhancement in tensile
strength.
- The elastic and crystallinity properties of the
nanocomposites improved with increasing of
NCC loadings.
[214]
Coffee silver skin Polylactic acid
- NCC with loading of 3 wt% in PLA film
enhanced water barrier and mechanical
properties of nanocomposites.
- Nanocomposite with NCC can overcome
drawbacks of biopolymer film.
[215]
Microcrystalline cellulose Polylactic acid
- NCC-reinforced PLA exhibited improvement
in thermal, mechanical, and UV barrier
properties. [216]
Microcrystalline cellulose Polylactic acid
- The reinforcement of the polyethylene glycol
(PEG) and NCC improved the crystallinity of
the PLA.
- The impact and the elongation at break
increased from 0.864 to 2.64 kJ, and 22 from
11% to 106.0%, respectively.
[217]
Polymers 2021,13, 231 17 of 35
Table 3. Cont.
Source of Nanocrystalline
Cellulose Source of Biopolymer The Effect of the Reinforcement Ref.
Microcrystalline cellulose Polylactic acid
- The addition of NCC into the PLA showed an
increment on tensile strength of PLA and
PLA-g-silane nanofiber.
- The modified PLA nanocomposite considered
as a practical candidate for hard tissue
engineering applications according to
cytotoxicity results.
[218]
Microcrystalline cellulose Polylactic acid
- Acetylation can improve the performance of
the composite by enabling linkages between
carbonyl groups, helping to establish a good
stress transfer between the fiber and the
matrix.
[219]
Nanocrystalline cellulose Polyhydroxy acids
-
Nanocellulose-reinforced PHA films improved
the mechanical properties by 23% compared to
neat PHA samples.
-
Increase of the crystallinity and stiffness of the
nanocomposites.
- Surface roughness of the nanocomposites was
increased, which contributed to better
interlaminar bonding in multi-layer
composites applications.
- Presence of UV blocking effect.
[220]
Kenaf Polyhydroxy acids - Improve the conductivity of the polymer
nanocomposites. [221]
Bleached pulp board Polyhydroxybutyrate
- Nanocellulose worked as heterogeneous
nucleating agent in PHB.
- Crystallinity of polymer was reduced and
improved the toughness of PHB.
- The mechanical properties of the
nanocomposites such as Young’s modulus and
elongation at break increased by 18.4% and
91.2%, respectively.
[222]
Nanocrystalline cellulose
Poly(3-hydroxybutyrate-
co-3-hydroxyvalerate)
(PHBV)
- NCC was dispersed evenly in GMA-g-PHBV.
- Limited reinforcement observed despite
enhanced dispersion relative to the neat PHBV
matrix due to the hydrophobization surface of
NCC.
[223]
Polymers 2021,13, 231 18 of 35
Table 3. Cont.
Source of Nanocrystalline
Cellulose Source of Biopolymer The Effect of the Reinforcement Ref.
Nanocrystalline cellulose Polybutylene succinate
- Restricted the mobility of polymer chains and
promoted nucleation and recrystallization of
polymer.
- Degree of crystallinity increased from 65.9 to
75.6%.
- The tensile strength increased from 23.2 MPa
to 32.9 MPa.
- Oxygen transmission rate of PBS films was
decreased from 737.7 to 280 cc/m2/day.
- Water vapor transmission rate (WVTR) of PBS
films decreased from 83.8 to 49.4 g/m2/day.
[224]
Microcrystalline cellulose
(MCC)
Poly(butylene succinate)
(PBS)/polylactic acid
(PLA)
- Impact strength, moduli, and crystallinity of
the nanocomposites increased.
- Thermal stability, storage modulus, glass
translation temperature of nanocomposites
increased.
[225]
Cotton Cationic starch
- Increased in tensile strength, oil, and air
resistance of the coated paper composites with
the optimized amount for the NCC
nanoparticles was 5 wt%.
- Water absorption of the coated paper
composite decreased by 50% at 5 wt% NCC
concentration.
[226]
Eggshell Corn starch
- The eggshell nanofiller was uniformly
dispersed and reinforced within film matrix.
- Tensile properties, thermal stability, water
vapor, and oxygen barrier properties were also
tremendously improved as compared to pure
starch film.
[227]
Orange peel Starch
- Significant improvement of water barrier
properties at 2 wt% concentration.
- Well dispersed in starch matrix during
formation of biofilms.
- Act as compatibilizer.
[228]
Rattan biomass Sago starch - Rattan NCC filler decreased the water uptake
of the bionanocomposite. [229]
Water hyacinth Bengkuang (Pachyrhizus
erosus) starch
- 60 min vibrated water hyacinth
NCC/bengkuang bionanocomposites had slow
biodegradation rate.
- The optimum 60 min vibrated samples
resulted in high thermal stability and low
moisture absorption rate.
[230]
Polymers 2021,13, 231 19 of 35
Table 3. Cont.
Source of Nanocrystalline
Cellulose Source of Biopolymer The Effect of the Reinforcement Ref.
Sugar palm fiber Sugar palm
starch
- Improvement in the water barrier property
and water vapor permeability (WVP) of the
nano composite film by 19.94%.
- Improvement in mechanical, thermal, and
physical properties.
[231,232]
Kenaf fibers Cassava starch
- Enhancement in the tensile strength and
modulus of the biocomposite films.
- Decreased the water absorption by the
biocomposite or the water sensitivity.
[233]
Garlic stalks Corn starch
- Scanning electron micrographs of the films
showed homogeneous dispersion of
nanocrystalline cellulose in the starch matrix.
- Improvement in tensile strength and modulus
and improvement in moisture property.
[234]
Kenaf fibers k-carrageenan
- The biocomposite showed enhancement in
mechanical properties and thermal stability.
- The biocomposite film shows a good
dispersion of the cellulosic fiber on the starch
matrix.
[235]
Sugarcane bagasse fiber Maize starch -
Improvement in water vapor barrier properties
with addition of nanocrystalline cellulose. [236]
Cotton cellulosepowders Plasticized starch - Improvement in the thermal stability,
mechanical properties, and air permeability. [171]
Potato peelwaste Potato starch - Enhancement in tensile modulus and water
permeability property. [237]
Sugarcane Bagasse Tapioca Starch
- The nanocellulose was found in good
dispersion in starch-based tapioca
biocomposite.
- Resulting in good adhesion bonding.
- Improved tensile strength up to 20.84 MPa
with the incorporation of 4% nanocellulose.
[238]
Table 4. Illustration of the effect of nanofibrillated cellulose on the biopolymer matrix.
Source of Nanofabrillated
Cellulose Biopolymer The Effect of the Reinforcement Ref.
Kenaf pulp Polylactic acid
- The tensile properties of nanocomposites
indicated that strength and modulus were
improved with increasing NFC contents. [239]
Polymers 2021,13, 231 20 of 35
Table 4. Cont.
Source of Nanofabrillated
Cellulose Biopolymer The Effect of the Reinforcement Ref.
Banana waste Polylactic acid
- The incorporation of 20 wt% of glycerol
triacetate and 1 wt% of nanocellulose doubled
the degree of crystallinity.
-
Dynamic mechanical thermal analysis (DMTA)
exhibited a 30 to 50% reduction in storage
modulus (stiffness) when compared to neat
PLA.
[240]
Nata-de-coco Polylactic acid
- The tensile modulus of the laminated
nanocellulose composites was found
increasing (from 12.5–13.5 GPa), insensitive to
the number of sheets of nanocellulose in the
composites.
-
Tensile strength of the laminated nanocellulose
composites decreased by 21% (from 121 MPa
to 95 MPa) when the number of reinforcing
nanocellulose sheets increased from 1 to 12
sheets.
[241]
Linter pulp Polylactic acid
- The impact strength, tensile strength, and
Young’s modulus of nanocomposites
(PLA/CNF5/PLAgMA5) increased by 131%,
138%, and 40%, respectively, compared to neat
PLA, with increasing of nanocellulose.
[242]
Kenaf Polylactic acid
- The strength and tensile modulus increased
from 58 MPa to 71 MPa, and from 2.9 GPa to
3.6 GPa, respectively, for nanocomposites with
loading of 5 wt% NFC.
- The storage modulus of the nanocomposites
increased compared to neat PLA.
-
The addition of NFC shifted the tan delta peak
towards higher temperatures.
- The tan delta peak of the PLA shifted from 70
C to 76 C for composites with 5 wt% CNF.
[243]
Carrot pomace Polylactic acid
- The incorporation of nanocellulose increased
hydrophilicity.
- The transmission rates of oxygen, carbon
dioxide, and nitrogen increased after
incorporating nanocellulose into PLA.
[243]
Bleached birch Kraft pulp Polyhydroxyalkanoates
(PHA)
- The reinforcement of nanocellulose with
polymers improved mechanical properties,
water contact resistance, and higher barrier
performance against water vapor compared to
the neat nanopapers.
[244]
Polymers 2021,13, 231 21 of 35
Table 4. Cont.
Source of Nanofabrillated
Cellulose Biopolymer The Effect of the Reinforcement Ref.
Ethyl cellulose Poly(ethylene
glycol)dimethacrylate
- Improved the compressive strength of
nanocomposites. [245]
Bleached pulp board Polyhydroxybutyrate
(PHB)
- The light transmittance, tensile strength, and
elongation at break were reduced.
- The crystallinity, thermal properties, and
Young’s modulus were increased.
[222]
Bleached Kraft eucalyptus
fibers
poly (3-hydroxybutyrate-
co-3-hydroxyvalerate,
PHBV)
- Incorporation of nanocellulose increased
tensile modulus, thermal degradation, and
storage modulus.
- Nanocellulose promotes the early onset of
crystallization.
- Inhibit foaming.
- Decreased the solubility of CO2and increased
desorption diffusivity.
[246]
Regenerated cellulose poly(3 hydroxybutyrate)
(PHB)
- Increased loading of regenerated cellulose
decreased the tensile strength and elongation
at break. [247]
Nanofibrillated cellulose Polybutylene succinate
(PBS)
- Incorporation of nanocellulose has drastically
increased the crystallinity of nanocomposites,
thus acting as nucleating agents.
- Form flexible nanocomposite films.
- Improve the mechanical properties of
nanocomposite films.
[246]
Wood cellulose pulps Chitosan
- The mechanical properties and thermal
stability of chitosan nanocomposite foams
increased.
- The chitosan nanocomposite foams displayed
a highly efficient water/oil separation capacity
even at 90 C.
- Goof biocompatibility with L929 mouse
fibroblasts.
[248]
Bleached pine sulfite
dissolving pulp Chitosan
- The incorporation of NFC improved the
mechanical properties of composites of
chitosan hydrogel matrices. [249]
Agave tequilana Weber Corn starch - Drastic improvement in term of tensile,
flexural, and impact performance. [250]
Bamboo helocellulose Thermoplastic starch
- Bamboo NFC filler well-dispersed in TPS
matrix which contributed high in tensile
properties.
- Lesser in water uptake of bionanocomposite.
[251]
Polymers 2021,13, 231 22 of 35
Table 4. Cont.
Source of Nanofabrillated
Cellulose Biopolymer The Effect of the Reinforcement Ref.
Cassava residue cellulose Cassava starch
- Improved the tensile strength, hydrophobicity,
and water vapor transmission coefficient of the
bionanocomposite films by 1034%, 129.4%, and
35.95%, respectively.
- Improved dispersibility with those fibrils that
were detached from each other.
[252]
Eucalyptus Waxy corn starch
- The moisture content, water solubility, and
water vapor permeability were significantly
reduced by the presence of NFC filler for both
regular and waxy starch films.
- Thermal and tensile properties also increased
at only 1% of suspension.
[253]
Pineapple leaf Thermoplastic potato
starch
- Polymer chain confinement around NFC filler
had excellent dispersion and superior
interaction between matrix and NFC filler.
- Barrier properties were enhanced.
[254]
Softwood alpha cellulose pulp
Cationic starch
-
Improved the tensile and burst strengths of the
paper composites.
- Contributed to enhancement of retention and
drainage of pulp paper due to interaction
between fillers and polymers.
- Improved brightness of paper.
[255]
Softwood cellulose pulp Modified starch
- The implementation of acetyl oxides starch
with TEMPO-oxidized cellulose nanofibre
(TNFC) filler resulted in less swelling inside
water and highest wet tensile behaviors.
[256]
Kenaf fibers Maize starch
- Addition of nanofibrillated cellulose to the
starch enhanced the mechanical properties (in
terms of tensile strength and Young’s modulus)
and the thermal stability of the
nanocomposites.
- Reduced moisture absorption.
- Decreased water sensitivity.
[257,258]
Bamboo nanofibers Cassava starch
- The nanofibrillated cellulose increased tensile
strength of 50% of starch films, while the
elongation at break showed similar increase
(66%) at concentration of 1.0 g/100 g of
nanofibrillated cellulose.
-
Improvement in the structure of the composite.
[259]
Polymers 2021,13, 231 23 of 35
Table 4. Cont.
Source of Nanofabrillated
Cellulose Biopolymer The Effect of the Reinforcement Ref.
Rice straw Potato starch
-
The yield strength and Young’s modulus of the
nanocomposite enhanced after adding the
nanofibrillated cellulose to the starch.
- The glass transition temperature increased.
- The humidity absorption resistance of films
was significantly enhanced by using 10 wt%
cellulose nanofibers.
- The transparency of the nanocomposites was
reduced compared to the pure starch
composite.
[260]
Sugar palm Sugar palm starch
- Improved water absorption and water
solubility properties of the nanocomposite
films by 18.84% and 39.38%, respectively.
- Good compatibility between the
nanofibrillated cellulose and the sugar palm
fiber, the composition created intermolecular
hydrogen bonds between them.
[261264]
Vaezi et al. [
226
] evaluated the effect of coatings on bionanocomposites of cationic
starch/cotton NCC in the application of paper packaging. They discovered that the increase
in NCC nanoparticle loading increased the tensile strength, oil resistance, and air resistance
of the coated paper, with the optimized amount of 5 wt% of NCC nanoparticles. In terms
of physical properties, the water absorption of the coated paper decreased by 50% at 5 wt%
NCC concentration. Jiang et al. [
227
] performed a study regarding the impact of eggshell
NCC filler on the overall properties of corn starch films. They showed that the inclusion of
the fillers noteworthily boosted the tensile behavior, thermal stability, oxygen, and water
vapor barrier properties compared to the pure corn starch film. This happened due to the
inductive effect between the C–C bonds on cornstarch skeleton and the O–C–O bond on
calcium carbonate in the eggshell nanofiller, which later contributed to strong interaction
and biocompatibility between the two components.
Another well-known source of NCC filler reinforced in biocomposites is orange peel.
It has been widely used in the packaging industry in order to develop high durability
biofilm. A study conducted by Fath et al. [
228
] found that orange peel NCC filler had
good dispersion in their matrix. The filler also acted as a compatibilizer, which improved
the physical interaction between NCC filler-starch films. The inclusion of the fillers also
decreased the water vapor transmission rate, especially at 2 wt% concentration. A research
study carried out by Nasution et al. [
229
] measured the effect of filler loading and co-
plasticizer addition on rattan NCC-filled sago starch bionanocomposite. They stressed that
the lowest water absorption rate was 9.37% at an additional of 3 wt% rattan NCC filler and
10 wt% acetic acid as compared to pure films.
For the effect of acetyl treatment of starch NCC filler to overall properties of the PLA
based nanocomposites, the nanocomposites of 1 and 3 wt% fiber loadings of untreated and
treated fiber as well as control were tested for the morphological, barrier, and mechanical
properties [
212
]. It was observed that the PLA nanocomposites with treated starch NCC
filler provided better filler dispersion and interaction with the matrix. Moreover, it also had
a high potential to improve the oxygen barrier and tensile properties of PLA nanocomposite.
Syafri et al. [
230
] studied the influence of sonication time on thermal stability, biodegra-
dation, and moisture absorption performance of water hyacinth NCC/bengkuang starch
bionanocomposites. The result showed that the bionanocomposite vibrated for 60 min
Polymers 2021,13, 231 24 of 35
had the highest thermal stability and presented with low moisture absorption capability.
Furthermore, the 60 min-vibrated nanobiocomposite had low porosity formation and a
coarse surface. The bionanocomposite also had a slow biodegradation rate, which is highly
suitable for the application of food packaging bags. As well as NCC-filled bionanocom-
posites, NFC filler can be implemented in a biopolymer matrix to form durable composite
films. Table 4illustrates some examples of NFC-filled biopolymer composite and shows
the effect of NFC fillers.
A study conducted by Lomelí-Ramírez et al. [
250
] studied the mechanical properties
of dried and hydrated Agave tequilana Weber NFC filled in corn starch bionanocomposite.
It showed tremendous improvement in terms of tensile, flexural, and impact performance
due to a small amount of 1 wt% of NFC filler. This could be attributed to an increase of
stiffness caused by NFC filler in the bionanocomposite. Pitiphatharaworachot et al. [
251
]
evaluated the bamboo holocellulose NFC fillers reinforced in thermoplastic starch (TPS)
bionanocomposites. The NFC filler was prepared from bamboo holocellulose powder using
TEMPO-mediated oxidation. They established that bamboo NFC filler was individually
dispersed with TPS matrix, which subsequently contributed to less water uptake, high
transparency, better tensile strength, and better modulus as compared to pure TPS films.
The optimum concentration of NFC filler was 1.5 wt%.
In accordance with Huang et al. [
221
], they carried out research to compare the effects
of various modification methods (silane and malic acid treatment) on physical and chemical
properties of cassava-filled cassava starch bionanocomposites. In their findings, modified
NFC filler significantly improved dispersibility with those fibrils that were detached from
each other. Morphologically, they formed three-dimensional network structures with
no occurrence of coarse fiber aggregation. The inclusion of modified cassava NFC filler
improved the tensile strength, hydrophobicity, and water vapor transmission coefficient of
the bionanocomposite films by 1034%, 129.4%, and 35.95%, respectively.
Furthermore, De Almeida et al. [
253
] performed a study on thermal, physical, and
mechanical behaviors of regular and waxy corn starch films reinforced with eucalyptus
NFC filler. The moisture content, water solubility, and water vapor permeability were
significantly reduced by the presence of NFC filler for both regular and waxy starch
films. It was possible to observe that the addition of NFC filler enhanced the thermal
and tensile properties of the bionanocomposite since only 1% of the suspension was
added. Balakrishnanan et al. [
254
] studied the effect of filler loading on morphological,
transport property, and viscoelastic polymer chain confinement of pineapple leaf NFC
filled in thermoplastic potato starch bionanocomposites. They confirmed that the polymer
chain confinement around the NFC filler has excellent dispersion and superior interaction
between matrix and NFC filler. The bionanocomposites obey the pseudo-fickian property.
In terms of barrier properties, the addition of NFC filler at 3 wt% concentration also
resulted in its enhancement. However, further increased filler content would depreciate
the properties due to the agglomeration of fiber.
Another study by Tajik et al. [
255
] evaluated the impact of cationic starch in the pres-
ence of NFC filler on the structural, optical, and strength properties of paper. They found
retention and reinforcing effects of the additives on the paper network. In this manner, the
mechanical properties such as tensile and burst strengths were drastically increased as the
increasing levels of the additives were up to 33% and 23%, respectively, for 0.6 CS/2% NFCs
filler paper bionanocomposites. The bagasse NFCs filler also contributed to the improve-
ment of retention and drainage of pulp at lower levels because of the interaction between
filler and starch polymer. In addition, a higher concentration of the nanofillers improved
the brightness of paper. Research work by Soni et al. [
256
] blended TEMPO-oxidized NFCs
fillers inside three modified starches, namely hydroxypropyl starch (HPS), acetyl starch
(AS), and acetyl oxidized starch (AOS), to evaluate their mechanical strength and durability
in water. They discovered that TNFC/acetyl oxidized starch biofilm displayed less water
swelling and improved wet tensile properties due to the formation of hemiacetal between
Polymers 2021,13, 231 25 of 35
nanofiller and starch polymer. TNFC/HPS biofilm illustrates the highest wet stiffness with
the minimum swelling in water.
11. Conclusions
There is no escape from stopping or at least minimizing the usage of non-degradable
petroleum-sourced materials to protect the environment. Micro- and nanocellulose are
good alternatives to manufacture composites with either natural starch or synthetic matrix.
The cellulose classification and extraction methods have been highlighted. The develop-
ment of composite materials containing cellulose has improved in terms of mechanical,
thermal, and aging resistance. Further improvements to enhance the dispersion and com-
patibility of cellulose have been discussed. These advancements have been highlighted to
reveal the high potential of cellulose-based composite for a large number of applications.
The increment of using plastic materials in the world requires more development in bio-
plastic materials to replace petroleum-based synthesis polymers. Processes of extracting,
isolating, and injecting cellulose require more studies and adjustments to improve the
properties of the biopolymers and to have broader applications.
Author Contributions:
Data curation, A.A.B.O. and S.M.S.; Formal analysis, R.A.I. and M.R.M.A.;
Funding acquisition, A.A.B.O., S.M.S., S.S.R.K., and M.P.; Project administration, A.A.B.O., S.M.S.,
S.S.R.K., and M.P.; Writing—original draft, A.A.B.O., R.A.I., A.A.B.A.M., and M.R.M.A.; Writing—review
& editing, R.A.I., A.A.B.A.M., M.R.M.A., S.S.R.K., and M.P. All authors have read and agreed to the
published version of the manuscript.
Funding:
This research was funded by BOLD2020 grant coded RJO10517844/112 by Innovation & Re-
search Management Center (iRMC), Universiti Tenaga Nasional, Malaysia. This research was also sup-
ported by the Ministry of Education, Youth and Sports of the Czech Republic and the European Union
(European Structural and Investment Funds—Operational Programme Research, Development and
Education) in the frames of the project “Modular platform for autonomous chassis of specialized elec-
tric vehicles for freight and equipment transportation”, Reg. No. CZ.02.1.01/0.0/0.0/16_025/0007293.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement:
The data presented in this study are available on request from the
corresponding author.
Acknowledgments:
The authors would like express gratitude for the financial support received
from BOLD2020 grant (RJO10517844/112) by Innovation & Research Management Center (iRMC),
Universiti Tenaga Nasional, Malaysia. Ministry of Education, Youth and Sports of the Czech Republic
and the European Union (European Structural and Investment Funds—Operational Programme
Research, Development and Education).
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
Sapuan, S.M.; Aulia, H.S.; Ilyas, R.A.; Atiqah, A.; Dele-Afolabi, T.T.; Nurazzi, M.N.; Supian, A.B.M.; Atikah, M.S.N. Mechanical
properties of longitudinal basalt/woven-glass-fiber-reinforced unsaturated polyester-resin hybrid composites. Polymers
2020
,12,
2211. [CrossRef] [PubMed]
2.
Asyraf, M.R.M.; Ishak, M.R.; Sapuan, S.M.; Yidris, N.; Ilyas, R.A. Woods and composites cantilever beam: A comprehensive
review of experimental and numerical creep methodologies. J. Mater. Res. Technol. 2020,9, 6759–6776. [CrossRef]
3.
Ilyas, R.A.; Sapuan, S.M.; Atiqah, A.; Ibrahim, R.; Abral, H.; Ishak, M.R.; Zainudin, E.S.; Nurazzi, N.M.; Atikah, M.S.N.; Ansari,
M.N.M.; et al. Sugar palm (Arenga pinnata [Wurmb.]Merr) starch films containing sugar palm nanofibrillated cellulose as
reinforcement: Water barrier properties. Polym. Compos. 2020,41, 459–467. [CrossRef]
4.
Abdul Khalil, H.P.S.; Bhat, A.H.; Yusra, A.F.I. Green composites from sustainable cellulose nanofibrils: A review. Carbohydr.
Polym. 2012,87, 963–979. [CrossRef]
5.
Syafiq, R.; Sapuan, S.M.; Zuhri, M.Y.M.; Ilyas, R.A.; Nazrin, A.; Sherwani, S.F.K.; Khalina, A. Antimicrobial activities of starch-
based biopolymers and biocomposites incorporated with plant essential oils: A review. Polymers
2020
,12, 2403. [CrossRef]
[PubMed]
Polymers 2021,13, 231 26 of 35
6.
Shankar, S.; Rhim, J.W. Preparation of nanocellulose from micro-crystalline cellulose: The effect on the performance and properties
of agar-based composite films. Carbohydr. Polym. 2016,135, 18–26. [CrossRef] [PubMed]
7.
Tian, C.; Yi, J.; Wu, Y.; Wu, Q.; Qing, Y.; Wang, L. Preparation of highly charged cellulose nanofibrils using high-pressure
homogenization coupled with strong acid hydrolysis pretreatments. Carbohydr. Polym.
2016
,136, 485–492. [CrossRef] [PubMed]
8.
Nair, S.S.; Zhu, J.Y.; Deng, Y.; Ragauskas, A.J. Characterization of cellulose nanofibrillation by micro grinding. J. Nanopart. Res.
2014,16, 2349. [CrossRef]
9.
Mocchiutti, P.; Schnell, C.N.; Rossi, G.D.; Peresin, M.S.; Zanuttini, M.A.; Galván, M.V. Cationic and anionic polyelectrolyte
complexes of xylan and chitosan. Interaction with lignocellulosic surfaces. Carbohydr. Polym. 2016,150, 89–98. [CrossRef]
10.
Zeng, J.; Tong, Z.; Wang, L.; Zhu, J.Y.; Ingram, L. Isolation and structural characterization of sugarcane bagasse lignin after dilute
phosphoric acid plus steam explosion pretreatment and its effect on cellulose hydrolysis. Bioresour. Technol.
2014
,154, 274–281.
[CrossRef]
11.
Wang, W.; Liang, T.; Bai, H.; Dong, W.; Liu, X. All cellulose composites based on cellulose diacetate and nanofibrillated cellulose
prepared by alkali treatment. Carbohydr. Polym. 2018,179, 297–304. [CrossRef] [PubMed]
12.
Kaur, V.; Chattopadhyay, D.P.; Kaur, S. Study on Extraction of Bamboo Fibres from Raw Bamboo Fibres Bundles Using Different
Retting Techniques. Text. Light Ind. Sci. Technol. 2013,2, 174–179.
13.
Li, Z.; Meng, C.; Yu, C. Analysis of oxidized cellulose introduced into ramie fiber by oxidation degumming. Text. Res. J.
2015
,85,
2125–2135. [CrossRef]
14.
Abitbol, T.; Rivkin, A.; Cao, Y.; Nevo, Y.; Abraham, E.; Ben-Shalom, T.; Lapidot, S.; Shoseyov, O. Nanocellulose, a tiny fiber with
huge applications. Curr. Opin. Biotechnol. 2016,39, 76–88. [CrossRef] [PubMed]
15.
Klemm, D.; Cranston, E.D.; Fischer, D.; Gama, M.; Kedzior, S.A.; Kralisch, D.; Kramer, F.; Kondo, T.; Lindström, T.; Nietzsche, S.;
et al. Nanocellulose as a natural source for groundbreaking applications in materials science: Today’s state. Mater. Today
2018
,21,
720–748. [CrossRef]
16.
Thompson, L.; Azadmanjiri, J.; Nikzad, M.; Sbarski, I.; Wang, J.; Yu, A. Cellulose nanocrystals: Production, functionalization and
advanced applications. Rev. Adv. Mater. Sci. 2019,58, 1–16. [CrossRef]
17. Dufresne, A. Nanocellulose: A new ageless bionanomaterial. Mater. Today 2013,16, 220–227. [CrossRef]
18.
Chirayil, C.J.; Mathew, L.; Thomas, S. Review of recent research in nano cellulose preparation from different lignocellulosic fibers.
Rev. Adv. Mater. Sci. 2014,37, 20–28.
19.
Ferreira, F.V.; Otoni, C.G.; De France, K.J.; Barud, H.S.; Lona, L.M.F.; Cranston, E.D.; Rojas, O.J. Porous nanocellulose gels and
foams: Breakthrough status in the development of scaffolds for tissue engineering. Mater. Today 2020,37, 126–141. [CrossRef]
20.
Ansari, F.; Ding, Y.; Berglund, L.A.; Dauskardt, R.H. Toward Sustainable Multifunctional Coatings Containing Nanocellulose in a
Hybrid Glass Matrix. ACS Nano 2018,12, 5495–5503. [CrossRef]
21.
Erlandsson, J.; López Durán, V.; Granberg, H.; Sandberg, M.; Larsson, P.A.; Wågberg, L. Macro- and mesoporous nanocellulose
beads for use in energy storage devices. Appl. Mater. Today 2016,5, 246–254. [CrossRef]
22.
Wang, Z.; Lee, Y.H.; Kim, S.W.; Seo, J.Y.; Lee, S.Y.; Nyholm, L. Why Cellulose-Based Electrochemical Energy Storage Devices? Adv.
Mater. 2020,2000892, 1–18. [CrossRef] [PubMed]
23.
Ilyas, R.A.; Sapuan, S.M.; Atikah, M.S.N.; Asyraf, M.R.M.; Rafiqah, S.A.; Aisyah, H.A.; Nurazzi, N.M.; Norrrahim, M.N.F. Effect of
hydrolysis time on the morphological, physical, chemical, and thermal behavior of sugar palm nanocrystalline cellulose (Arenga
pinnata (Wurmb.)Merr). Text. Res. J. 2021,91, 152–167. [CrossRef]
24.
Ilyas, R.A.; Sapuan, S.M.; Ishak, M.R.; Zainudin, E.S. Sugar palm nanofibrillated cellulose (Arenga pinnata (Wurmb.)Merr):
Effect of cycles on their yield, physic-chemical, morphological and thermal behavior. Int. J. Biol. Macromol.
2019
,123, 379–388.
[CrossRef] [PubMed]
25.
Azammi, A.M.N.; Ilyas, R.A.; Sapuan, S.M.; Ibrahim, R.; Atikah, M.S.N.; Asrofi, M.; Atiqah, A. Characterization studies of
biopolymeric matrix and cellulose fibres based composites related to functionalized fibre-matrix interface. In Interfaces in Particle
and Fibre Reinforced Composites; Elsevier: London, UK, 2020; pp. 29–93. ISBN 9780081026656.
26.
Abral, H.; Basri, A.; Muhammad, F.; Fernando, Y.; Hafizulhaq, F.; Mahardika, M.; Sugiarti, E.; Sapuan, S.M.; Ilyas, R.A.; Stephane,
I. A simple method for improving the properties of the sago starch films prepared by using ultrasonication treatment. Food
Hydrocoll. 2019,93, 276–283. [CrossRef]
27.
Abral, H.; Atmajaya, A.; Mahardika, M.; Hafizulhaq, F.; Kadriadi; Handayani, D.; Sapuan, S.M.; Ilyas, R.A. Effect of ultrasonication
duration of polyvinyl alcohol (PVA) gel on characterizations of PVA film. J. Mater. Res. Technol. 2020,9, 2477–2486. [CrossRef]
28.
Shatkin, J.A.; Wegner, T.H.; Bilek, E.M.; Cowie, J. Market projections of cellulose nanomaterial-enabled products—Part 1:
Applications. TAPPI J. 2014,13, 9–16. [CrossRef]
29.
Ventrapragada, L.K.; Creager, S.E.; Rao, A.M.; Podila, R. Carbon Nanotubes Coated Paper as Current Collectors for Secondary
Li-ion Batteries. Nanotechnol. Rev. 2019,8, 18–23. [CrossRef]
30.
Kalwar, K.; Shen, M. Electrospun cellulose acetate nanofibers and Au@AgNPs for antimicrobial activity—A mini review.
Nanotechnol. Rev. 2019,8, 246–257. [CrossRef]
31.
Jiao, Z.; Zhang, B.; Li, C.; Kuang, W.; Zhang, J.; Xiong, Y.; Tan, S.; Cai, X.; Huang, L. Carboxymethyl cellulose-grafted graphene
oxide for efficient antitumor drug delivery. Nanotechnol. Rev. 2018,7, 291–301. [CrossRef]
32.
Zhe, Z.; Yuxiu, A. Nanotechnology for the oil and gas industry—An overview of recent progress. Nanotechnol. Rev.
2018
,7,
341–353. [CrossRef]
Polymers 2021,13, 231 27 of 35
33.
Asyraf, M.R.M.; Ishak, M.R.; Sapuan, S.M.; Yidris, N.; Ilyas, R.A.; Rafidah, M.; Razman, M.R. Potential Application of Green
Composites for Cross Arm Component in Transmission Tower: A Brief Review. Int. J. Polym. Sci.
2020
,2020, 8878300. [CrossRef]
34.
Johari, A.N.; Ishak, M.R.; Leman, Z.; Yusoff, M.Z.M.; Asyraf, M.R.M. Influence of CaCO
3
in pultruded glass fibre/unsaturated
polyester composite on flexural creep behaviour using conventional and TTSP methods. Polimery 2020,65, 46–54. [CrossRef]
35.
Asyraf, M.R.M.; Rafidah, M.; Ishak, M.R.; Sapuan, S.M.; Ilyas, R.A.; Razman, M.R. Integration of TRIZ, Morphological Chart and
ANP method for development of FRP composite portable fire extinguisher. Polym. Compos. 2020,41, 2917–2932. [CrossRef]
36.
Klemm, D.; Kramer, F.; Moritz, S.; Lindström, T.; Ankerfors, M.; Gray, D.; Dorris, A. Nanocelluloses: A new family of nature-based
materials. Angew. Chem. Int. Ed. 2011,50, 5438–5466. [CrossRef] [PubMed]
37.
Ferrer, A.; Pal, L.; Hubbe, M. Nanocellulose in packaging: Advances in barrier layer technologies. Ind. Crops Prod.
2017
,95,
574–582. [CrossRef]
38.
Ilyas, R.A.; Sapuan, S.M.; Norrrahim, M.N.F.; Yasim-Anuar, T.A.T.; Kadier, A.; Kalil, M.S.; Atikah, M.S.N.; Ibrahim, R.; Asrofi, M.;
Abral, H.; et al. Nanocellulose/starch biopolymer nanocomposites: Processing, manufacturing, and applications. In Advanced
Processing, Properties, and Application of Strach and Other Bio-Based Polymer; Al-Oqla, F.M., Ed.; Elsevier Inc.: Amsterdam, The
Netherlands, 2021.
39.
Islam, M.T.; Alam, M.M.; Patrucco, A.; Montarsolo, A.; Zoccola, M. Preparation of nanocellulose: A review. AATCC J. Res.
2014
,1,
17–23. [CrossRef]
40.
Kumar, V.; Bollström, R.; Yang, A.; Chen, Q.; Chen, G.; Salminen, P.; Bousfield, D.; Toivakka, M. Comparison of nano- and
microfibrillated cellulose films. Cellulose 2014,21, 3443–3456. [CrossRef]
41.
Siqueira, G.; Bras, J.; Dufresne, A. Cellulosic Bionanocomposites: A Review of Preparation, Properties and Applications. Polymers
2010,2, 728–765. [CrossRef]
42.
Lavoine, N.; Desloges, I.; Sillard, C.; Bras, J. Controlled release and long-term antibacterial activity of chlorhexidine digluconate
through the nanoporous network of microfibrillated cellulose. Cellulose 2014,21, 4429–4442. [CrossRef]
43.
Moon, R.J.; Martini, A.; Nairn, J.; Simonsen, J.; Youngblood, J. Cellulose nanomaterials review: Structure, properties and
nanocomposites. Chem. Soc. Rev. 2011,40, 3941. [CrossRef] [PubMed]
44.
Brinchi, L.; Cotana, F.; Fortunati, E.; Kenny, J.M. Production of nanocrystalline cellulose from lignocellulosic biomass: Technology
and applications. Carbohydr. Polym. 2013,94, 154–169. [CrossRef] [PubMed]
45.
Santos-Ebinuma, V.C.; Roberto, I.C.; Simas Teixeira, M.F.; Pessoa, A. Improving of red colorants production by a new Penicillium
purpurogenum strain in submerged culture and the effect of different parameters in their stability. Biotechnol. Prog.
2013
,29,
778–785. [CrossRef] [PubMed]
46.
Jozala, A.F.; de Lencastre-Novaes, L.C.; Lopes, A.M.; de Carvalho Santos-Ebinuma, V.; Mazzola, P.G.; Pessoa, A., Jr.; Grotto, D.;
Gerenutti, M.; Chaud, M.V. Bacterial nanocellulose production and application: A 10-year overview. Appl. Microbiol. Biotechnol.
2016,100, 2063–2072. [CrossRef]
47.
Krystynowicz, A.; Czaja, W.; Wiktorowska-Jezierska, A.; Gonçalves-Mi´skiewicz, M.; Turkiewicz, M.; Bielecki, S. Factors affecting
the yield and properties of bacterial cellulose. J. Ind. Microbiol. Biotechnol. 2002,29, 189–195. [CrossRef]
48.
Chawla, P.R.; Bajaj, I.B.; Survase, S.A.; Singhal, R.S. Microbial cellulose: Fermentative production and applications. Food Technol.
Biotechnol. 2009,47, 107–124.
49.
Lee, K.Y.; Buldum, G.; Mantalaris, A.; Bismarck, A. More than meets the eye in bacterial cellulose: Biosynthesis, bioprocessing,
and applications in advanced fiber composites. Macromol. Biosci. 2014,14, 10–32. [CrossRef]
50.
Jeon, S.; Yoo, Y.M.; Park, J.W.; Kim, H.J.; Hyun, J. Electrical conductivity and optical transparency of bacterial cellulose based
composite by static and agitated methods. Curr. Appl. Phys. 2014,14, 1621–1624. [CrossRef]
51.
Tyagi, N.; Suresh, S. Production of cellulose from sugarcane molasses using Gluconacetobacter intermedius SNT-1: Optimization
& characterization. J. Clean. Prod. 2016,112, 71–80. [CrossRef]
52.
Huang, X.; Netravali, A. Biodegradable green composites made using bamboo micro/nano-fibrils and chemically modified soy
protein resin. Compos. Sci. Technol. 2009,69, 1009–1015. [CrossRef]
53.
Mohammed, L.; Ansari, M.N.M.; Pua, G.; Jawaid, M.; Islam, M.S. A Review on Natural Fiber Reinforced Polymer Composite and
Its Applications. Int. J. Polym. Sci. 2015,2015, 243947. [CrossRef]
54.
Nurazzi, N.M.; Khalina, A.; Sapuan, S.M.; Ilyas, R.A. Mechanical properties of sugar palm yarn/woven glass fiber reinforced
unsaturated polyester composites: Effect of fiber loadings and alkaline treatment. Polimery 2019,64, 12–22. [CrossRef]
55.
Mazani, N.; Sapuan, S.M.; Sanyang, M.L.; Atiqah, A.; Ilyas, R.A. Design and fabrication of a shoe shelf from kenaf fiber reinforced
unsaturated polyester composites. In Lignocellulose for Future Bioeconomy; Elsevier: Amsterdam, The Netherlands, 2019; pp.
315–332. ISBN 9780128163542.
56.
Nurazzi, N.M.; Khalina, A.; Sapuan, S.M.; Ilyas, R.A.; Rafiqah, S.A.; Hanafee, Z.M. Thermal properties of treated sugar palm
yarn/glass fiber reinforced unsaturated polyester hybrid composites. J. Mater. Res. Technol. 2020,9, 1606–1618. [CrossRef]
57.
Norizan, M.N.; Abdan, K.; Ilyas, R.A.; Biofibers, S.P. Effect of fiber orientation and fiber loading on the mechanical and thermal
properties of sugar palm yarn fiber reinforced unsaturated polyester resin composites. Polimery 2020,65, 34–43. [CrossRef]
58.
Ayu, R.S.; Khalina, A.; Harmaen, A.S.; Zaman, K.; Isma, T.; Liu, Q.; Ilyas, R.A.; Lee, C.H. Characterization Study of Empty Fruit
Bunch (EFB) Fibers Reinforcement in Poly(Butylene) Succinate (PBS)/Starch/Glycerol Composite Sheet. Polymers
2020
,12, 1571.
[CrossRef] [PubMed]
Polymers 2021,13, 231 28 of 35
59.
Morán, J.I.; Alvarez, V.A.; Cyras, V.P.; Vázquez, A. Extraction of cellulose and preparation of nanocellulose from sisal fibers.
Cellulose 2008,15, 149–159. [CrossRef]
60.
Jumaidin, R.; Khiruddin, M.A.A.; Asyul Sutan Saidi, Z.; Salit, M.S.; Ilyas, R.A. Effect of cogon grass fibre on the thermal,
mechanical and biodegradation properties of thermoplastic cassava starch biocomposite. Int. J. Biol. Macromol.
2020
,146, 746–755.
[CrossRef]
61.
Sari, N.H.; Pruncu, C.I.; Sapuan, S.M.; Ilyas, R.A.; Catur, A.D.; Suteja, S.; Sutaryono, Y.A.; Pullen, G. The effect of water immersion
and fibre content on properties of corn husk fibres reinforced thermoset polyester composite. Polym. Test.
2020
,91, 106751.
[CrossRef]
62.
Jumaidin, R.; Saidi, Z.A.S.; Ilyas, R.A.; Ahmad, M.N.; Wahid, M.K.; Yaakob, M.Y.; Maidin, N.A.; Rahman, M.H.A.; Osman, M.H.
Characteristics of Cogon Grass Fibre Reinforced Thermoplastic Cassava Starch Biocomposite: Water Absorption and Physical
Properties. J. Adv. Res. Fluid Mech. Therm. Sci. 2019,62, 43–52.
63.
Aisyah, H.A.; Paridah, M.T.; Sapuan, S.M.; Khalina, A.; Berkalp, O.B.; Lee, S.H.; Lee, C.H.; Nurazzi, N.M.; Ramli, N.; Wahab, M.S.;
et al. Thermal Properties of Woven Kenaf/Carbon Fibre-Reinforced Epoxy Hybrid Composite Panels. Int. J. Polym. Sci.
2019
,
2019, 5258621. [CrossRef]
64.
Jumaidin, R.; Ilyas, R.A.; Saiful, M.; Hussin, F.; Mastura, M.T. Water Transport and Physical Properties of Sugarcane Bagasse Fibre
Reinforced Thermoplastic Potato Starch Biocomposite. J. Adv. Res. Fluid Mech. Therm. Sci. 2019,61, 273–281.
65.
Trache, D.; Hussin, M.H.; Hui Chuin, C.T.; Sabar, S.; Fazita, M.R.N.; Taiwo, O.F.A.; Hassan, T.M.; Haafiz, M.K.M. Microcrystalline
cellulose: Isolation, characterization and bio-composites application—A review. Int. J. Biol. Macromol.
2016
,93, 789–804.
[CrossRef] [PubMed]
66.
Kargarzadeh, H.; Ioelovich, M.; Ahmad, I.; Thomas, S.; Dufresne, A. Methods for Extraction of Nanocellulose from Various
Sources. Handb. Nanocellulose Cellul. Nanocompos. 2017, 1–51. [CrossRef]
67.
Kian, L.K.; Jawaid, M.; Ariffin, H.; Alothman, O.Y. Isolation and characterization of microcrystalline cellulose from roselle fibers.
Int. J. Biol. Macromol. 2017,103, 931–940. [CrossRef] [PubMed]
68.
Phanthong, P.; Reubroycharoen, P.; Hao, X.; Xu, G.; Abudula, A. Nanocellulose: Extraction and application. Carbon Resour.
Convers. 2018,1, 32–43. [CrossRef]
69.
Haafiz, M.K.M.; Hassan, A.; Zakaria, Z.; Inuwa, I.M. Isolation and characterization of cellulose nanowhiskers from oil palm
biomass microcrystalline cellulose. Carbohydr. Polym. 2014,103, 119–125. [CrossRef] [PubMed]
70.
Sabaruddin, F.A.; Tahir, P.M.; Sapuan, S.M.; Ilyas, R.A.; Lee, S.H.; Abdan, K.; Mazlan, N.; Roseley, A.S.M.; Abdul Khalil, H.P.S.
The Effects of Unbleached and Bleached Nanocellulose on the Thermal and Flammability of Polypropylene-Reinforced Kenaf
Core Hybrid Polymer Bionanocomposites. Polymers 2021,13, 116. [CrossRef]
71.
Saelee, K.; Yingkamhaeng, N.; Nimchua, T.; Sukyai, P. An environmentally friendly xylanase-assisted pretreatment for cellulose
nanofibrils isolation from sugarcane bagasse by high-pressure homogenization. Ind. Crops Prod. 2016,82, 149–160. [CrossRef]
72.
Rebouillat, S.; Pla, F. State of the Art Manufacturing and Engineering of Nanocellulose: A Review of Available Data and Industrial
Applications. J. Biomater. Nanobiotechnol. 2013,04, 165–188. [CrossRef]
73.
Siró, I.; Plackett, D. Microfibrillated cellulose and new nanocomposite materials: A review. Cellulose
2010
,17, 459–494. [CrossRef]
74.
Jonoobi, M.; Harun, J.; Tahir, P.M.; Zaini, L.H.; SaifulAzry, S.; Makinejad, M.D. Characteristics of nanofibers extracted from kenaf
core. BioResources 2010,5, 2556–2566. [CrossRef]
75.
Alemdar, A.; Sain, M. Isolation and characterization of nanofibers from agricultural residues—Wheat straw and soy hulls.
Bioresour. Technol. 2008,99, 1664–1671. [CrossRef] [PubMed]
76.
Deshpande, A.P.; Bhaskar Rao, M.; Lakshmana Rao, C. Extraction of bamboo fibers and their use as reinforcement in polymeric
composites. J. Appl. Polym. Sci. 2000,76, 83–92. [CrossRef]
77.
Kushwaha, P.K.; Kumar, R. The studies on performance of epoxy and polyester-based composites reinforced with bamboo and
glass fibers. J. Reinf. Plast. Compos. 2010,29, 1952–1962. [CrossRef]
78.
Kushwaha, P.K.; Kumar, R. Studies on performance of acrylonitrile-pretreated bamboo-reinforced thermosetting resin composites.
J. Reinf. Plast. Compos. 2010,29, 1347–1352. [CrossRef]
79.
Abral, H.; Ariksa, J.; Mahardika, M.; Handayani, D.; Aminah, I.; Sandrawati, N.; Sapuan, S.M.; Ilyas, R.A. Highly transparent and
antimicrobial PVA based bionanocomposites reinforced by ginger nanofiber. Polym. Test. 2019,81, 106186. [CrossRef]
80.
Hyojin, K.; Kazuya, O.; Toru, F.; Kenichi, T. Influence of fiber extraction and surface modification on mechanical properties of
green composites with bamboo fiber. J. Adhes. Sci. Technol. 2013,27, 1348–1358.
81.
Zakikhani, P.; Zahari, R.; Sultan, M.T.H.; Majid, D.L. Extraction and preparation of bamboo fibre-reinforced composites. Mater.
Des. 2014,63, 820–828. [CrossRef]
82.
Rao, K.M.M.; Rao, K.M. Extraction and tensile properties of natural fibers: Vakka, date and bamboo. Compos. Struct.
2007
,77,
288–295. [CrossRef]
83.
Phong, N.T.; Fujii, T.; Chuong, B.; Okubo, K. Study on How to Effectively Extract Bamboo Fibers from Raw Bamboo and
Wastewater Treatment. J. Mater. Sci. Res. 2011,1, 144–155. [CrossRef]
84.
Sharma, A.; Thakur, M.; Bhattacharya, M.; Mandal, T.; Goswami, S. Commercial application of cellulose nano-composites—A
review. Biotechnol. Rep. 2019,21, e00316. [CrossRef] [PubMed]
85.
Kalia, S.; Boufi, S.; Celli, A.; Kango, S. Nanofibrillated cellulose: Surface modification and potential applications. Colloid Polym.
Sci. 2014,292, 5–31. [CrossRef]
Polymers 2021,13, 231 29 of 35
86.
Ling, S.; Chen, W.; Fan, Y.; Zheng, K.; Jin, K.; Yu, H.; Buehler, M.J.; Kaplan, D.L. Biopolymer nanofibrils: Structure, modeling,
preparation, and applications. Prog. Polym. Sci. 2018,85, 1–56. [CrossRef] [PubMed]
87. Nishiyama, Y. Structure and properties of the cellulose microfibril. J. Wood Sci. 2009,55, 241–249. [CrossRef]
88.
Missoum, K.; Belgacem, M.N.; Bras, J. Nanofibrillated cellulose surface modification: A review. Materials
2013
,6, 1745–1766.
[CrossRef] [PubMed]
89.
Andberg, M.; Penttilä, M.; Saloheimo, M. Swollenin from Trichoderma reesei exhibits hydrolytic activity against cellulosic
substrates with features of both endoglucanases and cellobiohydrolases. Bioresour. Technol. 2015,181, 105–113. [CrossRef]
90.
Long, L.; Tian, D.; Hu, J.; Wang, F.; Saddler, J. A xylanase-aided enzymatic pretreatment facilitates cellulose nanofibrillation.
Bioresour. Technol. 2017,243, 898–904. [CrossRef]
91.
Henriksson, M.; Henriksson, G.; Berglund, L.A.; Lindström, T. An environmentally friendly method for enzyme-assisted
preparation of microfibrillated cellulose (MFC) nanofibers. Eur. Polym. J. 2007,43, 3434–3441. [CrossRef]
92.
Fraschini, C.; Chauve, G.; Bouchard, J. TEMPO-mediated surface oxidation of cellulose nanocrystals (CNCs). Cellulose
2017
,24,
2775–2790. [CrossRef]
93.
Gamelas, J.A.F.; Pedrosa, J.; Lourenço, A.F.; Mutjé, P.; González, I.; Chinga-Carrasco, G.; Singh, G.; Ferreira, P.J.T. On the
morphology of cellulose nanofibrils obtained by TEMPO-mediated oxidation and mechanical treatment. Micron
2015
,72, 28–33.
[CrossRef]
94.
Medronho, B.; Andrade, R.; Vivod, V.; Ostlund, A.; Miguel, M.G.; Lindman, B.; Voncina, B.; Valente, A.J.M. Cyclodextrin-grafted
cellulose: Physico-chemical characterization. Carbohydr. Polym. 2013,93, 324–330. [CrossRef] [PubMed]
95.
Aulin, C.; Shchukarev, A.; Lindqvist, J.; Malmström, E.; Wågberg, L.; Lindström, T. Wetting kinetics of oil mixtures on fluorinated
model cellulose surfaces. J. Colloid Interface Sci. 2008,317, 556–567. [CrossRef] [PubMed]
96.
Kaboorani, A.; Riedl, B. Surface modification of cellulose nanocrystals (CNC) by a cationic surfactant. Ind. Crops Prod.
2015
,65,
45–55. [CrossRef]
97. Eyley, S.; Thielemans, W. Surface modification of cellulose nanocrystals. Nanoscale 2014,6, 7764–7779. [CrossRef]
98.
Larsson, E.; Sanchez, C.C.; Porsch, C.; Karabulut, E.; Wågberg, L.; Carlmark, A. Thermo-responsive nanofibrillated cellulose by
polyelectrolyte adsorption. Eur. Polym. J. 2013,49, 2689–2696. [CrossRef]
99.
Hokkanen, S.; Bhatnagar, A.; Sillanpää, M. A review on modification methods to cellulose-based adsorbents to improve adsorption
capacity. Water Res. 2016,91, 156–173. [CrossRef]
100.
Kalia, S.; Sabaa, M.W. Polysaccharide Based Graft Copolymers; Springer: Berlin, Germany, 2013; Volume 9783642365, ISBN
9783642365669.
101.
Wei, L.; McDonald, A.G.; Stark, N.M. Grafting of Bacterial Polyhydroxybutyrate (PHB) onto Cellulose via In Situ Reactive
Extrusion with Dicumyl Peroxide. Biomacromolecules 2015,16, 1040–1049. [CrossRef]
102.
Tee, Y.B.; Talib, R.A.; Abdan, K.; Chin, N.L.; Basha, R.K.; Md Yunos, K.F. Thermally Grafting Aminosilane onto Kenaf-Derived
Cellulose and Its Influence on the Thermal Properties of Poly(Lactic Acid) Composites. BioResources
2013
,8, 4468–4483. [CrossRef]
103.
Cheng, D.; Wen, Y.; Wang, L.; An, X.; Zhu, X.; Ni, Y. Adsorption of polyethylene glycol (PEG) onto cellulose nano-crystals to
improve its dispersity. Carbohydr. Polym. 2015,123, 157–163. [CrossRef]
104.
Alves Fidelis, M.E.; Pereira, T.V.C.; Gomes, O.D.F.M.; De Andrade Silva, F.; Toledo Filho, R.D. The effect of fiber morphology on
the tensile strength of natural fibers. J. Mater. Res. Technol. 2013,2, 149–157. [CrossRef]
105.
Papageorgiou, D.G.; Kinloch, I.A.; Young, R.J. Mechanical properties of graphene and graphene-based nanocomposites. Prog.
Mater. Sci. 2017,90, 75–127. [CrossRef]
106.
Chuayjuljit, S.; Su-Uthai, S.; Charuchinda, S. Poly(vinyl chloride) film filled with microcrystalline cellulose prepared from cotton
fabric waste: Properties and biodegradability study. Waste Manag. Res. 2010,28, 109–117. [CrossRef] [PubMed]
107.
Sun, X.; Lu, C.; Liu, Y.; Zhang, W.; Zhang, X. Melt-processed poly(vinyl alcohol) composites filled with microcrystalline cellulose
from waste cotton fabrics. Carbohydr. Polym. 2014,101, 642–649. [CrossRef] [PubMed]
108.
Samat, N.; Marini, C.D.; Maritho, M.A.; Sabaruddin, F.A. Tensile and impact properties of polypropylene/microcrystalline
cellulose treated with different coupling agents. Compos. Interfaces 2013,20, 497–506. [CrossRef]
109.
Chen, J.; Long, Z.; Wang, J.; Wu, M.; Wang, F.; Wang, B.; Lv, W. Preparation and properties of microcrystalline cellu-
lose/hydroxypropyl starch composite films. Cellulose 2017,24, 4449–4459. [CrossRef]
110.
Suzuki, K.; Sato, A.; Okumura, H.; Hashimoto, T.; Nakagaito, A.N.; Yano, H. Novel high-strength, micro fibrillated cellulose-
reinforced polypropylene composites using a cationic polymer as compatibilizer. Cellulose 2014,21, 507–518. [CrossRef]
111.
Gabr, M.H.; Phong, N.T.; Abdelkareem, M.A.; Okubo, K.; Uzawa, K.; Kimpara, I.; Fujii, T. Mechanical, thermal, and moisture
absorption properties of nano-clay reinforced nano-cellulose biocomposites. Cellulose 2013,20, 819–826. [CrossRef]
112.
Noorbakhsh-Soltani, S.M.; Zerafat, M.M.; Sabbaghi, S. A comparative study of gelatin and starch-based nano-composite films
modified by nano-cellulose and chitosan for food packaging applications. Carbohydr. Polym. 2018,189, 48–55. [CrossRef]
113.
Ching, Y.C.; Rahman, A.; Ching, K.Y.; Sukiman, N.L.; Chuah, C.H. Preparation and characterization of polyvinyl alcohol-based
composite reinforced with nanocellulose and nanosilica. BioResources 2015,10, 3364–3377. [CrossRef]
114.
Chirayil, C.J.; Mathew, L.; Hassan, P.A.; Mozetic, M.; Thomas, S. Rheological behaviour of nanocellulose reinforced unsaturated
polyester nanocomposites. Int. J. Biol. Macromol. 2014,69, 274–281. [CrossRef]
115.
Dehnad, D.; Emam-Djomeh, Z.; Mirzaei, H.; Jafari, S.M.; Dadashi, S. Optimization of physical and mechanical properties for
chitosan- nanocellulose biocomposites. Carbohydr. Polym. 2014,105, 222–228. [CrossRef] [PubMed]
Polymers 2021,13, 231 30 of 35
116.
Anju, T.R.; Ramamurthy, K.; Dhamodharan, R. Surface modified microcrystalline cellulose from cotton as a potential mineral
admixture in cement mortar composite. Cem. Concr. Compos. 2016,74, 147–153. [CrossRef]
117.
Parveen, S.; Rana, S.; Ferreira, S.; Filho, A.; Fangueiro, R. Ultrasonic dispersion of micro crystalline cellulose for developing
cementitious composites with excellent strength and stiffness. Ind. Crops Prod. 2018,122, 156–165. [CrossRef]
118.
Boran, S.; Dizman Tomak, E.; Dönmez Çavdar, A. Biological properties of nutshell and microcrystalline cellulose (MCC) filled
high density polyethylene composites. Arch. Mater. Sci. Eng. 2016,78, 71–77. [CrossRef]
119.
Erbas Kiziltas, E.; Kiziltas, A.; Bollin, S.C.; Gardner, D.J. Preparation and characterization of transparent PMMA-cellulose-based
nanocomposites. Carbohydr. Polym. 2015,127, 381–389. [CrossRef] [PubMed]
120.
Cao, Y.; Zavaterri, P.; Youngblood, J.; Moon, R.; Weiss, J. The influence of cellulose nanocrystal additions on the performance of
cement paste. Cem. Concr. Compos. 2015,56, 73–83. [CrossRef]
121.
Fu, T.; Montes, F.; Suraneni, P.; Youngblood, J.; Weiss, J. The influence of cellulose nanocrystals on the hydration and flexural
strength of Portland cement pastes. Polymers 2017,9, 424. [CrossRef]
122.
Gabr, M.H.; Phong, N.T.; Okubo, K.; Uzawa, K.; Kimpara, I.; Fujii, T. Thermal and mechanical properties of electrospun
nano-celullose reinforced epoxy nanocomposites. Polym. Test. 2014,37, 51–58. [CrossRef]
123.
Cataldi, A.; Dorigato, A.; Deflorian, F.; Pegoretti, A. Thermo-mechanical properties of innovative microcrystalline cellulose filled
composites for art protection and restoration. J. Mater. Sci. 2014,49, 2035–2044. [CrossRef]
124.
Dos Santos, F.A.; Tavares, M.I.B. Development and characterization of hybrid materials based on biodegradable PLA matrix,
microcrystalline cellulose and organophilic silica. Polimeros 2014,24, 561–566. [CrossRef]
125.
Sabo, R.; Jin, L.; Stark, N.; Ibach, R.E. Effect of environmental conditions on the mechanical properties and fungal degradation of
polycaprolactone/microcrystalline cellulose/wood flour composites. BioResources 2013,8, 3322–3335. [CrossRef]
126.
Lee, J.A.; Yoon, M.J.; Lee, E.S.; Lim, D.Y.; Kim, K.Y. Preparation and characterization of cellulose nanofibers (CNFs) from
microcrystalline cellulose (MCC) and CNF/polyamide 6 composites. Macromol. Res. 2014,22, 738–745. [CrossRef]
127.
Arjmandi, R.; Hassan, A.; Eichhorn, S.J.; Mohamad Haafiz, M.K.; Zakaria, Z.; Tanjung, F.A. Enhanced ductility and tensile
properties of hybrid montmorillonite/cellulose nanowhiskers reinforced polylactic acid nanocomposites. J. Mater. Sci.
2015
,50,
3118–3130. [CrossRef]
128.
Jabbar, A.; Militký, J.; Wiener, J.; Kale, B.M.; Ali, U.; Rwawiire, S. Nanocellulose coated woven jute/green epoxy composites:
Characterization of mechanical and dynamic mechanical behavior. Compos. Struct. 2017,161, 340–349. [CrossRef]
129.
Kiziltas, A.; Gardner, D.J.; Han, Y.; Yang, H.S. Mechanical Properties of Microcrystalline Cellulose (MCC) Filled Engineering
Thermoplastic Composites. J. Polym. Environ. 2014,22, 365–372. [CrossRef]
130.
Parveen, S.; Rana, S.; Fangueiro, R.; Paiva, M.C. A novel approach of developing micro crystalline cellulose reinforced cementitious
composites with enhanced microstructure and mechanical performance. Cem. Concr. Compos. 2017,78, 146–161. [CrossRef]
131.
Dai, X.; Xiong, Z.; Ma, S.; Li, C.; Wang, J.; Na, H.; Zhu, J. Fabricating highly reactive bio-based compatibilizers of epoxidized citric
acid to improve the flexural properties of polylactide/microcrystalline cellulose blends. Ind. Eng. Chem. Res.
2015
,54, 3806–3812.
[CrossRef]
132.
Saba, N.; Mohammad, F.; Pervaiz, M.; Jawaid, M.; Alothman, O.Y.; Sain, M. Mechanical, morphological and structural properties
of cellulose nanofibers reinforced epoxy composites. Int. J. Biol. Macromol. 2017,97, 190–200. [CrossRef]
133.
Zhang, Y.; Wu, J.; Wang, B.; Sui, X.; Zhong, Y.; Zhang, L.; Mao, Z.; Xu, H. Cellulose nanofibril-reinforced biodegradable polymer
composites obtained via a Pickering emulsion approach. Cellulose 2017,24, 3313–3322. [CrossRef]
134.
Chaiwutthinan, P.; Pimpan, V.; Chuayjuljit, S.; Leejarkpai, T. Biodegradable Plastics Prepared from Poly(lactic acid), Poly(butylene
succinate) and Microcrystalline Cellulose Extracted from Waste-Cotton Fabric with a Chain Extender. J. Polym. Environ.
2015
,23,
114–125. [CrossRef]
135.
Soheilmoghaddam, M.; Wahit, M.U.; Ibrahim Akos, N. Regenerated cellulose/epoxidized natural rubber blend film. Mater. Lett.
2013,111, 221–224. [CrossRef]
136.
El-Hadi, A.M. Increase the elongation at break of poly (lactic acid) composites for use in food packaging films. Sci. Rep.
2017
,7,
46767. [CrossRef] [PubMed]
137.
Niazi, M.B.K.; Jahan, Z.; Berg, S.S.; Gregersen, Ø.W. Mechanical, thermal and swelling properties of phosphorylated nanocellulose
fibrils/PVA nanocomposite membranes. Carbohydr. Polym. 2017,177, 258–268. [CrossRef] [PubMed]
138.
Gu, J.; Chen, W.J.; Lin, L.; Luo, Y.F.; Jia, D.M. Effect of nanocrystalline cellulose on the curing characteristics and aging resistance
properties of carbon black reinforced natural rubber. Chin. J. Polym. Sci. Engl. Ed. 2013,31, 1382–1393. [CrossRef]
139.
Dai, L.; Wang, L.Y.; Yuan, T.Q.; He, J. Study on thermal degradation kinetics of cellulose-graft-poly(l-lactic acid) by thermogravi-
metric analysis. Polym. Degrad. Stab. 2014,99, 233–239. [CrossRef]
140.
Fiore, V.; Scalici, T.; Badagliacco, D.; Enea, D.; Alaimo, G.; Valenza, A. Aging resistance of bio-epoxy jute-basalt hybrid composites
as novel multilayer structures for cladding. Compos. Struct. 2017,160, 1319–1328. [CrossRef]
141.
Movilla-Quesada, D.; Raposeiras, A.C.; Muñoz, O.; Castro-Fresno, D. Reduction in the use of mineral aggregate by recycling
cellulose ashes to decrease the aging of hot asphalt mixtures. Constr. Build. Mater. 2017,143, 547–557. [CrossRef]
142.
Abdullah, M.E.; Rosni, N.N.M.; Jaya, R.P.; Yaacob, H.; Hassan, N.A. Agussabti Effect of charcoal ash coconut shell from waste
material at different size on the physical properties of bitumen. Key Eng. Mater. 2017,744, 121–125. [CrossRef]
Polymers 2021,13, 231 31 of 35
143.
Bhasney, S.M.; Bhagabati, P.; Kumar, A.; Katiyar, V. Morphology and crystalline characteristics of polylactic acid [PLA]/linear
low density polyethylene [LLDPE]/microcrystalline cellulose [MCC] fiber composite. Compos. Sci. Technol.
2019
,171, 54–61.
[CrossRef]
144.
Zhuravleva, N.; Reznik, A.; Kiesewetter, D.; Stolpner, A.; Khripunov, A. Possible applications of bacterial cellulose in the
manufacture of electrical insulating paper. J. Phys. Conf. Ser. 2018,1124, 031008. [CrossRef]
145.
Liu, Y.; Zhang, S.; Wang, X.; Pan, Y.; Zhang, F.; Huang, J. Mechanical and aging resistance properties of polypropylene (PP)
reinforced with nanocellulose/attapulgite composites (NCC/AT). Compos. Interfaces 2020,27, 73–85. [CrossRef]
146.
Morán, J.I.; Vázquez, A.; Cyras, V.P. Bio-nanocomposites based on derivatized potato starch and cellulose, preparation and
characterization. J. Mater. Sci. 2013,48, 7196–7203. [CrossRef]
147.
Yang, B.; Zhang, M.; Lu, Z.; Tan, J.; Luo, J.J.; Song, S.; Ding, X.; Wang, L.; Lu, P.; Zhang, Q. Comparative study of aramid nanofiber
(ANF) and cellulose nanofiber (CNF). Carbohydr. Polym. 2019,208, 372–381. [CrossRef] [PubMed]
148.
Balakrishnan, P.; Gopi, S.; Sreekala, M.S.; Thomas, S. UV resistant transparent bionanocomposite films based on potato
starch/cellulose for sustainable packaging. Starch-Stärke 2018,70, 1700139. [CrossRef]
149.
Li, Z.; Liu, J.; Jiang, K.; Thundat, T. Carbonized nanocellulose sustainably boosts the performance of activated carbon in ionic
liquid supercapacitors. Nano Energy 2016,25, 161–169. [CrossRef]
150.
Le Bras, D.; Strømme, M.; Mihranyan, A. Characterization of dielectric properties of nanocellulose from wood and algae for
electrical insulator applications. J. Phys. Chem. B 2015,119, 5911–5917. [CrossRef]
151.
Willgert, M.; Boujemaoui, A.; Malmström, E.; Constable, E.C.; Housecroft, C.E. Copper-based dye-sensitized solar cells with
quasi-solid nano cellulose composite electrolytes. RSC Adv. 2016,6, 56571–56579. [CrossRef]
152.
Zhao, Y.; Dang, W.; Lu, Z.; Deng, J.; Hao, Y.; Su, Z.; Zhang, M. Fabrication of mechanically robust and UV-resistant aramid
fiber-based composite paper by adding nano-TiO2and nanofibrillated cellulose. Cellulose 2018,25, 3913–3925. [CrossRef]
153.
Jang, Y.W.; Min, B.G.; Yoon, K.H. Enhancement in compressive strength and UV ageing-resistance of poly(p-phenylene ben-
zobisoxazole) nanocomposite fiber containing modified polyhedral oligomeric silsesquioxane. Fibers Polym.
2017
,18, 575–581.
[CrossRef]
154.
Chen, L.; Hu, Z.; Wu, Z.; Wu, G.; Ma, L.; Zhang, C.; Huang, Y. POSS-bound ZnO nanowires as interphase for enhancing interfacial
strength and hydrothermal aging resistance of PBO fiber/epoxy resin composites. Compos. Part A Appl. Sci. Manuf.
2017
,96, 1–8.
[CrossRef]
155.
Feng, Y.; Zhang, J.; He, J.; Zhang, J. Transparent cellulose/polyhedral oligomeric silsesquioxane nanocomposites with enhanced
UV-shielding properties. Carbohydr. Polym. 2016,147, 171–177. [CrossRef] [PubMed]
156.
De Amorim, J.D.P.; de Souza, K.C.; Duarte, C.R.; da Silva Duarte, I.; de Assis Sales Ribeiro, F.; Silva, G.S.; de Farias, P.M.A.; Stingl,
A.; Costa, A.F.S.; Vinhas, G.M.; et al. Plant and bacterial nanocellulose: Production, properties and applications in medicine, food,
cosmetics, electronics and engineering. A review. Environ. Chem. Lett. 2020,18, 851–869. [CrossRef]
157.
Pacheco, G.; de Mello, C.V.; Chiari-Andréo, B.G.; Isaac, V.L.B.; Ribeiro, S.J.L.; Pecoraro, É.; Trovatti, E. Bacterial cellulose skin
masks—Properties and sensory tests. J. Cosmet. Dermatol. 2018,17, 840–847. [CrossRef] [PubMed]
158.
Reiniati, I.; Hrymak, A.N.; Margaritis, A. Recent developments in the production and applications of bacterial cellulose fibers and
nanocrystals. Crit. Rev. Biotechnol. 2017,37, 510–524. [CrossRef]
159.
Wei, H.; Rodriguez, K.; Renneckar, S.; Vikesland, P.J. Environmental science and engineering applications of nanocellulose-based
nanocomposites. Environ. Sci. Nano 2014,1, 302–316. [CrossRef]
160.
De Oliveira Barud, H.G.; da Silva, R.R.; da Silva Barud, H.; Tercjak, A.; Gutierrez, J.; Lustri, W.R.; de Oliveira, O.B.; Ribeiro, S.J.L.
A multipurpose natural and renewable polymer in medical applications: Bacterial cellulose. Carbohydr. Polym.
2016
,153, 406–420.
[CrossRef]
161.
Azeredo, H.M.C.; Rosa, M.F.; Mattoso, L.H.C. Nanocellulose in bio-based food packaging applications. Ind. Crops Prod.
2017
,97,
664–671. [CrossRef]
162.
Gama, M.; Dourado, F.; Bielecki, S. Bacterial NanoCellulose: From Biotechnology to Bio-Economy; Elsevier: Amsterdam, The
Netherlands, 2016; ISBN 9780444634580.
163.
Boufi, S.; González, I.; Delgado-Aguilar, M.; Tarrès, Q.; Pèlach, M.À.; Mutjé, P. Nanofibrillated cellulose as an additive in
papermaking process: A review. Carbohydr. Polym. 2016,154, 151–166. [CrossRef]
164.
Grüneberger, F.; Künniger, T.; Zimmermann, T.; Arnold, M. Rheology of nanofibrillated cellulose/acrylate systems for coating
applications. Cellulose 2014,21, 1313–1326. [CrossRef]
165.
Naderi, A.; Lindström, T.; Sundström, J. Carboxymethylated nanofibrillated cellulose: Rheological studies. Cellulose
2014
,21,
1561–1571. [CrossRef]
166.
Missoum, K.; Martoïa, F.; Belgacem, M.N.; Bras, J. Effect of chemically modified nanofibrillated cellulose addition on the
properties of fiber-based materials. Ind. Crops Prod. 2013,48, 98–105. [CrossRef]
167.
Gupta, P.; Singh, B.; Agrawal, A.K.; Maji, P.K. Low density and high strength nanofibrillated cellulose aerogel for thermal
insulation application. Mater. Des. 2018,158, 224–236. [CrossRef]
168.
Wang, X.; Cheng, F.; Liu, J.; Smått, J.H.; Gepperth, D.; Lastusaari, M.; Xu, C.; Hupa, L. Biocomposites of copper-containing
mesoporous bioactive glass and nanofibrillated cellulose: Biocompatibility and angiogenic promotion in chronic wound healing
application. Acta Biomater. 2016,46, 286–298. [CrossRef] [PubMed]
Polymers 2021,13, 231 32 of 35
169.
Brown, E.E.; Hu, D.; Abu Lail, N.; Zhang, X. Potential of nanocrystalline cellulose-fibrin nanocomposites for artificial vascular
graft applications. Biomacromolecules 2013,14, 1063–1071. [CrossRef]
170.
George, J.; Sabapathi, S.N. Cellulose nanocrystals: Synthesis, functional properties, and applications. Nanotechnol. Sci. Appl.
2015
,
8, 45–54. [CrossRef] [PubMed]
171.
Yang, S.; Tang, Y.; Wang, J.; Kong, F.; Zhang, J. Surface Treatment of Cellulosic Paper with Starch-Based Composites Reinforced
with Nanocrystalline Cellulose. Ind. Eng. Chem. Res. 2014,53, 13980–13988. [CrossRef]
172.
Jawaid, M.; Boufi, S.; Abdul Khalil, H.P.S. Cellulose-Reinforced Nanofibre Composites: Production, Properties and Applications;
Woodhead Publishing: Duxford, UK, 2017; ISBN 9780081009659.
173.
Lalia, B.S.; Samad, Y.A.; Hashaikeh, R. Nanocrystalline cellulose-reinforced composite mats for lithium-ion batteries: Electro-
chemical and thermomechanical performance. J. Solid State Electrochem. 2013,17, 575–581. [CrossRef]
174.
Ilyas, R.A.; Sapuan, M.S.; Norizan, M.N.; Norrrahim, M.N.F.; Ibrahim, R.; Atikah, M.S.N.; Huzaifah, M.R.M.; Radzi, A.M.; Izwan,
S.; Azammi, A.M.N.; et al. Macro to nanoscale natural fiber composites for automotive components: Research, development,
and application. In Biocomposite and Synthetic Composites for Automotive Applications; Sapuan, M.S., Ilyas, R.A., Eds.; Woodhead
Publishing: Amsterdam, The Netherlands, 2020.
175.
Ilyas, R.A.; Sapuan, S.M.; Asyraf, M.R.M.; Atikah, M.S.N.; Ibrahim, R.; Dele-Afolabia, T.T. Introduction to biofiller reinforced
degradable polymer composites. In Biofiller Reinforced Biodegradable Polymer Composites; Sapuan, S.M., Jumaidin, R., Hanafi, I.,
Eds.; CRC Press: Boca Raton, FL, USA, 2020.
176.
Lynch, M. Decorative Skin and Hair Cosmetics Containing Microcrystalline Cellulose As Enhancing Agent. U.S. Patent 10/752173,
12 August 2004.
177.
Norrrahim, M.N.F.; Norizan, M.N.; Ilyas, R.A.; Sapuan, M.S.; Jenol, M.A. Performance Evaluation of Cellulose Nanofiber
Reinforced Polypropylene Biocomposites for Automotive Applications. In Biocomposite and Synthetic Composites for Automotive
Applications; Sapuan, M.S., Ilyas, R.A., Eds.; Woodhead Publishing: Amsterdam, The Netherlands, 2020.
178.
Miao, C.; Hamad, W.Y. In-situ polymerized cellulose nanocrystals (CNC)—poly(L-lactide) (PLLA) nanomaterials and applications
in nanocomposite processing. Carbohydr. Polym. 2016,153, 549–558. [CrossRef]
179.
Dufresne, A. Cellulose nanomaterial reinforced polymer nanocomposites. Curr. Opin. Colloid Interface Sci.
2017
,29, 1–8. [CrossRef]
180.
Kuboki, T. Mechanical properties and foaming behavior of injection molded cellulose fiber reinforced polypropylene composite
foams. J. Cell. Plast. 2014,50, 129–143. [CrossRef]
181.
Zhang, W.; He, X.; Li, C.; Zhang, X.; Lu, C.; Zhang, X.; Deng, Y. High performance poly (vinyl alcohol)/cellulose nanocrystals
nanocomposites manufactured by injection molding. Cellulose 2014,21, 485–494. [CrossRef]
182. Shokrieh, M.M. Residual Stresses in Composite Materials; Woodhead Publishing: Duxford, UK, 2014; ISBN 9780857092700.
183.
Heim, H.P.; Tromm, M. General aspects of foam injection molding using local precision mold opening technology. Polymer
2015
,
56, 111–118. [CrossRef]
184.
Wang, L.; Okada, K.; Hikima, Y.; Ohshima, M.; Sekiguchi, T.; Yano, H. Effect of cellulose nanofiber (CNF) surface treatment on
cellular structures and mechanical properties of polypropylene/CNF nanocomposite foams via core-back foam injection molding.
Polymers 2019,11, 249. [CrossRef] [PubMed]
185.
Peng, J.; Walsh, P.J.; Sabo, R.C.; Turng, L.S.; Clemons, C.M. Water-assisted compounding of cellulose nanocrystals into polyamide
6 for use as a nucleating agent for microcellular foaming. Polymer 2016,84, 158–166. [CrossRef]
186.
Borkotoky, S.S.; Dhar, P.; Katiyar, V. Biodegradable poly (lactic acid)/Cellulose nanocrystals (CNCs) composite microcellular
foam: Effect of nanofillers on foam cellular morphology, thermal and wettability behavior. Int. J. Biol. Macromol.
2018
,106,
433–446. [CrossRef]
187.
Wang, L.; Ishihara, S.; Hikima, Y.; Ohshima, M.; Sekiguchi, T.; Sato, A.; Yano, H. Unprecedented Development of Ultrahigh
Expansion Injection-Molded Polypropylene Foams by Introducing Hydrophobic-Modified Cellulose Nanofibers. ACS Appl.
Mater. Interfaces 2017,9, 9250–9254. [CrossRef]
188.
Standau, T. Batch Foaming and Foam Extrusion. University of Bayreuth. 2010, p. 1. Available online: http://www.polymer-
engineering.de/en/research/fields/lightweight-materials/batch-foaming- and-foam-extrusion.html (accessed on 1 August
2020).
189.
Kuboki, T.; Park, C.B. Chapter 13 Foaming Technology of Wood Fiber/Plastic Composites. Mater. Energy
2014
, 197–218. [CrossRef]
190.
Li, Q.; Matuana, L.M. Foam Extrusion of High Density Polyethylene/Wood-Flour Composites Using Chemical Foaming Agents.
J. Appl. Polym. Sci. 2003,88, 3139–3150. [CrossRef]
191.
Yang, Z.; Peng, H.; Wang, W.; Liu, T. Crystallization behavior of poly(
ε
-caprolactone)/layered double hydroxide nanocomposites.
J. Appl. Polym. Sci. 2010,116, 2658–2667. [CrossRef]
192.
Romeo RIM. What Is Structural Foam Molding? Romeorim Reaction Injection Molding. 2020, p. 1. Available online: https:
//romeorim.com/structural-foam-molding/ (accessed on 1 August 2020).
193.
Bledzki, A.K.; Faruk, O. Influence of processing temperature on microcellular injection-moulded wood-polypropylene composites.
Macromol. Mater. Eng. 2006,291, 1226–1232. [CrossRef]
194.
Yoon, J.D.; Kuboki, T.; Jung, P.U.; Wang, J.; Park, C.B. Injection molding of wood-fiber/plastic composite foams. Compos. Interfaces
2009,16, 797–811. [CrossRef]
195.
Ilyas, R.A.; Sapuan, S.M. Biopolymers and Biocomposites: Chemistry and Technology. Curr. Anal. Chem.
2020
,16, 500–503.
[CrossRef]
Polymers 2021,13, 231 33 of 35
196.
Zhu, L.; Zong, L.; Wu, X.; Li, M.; Wang, H.; You, J.; Li, C. Shapeable Fibrous Aerogels of Metal–Organic-Frameworks Templated
with Nanocellulose for Rapid and Large-Capacity Adsorption. ACS Nano 2018,12, 4462–4468. [CrossRef]
197.
Roohani, M.; Habibi, Y.; Belgacem, N.M.; Ebrahim, G.; Karimi, A.N.; Dufresne, A. Cellulose whiskers reinforced polyvinyl alcohol
copolymers nanocomposites. Eur. Polym. J. 2008,44, 2489–2498. [CrossRef]
198.
Cao, X.; Xu, C.; Wang, Y.; Liu, Y.; Liu, Y.; Chen, Y. New nanocomposite materials reinforced with cellulose nanocrystals in nitrile
rubber. Polym. Test. 2013,32, 819–826. [CrossRef]
199.
Mahendra, I.P.; Wirjosentono, B.; Tamrin, T.; Ismail, H.; Mendez, J.A.; Causin, V. The effect of nanocrystalline cellulose and
TEMPO-oxidized nanocellulose on the compatibility of polypropylene/cyclic natural rubber blends. J. Thermoplast. Compos.
Mater. 2020. [CrossRef]
200.
Dai, H.; Zhang, H.; Ma, L.; Zhou, H.; Yu, Y.; Guo, T.; Zhang, Y.; Huang, H. Green pH/magnetic sensitive hydrogels based on
pineapple peel cellulose and polyvinyl alcohol: Synthesis, characterization and naringin prolonged release. Carbohydr. Polym.
2019,209, 51–61. [CrossRef]
201.
Jain, M.; Pradhan, M.K. Morphology and mechanical properties of sisal fiber and nano cellulose green rubber composite: A
comparative study. Int. J. Plast. Technol. 2016,20, 378–400. [CrossRef]
202.
Karmaker, N.; Islam, F.; Islam, M.N.; Razzak, M.; Koly, F.A.; Chowdhury, A.M.S.; Khan, R.A. Fabrication and Characterization of
PVA-Gelatin-Nano Crystalline Cellulose based Biodegradable Film: Effect of Gamma Radiation. J. Res. Updates Polym. Sci.
2019
,
8, 7–14. [CrossRef]
203.
Xu, S.; Girouard, N.; Schueneman, G.; Shofner, M.L.; Meredith, J.C. Mechanical and thermal properties of waterborne epoxy
composites containing cellulose nanocrystals. Polymer 2013,54, 6589–6598. [CrossRef]
204.
Zhang, K.; Lu, Y.; Hao, N.; Nie, S. Enhanced thermal conductivity of cellulose nanofibril/aluminum nitride hybrid films by
surface modification of aluminum nitride. Cellulose 2019,26, 8669–8683. [CrossRef]
205.
Pandurangan, M.T.; Kanny, K. Study of curing characteristics of cellulose nanofiber-filled epoxy nanocomposites. Catalysts
2020
,
10, 831. [CrossRef]
206.
Khan, A.; Asiri, A.M.; Jawaid, M.; Saba, N. Inamuddin Effect of cellulose nano fibers and nano clays on the mechanical,
morphological, thermal and dynamic mechanical performance of kenaf/epoxy composites. Carbohydr. Polym.
2020
,239, 116248.
[CrossRef] [PubMed]
207.
Nair, S.S.; Kuo, P.-Y.; Chen, H.; Yan, N. Investigating the effect of lignin on the mechanical, thermal, and barrier properties of
cellulose nanofibril reinforced epoxy composite. Ind. Crops Prod. 2017,100, 208–217. [CrossRef]
208.
Vu, C.M.; Sinh, L.H.; Choi, H.J.; Pham, T.D. Effect of micro/nano white bamboo fibrils on physical characteristics of epoxy resin
reinforced composites. Cellulose 2017,24, 5475–5486. [CrossRef]
209.
Guimarães Junior, M.; Teixeira, F.G.; Tonoli, G.H.D. Effect of the nano-fibrillation of bamboo pulp on the thermal, structural,
mechanical and physical properties of nanocomposites based on starch/poly(vinyl alcohol) blend. Cellulose 2018. [CrossRef]
210.
Ilyas, R.A.; Sapuan, S.M. The Preparation Methods and Processing of Natural Fibre Bio-polymer Composites. Curr. Org. Synth.
2020,16, 1068–1070. [CrossRef]
211.
Nazrin, A.; Sapuan, S.M.; Zuhri, M.Y.M.; Ilyas, R.A.; Syafiq, R.; Sherwani, S.F.K. Nanocellulose Reinforced Thermoplastic Starch
(TPS), Polylactic Acid (PLA), and Polybutylene Succinate (PBS) for Food Packaging Applications. Front. Chem.
2020
,8, 213.
[CrossRef]
212.
Takkalkar, P.; Griffin, G.; Kao, N. Enhanced Mechanical and Barrier Performance of Poly (Lactic Acid) Based Nanocomposites
Using Surface Acetylated Starch Nanocrystals. J. Polym. Environ. 2019,27, 2078–2088. [CrossRef]
213.
Blaker, J.J.; Lee, K.-Y.; Walters, M.; Drouet, M.; Bismarck, A. Aligned unidirectional PLA/bacterial cellulose nanocomposite fibre
reinforced PDLLA composites. React. Funct. Polym. 2014,85, 185–192. [CrossRef]
214.
Dhar, P.; Tarafder, D.; Kumar, A.; Katiyar, V. Thermally recyclable polylactic acid/cellulose nanocrystal films through reactive
extrusion process. Polymer 2016,87, 268–282. [CrossRef]
215.
Sung, S.H.; Chang, Y.; Han, J. Development of polylactic acid nanocomposite films reinforced with cellulose nanocrystals derived
from coffee silverskin. Carbohydr. Polym. 2017,169, 495–503. [CrossRef] [PubMed]
216.
Kale, R.D.; Gorade, V.G.; Madye, N.; Chaudhary, B.; Bangde, P.S.; Dandekar, P.P. Preparation and characterization of biocomposite
packaging film from poly(lactic acid) and acylated microcrystalline cellulose using rice bran oil. Int. J. Biol. Macromol.
2018
,118,
1090–1102. [CrossRef] [PubMed]
217.
Zhang, P.; Gao, D.; Zou, P.; Wang, B. Preparation and thermomechanical properties of nanocrystalline cellulose reinforced
poly(lactic acid) nanocomposites. J. Appl. Polym. Sci. 2017,134. [CrossRef]
218.
Rahmat, M.; Karrabi, M.; Ghasemi, I.; Zandi, M.; Azizi, H. Silane crosslinking of electrospun poly (lactic acid)/nanocrystalline
cellulose bionanocomposite. Mater. Sci. Eng. C 2016,68, 397–405. [CrossRef]
219.
Mukherjee, T.; Tobin, M.J.; Puskar, L.; Sani, M.-A.; Kao, N.; Gupta, R.K.; Pannirselvam, M.; Quazi, N.; Bhattacharya, S. Chemically
imaging the interaction of acetylated nanocrystalline cellulose (NCC) with a polylactic acid (PLA) polymer matrix. Cellulose
2017
,
24, 1717–1729. [CrossRef]
220.
Mármol, G.; Gauss, C.; Fangueiro, R. Potential of Cellulose Microfibers for PHA and PLA Biopolymers Reinforcement. Molecules
2020,25, 4653. [CrossRef]
Polymers 2021,13, 231 34 of 35
221.
Hazwan Hussin, M. Physicochemical Studies of Kenaf Nanocrystaline Cellulose and Poly (3-hydroxybutyrate-co-3-
hydroxyhexanoate) as Filler for Lithium Perchlorate based Polymer Electrolyte. Int. J. Electrochem. Sci.
2019
,14, 1620–1633.
[CrossRef]
222.
Zhang, B.; Huang, C.; Zhao, H.; Wang, J.; Yin, C.; Zhang, L.; Zhao, Y. Eects of cellulose nanocrystals and cellulose nanofibers on
the structure and properties of polyhydroxybutyrate nanocomposites. Polymers 2019,11, 2063. [CrossRef]
223.
Zheng, T.; Zhang, Z.; Shukla, S.; Agnihotri, S.; Clemons, C.M.; Pilla, S. PHBV-graft-GMA via reactive extrusion and its use in
PHBV/nanocellulose crystal composites. Carbohydr. Polym. 2019,205, 27–34. [CrossRef]
224.
Xu, J.; Manepalli, P.H.; Zhu, L.; Narayan-Sarathy, S.; Alavi, S. Morphological, barrier and mechanical properties of films from
poly (butylene succinate) reinforced with nanocrystalline cellulose and chitin whiskers using melt extrusion. J. Polym. Res.
2019
,
26, 188. [CrossRef]
225.
Zhang, X.; Zhang, Y. Reinforcement effect of poly(butylene succinate) (PBS)-grafted cellulose nanocrystal on toughened
PBS/polylactic acid blends. Carbohydr. Polym. 2016,140, 374–382. [CrossRef] [PubMed]
226.
Vaezi, K.; Asadpour, G.; Sharifi, S.H. Effect of coating with novel bio nanocomposites of cationic starch/cellulose nanocrystals on
the fundamental properties of the packaging paper. Polym. Test. 2019,80, 106080. [CrossRef]
227.
Jiang, B.; Li, S.; Wu, Y.; Song, J.; Chen, S.; Li, X.; Sun, H. Preparation and characterization of natural corn starch-based composite
films reinforced by eggshell powder. CYTA J. Food 2018,16, 1045–1054. [CrossRef]
228.
Al Fath, M.T.; Nasution, H.; Harahap, H.; Ayu, G.E. Biocomposite of pectin and starch filled with nanocrystalline cellulose (NCC):
The effect of filler loading and glycerol addition. AIP Conf. Proc. 2019,2175, 020012. [CrossRef]
229.
Nasution, H.; Harahap, H.; Al Fath, M.T.; Afandy, Y. Physical properties of sago starch biocomposite filled with Nanocrystalline
Cellulose (NCC) from rattan biomass: The effect of filler loading and co-plasticizer addition. IOP Conf. Ser. Mater. Sci. Eng.
2018
,
309, 012033. [CrossRef]
230.
Syafri, E.; Sudirman; Mashadi; Yulianti, E.; Deswita; Asrofi, M.; Abral, H.; Sapuan, S.M.; Ilyas, R.A.; Fudholi, A. Effect of sonication
time on the thermal stability, moisture absorption, and biodegradation of water hyacinth (Eichhornia crassipes) nanocellulose-filled
bengkuang (Pachyrhizus erosus) starch biocomposites. J. Mater. Res. Technol. 2019,8, 6223–6231. [CrossRef]
231.
Ilyas, R.A.; Sapuan, S.M.; Ishak, M.R.; Zainudin, E.S. Sugar palm nanocrystalline cellulose reinforced sugar palm starch composite:
Degradation and water-barrier properties. IOP Conf. Ser. Mater. Sci. Eng. 2018,368, 012006. [CrossRef]
232.
Ilyas, R.A.; Sapuan, S.M.; Ishak, M.R.; Zainudin, E.S. Development and characterization of sugar palm nanocrystalline cellulose
reinforced sugar palm starch bionanocomposites. Carbohydr. Polym. 2018,202, 186–202. [CrossRef]
233.
Zainuddin, S.Y.Z.; Ahmad, I.; Kargarzadeh, H. Cassava starch biocomposites reinforced with cellulose nanocrystals from kenaf
fibers. Compos. Interfaces 2013,20, 189–199. [CrossRef]
234.
Barkoula, N.M.; Alcock, B.; Cabrera, N.O.; Peijs, T. Flame-Retardancy Properties of Intumescent Ammonium Poly(Phosphate)
and Mineral Filler Magnesium Hydroxide in Combination with Graphene. Polym. Polym. Compos.
2008
,16, 101–113. [CrossRef]
235.
Zarina, S.; Ahmad, I. Biodegradable composite films based on
κ
-carrageenan reinforced by cellulose nanocrystal from kenaf
fibers. BioResources 2015,10, 256–271. [CrossRef]
236.
Slavutsky, A.M.; Bertuzzi, M.A. Water barrier properties of starch films reinforced with cellulose nanocrystals obtained from
sugarcane bagasse. Carbohydr. Polym. 2014,110, 53–61. [CrossRef] [PubMed]
237.
Chen, D.; Lawton, D.; Thompson, M.R.; Liu, Q. Biocomposites reinforced with cellulose nanocrystals derived from potato peel
waste. Carbohydr. Polym. 2012,90, 709–716. [CrossRef] [PubMed]
238.
Asrofi, M.; Sujito; Syafri, E.; Sapuan, S.M.; Ilyas, R.A. Improvement of Biocomposite Properties Based Tapioca Starch and
Sugarcane Bagasse Cellulose Nanofibers. Key Eng. Mater. 2020,849, 96–101. [CrossRef]
239.
Ahsan, Q.; Carron, T.S.S.; Mustafa, Z. On the use of nano fibrillated kenaf cellulose fiber as reinforcement in polylactic acid
biocomposites. J. Mech. Eng. Sci. 2019,13, 4970–4988. [CrossRef]
240.
Farid, T.; Herrera, V.N.; Kristiina, O. Investigation of crystalline structure of plasticized poly (lactic acid)/Banana nanofibers
composites. IOP Conf. Ser. Mater. Sci. Eng. 2018,369, 012031. [CrossRef]
241.
Hervy, M.; Blaker, J.J.; Braz, A.L.; Lee, K.-Y. Mechanical response of multi-layer bacterial cellulose nanopaper reinforced
polylactide laminated composites. Compos. Part A Appl. Sci. Manuf. 2018,107, 155–163. [CrossRef]
242.
Ghasemi, S.; Behrooz, R.; Ghasemi, I.; Yassar, R.S.; Long, F. Development of nanocellulose-reinforced PLA nanocomposite by
using maleated PLA (PLA-g-MA). J. Thermoplast. Compos. Mater. 2018,31, 1090–1101. [CrossRef]
243.
Jonoobi, M.; Harun, J.; Mathew, A.P.; Oksman, K. Mechanical properties of cellulose nanofiber (CNF) reinforced polylactic acid
(PLA) prepared by twin screw extrusion. Compos. Sci. Technol. 2010,70, 1742–1747. [CrossRef]
244.
Cherpinski, A.; Torres-Giner, S.; Vartiainen, J.; Peresin, M.S.; Lahtinen, P.; Lagaron, J.M. Improving the water resistance of
nanocellulose-based films with polyhydroxyalkanoates processed by the electrospinning coating technique. Cellulose
2018
,25,
1291–1307. [CrossRef]
245.
Widiyanti, P.; Amali, M.A. Aminatun Poly(ethylene glycol)dimethacrylate—Nanofibrillated cellulose bionanocomposites as
injectable hydrogel for therapy of herniated nucleus pulposus patients. J. Mater. Res. Technol. 2020,9, 12716–12722. [CrossRef]
246.
Srithep, Y.; Ellingham, T.; Peng, J.; Sabo, R.; Clemons, C.; Turng, L.S.; Pilla, S. Melt compounding of poly (3-hydroxybutyrate-co-3-
hydroxyvalerate)/nanofibrillated cellulose nanocomposites. Polym. Degrad. Stab. 2013,98, 1439–1449. [CrossRef]
247.
Yeng, L.C. Characterization and Properties of Sodium Hydroxide/Urea Regenerated Cellulose Reinforced Poly(3 Hydroxybutyrate) Biocom-
posites; Universiti Teknologi Malysia: Skudai, Malaysia, 2016.
Polymers 2021,13, 231 35 of 35
248.
Wang, Y.; Uetani, K.; Liu, S.; Zhang, X.; Wang, Y.; Lu, P.; Wei, T.; Fan, Z.; Shen, J.; Yu, H.; et al. Multifunctional Bionanocomposite
Foams with a Chitosan Matrix Reinforced by Nanofibrillated Cellulose. ChemNanoMat 2017,3, 98–108. [CrossRef]
249.
Doench, I.; Tran, T.; David, L.; Montembault, A.; Viguier, E.; Gorzelanny, C.; Sudre, G.; Cachon, T.; Louback-Mohamed,
M.; Horbelt, N.; et al. Cellulose Nanofiber-Reinforced Chitosan Hydrogel Composites for Intervertebral Disc Tissue Repair.
Biomimetics 2019,4, 19. [CrossRef]
250.
LomelíRamírez, M.G.; Satyanarayana, K.G.; Iwakiri, S.; De Muniz, G.B.; Tanobe, V.; Flores-Sahagun, T.S. Study of the properties
of biocomposites. Part I. Cassava starch-green coir fibers from Brazil. Carbohydr. Polym. 2011,86, 1712–1722. [CrossRef]
251.
Pitiphatharaworachot, S.; Chitbanyong, K.; Sungkaew, S.; Pisutpiched, S.; Khantayanuwong, S.; Puangsin, B. Starch nanocompos-
ites reinforced with TEMPO-oxidized cellulose nanofibrils derived from bamboo holocellulose. BioResources
2019
,14, 2784–2797.
[CrossRef]
252.
Huang, L.; Zhao, H.; Yi, T.; Qi, M.; Xu, H.; Mo, Q.; Huang, C.; Wang, S.; Liu, Y. Reparation and properties of cassava residue
cellulose nanofibril/cassava starch composite films. Nanomaterials 2020,10, 755. [CrossRef]
253.
De Almeida, V.S.; Barretti, B.R.V.; Ito, V.C.; Malucelli, L.; da Silva Carvalho Filho, M.A.; Demiate, I.M.; Pinheiro, L.A.; Lacerda,
L.G. Thermal, Morphological, and Mechanical Properties of Regular and Waxy Maize Starch Films Reinforced with Cellulose
Nanofibers (CNF). Mater. Res. 2020,23. [CrossRef]
254.
Balakrishnan, P.; Sreekala, M.S.; Kunaver, M.; Huski´c, M.; Thomas, S. Morphology, transport characteristics and viscoelastic
polymer chain confinement in nanocomposites based on thermoplastic potato starch and cellulose nanofibers from pineapple leaf.
Carbohydr. Polym. 2017,169, 176–188. [CrossRef]
255.
Tajik, M.; Torshizi, H.J.; Resalati, H.; Hamzeh, Y. Effects of cationic starch in the presence of cellulose nanofibrils on structural,
optical and strength properties of paper from soda bagasse pulp. Carbohydr. Polym. 2018,194, 1–8. [CrossRef] [PubMed]
256.
Soni, R.; Asoh, T.A.; Uyama, H. Cellulose nanofiber reinforced starch membrane with high mechanical strength and durability in
water. Carbohydr. Polym. 2020,238, 116203. [CrossRef] [PubMed]
257.
Babaee, M.; Jonoobi, M.; Hamzeh, Y.; Ashori, A. Biodegradability and mechanical properties of reinforced starch nanocomposites
using cellulose nanofibers. Carbohydr. Polym. 2015,132, 1–8. [CrossRef] [PubMed]
258.
Karimi, S.; Tahir, P.; Dufresne, A.; Karimi, A.; Abdulkhani, A. A comparative study on characteristics of nanocellulose reinforced
thermoplastic starch biofilms prepared with different techniques. Nord. Pulp Pap. Res. J. 2014,29, 41–45. [CrossRef]
259.
Llanos, J.H.R.; Tadini, C.C. Preparation and characterization of bio-nanocomposite films based on cassava starch or chitosan,
reinforced with montmorillonite or bamboo nanofibers. Int. J. Biol. Macromol. 2018,107, 371–382. [CrossRef] [PubMed]
260.
Nasri-Nasrabadi, B.; Behzad, T.; Bagheri, R. Preparation and characterization of cellulose nanofiber reinforced thermoplastic
starch composites. Fibers Polym. 2014,15, 347–354. [CrossRef]
261.
Ilyas, R.A.; Sapuan, S.M.; Ishak, M.R.; Zainudin, E.S. Water transport properties of bio-nanocomposites reinforced by sugar palm
(arenga pinnata) nanofibrillated cellulose. J. Adv. Res. Fluid Mech. Therm. Sci. 2018,51, 234–246.
262.
Ilyas, R.A.; Sapuan, S.M.; Ibrahim, R.; Abral, H.; Ishak, M.R.; Zainudin, E.S.; Atikah, M.S.N.; Mohd Nurazzi, N.; Atiqah, A.;
Ansari, M.N.M.; et al. Effect of sugar palm nanofibrillated celluloseconcentrations on morphological, mechanical andphysical
properties of biodegradable films basedon agro-waste sugar palm (Arenga pinnata(Wurmb.)Merr) starch. J. Mater. Res. Technol.
2019,8, 4819–4830. [CrossRef]
263.
Atikah, M.S.N.; Ilyas, R.A.; Sapuan, S.M.; Ishak, M.R.; Zainudin, E.S.; Ibrahim, R.; Atiqah, A.; Ansari, M.N.M.; Jumaidin, R.
Degradation and physical properties of sugar palm starch/sugar palm nanofibrillated cellulose bionanocomposite. Polimery
2019
,
64, 27–36. [CrossRef]
264.
Rozilah, A.; Jaafar, C.N.A.; Sapuan, S.M.; Zainol, I.; Ilyas, R.A. The Effects of Silver Nanoparticles Compositions on the
Mechanical, Physiochemical, Antibacterial, and Morphology Properties of Sugar Palm Starch Biocomposites for Antibacterial
Coating. Polymers 2020,12, 2605. [CrossRef]
... Mechanical fibrillation of cellulose aided by different pretreatments produces promising materials such as cellulose microfibrils (CMFs) and nanofibrils (CNFs) (Khalil et al., 2014;Wang et al., 2021). However, despite the numerous applications of CNFs and CMFs across various sectors, such as papermaking , biomedicine (Jorfi & Foster, 2015;Seddiqi et al., 2021), water and wastewater treatments (Sayyed et al., 2021), energy storage (Ani, Nzereogu, Agbogu, Ezema, & Nwanya, 2022), composites (Omran et al., 2021) or food industry (Ahankari, Subhedar, Bhadauria, & Dufresne, 2021;Zhang, Shen, Li, Yang & Lin, 2022), among others, several challenges limit their industrial-scale production. Three main challenges need to be addressed include: i) reduction of costs through the search for more efficient and faster production techniques; ii) upscaling the production process; and iii) reduction of the environmental impact Djafari Petroudy et al., 2021). ...
... While significant advancements have been made in the development of nano cellulose-based functional nanocomposites, ongoing research aims to further optimize dispersion techniques, tailor surface modifications, and explore novel polymer/nano cellulose combinations. These efforts will likely lead to even more diverse and advanced applications in fields such as smart materials, 3D printing, and multifunctional coatings [69]. The incorporation of nature-inspired nano cellulose into polymer composites for mechanical, thermal, and electrical enhancements holds immense promise for a wide range of applications. ...
Chapter
Full-text available
A promising ecofriendly, cost-effective biomaterial derived from natural sources, i.e., lignocellulose natural fibers from trees, plants, agri waste, fruits, vegetables, seeds, and leaves. It clicked the attention of the researchers due to promising properties and versatility. The aim of the study is to examine the recent developments and preparation methods and extraction techniques for nanolignocellulose materials from nature. It also discusses the wide range of applications that use nano cellulose’s remarkable properties for a variety of research fields. Current work discussed advancements in nano cellulose preparation techniques, innovative methods, and cutting-edge applications. The ease of nano cellulose excels as a material for tissue engineering scaffolds, wound dressings, flexible and sustainable electronics, and drug delivery systems in the biomedical industry. It is desirable component in composites due to its remarkable reinforcing abilities in polymers. The use of environmental applications such as water purification, oil spill cleanup, and biodegradable packaging is also highlighted in this research. The studies emphasize the need for more investigation and optimization of extraction processes, characterization, and applications. This multidisciplinary study intends to motivate academics and scientists to fully utilize nano cellulose and contribute to the creation of environmentally friendly and sustainable solutions across a range of industries.
... CNFs are typically produced by mechanical defibrillation of cellulose fibers and sometimes complemented by optional chemical or enzymatic pretreatments [2]. Compared to cellulose fibers, CNFs exhibit a significantly increased surface area, leading to a higher concentration of hydroxyl groups on their surface, resulting in a highly hydrophilic material [3,4]. While the hydrophilic nature of CNFs proves advantageous for various applications, including hydrogels [5], scaffolds [6] and coatings [7], it imposes limitations when applications involve hydrophobic polymers or nonpolar environments due to their chemical incompatibility [8,9]. ...
Article
Full-text available
Cellulose nanofibrils are distinguished bionanomaterials known for their unique morphology, thermal stability, and ability to form networks, yet they encounter challenges in compatibility with hydrophobic matrices, limiting their application in various applications. This study introduces an innovative surface modification method to address the high hydrophilicity of CNFs. The novelty lies in the use of lipase as a biocatalyst in combination with renewable grafting agents, specifically butanoic and oleic acids. The lipase successfully esterified both acids onto the CNFs, with butanoic acid exhibiting a higher surface concentration, resulting in a more substantial reduction in hydrophilicity. Contact angle measurements demonstrated a notable shift, from 10.84° for untreated CNF to 68.4° and 55.1° for CNFs grafted with butanoic and oleic acid residues, respectively. While there were only slight alterations in crystallinity, thermal stability, and brittleness, lipase proved to be an effective catalyst for modifying the CNF surface with fatty acids. This approach offers a method to mitigate the high hydrophilicity of CNFs without compromising their key properties. Furthermore, it can be proposed as a means to tailor CNF for water-resistant applications in fields such as electronics, packaging, and Pickering emulsions. Graphical Abstract The structure of the lipase protein was sourced from the Protein Data Bank (PDB), first referenced by Xie et al. [38]
... Studies of polymeric composites reinforced with nanocellulose have been on the rise since the beginning of the 21st century. This growing interest is driven not only by the economic and environmental advantages offered by these materials, but also by the mechanical properties and biodegradability they provide to the polymeric matrix (Omran et al., 2021). Composites with nanocellulose can be applied in several fields, such as the pharmaceutical, cosmetics, packaging, biomedical, and automobile industries, among others Bhardwwaj, 2019;Dufresne, 2017). ...
Article
Nanocellulose-reinforced polymeric composites can be produced in situ, where polymerization occurs in the presence of reinforcing fillers. Moreover, it can also be produced ex situ, when, the polymeric matrix is mixed with a nanocellulose suspension. This paper provides a review covering technical information, including physical, chemical, and procedural characteristics of the polymerization process. It explores the use of three different in situ techniques: suspension, semi-suspension, and emulsion in the presence of nanocellulose suspensions to produce nanocellulose-reinforced composites. The polymerization techniques addressed in this work, when in the presence of nanocellulose, generally yield biodegradable materials from a renewable source and with wide applicability.
... As well as this, stirred cultures are more likely to cause mutations in microorganisms, which may have a negative impact on the production of BNC. 69 However, two primary issues with static culture methods are high costs and slow output rates. That is why use of a stirred culture has been proposed as a solution to these issues. ...
Article
Full-text available
Cellulose is the most abundant renewable polymer on Earth which is extensively distributed in diverse ecosystems. It is present in higher plants, marine organisms, and also produced through microbial processes in organisms like algae, fungi, and bacteria. From an industrial perspective, the semicrystalline nature of cellulose present in different plant and microbial sources enables the fabrication of various types of nanocellulose, such as nanofibre and nanocrystals, through mechanical disintegration and chemical methods, respectively. Nanocellulose distinguishes itself as a sustainable, nonharmful, and biodegradable polymer. It will enable sustainable development for responsible consumption and production. Possessing a range of excellent properties, it can be seamlessly integrated into various materials. Research on nanocellulose is gaining momentum in response to current issues related to fossil fuels, including concerns about CO 2 emissions, plastic pollution, and the need for renewable energy sources. This review addresses nanocrystals production method from cellulose found in agricultural, microbial sources, and its applications in fields such as materials science, electronics, medicine, and environmental science.
Article
Full-text available
Nanocellulose contains a large number of hydroxyl groups that can be used to modify its surface due to its structure. Owing to its appealing features, such as high strength, great stiffness, and high surface area, nanocellulose is currently gaining popularity in research and industry. The extraction of nanocellulose from the leftover bagasse fiber from sugarcane production by alkaline and acid treatment was successful in this study, with a production yield of 55.6%. The FTIR and XPS results demonstrated a difference in the functional and chemical composition of untreated sugarcane bagasse and extracted nanocellulose. SEM imaging was used to examined the size of the nanocellulose with ImageJ software v1.8.0. TGA, DTG, and XRD analyses were also performed to demonstrate the successful extraction of nanocellulose in terms of its morphology, thermal stability, and crystal structure before and after extraction. The anti-S. aureus activity of the extracted nanocellulose was discovered by using an OD600 test and a colony counting method, and an inhibitory rate of 53.12% was achieved. According to the results, nanocellulose produced from residual sugarcane bagasse could be employed as an antibacterial agent.
Article
Full-text available
In the era of smart and sustainable technology driven by naturally occurring materials, various nanocellulose-based materials play a crucial role. Shape memory behaviour and self-healing capabilities of nanocelluloses are emerging as focal points in numerous research domains. Nanocellulose and its derivatives such as cellulose nanocrystals (CNC) and cellulose nanofibers (CNF), are currently in the limelight due to their excellent shape-memory and self-healing properties, making them suitable for multifunctional devices. In this regard, CNF, as a cutting-edge material, has spurred researchers to explore its potential in developing contemporary multifunctional and personalized health devices. Therefore, a timely and comprehensive review is essential to gain deep insights into the effectiveness of shape-memory and self-healing capabilities of CNF for multifunctional devices. Herein, we first provide a succinct introduction to all nanocellulose materials. This review also depicts recent advancements and breakthroughs in the large and effective synthesis of CNF-based hybrid materials. Next, focusing on their self-healing and shape-memory performance, this review sheds new light on the advanced applications of CNF materials. Finally, perspectives on the current challenges and opportunities in this field are summarized for future researchers to gain an in-depth understanding of "CNF-based smart and sustainable materials."
Article
Tissue engineering has emerged as a remarkable field aiming to restore or replace damaged tissues through the use of biomimetic constructs. Among the diverse materials investigated for this purpose, nanocellulose-based...
Article
Micro‐sized luffa natural fibers (MLNFs) were derived from raw luffa natural fibers from Vietnam by undergoing treatment with sodium hydroxide. This study examined the impact of sodium hydroxide concentration, temperature, and treatment duration on the fracture toughness of epoxy resin 828 reinforced with MLNFs. The results showed that the fracture toughness of the composite, as measured by the critical‐stress‐intensity factor ( K IC ) and Izod impact strength, was improved after the treatment of the MLNFs with sodium hydroxide. Specifically, with the optimal sodium hydroxide treatment parameters of 6% NaOH at 70 °C for 6 h, the critical‐stress‐intensity factor increased by 93% (from 2.0 to 3.86 MPa m 1/2 ) and the Izod impact strength rose by 44.6% (from 3.81 to 5.51 kJ/m ² ). The thermal properties and surface structure of MLNFs treated with NaOH were analyzed using thermogravimetric analysis (TGA) and scanning electron microscopy (SEM) images.
Article
Full-text available
The thermal, thermo-mechanical and flammability properties of kenaf core hybrid polymer nanocomposites reinforced with unbleached and bleached nanocrystalline cellulose (NCC) were studied. The studied chemical composition found that unbleached NCC (NCC-UB) had 90% more lignin content compared to bleached NCC (NCC-B). Nanocelluloses were incorporated within polypropylene (PP) as the matrix, together with kenaf core as a main reinforcement and maleic anhydride grafted polypropylene (MAPP) as a coupling agent via a melt mixing compounding process. The result showed that the thermal stability of the nanocomposites was generally affected by the presence of lignin in NCC-UB and sulfate group on the surface of NCC-B. The residual lignin in NCC-UB appeared to overcome the poor thermal stability of the composites that was caused by sulfation during the hydrolysis process. The lignin helped to promote the late degradation of the nanocomposites, with the melting temperature occurring at a relatively higher temperature of 219.1 °C for PP/NCC-UB, compared to 185.9 °C for PP/NCC-B. Between the two types of nanocomposites, PP/NCC-B had notably lower thermo-mechanical properties, which can be attributed to the poor bonding and dispersion properties of the NCC-B in the nanocomposites blend. The PP/NCC-UB showed better thermal properties due to the effect of residual lignin, which acted as a compatibilizer between NCC-UB and polymer matrix, thus improved the bonding properties. The residual lignin in PP/NCC-UB helped to promote char formation and slowed down the burning process, thus increasing the flame resistance of the nanocomposites. Overall, the residual lignin on the surface of NCC-UB appeared to aid better stability on the thermal and flammability properties of the nanocomposites.
Article
Full-text available
The highest disease caused by low back pain is herniated nucleus pulposus (HNP) where nucleus pulposus (NP) protrudes out through the annulus fibrosis. Injectable hydrogel is a method to restore NP function with minimal invasive surgery. Variation concentration of NFC were studied in detail toward affects on PEGDMA (poly(ethylene glycol)dimethacrylate):NFC hydrogel mechanically and injectability. The characteristics of PEGDMA and nanofibrillated cellulose (NFC) based injectable hydrogels through the photoinitiator Irgacure 2959 in the formation of hydrogels with variations composition 1:0; 1:0.25; 1:0.5; 1:0.75. In the morphological test on NFC a fibre with a diameter of 100 nm was formed from the results of electrospinning on ethyl cellulose (EC). The FTIR showed of photopolymerization of the conjugation bond between Irgacure 2959 and PEGDMA producing C-C at a frequency of 1349.20 cm⁻¹. Swelling test used phosphate buffer saline (PBS) fulfilled the standard at a variation of 1:0.25; 1:0.5; and 1:0.75, 87.07%, 84.84% and 80.42% respectively. Viscosity test in the injectability of hydrogels showed a result of 73.67 dPa.s. The compressive strength test obtained good results in samples 1:0.75 showed a result of 3.42 kPa. In vitro injection model showed increasing of NFC concentration indicates the gel state when released from agarose will be less rupture. PEGDMA/NFC hydrogel biocomposite has the potential to be applied as an injectable hydrogel for the treatment of HNP patients based on the results.
Article
Full-text available
Recently, advanced technologies exploit materials from nonrenewable resources such as petroleum, natural gas, metal ores, and minerals. Since the depletion of these resources and environmental issues, it has brought attention to researchers to progress in the development of biodegradable materials from green composites. Most biofibres and biopolymers are obtained from agricultural waste products either from stem, leaf, stalk, or fruit. Nowadays, green composites with well-regulated life span have been widely discussed in numerous fields and applications. Some studies have shown that biofibres and biopolymers have comparable mechanical, thermal, and physical properties with glass fibre and other synthetic polymers. Thus, researchers are progressively narrowing down the development of green composite materials in many high strength applications, such as house deck and automotive components. This review focuses on the background of green composites (natural fibres and biopolymers), the manufacturing processes, potential applications in cross arm structures, and testing evaluations. This article also focuses on the specific current cross arm configurations and the pultrusion process to form squared hollow section beams. Many open issues and ideas for potential applications of green composites are analysed, and further emphases are given on the development of environmentally friendly material structures. Hence, the article is expected to deliver a state-of-art review on manufacturability and perspectives of natural fibre reinforced biopolymer composite cross arms for transmission towers. © 2020 M. R. M. Asyraf et al. This is an open access article distributed under the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
Article
Full-text available
Antibacterial sugar palm starch biopolymer composite films were developed and derived from renewable sources and inorganic silver nanoparticles (AgNPs) as main ingredients for antibacterial coatings. The composite films were produced by solution casting method and the mechanical and physicochemical properties were determined by tensile test, Fourier Transform Infrared (FTIR) analysis, thermal gravimetric analysis (TGA), antibacterial screening test and field emission scanning electron microscopy (FESEM) images. It was found that mechanical and antibacterial properties of biocomposite films were improved after the addition of AgNPs compared with the film without active metals. The weakness of neat biocomposite films was improved by incorporating inorganic AgNPs as a nanofiller in the films' matrix to avoid bacterial growth. The results showed that the tensile strength ranged between 8 kPa and 408 kPa and the elasticity modulus was between 5.72 kPa and 9.86 kPa. The addition of AgNPs in FTIR analysis decreased the transmittance value, caused small changes in the chemical structure, caused small differences in the intensity peaks, and produced longer wavelengths. These active films increased the degradation weight and decomposition temperature due to the more heat-stable AgNPs. Meanwhile, the average inhibited areas measured were between 7.66 and 7.83 mm (Escherichia coli), 7.5 and 8.0 mm (Salmonella cholerasuis), and 0.1 and 0.5 mm for Staphylococcus aureus. From the microscopic analysis, it was observed that the average size of all microbes for 1 wt% and 4 wt% AgNPs ranged from 0.57 to 2.90 mm. Overall, 3 wt% AgNP nanofiller was found to be the best composition that fulfilled all the mechanical properties and had better antimicrobial properties. Thus, the development of an organic-inorganic hybrid of antibacterial biopolymer composite films is suitable for antibacterial coatings.
Article
Bamboo holocellulose was oxidized by 2,2,6,6-tetramethylpiperidine-1-oxyl radical (TEMPO) in water at pH 10 with an addition of 7.5 mmol g-1 of NaClO for 2 h. After oxidation, the weight recovery ratio of TEMPO-oxidized bamboo holocellulose (TOBH) was 73% with a carboxylate content of 0.65 mmol g-1. The water-insoluble TOBH was converted to an aqueous dispersion of bamboo TEMPO-oxidized cellulose nanofibrils (TOBCN) through mechanical defibrillation and centrifugation, with the nanofibrillation yield of around 90%. Transmission electron microscopy (TEM) indicated that the widths of TOBCN were estimated to be between 3 and 4 nm and lengths of several micrometers. The TOBCN dispersions had high light-transparencies of 99% at a wavelength of 600 nm. These nanofibrils were used as reinforcing nanoparticles in thermoplastic starch (TPS) films. The TOBCN dispersions were added in dosages of 0, 0.3, 0.6, 0.9, 1.2, and 1.5 wt%. Glycerol was used as a plasticizer. Nanocomposite films were prepared through a solution casting process. The TOBCN/TPS nanocomposite films exhibited high optical transparencies and their tensile strength, Young’s modulus, elongation around failure, and work to cause a fracture increased, compared to the TPS films without TOBCN. The presence of TOBCNs, at 1.5 wt%, improved moisture resistance.
Chapter
Polypropylene (PP) is the most frequently used plastic in automotive manufacturing. Being a thermoplastic polymer, it can easily be molded into several shapes. Besides that, it has excellent heat and chemical resistance. The current research interest is to improve the PP properties to meet the automotive market demand. Cellulose nanofiber (CNF) has been widely studied as a reinforcement material in biocomposites. Recently, CNF has emerged as a subject of tremendous scientific interest and has attracted a great deal of attention across disciplines. In this study, biocomposites were prepared by the addition of 1–5 wt% CNF into PP using an internal blending technique. There are many factors that influence the properties of biocomposites such as concentration and the effect of drying on CNF. A homogenous dispersion of CNF in the PP matrix was obtained from biocomposites containing 1%–3% of CNF, which increased their tensile and flexural properties. Meanwhile, aggregations of CNF in the PP matrix occurred at a high amount of CNF loading, and freeze-dried prepared biocomposites resulted in the decrease of their mechanical properties. All in all, it is shown that the drying process of CNF can be eliminated in the production of biocomposites. Therefore, this proved that the biocomposites of PP/CNF are useful as automotive compartments with improved mechanical properties over neat PP.
Chapter
Currently, high-strength fibers such as kevlar, carbon, spectra, glass fiber, carbon nanotube and zylon reinforced composite materials are commonly used in advanced transport structures. However, the world is changing and green material is at the forefront due to the depletion of inorganic materials such as petroleum and other mineral sources. Natural fibers can be classified as green materials, that is, defined as renewable materials. This chapter describes the recent advancements in biocomposites, specifically natural fiber polymer composites (NFPC). This is a material with significant mechanical and barrier properties that is also economical, biodegradable, renewable, and has a low density character. The above properties make NFPC promising for various advanced applications in such fields as automotive parts, papermaking, flexible optoelectronics, scaffolds, optical devices, pharmaceutical products, substitutes/medical biomaterials, spacecraft, aircraft, and tissue repair (wound dressing). In this chapter, the automotive applications of the natural fiber reinforced polymer composites are summarized. The present review covers reinforcement of macro- to nanoscale natural fibers with polymer composites for automotive applications.
Article
The effect of nanocrystalline cellulose (NCC) and nanofiber cellulose (NFC) was estimated as a means of reinforcing and compatibilizing agent of polypropylene/ cyclic natural rubber (PP/CNR) blend in terms of mechanical and thermal properties. The morphological effect of NCC and NFC on the PP/CNR blend property was determined through several characterization techniques, i.e. SEM, contact and TGA/DTG. Scanning electron microscopy (SEM) images revealed that the addition of NCC and NFC became more homogenous than without the addition of nanocellulose. The improvement of nanocomposites was also observed on the result of interphase surface tension and thermal stability. This improvement was assumed as the result of physically/ chemically interaction of nanocellulose with the backbone of PP and CNR, in which the nanocellulose can be imagined acting as a bridge to link the PP and CNR’s backbone.