Trache2016
Trache2016
Trache2016
PII: S0141-8130(16)31016-9
DOI: http://dx.doi.org/doi:10.1016/j.ijbiomac.2016.09.056
Reference: BIOMAC 6523
Please cite this article as: Djalal Trache, M.Hazwan Hussin, Caryn Tan Hui Chuin,
Sumiyyah Sabar, M.R.Nurul Fazita, Owolabi F.A.Taiwo, T.M.Hassan, M.K.Mohamad
Haafiz, Microcrystalline cellulose: isolation, characterization and bio-composites
application − A review, International Journal of Biological Macromolecules
http://dx.doi.org/10.1016/j.ijbiomac.2016.09.056
This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Microcrystalline cellulose: isolation, characterization and bio-composites
application – A review
Djalal Trachea,*, M. Hazwan Hussinb, Caryn Tan Hui Chuinb, Sumiyyah Sabarc, M.R. Nurul Fazitad,
Owolabi F.A. Taiwod,e, T.M. Hassanc, M.K. Mohamad Haafizd
a
UER Chimie Appliquée, Ecole Militaire Polytechnique, BP 17, Bordj El-Bahri, Algiers, Algeria.
b
Lignocellulosic Research Group, School of Chemical Sciences, Universiti Sains Malaysia, 11800 Minden,
Penang, Malaysia.
c
School of Distance Education, Universiti Sains Malaysia, 11800 Minden, Penang, Malaysia.
d
School of Industrial Technology, Universiti Sains Malaysia, 11800 Minden, Penang, Malaysia.
e
Federal Institute of Industrial Research, Oshodi, Nigeria.
*
Corresponding author. Address: UER Chimie Appliquée, Ecole Militaire Polytechnique, P.O Box 17, Bordj El-
Bahri, Algiers, Algeria. Tel: +213 661808275, Fax: +213 21863204
E-mail Address: [email protected]
1
Graphical abstract
2
Highlights
This work provides an up-to date review on methods for isolation and characterization
of microcrystalline cellulose (MCC).
Interest for MCC is exponentially increasing.
New sources, new isolation processes, new treatments induce new types of MCC
materials for industrialization.
Opportunities of MCC-based composite polymers for the future are presented and
discussed.
Abstract:
Considering its widespread usage in various fields, such as food, pharmaceutical, medical,
materials. Although it still suffers from some drawbacks, MCC has recently gained more
mechanical properties, high surface area and biocompatibility. New sources, new isolation
processes, and new treatments are currently under development to satisfy the increasing
this review assembles the current knowledge on the isolation of MCC from different sources
3
1. Introduction
Different kinds of materials play a prominent role in the modern civilization. Among
others, polymer-based materials are usually used in various industries related to chemical
storage in batteries and supercapacitors [1-4]. These polymeric materials have superseded and
substituted the utilization of the traditional ones such as ceramics, glasses and metals in
several fields due to their low cost, availability, low weight, chemical inertness, strength, and
ease of processing [5, 6]. These polymers allow the broad development of various types of
materials such as plastics, elastomers, artificial fibers, and composites. However, the
depletion of fossil resources coupled with the increase in environmental awareness and health
concerns are acting synergistically in providing the impetus for emerging products and
ecology, green chemistry and engineering are being introduced into the development of next
During the past two decades, more attention was being given in the usage of bio-based
materials that can prevent the widespread dependence on fossil fuels [1, 4, 9-13].
of some new alternatives for sustainability by exploring the widely available natural
resources. Thus, there exists a growing interest in bio-composites research to reduce the
composites are non-biodegradable. It often takes many years, even centuries, for nature to
degrade these materials. Bio-based polymer composites are of particular interest. The
potential of their uses in several applications. Recently, the tailoring of physicochemical and
4
mechanical properties of such composites has been proven to be dependent on their intrinsic
characteristics, the spatial arrangement of phases and interface behavior [1, 6, 9, 14-16]. They
can be constituted by biopolymer matrix embedded with fibers whose origins can be found
either as natural minerals or as renewable plant raw materials. Advantages of natural fibers
over artificial ones like aramid, carbon, and glass, are as follows: renewability,
Mankind has employed cellulose for thousands of years as an important material for
hosing and clothing. In contrast with the uses of natural cellulose, the application of this
biopolymer as a chemical feedstock started just 170 years ago with the discovery of the first
cellulose derivative by Braconnot [4, 17], but subsequently developed into a production
volume of more than 5 million tons per year during this century . As reported by Berglund
[18], Boldizar et al. were the first to show that cellulose could be used as a reinforcement for a
cellulose [7, 20-24]. MCC has been widely used especially in cosmetic, food, suspension
other polymeric composites [20, 21, 23, 25-30]. Ecofriendly bio-composites have the potential
to be new products of the current century and partial remediation of several environmental
developed to meet the diverse needs for eco-friendly and economical commodity products
with high performance, durability and fully sustainable technology. The next generation of
bio-composites requires the manufacture of products and materials that far exceed current
materials must possess advanced performances, lower cost, reliable and adaptable properties.
5
Due to their eco-friendly benefits, MCC products are now easily found throughout the
life cycle (Fig. 1). The isolation of MCC fibers and their application in composite materials
have drawn much attention due to their high mechanical strength and stiffness combined with
scientific reports published in the last two decades. This is due to a number of unique
biocompatibility, etc. Fig.2 reveals that such investigations on MCC are on the rise with an
the number of articles is still modest with respect to publications dealing with other organic
and inorganic fillers. The limited application of MCC fibers to date may be partly due to their
separation processes from natural sources and the inherent incompatibility and dispersibity
between the hydrophilic and highly polar cellulose fibers and the hydrophobic and non-polar
Compared to other natural polysaccharide polymers, MCC also suffers some drawbacks
that are desired for some applications, such as, poor wettability, moisture absorption,
incompatibility with most polymeric matrices and limitation in processing temperature. Thus,
various solutions have been suggested in literature in overcoming these disadvantages [20, 21,
25-27, 29, 31]. However, no comprehensive review is available yet regarding the different
methods used to isolate MCC particles and their application in bio-composites. This present
review article revises and discusses the different procedures employed to extract MCC from
different sources, as well as the production of MCC-based composites and their current
applications.
6
2. Cellulose
one of the most important organic compounds produced in the biosphere. It is biosynthesized
by a number of living organisms ranging from lower to higher plants, sea animals, bacteria
and fungi [3, 4, 14, 32-35]. It is regarded as an inexhaustible source of feedstock for the
increasing demand for biocompatible and environmental friendly materials and products,
since it is regularly regenerated by nature in relatively short periods of time where the annual
fabrication is estimated to be over 7.5 x 1010 tons [14, 36, 37]. A number of reviews and
books have been already been published which reports the state of knowledge of this
fascinating polymer. Therefore only some important details are presented to avoid duplication
and the cellulose present in nature is graphically illustrated in Fig. 3 [1, 4, 14, 33, 35-39].
Cellulose can be effectively obtained from a top-down approach, in which wood, cotton,
annual plant or other agricultural residue can be used to produce a desired size of cellulose, or
from a bottom-up approach, where cellulose is biosynthesized from glucose using bacteria
(eg. Acetobacter Xylinam). The amounts and the properties of cellulose depend on the
extraction process, the origin and the lifetime of the natural source. Commonly, these sources
are composed of cellulose, hemicellulose, lignin, extractives and trace elements. In their cell
walls, the spirally oriented cellulose plays the role of reinforcements in the soft hemicellulose
and lignin matrix. An effective removal process of hemicellulose, lignin and other impurities
chiral, renewable and biodegradable. It is insoluble in water and in most organic solvents,
This biopolymer has gained a remarkable place in the annals of polymers. As early as
1839, the French Chemist Anselme Payen coined the name “cellulose” for a white powder
that he isolated from plant tissue [36, 37, 40, 41]. The chemical structure of cellulose was
7
determined later, in 1920, by Hermann Staudinger [40, 42]. Nowadays, its outstanding
hierarchical structure no longer holds any secret; cellulose is a linear carbohydrate polymer
size in terms of the number of anhydroglucose units per cellulose chain. In nature, the DP of
cellulose chains ranges from 100 for MCC to 15,000 for native cotton in relation to their
Each monomeric unit in the cellulose chain contains three carbons which are attached to
the hydroxyl groups: a primary C-6 atom in the methylol group (–CH2OH) and two secondary
carbons (C-2 and C-3) of the anhydroglucose unit [43]. The ability of these OH groups to
form hydrogen bonds plays an important role in controlling the physicochemical properties
and also directing the ultrastructure of cellulose. The cellulosic chains (20–300) are bundled
together to generate micro-fibrils, which are jointly grouped to form cellulose fibers. Each
polymeric chain is asymmetric, possessing two different end-units: a reducing and a non-
reducing end. The reducing end has carbonyl functionality, whereas the non-reducing end
features a supplementary OH group in position C-4. Hydrogen bonds within a cellulose chain
hinder the free rotation of the rings along their linked-glycosidic bonds, resulting in the
stiffening of the chain. However, the interchain hydrogen bonds and the van der Waals
interactions lead to the ordered (crystalline) or the disordered regions (amorphous) of the
cellulose structure. This biopolymer exists in different polymorphs, differing in unit cell
dimensions, in geometry, in chain orientation and polarity [35, 38, 44]. In nature, cellulose is
(triclinic structure) and Iβ (monoclinic structure). It can be transformed irreversibly into the
8
treatment with concentrated sodium hydroxide solution) or precipitated (regenerated) from
solution. Four other crystal polymorphs of cellulose (IIII, IIIII, IVI and IV), in addition to
cellulose I and cellulose II, can be produced by different treatments. Within the framework of
this review, only native cellulose I is considered. This semicrystalline fibrillar structure is the
main source of MCC and is responsible for the mechanical properties due to its high modulus
and crystallinity.
excessive amount of mineral acids. The microfibrils which make up the alpha cellulose
consist of paracrystalline and crystalline phases at nanometer range. The paracrystalline area
is an amorphous mass of cellulose chains whereas the crystalline regions comprise tight
bundles of microcrystals in a rigid linear arrangement. The crystalline regions are called
cellulose crystallites and they are formed by cellulose chains due to van der Waals
interactions and hydrogen bonding. The diameter of these crystallites is of the same order as
the diameter of the cellulose microfibrils [45]. The amorphous phase is readily hydrolyzed
when subjected to acid hydrolysis. This results in shorter and more crystalline fragments, i.e.
MCC. Consequently, the degree of polymerization of the cellulose chain is reduced with
Interestingly, the discovery of MCC commenced from a failed experiment [46]. The
particles in water. They thought the sharp blades of the Waring Blender would sliver off very
small fragments of the agglomerated microcrystals in the hydrolyzed cellulose, and the
obtained microcrystalline fragments would settle out of water. Instead, they obtained a stable
colloidal suspension, which is known commercially as Avicel [5]. MCC was first
9
commercialized in 1962 under the name Avicel®, which is marketed by FMC Corporation.
Since then, an exponential number of researchers have focused their work on such material.
Industrial scale MCC is manufactured through hydrolysis of cotton and wood cellulose using
dilute mineral acids. MCC is characterized by a high degree of crystallinity, and the values
MCC can be dried to a pure, fine-particle form for powdered grade or co-processed with
a water-soluble polymer to deliver a colloidal form (Fig. 5). It possesses the merits of
cellulose. It has a high potential to be used in several areas such as pharmaceutical, cosmetic,
food and polymer composites industries. In the powder form, it is employed, for instance as a
binder and filler in food, medical tablets, and particularly as reinforcement agent in the
stabilizer, a water retainer, a viscosity regulator, and emulsifier in different pastes and creams.
Features like lightness, stiffness, strength, fibrous nature, non-toxicity, water insolubility,
MCC is currently produced from a number of different cellulosic sources. Wood and
cotton are obviously the main industrial sources of cellulosic fibers, and are thus the most
important feedstock employed in the manufacture of MCC [45, 48, 51-53]. However,
competition among many areas such as furniture, pulp and paper industries, building products,
as well as the combustion of wood for energy and the employment of cotton for textile
industry, renders them challenging to provide all sectors with the required quantities of wood
and cotton at a suitable price. Besides that, wood and cotton are not available in many regions,
thus tuning its options to non-woody cellulose. Therefore, interest in other sources such as
herbaceous plants, grass, aquatic plants, agricultural crops and their by-products has
10
extensively stimulated considerable interest. These non-woody plants usually comprise less
lignin than wood. Therefore, bleaching methods are less chemical and energy demanding.
The extraction of MCC from different sources is relevant since it dictates the overall
performance of MCC. It can be produced from any material that is high in cellulose. In
literature, we find that diverse sources are already being used to produce MCC. For example,
it can be extracted from groundnut shell, cereal straw [54, 55], bagasse and corn cob [56],
bamboo [57], sugar beet pulp [58], luffa cylindrical [53], orange mesocarps [59], jute [60],
rice and bean hulls [61], hemp stalks and rice husks [62], newsprint [52], oil palm biomass
[24], fodder grass [63], wastepaper [64], filter paper [65] and cotton waste [66]. Esparto grass
[67] and soybean hulls [47] have also been investigated as potential sources of MCC.
Bacterial cellulose (e.g. Gluconacetobacter xylinus) has been used as the starting material for
research on MCC production as well [68, 69]. Microcrystalline cellulose obtained from
methods and conditions commonly differ in molecular weight, particle size, crystallinity,
The amount of cellulose in different natural sources can vary depending on the species
and age of the plant. In nature, lignocellulosic is a bio-composite which results from a
Table 1 shows the chemical composition of some common natural sources. From
optimize the pretreatment process required to isolate a pure cellulose pulp. Indeed, lignin is
However, there are various methods for the extraction of cellulose from plants using
chemical, physical, biological and combined processes [32, 70, 71]. These processes have
11
often been employed as a pretreatment to facilitate the hydrolysis procedure for the
production of MCC.
neutralized aqueous slurry obtained from the hydrolysis of cellulose procedure [72]. The main
commercial grades are achieved by controlling the drying steps to well manage the particle
size distribution, the moisture content, binding ability, etc. Other drying methods may be
applied such as freeze-drying, fluidized bed, microwaves and classical oven drying followed
Fig. 5 displays the different steps which are usually followed to produce MCC from its
feedstock using different procedures. For lignocellulosic materials, it involves firstly the
removal of lignin, hemicellulose, etc. and isolation of cellulosic fibers. Secondly, a controlled
hydrolysis treatment (generally acid hydrolysis) to remove the paracrystalline regions of the
cellulose polymer; and finally, post-treatments were performed to recover the final MCC
product. However, some authors had published methods using only one main step as it will be
characteristics of MCC exhibit variations according to the origin of the raw material and the
extraction process. Currently, MCC is a material available in the market with more than 60
years of history and the price of MCC (≈$4/kg), which is comparable to or less than some
other engineering should be fillers (glass or aramid) [74]. Over half a century later, MCC is
produced worldwide by more than 10 companies [48]. Several methods such as acid and
alkali hydrolysis, steam explosion, extrusion and radiation-enzymatic have been employed to
isolate highly purified MCC from plant cell walls. This section, therefore, describes in detail
the production of MCC processes based on the aforementioned techniques, including their
12
advantages and disadvantages, as well as some important issues regarding the methods used
Regarding the isolation of crystalline cellulosic region, in the form of MCC, a common
process widely used is based on acid hydrolysis because it requires shorter reaction time than
other processes. MCC can be produced by a continuous process rather than a batch-type
process and it uses limited amount of acid and produces small particles. Ranby [75],
considered as the pioneer in the cellulose hydrolysis process, produced stable suspensions of
colloidal-sized cellulose crystals by sulfuric acid hydrolysis of wood and cotton cellulose in
the early 1950s. The preparation of cellulose in microcrystalline form was first described in
several papers by Batistta and Smith of the American Viscose Corporation [46]. They had
isolated MCC fibers by controlled acid hydrolysis of cellulosic fibers to a level-off degree of
polymerization (LODP). The principle of their method was to pass dilute cellulosic pulp-
paracrystalline domains located at the surface and along their main axis. Upon contact with
acidic solutions, the paracrystalline or amorphous regions are preferentially cleaved; whereas
the crystalline domains that have a higher resistance to acid attack remain essentially intact.
Fig. 6 illustrates the preferential action of the amorphous regions of cellulose microfibrils,
resulting in MCC fibers. It is worth mentioning that the diameter of these MCC fibers (which
physicochemical, thermal and mechanical properties. The temperature and time of hydrolysis
13
procedure, nature and concentration of acid as well as the fiber-to-acid ratio play the
important roles in the particle size, morphology, crystallinity, thermal stability and
mechanical properties of MCC [61, 76, 77]. The acid hydrolysis processes with HCl and
H2SO4 from various cellulosic sources to produce MCC are tabulated in Table 2.
El-Sakhawy and Hassan isolated MCC from agricultural residues using HCl or H2SO4
and explored the effect of acid on the properties of the produced MCC [76]. They
demonstrated that the kind of acid used had no effect on both the crystallinity and crystallite
size. The dynamic laser light scattering particle size analyzer revealed that the nature of the
acid can affect the particle size of MCC depending on the natural sources used. Once the
thermal stability. In another work, based on the hydrolysis method with 1–2.5 M HCl at 105
°C, Leppänen et al. produced MCC from cotton, flux and various wood species (hardwood
and softwood) using different pulping processes (kraft and sulfite) [45]. They indicated that
the MCC obtained had very similar nanostructures. The method of producing MCC from
bacterial cellulose was reported by De Oliveira et al. [69]. They used sulfuric acid at a reflux
system under constant stirring. The obtained degree of polymerization (DP) of MCC was 205
and the particles size was between 70–90 μm. Haafiz et al. produced MCC from oil palm
empty fruit bunch [24]. They isolated MCC using an acid hydrolysis method based on original
morphological structure and a rough surface with 87% crystallinity. Trache et al. achieved the
desired MCC fibers by applying an HCl hydrolysis method using esparto grass fibers as raw
material [67]. The resultant MCC was transformed into fine powder by grinding. The authors
described that the DP of MCC was 318 and the crystallinity index was 81%.
Besides simple hydrolysis methods, other procedures were also explored to produce
MCC. Schaible and Sherwood reported the preparation of MCC from unbleached kraft pulp
14
using one-step hydrolysis and bleaching process using a sufficient amount of active oxygen in
an acidic environment [78]. In their method, the pulp of southern pine was hydrolyzed
respectively with 2 N HCl and [2 N H2SO4 + 0.2 M ozone] at boiling temperature for 60
minutes; the product was filtered out and washed with hot water, then freeze-dried. The
authors mentioned that this method allows the preparation of MCC from a variety of pulp
grades and it can eliminate the need to perform multiple steps with respect to hydrolysis and
bleaching to achieve a satisfactory product. Such a method would ideally perform these two
On the other hand, the isolation of MCC from different natural sources using the acid
and process MCC into functional and high-value added products. However, some drawbacks
such as the cost, amount of reagents, corrosivity hazards, and effluent treatment required to be
lignocellulosic materials [1, 5, 79]. This treatment is applied to disrupt the lignin structure and
to allow the separation of the structural linkages between lignin and carbohydrates. Therefore,
in order to overcome some problems caused by the use of acid hydrolysis process, some
authors utilized the alkali treatment to produce MCC. Trusovs mixed a cellulose source
material with an alkaline solution [79]. Once the material is completely swollen, hydrogen
peroxide was added to the mixing suspension to depolymerize the cellulose thereby reducing
viscosity. Finally, the solution was filtered, neutralized, washed and dried. The author
demonstrated that this process presented several advantages. This method does not involve
manner utilizing readily available chemicals and cellulose source materials. Few years later,
15
Nguyen reported the use of this alkali process jointly with an acidic treatment [80]. He
disclosed that this method employs less alkali and acid than previous processes and it is a
that the employment of this process needs to be carefully controlled to avoid undesirable
The steam explosion treatment process has been and has always been widely
investigated as a promising mechanical pulping method [70, 81, 82]. It offers many attractive
impact, fewer hazardous process chemicals and conditions and more potential for energy
efficiency, etc. This process was invented by Mason in 1927 as a process to produce fibers for
board production. It was used for the first time by DeLong to produce MCC [83]. In his
experiment, he treated the lignocellulose raw material in steam in two steps, with the steam-
treated lignocellulose from the first step being extracted to recover a low degree of
polymerization cellulose which is then impregnated with strong mineral acid to obtain MCC
during the second steam treatment. Ha and Landi patented a very interesting process using an
only one-step steam explosion treatment process without the need for conventional acid
hydrolysis as a supplement to steam treatment [82]. In this case, MCC was obtained by a
process of introducing a cellulose source material into a pressurized reactor; subjecting the
cellulosic material to a steam explosion treatment under controlled conditions. Some authors
reported that the steam explosion allows the breakdown of lignocellulosic material
components by steam heating, shearing forces owing to the expansion of the moisture and
hydrolysis of glycosidic bonds by an organic acid formed during the process [81].
2.1.2.4. Extrusion
16
Extrusion technology is a short-duration, high-temperature hydrolysis method with the
advantages of high flexibility and nonexistence of effluents [47]. Very few investigations
have been reported concerning the production of MCC by extrusion methods. Extrusion
mechanical and chemical treatments for the cellulosic feedstock [84]. Therefore, this
utilizing a process with lower moisture content. Hanna et al. produced MCC particles by
reactive extrusion [58]. The raw material was fed through an extruder, which consisted of a
screw and a barrel, in the presence of a basic aqueous solution so as to break down the
lignocellulosic complex into lignin, hemicellulose and cellulose. The lignin and hemicellulose
were extracted and the residual cellulose underwent acid hydrolysis to form microcrystalline
cellulose. It was disclosed that this method can provide a quicker process in an efficient
manner and it does not require any mechanical grinding to produce small particles of MCC.
In another paper, Merci et al. reported the use of a simple procedure based on reactive
extrusion to produce MCC from soybean hulls [47]. The method was based on a two-step
extrusion process. The soybean hull was first extruded with sodium hydroxide, followed by
second extrusion with sulfuric acid. These authors revealed that the MCC manufactured by
this method composed of short and rod-shaped fibers, with a cellulose content of 83.79% and
a crystallinity index of 70%. This method was previously scaled up by Hanna et al. and they
reported that the reactive extrusion procedure is an environmental friendly process. However,
the application of this process to produce MCC has received little attention from the scientific
community.
2.1.2.5. Radiation-enzymatic
Finally, MCC can be also produced by a radiation-enzymatic process, but only few
17
bleached dissolving pulp isolated from mountain spruce using an environmental friendly,
irradiation of cellulose pulp was accomplished by employing an electron beam. The sample,
which was swollen in a chemical solution and washed, was subjected to the action of
enzymes. After washing and filtration, the MCC obtained was dried up at ambient
atmosphere. These authors demonstrated that the MCC obtained was characterized by a
know, the enzymatic processes to produce MCC are more expensive to be economically
feasible for applications in practical use in one hand, and on the other hand they create MCC
atoms involved in the reaction process. During the transformation, some bonds or linkages are
broken and new ones are formed. Thus, this reaction mechanism may involve small series of
by-products such as atoms, molecules, electrons, free radicals, and electrons between
substances.
units into a long chain. Cellulose breaks down into glucose and reducing sugars
solution such as HCl, cellulose becomes solubilized in the reaction medium. The H+ moves
towards the β-glycosidic bond while the Cl- weakens the glycosidic bond to facilitate
hydrolysis. When the glycosidic linkage breaks, the H-bonded structure of cellulose begins to
open. From this reaction, glucose and reducing sugar are produced with equal or higher yield
18
[87]. Hydronium ions (H3O+) will be formed by protonation from the acid during glucose
Kinetics show the in-depth reaction mechanism form as it deals with rates during the
chemical reactions [89]. Regarding kinetics, the hydrolysis of cellulose involved the
the concentration of H3O+ which was initially reported by Saeman [90], and the reaction
formic acid, khyd = kinetic constant of cellulose hydrolysis, and kdeg = kinetic constant of
glucose degradation.
From Fig. 7, the protonation of the glycosidic oxygen took place in pathway I whereas the
protonation of pyranic oxygen took place in pathway II. A partial protonation from both
oxygen atoms occurred during the conformational restriction of cellulosic chain along the
glycosidic bond. During acid hydrolysis, the formation of a carbocation via a unimolecular
step was one of the crucial steps as it could be either cycle in pathway I or pathway II formed
from the mechanism. Cyclic carbonation, where the hydrogen and hydroxyl groups were
omitted for clarity also occurred throughout the mechanism which changes the
The first scientific knowledge for acid hydrolysis of cellulose with sulfuric acid was
known as the Scholler process, which was discovered in the 1920s, while the Berglus process
referred to the usage of hydrochloric acid in the hydrolysis of cellulose [86]. Li and Zhao
reported that phosphoric acid was less effective for acid hydrolysis while nitric acid and
hydrochloric acid had similar behaviors towards sulfuric acid without pretreatment [87].
Strong acid had the potential in the protonation of glycosidic oxygen because of its lower pKa
19
value compared to weak acid as glycosidic oxygen was a weak base [86, 88]. For a better acid
In order to study the structure and properties of MCC various complementary techniques
have been employed by many researchers. Such analyses include FTIR, CP/MAS 13C NMR,
and rapid as well as inexpensive that is used to study polymers [91]. Fourier transform
IR radiation by a sample. In this modern era, Fourier transform infrared spectroscopy has
been used more in depth by monitoring the chemical structure and functional groups of
lignocellulosic compounds [67, 91, 92]. For instance, various applications for biological,
biochemical, and food industries are widely used in the mid-region at 4000-400 cm-1 where
the band absorption involves transitions between vibrational energy states and rotational
A broad absorption band will be observed at 3400–3500 cm-1 or 3000-3600 cm-1 in the
FTIR spectra, which indicates the presence of –OH stretching vibrations from α-cellulose
when the chemical structure of MCC is analyzed from different biomass such as rice hulls,
bean hulls, Alfa fibers, jute, cotton silver, oil palm empty fruit bunch, fodder grass, banana
fiber waste, groundnut shells, and corn cob [23, 52, 60, 61, 63, 93-97]. For example in Alfa
fibers [67], the peak located at 1512 cm-1 shows the presence of lignin. However, the alkaline
treatment breaks the glycosidic bond between lignin and other carbohydrate fractions, which
results in disruption of the lignin structure in cellulose (Alfa-C). Thus, the presence of more
crystalline order in the microcrystalline cellulose samples (from Alfa-MCC and C-MCC) can
20
be observed at the peak from 2901 - 2905 cm-1. Absorption of water by cellulose displays an
absorption peak at 1639 cm-1 in Alfa-MCC while MCC in cotton silver, jute, rice hulls, and
bean hulls shows the water absorption peak at 1640 and 1645 cm-1 in oil palm empty fruit
bunch [23, 60, 67, 96, 98]. The occurrence of the absorption of water molecules is due to the
strong interaction between cellulose and water as well as the presence of small amounts of
The absorption bands present at 1425 cm-1, 1430 cm-1, 1432 cm-1 and 1436 cm-1 are due
to intermolecular hydrogen attraction at C6 of the aromatic ring group which increase in the
microcrystalline cellulose sample in rice hulls, bean hulls, oil palm empty fruit bunch, and
Alfa-fibers, respectively [23, 61, 67, 96]. Various studies also revealed that the band between
1157 - 1164 cm-1 corresponds to –C-O-C- stretching of β-1,4-glycosidic linkage [24, 60, 61,
91, 92, 100-102]. C-H vibration of cellulose which corresponds to the β-glycosidic linkage
4.2. Solid-state cross-polarization 13C nuclear magnetic resonance (CP/MAS 13C NMR)
radiofrequency radiation by atomic nuclei together with a nonzero spins in a strong magnetic
field [103]. NMR is a non-destructive method which is able to investigate the structural
behavior under solid phase and solution phase. Differentiation between molecules with
industries utilize NMR to study the final products from pulping processes as detailed
compositional information on the molecular structure of samples such as cellulose and lignin
The peaks of carbon cellulose C1, C4, and C6 appear at the region between 60-105
ppm, which are around δ 102-108 ppm, δ 80-92 ppm, and δ 57-67 ppm, respectively. These
peaks are broad and overlap one another in the NMR spectrum as reported by several reviews
21
[67, 105-107]. The peaks of carbon cellulose C2, C3, and C5 also appear within the region of
60-105 ppm in the spectrum, that is between 72-79 ppm [107]. The appearance of these peaks
within the spectrum of cellulose correspond to the accessible para-crystalline and amorphous
cellulose, as well as inaccessible fibril surfaces [105]. Sharper resonance at δ 86-92 ppm
results from the presence of C4 carbons in the crystalline regions as well as para-crystalline
domain, which is most commonly found in loblolly pine cellulose and MCC from bagasse,
Alfa fiber and cottonwood while the presence of the amorphous region corresponds to the
broad up-field resonance at 80-86 ppm. Thus, C4 carbons have both relative areas of
molecular weight distribution and relative molecular weight of the cellulose polymers [108].
Molecular weight of the cellulose polymer can be characterized by two types which are
number-average molecular weight (Mn) and weight-average molecular weight, (Mw). Apart
from that, polydispersity index (PDI) (Mw/Mn) is also important during the GPC analysis and
corresponds to a molecular weight of about 2 000 000 g mol-1. Depending on the method of
and consequently has an average molecular weight of 50 000–500 000 g mol-1. During
hydrolysis process, due to chain breakage, the molecular weight of MCC decreases compared
to that of the original cellulose and this decrease ranges between 30 000 and 50 000 g mol-1.
The degree of polymerization is typically less than 400 [48]. The determination of the degree
22
reproducibility. This GPC method is usually used to characterize the derived cellulose
using phenyl isocyanate and tetrahydrofuran (THF), which act as an eluting agent in the GPC
system [67, 105, 109]. The possible number-average (DPn) and weight-average (DPw) are
calculated by dividing Mn and Mw by 519 which is the weight corresponded to the monomer
of CTC. Depolymerization of cellulose after acid treatment through the cleavage of glycosidic
are commonly used to investigate the thermal and degradation properties of cellulose sample
[94, 110]. Generally, the TGA curves of MCC show two degradation stages, which indicate
the weight loss of the sample. The first stage corresponds to the removal of water within the
cellulose at the region between 60-140 °C. The second stage concerns the dehydration,
by the char residue formation in the range of 250-450 °C (Table 4). DSC analysis of MCC
indicates the presence of a common endothermic peak, which corresponds to the main
Compared to cellulose, MCC necessitates high energy to be degraded because of the high
degree of molecular ordering. The presence of crystalline region in MCC results in higher
residual weight which is due to the higher amount of crystalline domains in MCC which are
intrinsically flame resistant [67]. According to Adel et al. [98], the degradation of cellulose
23
that MCC exhibits interesting thermal stability and it is a prominent candidate in the
X-ray diffraction (XRD) is commonly used to analyze the identity of crystalline solids
based on their atomic-scale structure of materials [103, 111]. Apart from that, XRD is a
versatile technique that provides in-depth information regarding the chemical composition
and crystallographic structure of the materials. X-rays have wavelengths (λ) in systematic
order of a few angstroms which is same as the interatomic distances which causes diffraction
to occur in crystalline solids [103]. The diffraction pattern displays numerous sharp spots
known as Bragg diffraction peaks [111]. Crystallinity index can be calculated from the height
ratio between intensity of crystalline peak and total intensity of non-crystalline peak using the
I 200 I am (1)
CrI
I 200
where, CrI is crystallinity index, I200 is maximum intensity of the peak, and Iam is intensity of
High degree crystallinity index is obtained for acid hydrolyzed fibers and the average
size of the hydrolyzed fibers is approximately 25 nm [113]. The crystallite sizes are calculated
k
L (2)
cos
where, L is the size of the crystallite (nm), k is the Scherrer constant (0.94), λ is the X-ray
wavelength (0.15418nm), β is the full width half maximum of lattice plane reflection in
24
Various studies demonstrated that crystallinity index and sizes of microcrystalline
cellulose showed the highest value compared to cellulose when tested under X-ray diffraction
(XRD) with different lignocellulosic biomass (Table 5) [24, 52, 61, 67, 80, 93, 94, 115, 116].
hydrolysis process which prompts hydrolytic cleavage of glycosidic bonds and leads to
rearrangement of cellulose molecules [24, 98, 117, 118]. When the crystallinity size and index
increase, the toughness of cellulose structure exhibits better tensile strength towards fiber
[117]. This higher tensile strength is expected to increase the mechanical properties of
composite materials [99]. On the other hand, according to Kumar et al. [119], major amount
of lignin was eliminated during alkaline treatment which was then used for cellulose
extraction while the residual quantity of amorphous lignin was removed during acid
hydrolysis of MCC.
electron microscope (SEM). The image formed is in a three-dimensional structure and thus it
is useful for analysis of the sample. SEM consists of a shadow-relief effect of both secondary
and backscattered electron contrast, and is able to scan bulky images better with a greater
view, as well as produce a good image of the 3D structure of the sample [103, 120]. On the
other hand, transmission electron microscopy (TEM) is a powerful and unique nanoscale
imaging technique which uses an even higher resolution to construct the sample’s structure
[121]. It can control the size distribution and surface morphology of the sample, but the
analyzing process is time-consuming. By comparing SEM with TEM, SEM has the advantage
of scanning larger amount of sample once at higher magnification and greater resolution
[103].
25
There is not much difference in the MCC surface morphology as compared with several
types of sources such as MCC from Alfa fiber, rice hulls, bean hulls, bagasse, cotton silver,
jute, rice straw, cotton straw, and oil palm empty fruit brunch, as well as commercial MCC as
shown in Table 6 [52, 61, 67, 76, 96]. It seems that when the cellulose sample underwent
hydrolysis treatment, the surface morphology of MCC changed in terms of size and level of
smoothness. Before the treatment, cellulose exhibited an irregular shape in fibrils, rough
surface area, and was often aggregated as reported by Hemsri et al. [122]. According to
Trache et al. [67] and Adel et al. [98], the average diameter of commercial MCC obtained was
Fewer researchers had used TEM to examine the surface morphology of MCC. The
MCC extracted from oil palm empty fruit bunch revealed that MCC particles have the
tendency to agglomerate, thus forming larger particles in the micron range, while the MCC
The growing global interest to produce green materials in the recent past that can reduce
Polymer bio-composites filled with MCC is a relatively new research area with respect to
such as high strength and stiffness combined with high surface area, unique morphology, low
incompatibility with most polymeric matrices and limitation of processing temperature [20,
21, 26, 31]. For instance, cellulosic materials start to degrade near 220 °C, thus restricting the
type of matrix that can be used in association with cellulosic fillers [67].
26
The potential of MCC in various sectors of research and application is promising and
attracting increasing investments. There has been an increased interest in research and product
polyethylene reinforced with natural fibers. These composite materials are used extensively in
automotive applications, building materials, and household products [124]. This approach of
Among the factors contributing to this preference include, their numerous ecological
benefits and diversified raw materials [125]. Extensive research progress report has been
made in the last two decades on the production of biodegradable composites with biomaterials
at a reduced cost while maintaining the original properties of the bio-composite [126]. MCC
is viewed as a readily available baseline reinforcement [20, 21, 23, 25, 26, 28-30]. The
application of MCC as a reinforcing agent with excellent properties has attracted increasing
attention and interest in the past two decades due to its potential advantages such as
renewability, biodegradability and high surface area for bonding. Current global research in
bio-composite materials technology is geared towards replacing the current pure polymers or
glass fiber composites with lignocellulosic biomass fibers and natural fiber reinforced
polymer composites to create low cost, high performance, and low weight composite
materials [127]. In addition, these researches have led to bio-composite products with low
To obtain better interaction between the matrix and reinforcement, materials that
provide large surface area are considered [128]. Application of such materials leads to better
mechanical properties, heat resistance and dimensional stability [129, 130]. Such cellulosic
27
amorphous regions have been removed by hydrolysis process. Wood pulp has been the major
source of MCC, but research findings have confirmed the application of vegetative plants and
lignocellulosic biomass suitable for the preparation of MCC that is expected to disintegrate
into cellulose whiskers after a complete hydrolysis The major use of reinforcing filler is
important to enhance its mechanical properties and materials from biomass and it has been of
great interest because of their renewable nature, biocompatibility, and low energy
The high specific surface area of MCC compared to other conventional cellulose fibers
is an added advantage of MCC. MCC has generated much attention and interest in both
academic and industrial fields. Besides that, MCC also exhibits a unique capacity to improve
the morphology, thermal and mechanical properties of the composite [20, 21, 23, 25, 26, 28-
30, 132-134]. Its average microfibril aspect ratio and fibrillar structure control the
reinforcement effects. MCC forms a stable gel in water, and this is a major reason for its use.
Among the challenges of MCC is that it is difficult to disintegrate once it is dried. This
development is attributed to the presence of hydrogen bonds that cause strong adhesion
favorably to the mechanical behavior of most polymer bio-composites due to their short
microfibrils are expected to contribute favorably to the mechanical behavior of most polymer
weight, and biodegradability make them ideal candidates for the processing of polymer
nanocomposites [135, 136]. MCC have been reported to possess a Young’s modulus of
150 GPa and a surface area of several hundred [137], and they have the potential to
28
6. Current applications of microcrystalline cellulose
binders, adsorbents, flowability), food (as stabilizers, anti-caking agents, fat substitutes and
emulsifiers), beverage (as gelling agents, stabilizers, anti-caking agents and suspending
agents) cosmetic (as fat substitutes, thickeners, binders) and other industries (as binders) [20,
21, 39, 47, 48]. Among the unique properties are its chemical inactivity and hygroscopicity.
Furthermore, the absence of toxicity and high sorption has made MCC possible to be applied
in various industries [138]. The higher thermal stability of MCC over natural fiber has
[136].
Recently, successful preparation of MCC composite with other polymeric matrices has
been reported. Among the bio-composites with MCC are starch [98], chitosan [139],
poly(lactic acid) [23, 26], poly (3-hydroxybutyrate) [140] and poly(vinyl alcohol) [21]. The
dispersion levels of the filler within the polymeric matrix. In the interim, a palliative measure
in the use of latex in an aqueous suspension of the polymer to form a matrix has been reported
[135]. Besides that, owing to the emergence of new applications of bio-composites, extensive
research is currently being carried out by many research laboratories and companies with
MCC fiber and polymers to develop parts of miscellaneous sectors such as construction, food,
29
7. Concluding remarks
for further understanding of this biopolymer, is described. Here, we mainly focus on the MCC
techniques and hydrolysis methods are discussed in the preparation of MCC from different
sources. Several properties of this renewable material are considered, such as chemical
composition, molecular weight, crystallinity, morphology and thermal properties. Owing to its
interesting characteristics, various uses continue to be developed by scientists all around the
world. With regards to MCC-based composites applications, we only report and discuss
several selected examples since the number of polymers used and their applications in various
isolation procedures, treatments and drying. Furthermore, to date, the engineering properties
and MCC-based composites performances are still being developed. In order to satisfy the
criteria of employing MCC for widespread use with higher efficiency, more effort and
developments are required to expand the use of MCC for science and technology.
Acknowledgements
Authors wish to thanks their parental institutes (Universiti Sains Malaysia through USM Short
Term Grant–304/PKIMIA/6313216 and 304/PTEKIND/6313194) for providing the necessary
facility to accomplish this work.
References
[1] S. Kalia, B. Kaith, I. Kaur, Cellulose fibers: bio-and nano-polymer composites: green
chemistry and technology, Springer Science & Business Media 2011.
[2] H. Abdul Khalil, A. Bhat, A.I. Yusra, Green composites from sustainable cellulose
nanofibrils: a review, Carbohyd. Polym. 87(2) (2012) 963-979.
30
[3] V.K. Thakur, M.K. Thakur, Processing and characterization of natural cellulose
fibers/thermoset polymer composites, Carbohyd. Polym. 109 (2014) 102-117.
[4] V.K. Thakur, Nanocellulose Polymer Nanocomposites: Fundamentals and Applications,
John Wiley & Sons2015.
[5] C. Miao, W.Y. Hamad, Cellulose reinforced polymer composites and nanocomposites: a
critical review, Cellulose 20(5) (2013) 2221-2262.
[6] V.K. Thakur, Lignocellulosic polymer composites: Processing, characterization, and
properties, John Wiley & Sons2015.
[7] X. Ma, P.R. Chang, J. Yu, Properties of biodegradable thermoplastic pea
starch/carboxymethyl cellulose and pea starch/microcrystalline cellulose composites,
Carbohyd. Polym. 72(3) (2008) 369-375.
[8] K.G. Satyanarayana, G.G. Arizaga, F. Wypych, Biodegradable composites based on
lignocellulosic fibers—An overview, Prog. Polym. Sci. 34(9) (2009) 982-1021.
[9] E.M. Fernandes, R.A. Pires, J.F. Mano, R.L. Reis, Bionanocomposites from
lignocellulosic resources: Properties, applications and future trends for their use in the
biomedical field, Prog. Polym. Sci. 38(10) (2013) 1415-1441.
[10] A. Pappu, V. Patil, S. Jain, A. Mahindrakar, R. Haque, V.K. Thakur, Advances in
industrial prospective of cellulosic macromolecules enriched banana biofibre resources: A
review, Int. J. Biol. Macromol. 79 (2015) 449-458.
[11] M.C. Corobea, O. Muhulet, F. Miculescu, I.V. Antoniac, Z. Vuluga, D. Florea, D.M.
Vuluga, M. Butnaru, D. Ivanov, S.I. Voicu, Novel nanocomposite membranes from cellulose
acetate and clay‐silica nanowires, Polymer. Adv. Tech. (2016) 10.1002/pat.3835.
[12] S.I. Voicu, R.M. Condruz, V. Mitran, A. Cimpean, F. Miculescu, C. Andronescu, M.
Miculescu, V.K. Thakur, Sericin Covalent Immobilization onto Cellulose Acetate Membrane
for Biomedical Applications, ACS Sustain. Chem. Eng. 4(3) (2016) 1765-1774.
[13] V.K. Thakur, M.K. Thakur, R.K. Gupta, Rapid synthesis of graft copolymers from
natural cellulose fibers, Carbohyd. Polym. 98(1) (2013) 820-828.
[14] L. Brinchi, F. Cotana, E. Fortunati, J. Kenny, Production of nanocrystalline cellulose
from lignocellulosic biomass: technology and applications, Carbohyd. Polym. 94(1) (2013)
154-169.
[15] H. Abdul Khalil, Y. Davoudpour, M.N. Islam, A. Mustapha, K. Sudesh, R. Dungani, M.
Jawaid, Production and modification of nanofibrillated cellulose using various mechanical
processes: a review, Carbohyd. Polym. 99 (2014) 649-665.
31
[16] V.K. Thakur, S.I. Voicu, Recent advances in cellulose and chitosan based membranes for
water purification: A concise review, Carbohyd. Polym. 146 (2016) 148-165.
[17] D. Trache, K. Khimeche, A. Mezroua, M. Benziane, Physicochemical properties of
microcrystalline nitrocellulose from Alfa grass fibres and its thermal stability, J. Therm. Anal.
Calorim. 124(3) (2016) 1485-1496.
[18] L. Berglund, Cellulose-based nanocomposites, in: A.K. Mohanty, M. Misra, L.T. Drzal
(Eds.), Natural fibers, biopolymers, and biocomposites, CRC Press2005, pp. 807-832.
[19] A. Boldizar, C. Klason, J. Kubat, P. Näslund, P. Saha, Prehydrolyzed cellulose as
reinforcing filler for thermoplastics, Int. J. Polym. Mater. 11(4) (1987) 229-262.
[20] N. Izzati Zulkifli, N. Samat, H. Anuar, N. Zainuddin, Mechanical properties and failure
modes of recycled polypropylene/microcrystalline cellulose composites, Mater. Design 69
(2015) 114-123.
[21] X. Sun, C. Lu, Y. Liu, W. Zhang, X. Zhang, Melt-processed poly (vinyl alcohol)
composites filled with microcrystalline cellulose from waste cotton fabrics, Carbohyd. Polym.
101 (2014) 642-649.
[22] C.G. Hoyos, E. Cristia, A. Vázquez, Effect of cellulose microcrystalline particles on
properties of cement based composites, Mater. Design 51 (2013) 810-818.
[23] M.M. Haafiz, A. Hassan, Z. Zakaria, I.M. Inuwa, M.S. Islam, M. Jawaid, Properties of
polylactic acid composites reinforced with oil palm biomass microcrystalline cellulose,
Carbohyd. Polym. 98(1) (2013) 139-145.
[24] M.M. Haafiz, S. Eichhorn, A. Hassan, M. Jawaid, Isolation and characterization of
microcrystalline cellulose from oil palm biomass residue, Carbohyd. Polym. 93(2) (2013)
628-634.
[25] X. Xiao, S. Lu, B. Qi, C. Zeng, Z. Yuan, J. Yu, Enhancing the thermal and mechanical
properties of epoxy resins by addition of a hyperbranched aromatic polyamide grown on
microcrystalline cellulose fibers, RSC Adv. 4(29) (2014) 14928-14935.
[26] X. Dai, Z. Xiong, H. Na, J. Zhu, How does epoxidized soybean oil improve the
toughness of microcrystalline cellulose filled polylactide acid composites?, Compos. Sci.
Technol. 90 (2014) 9-15.
[27] A. Cataldi, A. Dorigato, F. Deflorian, A. Pegoretti, Thermo-mechanical properties of
innovative microcrystalline cellulose filled composites for art protection and restoration, J.
Mater. Sci. 49(5) (2014) 2035-2044.
[28] Z. Rafiee, V. Keshavarz, Synthesis and characterization of polyurethane/microcrystalline
cellulose bionanocomposites, Prog. Org. Coat. 86 (2015) 190-193.
32
[29] A. Cataldi, A. Dorigato, F. Deflorian, A. Pegoretti, Innovative microcrystalline cellulose
composites as lining adhesives for canvas, Polym. Eng. Sci. 55(6) (2015) 1349-1354.
[30] A. Cataldi, F. Deflorian, A. Pegoretti, Poly 2-ethyl-2-oxazoline/microcrystalline cellulose
composites for cultural heritage conservation: Mechanical characterization in dry and wet
state and application as lining adhesives of canvas, Int. J. Adhes. Adhes. 62 (2015) 92-100.
[31] S. Spoljaric, A. Genovese, R.A. Shanks, Polypropylene–microcrystalline cellulose
composites with enhanced compatibility and properties, Compos. Part. A Appl. Sci. Manuf.
40(6) (2009) 791-799.
[32] S. Ummartyotin, H. Manuspiya, A critical review on cellulose: from fundamental to an
approach on sensor technology, Renew. Sust. Energ. Rev. 41 (2015) 402-412.
[33] N. Lavoine, I. Desloges, A. Dufresne, J. Bras, Microfibrillated cellulose–Its barrier
properties and applications in cellulosic materials: A review, Carbohyd. Polym. 90(2) (2012)
735-764.
[34] D. Klemm, F. Kramer, S. Moritz, T. Lindström, M. Ankerfors, D. Gray, A. Dorris,
Nanocelluloses: A new family of nature‐based materials, Angew. Chem. Int. Edit. 50(24)
(2011) 5438-5466.
[35] J.-L. Wertz, J.P. Mercier, O. Bédué, Cellulose science and technology, CRC Press,
Switzerland, 2010.
[36] G. Siqueira, J. Bras, A. Dufresne, Cellulosic bionanocomposites: a review of preparation,
properties and applications, Polymers 2(4) (2010) 728-765.
[37] Y. Habibi, L.A. Lucia, O.J. Rojas, Cellulose nanocrystals: chemistry, self-assembly, and
applications, Chem. Rev. 110(6) (2010) 3479-3500.
[38] D. Klemm, B. Heublein, H.P. Fink, A. Bohn, Cellulose: fascinating biopolymer and
sustainable raw material, Angew. Chem. Int. Edit. 44(22) (2005) 3358-3393.
[39] M.A.S. Azizi Samir, F. Alloin, A. Dufresne, Review of recent research into cellulosic
whiskers, their properties and their application in nanocomposite field, Biomacromolecules
6(2) (2005) 612-626.
[40] S. Hokkanen, A. Bhatnagar, M. Sillanpää, A review on modification methods to
cellulose-based adsorbents to improve adsorption capacity, Water Res. 91 (2016) 156-173.
[41] M.M. de Souza Lima, R. Borsali, Rodlike cellulose microcrystals: structure, properties,
and applications, Macromol. Rapid Comm. 25(7) (2004) 771-787.
[42] J. Borges, J. Canejo, S. Fernandes, P. Brogueira, M. Godinho, Cellulose-Based Liquid
Crystalline Composite Systems, in: V.K. Thakur (Ed.), Nanocellulose Polymer
Nanocomposites: Fundamentals and Applications, Wiley-Scrivener2015, pp. 215-235.
33
[43] D.W. O’Connell, C. Birkinshaw, T.F. O’Dwyer, Heavy metal adsorbents prepared from
the modification of cellulose: A review, Bioresource Technol. 99(15) (2008) 6709-6724.
[44] P. Zugenmaier, Crystalline Cellulose and Cellulose derivatives: Characterization and
structures, Springer, Germany, 2008.
[45] K. Leppänen, S. Andersson, M. Torkkeli, M. Knaapila, N. Kotelnikova, R. Serimaa,
Structure of cellulose and microcrystalline cellulose from various wood species, cotton and
flax studied by X-ray scattering, Cellulose 16(6) (2009) 999-1015.
[46] O. Battista, P. Smith, Microcrystalline cellulose, Ind. Eng. Chem. 54(9) (1962) 20-29.
[47] A. Merci, A. Urbano, M.V.E. Grossmann, C.A. Tischer, S. Mali, Properties of
microcrystalline cellulose extracted from soybean hulls by reactive extrusion, Food Res. Int.
73 (2015) 38-43.
[48] G. Thoorens, F. Krier, B. Leclercq, B. Carlin, B. Evrard, Microcrystalline cellulose, a
direct compression binder in a quality by design environment—A review, Int. J. Pharm.
473(1) (2014) 64-72.
[49] M.H. Hussin, N.A. Pohan, Z.N. Garba, M.J. Kassim, A.A. Rahim, N. Brosse, M.
Yemloul, M.N. Fazita, M.M. Haafiz, Physicochemical of microcrystalline cellulose from oil
palm fronds as potential methylene blue adsorbents, Int. J. Biol. Macromol. 92 (2016) 11-19.
[50] K. Vijayalakshmi, T. Gomathi, S. Latha, T. Hajeeth, P. Sudha, Removal of copper (II)
from aqueous solution using nanochitosan/sodium alginate/microcrystalline cellulose beads,
Int. J. Biol. Macromol. 82 (2016) 440-452.
[51] T. Shcherbakova, N. Kotelnikova, Y.V. Bykhovtseva, Comparative study of powdered
and microcrystalline cellulose samples of a various natural origins: Physical and chemical
characteristics, Russ. J. Bioorgan. Chem. 38(7) (2012) 689-696.
[52] K. Das, D. Ray, N. Bandyopadhyay, S. Sengupta, Study of the properties of
microcrystalline cellulose particles from different renewable resources by XRD, FTIR,
nanoindentation, TGA and SEM, J. Polym. Environ. 18(3) (2010) 355-363.
[53] F. Ohwoavworhua, O. Kunle, S. Ofoefule, Extraction and characterization of
microcrystalline cellulose derived from Luffa cylindrica plant, Afr. J. Pharm. Res. Dev. 1(1)
(2004) 1-6.
[54] J. Jain, V. Dixit, K. Varma, Preparation of microcrystalline cellulose from cereal straw
and its evaluation as a tablet excipient, Indian J. Pharm. Sci. 45(3) (1983) 83-85.
[55] A. Okhamafe, A. Igboechi, T. Obaseki, Celluloses extracted from groundnut shell and
rice husk 1: Preliminary physicochemical characterization, Pharm. World J. 8(4) (1991) 120-
130.
34
[56] A. Okhamafe, E. Ejike, F. Akinrinola, D. Ubane-Ine, Aspect of tablet disintegrant
properties of cellulose derived from bagasse and maize cob, West Afr. J. Pharm. 9(1) (1995)
8-13.
[57] S. Ofoefule, A. Chukwu, Application of blends of MCC–Cissus gum in the formation of
aqueous suspensions, Boll. Chim. Farm. 138(5) (1999) 217–222.
[58] M. Hanna, G. Biby, V. Miladinov, Production of microcrystalline cellulose by reactive
extrusion, US Patent 6228213, 2001.
[59] P.M. Ejikeme, Investigation of the physicochemical properties of microcrystalline
cellulose from agricultural wastes I: Orange mesocarp, Cellulose 15(1) (2008) 141-147.
[60] M.S. Jahan, A. Saeed, Z. He, Y. Ni, Jute as raw material for the preparation of
microcrystalline cellulose, Cellulose 18(2) (2011) 451-459.
[61] A.M. Adel, Z.H.A. El-Wahab, A.A. Ibrahim, M.T. Al-Shemy, Characterization of
microcrystalline cellulose prepared from lignocellulosic materials. Part II: Physicochemical
properties, Carbohyd. Polym. 83(2) (2011) 676-687.
[62] T. Virtanen, K. Svedström, S. Andersson, L. Tervala, M. Torkkeli, M. Knaapila, N.
Kotelnikova, S.L. Maunu, R. Serimaa, A physico-chemical characterisation of new raw
materials for microcrystalline cellulose manufacturing, Cellulose 19(1) (2012) 219-235.
[63] R.D. Kalita, Y. Nath, M.E. Ochubiojo, A.K. Buragohain, Extraction and characterization
of microcrystalline cellulose from fodder grass; Setaria glauca (L) P. Beauv, and its potential
as a drug delivery vehicle for isoniazid, a first line antituberculosis drug, Colloids Surf. B
Biointerfaces 108 (2013) 85-89.
[64] O.O. Okwonna, The effect of pulping concentration treatment on the properties of
microcrystalline cellulose powder obtained from waste paper, Carbohyd. Polym. 98(1) (2013)
721-725.
[65] M. Ahmadi, A. Madadlou, A.A. Sabouri, Isolation of micro-and nano-crystalline
cellulose particles and fabrication of crystalline particles-loaded whey protein cold-set gel,
Food Chem. 174 (2015) 97-103.
[66] P. Chaiwutthinan, V. Pimpan, S. Chuayjuljit, T. Leejarkpai, 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. 23(1) (2015)
114-125.
[67] D. Trache, A. Donnot, K. Khimeche, R. Benelmir, N. Brosse, Physico-chemical
properties and thermal stability of microcrystalline cellulose isolated from Alfa fibres,
Carbohyd. Polym. 104 (2014) 223-230.
35
[68] S.M. Keshk, M.A. Haija, A new method for producing microcrystalline cellulose from
Gluconacetobacter xylinus and kenaf, Carbohyd. Polym. 84(4) (2011) 1301-1305.
[69] R.L. de Oliveira, H. da Silva Barud, R.M. de Assunçao, C. da Silva Meireles, G.O.
Carvalho, G.R. Filho, Y. Messaddeq, S.J.L. Ribeiro, Synthesis and characterization of
microcrystalline cellulose produced from bacterial cellulose, J. Therm. Anal. Calorim. 106(3)
(2011) 703-709.
[70] J. Pandey, H. Takagi, A. Nakagaito, H. Kim, Handbook of polymer nanocomposites.
Processing, performance and application, Springer2015.
[71] V.B. Agbor, N. Cicek, R. Sparling, A. Berlin, D.B. Levin, Biomass pretreatment:
fundamentals toward application, Biotechnology advances 29(6) (2011) 675-685.
[72] G. Thoorens, F. Krier, E. Rozet, B. Carlin, B. Evrard, Understanding the impact of
microcrystalline cellulose physicochemical properties on tabletability, Int. J. Pharm. 490(1)
(2015) 47-54.
[73] M. Balaxi, I. Nikolakakis, K. Kachrimanis, S. Malamataris, Combined effects of wetting,
drying, and microcrystalline cellulose type on the mechanical strength and disintegration of
pellets, J. Pharm. Sci. 98(2) (2009) 676-689.
[74] A. Kiziltas, D.J. Gardner, Y. Han, H.-S. Yang, Dynamic mechanical behavior and
thermal properties of microcrystalline cellulose (MCC)-filled nylon 6 composites,
Thermochim. Acta 519(1) (2011) 38-43.
[75] B.G. Ranby, Aqueous colloidal solutions of cellulose micelles, Acta. Chem. Scand. 3(5)
(1949) 649-650.
[76] M. El-Sakhawy, M.L. Hassan, Physical and mechanical properties of microcrystalline
cellulose prepared from agricultural residues, Carbohyd. Polym. 67(1) (2007) 1-10.
[77] H. Håkansson, P. Ahlgren, Acid hydrolysis of some industrial pulps: effect of hydrolysis
conditions and raw material, Cellulose 12(2) (2005) 177-183.
[78] D. Schaible, B. Sherwood, Treatment of pulp to produce microcrystalline cellulose, US
Patent 0131957 A1, 2004.
[79] S. Trusovs, Microcrystalline cellulose, US Patent 6392034 B1, 2002.
[80] X.T. Nguyen, Process for preparing microcrystalline cellulose, US Patent 7005514 B2,
2006.
[81] N. Jacquet, C. Vanderghem, S. Danthine, N. Quievy, C. Blecker, J. Devaux, M. Paquot,
Influence of steam explosion on physicochemical properties and hydrolysis rate of pure
cellulose fibers, Bioresource Technol. 121 (2012) 221-227.
36
[82] E.Y. Ha, C.D. Landi, Method for producing microcrystalline cellulose, US Patent
5769934, 1998.
[83] E.A. DeLong, Method of producing level off DP microcrystallinecellulose and glucose
from lignocellulosic material, US Patent 4645541, 1987.
[84] B. Lamsal, J. Yoo, K. Brijwani, S. Alavi, Extrusion as a thermo-mechanical pre-
treatment for lignocellulosic ethanol, Biomass Bioenrg. 34(12) (2010) 1703-1710.
[85] H. Stupińska, E. Iller, Z. Zimek, D. Wawro, D. Ciechańska, E. Kopania, J. Palenik, S.
Milczarek, W. Stęplewski, G. Krzyżanowska, An environment-friendly method to prepare
microcrystalline cellulose, Fibres Text. East. Eur. (5-6 (64)) (2007) 167--172.
[86] R. Rinaldi, F. Schüth, Acid hydrolysis of cellulose as the entry point into biorefinery
schemes, ChemSusChem 2(12) (2009) 1096-1107.
[87] C. Li, Z.K. Zhao, Efficient Acid‐Catalyzed Hydrolysis of Cellulose in Ionic Liquid, Adv.
Syn. Catal. 349(11‐12) (2007) 1847-1850.
[88] N.S. Mosier, C.M. Ladisch, M.R. Ladisch, Characterization of acid catalytic domains for
cellulose hydrolysis and glucose degradation, Biotechnol. Bioeng. 79(6) (2002) 610-618.
[89] A. Frost, R. Pearson, Kinetics and mechanism, J. Phys. Chem. 65(2) (1961) 384-384.
[90] J.F. Saeman, Kinetics of wood saccharification-hydrolysis of cellulose and
decomposition of sugars in dilute acid at high temperature, Ind. Eng. Chem. 37(1) (1945) 43-
52.
[91] M. Kačuráková, R. Wilson, Developments in mid-infrared FT-IR spectroscopy of
selected carbohydrates, Carbohyd. Polym. 44(4) (2001) 291-303.
[92] S.F. Sim, M. Mohamed, N.A.L.M.I. Lu, N.S.P. Sarman, S.N.S. Samsudin, Computer-
assisted analysis of fourier transform infrared (FTIR) spectra for characterization of various
treated and untreated agriculture biomass, BioResources 7(4) (2012) 5367-5380.
[93] C.P. Azubuike, A.O. Okhamafe, Physicochemical, spectroscopic and thermal properties
of microcrystalline cellulose derived from corn cobs, Int. J. Recycl. Org. Waste Agricult. 1(1)
(2012) 1-7.
[94] C. Azubuike, Physicochemical, spectroscopic and thermal properties of powdered
cellulose and microcrystalline cellulose derived from groundnut shells, J. Excipients Food
Chem. 3(3) (2012) 106-115.
[95] S. Elanthikkal, U. Gopalakrishnapanicker, S. Varghese, J.T. Guthrie, Cellulose
microfibres produced from banana plant wastes: Isolation and characterization, Carbohyd.
Polym. 80(3) (2010) 852-859.
37
[96] M.M. Haafiz, A. Hassan, Z. Zakaria, I. Inuwa, Isolation and characterization of cellulose
nanowhiskers from oil palm biomass microcrystalline cellulose, Carbohyd. Polym. 103 (2014)
119-125.
[97] A. El Ghali, I.B. Marzoug, M.H.V. Baouab, M.S. Roudesli, Separation and
characterization of new cellulosic fibres from Juncus acutus plant, BioResources 7(2) (2012)
2002-2018.
[98] A.M. Adel, N.A. El-Shinnawy, Hypolipidemic applications of microcrystalline cellulose
composite synthesized from different agricultural residues, Int. J. Biol. Macromol. 51(5)
(2012) 1091-1102.
[99] S.M. Rosa, N. Rehman, M.I.G. de Miranda, S.M. Nachtigall, C.I. Bica, Chlorine-free
extraction of cellulose from rice husk and whisker isolation, Carbohyd. Polym. 87(2) (2012)
1131-1138.
[100] M.Z. Karim, Z.Z. Chowdhury, S.B.A. Hamid, M.E. Ali, Statistical optimization for acid
hydrolysis of microcrystalline cellulose and its physiochemical characterization by using
metal ion catalyst, Materials 7(10) (2014) 6982-6999.
[101] J. Shi, J. Li, Metabolites and chemical group changes in the wood-forming tissue of
Pinus koraiensis under inclined conditions, BioResources 7(3) (2012) 3463-3475.
[102] W. Zhou, D. Zhu, A. Langdon, L. Li, S. Liao, L. Tan, The structure characterization of
cellulose xanthogenate derived from the straw of Eichhornia crassipes, Bioresource Technol.
100(21) (2009) 5366-5369.
[103] D.M. Luykx, R.J. Peters, S.M. van Ruth, H. Bouwmeester, A review of analytical
methods for the identification and characterization of nano delivery systems in food, J. Agr.
Food Chem. 56(18) (2008) 8231-8247.
[104] J.F. Haw, G.E. Maciel, H.A. Schroeder, Carbon-13 nuclear magnetic resonance
spectrometric study of wood and wood pulping with cross polarization and magic-angle
spinning, Anal. Chem. 56(8) (1984) 1323-1329.
[105] M.B. Foston, C.A. Hubbell, A.J. Ragauskas, Cellulose isolation methodology for NMR
analysis of cellulose ultrastructure, Materials 4(11) (2011) 1985-2002.
[106] P. Sannigrahi, A.J. Ragauskas, S.J. Miller, Effects of two-stage dilute acid pretreatment
on the structure and composition of lignin and cellulose in loblolly pine, Bioenerg. Res. 1(3-4)
(2008) 205-214.
[107] D. Bhattacharya, L.T. Germinario, W.T. Winter, Isolation, preparation and
characterization of cellulose microfibers obtained from bagasse, Carbohyd. Polym. 73(3)
(2008) 371-377.
38
[108] P. Engel, L. Hein, A.C. Spiess, Derivatization-free gel permeation chromatography
elucidates enzymatic cellulose hydrolysis, Biotechnol. Biofuels 5(1) (2012) 1.
[109] C.A. Hubbell, A.J. Ragauskas, Effect of acid-chlorite delignification on cellulose degree
of polymerization, Bioresource Technol. 101(19) (2010) 7410-7415.
[110] D. Trache, K. Khimeche, A. Donnot, R. Benelmir, Thermal analysis of microcrystalline
cellulose prepared from esparto grass, MATEC Web of Conferences, EDP Sciences, 2013, p.
01067.
[111] V. Petkov, Nanostructure by high-energy X-ray diffraction, Mater. Today 11(11) (2008)
28-38.
[112] L. Segal, J. Creely, A. Martin, C. Conrad, An empirical method for estimating the
degree of crystallinity of native cellulose using the X-ray diffractometer, Text. Res. J. 29(10)
(1959) 786-794.
[113] R. Chandrahasa, N. Rajamane, Jeyalakshmi., Development of cellulose nanofibres from
coconut husks, Int. J. Emerg.Technol. Adv. Eng. 4(4) (2014) 2250-2259.
[114] J. He, Y. Tang, S.-Y. Wang, Differences in Morphological Characteristics of Bamboo
Fibres and other Natural Cellulose Fibres: Studies on X-Ray Diffraction, Solid State^ 1^ 3C-
CP/MAS NMR, and Second Derivative FTIR Spectroscopy Data, Iran. Polym. J. 16(12)
(2007) 807.
[115] Y. Wang, X. Cao, L. Zhang, Effects of cellulose whiskers on properties of soy protein
thermoplastics, Macromol. Biosci. 6(7) (2006) 524-531.
[116] M. Pracella, D. Chionna, I. Anguillesi, Z. Kulinski, E. Piorkowska, Functionalization,
compatibilization and properties of polypropylene composites with hemp fibres, Compos. Sci.
Technol. 66(13) (2006) 2218-2230.
[117] M. Beg, M. Rosli, R. Ramli, N. Junadi, Microcrystalline Cellulose (MCC) from Oil
Palm Empty Fruit Bunch (EFB) Fiber via Simultaneous Ultrasonic and Alkali Treatment, Int.
J. Chem. Nucl. Mater. Metall. Eng. 9(1) (2015) 8-11.
[118] D. Trache, K. Khimeche, A. Donnot, R. Benelmir, FTIR spectroscopy and X-ray
powder diffraction characterization of microcrystalline cellulose obtained from alfa fibers,
MATEC Web of Conferences, EDP Sciences, 2013, p. 01023.
[119] R. Kumar, F. Hu, C.A. Hubbell, A.J. Ragauskas, C.E. Wyman, Comparison of
laboratory delignification methods, their selectivity, and impacts on physiochemical
characteristics of cellulosic biomass, Bioresource Technol. 130 (2013) 372-381.
39
[120] J. Goldstein, D.E. Newbury, P. Echlin, D.C. Joy, A.D. Romig Jr, C.E. Lyman, C. Fiori,
E. Lifshin, Scanning electron microscopy and X-ray microanalysis: a text for biologists,
materials scientists, and geologists, Springer Science & Business Media2012.
[121] Z. Wang, Transmission electron microscopy of shape-controlled nanocrystals and their
assemblies, J. Phys. Chem. B 104(6) (2000) 1153-1175.
[122] S. Hemsri, K. Grieco, A.D. Asandei, R.S. Parnas, Wheat gluten composites reinforced
with coconut fiber, Compos. Part. A Appl. Sci. Manuf. 43(7) (2012) 1160-1168.
[123] D. Klemm, D. Schumann, F. Kramer, N. Heßler, M. Hornung, H.-P. Schmauder, S.
Marsch, Nanocelluloses as innovative polymers in research and application, Polysaccharides
Ii, Springer2006, pp. 49-96.
[124] B. Ghanbarzadeh, S.A. Oleyaei, H. Almasi, Nanostructured Materials Utilized in
Biopolymer-based Plastics for Food Packaging Applications, Crit. Rev. Food Sci. Nutr.
55(12) (2015) 1699-1723.
[125] W.-m. Wang, Z.-s. Cai, J.-y. Yu, Study on the chemical modification process of jute
fiber, J. Eng. Fiber. Fabr. 3(2) (2008) 1-11.
[126] W. Wang, M.D. Mozuch, R.C. Sabo, P. Kersten, J. Zhu, Y. Jin, Production of cellulose
nanofibrils from bleached eucalyptus fibers by hyperthermostable endoglucanase treatment
and subsequent microfluidization, Cellulose 22(1) (2015) 351-361.
[127] X. Xu, H. Wang, L. Jiang, X. Wang, S.A. Payne, J. Zhu, R. Li, Comparison between
cellulose nanocrystal and cellulose nanofibril reinforced poly (ethylene oxide) nanofibers and
their novel shish-kebab-like crystalline structures, Macromolecules 47(10) (2014) 3409-3416.
[128] S. Yang, Y. Tang, J. Wang, F. Kong, J. Zhang, Surface treatment of cellulosic paper
with starch-based composites reinforced with nanocrystalline cellulose, Ind. Eng. Chem. Res.
53(36) (2014) 13980-13988.
[129] J. Yu, C. Wang, J. Wang, F. Chu, In situ development of self-reinforced cellulose
nanocrystals based thermoplastic elastomers by atom transfer radical polymerization,
Carbohyd. Polym. 141 (2016) 143-150.
[130] G.Y. Yun, H.S. Kim, J. Kim, K. Kim, C. Yang, Effect of aligned cellulose film to the
performance of electro-active paper actuator, Sensor. Actua. A Phys. 141(2) (2008) 530-535.
[131] M. Wang, X.-W. Han, L. Liu, X.-F. Zeng, H.-K. Zou, J.-X. Wang, J.-F. Chen,
Transparent aqueous Mg (OH) 2 nanodispersion for transparent and flexible polymer film
with enhanced flame-retardant property, Ind. Eng. Chem. Res. 54(51) (2015) 12805-12812.
40
[132] L. Petersson, I. Kvien, K. Oksman, Structure and thermal properties of poly (lactic
acid)/cellulose whiskers nanocomposite materials, Compos. Sci. Technol. 67(11) (2007)
2535-2544.
[133] D.M. Panaitescu, D. Donescu, C. Bercu, M.V. Dumitru, Polymer composites with
cellulose microfibrils, Polym. Eng. Sci. 47(8) (2007) 1228-1234.
[134] A.P. Mathew, K. Oksman, M. Sain, Mechanical properties of biodegradable composites
from poly lactic acid (PLA) and microcrystalline cellulose (MCC), J. Appl. Polym. Sci. 97(5)
(2005) 2014-2025.
[135] R. Arjmandi, A. Hassan, M. Haafiz, Z. Zakaria, M.S. Islam, Effect of hydrolysed
cellulose nanowhiskers on properties of montmorillonite/polylactic acid nanocomposites, Int.
J. Biol. Macromol. 82 (2016) 998-1010.
[136] Y. Davoudpour, S. Hossain, H.A. Khalil, M.M. Haafiz, Z.M. Ishak, A. Hassan, Z.I.
Sarker, Optimization of high pressure homogenization parameters for the isolation of
cellulosic nanofibers using response surface methodology, Ind. Crop. Prod. 74 (2015) 381-
387.
[137] M.M. Haafiz, A. Hassan, H.A. Khalil, I. Khan, I. Inuwa, M.S. Islam, M.S. Hossain, M.
Syakir, M.N. Fazita, Bionanocomposite based on cellulose nanowhisker from oil palm
biomass-filled poly (lactic acid), Polym. Test. 48 (2015) 133-139.
[138] S.J. Eichhorn, Cellulose nanowhiskers: promising materials for advanced applications,
Soft Matter 7(2) (2011) 303-315.
[139] C.G. Li, X.P. Wang, L. Liu, J.H. Cui, R. Zhang, The effect of corn stalk
microcrystalline cellulose on thermal and mechanical properties of chitosan composites,
Applied Mechanics and Materials, Trans Tech Publ, 2012, pp. 1038-1041.
[140] A.M. El-Hadi, Influence of microcrystalline cellulose fiber (MCCF) on the morphology
of poly (3-hydroxybutyrate)(PHB), Colloid. Polym. Sci. 291(3) (2013) 743-756.
[141] B. Philipp, V. Jacopian, F. Loth, W. Hirte, G. Schulz, Influence of cellulose physical
structure on thermohydrolytic, hydrolytic, and enzymatic degradation of cellulose, Adv.
Chem. 181(6) (1979) 127-143.
[142] T.P. Nevell, W.R. Upton, The hydrolysis of cotton cellulose by hydrochloric acid in
benzene, Carbohyd. Res. 49 (1976) 163-174.
[143] L. Averous, F. Le Digabel, Properties of biocomposites based on lignocellulosic fillers,
Carbohyd. Polym. 66(4) (2006) 480-493.
[144] V. Alvarez, E. Rodriguez, A. Vázquez, Thermal degradation and decomposition of
jute/vinylester composites, J. Therm. Anal. Calorim. 85(2) (2006) 383-389.
41
[145] A. Bochek, I. Shevchuk, V. Lavrent'ev, Fabrication of microcrystalline and powdered
cellulose from short flax fiber and flax straw, Russ. J. Appl. Chem. 76(10) (2003) 1679-1682.
[146] S.H. Ghaffar, M. Fan, Lignin in straw and its applications as an adhesive, Int. J. Adhes.
Adhes. 48 (2014) 92-101.
[147] H.S. Abdul Khalil, M.S. Alwani, A.K.M. Omar, Chemical composition, anatomy, lignin
distribution, and cell wall structure of Malaysian plant waste fibers, BioResources 1(2) (2007)
220-232.
[148] K. Goda, M. Sreekala, A. Gomes, T. Kaji, J. Ohgi, Improvement of plant based natural
fibers for toughening green composites—Effect of load application during mercerization of
ramie fibers, Compos. Part. A Appl. Sci. Manuf. 37(12) (2006) 2213-2220.
[149] M.F. Rosa, B.s. Chiou, E.S. Medeiros, D.F. Wood, L.H. Mattoso, W.J. Orts, S.H.
Imam, Biodegradable composites based on starch/EVOH/glycerol blends and coconut fibers,
J. Appl. Polym. Sci. 111(2) (2009) 612-618.
[150] P. Joseph, K. Joseph, S. Thomas, C. Pillai, V. Prasad, G. Groeninckx, M. Sarkissova,
The thermal and crystallisation studies of short sisal fibre reinforced polypropylene
composites, Compos. Part. A Appl. Sci. Manuf. 34(3) (2003) 253-266.
[151] J.C. Caraschi, A.L. Leãto, Characterization of curaua fiber, Mol. Cryst. Liq. Cryst.
353(1) (2000) 149-152.
[152] N. Terinte, R. Ibbett, K.C. Schuster, Overview on native cellulose and microcrystalline
cellulose I structure studied by X-ray diffraction (WAXD): Comparison between
measurement techniques, Lenzinger Berichte 89 (2011) 118-131.
42
List of Figures
Fig. 2 Illustration of the annual number of scientific publications since 1983, using the search
terms “microcrystalline cellulose and composite”. Data analysis completed using
Scopus search system on 19 August, 2016.
Fig. 3 Structural levels of organization of cellulose from the source to the molecule.
Fig. 4 Molecular structure of cellulose showing the numbering of carbon atoms, the reducing
end with a hemiacetal, and the non-reducing end with a free hydroxyl at C4.
Fig. 5 Scheme of the common steps needed to produce MCC from cellulose source materials.
43
Fig.1
44
Fig.2
45
Fig.3
46
Fig.4
47
Fig.5
48
Fig.6
49
O
H3 O
O
O H
O O
O
(1)
Pathway I
H3O
(1) (2)
O
O H H
O O O
O O
OH O
O
H H
OH OH
O (3) O
O O
H2O H3O
Fig. 7
50
Table 1 Chemical composition of cellulose-containing materials.
51
Table 2 Different acid hydrolysis procedures to produce MCC from various sources.
52
Table 3 GPC analyses for the cellulose and microcrystalline cellulose samples.
Mn Mw
Sample DPI DPn DPw References
(RSD%) (RSD%)
Alfa-C 78,215 (10) 438,004 (9) 5.59 150 844 [67] (Trache et al., 2014)
Alfa-MCC 40,208 (8) 164,853 (6) 4.10 77 318 [67] (Trache et al., 2014)
C-MCC 39179 (7) 152,798 (8) 3.89 75 294 [67] (Trache et al., 2014)
53
Table 4 Comparison of thermal properties using TGA and DSC from different biomass.
TGA analysis
DSC
Onset Degradation Residual weight loss DTG peak
Samples analysis References
(°C) Temperature (°C) at 400 °C (%) temperature DTGmax
Tm (°C)
T10 T50 (°C)
B-MCC 251 285 320 22 320 N/A [76] (El-Sakhawy and Hassan,
2007)
RS-MCC 250 320 330 22 335 N/A [76] (El-Sakhawy and Hassan,
2007)
CS-MCC 270 260 325 28 320 N/A [76] (El-Sakhawy and Hassan,
2007)
J-MCC 250 280 475 61 280 N/A [52] (Das et al., 2010)
CS*- 285 340 475 57 340 N/A [52] (Das et al., 2010)
MCC
BW-MCC 230 290 350 18 N/A N/A [95] (Elanthikkal et al., 2010)
RH-MCC 223 270 290 23 283 N/A [61] (Adel et al., 2011)
BH-MCC 230 260 285 25 281 N/A [61] (Adel et al., 2011)
CC-MCC 270 280 340 22 N/A 335 [94] (Azubuike and Okhamafe,
2012a)
GS-MCC 260 275 320 20 N/A 320 [93] (Azubuike and Okhamafe,
2012b)
OPEFB- 275 293 327 20 326 N/A [23, 96] (Haafiz et al., 2013b,
MCC 2014)
FG-MCC 286 300 360 30 N/A N/A [63] (Kalita et al., 2013)
Alfa- 322 330 360 20 351 349 [67] (Trache et al., 2014)
MCC
C-MCC 258 310 333 19 332 N/A [100] (Karim et al., 2014)
Where C-MCC means (commercial MCC); OPEFB-MCC means (oil palm empty fruit brunch-MCC); RH-MCC means (rice hulls-MCC); BH-
MCC means (bean hulls-MCC); B-MCC means (bagasse-MCC); RS-MCC means (rice straw-MCC); CS-MCC means (cotton stalk-MCC); CC-
MCC means (corn cob-MCC) and GS-MCC means (groundnut shell-MCC); J-MCC means (jute-MCC); CS*-MCC means (cotton silver-MCC);
BW-MCC means (banana waste-MCC); FG-MCC means (fodder grass-MCC).
54
Table 5 XRD analysis for crystallinity index (CrI) and crystallite size (L) of microcrystalline
cellulose (MCC) from different biomass.
Where C-MCC means (commercial MCC); OPEFB-MCC means (oil palm empty fruit
brunch-MCC); RH-MCC means (rice hulls-MCC); BH-MCC means (bean hulls-MCC); B-
MCC means (bagasse-MCC); RS-MCC means (rice straw-MCC); CS-MCC means (cotton
stalk-MCC); CL-MCC means (cotton linters-MCC); FG-MCC means (fodder grass-MCC); J-
MCC means (jute-MCC); CC-MCC means (corn cob-MCC); GS-MCC means (groundnut
shell-MCC); CS*-MCC means (cotton silver-MCC); J*-MCC means (jute-MCC).
55
Table 6 SEM analysis for surface morphology of microcrystalline cellulose (MCC) from
different biomass.
Where C-MCC means (commercial MCC); OPEFB-MCC means (oil palm empty fruit
brunch-MCC); RH-MCC means (rice hulls-MCC); BH-MCC means (bean hulls-MCC); B-
MCC means (bagasse-MCC); RS-MCC means (rice straw-MCC); CS-MCC means (cotton
stalk-MCC); FG-MCC means (fodder grass-MCC); J-MCC means (jute-MCC); CS*-MCC
means (cotton silver-MCC); J*-MCC means (jute-MCC); BW-MCC means (banana waste-
MCC).
56