Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

11

Cellulose and its Derivatives:


Properties and Applications
Rafael de Avila Delucis1*, Pedro Henrique Gonzalez de Cademartori2,
André Ricardo Fajardo3 and Sandro Campos Amico4

Programa de Pós-Graduação em Ciência e Engenharia de Materiais (PPGCEM),


1

Universidade Federal de Pelotas (UFPEL), Pelotas, Brazil


2
Programa de Pós-Graduação em Engenharia Florestal (PPGEF), Programa de Pós-Graduação em
Engenharia e Ciência dos Materiais (PIPE), Universidade Federal do Paraná (UFPR), Curitiba, Brazil
3
Laboratório de Tecnologia e Desenvolvimento de Compósitos e Materiais Poliméricos (LaCoPol),
Universidade Federal de Pelotas (UFPel), Brazil
4
Programa de Pós-Graduação em Engenharia de Minas, Metalúrgica e de Materiais (PPGE3M),
Universidade Federal do Rio Grande do Sul (UFRGS), Porto Alegre, Brazil

Abstract
Cellulose, the most abundant organic compound or natural polymer on Earth, constitutes the
main support for several living vegetables, being also found in some algae and marine creatures
(tunicates). There are many cellulose types and configurations, and the chemical and physical
properties of cellulosic derivatives depend on the biosynthesis of the crystalline cellulose microfi-
brils and the extraction process. Chemical and mechanical methods for the deconstruction of the
cell wall can yield many products with high added-value and wide commercial application. This
chapter focuses on the characteristics of cellulose sources, their classification, constitution and
properties. It also describes common strategies to perform the extraction of cellulosic materials,
and finally gives a deeper look into most of the cellulosic derivatives, regarding their properties
and applications.

Keywords: Cellulose, cellulose derivatives, extraction, structure, properties, applications

11.1 Introduction
Cellulose, the most abundant organic compound or natural polymer on Earth, belongs to
a large group of biomolecules called carbohydrates and constitutes the main support for
several living vegetables. It is also present in some algae and marine creatures (tunicates),

*Corresponding author: rafael.delucis@ufpel.edu.br

Inamuddin, Mohd Imran Ahamed, Rajender Boddula and Tariq Altalhi (eds.) Polysaccharides: Properties and Applications,
(221–252) © 2021 Scrivener Publishing LLC

221
222 Polysaccharides

playing an important role in their survival. It is a stable polymer with high longitudinal
stiffness due to its inter- and intra-chain bonding network [1].
There are many cellulose types and configurations, but most chemical and physical
properties of cellulosic derivatives are dependent on two major factors: biosynthesis of the
cellulose microfibrils (that differ for each source material) and extraction process (encom-
passing pretreatments, disintegration, and/or deconstruction). Indeed, different extraction
approaches can yield a wide range of degree of polymerizations, geometric dimensions,
morphologies, surface charges and areas, porosities, crystallinities, thermal stabilities and
mechanical properties.
Cellulosic materials derived from lignocellulosic materials are obtained by top-down
chemical and/or mechanical treatment. In opposite, some celluloses are biosynthesized by
some bacteria, resulting in bacterial cellulose, which is directly available as a fibrous net-
work, free from pectin, hemicelluloses, lignin or other biogenic products.
Each cellulose derivative is tailored to address a particular application. For instance,
microcrystalline cellulose (MCC) is available as powdered and colloidal MCC. The for-
mer is a spray-dried product with high α-cellulose concentration largely used in the food
industry [2, 3], whereas colloidal MCC is water dispersed and have functional properties
somewhat similar to those of hydrocolloids. In addition, nanofibrillated celluloses (CNF)
are outstanding renewable materials extracted from lignocellulosic biomass that have been
recently used in a series of innovative products. They are capable of increasing mechanical
strength and stiffness, as well as transparency and the number of recycling cycles, of paper
and cellulose-based films [4]. Cellulose esters present high commercial importance mainly
due to characteristics such as neutral pH, the ability to be incorporated into translucent
films and low cost [5, 6]. And carboxymethyl cellulose (CMC), due to its low cost, rheolog-
ical properties, low environmental impact and solubility in hot and cold water, is the most
important cellulose ether, with many applications [7, 8].
An accurate overview on cellulose and its derivatives is exhaustive, and a wide and
non-standardized terminology is found in the literature. Indeed, this is one of the most
indexed term in several fields, such as utilization of forest-based resources, materials
science and engineering and biochemistry. To tackle this challenge, this chapter was
divided into sections which separately describe the main raw materials used, the compo-
sition and chemical structure of lignocellulosic materials, the chemical and mechanical
methods for cellulose extraction, the cellulose products and its derivatives, and the main
applications.

11.2 Main Raw Materials


The demand for cellulose and its derivatives is clearly linked to the need for new uses
of some lignocellulosic biomasses, i.e. the exploitation of natural resources from plants,
including waste leftovers from various industrial activities. The leftovers, in particular, have
a special appeal related to sustainability, since this possible feedstock is renewable and pres-
ents huge availability.
When certain biomass is considered for a particular application, its economic feasibility
must be taken into account, mainly based on costs related to equipment, taxes, profit, labor
and raw material. Regarding commodities, quantity, seasonality and location of the biomass
Cellulose and its Derivatives 223

are also important. For biomasses spread over a large area, which is the case in many coun-
tries, transportation may become an issue due to their low density and the need to transfer
large volumes. In most cases, competitiveness against energy production and other appli­
cations should be identified [9].
a) Wood: Wood is the first cellulosic raw material. In 1838, Anselme Payen (a French
scientist) pioneered the isolation of cellulosic material from the cell wall of wood using
nitric acid. Currently, wood is the main industrially used feedstock for cellulosic pulp
production and has been exploited since the 80s [10]. In natural woods, cellulose plays a
mechanical role since cellulose fibers act as reinforcements, depending on features of the
microfibrillar layers belonging to their secondary structure.
From a chemical standpoint, there are significant differences between wood species.
Following the so-called Engler system (idealized by Adolf Engler, a German botanist),
woods can be classified in two groups: softwoods (i.e. gymnosperms) and hardwoods (i.e.
angiosperms). This way, woods from the same botanical group present similar chemical
compositions. In general, the cellulose content of a hardwood fiber is higher than that of a
softwood fiber and, because of that, hardwoods are historically more exploited for cellulose
extraction. Either way, most cellulosic products are extracted via organosolv pulping meth-
ods [11]. Nevertheless, intense scientific efforts over the last decades in the field of seed
genetics and silvicultural interventions have been able to increase productivity of cellulose
extraction from hardwoods.
b) Forestry wastes: More than 200 billion m³ of forestry wastes are annually generated
worldwide and some of that become urban residues and occupy large areas in landfills and
dumps, leading to concerns related to the pollution of air, soil and groundwater [9]. To
access the wood, bark, fruits, small branches and leaves, which compose up to 40% of the
total tree weight, are often left on the ground in forests. However, the pruning and cleaning
of forest wastes represent a source of biomass that can be used in the production of waxes,
essential oils, plant extracts and feed [11]. Industrial wastes, e.g. from the solid wood mill,
may comprise as much as 30% of woody raw material and can also be exploited, also help-
ing to minimize disposal issues.
c) Other vegetable fibers: There are numerous vegetable fibers, and they can be
grouped on the basis of their origin and location in the plant. According to Jawaid and
Khalil [12], vegetable fibers can be obtained from seeds (from flosses of seeds, like cotton),
fruits (i.e. extracted from the pulp or rusk, like coir and oil palm), leaves (i.e. from the leaf
of ribs, e.g. banana, sisal and raphia), grass/reeds (e.g. bamboo, corn and bagasse), wood
and roots (i.e. from hardwood and softwood trees), bast (such as hemp, flax, and kenaf) and
stalk (i.e. from straws of cultivated plants like oat, wheat, barley, rice or maize). The choice
of a particular fiber is basically dependent on its cost, availability in each region and specific
properties required.
d) Agricultural Crops: Crop residues can help to minimize CO2 emission and defor-
estation, relieving pressure on forest resources. Several authors already referred to these
resources as feedstock for the production of paper pulp and cellulose fiber-based products
since the utilization of non-edible lignocellulosic materials is also a strategy for producing
fuels and chemicals, alternatively to fossil resources [13].
According to Garcia et al. [11], crop residues used to obtain cellulose products include
wheat straw, sugar beet pulp, potato tubers, rutabaga, soy hulls, soybean stock, sorghum
fibers, soybean straw, rice straw, banana fibers, cassava, cornstalks corn husks, pineapple
224 Polysaccharides

leaves, grape hulls, oil palm biomass and bagasse. Herbaceous residues, from stems or stalks
usually abandoned during harvesting, have recently been considered for bioethanol pro-
duction. Non-wood plants typically have less lignin than wood and may demand less chem-
icals to be bleached, which is an important economic advantage. Also, alkali pre-treatments
tend to be more efficient in the production of cellulosic derivatives from both agricultural
residues and herbaceous crops [14].
e) Algae: Several algae species are able to produce cellulose microfibrils within their
cell wall, such as yellow-green, green, red and gray ones. They are commonly found in
freshwater and seawater ecosystems. Certain algae grow excessively sometimes, which
leads to an ecological imbalance and difficulties to the local population [15], and a bal-
anced exploitation of these plants can help to solve this issue. Green algae are the most
exploited, including Micrasterias rotata, Micrasterias denticulata, Valonia, Boergesenia and
Cladophora.
Each aquatic microorganism may yield a different cellulosic material since there are
important differences regarding the biosynthesis processes. Compared to traditional
sources, cellulose from algae is easily accessed due to the lack of typical physicochemical
barriers (like a lignocellulosic structure), which prevents the use of several chemical treat-
ments [16]. Certain algae can be largely obtained as a by-product of agar production. This
source could partially meet the demand for micro- and nanocellulosic materials since the
agar market is growing in niches like culinary, vegetable biology, molecular biology and
microbiology [17]. The research in this field is mostly related to genetic engineering, for
the development of improved algae able to produce useful cellulosic substances on acces-
sible culture media. Cellulosic substances from algae have high crystallinity, and also high
pore volume, large surface area and high thermal degradation resistance, depending on the
chemistry of the culture medium [18, 19].
f) Tunicates: Tunicates are marine invertebrate filter feeders, also called sea squirts.
They may look like small and colored blobs and are usually found attached to rocks, docks
or the hull of boats. The natural tunic of these sea animals is composed of fine networks of
cellulose fibers, which act as a protective barrier, in what occurs a certain chemical puri-
fication of the water. The cellulosic substances extracted from tunicates are called tunicin
cellulose. They can be highly crystalline and possess high biocompatibility; the latter drives
their use for the production of membranes for bone regeneration and the reconstruction of
other tissues [20].

11.3 Composition and Chemical Structure of Lignocellulosic


Materials
Lignocellulose refers to three-dimensional biopolymers. Most of the plants present this
chemical structure, marked by an interconnected structure of cellulose, lignin and hemicel-
luloses, with a small amount of extractives and inorganics. These lignocellulosic materials
are likely the most important renewable raw materials and many alternatives naturally grow
worldwide.
Several synthetic materials have been replaced by alternative ones obtained from ligno-
cellulosic sources. This movement is associated with the relatively recent recognition of the
Cellulose and its Derivatives 225

interdependence between economic and natural ecosystems, and much of the importance
attributed to lignocellulosic materials is associated with the exploitation of cellulose and its
derivatives.
In a plant cell, the cellulose chains are organized as crystalline (highly ordered) regions
interspersed by amorphous or semi-crystalline (less ordered) segments. The latter is
believed to be found on the fibril surface or in amorphous segments of fibril cellulose [21].
In general, the properties of extracted cellulosic materials depend on anatomical structure,
chemical composition, cell dimension and microfibrillar angle of the lignocellulosic source.
Climatic conditions, age and the degradation process also affect the fibers structure and
their chemical composition [22].
To describe the lignocellulose ultrastructure in detail is challenging. Thus, natural
fibers are usually designated as a natural composite, in which rigid cellulose fibers are
immersed into a flexible matrix constituted of hemicelluloses and lignin. The cell wall
structure of nearly all lignocellulosic plants has been well-studied and several models have
been reported since the 50s based on light and electron microscopy and X-ray diffraction
results. The cell wall is a multilayer system with 3–4 individual layers of crystalline cel-
lulose microfibrils in a particular orientation, which govern most mechanical features of
natural fibers [23].
Regarding chemical composition, the content of each macromolecular compound varies
in a wide range. Regardless of the physical and chemical treatments performed on natural
fibers to extract cellulose, it is important to know the content and distribution of all chemi-
cal compounds of their cell wall so as to control particle size and crystallinity of each cellu-
losic product. Table 11.1 presents a wide compilation of the reported chemical composition
of some of these resources.

11.4 Cellulose: Chemical Backbone and Crystalline Formats


Cellulose is a linear homopolysaccharide composed of glucose units connected to each
other by β-1,4-glycosidic bonds, producing very long linear and unbranched chains of
d-glucopyranose conformed in a flat ribbon-like structure [26]. Depending on the source,
the number of linear units can reach 1,000 [17]. The glucan ring is folded into an arm-
chair configuration with all methine groups (C–H) located on axial positions and hydroxyl
groups (O–H) on equatorial positions [27]. The chains are tightly packed in crystalline
microfibrils with a degree of polymerization within 300–1,700 for wood pulp and up to
10,000 for cotton and flax [28].
Microfibrils are 3–5 nm in diameter, being ordered as three-dimensional crystals,
although recent research indicates that the fibrils are not fully crystalline [26]. The final
length of the fibril is given by covalent bonds along the cellulose chains and the third
dimension exists due to van der Waals bonds, which act as bridges between adjacent cellu-
lose sheets linked together as fibril-like structures [1].
In pristine lignocellulosic materials, the crystalline domains and the less ordered seg-
ments (called amorphous regions) are present in variable ratios, depending on the plant
species, growing conditions, among other factors. The crystalline cellulose presents four
major polymorphs, Cellulose I, Cellulose II, Cellulose III, and Cellulose IV, described
below. All allomorphs are chemically identical, but vary in the 3-D arrangement of cellulose
226

Table 11.1 Natural resources from different origins and the content, in wt%, of the main components.
Raw material Classification Cellulose Hemicelluloses Lignin Organic extractives Inorganics Ref.
Hardwood Wood 31–64 25–40 14–34 0.1–7.7 <1 [24]
Softwood Wood 30–60 20–30 21–37 0.2–8.5 <1 [24]
Polysaccharides

Apple tree pruning Forestry waste 27.3 30.0 26.7 9.8 2.8 [11]
Eucalyptus bark Forestry waste 47.2 22.0 21.9 1.3 4.0 [11]
Olive tree pruning Forestry waste 32.6 28.6 28.0 3.9 2.8 [11]
Jute Vegetable fiber 61–71 14–20 12–13 0.5 n.d. [22]
Flax Vegetable fiber 71 18–21 n.d. n.d. n.d. [22]
Ramie Vegetable fiber 68.6–76.2 13–16 0.6–0.7 0.3 n.d. [22]
Sisal Vegetable fiber 65 12 9.9 2 n.d. [22]
Abaca Vegetable fiber 56–63 20–25 7-9 3 n.d. [22]
Hemp Vegetable fiber 68 15 10 0.8 n.d. [22]
Kenaf Vegetable fiber 72 20–21 n.d. n.d. n.d. [22]
Coir Vegetable fiber 32–43 0.1–0.3 40-45 n.d. n.d. [22]
Pineapple Vegetable fiber 81 n.d. 12.7 n.d. n.d. [22]
Curaua Vegetable fiber 73.6 9.9 7.5 n.d. n.d. [22]
(Continued)
Table 11.1 Natural resources from different origins and the content, in wt%, of the main components. (Continued)
Raw material Classification Cellulose Hemicelluloses Lignin Organic extractives Inorganics Ref.
Okra Agricultural Crop 36.3 28.7 17.9 3.0 4.4 [11]
Oil Palm Agricultural Crop 65 n.d. 29 n.d. n.d. [22]
Grapevine stems Agricultural Crop 43.1 19.4 26.6 5.3 3.9 [11]
Bean stalks Agricultural Crop 31.1 26.0 16.7 5.9 7.3 [11]
Chilli stalks Agricultural Crop 37.5 28.3 17.3 3.0 8.0 [11]
Pepper stalks Agricultural Crop 35.7 26.2 18.3 3.0 8.8 [11]
Cotton stalks Agricultural Crop 41.7 27.3 18.7 4.7 4.5 [11]
Cotton Agricultural Crop 82.7 5.7 n.d. 6.3 n.d. [22]
Rice husk Agricultural Crop 35–45 19–25 20 14–17 n.d. [22]
Rice straw Agricultural Crop 41–57 33 8–19 8–38 n.d. [22]
Wheat straw Agricultural Crop 38–45 15–31 12–20 n.d. n.d. [22]
Miscanthus Agricultural Crop 36.2 30.8 15.5 7.9 4.2 [11]
Paulownia Agricultural Crop 37.7 31.9 18.9 4.5 1.8 [11]
Algae biomass Algae 7.1 16.3 1.5 n.d. 1.8 [25]
n.d.: not determined.
Cellulose and its Derivatives
227
228 Polysaccharides

chains inside the unit cell. The crystal structures mostly differ in the CH2OH conformation
and chain polarity [26].
a) Cellulose I: This is the crystalline phase from the cellulose chain known as “nat-
ural” or “native” cellulose. In living plants, cellulose I is the most widespread crystal-
line form, although it can also be naturally produced by algae, tunicates and bacteria. It
is a thermodynamically metastable structure and has the highest axial elastic modulus
among all polymorphs since its crystallites follow the parallel chain polarity within the
unit cell.
Cellulose I has a triclinic structure (Iα) or a monoclinic structure (Iβ) [29] and coexist
in variable ratios depending on the cellulose source. Both. According to Makarem et al.
[26], celluloses Iα and Iβ present the largest crystals in nature. Phase Iα, with one chain, is
dominant in most algae and bacteria, whereas cellulose Iβ has two parallel chains and is the
most important polymorph for the majority of plants and tunicates [30].
Partial conversion of cellulose Iα into Iβ may be achieved by an activated process, that
requires high temperatures and slow largescale molecular motions [31, 32]. The conversion
of cellulose Iα into Iβ is considered irreversible since the Iα phase has higher free energy
than the Iβ phase [32]. Cellulose Iβ allomorph is also mostly found in tunicates [17].
b) Cellulose II: Cellulose II has the highest chemical reactivity and thermodynamic
stability among the cellulose polymorphs due to its monoclinic structure, with anti-parallel
chain polarity and a huge number of hydrogen bonds in the c direction in opposite to van
der Waals bonds found in other polymorphs [29, 33].
Cellulose II is quite useful, with a wide range of applications, including chemical indus-
tries in general, to produce rayon (among other synthetic textile fibers) and cellophane. It
can be produced from Cellulose I via regeneration followed by mercerization (i.e. treat-
ment with aqueous sodium hydroxide), which is not a reversible process and several inter-
mediate states may exist, depending on the NaOH concentration and temperature [34].
These mechanisms have not been fully clarified yet, especially the inversion of chain polar-
ity [26].
c) Cellulose III: It is obtained from soaking cellulose, in a way that an anhydrous
ammonia medium at about −80 °C is removed by washing and evaporation or vacuum
drying [35]. N-methylmorpholine-N-oxide, ethylenediamine and n-butylamine can
replace ammonia [36]. Thus, celluloses I and II are converted into celluloses III1 and III2,
respectively. Cellulose III can be back converted to their original forms by rehydration
[35, 26].
d) Cellulose IV: It is obtained by soaking cellulose in glycerol atmosphere at about 200
°C, with subsequent removal by washing with 2-propanol and water. This way, celluloses I
and II are converted into celluloses IV1 and IV2, respectively. Nevertheless, Wada et al. [35],
based on NMR, IR and XRD results, proposed that cellulose IV is not an actual allomorph.

11.5 Cellulose Extraction


11.5.1 Mechanical Methods
Mechanical treatments are among the most used methods to extract and/or isolate
cellulose from cellulose fibers, in macro-, micro- or nano-scale. Mechanical methods
Cellulose and its Derivatives 229

represent a variety of treatments widely used in papermaking factories and/or those


focused on developing new high added-value products, such as biocomposites or nano-
cellulose [37]. The description below of the common mechanical methods focuses on
the use of cellulose fibers as raw material to extract and/or isolate cellulose at the micro-
and nano-scale.
a) Beating: This refining process corresponds to the mechanical treatment of cellulose
chemical pulp in the presence of water [38], producing fine, internal and external fibrilla-
tion, fiber cutting and fiber deformation. Internal fibrillation is the most important primary
effect since it softens the fibers and improves their flexibility.
Common beaters used to refine cellulose pulp are PFI mill, Lampén mill and Valley
Hollander, where the former has efficient removal of material from the fibers’ surface, and
the latter is efficient in inner fibrillation and fiber cutting. The selection of the refining sys-
tem basically depends on the end-products and the type of pulp available. Chemical pulp
is widely treated by beating before its conversion to paper, since the mechanical processing
provides desirable papermaking properties, especially drainage resistance, tensile strength,
tensile stiffness, internal bonding strength and fracture toughness. Beating can also prevent
undesirable characteristics for paper sheets, such as weak bonding and non-uniform dis-
tribution of fibers. For instance, Chen et al. [38] reported positive aspects of PFI beating
in recycled eucalypt cellulose fibers. Longer beating times yielded higher tensile strength,
breaking length and stretch. Beating can also be applied as a pretreatment, promoting pro-
cessing efficiency in the subsequent isolation of cellulose nanofibrils via microfluidization,
as reported in Ref. [39].
b) High pressure homogenization (HPH): The HPH method uses high pressure (50–
2,000 MPa) to pass cellulose solution through a tiny nozzle [40], promoting a high degree
of fibrillation and producing microfibrillated cellulose (MFC) [41] and cellulose nanofibrils
(CNF), either at lab-scale or large-scale production. It is considered an efficient method
for cellulose fibers [42], including cotton [43], bamboo [44] and wood [45]. The degree of
cellulose fibrillation varies with applied pressure and number of homogenization cycles
[46], and this may translate into high energy demand [41], not necessarily promoting better
fibrillation of cellulose [47]. This can be overcome by using chemical purification or pre-
treatments in the cellulosic pulp.
c) Grinding: Different kinds of grinding have been used to degrade the cell wall and
separate cellulose fibers into micro- and/or nanofibers. Masuko© [48] pioneered an equip-
ment to process the cellulose slurry between two grind stones (disks), one of them is static
and the other rotates at 1,500 rpm, generating shearing forces able to disrupt hydrogen
bonding and the cell wall of fibers [42]. Many factors influence the grinding process, such
as the number of passes, the distance between grinding stones and characteristics of the
raw materials, resulting in nanofibrils with distinct physical and mechanical characteris-
tics, sometimes reaching enhanced mechanical properties [49]. Iwamoto et al. [50] recom-
mended 10 passes in the grinder to produce a TEMPO-oxidized nanocellulose suspension
of Halocynthia papillosa with uniform-sized fibrils.
Cellulose nanofibrils of 4–40 nm diameter were isolated from eucalypt wood by grinding
with 25 passes at 1,500 rpm [51]. They also observed that above ten passes in the grinder
could turn nanocellulose more accessible to thermal degradation, while 20 passes can
improve mechanical properties of nanopapers due to better homogenization and asso-
ciation of nanofibrils. Regarding the raw material, high content of hydrophobic lignin in
230 Polysaccharides

the CNF suspension can significantly influence morphology of fibrils with a tendency for
agglomeration [52].
d) Cryocrushing: This mechanical treatment fibrillates cellulose frozen in liquid
nitrogen [53]. Cellulose fibers are first soaked in water then frozen with liquid nitro-
gen and later crushed using mortar/pestle. The freezing step produces ice crystals in
the cell wall that exert mechanical pressure on frozen fibers when they are subjected to
the impact by mortar/pestle [54], which is essential to break the microfibrils and pro-
duce individual fibrils of low diameter [55]. The cryocrushed cellulose fibrils are usually
soaked in water for mechanical fibrillation using disintegrator systems or ultrasonica-
tors, which can reduce further diameter and length of the fibrils, reaching a nanometric
size [53, 56].
Cryocrushing is commonly used individually or in combination with other methods to
isolate MFC and/or CNF from biomasses such as wheat straw and soy hulls [53], aquatic
weed plant water hyacinth [57], flax fibers (Linum usitatissimum) [56], soybean pods [58],
kenaf bast (Hibiscus cannabinus) [59] and black spruce wood pulp [55]. CNF isolated from
flax, hemp, and rutabaga fibers can reach 5–80 nm diameter depending on the operational
conditions, as observed by Bhatnagar and Sain [60]. Cryocrushing can also be used as a
pretreatment, as in [61], who observed an increase in saccharification of holocellulose
from 4.3 to 54.1% for eucalypt biomass and from 3.9 to 40.6% for rice hull biomass [61].
Nevertheless, cryocrushing process has low productivity and demands high energy, which
hampers its wide application in commercial-scale [62].
e) Microfluidization: This mechanical treatment is similar to HPH due to the use of
high pressure to fibrillate cellulose, but makes use of an intensifier pump for better results
[63, 64]. Usually, the aqueous suspension of cellulose fibers is exposed to high pressure
(up to 276 MPa) at a constant shear rate (up to 107 s−1) in a z-shaped chamber in which
fibrillation occurs, The shear forces and impact of cellulose fibers with the wall produces
slurries of cellulose nanofibers with a gel-like aspect [39]. Microfluidization yields more
uniformly-sized cellulose nanofibers [65], but different number of passes are required in
the mechanical disintegration depending on the raw material, which can limit the scal-
ing-up of the process, due to the increase in energy consumption [48].
Microfluidization of cellulose fibers can be facilitated by enzymatic pretreatment with
endoglucanases, which also results in better dispersion and transparency of fibrils’ sus-
pension, as well as lower energy consumption [66]. It can also represent a useful tool
to break down CNC aggregates and improve CNC dispersion in a polymer matrix [67].
These microfluidized nanofibers can reach 5–100 nm diameter [39, 66, 68, 69], depend-
ing on operational conditions. For instance, the number of passes influences homoge-
neity of MFC/CNF suspensions [68] and tensile strength and modulus of CNC/chitosan
films [67].
f) Ball milling: Ball milling is based on the collision of rigid balls—like ceramic, stain-
less steel and flint pebbles balls—in a container generating high pressure, resulting in an
increase in surface area [70]. This process is usually applied as pretreatment or in combi-
nation with other methods to modify cellulose fibrils and, more recently, to obtain cellu-
lose nanocrystals or nanofibrils. Ball milling has many advantages when used for cellulose,
being an easy, economic and environmentally friendly process that does not use organic
solvents. Some disadvantages include irregular shape of nanomaterials produced, long time
required in some cases and possible contamination [71].
Cellulose and its Derivatives 231

Regarding the effect of ball milling under wet conditions, Ago et al. [72] observed vari-
able morphological aspect and crystallinity of microcrystalline cellulose depending on
the media used (water, toluene or 1-butanol), which was attributed to its influence on the
hydrogen bonding among cellulose fibrils. Regarding the isolation of CNF or production
of CNC, ball milling should be used with caution, because processing parameters like
speed and time of milling can influence morphology and crystallinity of CNC produced
by acid hydrolysis [71]. Qua et al. [73] used ball milling (600 rpm for 2 h) as pretreatment
followed by acid hydrolysis and ultrasonication to isolate CNF from microcrystalline cel-
lulose and flax fibers. Yang et al. [74] also applied ball milling (1,000 rpm for 24 h) as
pretreatment to produce CNC via acid hydrolysis, but used linter as raw material, and
reported a reduction in cellulose crystallinity to 21.9% and increased cellulose accessi-
bility. They also observed that ball milling is a good pretreatment before acid hydroly-
sis in the production of sphere-like CNC. This technique can also be a pretreatment in
the isolation of CNF in wet conditions since it causes fiber swelling, which facilitates
mechanical fibrillation [71]. In another study, Nge et al. [75] obtained spherical CNC
using high-pressure homogenization and ball milling, or rod-like CNC for high-pressure
homogenization.
g) Ultrasonication: Sonication refers to the application of sound energy to materials
dispersed in a liquid using an ultrasonic probe or a bath system. The term ultrasonication,
also called high-intensity sonication, is used when the applied frequency is higher than 20
kHz. Ultrasonication is commonly used as a pretreatment, combined with other methods,
to modify cellulose fibrils or to isolate/produce CNF or CNC. But it can also be used as the
main treatment, as in the work of Chen and coworkers [76], who isolated CNF from poplar
wood with 5–20 nm diameter and 69.34% crystallinity by chemical pretreatment followed
by ultrasonication (at 20–25 kHz frequency, 400–1,200 W power for 30 min). The high-
intensity energy applied is transferred to the cellulose chains via cavitation, which results in
fibrillation [77]. Similar characteristics of nanofibrils diameter and efficiency of ultrasoni-
cation to obtain nanocellulose were reported in [78], whereas [79] used this technique for
the dispersion of CNF in the cement slurry.

11.5.2 Chemical Methods


Chemical methods are an efficient way to facilitate fibrillation of cellulose fibers and focus
on the partial or complete removal of hemicelluloses and lignin [80]. Pretreatments of
cellulose fibers to extract and/or isolate cellulose at micro- and nano-scale enable 20–30
times reduction in energy required [42] and the most common methods are described
below.
a) Acid hydrolysis: Acid hydrolysis can be applied as a pretreatment or a single
treatment to extract and/or isolate CNC or cellulose nanowhiskers, with 4–20 nm diam-
eter, 50–60 nm length and specific gravity within 1.57–1.59 g cm−3 [81]. IN fact, this is
the most known way to produce CNC from any biomass. In general, acid hydrolysis is
performed using sulfuric acid (55–65 wt%) or hydrochloric acid (25–32 wt%) at moder-
ate temperatures, for the selective degradation of the amorphous region of cellulose fibers.
Hydrochloric acid produces CNC with minimal surface charge, while sulfuric acid (H2SO4)
results in a highly stable CNC aqueous suspension as a consequence of the esterification of
hydroxyl groups on the surface, producing charged sulfate groups [81, 82]. Indeed, H2SO4
232 Polysaccharides

is the most used media for the acid hydrolysis of cellulose and production of CNC [83].
Post-treatments like mechanical or ultrasound disintegration are usually applied to avoid
agglomeration of cellulose nanocrystals [84].
Heterogeneous acid hydrolysis of cellulose results in micro and nanoparticles with high
crystallinity [85]. Indeed, acid hydrolysis has been widely applied to produce CNC through
chemical degradation of noncrystalline region of cellulose fibers [81, 86–89]. For instance,
dos Santos et al. [90] produced CNC from an aqueous suspension of microcrystalline cel-
lulose using 64% H2SO4 at 40 °C for 2 h. The CNC was later incorporated into PLA (poly-
lactic acid) films and acted as nucleating agent for the PLA crystallization, also improving
mechanical properties. The optimal acid hydrolysis conditions depend on parameters like
acid concentration, and reaction temperature and time. A long time leads to full hydrolysis
[81] and yields shorter nanocrystals [91].
b) Enzymatic hydrolysis: Pretreatments with enzymatic hydrolysis result in degra-
dation of the amorphous phase of cellulose fibers [40]. This process may reduce energy
consumption in the production of CNF and/or MFC, as well as facilitate its mechanical
fibrillation. Also, this pretreatment is less aggressive than acid hydrolysis [92, 93], and pro-
motes cell wall delamination of the fibers [48] with little effect on the size of isolated cel-
lulose nanofibers [45]. Among the many enzymes used, ligninases and xylanases degrade
hemicelluloses and lignin, while cellulases hydrolyze the cellulose fibers [94]. Enzymatic
hydrolysis is largely used prior to the mechanical treatment of cellulose fibers, and com-
bined treatments may be an effective way to control fibrillation and produce networks of
long and intertwined cellulose I elements.
c) TEMPO Oxidation pretreatment: Oxidation pretreatment of cellulose fibers refers
to the modification of the cellulose hydroxyl groups with specific agents, such as TEMPO
(2,2,6,6-tetramethylpiperidine-N-oxyl). TEMPO in combination with NaClO can acceler-
ate the oxidation reaction of primary alcohol groups of cellulose [95]. Like enzymatic pre-
treatments, TEMPO oxidation can facilitate fibrillation and reduce the energy required to
isolate nanocellulose [96].
d) Carboxymethylation: Carboxymethylation can be used as a pretreatment to
increase anionic charges by forming carboxyl groups on the nanocellulose surface [48].
Partial carboxymethylation can be useful to create an electrostatic repulsive force between
cellulose fibrils. The efficiency of carboxymethylation was studied by Chinga-Carrasco
et al. [95, 97] who reported a significant effect on fibrillation, facilitating nanofibrils’
individualization.

11.6 Cellulose Products and its Derivatives


Chemical and mechanical methods for the deconstruction of the cell wall can yield many
products with high added-value and wide commercial application. The range of derivatives
obtained depends on source, origin, maturity and processing conditions, and commonly
vary in degree of geometric dimensions, polymerization, surface charge and area, morphol-
ogy, crystallinity, thermal stability, porosity and mechanical properties. The main cellulose
products and its derivatives are described below in relation to their processing, properties
and applications, along with the applied nomenclature.
Cellulose and its Derivatives 233

a) Cellulose nanocrystals (CNC): Cellulose nanocrystals, or nanocrystalline cellulose,


are defined as high-purity single crystals [98], similarly to rod-like nanoparticles, produced
by acid hydrolysis under controlled time, stirring and temperature conditions, and con-
sidering the acid type and acid-cellulose fiber ratio [83, 48]. CNC has high crystallinity,
around 90% [99], and its geometry also depends on the cellulose origin, e.g. softwood pulp,
cotton, tunicate and microcrystalline or bacterial cellulose [98, 100]. Typical CNC geome-
try is a few nanometers (5–20 nm) in width, and with length ranging from 100 to 500 nm
[99, 100]. In general, CNC presents interesting properties for high value-added materials,
such as modulus in the 100–150 GPa range, high crystallinity and surface area, and low
density [99].
b) Cellulose nanofibrils (CNF) and microfibrillated cellulose (MFC): Many labels
are used in the literature for CNF and MFC, including microfibril, microfibrillar cellulose,
microfibril aggregates, nanofiber, nanofibrillar and fibril aggregates. Cellulose nanofibrils
(CNF) and microfibrillated cellulose (MFC) are micrometer-long entangled fibrils with
amorphous and crystalline regions [63, 99] extracted from a variety of sources, such as
wood, cotton and any natural fiber. The most correct terms—CNF or MFC—are defined
based on the particle size in the aqueous suspension. When most fibrils are micrometric,
the term MFC is more appropriate [101]. The typical CNF/MFC diameter is in the 3–50 nm
range [100].
Extraction and isolation of CNF/MFC occurs by mechanical methods—grinding,
milling or homogenization—applied individually or in combination with chemical or
enzymatic methods. The entangled long fibers result in a viscous aqueous suspension,
usually with at low concentration (1–3%) [99], although industrial processes reach up
to 10%.
The significant interest in CNF/MFC for high added-value materials is driven by its dis-
tinctive properties, including low density, good reactivity with other materials and chem-
icals, high mechanical strength and surface area, chemical inertness, biocompatibility
and interesting barrier and optical properties [99]. For instance, CNF/MFC has attracted
the interest of the pulp and paper sector as a new paper component due to the cited char-
acteristics, as well as the inherent tendency to form entangled three-dimensional network
and the high interaction with cellulose fibers via hydrogen bonds [101]. CNF/MFC have
also been used in applications like food packing, nanocomposites, paper, printing, bio-
medical materials [63], reinforcements in composites [102] and production of aerogels
[103].
c) Nanowhiskers: Cellulose whiskers are produced by acid hydrolysis, where the
amorphous regions of cellulose microfibrils undergo transverse cleavage and, after soni-
cation, result in a rod-like material. Their typical diameter is around 2–20 nm, with a wide
length range (typically 100–600 nm), sometimes even above 1 μm [80]. Because of the
near-perfect crystalline arrangement of whiskers, this type of nanocellulose shows high
modulus. Alternative terms used for “cellulose whiskers” are: nanorods, nanowhiskers,
nanowires, rod-like cellulose crystals.
d) Bacterial cellulose (BC): Some types of bacteria can produce cellulose from some
plants. The biosynthesis of bacterial cellulose, also called as biocellulose or microbial cellu-
lose [104], occurs when cellulosic substances are extracellularly secreted from bacteria in a
suitable culture media. These bacteria are naturally found where sugar fermentation occurs
234 Polysaccharides

in plant carbohydrates, as unpasteurized or non-sterilized beer, juice and wine, or in dam-


aged flowers and fruits.
There are different genres of bacteria, including Alcaligenes, Acetobacter, Agrobacterium,
Rhizobium, Pseudomonas and Sarcina. Acetobacter xylinum is the most efficient bacteria,
capable of producing a certain nanocellulose at nano to macro length scales depending on
the biosynthesis route [104]. In general, bacterial cellulose has distinctive characteristics
including extremely fine and pure network structure with a degree of polymerization of up
to 8,000, high mechanical strength, water holding capability and high biocompatibility, an
important feature for its use as a biomaterial [28].
e) Cellulose esters: Cellulose acetate (CA) is considered the most important cellulose
ester. It is a thermoplastic polymer produced via esterification of cellulose fibers [105] with
different degrees of substation (DS), commonly 2.5 to combine good solubility in common
solvents and adequate melt properties and molecular weight for applications such as films,
plastics and textiles [106].
One of the main obstacles in the production of plastics with CA is the melting tem-
perature required for processing is higher than its decomposition temperature [105].
Also, it has low resistance to organic solvents, poor creep resistance and easy growth of
­microorganisms—mainly fungi—on its surface [107, 108]. Some of these limitations can be
mitigated through the production of nanocomposites, such as those with highly branched
alkoxysilane [109] or reinforced with clay [105]. Even so, CA is widely applied in the devel-
opment of new materials due to interesting properties, especially its biodegradability [106],
antifouling and hydrophilicity [110]. The most common application is in the production of
ultrafiltration membranes [111] for wastewater treatment, water desalination, water purifi-
cation and concentration of fruit juices [106, 112, 113].
Cellulose sulfate (CS) and cellulose nitrate (CN) are also important cellulose esters.
CS is a polyanionic carbohydrate polymer produced through the complete or partial
substitution of hydroxyl groups (OH) in the chemical chain of cellulose fibers (mainly
cotton and cotton linter) for sulfate groups (SO3H) using reagents like H2SO4 [6, 114],
pyridine, chlorosulfuric acid and N,N-dimethylformamide [115]. CS can be produced
via heterogeneous or homogenous sulfation, based on the selected reagents. The former
is a direct sulfonating process [116] using isopropylalcohol and sulfuric acid [117]. The
typical DS is 0.7 [116]. Pattern of substitution for cellulose derivates—like CS—with DS
below 3 depends on the distribution within the anhydroglucose unit at the O-2, O-3 and
O-6 sites along the polymer chain and between polymer chains [118]. The conversion
ratio of cellulose into CS and the yield of reaction increase with the concentration of
H2SO4 [117], but excess H2SO4 can result in a significant reduction in molecular weight
[114].
CS is particularly used for medical and biotechnological technologies, since this deriva-
tive of cellulose has hydrophilicity, water-solubility, biocompatibility and biodegradability
[114, 119]. CS can also be used to produce films incorporated with mustard essential oil and
β-cyclodextrin to improve antimicrobial activity [120] and NaCS films incorporated with
starch with good transparency and anti-oil properties [5]. As another example, Chen et al.
[6] used CS as raw material together with glycerol for the production of a homogeneous,
flexible and transparent film with interesting water barrier and mechanical properties, that
can improve the shelf life of coated bananas [6]. CS can also interact with chitosan since
Cellulose and its Derivatives 235

they (NaCS and chitosan) form a water-insoluble complex able to produce microcapsules
for oral controlled release systems [121]. Despite the interest and acceptance of CS for many
high added-value applications, its production generates a considerable amount of waste
after sulfation reaction (approximately 75% of the reaction solution), resulting in high costs
and non-adequate environmental conditions [117]. Also, although acidic conditions lead to
considerable degradation of cellulose, they can avoid the use of toxic reagents, facilitating
the industrial production of CS.
CN, another widely used cellulose ester, is a biocompatible polymer [122–124]. It is pro-
duced from the controlled reaction of cellulose and nitric acid, where the hydroxyl groups
are substituted for nitrate groups [125]. NC is applied in cosmetics, paints and printing
inks, if the nitrogen content is below 12.2%, and as energetic material in guns, if above
12.2%. The source of cellulose and its history also influences the NC application. CN is very
sensible to light and heat, and it necessary to control its degradation to stabilize its proper-
ties [126, 127].
CN can be used to produce derivatives, such as azidodeoxy cellulose nitrate, an energetic
binder [124]. The CN applications are mainly in the biomedical field, like biosensors and
tissue engineering [123], partly due to its high mechanical strength, high solubility and
good compatibility with many plasticizers [122]. The CN applications are supported by the
strong interaction between CN and polycaprolactone, as in the production of composite
nanofibers by electrospinning [123], superhydrophobic and self-cleaning CN nanocom-
posites by blending with natural rubber and nanoclay [128], and in multilayer composite
membrane for gas separation [129]. However, NC is a significant pollutant in industrial
effluents of industries related to energetic materials [130].
f) Cellulose ethers: Sodium carboxymethylcellulose (NaCMC), a water-soluble poly-
saccharide with a linear and long chain and negatively charged (anionic) [8], is the most
valuable cellulose ether. It is a granular substance obtained from the chemical reaction of
cellulose, sodium hydroxide (NaOH) and chloroacetic acid (ClCH2CO2H), which encom-
passes the substitution of hydroxyl groups (OH) of cellulose by methoxyl groups (OCH3).
Undesired reaction byproducts, like sodium chloride, should be removed to attend specific
purity criteria depending on the application. The NaCMC properties are mainly defined
by molecular weight, distribution of carboxyl groups along the cellulose chain and the
degree of substitution (number of carboxyl groups in the anhydroglucose units) [8], which
is within 0.4–1.5 [131].
Sodium carboxymethyl cellulose is among the most used cellulose derivatives, especially
for oil drilling [7], as matrix for drug delivery systems [132], biocomposite scaffolds [133],
superabsorbent hydrogels [134], stabilizer for oil-in-water emulsions [135], additive in pas-
tas (replacing semolina) and for the production of functional foods [136]. The wide use of
NaCMC is based on positive aspects like low cost, rheological properties, low environmen-
tal impact [7], solubility in both hot and cold water and polyelectrolyte character [8]. A
number of cellulose ethers have been studied and used, especially methyl cellulose (MC),
ethylhydroxyethyl cellulose (EHEC), hydroxyethyl cellulose (HEC), CMC, hydroxypropyl
cellulose (HPC) and cationic HECs (catHEC).
g) Regenerated cellulose: Regenerated cellulose is a cellulose derivative produced by
the conversion of natural cellulose—mainly from plant fibers or wood—into a soluble cel-
lulosic material, followed by its regeneration through the formation of fibers by spinning
236 Polysaccharides

process [137]. Common dissolution methods include acid hydrolysis [85], dissolution in
ionic liquids (1-ethyl-3-methylimidazolium chloride) [138, 139], in lithium hydroxide–urea
solution [140], in sodium hydroxide–carbon disulfide solution followed by acid hydrolysis
[141], in N-methylmorpholine-N-oxide, in cuprammonium, and the cellulose acetylation
with acetic anhydride and using sulfuric acid as catalyst [137]. Wet and/or dry spinning is
later used to obtain regenerated cellulose fibers, like in the Lyocell process, which is consid-
ered eco-friendly due to the solvent recovery.
Regenerated cellulose has a glossy and smooth appearance, and water absorption char-
acteristic similar to cotton [137]. Thus, it is used as raw material for several products,
including soft and highly drapable fabrics [137]. Rayon, an artificial textile material or
semi-synthetic fiber, represents a class of commercial reconstituted, regenerated and
purified cellulose fibers, such as viscose, modal and lyocell (from Lenzing AG Company)
[142], commonly present in clothing furnishing and hygiene products. Despite the nat-
ural cellulose source and its biodegradability compared to other derivatives like acetate
[143], rayon fibers significantly contribute to microplastics found in seas and oceans
[144]. High added-value applications can also be obtained from regenerated cellulose.
Oxidized regenerated cellulose serves as bioabsorbable topical hemostatic material to
control bleeding; and regenerated cellulose can be applied to prevent contamination
against antibiotic-resistant microorganisms, as raw material to produce amorphous
nanocellulose (ANC) [85], as a matrix for composites [138, 145, 146], as raw material
for membranes used in separation and adsorption processes [147], and as scaffolds for
biological applications [141].

11.7 Main Applications


Cellulose is probably the most processed polysaccharide in the world since it shows numer-
ous attractive physicochemical, biological, mechanical and economic characteristics, which
are paramount for different kinds of applications. Moreover, due to its versatile chemical
structure, cellulose has been physically and chemically modified, resulting in numerous
derivatives that broaden its range of applications. Indeed, cellulose itself and its derivatives
(including nanocellulose) are currently used in many technological, biomedical and indus-
trial fields, as depicted in Figure 11.1.
For centuries, fibrous materials like cellulose have been applied to produce paper and
woven fabrics. This successful use of cellulose can be mainly attributed to the intrinsic prop-
erties of cellulose fibers, such as malleability and mechanical resistance. Another advantage
is the high recyclability rate of materials produced from cellulose fibers. Although these
properties are shared by cellulose fibers extracted from different sources, their magnitude
can vary considerably because the botanical origin of the plant selected to extract the cel-
lulose fibers influences their features (e.g. cellulose content, crystallinity index, dimension
and degree of polymerization) [23]. For example, cellulose comprises 90% of cotton fibers,
but only 40–50% of wood pulp. Cotton and wood pulp are significant sources of cellulose
for the manufacturing of paper, textiles, construction materials, cardboard, etc. [2].
The technology of paper production from cellulose has advanced significantly since its
discovery in China around 105 A.D. It currently attracts increasing interest in other fields,
such as early-stage diagnostics, 3D scaffolds for cell growth, microfluidics and, especially,
Cellulose and its Derivatives 237

Building and
construction Food
Paper and industry
paper board

Biomedicine
Texile
industry Application
of cellulose,
Hygiene and
cellulose-derivatives
cosmetics
and
Paints nanocellulose

Pharmaceutics

Surface
coatings
Absorbing
materials
Sensors Energy
storage

Figure 11.1 Main applications fields of cellulose, cellulose-derivatives and nanocellulose (nanowhiskers and
nanofibrils).

flexible electronics [148, 149]. This prominent role can be credited to the use of cellulose-­
derivatives and nanocellulose (cellulose nanofibrils (CNF) and nanowhiskers (CNW))
for the production of composites [150, 151]. Overall, the use of these materials in the
paper and board industry aims at enhancing or bring new properties to such products.
For instance, the use of cationic cellulose-derivatives containing quaternary ammonium
groups has found application in the paper industry due to bactericidal properties [152].
In parallel, CNF and CNW have been used as filler materials to strengthen the relative
bonding area in paper sheets [153]. Also, their addition increases sheet density, enhances
paper strength and hydroexpansion, and reduces light scattering and air permeability (i.e.
barrier effect).
Nanocellulose has been used as surface films or coatings on paper and board to improve
printability [4], since it enhances surface smoothness and ink absorption [154]. More
recently, modification of nanocellulose is cited as a suitable strategy to obtain more func-
tional fillers. Feese et al. [155] synthesized fillers based on CNW modified with porphy-
rins which exhibited remarkable photobactericidal effects against Escherichia coli and
Staphylococcus aureus. These photos bactericidal CNW could also be applied in the pro-
duction of woven fabrics, food packaging, and of healthcare products, for instance. In con-
trast, bactericidal papers and boards have been produced by impregnating cellulose fibers
with silver nanoparticles, although some issues related to the leaching of silver have been
mentioned [156].
Composites based on cellulose and other classes of materials (e.g. synthetic polymers,
carbon derivatives and clay) with conductive properties have also been studied [157].
238 Polysaccharides

In this case, cellulose, a strong and flexible non-conductive polymer, is combined with
materials that bring electrical conductivity, optical and chemical properties. The strategies
to prepare conductive papers encompass the grafting of conductive polymers such as poly-
pyrrole or polyaniline in cellulose fibers or their use to coat the fibers [158, 159]. The poor
solubility of cellulose hampers these processes, but this may be overcome, for instance, by
using ionic liquids [160, 161].
Likewise, nanocellulose has been used to prepare conductive papers. It has also found
applications in other electrical devices, such as supercapacitors [162], conductive films and
coatings [163], sensors [164], and also in substrates [165] and separators for energy storage
devices [166]. For example, carboxymethyl-CNF can be used as a substrate for the develop-
ment of transparent papers for dielectric and insulator applications due to its large forward
light scattering ability and high optical transparency [162, 163].
The automotive industry has also taken advantage of the cellulose properties, especially
its reinforcing ability [167]. Compared to synthetic fibers, cellulose fibers and their deriva-
tives have important advantages like low density, low cost, recyclability and biodegradabil-
ity [168]. So, they are promising substituents of glass fibers, generally used in automotive
components (roofs, dashboard, coverings, seats, trunk lids, etc.). More recently, the use
of nanocellulose (CNF and CNW) for this kind of application has been promoted mainly
due to its superior reinforcing ability compared to microfibrillated cellulose, justified by a
stronger polymer matrix/filler bonding and a more effective load transfer [169]. Naturally,
the reinforcing effect can be impaired based on quality, physicochemical and surface
properties of nanocellulose. So, appropriate interfacial adhesion between cellulose (and
its derivatives) and the thermoplastic matrix used to produce these automotive parts is
necessary [167].
Overall, thermoplastics (e.g. polyethylene, polypropylene, and polystyrene) exhibit
hydrophobic nature, which is incompatible with hydrophilic cellulose fibers. This incom-
patibility impairs the dispersion of fibers in the thermoplastic matrix resulting in com-
posites with inferior quality and weak mechanical properties. Coupling agents and fiber
surface modification have been utilized to improve adhesion between cellulose fibers and
thermoplastic matrix [170, 171]. Saeed et al. [171], for instance, recently demonstrated
that composites based on polystyrene reinforced with cellulose-triacetate fibers exhibited
greater interfacial adhesion compared to those with raw cellulose fibers, improving the
mechanical properties of such composites. Despite these promising strategies, the wide-
spread use of composites based on cellulose fibers is hampered by drawbacks associated
with their cost, limited life-cycle and uncontrolled microstructure [172]. According to
Ilyas et al. [173], the replacement of synthetic fibers by cellulose in automotive applica-
tions still needs to address crucial characteristics, such as high-quality and performance,
durability and reliability.
The use of cellulose and cellulose-derivatives in building products (e.g. concrete, wall
plasters, dry-mix mortar, tile adhesives and paints) is also widely investigated since these
materials can have different roles, as shown in Table 11.2 [174, 175]. Moreover, with the gen-
erally growing environmental awareness directed to building materials, the use of cellulose
and its derivatives is expected to expand. A crucial aspect of building products prepared
with biopolymers and derivatives is their impact on the final price of these products, which
is especially worrying for cellulose derivatives, such as methyl hydroxypropylcellulose
(MHPC), hydroxyethylcellulose (HEC) or nanocellulose (CNF and CNW) [175], which are
Cellulose and its Derivatives 239

Table 11.2 Cellulose and cellulose derivatives used in building products.


Building product Cellulosic component Function
Concrete Nanocellulose Reinforcing agent
Cellulose fiber Plasticizer
Adhesion agent
Tile adhesives Nanocellulose Water-retention agent
Cellulose fiber Reinforcing agent
Methylcellulose
Joint filler and components Methyl hydroxyethylcellulose Water-retention agent
Paints and coatings Hydroxyethylcellulose Water-retention agent
Methyl hydroxyethylcellulose Viscosifier
Ethyl hydroxyethylcellulose
Nanocellulose
Oil well construction Hydroxyethylcellulose Water-retention agent

more expensive than cellulose. This suggests that their contribution to the building prod-
ucts must be highly valuable [174]. One strategy to obtain cost reduction or products with
specific properties is to blend cellulosic materials with a small content of a synthetic poly-
mer (e.g. poly(acrylamide), poly(vinyl alcohol) or poly(ethylene oxide)), tailoring specific
binders for major users.
While etherified cellulose-derivatives have a remarkable role in building products, ester-
ified cellulose derivatives are used in various separation techniques in the field of food and
beverages, pharmacy, wastewater treatment, among others [2, 161, 176]. Cellulose esters
perform all types of filtration, such as ultrafiltration, filtration of particles, nanoparticles,
microparticles, and hyperfiltration. They show high viscosity at low concentrations and can
also work as anti-foaming surfactant agents.
In the food industry, cellulose and its derivatives are mostly used as anticaking, dispers-
ing, emulsifying, gellyfing, and texturizing agents [2]. The most utilized cellulose deriva-
tives for foodstuff are acetate and nitrate esters of cellulose, cellulose ethers, and sodium
carboxymethyl cellulose, which are water-soluble derivatives [2]. According to Hamad et al.
[2], their suitability for foodstuff must be guided by their physical and chemical charac-
teristics. The following parameters should also be considered: (i) chemical nature of the
cellulose derivative, (ii) molecular weight, (iii) presence of other active additives, (iv) food
production process, and (v) physical properties, including size and surface charge of the
cellulose derivative.
More recently, cellulose and its derivatives (mainly nanocellulose) have been investigated
for eco-friendly food packaging generally comprised of biocomposites materials, bio-nano
composites, or nanopapers engineered with cellulose fibers, nanocellulose, and other bio-
degradable and renewable materials [3, 177, 178]. Due to their toxicity-free feature, these
packaging materials are claimed to be more efficient in preserving and protecting the food
from contaminants and microbes, extending its shelf life [3]. Additional advantages asso-
ciated with the use of cellulose and nanocellulose include better mechanical properties,
240 Polysaccharides

lightness and barrier effect against water and gases [4]. Moustafa et al. [3] stated that uti-
lization of cellulosic nanofibers in packaging could decrease the cost of products and this
optimist viewpoint is shared by other researchers [179, 180].
The eco-friendly and sustainable label associated with the use of cellulose and its deriva-
tives in products has also stimulated their use in personal care formulations and cosmetics
[181]. Modification of the cellulose backbone has enabled the obtaining of water-soluble
derivatives with new and unique benefits for a variety of products. For example, nonionic
derivatives of cellulose, such as hydroxyethylcellulose (HEC), are used as thickeners and
rheology modifiers in personal care products including shampoos, conditioners, styl-
ing products, personal cleansers and skincare products. Some cellulose derivatives (e.g.
hydroxypropylmethylcellulose—HMPC) are extensively used as excipients in topical cos-
metics due to non-toxic and nonirritating properties [182]. In pharmaceutical applications,
cellulose-derivatives have large utilization, as in extended and controlled release matrices,
osmotic drug delivery systems, bioadhesives and mucoadhesives, thickening agents and
stabilizers, among many others [176, 182]. Similarly, nanocellulose has been used as a deliv-
ery and barrier agent in cosmetic products used for gentle care and effective skin treatment
[183, 184].
The use of cellulose derivatives and nanocellulose in this kind of application is highly
attractive since some of them can yield a gel-like material, which allows the delivery of bio­
active compounds (e.g. cosmetics and drugs) in a localized and controlled manner [185].
Gels are colloidal systems where a continuous solid phase (i.e. a polymeric matrix) is filled
with a disperse phase (gas or liquid) [186]. Hydrogels, for example, are produced with an
aqueous disperse phase. These semisolid materials have hydrophilic properties and are
highly valued in medical and biomedical applications since they are usually transparent,
water washable, thixotropic, greaseless, allow the incorporation of hydrophobic compounds
or insoluble solids, and also show good bioadhesivity [187, 188]. Cellulose-based hydro-
gels also exhibit biocompatible and biodegradable properties [187, 189, 190]. Furthermore,
stimuli-responsive properties can be designed using specific cellulose-derivatives [181, 191,
192].
It is worth noticing that the applicability of cellulose-based hydrogels is directly related
to the preparation method (chemical or physical), and their composition, mechanical and
stimuli-responsive properties. For example, chemical hydrogels exhibit greater mechanical
strength than physical hydrogels, but the chemical reactions lead to permanent networks,
which may result in drawbacks related to biodegradability and toxicity. Recently, dual
crosslinked hydrogels have been exploited to obtain materials with an enhanced biological
performance [189, 193].
The use of raw cellulose fibers to prepare hydrogels is mainly limited by issues related to its
processability in aqueous medium. Although hydrogels can be produced using ionic liquids
[181, 194], water-soluble cellulose derivatives (e.g. carboxymethyl cellulose, hydroxyethyl-
cellulose and methylcellulose) are still preferred for biomedical and pharmaceutical appli-
cations. At least a dozen review papers summarize the utilization of nanocellulose-based
hydrogels for those applications [165, 183, 187, 195–197]. According to Du et al. [196],
the excitement in the use of nanocellulose for preparing hydrogels can be explained by
the unique properties of such nanostructured material, which yields more mechanically
stable self-assembled hydrogels with controlled morphology. Additionally, the incorpora-
tion of nanocellulose in synthetic or natural polymer matrices (as a filler, for instance) has
Cellulose and its Derivatives 241

increased the applicability of hydrogels for biomedical uses, especially due to their ability
to enhance mechanical properties [198], even if only a small amount of nanocellulose is
present (0.01–1 wt%) [196].
The modifiable surface of nanocellulose is another remarkable advantage since it allows
the anchorage of antibacterial and antiviral agents, biomarkers, gene vectors, drugs, among
others [199, 200]. This characteristic widens the range of applicability of nanocellulose-based
hydrogels in different pharmaceutical and biomedical areas (e.g. tissue engineering, bio-
sensors, delivery vehicles and biocatalysts) [183, 195, 173]. Moreover, the modification of
nanocellulose (with hydrophilic or charged groups, for example) can improve their dis-
tribution and interaction with the hydrogel matrix. This allows, for instance, the efficient
preparation by 3D printing of nanocellulose-based hydrogels with tailored 3D-shaped
structures [196, 201, 202].
The incorporation of other materials into nanocellulose-based hydrogels (resulting in
composite hydrogels) has been widely explored in tissue engineering applications. For
instance, the incorporation of minerals such as hydroxyapatite is considered an efficient
strategy to enhance biological activity, and scaffolds prepared with them show benefits in
bone tissue regeneration due to a benign synergistic effect [203]. In addition, injectable
nanocellulose-based hydrogels prepared in situ by chemical or physical crosslinking or by
self-assembling processes [198, 204] has gained widespread appreciation since it encom-
passes a minimally invasive injection into target sites (e.g. skin wounds or bone defects)
[205].
Although many papers have described the potential utilization of cellulose and
­cellulose-derivatives hydrogels in the biomedical field, most of them are still in progress
at a fundamental level and extensive clinical trial studies are still required. Without long-
term studies on toxicology and biocompatibility between these hydrogels and the biological
body, real clinical application of such materials is far from happening.
Cellulose and cellulose-derivative hydrogels have found huge applicability in other rel-
evant areas that are not as strict as biomedical. These hydrogels have been studied for use
in numerous wastewater treatment systems, such as absorption, adsorption, membrane fil-
tration, flocculation, among others [181, 206, 207]. Particularly, they have found extensive
applicability in adsorptive processes towards the removal of various toxic contaminants
(e.g. dyes, medicaments, metal ions, agrochemicals, cosmetics and hydrocarbons) from
water [208–212]. The use of composite hydrogels for this aim, with the incorporation of
materials such as clay, metal and oxide nanoparticles, graphene or carbon nanotubes, is also
commonly reported [211–215].

11.8 Conclusion
Cellulosic materials are abundant, biodegradable, bio-based and biocompatible. They pres-
ent a remarkable set of characteristics, including large surface area, high thermal stability
and water-retention ability, and they can also be readily modified.
Each cellulose source has particularities related to its historical utilization and cost, the
natural state of the cellulose, the required route for its transformation into cellulosic prod-
ucts, and its environmental impact. Indeed, there are important differences regarding the
composition of cellulose and its position in the lignocellulose network and this knowledge
242 Polysaccharides

is important for the development of novel or improved deconstruction methods, which also
depend on the intended final use of the cellulose-derivative. This field, in particular, has
seen significant advances over the last years.
The good balance between environmental and economic features and the excellent phys-
icochemical properties of cellulose and cellulose-derivatives have undoubtedly stimulated
their use. Their demand is also linked to the search for new uses for some lignocellulosic
biomasses, e.g. the exploitation of natural resources from plants, including waste leftovers
from various industrial activities.
As presented in this chapter, the market potential for cellulose and cellulose-derivatives
(particularly nanocellulose) in the electronics, energy, biomedicine, water treatment, and
other sectors is enormous, and there are other emerging technological areas (e.g. agricul-
ture, oil and gas and even aerospace). Their use in useful products can even help to reduce
the demand for petroleum-based polymers. Certainly, cellulose and cellulose-derivatives is
a key-material in the drive for ever-more sustainable and green products.

References
1. Oehme, D.P., Yang, H., Kubicki, J.D., An evaluation of the structures of cellulose generated by
the CHARMM force field: comparisons to in planta cellulose. Cellulose, 25, 3755, 2018.
2. Hamad, A.M.A., Ates, S., Durmaz, E., Evaluation of the possilities for cellulose derivatives in
food products. Kastamonu Univ. J. For. Fac., 16, 383, 2016.
3. Moustafa, H., Youssef, A.M., Darwish, N.A., Abou-Kandil, A.I., Eco-friendly polymer compos-
ites for green packaging: Future vision and challenges. Compos. B-Eng., 172, 16, 2019.
4. Hubbe, M.A., Ferrer, A., Tyagi, P., Yin, Y.Y., Salas, C., Pal, L., Rojas, O.J., Nanocellulose in thin
films, coatings, and plies for packaging applications: A review. Bioresources, 12, 2143, 2017.
5. Chen, G., Liu, B., Zhang, B., Characterization of composite hydrocolloid film based on sodium
cellulose sulfate and cassava starch. J. Food Eng., 125, 105, 2014.
6. Chen, G., Zhang, B., Zhao, J., Chen, H., Development and characterization of food packaging
film from cellulose sulfate. Food Hydrocoll., 35, 476, 2014.
7. Dolz, M., Jiménez, J., Hernández, M.J., Delegido, J., Casanovas, A., Flow and thixotropy of
non-contaminating oil drilling fluids formulated with bentonite and sodium carboxymethyl
cellulose. J. Pet. Sci. Eng., 57, 294, 2007.
8. Singh, R.K. and Khatri, O.P., A scanning electron microscope based new method for determin-
ing degree of substitution of sodium carboxymethyl cellulose. J. Microsc., 246, 43, 2012.
9. Delucis, R.A., Magalhães, W.L.E., Petzhold, C.L., Amico, S.C., Forest-based resources as fillers
in biobased polyurethane foams. J. Appl. Polym. Sci., 135, 45684, 2018.
10. Egüés, I., Sanchez, C., Mondragon, I., Labidi, J., Effect of alkaline and autohydrolysis processes
on the purity of obtained hemicelluloses from corn stalks. Bioresour. Technol., 103, 239, 2012.
11. Garcia, A., Alriols, M.G., Labidi, J., Evaluation of different lignocellulosic raw materials as
potential alternative feedstocks in biorefinery processes. Ind. Crops Prod., 53, 102, 2014.
12. Jawaid, M. and Abdul Khalil, H.P.S., Cellulosic/synthetic fiber reinforced polymer hybrid com-
posites: A review. Carbohydr. Polym., 86, 1, 2011.
13. Alriols, M.G., Tejado, A., Blanco, M., Mondragon, I., Labidi, J., Agricultural palm oil tree res-
idues as raw material for cellulose, lignin and hemicelluloses production by ethylene glycol
pulping process. Chem. Eng. J., 148, 106, 2009.
14. Galbe, M. and Zacchi, G., Pretreatment: The key to efficient utilization of lignocellulosic mate-
rials. Biomass Bioenergy, 46, 70, 2012.
Cellulose and its Derivatives 243

15. Sucaldito, M.R. and Camacho, D.H., Characteristics of unique HBr-hydrolyzed cellulose nano-
crystals from freshwater green algae (Cladophora rupestris) and its reinforcement in starch-
based film. Carbohydr. Polym., 169, 315, 2017.
16. Moon, R.J., Martini, A., Nairn, J., Simonsen, J., Youngblood, J., Cellulose nanomaterials review:
Structure, properties and nanocomposites. Chem. Soc. Rev., 40, 3941, 2011.
17. Mokhena, T.C. and John, M.J., Cellulose nanomaterials: New generation materials for solving
global issues. Cellulose, 27, 1149, 2020.
18. Mihranyan, A., Llagostera, A.P., Karmhag, R., Strømme, M., Ek, R., Moisture sorption by cellu-
lose powders of varying crystallinity. Int. J. Appl. Pharm., 269, 433, 2004.
19. Hai, L.V., Son, H.N., Seo, Y.B., Physical and bio-composite properties of nanocrystalline cellu-
lose from wood, cotton linters, cattail, and red algae. Cellulose, 22, 1789, 2015.
20. Kim, S.M., Park, J.M., Kang, T.Y., Kim, Y.S., Lee, S.K., Purification of squirt cellulose membrane
from the cystic tunic of Styela clava and identification of its osteoconductive effect. Cellulose,
20, 655, 2013.
21. Cosgrove, D.J., Re-constructing our models of cellulose and primary cell wall assembly. Curr.
Opin. Plant Biol., 22, 122, 2014.
22. Faruk, O., Bledzki, A.K., Fink, H-., Sain, M., Biocomposites reinforced with natural fibers:
2000–2010. Prog. Polym. Sci., 37, 1552, 2012.
23. Sorieul, M., Dickson, A., Hill, S., Pearson, H., Plant fiber: Molecular structure and biomechan-
ical properties, of a complex living material, influencing its deconstruction towards a biobased
composite. Materials, 9, 618, 2016.
24. Tsoumis, G., Science and technology of wood: Structure, properties and utilization, Van Nastrnd
Reinold, New York, 1991.
25. Ververis, C., Georghiou, K., Danielidis, D., Hatzinikolaou, D.G., Santas, P., Santas, R., Corleti,
V., Cellulose, hemicelluloses, lignin and ash content of some organic materials and their suit-
ability for use as paper pulp supplements. Bioresour. Technol., 98, 296, 2007.
26. Makarem, M., Lee, C.M., Kafle, K., Huang, S., Chae, I., Yang, H., Kubicki, J.D., Kim, S.H.,
Probing cellulose structures with vibrational spectroscopy. Cellulose, 26, 35, 2019.
27. McNamara, J., Morgan, J.L.W., Zimmer, J., A molecular description of cellulose biosynthesis.
Annu. Rev. Biochem., 84, 895, 2015.
28. Kamel, S., Nanotechnology and its applications in lignocellulosic composites, a mini review.
Express Polym. Lett., 1, 546, 2007.
29. Nishiyama, Y., Langan, P., Chanzy, H., Crystal structure and hydrogen-bonding system in cel-
lulose Iβ from synchrotron X-ray and neutron fiber diffraction. J. Am. Chem. Soc., 124, 9074,
2002.
30. Lee, C., Mittal, A., Barnette, A.L., Kafle, K., Park, Y.B., Shin, H., Johnson, D.K., Park, S., Kim,
S.H., Cellulose polymorphism study with sum-frequency-generation (SFG) vibration spectros-
copy: Identification of exocyclic CH2OH conformation and chain orientation. Cellulose, 20,
991, 2013.
31. Zugenmaier, P., Conformation and packing of various crystalline cellulose fibers. Prog. Polym.
Sci., 26, 1341, 2001.
32. Matthews, J.F., Himmel, M.E., Crowley, M.F., Conversion of cellulose Iα to Iβ via a high tem-
perature intermediate (I-HT) and other cellulose phase transformations. Cellulose, 19, 297,
2012.
33. Jones, D., Ormondroyd, G.O., Curling, S.F., Popescu, C.-M., Popescu, M.-C., Chemical compo-
sitions of natural fibres, in: Advanced High Strength Natural Fibre Composites in Construction,
M. Fan, and F. Fu (Eds.), pp. 23–58, Woodhead Publishing, United Kingdom, 2017.
34. Kobayashi, K., Kimura, S., Togawa, E., Wada, M., Crystal transition from Na–cellulose IV to
cellulose II monitored using synchrotron X-ray diffraction. Carbohydr. Polym., 83, 483, 2011.
244 Polysaccharides

35. Wada, M., Heux, L., Sugiyama, J., Polymorphism of cellulose I family: Reinvestigation of cellu-
lose IVI. Biomacromolecules, 5, 1385, 2004.
36. Qin, L., Li, W., Zhu, J., Liang, J., Li, B., Yuan, Y., Ethylenediamine pretreatment changes cellu-
lose allomorph and lignin structure of lignocellulose at ambient pressure. Biotechnol. Biofuels,
8, 174, 2015.
37. Hall, M., Bansal, P., Lee, J.H., Realff, M.J., Bommarius, A.S., Biological pretreatment of cellu-
lose: Enhancing enzymatic hydrolysis rate using cellulose-binding domains from cellulases.
Bioresour. Technol., 102, 2910, 2011.
38. Chen, Y., Wan, J., Zhang, X., Ma, Y., Wang, Y., Effect of beating on recycled properties of
unbleached eucalyptus cellulose fiber. Carbohydr. Polym., 87, 730, 2012.
39. Ferrer, A., Filpponen, I., Rodríguez, A., Laine, J., Rojas, O.J., Valorization of residual empty
palm fruit bunch fibers (EPFBF) by microfluidization: Production of nanofibrillated cellulose
and EPFBF nanopaper. Bioresour. Technol., 125, 249, 2012.
40. Abdul Khalil, H.P.S., Davoudpour, Y., Islam, M.N., Mustapha, A., Sudesh, K., Dungani, R.,
Jawaid, M., Production and modification of nanofibrillated cellulose using various mechanical
processes: A review. Carbohydr. Polym., 99, 649, 2014.
41. Nakagaito, A.N. and Yano, H., The effect of morphological changes from pulp fiber towards
nano-scale fibrillated cellulose on the mechanical properties of high-strength plant fiber based
composites. Appl. Phys. A, 78, 547, 2004.
42. Siró, I. and Plackett, D., Microfibrillated cellulose and new nanocomposite materials: A review.
Cellulose, 17, 459, 2010.
43. Wang, Y., Wei, X., Li, J., Wang, F., Wang, Q., Chen, J., Kong, L., Study on nanocellulose by high
pressure homogenization in homogeneous isolation. Fibers Polym., 16, 572, 2015.
44. Yuan, Z., Wei, W., Wen, Y., Improving the production of nanofibrillated cellulose from bamboo
pulp by the combined cellulase and refining treatment. J. Chem. Technol. Biotechnol., 94, 2178,
2019.
45. Siddiqui, N., Mills, R.H., Gardner, D.J., Bousfield, D., Production and characterization of cellu-
lose nanofibers from wood pulp. J. Adhes. Sci. Technol., 25, 709, 2011.
46. Kalia, S., Boufi, S., Celli, A., Kango, S., Nanofibrillated cellulose: Surface modification and
potential applications. Colloid Polym. Sci., 292, 5, 2014.
47. Iwamoto, S., Nakagaito, A.N., Yano, H., Nogi, M., Optically transparent composites reinforced
with plant fiber-based nanofibers. Appl. Phys. A, 81, 1109, 2005.
48. Lavoine, N., Desloges, I., Dufresne, A., Bras, J., Microfibrillated cellulose—Its barrier properties
and applications in cellulosic materials: A review. Carbohydr. Polym., 90, 735, 2012.
49. Spence, K.L., Venditti, R.A., Rojas, O.J., Habibi, Y., Pawlak, J.J., A comparative study of energy
consumption and physical properties of microfibrillated cellulose produced by different pro-
cessing methods. Cellulose, 18, 1097, 2011.
50. Iwamoto, S., Kai, W., Isogai, A., Iwata, T., Elastic modulus of single cellulose microfibrils from
tunicate measured by atomic force microscopy. Biomacromolecules, 10, 2571, 2009.
51. Malucelli, L.C., Matos, M., Jordão, C., Lacerda, L.G., Carvalho Filho, M.A.S., Magalhães,
W.L.E., Grinding severity influences the viscosity of cellulose nanofiber (CNF) suspensions
and mechanical properties of nanopaper. Cellulose, 25, 6581, 2018.
52. Lê, H.Q., Dimic-Misic, K., Johansson, L.-S., Maloney, T., Sixta, H., Effect of lignin on the mor-
phology and rheological properties of nanofibrillated cellulose produced from γ-valerolactone/
water fractionation process. Cellulose, 25, 179, 2018.
53. Alemdar, A. and Sain, M., Isolation and characterization of nanofibers from agricultural
­residues—Wheat straw and soy hulls. Bioresour. Technol., 99, 1664, 2008.
54. Wang, B. and Sain, M., Dispersion of soybean stock-based nanofiber in a plastic matrix. Polym.
Int., 56, 538, 2007.
Cellulose and its Derivatives 245

55. Chakraborty, A., Sain, M., Kortschot, M., Cellulose microfibrils: A novel method of prepara-
tion using high shear refining and cryocrushing. Holzforschung, 59, 102, 2005.
56. Islam, M.T., Montarsolo, A., Zoccola, M., Canetti, M., Cacciamani, A., Bertini, F., Effect of indi-
vidualized cellulose fibrils on properties of poly(methyl methacrylate) composites. J. Macromol.
Sci. Phys., 55, 867, 2016.
57. Thiripura Sundari, M. and Ramesh, A., Isolation and characterization of cellulose nanofibers
from the aquatic weed water hyacinth—Eichhornia crassipes. Carbohydr. Polym., 87, 1701,
2012.
58. Wang, B. and Sain, M., Isolation of nanofibers from soybean source and their reinforcing capa-
bility on synthetic polymers. Compos. Sci. Technol., 67, 2521, 2007.
59. Jonoobi, M., Harun, J., Shakeri, A., Misra, M., Oksmand, K., Chemical composition, crystallin-
ity, and thermal degradation of bleached and unbleached kenaf bast (Hibiscus cannabinus) pulp
and nanofibers. BioResources, 4, 626, 2009.
60. Bhatnagar, A. and Sain, M., Processing of cellulose nanofiber-reinforced composites. J. Reinf.
Plast. Compos., 24, 1259, 2005.
61. Castoldi, R., Correa, V.G., de Morais, G.R., de Souza, C.G.M., Bracht, A., Peralta, R.A., Peralta-
Muniz Moreira, R.F., Peralta, R.M., Liquid nitrogen pretreatment of eucalyptus sawdust and
rice hull for enhanced enzymatic saccharification. Bioresour. Technol., 224, 648, 2017.
62. Janardhnan, S. and Sain, M., Bio-treatment of natural fibers in isolation of cellulose nanofi-
bers: Impact of pre-refining of fibers on bio-treatment efficiency and nanofiber yield. J. Polym.
Environ., 19, 615, 2011.
63. Abdellaoui, H., Raji, M., Essabir, H., Qaiss, R.B., El Kacem, A., Bouhfid, R., Qaiss, el Kacem, A.,
Nanofibrillated cellulose-based nanocomposites, in: Bio-Based Polymers and Nanocomposites:
Preparation, Processing, Properties & Performance, M.L. Sanyang, and M. Jawaid (Eds.),
pp. 67–86, Springer Nature, Switzerland, 2019.
64. Wang, W., Sabo, R.C., Mozuch, M.D., Kersten, P., Zhu, J.Y., Jin, Y., Physical and mechanical
properties of cellulose nanofibril films from bleached eucalyptus pulp by endoglucanase treat-
ment and microfluidization. J. Polym. Environ., 23, 551, 2015.
65. Aulin, C., Ahola, S., Josefsson, P., Nishino, T., Hirose, Y., Österberg, M., Wågberg, L., Nanoscale
cellulose films with different crystallinities and mesostructures—Their surface properties and
interaction with water. Langmuir, 25, 7675, 2009.
66. Wang, W., Mozuch, M.D., Sabo, R.C., Kersten, P., Zhu, J.Y., Jin, Y., Production of cellulose
nanofibrils from bleached eucalyptus fibers by hyperthermostable endoglucanase treatment
and subsequent microfluidization. Cellulose, 22, 351, 2015.
67. Khan, A., Vu, K.D., Chauve, G., Bouchard, J., Riedl, B., Lacroix, M., Optimization of microflu-
idization for the homogeneous distribution of cellulose nanocrystals (CNCs) in biopolymeric
matrix. Cellulose, 21, 3457, 2014.
68. Taheri, H. and Samyn, P., Effect of homogenization (microfluidization) process parameters in
mechanical production of micro- and nanofibrillated cellulose on its rheological and morpho-
logical properties. Cellulose, 23, 1221, 2016.
69. Zimmermann, T., Pöhler, E., Geiger, T., Cellulose fibrils for polymer reinforcement. Adv. Eng.
Mater., 6, 754, 2004.
70. Ma, P.-C., Siddiqui, N.A., Marom, G., Kim, J.-K., Dispersion and functionalization of carbon
nanotubes for polymer-based nanocomposites: A review. Compos. Part A Appl. Sci. Manuf., 41,
1345, 2010.
71. Piras, C.C., Fernández-Prieto, S., De Borggraeve, W.M., Ball milling: a green technology for the
preparation and functionalisation of nanocellulose derivatives. Nanoscale Adv., 1, 937, 2019.
72. Ago, M., Endo, T., Okajima, K., Effect of solvent on morphological and structural change of
cellulose under ball-milling. Polym. J., 39, 435, 2007.
246 Polysaccharides

73. Qua, E.H., Hornsby, P.R., Sharma, H.S.S., Lyons, G., Preparation and characterisation of cellu-
lose nanofibers. J. Mater. Sci., 46, 6029, 2011.
74. Yang, D., Peng, X.-W., Zhong, L.-X., Cao, X.-F., Chen, W., Sun, R.-C., Effects of pretreatments
on crystalline properties and morphology of cellulose nanocrystals. Cellulose, 20, 2427, 2013.
75. Nge, T.T., Lee, S.-H., Endo, T., Preparation of nanoscale cellulose materials with different mor-
phologies by mechanical treatments and their characterization. Cellulose, 20, 1841, 2013.
76. Chen, W., Yu, H., Liu, Y., Chen, P., Zhang, M., Hai, Y., Individualization of cellulose nanofi-
bers from wood using high-intensity ultrasonication combined with chemical pretreatments.
Carbohydr. Polym., 83, 1804, 2011.
77. Chen, W., Li, Q., Cao, J., Liu, Y., Li, J., Zhang, J., Luo, S., Yu, H., Revealing the structures of
cellulose nanofiber bundles obtained by mechanical nanofibrillation via TEM observation.
Carbohydr. Polym., 117, 950, 2015.
78. Soni, B., Hassan, E.B., Mahmoud, B., Chemical isolation and characterization of different cellu-
lose nanofibers from cotton stalks. Carbohydr. Polym., 134, 581, 2015.
79. Konsta-Gdoutos, M.S. and Aza, C.A., Self sensing carbon nanotube (CNT) and nanofiber
(CNF) cementitious composites for real time damage assessment in smart structures. Cem.
Concr. Compos., 53, 162, 2014.
80. Hubbe, M.A., Rojas, O.J., Lucia, L.A., Sain, M., Cellulosic nanocomposites: A review.
Bioresourses, 3, 929, 2008.
81. Beck-Candanedo, S., Roman, M., Gray, D.G., Effect of reaction conditions on the properties
and behavior of wood cellulose nanocrystal suspensions. Biomacromolecules, 6, 1048, 2005.
82. Araki, J., Wada, M., Kuga, S., Okano, T., Influence of surface charge on viscosity behavior of
cellulose microcrystal suspension. J. Wood Sci., 45, 258, 1999.
83. Habibi, Y., Lucia, L.A., Rojas, O.J., Cellulose nanocrystals: Chemistry, self-assembly, and appli-
cations. Chem. Rev., 110, 3479, 2010.
84. Ioelovich, M., Cellulose as a nanostructured polymer: A short review. Bioresourses, 3, 1403,
2008.
85. Ioelovich, M., Nanoparticles of amorphous cellulose and their properties. Am. J. Nano Res.
Appl., 1, 41, 2013.
86. Khoo, R.Z., Ismail, H., Chow, W.S., Thermal and morphological properties of poly (lactic acid)/
nanocellulose nanocomposites. Procedia Chem., 19, 788, 2016.
87. Lee, S.-Y., Mohan, D.J., Kang, I.-A., Doh, G.-H., Lee, S., Han, S.O., Nanocellulose reinforced
PVA composite films: Effects of acid treatment and filler loading. Fibers Polym., 10, 77, 2009.
88. Maiti, S., Jayaramudu, J., Das, K., Reddy, S.M., Sadiku, R., Ray, S.S., Liu, D., Preparation and
characterization of nano-cellulose with new shape from different precursor. Carbohydr. Polym.,
98, 562, 2013.
89. Mandal, A. and Chakrabarty, D., Isolation of nanocellulose from waste sugarcane bagasse
(SCB) and its characterization. Carbohydr. Polym., 86, 1291, 2011.
90. dos Santos, F.A., Iulianelli, G.C.V., Tavares, M.I.B., Effect of microcrystalline and nanocrystals
cellulose fillers in materials based on PLA matrix. Polym. Test., 61, 280, 2017.
91. Dong, X., Revol, J.-F., Gray, D.G., Effect of microcrystallite preparation conditions on the for-
mation of colloid crystals of cellulose. Cellulose, 5, 19, 1998.
92. 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., 43, 3434, 2007.
93. Pääkkö, M., Ankerfors, M., Kosonen, H., Nykänen, A., Ahola, S., Österberg, M., Ruokolainen, J.,
Laine, J., Larsson, P.T., Ikkala, O., Lindström, T., Enzymatic hydrolysis combined with mechan-
ical shearing and high-pressure homogenization for nanoscale cellulose fibrils and strong gels.
Biomacromolecules, 8, 1934, 2007.
Cellulose and its Derivatives 247

94. Percival Zhang, Y.-H., Himmel, M.E., Mielenz, J.R., Outlook for cellulase improvement:
Screening and selection strategies. Biotechnol. Adv., 24, 452, 2006.
95. Chinga-Carrasco, G., Kuznetsova, N., Garaeva, M., Leirset, I., Galiullina, G., Kostochko, A.,
Syverud, K., Bleached and unbleached MFC nanobarriers: Properties and hydrophobisation
with hexamethyldisilazane. J. Nanopart. Res., 14, 1280, 2012.
96. Tejado, A., Alam, M.N., Antal, M., Yang, H., van de Ven, T.G.M., Energy requirements for the
disintegration of cellulose fibers into cellulose nanofibers. Cellulose, 19, 831, 2012.
97. Chinga-Carrasco, G. and Syverud, K., Pretreatment-dependent surface chemistry of wood
nanocellulose for pH-sensitive hydrogels. J. Biomater. Appl., 29, 423, 2014.
98. Azizi Samir, M.A.S., Alloin, F., Dufresne, A., Review of recent research into cellulosic whiskers,
their properties and their application in nanocomposite field. Biomacromolecules, 6, 612, 2005.
99. Abitbol, T., Rivkin, A., Cao, Y., Nevo, Y., Abraham, E., Ben-Shalom, T., Lapidot, S., Nanocellulose,
a tiny fiber with huge applications. Curr. Opin. Biotechnol., 39, 76, 2016.
100. Lavoine, N. and Bergström, L., Nanocellulose-based foams and aerogels: Processing, proper-
ties, and applications. J. Mater. Chem. A, 5, 16105, 2017.
101. 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., 154, 151, 2016.
102. Kim, S.-H., Kim, E.-S., Choi, K., Cho, J.K., Sun, H., Yoo, J.W., Park, I.-K., Lee, Y., Choi, H.R.,
Kim, T., Suhr, J., Yun, J.-H., Choi, H.J., Nam, J.-D., Rheological and mechanical properties of
polypropylene composites containing microfibrillated cellulose (MFC) with improved compat-
ibility through surface silylation. Cellulose, 26, 1085, 2019.
103. Zhang, T., Zhang, Y., Wang, X., Liu, S., Yao, Y., Characterization of the nano-cellulose aerogel
from mixing CNF and CNC with different ratio. Mater. Lett., 229, 103, 2018.
104. Czaja, W., Krystynowicz, A., Bielecki, S., Brown Jr., R.M., Microbial cellulose—The natural
power to heal wounds. Biomaterials, 27, 145, 2006.
105. Park, H.-M., Mohanty, A.K., Drzal, L.T., Lee, E., Mielewski, D.F., Misra, M., Effect of sequen-
tial mixing and compounding conditions on cellulose acetate/layered silicate nanocomposites.
J. Polym. Environ., 14, 27, 2006.
106. Puls, J., Wilson, S.A., Hölter, D., Degradation of cellulose acetate-based materials: A review.
J. Polym. Environ., 19, 152, 2011.
107. Shojaie, S.S., Rials, T.G., Kelley, S.S., Preparation and characterization of cellulose acetate
organic/inorganic hybrid films. J. Appl. Polym. Sci., 58, 1263, 1995.
108. Zoppi, R.A. and Gonçalves, M.C., Hybrids of cellulose acetate and sol–gel silica: Morphology,
thermomechanical properties, water permeability, and biodegradation evaluation. J. Appl.
Polym. Sci., 84, 2196, 2002.
109. da Silva, C.A., Favaro, M.M., Yoshida, I.V.P., Gonçalves, M.C., Nanocomposites derived from
cellulose acetate and highly branched alkoxysilane. J. Appl. Polym. Sci., 121, 2559, 2011.
110. Maheswari, P., Barghava, P., Mohan, D., Preparation, morphology, hydrophilicity and perfor-
mance of poly (ether-ether-sulfone) incorporated cellulose acetate ultrafiltration membranes.
J. Polym. Res., 20, 74, 2013.
111. Gorey, C. and Escobar, I.C., N-isopropylacrylamide (NIPAAM) modified cellulose acetate
ultrafiltration membranes. J. Membr. Sci., 383, 272, 2011.
112. El Badawi, N., Ramadan, A.R., Esawi, A.M.K., El-Morsi, M., Novel carbon nanotube–­cellulose
acetate nanocomposite membranes for water filtration applications. Desalination, 344, 79, 2014.
113. Soyekwo, F., Zhang, Q.G., Deng, C., Gong, Y., Zhu, A.M., Liu, Q.L., Highly permeable cellu-
lose acetate nanofibrous composite membranes by freeze-extraction. J. Membr. Sci., 454, 339,
2014.
114. Wu, Q.-X., Guan, Y.-X., Yao, S.-J., Sodium cellulose sulfate: A promising biomaterial used for
microcarriers’ designing. Front. Chem. Sci. Eng., 13, 46, 2019.
248 Polysaccharides

115. Brewer, R.J. and Tenn, K., Process for preparing cellulose sulfate esters, US Patent 4,480,091,
assigned to Eastman Kodak Company, 1983, https://patents.google.com/patent/US3528963A/en
116. Chen, G., Zhang, B., Zhao, J., Chen, H., Improved process for the production of cellulose sulfate
using sulfuric acid/ethanol solution. Carbohydr. Polym., 95, 332, 2013.
117. Yao, S., An improved process for the preparation of sodium cellulose sulphate. Chem. Eng. J.,
78, 199, 2000.
118. Schuldt, U., Wagenknecht, W., Richter, A., Electrosorption of sodium cellulose sulfates with
different substitution patterns. Cellulose, 9, 271, 2002.
119. Zhu, L.-Y., Lin, D.-Q., Yao, S.-J., Biodegradation of polyelectrolyte complex films composed of
chitosan and sodium cellulose sulfate as the controllable release carrier. Carbohydr. Polym., 82,
323, 2010.
120. Chen, G. and Liu, B., Cellulose sulfate based film with slow-release antimicrobial properties
prepared by incorporation of mustard essential oil and β-cyclodextrin. Food Hydrocoll., 55, 100,
2016.
121. Xie, Y.-L., Wang, M.-J., Yao, S.-J., Preparation and characterization of biocompatible microcap-
sules of sodium cellulose sulfate/chitosan by means of layer-by-layer self-assembly. Langmuir,
25, 8999, 2009.
122. Gensh, K.V., Kolosov, P.V., Bazarnova, N.G., Quantitative analysis of cellulose nitrates by
Fourier transform infrared spectroscopy. Russ. J. Bioorganic Chem., 37, 814, 2011.
123. Nartker, S., Hassan, M., Stogsdill, M., Electrospun cellulose nitrate and polycaprolactone
blended nanofibers. Mater. Res. Express, 2, 35401, 2015.
124. Pourmortazavi, S.M., Sadri, M., Rahimi-Nasrabadi, M., Shamsipur, M., Jabbarzade, Y., Khalaki,
M.S., Abdollahi, M., Shariatinia, Z., Kohsari, I., Atifeh, S.M., Thermal decomposition kinet-
ics of electrospun azidodeoxy cellulose nitrate and polyurethane nanofibers. J. Therm. Anal.
Calorim., 119, 281, 2015.
125. Koob, S.P., The instability of cellulose nitrate adhesives. Conserv., 6, 31, 1982.
126. Moniruzzaman, M., Bellerby, J.M., Mai, N., The effect of light on the viscosity and molecular
mass of nitrocellulose. Polym. Degrad. Stab., 96, 929, 2011.
127. Quye, A., Littlejohn, D., Pethrick, R.A., Stewart, R.A., Investigation of inherent degradation in
cellulose nitrate museum artefacts. Polym. Degrad. Stab., 96, 1369, 2011.
128. Bayer, I.S., Steele, A., Martorana, P., Loth, E., Robinson, S.J., Stevenson, D., Biolubricant induced
phase inversion and superhydrophobicity in rubber-toughened biopolymer/organoclay nano-
composites. Appl. Phys. Lett., 95, 63702, 2009.
129. Shieh, J.-J. and Chung, T.S., Cellulose nitrate-based multilayer composite membranes for gas
separation. J. Membr. Sci., 166, 259, 2000.
130. Berthumeyrie, S., Collin, S., Bussiere, P.-O., Therias, S., Photooxidation of cellulose nitrate:
New insights into degradation mechanisms. J. Hazard. Mater., 272, 137, 2014.
131. Du, B., Li, J., Zhang, H., Huang, L., Chen, P., Zhou, J., Influence of molecular weight and
degree of substitution of carboxymethylcellulose on the stability of acidified milk drinks. Food
Hydrocoll., 23, 1420, 2009.
132. Cai, X., Luan, Y., Dong, Q., Shao, W., Li, Z., Zhao, Z., Sustained release of 5-fluorouracil by
incorporation into sodium carboxymethylcellulose sub-micron fibers. Int. J. Pharm., 419, 240,
2011.
133. Grumezescu, A.M., Andronescu, E., Ficai, A., Bleotu, C., Mihaiescu, D.E., Chifiriuc, M.C.,
Synthesis, characterization and in vitro assessment of the magnetic chitosan–carboxymethyl-
cellulose biocomposite interactions with the prokaryotic and eukaryotic cells. Int. J. Pharm.,
436, 771, 2012.
134. Tang, H., Chen, H., Duan, B., Lu, A., Zhang, L., Swelling behaviors of superabsorbent chitin/
carboxymethylcellulose hydrogels. J. Mater. Sci., 49, 2235, 2014.
Cellulose and its Derivatives 249

135. Liu, L., Zhao, Q., Liu, T., Kong, J., Long, Z., Zhao, M., Sodium caseinate/carboxymethylcellu-
lose interactions at oil–water interface: Relationship to emulsion stability. Food Chem., 132,
1822, 2012.
136. Aravind, N., Sissons, M., Fellows, C.M., Effect of soluble fiber (guar gum and carboxymethyl-
cellulose) addition on technological, sensory and structural properties of durum wheat spa-
ghetti. Food Chem., 131, 893, 2012.
137. Chen, J., Synthetic textile fibers: Regenerated cellulose fibers, in: Textiles and Fashion, R. Sinclair,
(Ed.) pp. 79–95, Woodhead Publishing, United Kingdom, 2015.
138. Hanid, N.A., Wahit, M.U., Guo, Q., Mahmoodian, S., Soheilmoghaddam, M., Development of
regenerated cellulose/halloysites nanocomposites via ionic liquids. Carbohydr. Polym., 99, 91,
2014.
139. Zhang, H., Wu, J., Zhang, J., He, J., 1-allyl-3-methylimidazolium chloride room temperature
ionic liquid: A new and powerful nonderivatizing solvent for cellulose. Macromolecules, 38,
8272, 2005.
140. Isobe, N., Kim, U.-J., Kimura, S., Wada, M., Kuga, S., Internal surface polarity of regenerated
cellulose gel depends on the species used as coagulant. J. Colloid Interface Sci., 359, 194, 2011.
141. de Araújo Jr., A.M., Braido, G., Saska, S., Barud, H.S., Franchi, L.P., Assunção, M.N., Scarel-
Caminaga, R.M., Capote, T.S.O., Messaddeq, Y., Ribeiro, S.J.L., Regenerated cellulose scaf-
folds: Preparation, characterization and toxicological evaluation. Carbohydr. Polym., 136, 892,
2016.
142. Kauffman, G.B., Rayon: The first semi-synthetic fiber product. J. Chem. Educ., 70, 887, 1993.
143. Park, C.H., Kang, Y.K., Im, S.S., Biodegradability of cellulose fabrics. J. Appl. Polym. Sci., 94,
248, 2004.
144. Lusher, A.L., McHugh, M., Thompson, R.C., Occurrence of microplastics in the gastrointesti-
nal tract of pelagic and demersal fish from the English channel. Mar. Pollut. Bull., 67, 94, 2013.
145. Ren, F., Li, Z., Tan, W.-Z., Liu, X.-H., Sun, Z.-F., Ren, P.-G., Yan, D.-X., Facile preparation of
3D regenerated cellulose/graphene oxide composite aerogel with high-efficiency adsorption
towards methylene blue. J. Colloid Interface Sci., 532, 58, 2018.
146. Zhang, Y., Jiang, Y., Han, L., Wang, B., Xu, H., Zhong, Y., Zhang, L., Mao, Z., Sui, X., Biodegradable
regenerated cellulose-dispersed composites with improved properties via a pickering emulsion
process. Carbohydr. Polym., 179, 86, 2018.
147. Xiong, X., Duan, J., Zou, W., He, X., Zheng, W., A pH-sensitive regenerated cellulose mem-
brane. J. Membr. Sci., 363, 96, 2010.
148. Gao, B.B., Liu, H., Gu, Z.Z., Flourishing smart flexible membranes beyond paper. Analy. Chem.,
91, 4224, 2019.
149. Yetisen, A.K., Akram, M.S., Lowe, C.R., Paper-based microfluidic point-of-care diagnostic
devices. Lab Chip, 13, 2210, 2013.
150. Kargarzadeh, H., Huang, J., Lin, N., Ahmad, I., Mariano, M., Dufresne, A., Galeski, A., Recent
developments in nanocellulose-based biodegradable polymers, thermoplastic polymers, and
porous nanocomposites. Progr. Polym. Sci., 87, 197, 2018.
151. Nguyen, L.H., Naficy, S., Chandrawati, R., Dehghani, F., Nanocellulose for sensing applications.
Adv. Mater. Interfaces, 6, 1, 2019.
152. Coma, V., Polysaccharide-based biomaterials with antimicrobial and antioxidant properties.
Polimeros, 23, 287, 2013.
153. Bharimalla, A.K., Deshmukh, S.P., Patil, P.G., Vigneshwaran, N., Micro/nano-fibrillated cel-
lulose from cotton linters as strength additive in unbleached kraft paper: Experimental, semi-­
empirical, and mechanistic studies. Bioresources, 12, 5682, 2017.
154. Torvinen, K., Sievanen, J., Hjelt, T., Hellen, E., Smooth and flexible filler-nanocellulose com-
posite structure for printed electronics applications. Cellulose, 19, 821, 2012.
250 Polysaccharides

155. Feese, E., Sadeghifar, H., Gracz, H.S., Argyropoulos, D.S., Ghiladi, R.A., Photobactericidal
porphyrin-cellulose nanocrystals: Synthesis, characterization, and antimicrobial properties.
Biomacromolecules, 12, 3528, 2011.
156. Dankovich, T.A. and Gray, D.G., Bactericidal paper impregnated with silver nanoparticles for
point-of-use water treatment. Environ. Sci. Technol., 45, 1992, 2011.
157. He, M.H., Zhang, K.L., Chen, G.X., Tian, J.F., Su, B., Ionic gel paper with long-term bendable
electrical robustness for use in flexible electroluminescent devices. ACS Appl. Mater. Interfaces,
9, 16466, 2017.
158. Baptista, A.C., Ropio, I., Romba, B., Nobre, J.P., Henriques, C., Silva, J.C., Ferreira, I., Cellulose-
based electrospun fibers functionalized with polypyrrole and polyaniline for fully organic bat-
teries. J. Mater. Chem. A, 6, 256, 2019.
159. Sasso, C., Zeno, E., Petit-Conil, M., Chaussy, D., Belgacem, M.N., Tapin-Lingua, S., Beneventi,
D., Highly conducting polypyrrole/cellulose nanocomposite films with enhanced mechanical
properties. Macromol. Mater. Eng., 295, 934, 2010.
160. Isik, M., Sardon, H., Mecerreyes, D., Ionic liquids and cellulose: Dissolution, chemical modifi-
cation and preparation of new cellulosic materials. Int. J. Mol. Sci., 15, 11922, 2014.
161. Kang, H.L., Liu, R.G., Huang, Y., Cellulose derivatives and graft copolymers as blocks for func-
tional materials. Polym. Int., 62, 338, 2013.
162. Tayeb, P. and Tayeb, A.H., Nanocellulose applications in sustainable electrochemical and piezo-
electric systems: A review. Carbohydr. Polym., 224, 1, 2019.
163. Hu, L.B., Zheng, G.Y., Yao, J., Liu, N.A., Weil, B., Eskilsson, M., Cui, Y., Transparent and con-
ductive paper from nanocellulose fibers. Energy Environ. Sci., 6, 513, 2013.
164. Agate, S., Joyce, M., Lucia, L., Pal, L., Cellulose and nanocellulose-based flexible-hybrid printed
electronics and conductive composites—A review. Carbohydr. Polym., 198, 249, 2018.
165. Panchal, P., Ogunsona, E., Mekonnen, T., Trends in advanced functional material applications
of nanocellulose. Processes, 7, 1, 2019.
166. Wang, Z.H., Tammela, P., Stromme, M., Nyholm, L., Cellulose-based supercapacitors: material
and performance considerations. Adv. Energy Mater., 7, 1, 2017.
167. Campos, A.R., Cunha, A.M., Tielas, A., Mateos, A., Biodegradable composites applied to the
automotive industry: The development of a loudspeaker front. Adv. Mater. Forum, 587, 187, 2008.
168. Oksman, K., Mathew, A.P., Sain, M., Novel bionanocomposites: Processing, properties and
potential applications. Plast. Rubber Compos., 38, 396, 2009.
169. Kargarzadeh, H., Mariano, M., Huang, J., Lin, N., Ahmad, I., Dufresne, A., Thomas, S., Recent
developments on nanocellulose reinforced polymer nanocomposites: A review. Polymer, 132,
368, 2017.
170. Dufresne, A. and Belgacem, M.N., Cellulose-reinforced composites: From micro-to nanoscale.
Polímeros, 23, 277, 2013.
171. Saeed, U., Dawood, U., Ali, M.A., Cellulose triacetate fiber-reinforced polystyrene compos-
ite. J. Thermoplast. Compos. Mater., 2019, In https://journals.sagepub.com/doi/abs/10.1177/​
0892705719847249?journalCode=jtca#articleCitationDownloadContainer
172. Mohanty, A.K., Misra, M., Drzal, L.T., Sustainable bio-composites from renewable resources:
Opportunities and challenges in the green materials world. J. Polym. Environ., 10, 19, 2002.
173. Ilyas, R.A., Sapuan, S.M., Sanyang, M.L., Ishak, M.R., Zainudin, E.S., Nanocrystalline cellulose
as reinforcement for polymeric matrix nanocomposites and its potential applications: A review.
Curr. Anal. Chem., 14, 203, 2018.
174. Hospodarova, V., Stevulova, N., Sicakova, A., Possibilities of using cellulose fibers in building
materials, in: 2nd International Conference on Innovative Materials, Structures and Technologies
(IMST), Riga, Latvia, 2015.
Cellulose and its Derivatives 251

175. Plank, J., Applications of biopolymers and other biotechnological products in building materi-
als. Appl. Microbiol. Biotechnol., 66, 1, 2004.
176. Reis, D.T., Pereira, A.K.D., Scheidt, G.N., Pereira, D.H., Plant and bacterial cellulose: Production,
chemical structure, derivatives and applications. Orbital, 11, 321, 2019.
177. Youssef, A.M. and El-Sayed, S.M., Bionanocomposites materials for food packaging applica-
tions: Concepts and future outlook. Carbohydr. Polym., 193, 19, 2018.
178. de Azeredo, H.M.C., Nanocomposites for food packaging applications. Food Res. Int., 42, 1240,
2009.
179. Khalil, H., Davoudpour, Y., Saurabh, C.K., Hossain, M.S., Adnan, A.S., Dungani, R., Haafiz,
M.K.M., A review on nanocellulosic fibers as new material for sustainable packaging: Process
and applications. Renewable Sustainable Energy Rev., 64, 823, 2016.
180. Mondal, S., Preparation, properties and applications of nanocellulosic materials. Carbohydr.
Polym., 163, 301, 2017.
181. Kabir, S.M.F., Sikdar, P.P., Haque, B., Bhuiyan, M.A.R., Ali, A., Islam, M.N., Cellulose-based
hydrogel materials: Chemistry, properties and their prospective applications. Prog. Biomater.,
7, 153, 2018.
182. Kamel, S., Ali, N., Jahangir, K., Shah, S.M., El-Gendy, A.A., Pharmaceutical significance of cel-
lulose: A review. Express Polym. Lett., 2, 758, 2008.
183. Bacakova, L., Pajorova, J., Bacakova, M., Skogberg, A., Kallio, P., Kolarova, K., Svorcik, V.,
Versatile application of nanocellulose: from industry to skin tissue engineering and wound
healing. Nanomaterials, 9, 1, 2019.
184. Klemm, D., Cranston, E.D., Fischer, D., Gama, M., Kedzior, S.A., Kralisch, D., Rauchfuss, F.,
Nanocellulose as a natural source for groundbreaking applications in materials science: Today’s
state. Mater. Today, 21, 720, 2018.
185. Shokri, J. and Adbkia, K., Application of cellulose and cellulose derivatives in pharmaceutical
industries, in: Cellulose—Medical, Pharmaceutical and Electronic Applications, T. van de Ven,
(Ed.), Intech Open, London, UK, 2013.
186. Marr, P.C. and Marr, A.C., Ionic liquid gel materials: applications in green and sustainable
chemistry. Green Chem., 18, 105, 2016.
187. Mohammadinejad, R., Maleki, H., Larrañeta, E., Fajardo, A.R., Nik, A.B., Shavandi, A., Thakur,
V.K., Status and future scope of plant-based green hydrogels in biomedical engineering. Appl.
Mater. Today, 16, 213, 2019.
188. Ullah, F., Othman, M.B.H., Javed, F., Ahmad, Z., Akil, H.M., Classification, processing and
application of hydrogels: A review. Mater. Sci. Eng. C, 57, 414, 2015.
189. Fu, L.H., Qi, C., Ma, M.G., Wan, P.B., Multifunctional cellulose-based hydrogels for biomedical
applications. J. Mater. Chem. B, 7, 1541, 2019.
190. Rodrigues, F.H.A., Spagnol, C., Pereira, A.G.B., Martins, A.F., Fajardo, A.R., Rubira, A.F.,
Muniz, E.C., Superabsorbent hydrogel composites with a focus on hydrogels containing nano-
fibers or nanowhiskers of cellulose and chitin. J. Appl. Polym. Sci., 131, 1, 2014.
191. Chang, C.Y., Duan, B., Cai, J., Zhang, L.N., Superabsorbent hydrogels based on cellulose for
smart swelling and controllable delivery. Eur. Polym. J., 46, 92, 2010.
192. Qiu, X.Y. and Hu, S.W., “Smart” materials based on cellulose: A review of the preparations,
properties, and applications. Materials, 6, 738, 2013.
193. Fajardo, A.R., Favaro, S.L., Rubira, A.F., Muniz, E.C., Dual-network hydrogels based on chem-
ically and physically crosslinked chitosan/chondroitin sulfate. React. Funct. Polym., 73, 1662,
2013.
194. Peng, H.F., Wang, S.P., Xu, H.Y., Dai, G.L., Preparations, properties, and formation mechanism
of novel cellulose hydrogel membrane based on ionic liquid. J. Appl. Polym. Sci., 135, 1, 2018.
252 Polysaccharides

195. de France, K.J., Hoare, T., Cranston, E.D., Review of hydrogels and aerogels containing nano-
cellulose. Chem. Mater., 29, 4609, 2017.
196. Du, H.S., Liu, W.M., Zhang, M.L., Si, C.L., Zhang, X.Y., Li, B., Cellulose nanocrystals and cellu-
lose nanofibrils based hydrogels for biomedical applications. Carbohydr. Polym., 209, 130, 2019.
197. Dugan, J.M., Gough, J.E., Eichhorn, S.J., Bacterial cellulose scaffolds and cellulose nanowhis-
kers for tissue engineering. Nanomedicine, 8, 287, 2013.
198. Curvello, R., Raghuwanshi, V.S., Garnier, G., Engineering nanocellulose hydrogels for biomed-
ical applications. Adv. Colloid Interface Sci., 267, 47, 2019.
199. Aziz, T., Fan, H., Haq, F., Khan, F.U., Numan, A., Ullah, A., Wazir, N., Facile modification and
application of cellulose nanocrystals. Iran. Polym. J., 28, 707, 2019.
200. Hu, W.L., Chen, S.Y., Yang, J.X., Li, Z., Wang, H.P., Functionalized bacterial cellulose derivatives
and nanocomposites. Carbohydr. Polym., 101, 1043, 2014.
201. Wang, Q.Q., Sun, J.Z., Yao, Q., Ji, C.C., Liu, J., Zhu, Q.Q., 3D printing with cellulose materials.
Cellulose, 25, 4275, 2018.
202. Wang, Y.G., Wang, X.J., Xie, Y.J., Zhang, K., Functional nanomaterials through esterification of
cellulose: A review of chemistry and application. Cellulose, 25, 3703, 2018.
203. Lin, N. and Dufresne, A., Nanocellulose in biomedicine: Current status and future prospect.
Eur. Polym. J., 59, 302, 2014.
204. del Valle, L.J., Diaz, A., Puiggali, J., Hydrogels for biomedical applications: cellulose, chitosan,
and protein/peptide derivatives. Gels, 3, 1, 2017.
205. Liu, M., Zeng, X., Ma, C., Yi, H., Ali, Z., Mou, X.B., He, N.Y., Injectable hydrogels for cartilage
and bone tissue engineering. Bone Res., 5, 1, 2017.
206. Shanmugarajah, B., Chew, I.M., Mubarak, N.M., Choong, T.S., Yoo, C., Tan, K., Valorization of
palm oil agro-waste into cellulose biosorbents for highly effective textile effluent remediation.
J. Clean. Prod., 210, 697, 2019.
207. Wang, D., A critical review of cellulose-based nanomaterials for water purification in industrial
processes. Cellulose, 26, 687, 2019.
208. Chen, P.P., Liu, X.Y., Jin, R.D., Nie, W.Y., Zhou, Y.F., Dye adsorption and photo-induced recy-
cling of hydroxypropyl cellulose/molybdenum disulfide composite hydrogels. Carbohydr.
Polym., 167, 36, 2017.
209. Liu, J.J., Chu, H.J., Wei, H.L., Zhu, H.Z., Wang, G., Zhu, J., He, J., Facile fabrication of car-
boxymethyl cellulose sodium/graphene oxide hydrogel microparticles for water purification.
RSC Adv., 6, 50061, 2016.
210. Ma, J.H., Liu, Y.T., Ali, O., Wei, Y.F., Zhang, S.Q., Zhang, Y.M., Luo, S.L., Fast adsorption of
heavy metal ions by waste cotton fabrics based double network hydrogel and influencing fac-
tors insight. J. Hazard. Mater., 344, 1034, 2018.
211. Melo, B.C., Paulino, F.A.A., Cardoso, V.A., Pereira, A.G.B., Fajardo, A.R., Rodrigues, F.H.A.,
Cellulose nanowhiskers improve the methylene blue adsorption capacity of chitosan-g-
poly(acrylic acid) hydrogel. Carbohydr. Polym., 181, 358, 2018.
212. Santoso, S.P., Kurniawan, A., Soetaredjo, F.E., Cheng, K.C., Putro, J.N., Ismadji, S., Ju, Y.H., Eco-
friendly cellulose-bentonite porous composite hydrogels for adsorptive removal of azo dye and
soilless culture. Cellulose, 26, 3339, 2019.
213. Anirudhan, T.S. and Tharun, A.R., Preparation and adsorption properties of a novel interpene-
trating polymer network (IPN) containing carboxyl groups for basic dye from aqueous media.
Chem. Eng. J., 181, 761, 2012.
214. Dai, H.J. and Huang, H.H., Synthesis, characterization and properties of pineapple peel
­cellulose-g-acrylic acid hydrogel loaded with kaolin and sepia ink. Cellulose, 24, 69, 2017.
215. Maity, J. and Ray, S.K., Removal of Cu (II) ion from water using sugar cane bagasse cellulose
and gelatin based composite hydrogels. Int. J. Biol. Macromol., 97, 238, 2017.

You might also like