You are on page 1of 46

View Article Online

View Journal

Nanoscale
Accepted Manuscript

This article can be cited before page numbers have been issued, to do this please use: Y. Xue, Z. Mou
and H. Xiao, Nanoscale, 2017, DOI: 10.1039/C7NR04994C.

Volume 8 Number 1 7 January 2016 Pages 1–660 This is an Accepted Manuscript, which has been through the
Royal Society of Chemistry peer review process and has been
accepted for publication.
Nanoscale Accepted Manuscripts are published online shortly after
www.rsc.org/nanoscale

acceptance, before technical editing, formatting and proof reading.


Using this free service, authors can make their results available
to the community, in citable form, before we publish the edited
article. We will replace this Accepted Manuscript with the edited
and formatted Advance Article as soon as it is available.

You can find more information about Accepted Manuscripts in the


author guidelines.

Please note that technical editing may introduce minor changes


to the text and/or graphics, which may alter content. The journal’s
ISSN 2040-3364 standard Terms & Conditions and the ethical guidelines, outlined
PAPER
in our author and reviewer resource centre, still apply. In no
Qian Wang et al.
TiC2: a new two-dimensional sheet beyond MXenes
event shall the Royal Society of Chemistry be held responsible
for any errors or omissions in this Accepted Manuscript or any
consequences arising from the use of any information it contains.

rsc.li/nanoscale
Page 1 of 45 Nanoscale

Nanocellulose as Sustainable Biomass Material: Structure, Properties, Present Status and


Future Prospects in Biomedical Applications
View Article Online
DOI: 10.1039/C7NR04994C

Yan Xue,*a,b Zihao Mou,a and Huining Xiao*b


a
School of Chemistry and Chemical Engineering, Oil & Gas Field Applied Chemistry Key Laboratory of Sichuan Province,
Southwest Petroleum University, Chengdu 610500, China
b
Department of Chemical Engineering, University of New Brunswick, Fredericton, NB E3B 5A3, Canada

Nanocellulose, extracted from the most abundant biomass material ―― cellulose, has proven
to be an environmentally friendly material with excellent mechanical performance owing to its
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

unique nano-scaled structure, and has been used in a variety of applications as engineering and

Nanoscale Accepted Manuscript


functional materials. The great biocompatibility and biodegradability, in particular, render
nanocellulose promising in biomedical applications. In this review, the structure, treatment
technology and properties of three different nanocellulose categories, i.e., nanofibrillated
cellulose (NFC), nanocrystalline cellulose (NCC) and bacterial nanocellulose (BNC), are
introduced and compared. The cytotoxicity, biocompatibility and frontier applications in
biomedicine of the three nanocellulose categories were the focus and detailed in each section.
Future prospects concerning the cytotoxicity in applications and industrial production of
nanocellulose are also discussed in the last section.

1. Introduction to celluloses and nanocelluloses


Cellulose, the most widely distributed and abundant natural polymers on earth, is a key
source of renewable material chiefly in the form of plant fibre. The annual production of
cellulose is estimated at more than 7.5 × 1010 tons 1. As an important energy source and
chemical raw material, cellulose has been extensively used due to their non-toxicity,
non-pollution, biodegradability, biocompatibility, facile modification and renewability.
Regarding to the molecular structure, cellulose is characterized as a polysaccharide consisting
of glucose units (Figure 1). Constituting the main component of the plant cell walls, cellulose
represents over 50% of the carbon content in the botanic field. The cellulose content of wood is
2, 3
40~50% , accompanying which is the hemicellulose and lignin. The cotton fibre is the
natural source of the purest cellulose, of which the cellulose content is over 90% 4, 5.
Nanoscale Page 2 of 45

View Article Online


DOI: 10.1039/C7NR04994C
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


Figure 1. Schematic diagram of cellulose hierarchical structure: from resources to molecule chains

In addition to green plants, the tunicates and some species of invertebrates, amoeba 1,
oomycetes 6 and bacteria can produce cellulose with the same structure at the molecular level.
The major difference is no accompanying hemicellulose or lignin obtained from the latter ones.
The elementary cellulose chains, i.e., the polysaccharide chains, assemble in the form of
ordered parallel layers into elementary fibrils, either in plant cell walls or produced by bacteria.
Linked by van der Walls forces and strong intra- and/or intermolecular hydrogen bonds, the
7, 8
cellulose chains are tightly aggregated together with a lateral dimension of 3 ~ 5 nm . Each
elementary fibril is a bundle of cellulosic crystals along the fibril axis alternated with
amorphous domains. Bundles of elementary fibrils further constitute cellulose microfibrils with
a cross-section diameter of 5 ~ 20 nm and length of several micrometers, depending on their
9, 10
origin . As the most basic unit for constituting the plant cell wall, the cellulose microfibrils
are enclosed by the matrix components, e.g., hemicellulose, pectin and lignin, and assemble
into larger bundles, i.e., the cellulose fibres. In plant cell walls, cellulose is constituted within
the cellular hierarchical organization as concentric layers. The cell walls are separated by
middle lamellas from each other, of which the outermost layer is the primary cell wall followed
by the secondary cell wall. Surrounding the lumen, the secondary cell wall is divided into the
11-15
inner layer, the middle layer and the outer layer in sequence . Within these layers, cellulose
microfibrils are packed in the pattern of helical thread and orient at characteristic angles, which
Page 3 of 45 Nanoscale

varies depending on the cell wall layer and the plant type (Figure 1).
As the skeletal component maintaining the framework of cell walls, cellulose presents a
View Article Online
DOI: 10.1039/C7NR04994C

certain extent of mechanical strength arising from the mentioned hydrogen bonds. The
hydroxyl groups on the cellulose chains, by forming hydrogen bonds, bundle the cellulose
microfibrils together and structure the crystalline component in the ordered domain. The
crystalline component can be isolated from plant cell walls via various chemical treatments.
The resulting crystalline fragments have a particle size of nanometer, named nanocrystalline
cellulose (NCC) (Figure 2a) 16, 17, for which “cellulose nanocrystal” 2, 18, “cellulose whisker” 19,
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

20-22 23, 24 25
“nanowisker” , “nanowire” and “nanorod”

Nanoscale Accepted Manuscript


are also used as synonyms due to their
slender and rod-like shape.
Among the cellulose hierarchical structures, the nanofibrillated cellulose (NFC) is assigned
to cellulose particles with nano-size as NCC. However, different from NCC, which renders
cellulose with stiffness, NFC provides certain flexibility generated from the amorphous
domains (Figure 2b). NFC can form entangled network with alternating amorphous and
crystalline domains. It is worth noting that the term “microfibril” represents the most basic
entity, which can be isolated from cell walls, while NFC is usually composed of aggregates of
cellulose microfibrils with diameters below 100 nm 26. In the literature, the terminologies such
27, 28 29-31
as “microfibrillated cellulose” , “microfibril aggregates”, “nanofibrillated cellulose” ,
32, 33 34, 35 36-38
“nanofibril” , “nanofibre” and “nanofibrillar cellulose” are also employed to
define NFC, and in addition “microfibril” is sometimes irregularly used to describe the same
material.
Nanoscale Page 4 of 45

View Article Online


DOI: 10.1039/C7NR04994C
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


Figure 2. Micro morphology of nanocellulose: (a1) SEM image of NCC treated via lyophilization 39;
(a2) TEM image of NCC 40; (a3) 2D AFM image of softwood NCC 41; (b1) FE-SEM image of NFC from
beech wood pulp 42; (b2) Cryo-TEM image of frozen NFC solution after cationization 43; (b3) AFM
image of NFC 42; (c1) SEM image of freeze-dried BNC 44; (c2) TEM image of BNC 45; (c3) AFM image
of BNC 46. Reprinted with permissions from the aforementioned references.

Structurally similar to NFC, bacterial cellulose is another primary source of cellulose with
nanoscale, consisting of microfibrils constituted by elementary fibrils. Within the bacterial
cellulose, the microfibrils assemble into ribbons with twists having a width of 20 ~ 50 nm
(Figure 2c). Accordingly, bacterial nanocellulose (BNC) is often termed to reflect the
nano-dimension of bacterial cellulose. BNC, a secretion of bacteria under specific culturing
conditions, is produced for their survival, e.g., as a barrier to ultraviolet light, or to defend itself
against fungi, yeasts and other organisms 47. Aside from the general properties of cellulose, the
nano-sized natural cellulose presents some unique features due to its nano-scale, including high
aspect ratio, large specific surface area, high stiffness and tensile strength, low density and low
48-50
thermal coefficient of expansion . Beyond possessing the advantageous performances of
nanomaterials, the natural nanocellulose is low cost, completely renewable and highly
biocompatible compared to the synthetic ones. The nanocellulose, triggered by its superior
features, has received tremendous attention for a variety of applications, including foods,
cosmetics, pulp and paper industry, biomedical implants, electronic chemistry, optical materials,
separation membranes and so forth. The nanocellulose can be used as an independent
functional material or reinforcement unit in composite materials. Moreover, the nanocellulose
Page 5 of 45 Nanoscale

appears to have significant advantages in biomedical fields such as medical packaging,


membrane for hemodialysis, vascular grafts, wound dressing, tissue engineering, drug delivery,
View Article Online
DOI: 10.1039/C7NR04994C

owing to its virtual ubiquity with biological utility in nature, e.g., intrinsic biodegradability and
biocompatibility 37, 51-55.
The following review presents the main preparation methods, structure and performance
characteristics of different varieties of nanocellulose, and their corresponding applications in
biomedical field through examples in particular. It has been mentioned that the nomenclatures
defining the nanocellulose are not being applied uniformly in the literature, even anomalies
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


exist. In this vein, the terms of NFC, NCC and BNC are employed to represent the
aforementioned three members of the nanocellulose family. Correspondingly, this review is
divided into three sections regarding to the three categories of nanocellulose.

2. Nanofibrillated cellulose (NFC)


The earliest production of NFC is traced back to 1983 when Turbak et al. and Herrick et al.
10
extracted NFC from wood via high pressure homogenization (HPH) . The most important
lignocellulosic material used for producing NFC is wood, while some other cellulosic sources,
such as agriculture crops, water plants, grasses and their by-products, are representing
alternative raw materials for NFC extraction (Table 1) 15, 56. Compared to wood, the latter ones
contain less lignin components; and the microfibrils could be more easily isolated from their
primary cell walls. Accordingly, environmental benefits resulting from the low energy
consumption for both cultivation and microfibril extraction are obvious for non-wood based
NFCs due to their short growth periods and easier delignification. Without regard to the
cellulosic sources, NFC is manufactured from fibre pulp mainly via mechanical treatments,
overcoming the interfibrillar hydrogen bonds and releasing the NFC. Although the HPH is the
most widely conducted technique for both lab-scale and industrial scale production of NFC,
other strategies including microfluidization, micro-grinding, cryocrushing and ultrasonication
29, 57
are also applied according to the fibre sources , production conditions and specific quality
requirements. Since the mechanical delamination of fibre results in pulp clogging and high
energy consumption, certain pretreatments might be conducted to address the problems. By
weakening the hydrogen bonds, modifying using repulsive charges, and/or decreasing the
degree of polymerization, the treatments such as mechanical refining, alkaline hydrolysis,
Nanoscale Page 6 of 45

enzymatic hydrolysis, acid hydrolysis, steam explosion, 2,2,6,6-tetramethylpiperidine-1-oxyl


(TEMPO)-mediated oxidation, carboxymethylation and acetylation have been developed to be
View Article Online
DOI: 10.1039/C7NR04994C

employed prior to the key mechanical treatments 27, 58, 59. Assisting the mechanical treatment by
performing the pre-treatment strategies can facilitate the fibre delamination and
correspondingly, lower the energy demand for fibrillation. The mechanical treatments and
pre-treatment strategies are detailed in Table 1. Further to these two treatments, the
post-treatment of NFC, known as the NFC modification, has been proposed for improving the
compatibility and/or homogeneous dispersion within matrices. In essence, the post-treatment
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


approach applied to NFC is based on its potential application, endowing it with specific
properties. Accordingly, the NFC post-treatments have been described in the following context
interspersed within individual biomedical applications, rather than being introduced
specifically.

Table 1. Typical sources of different nanocelluloses and technologies for extracting purified
nanocelluloses.
Type of nanocellulose Common sources Processing technology
60, 61 35, 62 63, 64
NFC Wood , cotton , hemp , High-pressure homogenization 83, 84
65 66
potato tuber , wheat straw , sugar Microfluidization 42, 85, 86
beet 67, 68, flax 69, 70, bagasse 71, palm 72, Ultrasonication 87-89
abaca 73, Swede root 74, soybean stock Cryocrushing 66
75
, banana rachis 76, alfa 77, pine 78, Micro-grinding 90-92
tunicate 79, 80, algea 81, 82 Refining 93-95
Steam explosion 96-98
TEMPO-mediated oxidation 99-101
Enzyme-assisted hydrolysis 102-104
Acid hydrolysis 105, 106
NCC Wood 13, 41, cotton 107-109, wheat straw Acid hydrolysis 107, 123-125
110
, sugar beet 111, flax 112, hemp 113,
palm 72, ramie 114, 115, sisal 116, 117, alfa
77
, jute 118, tunicate 119-122,
BNC Acetobacter 126-128, Biosynthesis via agitated fermentation
Acanthamoeba 52, 129, Agrobacterium 136-139

130, 131
, Rhodobacter 132, 133, Zoogloea Biosynthesis via stationary fermentation
134, 135 140-143

2.1. Cytotoxicity of NFC


With respect to biomedical applications, the toxicity of materials among others has to be
concerned and clarified. Although NFC is extracted from cellulosic sources, biomasses with no
or low toxicity, the fabrillation approach, fibre dimension, surface modification,
Page 7 of 45 Nanoscale

hydrophilicity/hydrophobicity and aggregation, etc., might influence their cytotoxicity and


biocompatibility. Above all, the nanoscale has been recognized as a potential factor generating
View Article Online
DOI: 10.1039/C7NR04994C
144, 145
the toxicity of materials . Some innoxious bulk materials gain toxicity since they are
miniaturized to nanoscale. For instance, carbon black is relatively inert in its bulkier forms
whereas its nanoparticles, e.g., fullerene and carbon nanotubes, have high pulmonary toxicity
146,
probably due to their facile uptake and more extensive interaction with biological systems
147
. Proceeding from this aspect, the biocompatibility and toxicity to human body and
environment of the nanocellulosic materials cannot be predicted only based on their chemical
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


structures, and specific studies have to be conducted to redetermine their human toxicity and
ecotoxicity. Regarding the potential risks of NFC, some European projects, e.g., “NanoFun”
148, 149
and “Scale-up Nanoparticles in Modern Papermaking” , showed deep concerns, and a
complete literature study was reported in 2010 150, where it was emphasized that no finding on
151, 152
the toxicity, toxicokinetics or ecotoxicology of NFC had been reported. Vartiainen et al.
tested the health and environmental safety of NFC processed by friction grinding and spray
drying using mouse, human cells and luminescent bacteria as the tested models, and concluded
that both the test NFC appeared hazard to neither the human health nor ecological environment.
A study on the safety of NFC from bleached hard wood kraft pulp showed neither sublethal
effect in the RNA inhibition assay nor genotoxic indication in the Ames test was detected in
vitro, and in vivo test with nematode presented no cytotoxic effect either 153, 154. There are some
other papers exclusively devoted to the toxicity issue of NFC, although only a few number, and
none of them showed evidence of toxic effect on human health or environment 155-157. Based on
the corresponding studies so far, all the results have been positive. Nonetheless, the toxicity and
biocompatibility of NFC products, especially those for biomedical applications, have to be
taken into account specifically due to the great variety in their chemical structures and
properties.
2.2. NFC used in antimicrobial composites
The large surface area to volume ratio and high surface charge density of NFC, especially
the NFC pre-treated by TEMPO-mediated oxidation of which partial hydroxyls are replaced by
carboxyls, facilitate it stabilizing metallic nanoparticles as a template. It has been widely
known that the colloidal Ag nanoparticles (AgNPs) present highly antibacterial activity, while
their aggregation under physiological conditions impedes their applications. Aiming at this
Nanoscale Page 8 of 45

problem, several researchers have taken advantage of NFC for stabilizing AgNPs. For example,
158
Wang et al. developed NFC-stabilized AgNPs (NFC-AgNPs) by reducing the Ag+Viewinto
Article Online
DOI: 10.1039/C7NR04994C

AgNPs in the presence of the NFCs capping agent, which were derived from rice straw
cellulose via TEMPO-mediated oxidation. Presenting excellent performance for dispersing and
stabilizing AgNPs in physiological growth medium, the NFCs were conducive to the AgNPs
exerting their antibacterial activity, by which the production of bacterial extracellular
polysaccharides was induced and bacteria clustered correspondingly. NFC was formed into
32 159
films by Díez et al. for incorporating Ag nanoclusters (AgNCs) . By immersing the NFC
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


films into AgNC solution, the NFC-AgNC composites were constituted from which the AgNC
was steadily released. The antibacterial halo test of the NFC-AgNC composites against E. coli
showed the released fluorescent AgNC completely inhibited the bacterial growth in an agar
61
area 5.4 times larger than the film area . Some other forms of NFC-AgNC composites, e.g.,
160 161, 162
NFC-AgNC paper , hydrogel and aerogel , have been developed and presented
antibacterial properties. Given the color issue caused by the AgNPs, Martins et al. 163 used ZnO
nanoparticles as an alternative to the AgNPs for preparing the antibacterial composite. Similarly,
the ZnO nanoparticles were located on the NFC surfaces assisted by the polyelectrolytic
macromolecular linkers. The NFC-ZnO composites, as fillers used in the starch-based coating
formulation, endowed the coated paper sheets with antibacterial activity accompanied by the
improvements in their mechanical performance and air permeability. The antibacterial activity
study elucidated that the antibacterial mechanism of the ZnO nanoparticle was derived from
both the photoactivity of the semiconductor and the reactive oxygen species on the particles
surfaces. Another metal oxide nanoparticle TiO2 nanoparticle has also been immobilized onto
the NFC surfaces for rendering NFC antibacterial. Missoum et al. prepared a series of anionic
164 165
NFC via esterification or carbanilation , all of which showed bacteriostatic activity or
even bactericidal effect owing to the fatty chains of the grafted molecules. In addition, the
functionalization of the grafted NFC with TiO2 generated a synergistically antibacterial effect,
166 167
significantly enhancing the antibacterial performance of the grafted NFC . Xiao et al.
developed TiO2/chitosan composite coated NFC via a layer-by-layer process followed by the
deposition of AgNPs via UV irradiation. The cable-like NFC composite exhibited extraordinary
antibacterial activity against both Gram-positive and Gram-negative bacteria due to the coated
TiO2, chitosan and AgNPs.
Page 9 of 45 Nanoscale

In addition to incorporating AgNPs as the antibacterial agent, some cationic compounds,


e.g., some quaternary ammonium salts (QAS), have been utilized for rendering NFC
View Article Online
DOI: 10.1039/C7NR04994C
168
antibacterial. Syverud et al. modified NFC films by adsorbing the cationic surfactant
cetyltrimethylammonium bromide (CTAB) onto the film surfaces, and quantified the surface
porosity of the NFC films and the surfactant distribution on the NFC surfaces. Both the
increase of NFC fibrillation degree and the introduction of the CTAB surfactant induced the
hydrophobicity increase of the NFC films. As a membrane-active agent targeting at the
bacterial cytoplasmic membrane, the CTAB compound was suggested to impart the NFC films
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

169

Nanoscale Accepted Manuscript


with antibacterial activity . By an adsorption-curing process, a quaternary ammonium salt
octadecyldimethyl(3-trimethoxysilylpropyl)ammonium chloride (ODDMAC) was grafted onto
170
the NFC surfaces, from which the antibacterial films were prepared . The QAS surface
modified NFC films presented excellent antibacterial performance against both Gram-positive
and Gram-negative bacteria. Meanwhile, no ODDMAC was released from the NFC films in the
inhibitory zone testing, revealing the contact disinfection mode of the QAS-modified NFC
films with nonleaching ODDMAC. Being different from the aforementioned QAS adsorbed or
modified NFC which underwent a post-treatment of cationization, Saini et al. modified
cellulose fibres using 2,3-epoxypropyl trimethylammonium chloride (EPTMAC) via a
171
cationization pre-treatment followed by a grinding fibrillation . Since the cationization
pre-treatment imparted the fibrils with highly swollen outer layer, the following mechanical
fibrillation was effectively facilitated, showing a fivefold reduction of the energy consumption
172
in fibrillation . Moreover, the EPTMAC-modified NFC derived from the pre-treatment
displayed substantial antibacterial capacity with the degree of substitution (DS) higher than
0.18. Such cationic NFC was promised to be applied as a contact antibacterial nanomaterial
with tunable antibacterial activity depending on the DS of NFC.
2.3. NFC-based cell culture scaffold
Possessing the advantages of biocompatibility, biodegradability, facile chemical
modification, ultralow density and high mechanical strength, the NFC has been proposed as an
173
excellent candidate for cell culture scaffold. Hua et al. modified Cladophora nanocellulose
(CC) and the NFC fibrillated from softwood and studied their biological response through
indirect and direct contact assays. The indirect cytotoxicity test showed that neither the anionic
nor cationic CC, modified by the TEMPO-mediated oxidation and EPTMAC condensation
Nanoscale Page 10 of 45

respectively, was cytotoxic to the human dermal fibroblasts. In comparison to the tissue culture
material Thermanox, the anionic CC promoted the fibroblast adhesion and exhibited cell
View Article Online
DOI: 10.1039/C7NR04994C

viability. Likewise, both the anionic and cationic NFC modified by the carboxymethylation and
EPTMAC condensation respectively presented non-cytotoxic, whereas the cationic NFC was
more cytocompatible compared to the unmodified or the anionic NFC in the cell adhesion and
viability evaluation. The different cytocompatibility between the CC and NFC with different
charges was predicted to be a result aggregated from the specific area, surface topology,
hydrophobicity/hydrophilicity and pore size distribution. In regard to the three-dimensional cell
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

174 175

Nanoscale Accepted Manuscript


culturing, Bhattacharya et al. and Cai et al. developed NFC hydrogel and NFC aerogel
microspheres as the cell culture scaffolds, respectively. The injectable NFC suspensions at high
stress allowed facile immersing cells into the system owing to their rheological properties and
reversible gelation, while after injection, the spontaneously formed NFC hydrogel was tuned
into the cell culture scaffold with certain mechanical strength for cell growth and differentiation.
The single component NFC scaffold without additive growth factors displayed cellular
biocompatibility and differentiation of the human hepatic progenitor HepaRG cells and retinal
pigment epithelial cells (Figure 3I). The NFC aerogel microspheres were fabricated by spraying
and atomizing the NFC aqueous gel using a steel nozzle followed by freeze-drying. With a bulk
density as low as 1.8 mg·cm-3, the ultralight and high porous aerogel microspheres possessed
high water uptake capacity, i.e., 100 g·g-1. In the cell culture assay the NFC aerogel
microspheres with an across-linked stable structure facilitated the 3T3 NIH cells attachment,
differentiation and proliferation, presenting non-cytotoxic and biocompatible (Figure 3II). As a
specific application of the NFC cell culture scaffold, a NFC nanocomposite scaffold was
176
developed as ligament and tendon substitute by Mathew et al. . The scaffold was fabricated
by partially dissolving the NFC networks in ionic liquid for different time intervals, where the
dissolved NFC constituted the matrix phase and the undissolved and/or partially dissolved NFC
constituted the reinforcing phase. With the mechanical strength as high as the natural ligaments
and tendons, the as-prepared NFC nanocomposites appeared to be a good candidate for
supporting the adhesion and growth of both the human ligament cells (HLC) and human
vascular endothelial cells (HEC) (Figure 3III), promising for their application of artificial
ligaments and tendons.
Page 11 of 45 Nanoscale

View Article Online


DOI: 10.1039/C7NR04994C
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


Figure 3. Cell adhesion and proliferation on NFC-based cell culture scaffold. (I) a. Confocal laser
scanning microscope images of HepaRG and HepG2 cells cultured in NFC hydrogel (Scale bar 10 μm);
b. Albumin secretion of HepaRG and (c) HepG2 cultured in NFC hydrogel and PuramatrixTM (PM).
(II) a. SEM image of aerogel microsphere with 1.5% of NFC; b. MTT data of 3T3 NIH cells seeded on
cellulose aerogel microspheres. (III) a. HLC proliferation on NFC scaffold after 1-day culture; b. HEC
proliferation on NFC scaffold after 3-day culture; c. Data of HLC proliferation on NFC scaffold.
Reprinted with permission (I) from ref. 174, (II) from ref. 175 and (III) from ref. 176.

2.4. NFC composites for DNA extraction


The NFC paper material with intrinsic large surface area and high mechanical strength was
177, 178
firstly explored by Mihranyan et al. to be employed as an electrochemically controlled
ion-exchange material, which was prepared by coating a thin polypyrrole (PPy) layer onto the
NFC paper derived from Cladophora or wood cellulose. Subsequent to the PPy-NFC
179-181
composites showing promising active ion-extraction performance , the biocompatible
composites were further investigated as hemodialysis materials and DNA extraction membrane.
Combining passive ultrafiltration and potential-controlled ion exchange, the PPy-NFC
composites were able to effectively extract uraemic toxins and improved the platelet adhesion
and thrombin generation due to a heparin coating, which was comparable to the
Nanoscale Page 12 of 45

182 183
haemocompatible polymer polysulphone . Ferraz et al. studied the effect of a series of
processing parameters including rinsing, extraction and aging on their in vitro DOI:
and10.1039/C7NR04994C
in View
vivoArticle Online

cytotoxicity and electroactivity of the PPy-NFC composites. Results demonstrated that the
composites aging exerted a significant negative effect on their biocompatibility, indicating the
degradation of PPy should be controlled for developing stable and biocompatible electroactive
hemodialysis membranes. To be applied as a DNA extraction membrane, the PPy-NFC
composites were studied in the term of inducing the extraction of 6-FAM tagged ss-DNA
hexamers from a borax buffer solution 184. The fluorescence measurements proved the ss-DNA
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


hexamers were successfully extracted upon the PPy oxidation using phosphomolybdic acid as
the oxidant. In addition, the energy-filtered transmission electron microscopy assay clearly
showed the ss-DNA hexamers penetrated into the PPy coating induced by the electrochemically
controlled extraction even in the presence of competing ions.
2.5. NFC-reinforced biocomposites
The excellent mechanical properties of NFC, i.e., high stiffness and strength, have favored
them promising reinforcement ability which might be improved via the formation of hydrogen
bonds 185. In addition, the rich hydroxyl groups on its surface favor the variety of modifications
towards NFC. In our previous work, NFC have been used to improve the compatibility,
186-188 189
morphology and mechanical performance of composites . Saito et al. attempted to
measure the mechanical strength of single NFC fibrils from wood and sea tunicate respectively,
which were obtained by the TEMPO-mediated topochemical modification. The mean strength
resulted from the sonication-induced cavitation of filamentous structures upon hydrodynamic
stresses was 1.6 ~ 3 GPa for the wood NFC and 3 ~ 6 GPa for the tunicate NFC, comparable to
that of the commercially available carbon nanotubes. The high-pressure steam treatment and
acid hydrolysis of pineapple leaf fibres produced NFC with widths ranging from 5 to 15 nm
under a high yield. Such fibrillated NFC was utilized as a reinforcement component in
polyurethane (PU)-NFC nanocomposites by stacking the NFC mats between PU films, which
showed a strength increase of 3 folds and stiffness enhancement of 26 folds due to the addition
190
of NFC . Furthermore, the as-prepared composites displayed haemocompatibility, fatigue
resistance, high biological durability both as heart valves in a rat implant model and vascular
prostheses in a patient with multiple endocrine neoplasia 2B. NFC was also employed as a
reinforcing agent in composite hydrogels for biomedical applications. For example, a
Page 13 of 45 Nanoscale

poly(N-vinyl-2-pyrrolidone) (polyNVP) hydrogel reinforced by carboxymethylated NFC


powder, as a potential alternative to the native human nucleus pulposus, was developed byViewUV
Article Online
DOI: 10.1039/C7NR04994C
191
polymerization of NVP using Tween 20 trimethacrylate as the cross-linker . In addition to
their biocompatibility 192, which showed 90% of cell confluence without any abnormal cellular
morphology being detected, the NFC-reinforced polyNVP hydrogels possessed adequate
swelling ratio and mechanical strength which were responsible for restoring the annulus
fibrosis loading and the intervertebral height 193. These NFC-reinforced biocomposites could be
employed as implant materials in a wide range of applications such as artificial skins, vascular
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


grafts, urethral catheters, articular cartilage, cardiovascular implants, etc., as well as medical
disposables such as wound dressings, medical bags, surgical gloves and gowns.
2.6. NFC used in drug delivery system
Study on the drug delivery capacity of NFC has received limited interest probably due to
the high hydrophilicity of NFC. With the aid of a hydrophobin coupled with two cellulosic
194
binding domains, Valo et al. bound hydrophobic drug nanoparticles onto the NFC surfaces
for drug stabilization during the process and storage of the nanoparticle formulation. With
itraconazole (ITR) as a model drug, the protein coated drug nanoparticles were able to be stored
stably for over 10 months by being immobilized into the NFC matrix; and the subsequent
freeze-drying process showed no morphological alternation of the drug due to the protection of
NFC. Furthermore, the release rate of ITR in the presence of NFC matrix was as high as from
the particles themselves, improving the in vivo performance of ITR. Also assisted by protein
molecules, a superparamagnetic hierarchical material was fabricated using the NFC of
195
commercially available filter paper as the template . By a facile silica replication approach,
ferritin molecules were firstly assembled onto the silica-coated NFC, followed by the ferrous
substitution of the ferric cores through a core construction process, and finally oxidation of the
ferrous cores into γ–Fe2O3 as well as removal of both the protein shells and NFC via a
196-198
post-calcination treatment . Such fabricated nanostructured material, constituted of
cross-linked hierarchical network of silica nanotubes with γ–Fe2O3 nanoparticles embedded in
the silica walls, presented significant superparamagnetism and was promising for biomedical
applications, e.g., targeted drug delivery and magnetic resonance imaging.

3. Nanocrystalline cellulose (NCC)


Nanoscale Page 14 of 45

Compared to NFC, the whisker shaped NCC has a high degree of crystallinity ranging from
49, 199
54% to 88% , therefore, a key stage during the NCC production is the destruction and
View Article Online
DOI: 10.1039/C7NR04994C

removal of the disordered or paracrystalline components alternated between crystalline


domains within cellulose fibres. Lignocellulosic biomass processed through alkali and
bleaching treatments was hydrolyzed by sulfuric acid, from which Rånby obtained NCC
suspensions in 1949 for the first time 1. The amorphous regions within the cellulosic fibres are
susceptible to acid which are preferentially removed, whereas the crystalline ones with a higher
200, 201
resistance to acid remain intact . Up to now, the main strategy for extracting NCC from
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

202

Nanoscale Accepted Manuscript


cellulosic fibres is still depending on the acid hydrolysis (Table 1), where sulfuric and
203, 204
hydrochloric acids are extensively used and alternatively, hydrobromic and phosphoric
205-207
acids have been also involved. Following the acid hydrolysis, dilution and repeated
washing with water accompanied by centrifugation and extensive dialysis are conducted for the
removal of free acidic molecules and/or impurities as thoroughly as possible. Thereafter, the
NCC particles are dispersedly stabilized as uniform suspensions by a mechanical treatment, e.g.,
sonication. As can be seen the crucial stage in the NCC fibrillation is a chemical treatment, i.e.,
the acid hydrolysis, whereas the most important process for the NFC extraction is a mechanical
treatment, i.e., high-pressure homogenization. In regard to the cellulosic sources for extracting
NCC, wood and plants are the most important ones as in the case of NFC, besides which
tunicates have become a favored source for NCC production due to their high crystallinity 48, 208,
209
(Table 1).
3.1. Toxicology and ecotoxicology of NCC
In order to be used in biomedical applications, NCC and its corresponding composites,
most important of all, have to be non-toxic to human body. Just like in the case of NFC, there
have been only a few studies conducted specially on the toxicity of NCC. Through
ecotoxicological testing with rainbow trout hepatocytes and several aquatic species, it was
found neither carboxyl methyl cellulose nor NCC presented genotoxicity, concluding NCC has
210, 211 212-214
low toxic potential towards environment . Sunasee et al. prepared two types of
polycation grafted NCC derived from 2-aminoethyl methacrylate hydrochloride and
N-(2-aminoethylmethacrylamide) hydrochloride respectively, of which the cytotoxicity results
evaluated via MTT assay against mouse and human cells were below the toxicity level for
biomedical materials. NCC also presented significantly lower intracellular toxic effects and
Page 15 of 45 Nanoscale

215
inflammatory response compared to multiwalled carbon nanotubes and crocidolite fibers .
Other studies regarding the cytotoxicity and/or ecotoxicity of NCC have notDOI:
revealed a
View Article Online
10.1039/C7NR04994C

potential risk to human health and environment either 216, 217.


3.2. NCC in antibacterial applications
In the field of antibacterial applications of NCC, AgNPs, possessing excellent antibacterial
218, 219
properties, are usually involved . As with NFC, NCC could reduce and stabilize the
220, 221
AgNPs as the capping or dispersing agent. Xiong et al. developed a facile approach for
preparing stable AgNPs using sulfated NCC, of which the sulfated groups played an important
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


role in stabilizing AgNPs and controlling their particle size. The morphology of the AgNPs
could be feasibly adjusted according to the AgNO3 concentration, resulting in silver
nanospheres at low concentration and dendritic AgNPs, showing a higher antibacterial activity
compared to the former one, at relatively high concentration. Besides silver, some other metals
222 223 224, 225 226-228 229
including nickel , palladium , platinum , gold and copper have also been
reported using NCC as the stabilizing template for forming NPs. Owing to the NCC, the
antibacterial metal NPs are stabilized and correspondingly, their antibacterial performance
reaches their full potential. Moreover, NCC possessing intrinsic high modulus and strength is
230
an excellent reinforcing agent. Concerning these two aspects, Liu et al. incorporated the
nanocomposites constituted of carboxylated NCCs and AgNPs within waterborne polyurethane
(WPU), improving the mechanical strength of and introducing antibacterial property to the
231, 232
WPU matrix as dual-functional fillers. Fortunati et al. developed a nanocomposite film
combined with surfactant-modified NCCs as the nucleating agents and AgNPs as the biocides
respectively via melt extrusion of poly(lactic acid) (PLA) followed by cast-filming. The
presence of the NCCs addressed the slow crystallization of PLA and imparted the PLA matrix
with highly mechanical and good thermal properties in both binary and ternary systems. In
addition to the AgNPs, widely studied for rendering NCCs antibacterial, a photobactericidal
233
porphyrin has been investigated for developing antibacterial NCCs . The porphyrin was
bonded onto the NCC surfaces via “click chemistry”, generating insoluble crystalline
NCC-Porphyrin which presented excellent photodynamic inhibition against Gram-positive
bacteria and relatively low activity against Gram-negative bacteria.
3.3. NCC-based drug carriers
Being nano-sized, biocompatible, biodegradable and facile-modified, the NCCs have been
Nanoscale Page 16 of 45

promising as carriers of bioactive molecules for biomedical applications. The hydrophilicity of


NCC surfaces is able to hamper the opsonic proteins adsorption, which interferes withViewthe
Article Online
DOI: 10.1039/C7NR04994C

phagocytosis for the removal of carriers from bloodstream 234. Accordingly, NCCs are expected
to possess a longer half-life in the serum than hydrophobic nanoparticles, e.g., hydrophobic
ligand nanoparticles and carbon nanotubes. Although the NCCs with size ranging from 50 to
150 nm are too large for their removal from bloodstream through the renal system, they are
235
small enough for their clearance via the mononuclear phagocytic system . Furthermore,
NCCs can be endogenously removed from the body by the systemic administration of
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

236-238

Nanoscale Accepted Manuscript


cellulolytic enzymes and in situ degradation to glucose . Since the NCCs have a large
number of hydroxyls on the surfaces, the electrical property and hydrophilicity/hydrophobicity
of NCCs can be converted via facile surface modification, according to the specification of
drug delivery application. Jackson et al. 239 studied the drug release performance of pristine and
CTAB-modified NCCs, respectively. The NCCs, obtained by acid hydrolysis of softwood, were
found to bind a large quantity of water soluble drugs, i.e., tetracycline and doxorubicin, and
release them over a one-day period. On the contrary, the CTAB-modified NCCs could bind
hydrophobic drugs, i.e., the anticancer docetaxel, paclitaxel and etoposide, and release them
with control over a two-day period. For evaluating the cellular uptake ability of the modified
NCCs, fluorescein was adsorbed onto their surfaces and the KU-7 bladder cancer cells were
used as the cell model, demonstrating highly efficient delivery of the CTAB-modified NCCs
240
into the cellular cytoplasm (Figure 4A). Roman et al. confirmed the potential of NCCs
employed in targeted drug delivery systems, which was different from the former one that the
limited uptake of fluorescein-5’-isothiocyanate (FITC) labeled NCCs by the non-activated
human brain microvascular endothelial cells indicated their lack of untargeted uptake. However,
the effect of surface charge characteristics of fluorescent labeled NCCs on their cellular uptake
was ignored in Roman’s study. By investigating the cellular uptake behavior and cytotoxicity of
two fluorescent-labeled NCCs, i.e., FITC and rhodamine B isothiocyanate (RBITC)-labeled
241, 242
NCCs, Mahmoud et al. found that the surface charge of NCCs significantly interfered
with their internalization pattern. The positive charges on the RBITC-labeled NCCs triggered
their escape from endosomes into the cytoplasm, while the negatively-charged FITC-labeled
NCCs were restrained within the cellular endosomes, being unable to achieve internalization.
To be employed in a multi-particulate oral drug delivery system, negatively charged NCCs
Page 17 of 45 Nanoscale

were combined with positively charged chitosan through turbidimetric titration, resulting in
structure-controllable polyelectrolyte-macroion complex (PMC) 243. The particle size and shape
View Article Online
DOI: 10.1039/C7NR04994C

of the as-prepared PMC depended on the NCC concentration, mixing sequence and mixing
ratio of the two components. Higher NCC concentration generated larger and/or highly
aggregated PMC particles. Titration of the chitosan solution with NCC suspensions eventually
resulted in a tightly coiled conformation of the PMC particles while the titration in reverse
order induced loose ends of the chitosan protruding from the PMC surfaces. Higher chitosan
content formed spherical particles whereas lower chitosan content formed non-spherical ones.
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


Therefore, by regulating the titration conditions, the conformation of the PMC particles was
able to be facilely tuned to meet the requirements for the multiparticulate oral drug delivery
application. Considering their excellent mechanical property, the NCCs have been used as a
bifunctional component in alginate-based nanocomposite microspheres for improving the
244
mechanical strength as well as controlling their drug release . The alginate-based
nanocomposite microspheres, with semi-interpenetrating polymeric networks (Semi-IPN)
formed via an ionic crosslinking, showed more consistent swelling patterns, increased
encapsulating efficiency and improved sustained release behavior of the model drug, as well as
higher mechanical strength owing to the introduction of NCCs (Figure 4B). In another drug
delivery system, the pharmaceutical acrylic beads, the NCCs were used as co-stabilizer during
245
the beads construction . The acrylic beads were prepared by suspension polymerization of
ethyl acrylate, methyl methacrylate and butyl methacrylate with or without NCCs, of which the
particle size and distribution were respectively smaller and narrower in the presence of NCCs.
Using propranolol hydrochloride as the model drug for the compression process of as-prepared
beads, the drug release tablets were obtained. The tablets containing NCCs presented
controlled-release profiles in the dissolution tests, suggesting those beads could be employed as
246
excipient in tablets for drug controlled release (Figure 4C). Lin et al. developed a
supramolecular hydrogel via in situ inclusion between Pluronic polymers and –cyclodextrins,
which was grafted on the NCC surfaces using epichlorohydrin as the coupling agent. The
uncovered poly(ethylene glycol) segments improved the dispersion and compatibility of NCCs,
and induced the -cyclodextrin inclusion for the formation of in situ hydrogels. It was found
that the introduction of NCCs generated a significant improvement of the structural and thermal
stability of the hydrogels. Moreover, the in vitro release of doxorubicin using the
Nanoscale Page 18 of 45

supramolecular hydrogels as drug carrier revealed the promising profile of prolonged drug
release, attributed to the “locking effect” and “obstruction effect” (Figure 4D). View Article Online
DOI: 10.1039/C7NR04994C
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


Figure 4. Various NCC-based drug delivery systems. (A) Confocal micrograph of KU-7 cells
incubated for 2 h with NCC/CTAB/fluorescein nanocomplexes and in vitro release profile of three
hydrophobic drugs from NCC/CTAB nanocomplexes in PBS. (B) SEM images of nanocomposite
microspheres of (a) SA, (b) SA/CN, (c) SA/CHW and (d) SA/SN and their drug release capacity under
different pH conditions (SA-sodium alginate, CN-cellulose nanocrystals, CHW-chitin whiskers,
SN-starch nanocrystals, scale bar-100 μm). (C) SEM image of NCC beads and release profiles of PHCl
in tablets using NCC beads as excipient. (D) Schematic diagram of supramolecular hydrogel from
NCC and cyclodextrin in situ inclusion and in vitro release profiles of doxorubicin∙HCl from hydrogels.
Reprinted with permission (A) from ref. 239, (B) from ref. 244, (C) from ref. 245 and (D) from ref. 246.

3.4. Fluorescein-modified NCC for bioimaging


Page 19 of 45 Nanoscale

As aforementioned, fluorescein can be combined with NCCs via covalent and/or


non-covalent bonding for tracking their cellular uptake behavior. Based on this, theDOI:
fluorescent
View Article Online
10.1039/C7NR04994C

NCCs are expected to be capable of being sensors and/or bioimaging for biological applications.
A wide variety of fluorescein has been used for developing fluorescent NCCs. For instance,
247
Huang et al. introduced hydrazine- and amino-substituted methylcoumarin onto NCCs
248
respectively to form methylcoumarin-modified NCCs. Yang et al. attached 1-pyrenebutyric
acid N-hydroxy succinimide ester (PSE) and FITC with NCCs via a silanization based on
249
3-aminopropyltrimethoxysilane (Figure 5A). Filpponen et al. grafted azide-containing
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

coumarin and anthracene chromophores onto NCCs through “click” reaction, obtaining

Nanoscale Accepted Manuscript


photo-responsive cellulosic nanomaterials based on the photochemical irradiation initiating
[2+2] and [4+4] cycloaddition between the bonded coumarin and anthracene units, respectively.
250
Hassan et al. developed a fluorescent NCC through the supramolecular assembly of
terpyridine-modified NCCs with 2,2’:6’,2”-terpyridine via RuIII/RuII reduction, which was
expected to be applied in bioimaging due to the facile reaction of the surface azide groups with
antigens through “click chemistry”. Two facile approaches for preparing dual fluorescent
labeled NCCs were developed by Nielsen et al., of which the first one was attaching the
isothiocyanates with cellulosic hydroxyls via forming the thiocarbamate bonds, and the second
one was carried out through a three-step procedure, i.e., introducing double bonds onto the
NCC surfaces by esterification, subsequent thiol-ene Michael addition followed by the
thiol-ene click reaction with the succinimidyl ester dyes 251. Both of the as-prepared fluorescent
NCCs presented pH-sensitivity which displayed critical fluorescence emission intensities at
specific pH conditions (Figure 5B). A study involving the fluorescent labeling of NCCs with
various surface charge densities revealed the surface charges of NCCs could interfere with the
252
attachment of fluorescein (Figure 5C) . Two sulfuric acid hydrolyzed NCCs with sulfur
contents of 0.47 wt% and 0.21 wt%, as well as hydrochloride acid hydrolyzed uncharged NCCs
were labeled with 5-(4, 6-dichlorotriazinyl) aminofluorescein (DTAF) via a one-step
nucleophilic attack of the triazinyl groups by the cellulosic hydroxyls. The surface charge
depended on labeling degree, implying the presence of the surface sulfate groups could impede
DTAF labeling. By loading the DTAF-labeled NCCs into electrospun polyvinyl alcohol (PVA)
fibres, it demonstrated that the cellulosic nanoparticles could be uniformly dispersed within the
polymeric matrix via the fluorescence microscopy detection. It is worth noting that none of the
Nanoscale Page 20 of 45

fluoresceins, introduced onto NCCs via covalent bonding, showed negative influences towards
the chemical structures and behaviors of the NCCs. A completely different DOI:
strategy for
View Article Online
10.1039/C7NR04994C

preparing fluorescent NCCs was developed by Wang et al., in which poly(9,9-dioctylfluorene)


was encapsulated within amphiphilic NCC aggregates for sensing hydrophobic analytes in
aqueous media 253. Such fluorescent NCC aggregates sensing system might be applied to detect
contaminants or in tissue engineering.
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


Figure 5. Fluorescent functionalization of NCC. (A) PSE and FITC fabricated fluorescent NCC
and corresponding confocal fluorescence images. (B) Succinimidyl ester dyes labeled NCCs and
pH-responsive fluorescence. (C) DTAF-labeled sulfated NCC and epifluorescence microscopy images
of fluorescent NCCs and electrospun PVA fibers loaded with fluorescent NCCs. Reprinted with
permission (A) from ref. 248, (B) from ref. 251 and (C) from ref. 252.

3.5. NCC for protein and/or DNA immobilization


Besides their antibacterial performance, the metal-NCC composites have also shown
protein and/or DNA immobilization capacity for detecting or stabilizing the biomacromolecules.
Page 21 of 45 Nanoscale

An AuNP-NCC composite was utilized as a scaffold for immobilizing enzymes, on which the
conjugated enzyme models, i.e., the cyclodextrin glycosyl transferase and alcohol oxidase,
View Article Online
DOI: 10.1039/C7NR04994C
254
presented high loading efficiency and well stability . The novel nanocomposites were
expected to be applicable in the industrial scale enzyme-catalyzed processes owing to the
improved enzyme loading, activity and stability, as well as their renewability and cost
effectiveness. Regarding to the DNA detection, an AgNP-NCC composite was prepared by
depositing AgNPs onto NCCs through reducing metallic cations, in which the NCC scaffolds
255
effectively prevented the aggregation of AgNPs . Working in detecting DNA, the
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


AgNP-NCC composites linked with probe DNA via amide bonds, which was subsequently
hybridized with the target DNA sequence labeled glass carbon electrodes, followed by being
immersed into a nitric acid solution to release the Ag+ from AgNPs. A linear relationship
between the logarithmic target DNA concentration and the differential pulse anodic stripping
voltammetry of the Ag attached on electrodes via electroreduction was then obtained. Also
based on the complementary sequences of DNA, a DNA-grafted NCC hybrid nanomaterial was
256
produced for potential tissue engineering applications . Theoretically, the single-strand
oligonucleotide grafted NCCs with two complementary sequences, prepared via carbodiimide
chemistry, were able to self-assemble according to the molecular recognition of the base pairs.
The formation of the DNA-NCC hybrid materials has been also confirmed by a variety of
characteristic methods, including DLS, UV-Vis and AFM. In addition, the NCCs themselves,
without metal combined, have shown excellent performance for aligning proteins in
biomolecular NMR 257.
3.6. NCC as reinforcing fillers in biocomposites
Since NCCs possess excellent mechanical properties, surface-modified and/or
non-modified NCCs have been used as reinforcing fillers in a wide range of polymeric matrices,
112, 258-266
such as PU , PLA 22, 267-276, polyethylene oxide (PEO) 277-283
and poly(ε-caprolactone)
(PCL) 284-289, etc., some of which were involved in the previous examples. The presence of the
NCCs significantly enhanced the mechanical strength of the matrices regardless of the
compositing approaches. Following are some advanced strategies for preparing biocomposites
containing NCCs and/or the corresponding promising properties for biomedical applications.
Habibi et al. 290 and Peltzer et al. 291 grafted PCL and PLA chains respectively onto the surfaces
of NCC via ring-opening polymerization, rendering NCCs highly hydrophobic, thus improving
Nanoscale Page 22 of 45

292 293, 294


their dispersion within the polymer matrices. Peresin et al. and Zhou et al. prepared
PVA/NCC, PEO/NCC and PLA/NCC composite fibres via an electrospinning process, among
View Article Online
DOI: 10.1039/C7NR04994C

which the fibrous PLA/NCC biocomposite presented cytocompatibility, biodegradability and


295
good mechanical property as scaffolds applied in bone tissue engineering. Rueda et al.
evaluated the cell response towards a PU/NCC composite formed by in situ polymerization and
confirmed the excellent biocompatibility of the composites, which was reflected from the good
cell adhesion and proliferation behavior of the L-929 fibroblasts on the PU/NCC surfaces.
Capadona et al. 296 developed a versatile strategy for preparing polymeric nanocomposite based
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


on NCC templates, in which a NCC template was firstly formed via a sol-gel process, followed
by immersing the template into a matrix polymer solution. Finally, the dried and compacted
nanocomposites were obtained. Using cis-polybutadiene and ethylene oxide–epichlorohydrin
copolymer as the model polymers, such strategy was proved applicable for constructing
polymeric nanocomposites involving nonpolar or polar polymers, significantly enhancing their
mechanical strength. By investigating the dispersion and/or aggregation behaviors of the
carbonate- and amine-modified NCCs under different pH conditions, Way et al. found that both
the carbonate and amine moieties imparted NCCs with pH-triggered aggregation due to the
formation of hydrogen bonds, where the carbonate-NCCs were dispersed in acidic medium and
formed hydrogels at higher pH. However, the pH-sensitive behavior was opposite for the
297
amine-NCCs . Based on the unique feature of pH-sensitivity, either of the modified NCCs
was incorporated into a PVA matrix; and the resulting nanocomposite film presented
mechanically adaptive pH-sensitivity.

4. Bacterial nanocellulose (BNC)


BNC, also known as microbial cellulose or MC, can be derived from various species
(Table 1) as extracellular secretion. Only Gluconacetobacter (Acetobacter) strains can produce
4, 298
BNC commercially . In molecular formula the BNC is identical to plant-originated
nanocellulose described above, while in dimensions the BNC is normally larger compared to
the latter one. Moreover, there is absence of hemicellulose and lignin within BNC. That is, the
nanocellulose from bacteria is higher crystalline (up to 84~89%) with less amorphous domains
299-301
than those from plant, generating larger bacterial nanocrystals . Presenting significantly
80
higher Young’s modulus (up to 150 GPa) and mechanical strength compared to nanofibers
Page 23 of 45 Nanoscale

46, 302-307
from plants, BNC has been widely used in various biomedical materials and other
308-312
engineering materials as reinforcing component. Owing to the high purityDOI:
of 10.1039/C7NR04994C
bacterial
View Article Online

cellulose, BNC is produced with no need of the cleavage of hemicellulose or lignin via
complicated mechanical and/or chemical treatments. Generally, BNC is biofabricated via a
cellulose-synthesizing complex from the bacterial cells in stationary or agitated cultures (Table
1). In both cases, the glucose units firstly assemble with the cellulose molecule growing inside
the bacterial cell, and elementary fibril extrudes out through pores on the cellular surface,
further assembling and crystallizing into microfibrils and then hierarchically structured ribbons
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

131, 313

Nanoscale Accepted Manuscript


. During stationary fermentation, these ribbons intersperse and intertwine into gelatinous
pellicles with a 3D network and significantly high water content, of which the thickness
increases with incubation time until the entrapped bacterial cells become inactive. The
stationary fermentation produces the BNC presenting high crystallinity and mechanical strength,
while the low yield has restricted its commercial applications. Accordingly, a variety of
approaches has been developed to improve the productivity of the stationary fermentation,
including the utilization of alternative carbon sources 314-317, in situ pH control 318, inhibition of
319 320, 321
side production , isolation of high yield stains . By contrast, aerated or agitated
fermentation has dramatically accelerated the BNC production process, by which most of the
commercial BNC has been produced. During agitated fermentation, the bacterial cells
sufficiently contact with the provided oxygen and therefore BNC with high yield is generated
in forms of small pellets or granules. Compared to BNC obtained from stationary fermentation,
199
those from agitated fermentation has lower crystallinity and mechanical strength . Several
aspects, including application, costing, capacity and process rate, etc., should be considered
comprehensively to select the most appropriate approach.
4.1. Biocompatibility of BNC
As bacterial secretion with nanofibril morphology, BNC is in a way similar to collagen, the
322-324
primary fibrous protein support in skins, bones and cartilages . In addition to the
similarity in size, both collagen and BNC contain a large amount of water like a hydrogel.
Sometimes, BNC is referred to a “collagen-like” component. Nevertheless, in biomedical
applications BNC possesses a great strength over collagen which is free of immune stimulation
325
, exhibiting excellent biocompatibility and tissue integration capacity. Along with much
higher mechanical strength, BNC have continuously inspired researchers attempting to apply it
Nanoscale Page 24 of 45

in various biomedical products due to its unique properties. There were several in vitro and in
vivo biocompatibility studies on BNC supporting them amenable for use in biomedical fields. A
View Article Online
DOI: 10.1039/C7NR04994C

genotoxicity evaluation of BNC produced via acid and ultrasonic treatments was carried out in
326
vitro . Both the comet cell assay and the Salmonella reversion assay demonstrated the BNC
had no genotoxicity. The proliferation assay showed a slight reduction of the fibroblasts
327
proliferation rate, while no cellular morphology alterations were observed. Jeong et al.
evaluated the toxicity of BNC in vitro on human endothelial cells and in vivo on a mouse model.
The results of the MTT tests and flow cytometric analysis demonstrated that neither
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


morphology alterations of endothelial cells nor apoptosis and necrosis were caused by BNC.
Besides, the blood collected from the BNC-treated mice did not show any biochemical changes.
328
A systematic in vivo biocompatibility study on BNC was conducted by Helenius et al. .
Implanting BNC into rats for 1, 4 and 12 weeks, no histologic inflammation and no foreign
body reaction were detected in the tissue surrounding the BNC specimen. Instead, the
fibroblasts with synthesized collagen were integrated into the BNC and new blood vessels were
formed.
4.2. BNC used as wound dressing and skin tissue repair materials
In the early 1980s, Johnson & Johnson Company was the first to devote efforts to
329, 330
commercialize BNC in large-scale for its biomedical application as wound dressings .
This is due to the distinguished mechanical and chemical properties of BNC generating strong
absorption capacity of liquids, great water retention ability, high wet strength and great
conformability, which promise BNC as an ideal wound healing material, particularly for the
burn wounds, chronic wounds like ulcers and surgical procedures (Figure 6I) 331. In the field of
wound dressings, the commercial application of BNC has been more extensive compared to
plant-derived nanocelluloses, based on which a number of commercial wound dressing
products have appeared, such as Biofill® , Gengiflex®, XCell®, Dermafill® 332, 333
, etc.. Wound
healing is a complex process, during which the proteolytic enzymes, reactive oxygen species
(ROS) and reactive nitrogen species (RNS) contained within the wound exudates have been
proven negatively effective to growth factors, leading to an imbalanced degradation/remodeling
processes. Therefore, attempts to reduce the proteolytic enzymes as well as the ROS and RNS
334, 335
are effective means to improve the wound healing. Wiegand et al. evaluated and
compared the capacity of antioxidation and reducing polymorphonuclear (PMN) elastase of
Page 25 of 45 Nanoscale

three different wound dressing materials, i.e., oxidized regenerated BNC (ORC), collagen and
ORC/collagen composite. Both the ORC and ORC/collagen composite were found to
View Article Online
DOI: 10.1039/C7NR04994C

completely deplete ROS and RNS, while the pure collagen showed a much lower ability in this
aspect. Regarding the in vitro PMN elastase binding, the pure ORC exhibited the greatest
capacity among the three tested products. All the results demonstrated BNC has an excellent
capacity for absorption of proteolytic enzymes and ROS/RNS, significantly improving the
336
wound healing process. Fu et al. developed a novel “multilayer fermentation method”,
which combined the stationary and agitated fermentation processes, for the preparation of BNC
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


wound dressings with a multilayer structure and controllable thickness by tuning the
337
fermentation period (Figure 6II). Qiu et al. incorporated a Chinese herbal medicinal
ingredient-vaccarin into the BNC system, of which the promoting blood coagulation effect
favored the BNC-vaccarin membrane as a potential material for wound healing. The
bacteria-biosynthesized poly(3-hydroxybutyrate) (PHB) has excellent biocompatibility,
biodegradability and blood clotting property, showing potential as high-performance
338
biomedical materials. Given this, Cai et al. prepared PHB-BNC and
339
poly(3-hydroxubutyrate-co-4-hydroxubutyrate) (P(3HB-co-4HB))-BNC composites and
evaluated their mechanical, biodegradable and biocompatible performances. The composites
presented higher mechanical strength and improved biodegradability compared to pure PHB
and P(3HB-co-4HB), respectively. Moreover, the cell proliferation assay using Chinese
Hamster Lung fibroblast cells as model cells demonstrated that the incorporation of BNC
rendered the materials much better biocompatible.
Nanoscale Page 26 of 45

View Article Online


DOI: 10.1039/C7NR04994C
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


Figure 6. BNC-based materials applied for skin tissue repair. (I) BNC dressings applied on
various skin wounds. (II) Skin repair effect of BNC dressing on the full-thickness skin lesion of
BALB/c mice. (III) a. SEM morphology of BNC-SSDs membrane (Scale bar-20 μm); b. Macroscopic
appearance of partial thickness wound excised on rat at 14th day; c. Macroscopic appearance of partial
thickness wound covered with BNC-SSDs membrane excised on rat at 14th day; d. Wound healing rate
of the wound. (IV) Cell morphology of fibroblast on (a) tissue culture polystyrene control, (b) BNC
membranes and (c) BNC-Chitosan membranes for 24 h incubation; d. Wound closure ratio of wound
healing (Ch-Chitosan). Reprinted with permission (I) from ref. 333 and ref. 304, (II) from ref. 336, (III)
from ref. 340 and (IV) from ref. 341.

In addition to the protein-degrading enzymes and ROS/RNS interfering with the wound
healing process, bacterial infection is another major concern which can seriously impact the
wound cicatrization. However, cellulose has no inherent bactericidal property. To address this
issue, a variety of antibacterial components has been attempted to be incorporated within the
cellulosic wound dressing materials. Among them, the AgNPs are most extensively studied due
to their broad-spectrum and high efficacy against bacteria. Studies have indicated the BNC, by
binding AgNPs, can greatly disperse and stabilize AgNPs in the bacterial culture medium, and
342-344
lead to a concentration-dependent antibacterial effect . Mostly, the AgNPs are formed on
the surface of BNC by in situ synthesis using AgNO3 as the silver source and NaBH4 as the
345, 346 347
reducing agent . Barud et al. used triethanolamine (TEA) and BNC itself as the
reducing agents for the synthesis of AgNPs and preparation of antibacterial BNC via hydrolytic
Page 27 of 45 Nanoscale

decomposition approach. Recently, a facile approach for fabricating BNC-AgNP composites


based on hydrothermal reaction without any introduced chemical reagents wasDOI:
developed,
View Article Online
10.1039/C7NR04994C
348-351
which could greatly control the particle size and size distribution . Using BNC as the
reducing and stabilizing agent for AgNPs commendably solved the problem of relying on toxic
reducing agents and organic solvents, creating a green route for AgNP synthesis. Zheng et al.
340, 352
developed a novel AgNP-BNC system, of which the AgNPs were prepared from silver
sulfadiazine (SSD) through ultrasonication and centrifugation. The suspended SSD particles
(SSD-s) with small particle size of 282 nm and precipitated ones (SSD-p) with large particle
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


size of 2150 nm were obtained and combined with BNC, respectively. Compared to the
BNC-SSD-s membrane, the BNC-SSD-p one presented a much larger inhibition zone against
bacteria, while the impregnated SSD particles were too large to completely permeate into the
BNC network. The in vivo analysis conducted on rat model showed the BNC-SSD-s membrane
dramatically promoted the epithelization and had a high healing rate on burn wounds (Figure
353
6III). A recent study by George et al. demonstrated the simultaneous incorporation of BNC
and AgNPs could significantly reduce the brittleness and improve the elongation of PVA films,
and the incorporated AgNPs might impart the PVA-BNC composites with additional
antimicrobial functionality.
Regarding to the antibacterial BNC including BNC wound dressings, a variety of
antimicrobial agents has been utilized for exerting BNC bactericidal in addition to the AgNPs.
A series of BNC-montmorillonite (MMT) composite films containing different inorganic metal
ions were prepared by anchoring modified MMTs onto the BNC surfaces 354. The incorporation
of MMT imparted high mechanical strength and certain disinfectant activity to BNC, while the
addition of Cu, compared to Na and Ca, exceptionally enhanced the antibacterial property of
the BNC-MMT composite film. Chitosan and chitin are versatile biomedical materials with
excellent biocompatibility and antibacterial activity. Considering this, several efforts have been
341, 355, 356
made to develop BNC-chitosan/chitin composites (Figure 6IV) . In the study on the
BNC combining with chitin nanocrystals (ChNCs) obtained through three different approaches,
i.e., TEMPO mediated oxidation, partial deacetylation and acid hydrolysis of α-chitin, Butchosa
357
et al. found that the composites post-modified by the partially deacetylated ChNCs
presented the greatest bactericidal/bacteriostatic performance and excellent mechanical
358 359, 360
property. Additionally, ZnO and TiO2 have been applied as the antibacterial
Nanoscale Page 28 of 45

components in the BNC wound dressings. In order to use BNC membranes as drug delivery
361, 362
system for the treatment of skin diseases, Almeida et al. evaluated the skin irritation
View Article Online
DOI: 10.1039/C7NR04994C

potential of BNC on human body. The transepidermal water loss tests and clinical score results
indicated that the BNC membrane possessed good skin tolerance. The incorporation of glycerin
imparted a moisturizing effect on the membranes. All the findings supported the BNC
membrane as a drug delivery system for topical and/or transdermal treatments.
4.3. BNC for cartilage replacement
Owing to the aformentioned unique properties of BNC, the mimicking-soft tissue
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


nanomaterial has been highly recommended for tissue engineering, especially as scaffolds to
support cellular proliferation and growth within the tissue. In the field of the cartilage repair,
several researchers have attempted to use BNC as the scaffold for temporarily substituting
cartilages and stimulating the formation of new tissues. A study demonstrated quine
chondrocytes robustly proliferated on the BNC materials and maintained the cellular phenotype
363 364
. Svensson et al. compared unmodified and sulfation/phosphorylation modified BNC
materials in terms of mechanical property and promotion of cellular growth for its potential
application as cartilage scaffold. It was found that the native BNC, with excellent mechanical
strength, significantly enhanced the chondrocyte growth, while the chemically modified BNCs
did not show any promotion for chondrocyte proliferation due to their porosity interfering with
the chondrocyte viability. To address the limited pore size of BNC materials for cartilage repair
365
application, Andersson et al. fabricated a BNC scaffold with pore size of 150 ~ 300 μm
using wax particles as the pore former, of which the porous construct allowed chondrocytes to
proliferate onto both the inside and outside pore surfaces of the BNC scaffold. BNC was also
evaluated as a potential meniscus material for cartilage repair, presenting similar mechanical
strength as native big meniscus and much higher strength compared to collagen material
(Figure 7A) 366. Meanwhile, the BNC material could be facilely fabricated into meniscus shape
and promote the cellular migration. To form micro-scaled porous structure, a novel approach
using pin templates comprised of optical fibers or polystyrene was developed by Rambo et al.
367
. By using different pin templates, pores with different diameters were structured (i.e., Ø =
60 μm with optical fibers and Ø = 300 μm with polystyrene beads), which could improve the
oxygenation, exudate absorption and delay on wound contracture of cartilage repairing tissues.
Page 29 of 45 Nanoscale

View Article Online


DOI: 10.1039/C7NR04994C
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


Figure 7. A variety of BNC-based bio-engineering materials. (A) BNC meniscus implant. (B)
BASYC® tubes with different sizes. (C) BNC sponge as tissue engineering scaffold and
phase-contrast microscopic graph of fibrous cells growing on the scaffold after in vitro culture for 5
days (S-scaffold, C-cells). (D) SEM image of porous BNC matrix and immunofluorescence analysis of
human urine-derived stem cells seeded BNC scaffold stained by α-smooth muscle actin (Scale bar-200
μm). Reprinted with permission (A) from ref. 366, (B) from ref. 368, (C) from ref. 369 and (D) from ref.
370
.

4.4. BNC-based artificial blood vessel


Vascular grafts provide an effective approach for the treatment of arterial diseases.
Previous studies have shown that the BNC was a potential vascular graft material for replacing
371, 372 373, 374
diseased blood vessels . In this vein, Bäckdahl et al. compared the mechanical
properties and cellular proliferation of BNC with porcine carotid arteries and
expanded-polytetraflourethylene (ePTFE). Apart from much higher mechanical strength
compared to the latter two, the BNC material exhibited a similar stress-strain response with the
carotid arteries due to their semblable architectures. Moreover, the human smooth muscle cells
attached and proliferated onto the BNC surfaces with a 40 μm of growth after 14-day culture.
375, 376
Millon et al. prepared a PVA-BNC nanocomposite via mixing the PVA and BNC
solutions followed by heating-refrigerating circulation. The mechanical properties of the
resulting composites, including mechanical strength, stress-stain slope and relaxation
performance, were well matched with both aorta and heart valve tissue. Due to their great
Nanoscale Page 30 of 45

moldability, BNC could be facilely formed into tubular shapes via various molding
368, 377, 378
technologies. Klemm et al. attempted to culture A. xylinum between two cylindrical
View Article Online
DOI: 10.1039/C7NR04994C

tubes with different diameters, constructing a tubular BNC material with a product name of
BASYC® (Figure 7B), which showed a high potential as a vascular graft in microsurgery. In a
study of comparing the angiogenesis properties of BNC, polyglutamic acid (PGA) and ePTFE,
379
the BNC presented lower activity than PGA material . To improve the attachment of human
microvascular endothelial cells to BNC, chimeric proteins including a cellulose-binding
module and adhesion peptides were used to improve the property of BNC as artificial blood
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

vessels 380.

Nanoscale Accepted Manuscript


4.5. BNC used for bone regeneration
Given the fact that in tissue engineering for bone healing bone cells need a biocompatible
environment with a nano-scaled porous and 3D network structure for cellular proliferation,
migration and differentiation, many nano-biomaterials have been developed and applied as
381 382
scaffolds for bone tissue engineering . Biomimetic ceramics and nanofibers electrospun
from synthetic polymers 383, 384 have been widely used for bone regeneration. However, the low
mechanical strength and stability are their greatest drawbacks. The appearance and application
of BNC elegantly solve the problem. First of all, BNC has shown excellent performance in
supporting bone cells growth and maintaining their morphology. The in vivo study conducted
by Mendes et al. 385 demonstrated the BNC repair membrane could prevent the fibroblasts from
intruding into the bone defects owing to its nano-scaled pores, provoking no or low
inflammatory responses. The too small pore size, on the other side, limited the infiltration of
the bone cells. To address, paraffin wax has been incorporated into the BNC system for
386
structuring micro-scaled pores . Results showed the porous BNC with a pore size of 300 ~
500 μm had a high mechanical strength and improved the ingrowth of osteoblast forming
mineralized tissues. Another study attempting to fabricate large-pore-sized BNC materials was
to prepare BNC sponge via emulsion freeze-drying approach, which generated a hierarchical
pore structure composed of 20 ~ 1000 μm sized macropores and 4 nm sized nano-pores (Figure
369
7C) . Since the bone tissue is mainly constituted by collagen and hydroxyapatite (HAP), of
which the HAP has intrinsic high bioactivity and osteoconductivity, BNC could be applied as a
collagen substitute for the fabrication of BNC-HAP composites with enhanced mechanical
387-390 44, 391
properties . For example, Wan et al. prepared BNC-HAP composites by soaking
Page 31 of 45 Nanoscale

phosphorylated and unphosphorylated BNC respectively into simulated body fluid. SEM
analysis indicated that the former one induced the uniform crystallization of HAP on its surface
View Article Online
DOI: 10.1039/C7NR04994C

while the latter one showed little activity, presumably due to the calcium phosphate complexes
392 393
acting as the nuclei of HAP. Hutchens et al. and Grande et al. used native BNC and
carboxymethylcellulose containing BNC to develop BNC-HAP materials containing
calcium-deficient HAP, respectively. Both the composites were proven to be ideal scaffolds for
the biomineralizaiton of bones. With consideration of the porous structure and HAP attachment,
Tsioptsias et al. 394 fabricated BNC-HAP composites using poly(methyl methacrylate) particles
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


as the porogen, where the HAP was incorporated via dispersing within the ionic liquid solution
of BNC followed by solvent evaporation.
4.6. BNC in other biomedical applications
The BNC has also been applied for oral tissue regeneration in the term of BNC-alginate
395
composite sponge . Prepared by a freeze drying technique, the sponge showed the excellent
mechanical property, great water uptake capacity and improved proliferation of human
keratinocytes and gingival fibroblasts, and was promising as artificial mucosal flaps. Bodin et
370
al. developed BNC scaffolds for use as conduits in urinary diversion. Human urine-derived
stem cells were seeded onto the BNC scaffolds and differentiated into smooth muscle and
urothelial cells with multilayers after two-week culture. Afterwards, the cell infiltrated scaffolds
were implanted into athymic mice subsequently, which showed the expression of urothelial and
smooth muscle cellular markers (Figure 7D).
5. Conclusion and future prospects
The development of science and technology along with the resource shortage and
environmental deterioration is seeking for more environmentally friendly and renewable raw
materials as alternatives for petroleum-based materials. Cellulose provides a sustainable
resource satisfying the requirements and has continuously inspired researchers to exploit
cellulose-based materials with novel functionalities. The nanocellulose, including NFC, NCC
and BNC, combines the intrinsic features of cellulose and unique properties of nano-scaled
materials deriving a vast variety of promising performances. Triggered by the promising
performances, especially the excellent mechanical properties, biocompatibility and
biodegradability, the nanocellulose has been extensively and profoundly investigated in the
applications related to antimicrobial materials, skin replacement product, drug delivery, tissue
Nanoscale Page 32 of 45

engineering, etc., over the last decade. The present review briefly introduced the structure,
properties, treatment technologies of three types of nanocellulose along with theDOI:
focuses on
View Article Online
10.1039/C7NR04994C

their cytotoxicity, biocompatibility and frontier study and applications in biomedical field.
Because of the different resources, treatment processes and structures of NFC, NCC and BNC,
these nanocellulose have their unique properties in addition to their common functionalities and
thus perform particularly well in some way. Exceptionally abundant raw material resources and
well-established chemical processes contribute to achieving the industrial production of NFC.
The tangled network and larger aspect ratio impart NFC higher strength and modulus compared
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


to NCC. As reinforcing agents in composites, NFC leads to excellent mechanical properties of
the composite with relatively low cost. Physiologically inactive feature and high binding
property due to the large amount of hydrogen bonds favor NCC with great performance as drug
delivery agent and tablet binder. In addition, NCC presents a unique optical property owing to
its crystalline nature which provides NCC significant opportunities for specific applications in
optical materials. The BNC is purely biosynthesized by bacteria which is completely different
from NFC or NCC, eliminating the removal process of hemicellulose, lignin and other matrix
components. The high purity and crystallinity as well as excellent biocompatibility offer BNC
more favorable criteria in applications of bioengineering and biomedical fields. In particular,
BNC has been widely used in high-quality skin tissue repair and wound dressing materials due
to its transparency, high water retention, great wet tensile strength and resembling properties of
human collagen.
With the intensive study and application development of conventional nanocellulose, novel
polysaccharide-based nanomaterials are emerging. Very recently, a new kind of cellulose
nanoparticles consisting of a nanocrystalline body and two-end attached amorphous chains has
396
been developed by van de Ven . The new nanocellulose, named hairy nanocrystalline
cellulose (HNCC), could be facilely obtained by periodate oxidation of cellulose. Owing to the
highly reactive aldehyde groups on the amorphous regions, further functionalization of the
periodate-treated cellulose could result in cationic NCC (CNCC) and/or sterically stabilized
NCC (ENCC). The unique colloidal structure endows HNCC with some superior features over
conventional NCC, such as high functionality and reactivity, tunable charge, and colloidal
397
stability. HNCC are showing excellent promise in many applications , including transparent
film, decontamination flocculant, rheology modifier, and biomimetic mineralization, etc.
Page 33 of 45 Nanoscale

Furthermore, the favorable physiochemical properties might qualify HNCC for biomedical
applications with the assurance of biological safety, e.g., drug/gene carrier. View Article Online
DOI: 10.1039/C7NR04994C

Regarding to the cytotoxicity of nanocellulose, much effort has been devoted to proving
the biocompatibility of the pure NFC, NCC and BNC as described in the previous sections.
Various surface and/or bulk modifications of the nanocellulose bring up new opportunities for
developing unique functional materials. Meanwhile, the modification leads to some alterations
of physico-chemical properties of materials, especially at the nanoscale. On this view, the
influence of the incorporation of foreign molecules on the cytotoxicity and/or biocompatibility
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Nanoscale Accepted Manuscript


of nanocellulose should be elaborated specifically.
Although the three categories of nanocellulose have been being extensively studied and led
to some commercial products in biomedical field, developing more facile and efficient
technologies for nanocellulose producing should be taken into consideration as well. “Green”
resources need “green” technologies to exploit and utilize.

Acknowledgement
The authors gratefully acknowledge National Natural Science Foundation of China
(NNSFC) under the Grant No. 51603174 & No. 51379077 and the Natural Sciences and
Engineering Research Council of Canada (NSERC) for the financial support.

References
1. Y. Habibi, L. A. Lucia and O. J. Rojas, Chemical Reviews, 2010, 110, 3479-3500.
2. B. L. Holt, S. D. Stoyanov, E. Pelan and V. N. Paunov, Journal of Materials Chemistry, 2010, 20,
10058-10070.
3. A. Dufresne, Materials Today, 2013, 16, 220-227.
4. Y. Huang, C. Zhu, J. Yang, Y. Nie, C. Chen and D. Sun, Cellulose, 2014, 21, 1-30.
5. M. M. de Souza Lima and R. Borsali, Macromolecular Rapid Communications, 2004, 25, 771-787.
6. P. A. Richmond, Biosynthesis & Biodegradation of Cellulose, 1991, 5-23.
7. C. Chang and L. Zhang, Carbohydrate Polymers, 2011, 84, 40-53.
8. C. Miao and W. Y. Hamad, Cellulose, 2013, 20, 2221-2262.
9. D. Klemm, B. Heublein, H. P. Fink and A. Bohn, ChemInform, 2005, 44, 3358-3393.
10. K. Missoum, Doctoral dissertation, Grenoble, 2012.
11. J. Huang and Y. Gu, Current Opinion in Colloid & Interface Science, 2011, 16, 470-481.
12. G. Siqueira, J. Bras and A. Dufresne, Polymers, 2010, 2, 728.
13. D. Klemm, F. Kramer, S. Moritz, T. Lindström, M. Ankerfors, D. Gray and A. Dorris, Angewandte Chemie,
2011, 50, 5438-5466.
14. S. Boufi, in Biomass and Bioenergy, Springer, 2014, pp. 267-305.
15. S. Kalia, S. Boufi, A. Celli and S. Kango, Colloid and Polymer Science, 2014, 292, 5-31.
16. X. Xu, F. Liu, L. Jiang, J. Y. Zhu, D. Haagenson and D. P. Wiesenborn, Acs Applied Materials & Interfaces,
Nanoscale Page 34 of 45

2013, 5, 2999-3009.
17. H. Rosilo, J. R. Mckee, E. Kontturi, T. Koho, V. P. Hytönen, O. Ikkala and M. A. Kostiainen, Nanoscale, 2014,
View Article Online
6, 11871-11881. DOI: 10.1039/C7NR04994C

18. Q. Zhou, H. Brumer and T. T. Teeri, Macromolecules, 2009, 42, 5430-5432.


19. P. Tingaut, T. Zimmermann and G. Sèbe, Journal of Materials Chemistry, 2012, 22, 20105-20111.
20. L. Petersson, I. Kvien and K. Oksman, Composites Science and Technology, 2007, 67, 2535-2544.
21. Y. Habibi, A.-L. Goffin, N. Schiltz, E. Duquesne, P. Dubois and A. Dufresne, Journal of Materials Chemistry,
2008, 18, 5002-5010.
22. K. Oksman, A. Mathew, D. Bondeson and I. Kvien, Composites science and technology, 2006, 66,
2776-2784.
23. Paul Podsiadlo, L. Sui, Yaseen Elkasabi, Peter Burgardt, Jaebeom Lee, Ashwini Miryala, Winardi
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Kusumaatmaja, M. R. C. ‖, Max Shtein and John Kieffer, Langmuir the Acs Journal of Surfaces & Colloids,
2007, 23, 7901-7906.

Nanoscale Accepted Manuscript


24. B. S. Shim, P. Podsiadlo, D. G. Lilly, A. Agarwal, J. Lee, Z. Tang, S. Ho, P. Ingle, D. Paterson, W. Lu and N. A.
Kotov, Nano Letters, 2007, 7, 3266-3273.
25. E. Dujardin, M. Blaseby and S. Mann, Journal of Materials Chemistry, 2003, 13, 696-699.
26. M. Cheng and L. Zhong, Nordic Pulp & Paper Research Journal, 2014, 29, 156-166.
27. N. Lavoine, I. Desloges, A. Dufresne and J. Bras, Carbohydrate Polymers, 2012, 90, 735–764.
28. W. Zhang, X. Zhang, C. Lu, Y. Wang and Y. Deng, Journal of Physical Chemistry C, 2012, 116, 9227-9234.
29. H. P. S. A. Khalil, A. H. Bhat and A. F. I. Yusra, Carbohydrate Polymers, 2011, 87, 963-979.
30. Y. Li, H. Zhu, F. Shen, J. Wan, X. Han, J. Dai, H. Dai and L. Hu, Advanced Functional Materials, 2014, 24,
7366-7372.
31. S. Boufi and A. Gandini, RSC Advances, 2015, 5, 3141-3151.
32. S. Ahola, J. Salmi, L.-S. Johansson, J. Laine and M. Ö sterberg, Biomacromolecules, 2008, 9, 1273-1282.
33. Y. H. Jung, T.-H. Chang, H. Zhang, C. Yao, Q. Zheng, V. W. Yang, H. Mi, M. Kim, S. J. Cho, D.-W. Park, H.
Jiang, J. Lee, Y. Qiu, W. Zhou, Z. Cai, S. Gong and Z. Ma, Nature Communications, 2015, 6, 7170.
34. A. Shimotoyodome, J. Suzuki, Y. Kumamoto, T. Hase and A. Isogai, Biomacromolecules, 2011, 12,
3812-3818.
35. E. D. M. Teixeira, A. C. Corrêa and A. Manzoli, Cellulose, 2010, 17, 595-606.
36. J. Wang, Q. Cheng, L. Lin and L. Jiang, ACS nano, 2014, 8, 2739-2745.
37. M. M. Malinen, L. K. Kanninen, A. Corlu, H. M. Isoniemi, Y.-R. Lou, M. L. Yliperttula and A. O. Urtti,
Biomaterials, 2014, 35, 5110-5121.
38. T. Hakkarainen, R. Koivuniemi, M. Kosonen, C. Escobedo-Lucea, A. Sanz-Garcia, J. Vuola, J. Valtonen, P.
Tammela, A. Mäkitie, K. Luukko, M. Yliperttula and H. Kavola, Journal of Controlled Release, 2016, 244,
Part B, 292-301.
39. Y. Liu, H. Wang, G. Yu, Q. Yu, B. Li and X. Mu, Carbohydrate Polymers, 2014, 110, 415-422.
40. D. Liu, X. Chen, Y. Yue, M. Chen and Q. Wu, Carbohydrate Polymers, 2011, 84, 316-322.
41. Y. Boluk, R. Lahiji, L. Zhao and M. T. Mcdermott, Colloids & Surfaces A Physicochemical & Engineering
Aspects, 2011, 377, 297-303.
42. T. Zimmermann, N. Bordeanu and E. Strub, Carbohydrate Polymers, 2010, 79, 1086-1093.
43. A. Olszewska, P. Eronen, L.-S. Johansson, J.-M. Malho, M. Ankerfors, T. Lindström, J. Ruokolainen, J. Laine
and M. Ö sterberg, Cellulose, 2011, 18, 1213.
44. Y. Wan, L. Hong, S. Jia, Y. Huang, Y. Zhu, Y. Wang and H. Jiang, Composites Science and Technology, 2006,
66, 1825-1832.
45. C. Boisset, C. Fraschini, M. Schülein, B. Henrissat and H. Chanzy, Applied and Environmental Microbiology,
2000, 66, 1444-1452.
46. H. Yano, J. Sugiyama, A. N. Nakagaito, M. Nogi, T. Matsuura, M. Hikita and K. Handa, Advanced Materials,
Page 35 of 45 Nanoscale

2005, 17, 153-155.


47. R. J. Moon, A. Martini, J. Nairn, J. Simonsen and J. Youngblood, Chemical Society Reviews, 2011, 40,
View Article Online
3941-3994. DOI: 10.1039/C7NR04994C

48. S. Eyley and W. Thielemans, Frontiers of Chemical Science and Engineering, 2007, 6, 7764-7779.
49. L. Brinchi, F. Cotana, E. Fortunati and J. M. Kenny, Carbohydrate Polymers, 2013, 94, 154-169.
50. K. Oksman, Y. Aitomäki, A. P. Mathew, G. Siqueira, Q. Zhou, S. Butylina, S. Tanpichai, X. Zhou and S.
Hooshmand, Composites Part A: Applied Science and Manufacturing, 2016, 83, 2-18.
51. N. Shah, M. Ul-Islam, W. A. Khattak and J. K. Park, Carbohydrate Polymers, 2013, 98, 1585-1598.
52. N. Petersen and P. Gatenholm, Applied Microbiology and Biotechnology, 2011, 91, 1277.
53. D. Klemm, D. Schumann, F. Kramer, N. Heßler, M. Hornung, H.-P. Schmauder and S. Marsch, in
Polysaccharides II, ed. D. Klemm, Springer Berlin Heidelberg, Berlin, Heidelberg, 2006, DOI:
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

10.1007/12_097, pp. 49-96.


54. J. Xie and J. Li, Journal of Bioresources and Bioproducts, 2017, 2, 1-3.

Nanoscale Accepted Manuscript


55. M.-J. Le Guen, R. H. Newman, A. Fernyhough, S. J. Hill and M. P. Staiger, in Natural Fibres: Advances in
Science and Technology Towards Industrial Applications: From Science to Market, eds. R. Fangueiro and S.
Rana, Springer Netherlands, Dordrecht, 2016, 35-47.
56. B. Wei, Q. Li, F. Jin, H. Li and C. Wang, Energy & Fuels, 2016, 30, 2882-2891.
57. K. Missoum, M. N. Belgacem and J. Bras, Materials, 2013, 6, 1745-1766.
58. I. Siró and D. Plackett, Cellulose, 2010, 17, 459-494.
59. S. Mondal, Carbohydrate Polymers, 2017, 163, 301-316.
60. M. A. Hubbe, O. J. Rojas, L. A. Lucia and M. Sain, BioResources, 2008, 3, 929-980.
61. I. Díez, P. Eronen, M. Ö sterberg, M. B. Linder, O. Ikkala and R. H. Ras, Macromolecular bioscience, 2011,
11, 1185-1191.
62. T. Saito, Y. Nishiyama, J. L. Putaux, M. Vignon and A. Isogai, Biomacromolecules, 2006, 7, 1687-1691.
63. D. Dai, M. Fan and P. Collins, Industrial Crops and Products, 2013, 44, 192-199.
64. D. Dai and M. Fan, RSC Advances, 2013, 3, 4659-4665.
65. A. Dufresne, D. Dupeyre and M. R. Vignon, Journal of Applied Polymer Science, 2000, 76, 2080-2092.
66. A. Alemdar and M. Sain, Bioresource Technology, 2008, 99, 1664-1671.
67. J. Leitner, B. Hinterstoisser, M. Wastyn, J. Keckes and W. Gindl, Cellulose, 2007, 14, 419-425.
68. Y. Habibi and M. R. Vignon, Cellulose, 2008, 15, 177-185.
69. J. I. Morán, V. A. Alvarez, V. P. Cyras and A. Vázquez, Cellulose, 2008, 15, 149-159.
70. J. Biagiotti, D. Puglia, L. Torre, J. M. Kenny, A. Arbelaiz, G. Cantero, C. Marieta, R. Llano-Ponte and I.
Mondragon, Polymer Composites, 2004, 25, 470-479.
71. B. Deepanjan, G. Louist and W. Williamt, Carbohydrate Polymers, 2008, 73, 371-377.
72. A. Bendahou, H. Kaddami and A. Dufresne, European Polymer Journal, 2010, 46, 609-620.
73. S. Alila, I. Besbes, M. R. Vilar, P. Mutjé and S. Boufi, Industrial Crops & Products, 2013, 41, 250–259.
74. D. M. Bruce, R. N. Hobson, J. W. Farrent and D. G. Hepworth, Composites Part A Applied Science &
Manufacturing, 2005, 36, 1486-1493.
75. B. Wang and M. Sain, Composites Science & Technology, 2007, 67, 2521-2527.
76. R. Zuluaga, J. L. Putaux, A. Restrepo, I. Mondragon and P. Gañán, Cellulose, 2007, 14, 585-592.
77. A. B. Mabrouk, H. Kaddami, S. Boufi, F. Erchiqui and A. Dufresne, Cellulose, 2012, 19, 843-853.
78. I. Besbes, M. R. Vilar and S. Boufi, Carbohydrate Polymers, 2011, 86, 1198-1206.
79. S. Kimura and T. Itoh, Cellulose, 2004, 11, 377-383.
80. S. Iwamoto, W. Kai, A. Isogai and T. Iwata, Biomacromolecules, 2009, 10, 2571-2576.
81. K. Hua, D. O. Carlsson, E. Ålander, T. Lindström, M. Strømme, A. Mihranyan and N. Ferraz, Rsc Advances,
2014, 4, 2892-2903.
82. D. L. Bras, M. Stromme and A. Mihranyan, Journal of Physical Chemistry B, 2015, 119, 5911-5917.
Nanoscale Page 36 of 45

83. R. H. Müller, C. Jacobs and O. Kayser, Advanced Drug Delivery Reviews, 2001, 47, 3-19.
84. W. Stelte and A. R. Sanadi, Industrial & Engineering Chemistry Research, 2009, 48, 11211-11219.
View Article Online
85. J. I. Park, A. Saffari, S. Kumar, A. Günther and E. Kumacheva, Annual Review of Materials Research, 2010,
DOI: 10.1039/C7NR04994C

40, 415-443.
86. C. Aulin, J. Netrval, L. Wågberg and T. Lindström, Soft Matter, 2010, 6, 3298-3305.
87. W. Chen, H. Yu, Y. Liu, Y. Hai, M. Zhang and P. Chen, Cellulose, 2011, 18, 433-442.
88. G. H. Tonoli, E. M. Teixeira, A. C. Corrêa, J. M. Marconcini, L. A. Caixeta, M. A. Pereira-Da-Silva and L. H.
Mattoso, Carbohydrate Polymers, 2012, 89, 80-88.
89. S. P. Mishra, A. S. Manent, B. Chabot and C. Daneault, Bioresources, 2012, 7, 422-436.
90. K. Abe, A. Shinichiro Iwamoto and H. Yano, Biomacromolecules, 2007, 8, 3276-3278.
91. S. Iwamoto, A. N. Nakagaito and H. Yano, Applied Physics A, 2007, 89, 461-466.
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

92. M. Pääkkö, J. Vapaavuori, R. Silvennoinen, H. Kosonen, M. Ankerfors, T. Lindström, L. A. Berglund and O.


Ikkala, Soft Matter, 2008, 4, 2492-2499.

Nanoscale Accepted Manuscript


93. M. Andresen, L. S. Johansson, B. S. Tanem and P. Stenius, Cellulose, 2006, 13, 665-677.
94. S. Iwamoto, A. N. Nakagaito, H. Yano and M. Nogi, Applied Physics A, 2005, 81, 1109-1112.
95. Q. Cheng, S. Wang, T. G. Rials and S. H. Lee, Cellulose, 2007, 14, 593-602.
96. C. Cara, E. Ruiz, M. Ballesteros, P. Manzanares, M. a. J. Negro and E. Castro, Fuel, 2008, 87, 692-700.
97. B. M. Cherian, A. L. Leão, S. F. D. Souza, S. Thomas, L. A. Pothan and M. Kottaisamy, Carbohydrate
Polymers, 2010, 81, 720-725.
98. B. Deepa, E. Abraham, B. M. Cherian, A. Bismarck, J. J. Blaker, L. A. Pothan, A. L. Leao, S. F. de Souza and
M. Kottaisamy, Bioresource Technology, 2011, 102, 1988-1997.
99. A. Isogai, T. Saito and H. Fukuzumi, Nanoscale, 2011, 3, 71-85.
100. T. Saito, M. Hirota, N. Tamura, S. Kimura, H. Fukuzumi, L. Heux and A. Isogai, Biomacromolecules, 2009,
10, 1992-1996.
101. I. Besbes, S. Alila and S. Boufi, Carbohydrate Polymers, 2011, 84, 975-983.
102. M. Henriksson, G. Henriksson, L. A. Berglund and T. Lindström, European Polymer Journal, 2007, 43,
3434-3441.
103. Janardhnan, S., Sain and M., Bioresources, 2006, 1, 176-188.
104. J. Y. Zhu, R. Sabo and X. L. Luo, Green Chemistry, 2011, 13, 1339-1344.
105. A. Chaker, S. Alila, P. Mutjé, M. R. Vilar and S. Boufi, Cellulose, 2013, 20, 2863-2875.
106. S. Iwamoto, K. Abe and H. Yano, Biomacromolecules, 2008, 9, 1022-1026.
107. S. Elazzouzi-Hafraoui, Y. Nishiyama, J. L. Putaux, L. Heux, F. Dubreuil and C. Rochas, Biomacromolecules,
2008, 9, 57-65.
108. S. Montanari, M. Roumani, A. Laurent Heux and M. R. Vignon, Macromolecules, 2005, 38, 1665-1671.
109. J. Araki, M. Wada, A. Shigenori Kuga and T. Okano, Langmuir, 2000, 16, 2413-2415.
110. W. Helbert, J. Y. Cavaillé and A. Dufresne, Polymer Composites, 1996, 17, 604-611.
111. M. A. S. A. Samir, F. Alloin, J. Y. Sanchez and A. Dufresne, Macromolecules, 2004, 37, 4839-4844.
112. X. Cao, H. Dong and C. M. Li, Biomacromolecules, 2007, 8, 899-904.
113. D. Dasong, Doctoral dissertation, Brunel University London, 2015.
114. Y. Habibi and A. Dufresne, Biomacromolecules, 2008, 9, 1974-1980.
115. A. J. D. Menezes, G. Siqueira, A. A. S. Curvelo and A. Dufresne, Polymer, 2009, 50, 4552-4563.
116. N. L. G. D. Rodriguez, W. Thielemans and A. Dufresne, Cellulose, 2006, 13, 261-270.
117. G. Siqueira, J. Bras and A. Dufresne, Biomacromolecules, 2009, 10, 425-432.
118. L. Yu, J. Lin, F. Tian, X. Li, F. Bian and J. Wang, Journal of Materials Chemistry A, 2014, 2, 6402-6411.
119. P. Terech, L. Chazeau, ‡ and and J. Y. Cavaille, Macromolecules, 1999, 32, 1872-1875.
120. M. N. A. And and A. Dufresne, Macromolecules, 2000, 33, 8344-8353.
121. H. Yuan, Y. Nishiyama, M. Wada and S. Kuga, Biomacromolecules, 2006, 7, 696-700.
Page 37 of 45 Nanoscale

122. O. V. D. Berg, M. Schroeter, J. R. Capadona and C. Weder, Journal of Materials Chemistry, 2007, 17,
2746-2753.
View Article Online
123. S. Beck-Candanedo, M. Roman and D. G. Gray, Biomacromolecules, 2005, 6, 1048-1054. DOI: 10.1039/C7NR04994C

124. D. Bondeson, A. Mathew and K. Oksman, Cellulose, 2006, 13, 171-180.


125. W. Bai, J. Holbery and K. Li, Cellulose, 2009, 16, 455-465.
126. Shinsuke Ifuku, †, Masaya Nogi, Kentaro Abe, Keishin Handa, A. Fumiaki Nakatsubo and H. Yano†,
Biomacromolecules, 2007, 8, 1973-1978.
127. S. Yano, H. Maeda, M. Nakajima, T. Hagiwara and T. Sawaguchi, Cellulose, 2008, 15, 111-120.
128. R. Kose, I. Mitani, W. Kasai and T. Kondo, Biomacromolecules, 2011, 12, 716-720.
129. J. L. Potter and R. A. Weisman, Developmental Biology, 1972, 28, 472-479.
130. R. M. Brown, Journal of Polymer Science Part A Polymer Chemistry, 2004, 42, 487-495.
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

131. R. Jonas and L. F. Farah, Polymer Degradation and Stability, 1998, 59, 101-106.
132. J. L. Morgan, J. Strumillo and J. Zimmer, Nature, 2012, 493, 181-186.

Nanoscale Accepted Manuscript


133. S. Tanskul, K. Amornthatree and N. Jaturonlak, Carbohydrate Polymers, 2013, 92, 421-428.
134. A. Cavalcante, L. Carvalho and M. Carneiro-da-Cunha, Biochemical engineering journal, 2006, 29,
258-261.
135. S. C. Falcão, A. R. d. B. Coelho and J. Evêncio Neto, Acta Cirurgica Brasileira, 2008, 23, 184-191.
136. K.-C. Cheng, J. M. Catchmark and A. Demirci, Journal of Biological Engineering, 2009, 3, 12.
137. J. Y. Jung, J. K. Park and H. N. Chang, Enzyme and Microbial Technology, 2005, 37, 347-354.
138. A. Krystynowicz, W. Czaja, A. Wiktorowska-Jezierska, M. Gonçalves-Miśkiewicz, M. Turkiewicz and S.
Bielecki, Journal of Industrial Microbiology and Biotechnology, 2002, 29, 189-195.
139. S. Bae and M. Shoda, Biotechnol Prog, 2004, 20, 1366-1371.
140. C. Huang, H.-J. Guo, L. Xiong, B. Wang, S.-L. Shi, X.-F. Chen, X.-Q. Lin, C. Wang, J. Luo and X.-D. Chen,
Carbohydrate Polymers, 2016, 136, 198-202.
141. C. Campano, A. Balea, A. Blanco and C. Negro, Cellulose, 2016, 23, 57-91.
142. H. C. Huang, L. C. Chen, S. B. Lin, C. P. Hsu and H. H. Chen, Bioresource Technology, 2010, 101,
6084-6091.
143. A. Bodin, H. Bäckdahl, H. Fink, L. Gustafsson, B. Risberg and P. Gatenholm, Biotechnology &
Bioengineering, 2007, 97, 425.
144. N. Andre, X. Tian, M. D. Lutz and L. Ning, Science, 2006, 311, 622-627.
145. A. Dowling, R. Clift, N. Grobert, D. Hutton, R. Oliver, O. O’neill, J. Pethica, N. Pidgeon, J. Porritt and J.
Ryan, London: The Royal Society & The Royal Academy of Engineering Report, 2004, 61, e64.
146. J. Carrero-Sanchez, A. Elias, R. Mancilla, G. Arrellin, H. Terrones, J. Laclette and M. Terrones, Nano Letters,
2006, 6, 1609-1616.
147. G. Oberdörster, E. Oberdörster and J. Oberdörster, Environmental Health perspectives, 2005, 113, 823.
148. C. Hohenthal, M. Ovaskainen, D. Bussini, P. Sadocco, T. Pajula, H. Lehtinen, J. Kautto and K. Salmenkivi,
VTT Technical Research Centre of Finland, 2012.
149. V. Väänänen, E. Rydman, M. Ilves, K. Hannukainen, H. Norppa, A. von Wright, U. Honkalampi, I. Tsitko
and J. Rouhiainen, Scale-up Nanoparticles in Modern Papermaking (SUNPAP) WP, 2012.
150. J. Rouhianen, I. Tsitko, M. Vippola and J. Koivisto, Scale-up Nanoparticles in Modern
Papermaking-SUNPAP FP7, 2010.
151. M. Pitkänen, H. Kangas and J. Vartiainen, Handbook of Green Materials. World Scientific, 2014, 181-205.
152. J. Vartiainen, T. Pöhler, K. Sirola, L. Pylkkänen, H. Alenius, J. Hokkinen, U. Tapper, P. Lahtinen, A. Kapanen
and K. Putkisto, Cellulose, 2011, 18, 775-786.
153. M. Pitkänen, A. Sneck, H. Hentze, J. Sievänen, J. Hiltunen, E. Hellén, U. Honkalampi and A. von Wright,
Tappi International conference on nanotechnology for the forest products industry, 2010.
154. M. Pitkänen, H. Kangas, O. Laitinen, A. Sneck, P. Lahtinen, M. S. Peresin and J. Niinimäki, Cellulose, 2014,
Nanoscale Page 38 of 45

21, 3871-3886.
155. H. Mertaniemi, C. Escobedo-Lucea, A. Sanz-Garcia, C. Gandia, A. Mäkitie, J. Partanen, O. Ikkala and M.
View Article Online
Yliperttula, Biomaterials, 2012, 82, 208-220. DOI: 10.1039/C7NR04994C

156. V. R. Lopes, C. Sanchez-Martinez, M. Strømme and N. Ferraz, Particle and Fibre Toxicology, 2017, 14,
1-13.
157. L. Alexandrescu, K. Syverud, A. Gatti and G. Chinga-Carrasco, Cellulose, 2013, 20, 1765-1775.
158. M. S. Wang, F. Jiang, Y.-L. Hsieh and N. Nitin, Journal of Materials Chemistry B, 2014, 2, 6226-6235.
159. I. Díez, M. Pusa, S. Kulmala, H. Jiang, A. Walther, A. S. Goldmann, A. H. Müller, O. Ikkala and R. H. Ras,
Angewandte Chemie International Edition, 2009, 48, 2122-2125.
160. N. C. T. Martins, C. S. R. Freire, R. J. B. Pinto, S. C. M. Fernandes, C. P. Neto, A. J. D. Silvestre, J. Causio, G.
Baldi, P. Sadocco and T. Trindade, Cellulose, 2012, 19, 1425-1436.
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

161. H. Dong, J. F. Snyder, D. T. Tran and J. L. Leadore, Carbohydrate polymers, 2013, 95, 760-767.
162. K. J. De France, T. Hoare and E. D. Cranston, Chemistry of Materials, 2017, 29, 4609-4631.

Nanoscale Accepted Manuscript


163. N. C. T. Martins, C. S. R. Freire, C. P. Neto, A. J. D. Silvestre, J. Causio, G. Baldi, P. Sadocco and T. Trindade,
Colloids & Surfaces A Physicochemical & Engineering Aspects, 2013, 417, 111-119.
164. K. Missoum, Soft Matter, 2012, 8, 8338-8349.
165. K. Missoum, J. Bras and M. N. Belgacem, Cellulose, 2012, 19, 1957-1973.
166. K. Missoum, P. Sadocco, J. Causio, M. N. Belgacem and J. Bras, Materials Science and Engineering: C,
2014, 45, 477-483.
167. W. Xiao, J. Xu, X. Liu, Q. Hu and J. Huang, Journal of Materials Chemistry B, 2013, 1, 3477-3485.
168. K. Syverud, K. Xhanari, G. Chinga-Carrasco, Y. Yu and P. Stenius, Molecular & Cellular Biology, 1991, 11,
84-92.
169. G. Mcdonnell, . and A. D. Russell, Clinical Microbiology Reviews, 1999, 12, 147-179.
170. Martin Andresen, †, Per Stenstad, Trond Møretrø, Solveig Langsrud, Kristin Syverud, A. Leenasisko
Johansson and Per Stenius†, Biomacromolecules, 2007, 8, 2149-2155.
171. A. Pei, N. Butchosa, L. A. Berglund and Q. Zhou, Soft Matter, 2013, 9, 2047-2055.
172. S. Saini, F. Ç. Yücel, M. N. Belgacem and J. Bras, Carbohydrate Polymers, 2016, 135, 239-247.
173. K. Hua, D. O. Carlsson, E. Ålander, T. Lindström, M. Strømme, A. Mihranyan and N. Ferraz, Rsc Advances,
2013, 4, 2892-2903.
174. B. Madhushree, M. M. Malinen, L. Patrick, L. Yan-Ru, S. W. Kuisma, K. Liisa, L. Martina, C. Anne, G. G. G.
Christiane and I. Olli, Journal of Controlled Release, 2012, 164, 291-298.
175. H. Cai, S. Sharma, W. Liu, W. Mu, W. Liu, X. Zhang and Y. Deng, Biomacromolecules, 2014, 15, 2540-2547.
176. A. P. Mathew, K. Oksman, D. Pierron and M. F. Harmand, Carbohydrate Polymers, 2012, 87, 2291-2298.
177. A. Mihranyan, L. Nyholm, A. E. Bennett and M. Strømme, Journal of Physical Chemistry B, 2008, 112,
12249-12255.
178. G. Nystrom, A. Mihranyan, A. Razaq, T. Lindström, L. Nyholm and M. Strømme, The Journal of Physical
Chemistry B, 2010, 114, 4178-4182.
179. A. Razaq, A. Mihranyan, K. Welch, L. Nyholm and M. Strømme, Journal of Physical Chemistry B, 2009, 113,
426-433.
180. G. r. Frenning, A. Razaq, K. Gelin, L. Nyholm and A. Mihranyan, The Journal of Physical Chemistry B, 2009,
113, 4582-4589.
181. K. Gelin, A. Mihranyan, A. Razaq, L. Nyholm and M. Strømme, Electrochimica Acta, 2009, 54, 3394–3401.
182. F. Natalia, D. O. Carlsson, J. Hong, L. Rolf, F. Bengt, N. Leif, S. Maria and M. Albert, Journal of the Royal
Society Interface, 2012, 9, 1943-1955.
183. F. Natalia, S. Maria, F. Bengt, P. Sulena, N. Leif and M. Albert, Journal of Biomedical Materials Research
Part A, 2012, 100, 2128-2138.
184. R. Stefano, R. Aamir, N. Leif, S. M. Maria, L. Klaus and M. Albert, Journal of Physical Chemistry B, 2011,
Page 39 of 45 Nanoscale

114, 13644-13649.
185. M. Auad, V. Contos, S. Nutt, M. Aranguren and N. Marcovich, Proceedings of COMAT, 2005, 35-36.
View Article Online
186. Z. Song, H. Xiao and Y. Zhao, Carbohydrate Polymers, 2014, 111, 442-448. DOI: 10.1039/C7NR04994C

187. P. Lu, H. Xiao and Y. Pan, PROCEEDINGS OF THE 2014 INTERNATIONAL CONFERENCE ON MATERIALS
SCIENCE AND ENERGY ENGINEERING, 2015.
188. O. Faruk, M. Sain, R. Farnood, Y. Pan and H. Xiao, Journal of Polymers and the Environment, 2014, 22,
279-288.
189. S. Tsuguyuki, K. Ryota, W. Jakob, L. A. Berglund and I. Akira, Biomacromolecules, 2013, 14, 248-253.
190. B. M. Cherian, A. L. Leão, S. F. D. Souza, L. M. M. Costa, G. M. D. Olyveira, M. Kottaisamy, E. R. Nagarajan
and S. Thomas, Carbohydrate Polymers, 2011, 86, 1790-1798.
191. C. Eyholzer, A. B. D. Couraca and F. Duc, Biomacromolecules, 2011, 12, 1419-1427.
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

192. A. C. Borges, P. E. Bourban, D. P. Pioletti and J. A. E. Månson, Composites Science & Technology, 2010, 70,
1847–1853.

Nanoscale Accepted Manuscript


193. A. C. Borges, C. Eyholzer, F. Duc, P. E. Bourban, P. Tingaut, T. Zimmermann, D. P. Pioletti and J. A. E.
Månson, Acta Biomaterialia, 2011, 7, 3412-3421.
194. H. Valo, M. Kovalainen, P. Laaksonen, M. Häkkinen, S. Auriola, L. Peltonen, M. Linder, K. Järvinen, J.
Hirvonen and T. Laaksonen, Journal of Controlled Release, 2011, 156, 390-397.
195. G. Yuanqing, L. Xiaoyan, N. Tao and H. Jianguo, Chemical Communications, 2010, 46, 6096-6098.
196. R. A. Caruso, Angewandte Chemie International Edition, 2004, 43, 2746–2748.
197. J. Huang and T. Kunitake, Journal of the American Chemical Society, 2003, 125, 11834-11835.
198. M. Uchida, M. T. Klem, M. Allen, P. Suci, M. Flenniken, E. Gillitzer, Z. Varpness, L. O. Liepold, M. Young
and T. Douglas, Advanced Materials, 2007, 19, 1025-1042.
199. W. Czaja, D. Romanovicz and R. M. Brown, Cellulose, 2004, 11, 403-411.
200. M. N. A. And and A. Dufresne, Macromolecules, 2001, 34, 2921-2931.
201. B. L. Peng, N. Dhar, H. L. Liu and K. C. Tam, The Canadian Journal of Chemical Engineering, 2011, 89,
1191-1206.
202. M. S. Reid, M. Villalobos and E. D. Cranston, Langmuir, 2017, 33, 1583-1598.
203. H. Sadeghifar, I. Filpponen, S. P. Clarke, D. F. Brougham and D. S. Argyropoulos, Journal of Materials
Science, 2011, 46, 7344-7355.
204. Q. Lu, W. Lin, L. Tang, S. Wang, X. Chen and B. Huang, Journal of Materials Science, 2015, 50, 611-619.
205. H. Boerstoel, H. Maatman, J. B. Westerink and B. M. Koenders, Polymer, 2001, 42, 7371-7379.
206. S. Wei, V. Kumar and G. S. Banker, International Journal of Pharmaceutics, 1996, 142, 175-181.
207. X. Hao, S. Wei, Z. Chen, J. Zhu, F. Li, Z. Wu, W. Peng, X. Zeng and W. Tao, Carbohydrate Polymers, 2015,
123, 297-304.
208. M. Khandelwal and A. H. Windle, Polymer, 2013, 54, 5199-5206.
209. M. M. D. S. Lima, J. T. Wong, M. Paillet, R. Borsali and R. Pecora, Langmuir, 2003, 19, 24-29.
210. T. Kovacs, V. Naish, B. O'Connor, C. Blaise, F. Gagné, L. Hall, V. Trudeau and P. Martel, Nanotoxicology,
2010, 4, 255-270.
211. N. Lin and A. Dufresne, European Polymer Journal, 2014, 59, 302-325.
212. Z. Deng, H. Bouchékif, K. Babooram, A. Housni, N. Choytun and R. Narain, Journal of Polymer Science
Part A: Polymer Chemistry, 2008, 46, 4984-4996.
213. U. D. Hemraz, A. Lu, R. Sunasee and Y. Boluk, Journal of colloid and interface science, 2014, 430, 157-165.
214. R. Sunasee, U. D. Hemraz, K. Ckless, J. S. Burdick and Y. Boluk, MRS Online Proceedings Library Archive,
2015, 1718, 91-96.
215. E. J. Foster, M. J. D. Clift, B. Rothen-Rutishauser and C. Weder, Biomacromolecules, 2011, 12, 3666-3673.
216. K. Kümmerer, J. Menz, T. Schubert and W. Thielemans, Chemosphere, 2011, 82, 1387-1392.
217. H. Ni, S. Zeng, J. Wu, X. Cheng, T. Luo, W. Wang, W. Zeng and Y. Chen, Bio-medical materials and
Nanoscale Page 40 of 45

engineering, 2012, 22, 121-127.


218. N. Drogat, R. Granet, V. Sol, A. Memmi, N. Saad, C. K. Koerkamp, P. Bressollier and P. Krausz, Journal of
View Article Online
Nanoparticle Research, 2011, 13, 1557-1562. DOI: 10.1039/C7NR04994C

219. S. Wang, J. Sun, Y. Jia, L. Yang, N. Wang, Y. Xianyu, W. Chen, X. Li, R. Cha and X. Jiang, Biomacromolecules,
2016, 17, 2472-2478.
220. R. Xiong, C. Lu, W. Zhang, Z. Zhou and X. Zhang, Carbohydrate polymers, 2013, 95, 214-219.
221. A. R. Lokanathan, K. M. A. Uddin, O. J. Rojas and J. Laine, Biomacromolecules, 2013, 15, 373-379.
222. Y. Shin, I. T. Bae, B. W. Arey and G. J. Exarhos, Materials Letters, 2007, 61, 3215-3217.
223. C. M. Cirtiu, A. F. Dunlop-Brière and A. Moores, Green Chemistry, 2011, 13, 288-291.
224. X. Lin, M. Wu, D. Wu, S. Kuga, T. Endo and Y. Huang, Green Chemistry, 2011, 13, 283-287.
225. K. Benaissi, L. Johnson, D. A. Walsh and W. Thielemans, Green Chemistry, 2010, 12, 220-222.
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

226. E. Lam, S. Hrapovic, E. Majid, J. H. Chong and J. H. Luong, Nanoscale, 2012, 4, 997-1002.
227. L. Zhonghao and T. Andreas, Molecules, 2009, 14, 4682-4688.

Nanoscale Accepted Manuscript


228. Y. Shin, I. T. Bae, B. W. A. And and G. J. Exarhos, J.phys.chem.c, 2008, 112, 4844-4848.
229. P. Sonal, J. R. Capadona, S. J. Rowan, C. Weder, ., W. Yu-Ho, L. A. Stanciu and R. J. Moon, Langmuir the
Acs Journal of Surfaces & Colloids, 2010, 26, 8497-8502.
230. H. Liu, J. Song, S. Shang, Z. Song and D. Wang, ACS applied materials & interfaces, 2012, 4, 2413-2419.
231. E. Fortunati, I. Armentano, Q. Zhou, A. Iannoni, E. Saino, L. Visai, L. A. Berglund and J. Kenny,
Carbohydrate Polymers, 2012, 87, 1596-1605.
232. E. Fortunati, I. Armentano, Q. Zhou, D. Puglia, A. Terenzi, L. A. Berglund and J. Kenny, Polymer
Degradation and Stability, 2012, 97, 2027-2036.
233. E. Feese, H. Sadeghifar, H. S. Gracz, D. S. Argyropoulos and R. A. Ghiladi, Biomacromolecules, 2011, 12,
3528-3539.
234. C. Lemarchand, R. Gref and P. Couvreur, European Journal of Pharmaceutics and Biopharmaceutics, 2004,
58, 327-341.
235. D. E. Owens and N. A. Peppas, International journal of pharmaceutics, 2006, 307, 93-102.
236. P. Béguin and J.-P. Aubert, FEMS microbiology reviews, 1994, 13, 25-58.
237. L. R. Lynd, P. J. Weimer, W. H. Van Zyl and I. S. Pretorius, Microbiology and molecular biology reviews,
2002, 66, 506-577.
238. N. S. Mosier, C. M. Ladisch and M. R. Ladisch, Biotechnology and bioengineering, 2002, 79, 610-618.
239. J. K. Jackson, K. Letchford, B. Z. Wasserman, L. Ye, W. Y. Hamad and H. M. Burt, International journal of
nanomedicine, 2011, 6, 321.
240. M. Romany, S. Dong, A. Hirani and Y. W. Lee, in Cellulose nanocrystals for drug delivery, 2009.
241. K. A. Mahmoud, J. A. Mena, K. B. Male, S. Hrapovic, A. Kamen and J. H. Luong, ACS applied materials &
interfaces, 2010, 2, 2924-2932.
242. S. Dong and M. Roman, Journal of the American Chemical Society, 2007, 129, 13810-13811.
243. H. Wang and M. Roman, Biomacromolecules, 2011, 12, 1585-1593.
244. N. Lin, J. Huang, P. R. Chang, L. Feng and J. Yu, Colloids and Surfaces B: Biointerfaces, 2011, 85, 270-279.
245. J. Villanova, E. Ayres, S. Carvalho, P. Patrício, F. Pereira and R. Oréfice, European Journal of
Pharmaceutical Sciences, 2011, 42, 406-415.
246. N. Lin and A. Dufresne, Biomacromolecules, 2013, 14, 871-880.
247. J.-L. Huang, C.-J. Li and D. G. Gray, ACS Sustainable Chemistry & Engineering, 2013, 1, 1160-1164.
248. Q. Yang and X. Pan, Journal of applied polymer science, 2010, 117, 3639-3644.
249. I. Filpponen, H. Sadeghifar and D. S. Argyropoulo, Nanomaterials and nanotechnology, 2011, 1, 34-43.
250. M. L. Hassan, C. M. Moorefield, H. S. Elbatal, G. R. Newkome, D. A. Modarelli and N. C. Romano,
Materials Science and Engineering: B, 2012, 177, 350-358.
251. L. J. Nielsen, S. Eyley, W. Thielemans and J. W. Aylott, Chemical Communications, 2010, 46, 8929-8931.
Page 41 of 45 Nanoscale

252. T. Abitbol, A. Palermo, J. M. Moran-Mirabal and E. D. Cranston, Biomacromolecules, 2013, 14,


3278-3284.
View Article Online
253. X. Wang, Y. Guo, D. Li, H. Chen and R.-c. Sun, Chem. Commun., 2012, 48, 5569-5571. DOI: 10.1039/C7NR04994C

254. K. A. Mahmoud, K. B. Male, S. Hrapovic and J. H. Luong, ACS applied materials & interfaces, 2009, 1,
1383-1386.
255. H. Liu, D. Wang, Z. Song and S. Shang, Cellulose, 2011, 18, 67-74.
256. A. P. Mangalam, J. Simonsen and A. S. Benight, Biomacromolecules, 2009, 10, 497-504.
257. A. Y. Denisov, E. Kloser, D. G. Gray and A. K. Mittermaier, Journal of biomolecular NMR, 2010, 47,
195-204.
258. N. Marcovich, M. Auad, N. Bellesi, S. Nutt and M. Aranguren, Journal of materials research, 2006, 21,
870-881.
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

259. A. Pei, J.-M. Malho, J. Ruokolainen, Q. Zhou and L. A. Berglund, Macromolecules, 2011, 44, 4422-4427.
260. X. Cao, Y. Habibi and L. A. Lucia, Journal of Materials Chemistry, 2009, 19, 7137-7145.

Nanoscale Accepted Manuscript


261. N. Lin, J. Huang, P. R. Chang, J. Feng and J. Yu, Carbohydrate Polymers, 2011, 83, 1834-1842.
262. Y. Wang, H. Tian and L. Zhang, Carbohydrate Polymers, 2010, 80, 665-671.
263. Q. Wu, M. Henriksson, X. Liu and L. A. Berglund, Biomacromolecules, 2007, 8, 3687-3692.
264. L. Rueda, A. Saralegui, B. F. d’Arlas, Q. Zhou, L. A. Berglund, M. Corcuera, I. Mondragon and A. Eceiza,
Carbohydrate polymers, 2013, 92, 751-757.
265. S. Lin, J. Huang, P. R. Chang, S. Wei, Y. Xu and Q. Zhang, Carbohydrate polymers, 2013, 95, 91-99.
266. W. Shang, J. Huang, H. Luo, P. R. Chang, J. Feng and G. Xie, Cellulose, 2013, 20, 179-190.
267. E. Fortunati, M. Peltzer, I. Armentano, L. Torre, A. Jiménez and J. Kenny, Carbohydrate polymers, 2012, 90,
948-956.
268. N. Lin, G. Chen, J. Huang, A. Dufresne and P. R. Chang, Journal of Applied Polymer Science, 2009, 113,
3417-3425.
269. M. P. Arrieta, E. Fortunati, F. Dominici, J. López and J. M. Kenny, Carbohydrate polymers, 2015, 121,
265-275.
270. D. Bondeson and K. Oksman, Composites Part A: Applied Science and Manufacturing, 2007, 38,
2486-2492.
271. R. E. Drumright, P. R. Gruber and D. E. Henton, Advanced Materials, 2000, 12, 1841-1846.
272. Q. Shi, C. Zhou, Y. Yue, W. Guo, Y. Wu and Q. Wu, Carbohydrate polymers, 2012, 90, 301-308.
273. M. D. Sanchez-Garcia and J. M. Lagaron, Cellulose, 2010, 17, 987-1004.
274. Y. Habibi, S. Aouadi, J.-M. Raquez and P. Dubois, Cellulose, 2013, 20, 2877-2885.
275. A. N. Frone, S. Berlioz, J.-F. Chailan and D. M. Panaitescu, Carbohydrate Polymers, 2013, 91, 377-384.
276. N. Bitinis, R. Verdejo, J. Bras, E. Fortunati, J. M. Kenny, L. Torre and M. A. López-Manchado, Carbohydrate
polymers, 2013, 96, 611-620.
277. K. Ben Azouz, E. C. Ramires, W. Van den Fonteyne, N. El Kissi and A. Dufresne, ACS Macro Letters, 2011, 1,
236-240.
278. C. Zhou, Q. Wang and Q. Wu, Carbohydrate Polymers, 2012, 87, 1779-1786.
279. J. Yang, C.-R. Han, J.-F. Duan, F. Xu and R.-C. Sun, ACS applied materials & interfaces, 2013, 5, 3199-3207.
280. X. Xu, H. Wang, L. Jiang, X. Wang, S. A. Payne, J. Zhu and R. Li, Macromolecules, 2014, 47, 3409-3416.
281. M. A. S. A. Samir, F. Alloin, J.-Y. Sanchez and A. Dufresne, Polymer, 2004, 45, 4149-4157.
282. S. A. Paralikar, J. Simonsen and J. Lombardi, Journal of Membrane Science, 2008, 320, 248-258.
283. E. Kloser and D. G. Gray, Langmuir, 2010, 26, 13450-13456.
284. G. Chen, A. Dufresne, J. Huang and P. R. Chang, Macromolecular Materials & Engineering, 2009, 294, 59–
67.
285. J. O. Zoppe, M. S. Peresin, Y. Habibi, R. A. Venditti and O. J. Rojas, ACS applied materials & interfaces,
2009, 1, 1996-2004.
Nanoscale Page 42 of 45

286. W. Yuan, J. Yuan, F. Zhang and X. Xie, Biomacromolecules, 2007, 8, 1101-1108.


287. H. Lönnberg, L. Fogelström, Q. Zhou, A. Hult, L. Berglund and E. Malmström, Composites Science and
View Article Online
Technology, 2011, 71, 9-12. DOI: 10.1039/C7NR04994C

288. M. Labet and W. Thielemans, Cellulose, 2011, 18, 607-617.


289. A.-L. Goffin, J.-M. Raquez, E. Duquesne, G. Siqueira, Y. Habibi, A. Dufresne and P. Dubois, Polymer, 2011,
52, 1532-1538.
290. H. Youssef and D. Alain, Biomacromolecules, 2008, 9, 1974-1980.
291. M. Peltzer, A. Pei, Q. Zhou, L. Berglund and A. Jiménez, Polymer International, 2014, 63, 1056–1062.
292. M. S. Peresin, Y. Habibi, J. O. Zoppe, J. J. Pawlak and O. J. Rojas, Biomacromolecules, 2010, 11, 674-681.
293. C. Zhou, R. Chu, R. Wu and Q. Wu, Biomacromolecules, 2011, 12, 2617-2625.
294. C. Zhou, Q. Shi, W. Guo, L. Terrell, A. T. Qureshi, D. J. Hayes and Q. Wu, ACS applied materials & interfaces,
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

2013, 5, 3847-3854.
295. L. Rueda, A. Saralegi, B. Fernández-d’Arlas, Q. Zhou, A. Alonso-Varona, L. A. Berglund, I. Mondragon, M.

Nanoscale Accepted Manuscript


Corcuera and A. Eceiza, Cellulose, 2013, 20, 1819-1828.
296. J. R. Capadona, O. Van Den Berg, L. A. Capadona, M. Schroeter, S. J. Rowan, D. J. Tyler and C. Weder,
Nature Nanotechnology, 2007, 2, 765-769.
297. A. E. Way, L. Hsu, K. Shanmuganathan, C. Weder and S. J. Rowan, ACS Macro Letters, 2012, 1, 1001-1006.
298. S. P. Lin, I. L. Calvar, J. M. Catchmark and J. R. Liu, Cellulose, 2013, 20, 2191-2219.
299. T. Zhang, W. Wang, D. Zhang, X. Zhang, Y. Ma, Y. Zhou and L. Qi, Advanced Functional Materials, 2010, 20,
1152-1160.
300. W. Helbert, H. Chanzy, T. L. Husum, M. Schülein and S. Ernst, Biomacromolecules, 2003, 4, 481-487.
301. H. S. Barud, T. Regiani, R. F. C. Marques, W. R. Lustri, Y. Messaddeq and S. J. L. Ribeiro, Journal of
Nanomaterials, 2011, 8, 99-110.
302. M. M. Abeer, M. Amin, M. C. Iqbal and C. Martin, Journal of Pharmacy and Pharmacology, 2014, 66,
1047-1061.
303. J. WANG, Y. ZHU and J. DU, Journal of Mechanics in Medicine and Biology, 2011, 11, 285-306.
304. W. K. Czaja, D. J. Young, M. Kawecki and R. M. Brown, Biomacromolecules, 2007, 8, 1-12.
305. E. E. Brown and M.-P. G. Laborie, Biomacromolecules, 2007, 8, 3074-3081.
306. K. Yasuda, J. Ping Gong, Y. Katsuyama, A. Nakayama, Y. Tanabe, E. Kondo, M. Ueno and Y. Osada,
Biomaterials, 2005, 26, 4468-4475.
307. Z. Cai and J. Kim, Cellulose, 2010, 17, 83-91.
308. A. F. Jozala, L. C. de Lencastre-Novaes, A. M. Lopes, V. de Carvalho Santos-Ebinuma, P. G. Mazzola, A.
Pessoa-Jr, D. Grotto, M. Gerenutti and M. V. Chaud, Applied Microbiology and Biotechnology, 2016, 100,
2063-2072.
309. R. T. Olsson, M. A. Samir, G. Salazar-Alvarez, L. Belova, V. Ström, L. A. Berglund, O. Ikkala, J. Nogues and U.
W. Gedde, Nature nanotechnology, 2010, 5, 584-588.
310. W. Hu, S. Chen, J. Yang, Z. Li and H. Wang, Carbohydrate Polymers, 2014, 101, 1043-1060.
311. C. Long, D. Qi, T. Wei, J. Yan, L. Jiang and Z. Fan, Advanced Functional Materials, 2014, 24, 3953-3961.
312. Z. Y. Wu, C. Li, H. W. Liang, J. F. Chen and S. H. Yu, Angewandte Chemie, 2013, 125, 2997-3001.
313. K. Dieter, S. Dieter, K. Friederike, H. Nadine, K. Daniel and S. Barno, Macromolecular Symposia, 2009, 280,
60-71.
314. F. Hong and K. Qiu, Carbohydrate Polymers, 2008, 72, 545-549.
315. O. Shezad, S. Khan, T. Khan and J. K. Park, Carbohydrate Polymers, 2010, 82, 173-180.
316. M. Hornung, M. Ludwig, A. M. Gerrard and H. P. Schmauder, Engineering in Life Sciences, 2006, 6,
546-551.
317. F. Mohammadkazemi, M. Azin and A. Ashori, Carbohydrate Polymers, 2015, 117, 518-523.
318. C.-H. Kuo, J.-H. Chen, B.-K. Liou and C.-K. Lee, Food Hydrocolloids, 2016, 53, 98-103.
Page 43 of 45 Nanoscale

319. J. H. Ha, N. Shah, M. Ul-Islam, T. Khan and J. K. Park, Process Biochemistry, 2011, 46, 1717-1723.
320. Y. Kojima, A. Seto, N. Tonouchi, T. Tsuchida and F. Yoshinaga, Bioscience, Biotechnology, and Biochemistry,
View Article Online
1997, 61, 1585-1586. DOI: 10.1039/C7NR04994C

321. H. J. Son, M. S. Heo, Y. G. Kim and S. J. Lee, Biotechnology and applied biochemistry, 2001, 33, 1-5.
322. L. Cooperman and D. Michaeli, Journal of the American Academy of Dermatology, 1984, 10, 647-651.
323. M. C. Gómez-Guillén, B. Giménez, M. E. López-Caballero and M. P. Montero, Food Hydrocolloids, 2011,
25, 1813-1827.
324. F. Guilak, D. Butler, S. Goldstein and D. Mooney, in Functional tissue engineering, 2006.
325. X. Kong, F. Cui, X. Wang, M. Zhang and W. Zhang, Journal of Crystal Growth, 2004, 270, 197-202.
326. S. Moreira, N. B. Silva, J. Almeida-Lima, H. A. O. Rocha, S. R. B. Medeiros, C. Alves Jr and F. M. Gama,
Toxicology Letters, 2009, 189, 235-241.
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

327. S. I. Jeong, S. E. Lee, H. Yang, Y. H. Jin, C. S. Park and S. P. Yong, Molecular & Cellular Toxicology, 2010, 6,
370-377.

Nanoscale Accepted Manuscript


328. G. Helenius, H. Bäckdahl, A. Bodin, U. Nannmark, P. Gatenholm and R. Bo, Journal of Biomedical
Materials Research Part A, 2006, 76, 431-438.
329. D. F. Ring, W. Nashed and T. Dow, U.S. Patent 4588400, 1986.
330. D. F. Ring, W. Nashed and T. Dow, U.S. Patent 4655758, 1987.
331. L. Fu, J. Zhang and G. Yang, Carbohydrate Polymers, 2013, 92, 1432-1442.
332. O. Portal, W. A. Clark and D. J. Levinson, Wounds: a compendium of clinical research and practice, 2009,
21, 1-3.
333. W. Czaja, A. Krystynowicz, S. Bielecki and R. M. Brown, Biomaterials, 2006, 27, 145-151.
334. U. Schönfelder, M. Abel, C. Wiegand, D. Klemm, P. Elsner and U.-C. Hipler, Biomaterials, 2005, 26,
6664-6673.
335. C. Wiegand, P. Elsner, U.-C. Hipler and D. Klemm, Cellulose, 2006, 13, 689-696.
336. L. Fu, Y. Zhang, C. Li, Z. Wu, Q. Zhuo, X. Huang, G. Qiu, P. Zhou and G. Yang, Journal of Materials
Chemistry, 2012, 22, 12349-12357.
337. Y. Qiu, L. Qiu, J. Cui and Q. Wei, Materials Science and Engineering: C, 2016, 59, 303-309.
338. C. Zhijiang, Y. Guang and J. Kim, Current Applied Physics, 2011, 11, 247-249.
339. C. Zhijiang, H. Chengwei and Y. Guang, Carbohydrate polymers, 2012, 87, 1073-1080.
340. X. Wen, Y. Zheng, J. Wu, L. Yue, C. Wang, J. Luan, Z. Wu and K. Wang, Progress in Natural Science:
Materials International, 2015, 25, 197-203.
341. W.-C. Lin, C.-C. Lien, H.-J. Yeh, C.-M. Yu and S.-h. Hsu, Carbohydrate polymers, 2013, 94, 603-611.
342. R. J. Pinto, P. A. Marques, C. P. Neto, T. Trindade, S. Daina and P. Sadocco, Acta Biomaterialia, 2009, 5,
2279-2289.
343. J. Wu, Y. Zheng, X. Wen, Q. Lin, X. Chen and Z. Wu, Biomedical materials, 2014, 9, 035005.
344. C. Liu, D. Yang, Y. Wang, J. Shi and Z. Jiang, Journal of Nanoparticle Research, 2012, 14, 1-12.
345. G. Yang, J. Xie, F. Hong, Z. Cao and X. Yang, Carbohydrate Polymers, 2012, 87, 839-845.
346. T. Maneerung, S. Tokura and R. Rujiravanit, Carbohydrate polymers, 2008, 72, 43-51.
347. H. S. Barud, C. Barrios, T. Regiani, R. F. Marques, M. Verelst, J. Dexpert-Ghys, Y. Messaddeq and S. J.
Ribeiro, Materials Science and Engineering: C, 2008, 28, 515-518.
348. G. Yang, J. Xie, Y. Deng, Y. Bian and F. Hong, Carbohydrate Polymers, 2012, 87, 2482-2487.
349. Z. Li, L. Wang, S. Chen, C. Feng, S. Chen, N. Yin, J. Yang, H. Wang and Y. Xu, Cellulose, 2015, 22, 373-383.
350. S. Ifuku, M. Tsuji, M. Morimoto, H. Saimoto and H. Yano, Biomacromolecules, 2009, 10, 2714-2717.
351. J. Feng, Q. Shi, W. Li, X. Shu, A. Chen, X. Xie and X. Huang, Cellulose, 2014, 21, 4557-4567.
352. J. Luan, J. Wu, Y. Zheng, W. Song, G. Wang, J. Guo and X. Ding, Biomedical Materials, 2012, 7, 065006.
353. J. George, V. A. Sajeevkumar, K. V. Ramana and S. N. Sabapathy, Journal of Materials Chemistry, 2012, 22,
22433-22439.
Nanoscale Page 44 of 45

354. M. Ul-Islam, T. Khan, W. A. Khattak and J. K. Park, Cellulose, 2013, 20, 589-596.
355. D. Ciechaoska, Fibres & Textiles in Eastern Europe, 2004, 12, 48.
View Article Online
356. J. Kim, Z. Cai, H. S. Lee, G. S. Choi, D. H. Lee and C. Jo, Journal of Polymer Research, 2011, 18,
DOI:739-744.
10.1039/C7NR04994C

357. N. Butchosa, C. Brown, P. T. Larsson, L. A. Berglund, V. Bulone and Q. Zhou, Green chemistry, 2013, 15,
3404-3413.
358. C. Katepetch, R. Rujiravanit and H. Tamura, Cellulose, 2013, 20, 1275-1292.
359. J. Geng, D. Yang, Y. Zhu, L. Cao, Z. Jiang and Y. Sun, Journal of Nanoparticle Research, 2011, 13,
2661-2670.
360. D. Sun, J. Yang and X. Wang, Nanoscale, 2010, 2, 287-292.
361. I. F. Almeida, T. Pereira, N. H. Silva, F. P. Gomes, A. J. Silvestre, C. S. Freire, J. M. Sousa Lobo and P. C.
Costa, European Journal of Pharmaceutics & Biopharmaceutics Official Journal of Arbeitsgemeinschaft
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

Für Pharmazeutische Verfahrenstechnik E V, 2014, 86, 332-336.


362. E. Trovatti, C. S. R. Freire, P. C. Pinto, I. F. Almeida, P. Costa, A. J. D. Silvestre, C. P. Neto and C. Rosado,

Nanoscale Accepted Manuscript


International Journal of Pharmaceutics, 2012, 435, 83-87.
363. A. Volk, M. Kim and G. Dodge, Osteoarthritis and Cartilage, 2009, 17, S105.
364. A. Svensson, E. Nicklasson, T. Harrah, B. Panilaitis, D. Kaplan, M. Brittberg and P. Gatenholm, Biomaterials,
2005, 26, 419-431.
365. J. Andersson, H. Stenhamre, H. Bäckdahl and P. Gatenholm, Journal of Biomedical Materials Research
Part A, 2010, 94, 1124-1132.
366. A. Bodin, S. Concaro, M. Brittberg and P. Gatenholm, Journal of tissue engineering and regenerative
medicine, 2007, 1, 406-408.
367. C. Rambo, D. Recouvreux, C. Carminatti, A. Pitlovanciv, R. Antônio and L. Porto, Materials Science and
Engineering: C, 2008, 28, 549-554.
368. P. Gatenholm and D. Klemm, MRS bulletin, 2010, 35, 208-213.
369. C. Gao, Y. Wan, C. Yang, K. Dai, T. Tang, H. Luo and J. Wang, Journal of Porous Materials, 2011, 18,
139-145.
370. A. Bodin, S. Bharadwaj, S. Wu, P. Gatenholm, A. Atala and Y. Zhang, Biomaterials, 2010, 31, 8889-8901.
371. A. Putra, A. Kakugo, H. Furukawa, J. P. Gong and Y. Osada, Polymer, 2008, 49, 1885-1891.
372. P. A. Charpentier, A. Maguire and W.-k. Wan, Applied Surface Science, 2006, 252, 6360-6367.
373. H. Bäckdahl, G. Helenius, A. Bodin, U. Nannmark, B. R. Johansson, B. Risberg and P. Gatenholm,
Biomaterials, 2006, 27, 2141-2149.
374. H. Bäckdahl, B. Risberg and P. Gatenholm, Materials Science and Engineering: C, 2011, 31, 14-21.
375. L. Millon and W. Wan, Journal of Biomedical Materials Research Part B: Applied Biomaterials, 2006, 79,
245-253.
376. L. Millon, H. Mohammadi and W. Wan, Journal of Biomedical Materials Research Part B: Applied
Biomaterials, 2006, 79, 305-311.
377. J. Wippermann, D. Schumann, D. Klemm, J. Albes, T. Wittwer, J. Strauch and T. Wahlers, The Thoracic and
Cardiovascular Surgeon, 2008, 56, P22.
378. D. Klemm, U. Udhardt, S. Marsch and D. Schumann, Polymer News, 1999, 24, 377-378.
379. M. Esguerra, H. Fink, M. W. Laschke, A. Jeppsson, D. Delbro, P. Gatenholm, M. D. Menger and B. Risberg,
Journal of Biomedical Materials Research Part A, 2010, 93, 140-149.
380. F. K. Andrade, R. Costa, L. Domingues, R. Soares and M. Gama, Acta Biomaterialia, 2010, 6, 4034-4041.
381. K. J. Burg, S. Porter and J. F. Kellam, Biomaterials, 2000, 21, 2347-2359.
382. A. J. Salgado, O. P. Coutinho and R. L. Reis, Macromolecular bioscience, 2004, 4, 743-765.
383. J. Wang, C. M. Valmikinathan, W. Liu, C. T. Laurencin and X. Yu, Journal of Biomedical Materials Research
Part A, 2010, 93, 753-762.
384. J.-H. Jang, O. Castano and H.-W. Kim, Advanced drug delivery reviews, 2009, 61, 1065-1083.
Page 45 of 45 Nanoscale

385. P. N. Mendes, S. C. Rahal, O. C. M. Pereira-Junior, V. E. Fabris, S. L. R. Lenharo, J. F. de Lima-Neto and F. da


Cruz Landim-Alvarenga, Acta Veterinaria Scandinavica, 2009, 51, 1.
View Article Online
386. M. Zaborowska, A. Bodin, H. Bäckdahl, J. Popp, A. Goldstein and P. Gatenholm, Acta Biomaterialia, 2010,
DOI: 10.1039/C7NR04994C

6, 2540-2547.
387. M. González, E. Hernández, J. Ascencio, F. Pacheco, S. Pacheco and R. Rodríguez, Journal of Materials
Chemistry, 2003, 13, 2948-2951.
388. L. Hong, Y. Wang, S. Jia, Y. Huang, C. Gao and Y. Wan, Materials Letters, 2006, 60, 1710-1713.
389. B. Fang, Y.-Z. Wan, T.-T. Tang, C. Gao and K.-R. Dai, Tissue Engineering Part A, 2009, 15, 1091-1098.
390. C. Gao, G. Y. Xiong, H. L. Luo, K. J. Ren, Y. Huang and Y. Z. Wan, Cellulose, 2010, 17, 365-373.
391. Y. Wan, Y. Huang, C. Yuan, S. Raman, Y. Zhu, H. Jiang, F. He and C. Gao, Materials Science and Engineering:
C, 2007, 27, 855-864.
Published on 05 September 2017. Downloaded by University of Windsor on 05/09/2017 16:36:39.

392. S. A. Hutchens, R. S. Benson, B. R. Evans, H. M. O’Neill and C. J. Rawn, Biomaterials, 2006, 27, 4661-4670.
393. C. J. Grande, F. G. Torres, C. M. Gomez and M. C. Bañó, Acta Biomaterialia, 2009, 5, 1605-1615.

Nanoscale Accepted Manuscript


394. C. Tsioptsias and C. Panayiotou, Carbohydrate polymers, 2008, 74, 99-105.
395. N. Chiaoprakobkij, N. Sanchavanakit, K. Subbalekha, P. Pavasant and M. Phisalaphong, Carbohydrate
polymers, 2011, 85, 548-553.
396. T. G. M. van de Ven and A. Sheikhi, Nanoscale, 2016, 8, 15101-15114.
397. A. Sheikhi and T. G. M. van de Ven, Current Opinion in Colloid & Interface Science, 2017, 29, 21-31.

You might also like