Anda di halaman 1dari 37

Accepted Manuscript

Functionalization of cellulose nanocrystals for advanced applications

Juntao Tang, Jared Sisler, Nathan Grishkewich, Kam Chiu Tam

PII: S0021-9797(17)30100-5
DOI: http://dx.doi.org/10.1016/j.jcis.2017.01.077
Reference: YJCIS 21983

To appear in: Journal of Colloid and Interface Science

Received Date: 22 November 2016


Revised Date: 15 January 2017
Accepted Date: 22 January 2017

Please cite this article as: J. Tang, J. Sisler, N. Grishkewich, K.C. Tam, Functionalization of cellulose nanocrystals
for advanced applications, Journal of Colloid and Interface Science (2017), doi: http://dx.doi.org/10.1016/j.jcis.
2017.01.077

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Functionalization of cellulose nanocrystals for advanced applications

Juntao Tang, Jared Sisler, Nathan Grishkewich, Kam Chiu Tam

Department of Chemical Engineering, Waterloo Institute for Nanotechnology, University of


Waterloo, 200 University Avenue, Waterloo, ON, N2L 3G1, Canada

Abstract

Replacing the widespread use of petroleum-derived non-biodegradable materials with


green and sustainable materials is a pressing challenge that is gaining increasing attention by the
scientific community. One such system is cellulose nanocrystals (CNCs) derived from acid
hydrolysis of cellulosic materials, such as plants, tunicates and agriculture biomass. The
utilization of colloidal CNCs can aid in the reduction of carbon dioxide that is responsible for
global warming and climate change. CNCs are excellent candidates for the design and
development of functional nanomaterials in many applications due to several attractive features,
such as high surface area, hydroxyl groups for functionalization, colloidal stability, low toxicity,
chirality and mechanical strength. Several large scale manufacturing facilities have been
commissioned to produce CNCs of up to 1 000 kg/day, and this has generated increasing
interests in both academic and industrial laboratories. In this feature article, we will describe the
recent development of functionalized cellulose nanocrystals for several important applications in
ours and other laboratories. We will highlight some challenges and offer perspectives on the
potentials of this sustainable nanomaterial.
List of abbreviations
Cellulose nanocrystals (CNCs)
Nanofibrillated cellulose (NFC)
Small angle neutron scattering (SANS)
High internal phase emulsion (HIPE)
Didecyldimethylammonium bromide (DMAB)
Cetyltrimethylammonium bromide (CTAB)
Poly(dimetheylamino ethylmethacrylate) (PDMAEMA)
Poly(oligoethylene glycol) methacrylate (POEGMA)
Poly(methacrylic acid) (PMAA)
Poly(N-isopropylacrylamine) (PNIPAM)
Hydroxyethyl cellulose (HEC)
Polyacrylamide (PAAM)
Polarized optical microscopy (POM)
Tetraethyl orthosilicate (TEOS)
Tetramethyl orthosilicate (TMOS)
Titanium(IV) ethoxide (TEOT)
Urea-formaldehyde (UF)
Poly(vinyl alcohol) (PVA)
Poly(amidoamine) (PAMAM)
Melamine-formaldehyde resin (MF)
Polyrhodanine (PR)
Poly(N-vinylpyrrolidone) (PVP)
Chitosan oligosaccharide (CSos)
Imipramine hydrochloride (IMI)
Procaine hydrochloride (PrHy)
Doxorubicin (DOX)
Poly(ethyl ethylene phosphate) (PEEP)
Vitamin C (VC)
Fluorescein-5-isothiocyanate (FITC)
Rhodamine B isothiocyanate (RBITC)
1. Introduction
In nanoscience and nanotechnology, the synthesis and modification of nanomaterials with
well-defined structure and functionalities have attracted growing interest due to their many
potential applications [1,2]. Recent advances in nanomaterials have led to the development of
functionalized nanoparticles that hold promise in various industrial sectors, such as medicine,
electronics, biomaterials and energy production [3,4]. However, a large proportion of chemicals
used to produce nanomaterials are derived from petroleum-based resources, and they involve the
use of toxic reagents that are harmful to the environment. Due to concerns on global warming
and sustainable development, there is an urgent need to replace traditional raw material supply
with those derived from renewable resources [5]. Furthermore, the ability to transform cheap and
abundant material to yield high value products will offer significant advantage.
It is well-known that cellulose is the most abundant naturally occurring polymer found in
this planet [6]. It represents about 1.5x1012 tons (metric tonne) of total annual biomass
production and is considered an inexhaustible source of raw material capable of meeting the
increasing demand for environmentally friendly and biocompatible products [6,7]. Economical
and environmentally friendly methods have been developed to process cellulosic materials by
dissolving them in NaOH/urea solution or ionic liquids, as reported by Zhangs [810] and
Rogers laboratories [11,12], respectively. Unfortunately, the approach of disassembling
cellulosic materials to their molecular entities sacrifices the attractive physical properties of the
crystalline domains formed by the inter- and intra-molecular hydrogen bonds. By careful control
of the disassociation of amorphous regions while retaining the crystalline domains, a new form
of crystalline cellulose commonly referred to as nanocellulose is produced [7]. These
nanocelluloses have size in the nanometer regime, and they possess many attractive
characteristics, such as versatile fiber morphology, hydrophilicity, easy surface modification,
large surface area and high aspect ratios [7]. Depending on their dimensions, origins and
processing conditions, nanocellulose can be divided into two main categories, namely cellulose
nanocrystals and cellulose nanofibrils. In this feature article, we will focus mainly on cellulose
nanocrystals, however reviews on cellulose nanofibrils can be found in several recent
publications [7,1316].
2. Cellulose nanocrystals
Cellulose nanocrystals (CNCs) are the crystalline domains extracted from wood fiber
through acid hydrolysis. They are rigid, rod-like particles with a width of several nanometers and
lengths of up to hundreds of nanometers [17,18]. The microscopic properties (physical and
surface chemistry) of CNCs have an important bearing on their macroscopic properties (rheology,
colloidal stability, etc), and these are summarized in Figure 1.

Figure 1 A summary of the physical and chemical properties of cellulose nanocrystals.

2.1 Physical properties


The main physical dimensions for cellulose nanocrystals include the length (L), diameter
(D) and aspect ratio (L/D), which are dependent on the source of cellulose or hydrolysis
conditions (acid type, reaction time and temperature). CNCs derived from wood and cotton is
usually shorter than that obtained from tunicate and bacterial cellulose because the latter possess
a higher degree of crystallinity [17,18]. Lower fractions of amorphous regions make them more
resistant to degradation from acid hydrolysis resulting in larger rod structures. Typically, the
aspect ratio ranges from 10-30 for CNCs derived from cotton and up to approximately 70 for
tunicate. Sulfuric and hydrochloric acids are the most commonly used acids in the hydrolysis
process, but other strong acids, such as phosphoric and hydrobromic acid have also been reported
[14,17]. Different acids may lead to significant differences in the dispersity and colloidal
stability of CNCs. For instance, CNCs derived from sulfuric acid hydrolysis disperses readily in
water due to the abundance of negatively charged sulfate ester groups on their surface, while
aqueous solutions of CNCs produced from hydrochloric acid hydrolysis display poor colloidal
stability [18]. In addition to the properties discussed above, cellulose nanocrystals also possess
other attractive features: large surface area (250~500 m2/g), and improved mechanical strength
(tensile strength 7 500 MPa and Youngs modulus of 100~140 GPa)[19]. The extremely high
Youngs modulus is an attractive characteristic for application in nanocomposites. Liquid
crystalline behavior has also been observed for non-flocculating cellulose nanocrystal
suspensions. In the dilute solution regime, CNCs are isotropic, and at higher concentrations the
nanoparticles align to form an anisotropic nematic phase [17]. Beyond this critical concentration,
CNCs dispersions display shear birefringence, and they can spontaneously phase separate into an
upper isotropic and a lower anisotropic phase. The chiral nematic or cholesteric structure in the
anisotropic phase possesses a helical twist along the main axis, with the orientation of each stack
planes being rotated about the perpendicular axis. The parallel alignment of the CNCs is
attributed to the well-known entropically driven self-orientation phenomenon, and the helix of
cellulose nanocrystals is left-handed, reflecting the intrinsic chirality of crystalline cellulose.
However, the pitch distance between the different planes can vary significantly, ranging from
less than 1 to more than 50 nm, and it is a function of temperature, sonication time and ionic
strength, but it is independent of the concentration of CNCs [20]. More interestingly, the chiral
nematic structure of the suspension can be preserved via slow and complete evaporation of the
water phase, yielding an iridescent film. The spectacular iridescent coloring originates from the
reflection of light by the chiral nematic phases in a Bragg-type manner. The reflected color of the
films can be manipulated by varying the pitch of the helical structure, and these iridescent
materials are of great interest in coatings, security features and sensors.

2.2 Surface chemistry properties


As the main chemical component of cellulose nanocrystals consists of cellulose chains,
all classical chemistry on cellulose or polysaccharides is applicable to cellulose nanocrystals.
However, CNCs is thought to be less reactive when compared to amorphous cellulose chains
since most of the polymer chains are buried within the inaccessible crystalline regions. The
monomeric glucose units of the cellulose chain possess three hydroxyl groups, which provide
reactive platforms for chemical modifications. Aside from abundant hydroxyl groups, the surface
of CNCs may contain other types of functional groups that are directly related to its preparation
and processing conditions. The common functional groups are sulfate groups ( ) and
-
carboxyl groups (-COO ). With additional mild post-hydrolysis reactions, aldehyde groups (-
CHO), amino groups (-NH2) or thiol groups (-SH) may also be introduced to the CNCs surface.
Depending on the specific functional groups on the surface, CNCs nanoparticles exhibit different
charge properties. CNCs bearing sulfate or carboxylate groups on the surface are negatively
charged over a wide range of pH conditions (above its pKa), while the amino groups are
positively charged below the pKa values of the weak base. In addition, modifying the cellulose
nanocrystals with quaternary ammonium groups will render their surface with permanent
cationic charges.

Figure 2 Evolution of the number of research publications and citations on cellulose


nanocrystals during the past ten years (2006-2016) according to ISI Web of Knowledge system.

3. Functional cellulose nanocrystals for advanced applications


In the past 20 years, the investigation and utilization of cellulose nanocrystals in
functional materials has become an active field of research activity and many researchers have
dedicated their efforts to the study of this remarkable material, which is reflected by the growing
number of publications and citations as summarized in Figure 2. The prospect of modifying and
functionalizing cellulose nanocrystals is attractive as it enables the creation of advanced
materials with new or improved properties [21]. By introducing the functional components
(materials or chemical groups) to the system, synergistic effects can be achieved, which can
impart electronic, magnetic, catalytic, fluorescence and optical properties. Thus, their
functionalities will be improved and potential applications in specific fields can be expanded. In
the current article, we will highlight recent developments in the functionalization of cellulose
nanocrystals and offer prospect and potentials of this component in novel sustainable
nanomaterials for the future.

3.1 Emulsion stabilizer


Currently, there is a growing market trend toward the formulation of products that can
maintain the consumer perception of being natural and green. This has motivated the
production of biobased nanoparticles for the formulation of Pickering emulsions for the food and
cosmetic industries [22,23]. Cellulose nanocrystals are ideal for this application, and they have
been shown to be effective Pickering emulsifiers [24,25]. They can stabilize monodispersed oil
(hexadecane) droplets of ~4 m in water phase against coalescence for 4 months (Figure 3A).
This is due to the partitioning of stable CNCs nanoparticles at the oil-water interface that
significantly enhance the stability of oil droplets [24]. Further research suggested that cellulose
nanocrystals with a charge density greater than 0.03 e/nm2 could not efficiently stabilize oil
droplets due to the strong electrostatic repulsions between the nanoparticles located at the oil-
water interface [25]. In addition, neutral CNCs extracted by HCl hydrolysis performed better
than sulfated CNCs at the oil/water interface. Kalashnikova and Carpon showed that the aspect
ratio has a direct impact on the interfacial coverage, where a low aspect ratio resulted in a dense
organization of short nanocrystals at the oil-water interface [26] (Figure 3B). Recently, the
packing characteristics of CNCs (195 nm long, 23 nm width and 6 nm thick) with different
surface charges at the interface was examined by Carpon and coworkers [27] using small angle
neutron scattering (SANS). They reported that the average thickness of the layer around the oil
droplets was determined to be 7 and 18 nm for charged and uncharged CNCs, respectively. This
result supported the postulate that the (2 0 0) crystalline plane of the nanoparticles directly
interacts with the interface. Due to the colloidal network structure that forms at the interface, oil-
in-water high internal phase emulsion (HIPE) systems [28] as well as water-in-water emulsions
[29] could be stabilized. The HIPE system displayed a gel-like behavior, which can only be
produced via a two-step method consisting of first the formation of the primary Pickering
emulsion and second a subsequent swelling.
Figure 3 (A) Scanning electron micrographs of polymerized styrene Pickering emulsion
stabilized by bacterial cellulose nanocrystals; Schematic representation of the stabilization of the
I cotton cellulose nanocrystals at the oil/water interface [25]; (B) Scanning electron
microscopy (SEM) images of polymerized styrenewater emulsions stabilized by CNCs with
different aspect ratios [26]; (C) Water-dodecane emulsions (1:1 by volume) stabilized by 0.25
wt.% CNCs and surfactant (a) CTAB and (b) DMAB with concentrations from 0 to 16 mM [30];
(D) Responsive behavior of emulsions (PDMAEMA-g-CNCs, 0.5 wt %) by adjusting the pH
values of the aqueous phase [37]; (E) Schematic illustrating the pH-responsive behavior of
Pickering emulsions stabilized by CNCs-POEGMA-PMAA and the demonstration for oil
harvesting application [38]; (F) Illustration of a redispersable Pickering emulsion using tannic
acid and HEC modified cellulose nanocrystals as stabilizer [40].

Aside from the investigation on emulsion systems stabilized by pristine CNCs, various
modification strategies have been used to manipulate the surface functionalities of CNCs in
controlling the physical properties of Pickering emulsions. This was achieved through the
incorporation of surfactants [30], functional groups or surface active polymers, for example
TEMPO oxidation and adsorption [31], periodate oxidation and amination [32], esterification
[33], as well as long alkyl chain grafting [34]. Two types of cationic charged surfactants
didecyldimethylammonium bromide (DMAB) and cetyltrimethylammonium bromide (CTAB)
were adsorbed on the CNCs surface to tailor the hydrophobicity of the nanoparticles, and their
capability to stabilize emulsions was investigated by Cranston and coworkers [30]. They
observed a double transitional phase inversion (from oil-in-water to water-in-oil and then back to
oil-in-water) for emulsions stabilized by CNCs with increasing amounts of DMAB (a more
hydrophobic molecule). However, no phase inversion could be induced for CNCs modified by
CTAB (Figure 3C). Similarly, Capron and coworkers reported a simple method to prepare
hydrophobic CNCs by adsorbing quaternary ammonium salts onto the TEMPO-oxidized CNCs,
and the modified nanoparticles were capable of stabilizing inverse water-in-oil emulsions [31].
Pelton and coworkers studied the effects of both surfactant and water-soluble polymers
(hydroxyethyl cellulose or methyl cellulose) on the properties of Pickering emulsions stabilized
by cellulose nanocrystals [35]. The polymer coated CNCs nanoparticles produced emulsions
with smaller droplet sizes, and the emulsions could resist coalescence when subjected to multiple
cycles of heating and cooling.
Compared to systems modified by physical adsorption, covalent chemical modifications
can provide a more robust and versatile platform. Capron and coworkers tailored the
hydrophobicity of CNCs and nanofibrillated cellulose (NFC) by chemical modification with
lauroyl chloride (C12) [34]. They observed that the Pickering emulsions stabilized by the
combination of two types of modified nanocellulose could be formulated into oil-in-water-in-oil
(o/w/o) emulsions depending on the degree of substitutions. Sebe and coworkers modified the
CNCs surface with vinyl acetate (VAc) and vinyl cinnamate (VCIn) and observed that the
esterification treatment may significantly impact their utilization as Pickering emulsifiers [33].
VCIn-treated particles could only stabilize the cyclohexane-in-water emulsions, while the acetyl
modified CNCs could be used to prepare stable ethyl acetate-in-water, toluene-in-water, and
cyclohexane-in-water emulsions. Furthermore, in order to tailor the control of Pickering
emulsions for specific applications, much research has been devoted to the development of
emulsifiers that activate and deactivate in response to external stimuli. Zoppe et al. grafted
thermo-responsive Poly(N-isopropylacrylamine) (PNIPAM) onto CNCs surface and compared
the different nanoparticles in stabilizing emulsion systems [36]. They found that modified
cellulose nanocrystals could stabilize the emulsions for a period of 4 months compared to
unmodified CNCs nanoparticles. Tang and coworkers reported a dual-responsive (pH and thermo)
system based on poly(dimetheylamino ethylmethacrylate) (PDMAEMA) grafted cellulose
nanocrystals [37]. They demonstrated the feasibility of stabilizing both toluene- and heptane-in-
water emulsions under basic conditions. Decreasing the pH values lead to the protonation of
tertiary amines on PDMAEMA chains, that promotes the electrostatic interaction between CNCs
particles and polymer chains, resulting in a reversible particle aggregation and emulsion
instability (Figure 3D). Following this, they grafted binary polymer brushes consisting of
poly(oligoethylene glycol) methacrylate (POEGMA) and poly(methacrylic acid) (PMAA) on the
surface of CNCs nanoparticles [38]. This would permit the control of the stability of Pickering
emulsions using two types of triggers, i.e., POEGMA for temperature and PMAA for the pH.
They demonstrated a reversible emulsification-demulsification process controlled by pH using
the binary brush grafted nanoparticles, where the emulsification and oil-water separation could
be repeated 5 times without any loss in efficiency (Figure 3E).
Cellulose nanocrystal based Pickering emulsions could be formulated for use in
encapsulation systems. Marquis and coworkers described a two-step approach to encapsulate oil
microdroplets within alginate microgels [39]. The microdroplets were oil-in-water Pickering
emulsions that were stabilized by cellulose nanocrystals and calcium carbonate. They further
demonstrated that the CNCs layer could provide an ideal shell to prevent the coalescence of oil
droplets. The Ca2+ released from the CaCO3 particles could be used for the gelation of alginate to
form the microgel. Nile Red was used as a model compound to trace the release profile after
encapsulation, and the double encapsulation protocol could provide better protection as well as
sustained release when compared to the traditional method of encapsulation. Hu et al. has also
reported another encapsulation system, which was based on emulsions stabilized by
hydroxyethyl cellulose (HEC) modified CNCs nanoparticles [40]. They further coated emulsified
corn oil in water emulsions with tannic acid, which can be transformed into solid dry emulsions
(powders) via freeze-drying (Figure 3F). This work extended the use of surfactant free emulsions
for food, cosmetic and pharmaceutical applications.

3.2 Templates for functional materials


Hard templating using preformed mesoporous materials (also termed nanocasting) has
emerged as a versatile technique to prepare materials that cannot be accessed through
conventional lyotropic template synthesis, e.g. due to hydrolytic instability of precursors.
Cellulose nanocrystal dispersions can exhibit lyotropic chiral nematic behavior at relatively low
concentrations (e.g., 1-7 wt%), have lower viscosities and they form over shorter time scales
when compared to several other cellulose derivatives, i.e. ethyl cellulose or hydropropyl
ethylcellulose. This has generated a strong interest in using evaporation-induced self-assembly
protocols to prepare functional mesoporous materials with chiral nematic order [20,41]. In the
templating approach, successive loading of precursors permeate through a stable mesoporous
support, often followed by calcination to construct an interconnected network to produce the
desired product. The remaining active components can either be templating materials (CNCs) or
functional materials introduced in the synthesizing steps (Figure 4A and 4B). Many kinds of
ordered mesoporous materials (e.g., carbon, metal oxides, and polymers) can be prepared
through hard templating approaches. A summary on the different functional materials is
documented in Table 1.
MacLachlan and coworkers have published detailed reviews on the use of CNCs as a
templating material [42,43], hence we will only discuss research reports published in the last 2
years. Nguyen et al. have demonstrated a representative method of fabricating new porous
semiconducting material with chiral nematic structures [44]. This was achieved by first casting
SiO2/C composite films by cocondensing SiO2 with cellulose nanocrystals and subsequently
pyrolyzing the material. Then, magnesiothermic reduction was applied to finally covert the
SiO2/C composite into silicon carbide. The chiral nematic hierarchical structure originating from
the evaporation induced self assembly of CNCs was retained. Aside from templating chiral
nematic structures in 2D films, cellulose nanocrystals have also been used to prepare three
dimensional structural materials. Wang et al. used an inverse emulsion polymerization method to
capture the chiral nematic structure of CNCs within microspheres [45] (Figure 4C). They
demonstrated that the CNCs tactoids first formed within the water droplets, and subsequent
polymerization would solidify the microspheres to retain their structure. By incorporating the
hydrolysis reaction of silica within the microsphere formation process, organic-silica
microspheres were also fabricated. Upon further removal of the organic matrix, the mesoporous
silica microspheres with chiral nematic structures were obtained, and may have potential
applications in optical devices and chiral separations.

3.3 Functional cellulose nanocrystal-Inorganic hybrids


Inorganic materials are attractive components to be incorporated into cellulose
nanocrystal systems due to their size-dependent magnetic, catalytic and optical properties. For
example, graphene or graphene oxide, noble metal nanoparticles, quantum dots as well as metal
oxide nanoparticles have generated increasing interest among both academic and industrial
laboratories. By taking advantage of cellulose nanocrystals, the aggregation behavior of
thermodynamically unstable inorganic nanoparticles can be minimized or eliminated. CNCs are
sustainable and ideal for use as supporting materials, and they are considered to be non-toxic and
environmentally friendly. The ability of CNCs to form stable colloidal dispersions allows post-
processing into versatile products such as 1D dimensional inks and dispersions, 2D films or 3D
composites. In addition, their structures and properties can be readily adjusted during the
pretreatment and chemical modification process.

Figure 4 A representative illustration of synthesizing chiral nematic mesoporous materials using


cellulose nanocrystals as templates [62]; (B) Optical characterization of CNCs/silica composite
films and the corresponding mesoporous silica films [41]; (C) Schematicillustration on the
preparation of PAAM/CNCs/Silica composites spheres and the characterization of composite
spheres using polarized optical microscopy (POM), scanning electron microscopy and 3D
reconstructed confocal fluorescence microscopy [45].
Table 1 Summary regarding functional materials using CNCs as templates
Ref. Precursors Method Properties or application

[41] TEOS or TMOS Remove CNCs Mesoporous silica film

[46] TEOS or TMOS Remove CNCs +Ag NPs filled Ag assembled in chiral nematic Silica film
[47] TEOS or TMOS Pyrolysis-remove SiO2 Mesoporous Carbon
Mesoporous silica+TiCl4
[48] TMOS and TiCl4 Mesoporous TiO2
infiltration + silica etching
Ethylene-bridged organosilica Improving mechanical properties and flexibility
[49] Remove CNCs
precursors compared to pure silica
[50] TMOS Remove CNCs Ionic strength for color changing
EISA (evaporation induced self-
[51] PAAm hydrogel precusors Hydrogel sensor
assembly)
[52] Phenol-formaldehyde Remove CNCs Chiral mesoporous photonic resin
Adding Polyols such as glucose to eliminate the
[53] TMOS Remove CNCs
crack
Detailed investigation on the conditions for film (pH
[54] TMOS Remove CNCs
and ratio)
[55] Phenol-formaldehyde Remove CNCs Chiral nematic Structures and Actuator Properties

[56] TMOS PVA to reduce the crack Remove CNCs+ CdS QDs Mesoporous, chiral nematic order and luminescence
Alkaline treatment to remove UF Mesoporous chiral cellulose material displaying
[57] Urea formaldehyde
resin dynamic photonic properties
Chiral nematic SiO2 for Gas Chromatographic
[58] TMOS Remove CNCs
separation
[59] TEOT (Ti) Remove CNCs Mesoporous TiO2 film for solar cell
Precusor for prussian blue, organosilica Remove CNCs and silica
[60] precursors, TMOS Mesoporous coordination polymers
Porous semiconducting material with chiral nematic
[44] TMOS Magnesiothermic reduction structures
Remove PF, Photonic Patterns
[61] Phenol-formaldehyde Printed (dry or wet) Anticounterfeiting materials
Inverse emulsion polymerization Noval mesoporous silica microsphere for optical
[45] PAAM, TMOS and romove CNCs device and chiral separation
Urea formaldehyde, cobalt ferrite
[62] precusor Remove UF and CNCs Electromagnetic interference (EMI) shielding

The deposition of nanoparticles onto the surface of cellulose nanocrystals can generate
new hybrid materials that are suitable for use as heterogeneous catalysts in engineering
applications, especially for wastewater treatment. By using rod-like pristine cellulose
nanocrystals as substrate, various types of catalysts have been prepared including nickel
nanocrystals [63], palladium nanoparticles [64,65], TiO2 nanocubes [66], and alloy nanoparticles
[67]. They have been widely used in chemical reduction reactions (e.g. 4-nitrophenol to 4-
aminophenol; oxygen reduction reaction; hydrogenation of C-C and C-O multiple bonds),
oxidations (benzyl alcohol to benzaldehyde), coupling reactions (Mizoroki-Heck coupling
reaction) and photo degradations (methylene blue or methyl orange). Two excellent review
articles on this topic have recently been published [68,69].
Figure 5 Schematic illustration on application of nano-catalysts (reducing 4-nitrophenol) using
modified cellulose nanocrystals as carriers via different protocols, (A) Mussel-inspired
polydopamine coating [70]; (B) PAMAM dendrimer grafting [71] and (C) Porous melamine-
formaldehyde resin coating [72].

Aside from the most developed approach on using pristine CNCs nanoparticles, other
modification methods have been developed that introduce metal affinity groups or polymers on
the surface of CNCs for the purpose of loading inorganic nanoparticles. Tang et al. reported a
simple and facile approach using mussel-inspired polydopamine to deposit silver nanoparticles
onto cellulose nanocrystals [70]. The catechol-rich polydopamine can be utilized as chelating
groups as well as reducing agents, which has the added benefit of not requiring harsh conditions
or toxic reducing agents to synthesize the nanoparticles (Figure 5A). By using the model reaction
of 4-nitrophenol to 4-aminophenol, they found that the catalytic rate constants were 6 times
faster than pristine silver nanoparticles reduced by dopamine. Chen et al. covalently grafted
poly(amidoamine) (PAMAM) dendrimers onto the surface of oxidized CNCs using peptidic
coupling [71]. The incorporated PAMAM dendrimers provided a well-defined hyper-branched
structure, from which the cavities could be used as nano-reactors to control the size of gold
nanoparticles (Figure 5B). By optimizing the synthesis conditions, the best performance of the
gold nanocatalyst to reduce 4-nitrophenol was reflected by a turnover frequency (TOF) of 5 400
h-1, which was exceptionally greater than other systems that used modified CNCs as supports.
Very recently, Wu et al. extended these ideas to prepare highly porous cellulose nanocrystal
support systems [72]. Mesoporous structures were achieved by coating a layer of melamine-
formaldehyde resin (MF) using polycondensation reactions. The nitrogen enriched resin
provided chelating sites that can improve metal ion binding and the porous structure can confine
the growth of nanoparticles with a uniform size distribution (Figure 5C). After depositing the Pt
or Pd nanoparticles onto MF-CNCs, the hybrid material displayed superior catalytic properties
towards the reduction of 4-nitrophenol with a TOF of up to 3168 h-1. In an effort to combat the
problem of water pollution, rapid and recyclable treatment materials based on cellulose
nanocrystals were prepared by Chen et al. [73]. They deposited superparamagnetic Fe3O4
nanoparticles onto the surface of cellulose nanocrystals, resulting in nanorods that could respond
to external magnetic triggers. In order to improve the stability of CNCs against oxidation, a
uniform silica layer was coated onto the hybrids. This coating layer could increase the onset
decomposition temperature by 60 oC when compared to pristine CNCs. Through further grafting
of -cyclodextrins (-CD), the resulting CNCs@Fe3O4@SiO2 @-CD hybrids exhibited good
adsorption properties towards model pharmaceutical residues in water, such as procaine
hydrochloride and imipramine hydrochloride, with an adsorption capacity of 13 mg/g and 14.8
mg/g, respectively.

Another example of inorganic-CNCs hybrid application is the use of such system as anti-
microbial agent. Silver or zinc oxide nanoparticles (Ag or ZnO) have been widely investigated
for this purpose. Drogat et al. reported an approach to deposit silver nanoparticles onto aldehyde
functionalized cellulose nanocrystals, where the aldehyde groups were used to reduce Ag+ to Ag0
in mild conditions [74]. The silver nanoparticles were in the size range of 20 to 45 nm and the
composite exhibited excellent antimicrobial properties. Shi et al. developed a method to load
silver nanoparticles using polydopamine coated cellulose nanocrystals [75]. The hybrid material
exhibited improved antimicrobial properties towards gram negative Escherichia coli, and gram
positive Bacillus subtilis bacteria with a minimal inhibition concentration (MIC) of 4 g/ml and
8 g/ml, respectively. TEM studies were conducted to understand the improved antimicrobial
properties of the hybrid materials. They attributed this result to the high concentrations of silver
ions released from the silver nanoparticles near the surface of the organisms, which kill the
bacteria. This process was assisted by the adhesive properties of polydopamine towards the cell
membranes. Yao and coworkers prepared ZnO/cellulose nanocrystal hybrids through a one-pot
green synthesis method, in which the modified carboxyl groups could function as stabilizing and
supporting agents for the deposition of ZnO nanoparticles [76]. Carboxyl groups were introduced
during the hydrolysis process using citric and hydrochloric mixed acids (C6H8O7/HCl). The
obtained ZnO nanoparticles with a hexagonal wurtzite structure and the smallest average
diameter of 42.6 nm displayed promising antimicrobial properties against the model bacteria;
Escherichia coli and Staphylococcus aureus.
Other applications could also be found in bioimaging/biosensing, such as quantum dots
(QDs) or carbon dots being the most well-known. Chen et al. demonstrated a one-pot synthesis
to prepare well-dispersed QDs in aqueous solution using oxidized cellulose nanocrystals [77]. A
co-precipitation method was introduced to synthesize the CdS@ZnS core-shell quantum dots.
The carboxylate groups provided the sites for coordinating Cd2+ ions that allow for the in-situ
nucleation and growth of QDs on the CNCs surface. The coating of a ZnS shell was introduced
to reduce the toxicity of hybrid materials as well as enhance the photo-emission intensity. They
further demonstrated that HeLa cells can uptake the composite nanoparticles effectively, with
intense red photoluminescence mostly observed in cytoplasmic regions (Figure 6A). In addition,
they reported a quantum dot-cellulose nanocrystal system that can be used for anti-counterfeiting
applications [78]. Structural colorful films were fabricated via layer-by-layer self-assembly of
oppositely charged CdS quantum dot modified cellulose nanocrystals on a flexible poly(ethylene
terephthalate) substrate. CNCs-COOH@CdS and CNCs-PEI@CdS were utilized in the assembly
process, with the emission peaks around 650 nm (size 4 nm) and 480 nm (size 2.1 nm),
respectively. The fabricated films displayed tunable structural colors from film interference and
adjustable emission colors from the quantum dots, which demonstrates their promising
applications in anti-counterfeiting devices (Figure 6B).
Figure 6 (A) Confocal fluorescence micrographs of HeLa cells showing the uptake of
CNCs/CdS@ZnS quantum dots with red emissions. The cytoskeleton actin was stained green
[77]. (B) Schematic illustration showing the quantum dot-cellulose nanocrystal films for the
purpose of anti-counterfeiting application. The film was fabricated via a layer-by-layer self-
assembly process and the thickness and roughness were characterized by scanning electron
microscopy [78].

3.4 Functional Cellulose nanocrystal-Organic hybrids


In this section, various types of functional groups, small organic molecules or organic
polymers will be discussed. Combining cellulose nanocrystals with both natural and synthetic
polymers expands their use in wastewater treatment, energy storage and biomedical applications.
Sodium alginate, chitosan, 4-vinylpyridine, pyrrole, rodanine and poly (ethyl ethylene phosphate)
are among several organic polymers that have been used to enhance the utility of CNCs. Many
recent reviews exist that discuss the use of CNCs in wastewater treatment, energy and
biomedical applications, and readers can refer to these for a deeper insight into these topics [5, 7,
16]. We have briefly summarized the application of cellulose nanocrystal-organic hybrids in
respect to three major challenges facing the world, such as environmental, energy and
biomedical sectors.

We will first discuss the application of CNCs in wastewater treatment. Environmental pollution
in developing countries is becoming more severe. Water is one of the basic necessities needed to
sustain and support life. CNCs is nontoxic and possesses high surface area, which is a key
characteristic for its application in wastewater treatment, especially as an adsorbent. Luong and
coworkers were the first to report the use of pristine CNCs to adsorb a cationic dye, methylene
blue, with the maximum capacity determined to be 101 mg/g determined from the Langmuir
isotherm [79]. Following that, Batmaz et al. modified the CNCs surface through TEMPO
oxidation to increase the surface charge density, which increased the maximum adsorption
capacity of methylene blue to 769 mg/g [80]. However, it should be noted that using
nanoparticles as adsorbents could be problematic for downstream separation processes (i.e.
coagulation or high speed centrifugation), which limit their large scale applications. Mohammed
et al. incorporated high surface area CNCs into a negatively charged biopolymer (alginate)
matrix [81,82] (Figure 7). A further ionic crosslinking process was introduced to produce macro-
size hydrogel beads. With this composite material, the CNCs adsorption capability was retained,
and the use of hydrogel beads facilitated the easy separation in batch adsorption or in a
continuous flow packed bed system [82]. Also, by impregnating CNCs into the hydrogel matrix,
it can contribute to a higher surface area for adsorption as well as enhanced mechanical
properties. Cranston and coworkers designed crosslinked cellulose nanocrystal aerogels ultilizing
the Schiff-base chemistry. The ultralightweight (5.6 mg/cm3) and porous (99.6%) aerogels
exhibited enhanced mechanical properties and shape recovery capability in water [83]. More
interestingly, they further demonstrated that the aerogels could be used as superabsorbents for
oil-water separations as they have strong affinity towards different solvents. Aside from the
decontamination techniques, such as adsorption or absorption, another promising large scale
process to explore and exploit is flocculation. Cranston and colleagues functionalized the surface
of CNCs with poly(4-vinylpyridine), a pH-responsive polymer that can tune the
hydrophilic/hydrophobic properties of the nanoparticle [84]. At pH values lower than the pka, the
pyridine motifs were protonated, resulting in a positive charge on the nanoparticle surface that
ensured a stable colloidal dispersion. While increasing the pH values, the deprotonation of the
pyridyl groups led to more exposed hydrophobic groups, causing the nanoparticles to flocculate
and precipitate from aqueous dispersions.

Figure 7 (A) Schematic showing the preparation of CNCs-alginate hydrogel beads [81]; (B)
Comparison of the adsorption capability of pure alginate and CNCs-alginate hydrogel beads
towards methylene blue [81]; (C) Schematic diagram of a fixed bed adsorption column process
for removing methylene blue from contaminated water [82].

There is immense value in using nanotechnology to rejuvenate the forest industry by


producing value added wood-derived products, especially in resolving issues, such as energy
storage [5]. Cellulose nanocrystals are ideal candidates for fabricating nanostructures with
enhanced properties, which has already been reflected by assembling them into supercapacitors
or batteries. Wu et al. modified the CNCs surface with carboxylic groups through a TEMPO
mediated oxidization process for such purposes [85]. These carboxylic groups formed strong
hydrogen bonds with pyrrole monomers, which were subsequently polymerized in-situ onto the
CNCs surface with the introduction of an oxidant. The optimized coating condition was achieved
through tuning the mass ratio of pyrrole to CNCs. By using the three-electrode system, they
further demonstrated that the optimized samples exhibited a supercapacitor behavior with a
capacitance of 248 F g-1 at the scan rate of 0.01 V s-1. However, they also encountered cycling
instability as revealed by the rapid capacitance decay over extended cycles due to the
inhomogeneous coating of conductive polymers as well as structural breakdown upon cycling. A
solution was advanced by introducing an amphiphilic polymer (poly(N-vinylpyrrolidone) (PVP))
that acted as a buffer layer for the favorable growth of polypyrrole [86]. This second generation
of conductive CNCs displayed a smooth and uniform coating, which increased the conductivity
from 4.5 to 36.9 S cm-1 (Figure 8A). They further demonstrated that the uniform coating
improved the cycling stabilities of the capacitative systems (Figure 8B). The material exhibited a
capacitance of 338.6 F g-1 at 2 A g-1 with a good capacity retention of 87.3% over 2 000 cycles
under the current density of 10 A g-1. As cellulose nanocrystals are a comparatively (in this
application field) low-cost biomaterial from nature, they could be considered as an abundant
carbon source for pyrolysis processes, and the carbon materials could be designed with versatile
nanostructures for fabricating electrode materials. Pang et al. described a method to prepare
nitrogen and sulfur dual-doped mesoporous carbon materials [87]. They first coated the CNC
surface with a conjugated polymer, polyrhodanine (PR), which provided a heteroatom source for
doping. Then, evaporation induced self-assembly, and the hydrolysis of tetraethyl orthosilicate
(TEOS) was used to fabricate PR-CNCs/silica composite films. Subsequent pyrolysis and
removal of the silica template generated a N, S co-doped mesoporous carbon that could be used
as a cathode material for lithium sulfur batteries (Figure 8C). Through doping N and S into the
carbon lattice, an enhanced chemisorption of lithium polysulfide was achieved, which could
deliver a high capacity of 1 370 mA h g-1 at C/20 and discharge/charge for 1 100 cycles at 2C
rate with a very low capacity fading of 0.052% per cycle. Liu et al. also reported a porous
nitrogen doped carbon derived from cellulose nanocrystals and urea [88]. The product displayed
superior catalytic performance toward oxygen reduction reactions in alkaline media when
compared to a commercial Pt/C catalyst. This is reflected by the comparable electrocatalytic
activity, a better tolerance to the methanol crossover effect and an improved long-term durability.
Figure 8 (A) Morphologies of polypyrrole coated onto CNCs surface are important for their
supercapacitor behavior, TEM images showing the morphology difference by using two different
coating protocols (TEMPO-CNCs and PVP-CNCs) [86]; (B) A comparison of capacitance (Cs)
loss between two generations of polypyrrole coated CNCs; also the cartoon to propose the
mechanism of better performance using PVP layer as a binder [86]; (C) Schematic illustration of
the synthesis of nitrogen/sulfur-doped mesoporous carbon for the utilization of lithium-sulfur
battery electrode [87].

The breadth of applications for nanoparticles in the biomedical sector is enormous,


ranging from drug delivery, antioxidant, antimicrobial, florescence biomarker and tissue
engineering [13,16,89]. For instance, Akhlaghi et al. developed a novel drug delivery system by
grafting the surface of oxidized CNCs with chitosan oligosaccharide (CSos) [90]. Using procaine
hydrochloride as a model drug compound, they determined that the drug loading can approach
up to 14% w/w and the modified system revealed a fast release over around 1 hour at pH 8. Later,
they used isothermal titration calorimetry (ITC) to further understand the interactions between
model drugs (imipramine hydrochloride (IMI) and procaine hydrochloride (PrHy)) and CS os [91].
They found that IMI exhibited a more dominant binding enthalpy compared to PrHy, indicating
that IMI had a higher binding ability toward CNCs carriers. Then, drug selective electrodes were
used to monitor the drug release profiles. They demonstrated that both PrHy and IMI displayed
fast response by tuning the pH values of the system, and the amount of IMI released from CNCs-
CSos was higher than PrHy. Wang et al. used a grafting to method to modify CNCs with
biodegradable polymers for drug delivery [92]. Propargyl-terminated poly(ethyl ethylene
phosphate) (propargyl-PEEP) was synthesized by ring-opening polymerization and subsequently
grafted onto azide-modified CNCs nanoparticles through Cu(I)-catalyzed azidealkyne
cycloaddition (CuAAC) click chemistry (Figure 9). The highly negative charged modified
nanoparticles could be used to load doxorubicin (DOX) via electrostatic interactions. When the
complex was internalized by tumor cells, the acidic environment disrupted the interactions
between the nanoparticle and DOX, leading to a pH-triggered release. They also demonstrated

Figure 9 (A) Schematic illustration of the synthesis pathway of CNCs-g-PEEP via (CuAAC)
click reaction and the formation of DOX-loaded cellulose nanocrystals; (B) In vitro drug
release profile of DOX-loaded CNCs-g-PEEP at pH 5.0 and 7.4 [92].

that the CNCs-based carrier showed good biocompatibility towards both HeLa cells and L929
cells, with average cell viabilities above 90%, even at high concentrations. Aside from specific
drugs for therapy or drug delivery, an antioxidant Vitamin C was formulated with modified
CNCs. CNCs-CSos was used to complex Vitamin C (VC) via a further ionic gelation through
triphosphate crosslinking [93]. The encapsulation efficiency of VC into a complex approached
up to 71.6% and 91% at pH 3 and 5, respectively. The system displayed a sustained release of
active components (Vitamin C) up to 20 days and showed promising radical scavenging and
antioxidant activity. Apart from drug delivery systems, Tang et al. reported on a stable
antimicrobial system that is based on a polyrhodanine coated cellulose nanocrystal [93]. The
coating procedure was easy and feasible using an in-situ oxidation polymerization procedure.
The optimized sample exhibited promising antimicrobial properties towards both E. coli (Gram
negative) and B. subtilis (Gram positive). Also, their toxicity towards HeLa cells were
demonstrated to be low within the concentrations that displayed favourable antimicrobial
properties. In addition to antimicrobial properties, they found that the modified nanoparticle
displayed a pH-dependent optical characteristic [94]. The color of the dispersion gradually
changed from pale red to blue violet when the pH was increased from 2.04 to 12.04. The redox
reversibility of the hybrid nanomaterial in response to pH was retained when transformed into
different geometries, such as 1D printable inks, 2D flat films and membranes, and 3D hydrogel
beads (Figure 10). They anticipated that the modified material could be processed into vacuum
packaging materials, with the effective antimicrobial properties providing long-term protection
for fresh meat or vegetables. The pH-dependent optical property may also be used to assess the
quality of food as the metabolism of microorganisms may lead to pH changes.

Figure 10 (A) Summary of the UV-Vis spectra of a 0.01 wt% CNCs@PR dispersion at different
pH values; (B) Sigmoidal plots of maximum absorption wavelength and absorbance@peak
versus pH. (C) Schematic illustrating the redox reversibility of the hybrid nanomaterial in
response to pH was retained when transformed into different geometries [94].
The purpose of CNCs as potential carriers for cell uptake originates from its size,
hydrophilic nature and chemical composition. Modifying the surface with fluorophores can
generate nanoscale biomarkers that could be used for the localization and quantification of
nanoparticles within the cell. Roman and coworkers were the first report the fluorescence
characteristic using modified CNCs by conjugating the surface with fluorescein-5-
isothiocyanate (FITC) [95]. They used spectrofluorometry, fluorescence microscopy and flow
cytometry to study the in-vivo interaction of modified CNCs within cells. Luong and co-workers
compared the application of negatively charged FITC and positively charged rhodamine B
isothiocyanate (RBITC) modified cellulose nanocrystals in cell internalization [96]. They found
that the negatively charged FITC-CNCs could not be internalized by the cells due to the strong
electrostatic repulsions between the anionic CNCs and anionic cellular membrane. However,
RBITC-CNCs exhibited low cytotoxicity and excellent membrane permeability in various cell
lines, which suggests that the surface charge and conjugated elements are essential factors when
functionalizing the nanoparticle as carriers to cells for bioimaging and drug delivery.

4. Conclusion and perspectives


The consumption of fossil fuels to produce energy and petrochemicals over the last 50
years has resulted in the rapid accumulation of green-house gases that is impacting the livelihood
of many communities around the globe. There is thus a compelling motivation to seek alternative
sources of energy and raw materials not derived from fossil fuels, but from renewable resources.
Research on the functionalization of CNCs is at its infancy, and the development of scalable
synthetic protocols for the chemical modification of CNCs, and fundamental understanding of
this new class of nanomaterial will create new opportunities and markets. We believe sustainable
nanomaterials will be extremely important in addressing two critical issues confronting our
world, namely energy and the environment. The utilization of cellulose nanocrystals will reduce
our dependence on conventional carbon sources (e.g. crude oil) and cellulosic materials are
excellent carbon sinks for capturing carbon dioxide.
This featured article outlines research activities that explore and exploit sustainable
nanomaterials for a variety of applications. We summarize the most recent developments on the
modification and utilization of functionalized cellulose nanocrystals. The modifications have
expanded the use of this remarkable nanomaterial to applications, such as emulsion stabilizers,
anti-microbial agents, controlled delivery systems, sustainable catalysts, templating agent for
mesoporous nanostructures, and various energy and electronic systems. Studies on the
modification and functionalization of cellulose nanocrystals have generated important
fundamental understanding necessary for designing novel functional systems for potential
applications.
Around the globe, several large scale production facilities to manufacture pristine
cellulose nanocrystals were commissioned. Some of these facilities include the following: (a)
Celluforce-1 000 kg/day (Canada), (b) American Process-500 kg/day (USA), (c) Holmen-100
kg/day (Sweden), (d) Alberta Innovates-20 kg/day (Canada), (e) US Forest Products Lab-10
kg/day (USA), (f) Blue Goose Biorefineries-10 kg/day (Canada), and (g) India Council for
Agricuture Research-10 kg/day (India). In addition, there are other smaller facilities in countries,
such as China, Brazil etc. that are not listed here. Consumers can now purchase significant
quantities of CNCs for field trials and large scale product formulations and evaluation. We are at
the stage where we will see the adoption of CNCs in product formulations in applications, such
as coatings, pulp and paper, consumer and personal care systems, wastewater treatment and
biomedical engineering. The future for CNCs is bright, and the research activities in academic
and industrial laboratories will generate many new discoveries and applications. This feature
article provides a glimpse amongst other of one such activity in the Laboratory of Functional
Colloids and Sustainable Nanomaterials at the University of Waterloo, Canada.

Acknowledgements
The research funding from CelluForce and AboraNano facilitated the research on CNCs. K. C.
Tam wishes to acknowledge funding from CFI and NSERC.
References

[1] I. Linkov, M.E. Bates, L.J. Canis, T.P. Seager, J.M. Keisler, A decision-directed approach
for prioritizing research into the impact of nanomaterials on the environment and human
health, Nat. Nanotechnol. 6 (2011) 784787. doi:10.1038/nnano.2011.163.
[2] V.L. Colvin, The potential environmental impact of engineered nanomaterials, Nat.
Biotechnol. 21 (2003) 11661170. doi:10.1038/nbt875.
[3] J. He, K. Sill, H. Xiang, Self-directed self-assembly of nanoparticle / copolymer mixtures,
Nature. 404 (2005) 5559. doi:10.1038/nature03361.1.
[4] G. Lu, S. Li, Z. Guo, O.K. Farha, B.G. Hauser, X. Qi, et al., Imparting functionality to a
metalorganic framework material by controlled nanoparticle encapsulation, Nat. Chem. 4
(2012) 310316. doi:10.1038/nchem.1272.
[5] H. Zhu, W. Luo, P.N. Ciesielski, Z. Fang, J.Y. Zhu, G. Henriksson, et al., Wood-Derived
Materials for Green Electronics, Biological Devices, and Energy Applications, Chem.
Rev. 116 (2016) 93059374. doi:10.1021/acs.chemrev.6b00225.
[6] D. Klemm, B. Heublein, H.-P. Fink, A. Bohn, Cellulose: Fascinating Biopolymer and
Sustainable Raw Material, Angew. Chem. Int. Ed. 44 (2005) 33583393.
doi:10.1002/anie.200460587.
[7] D. Klemm, F. Kramer, S. Moritz, T. Lindstrm, M. Ankerfors, D. Gray, et al.,
Nanocelluloses: A New Family of Nature-Based Materials, Angew. Chem. Int. Ed. 50
(2011) 54385466. doi:10.1002/anie.201001273.
[8] J. Cai, L. Zhang, Rapid dissolution of cellulose in LiOH/urea and NaOH/urea aqueous
solutions, Macromol. Biosci. 5 (2005) 539548. doi:10.1002/mabi.200400222.
[9] J. Cai, L. Zhang, Unique gelation behavior of cellulose in NaOH/urea aqueous solution,
Biomacromolecules. 7 (2006) 183189. doi:10.1021/bm0505585.
[10] C. Chang, L. Zhang, Cellulose-based hydrogels: Present status and application prospects,
Carbohydr. Polym. 84 (2011) 4053. doi:10.1016/j.carbpol.2010.12.023.
[11] R.P. Swatloski, S.K. Spear, J.D. Holbrey, R.D. Rogers, Dissolution of cellulose with
ionic liquids, J. Am. Chem. Soc. 124 (2002) 49744975. doi:ja025790m.
[12] H. Wang, G. Gurau, R.D. Rogers, Ionic liquid processing of cellulose, Chem. Soc. Rev.
41 (2012) 1519. doi:10.1039/c2cs15311d.
[13] T. Abitbol, A. Rivkin, Y. Cao, Y. Nevo, E. Abraham, T. Ben-Shalom, et al.,
Nanocellulose, a tiny fiber with huge applications, Curr. Opin. Biotechnol. 39 (2016) 76
88. doi:10.1016/j.copbio.2016.01.002.
[14] Y. Habibi, Key advances in the chemical modification of nanocelluloses, Chem. Soc.
Rev. 43 (2014) 15191542. doi:10.1039/C3CS60204D.
[15] C. Salas, T. Nypel, C. Rodriguez-Abreu, C. Carrillo, O.J. Rojas, Nanocellulose
properties and applications in colloids and interfaces, Curr. Opin. Colloid Interface Sci. 19
(2014) 383396. doi:10.1016/j.cocis.2014.10.003.
[16] N. Lin, A. Dufresne, Nanocellulose in biomedicine: Current status and future prospect,
Eur. Polym. J. 59 (2014) 302325. doi:10.1016/j.eurpolymj.2014.07.025.
[17] Y. Habibi, L.A. Lucia, O.J. Rojas, Cellulose nanocrystals: Chemistry, self-assembly, and
applications, Chem. Rev. 110 (2010) 34793500. doi:10.1021/cr900339w.
[18] B.L. Peng, N. Dhar, H.L. Liu, K.C. Tam, Chemistry and applications of nanocrystalline
cellulose and its derivatives: A nanotechnology perspective, Can. J. Chem. Eng. 89 (2011)
11911206. doi:10.1002/cjce.20554.
[19] S.J. Eichhorn, Cellulose nanowhiskers: promising materials for advanced applications,
Soft Matter. 7 (2011) 303315. doi:10.1039/C0SM00142B.
[20] J.P.F. Lagerwall, C. Schtz, M. Salajkova, J. Noh, J. Hyun Park, G. Scalia, et al.,
Cellulose nanocrystal-based materials: from liquid crystal self-assembly and glass
formation to multifunctional thin films, NPG Asia Mater. 6 (2014) e80.
doi:10.1038/am.2013.69.
[21] C. Moreau, A. Villares, I. Capron, B. Cathala, Tuning supramolecular interactions of
cellulose nanocrystals to design innovative functional materials, Ind. Crops Prod. 93
(2016) 96-107. doi:10.1016/j.indcrop.2016.02.028.
[22] M. Rayner, D. Marku, M. Eriksson, M. Sj, P. Dejmek, M. Wahlgren, Biomass-based
particles for the formulation of Pickering type emulsions in food and topical applications,
Colloids Surfaces A Physicochem. Eng. Asp. 458 (2014) 4862.
doi:10.1016/j.colsurfa.2014.03.053.
[23] J. Tang, P.J. Quinlan, K.C. Tam, Stimuli-responsive Pickering emulsions: recent
advances and potential applications, Soft Matter. 11 (2015) 35123529.
doi:10.1039/C5SM00247H.
[24] I. Kalashnikova, H. Bizot, B. Cathala, I. Capron, New Pickering Emulsions Stabilized by
Bacterial Cellulose Nanocrystals, Langmuir. 27 (2011) 74717479.
doi:10.1021/la200971f.
[25] I. Kalashnikova, H. Bizot, B. Cathala, I. Capron, Modulation of Cellulose Nanocrystals
Amphiphilic Properties to Stabilize Oil/Water Interface, Biomacromolecules. 13 (2012)
267275. doi:10.1021/bm201599j.
[26] I. Kalashnikova, H. Bizot, P. Bertoncini, B. Cathala, I. Capron, Cellulosic nanorods of
various aspect ratios for oil in water Pickering emulsions, Soft Matter. 9 (2013) 952959.
doi:10.1039/C2SM26472B.
[27] F. Cherhal, F. Cousin, I. Capron, Structural Description of the Interface of Pickering
Emulsions Stabilized by Cellulose Nanocrystals, Biomacromolecules. 17 (2016) 496502.
doi:10.1021/acs.biomac.5b01413.
[28] I. Capron, B. Cathala, Surfactant-free high internal phase emulsions stabilized by
cellulose nanocrystals, Biomacromolecules. 14 (2013) 291296. doi:10.1021/bm301871k.
[29] K.R. Peddireddy, T. Nicolai, L. Benyahia, I. Capron, Stabilization of Water-in-Water
Emulsions by Nanorods, ACS Macro Lett. 5 (2016) 283286.
doi:10.1021/acsmacrolett.5b00953.
[30] Z. Hu, S. Ballinger, R. Pelton, E.D. Cranston, Surfactant-enhanced cellulose nanocrystal
Pickering emulsions, J. Colloid Interface Sci. 439 (2015) 139148.
doi:10.1016/j.jcis.2014.10.034.
[31] D. Saidane, E. Perrin, F. Cherhal, F. Guellec, I. Capron, Some modification of cellulose
nanocrystals for functional Pickering emulsions, Philos. Trans. R. Soc. A Math. Phys.
Eng. Sci. 374 (2016) 20150139. doi:10.1098/rsta.2015.0139.
[32] M. Visanko, H. Liimatainen, J.A. Sirvi, J.P. Heiskanen, J. Niinimki, O. Hormi,
Amphiphilic Cellulose Nanocrystals from Acid-Free Oxidative Treatment:
Physicochemical Characteristics and Use as an OilWater Stabilizer, Biomacromolecules.
15 (2014) 27692775. doi:10.1021/bm500628g.
[33] G. Sbe, F. Ham-Pichavant, G. Pecastaings, Dispersibility and Emulsion-Stabilizing
Effect of Cellulose Nanowhiskers Esterified by Vinyl Acetate and Vinyl Cinnamate,
Biomacromolecules. 14 (2013) 29372944. doi:10.1021/bm400854n.
[34] A.G. Cunha, J.B. Mougel, B. Cathala, L.A. Berglund, I. Capron, Preparation of double
pickering emulsions stabilized by chemically tailored nanocelluloses, Langmuir. 30
(2014) 93279335. doi:10.1021/la5017577.
[35] Z. Hu, T. Patten, R. Pelton, E.D. Cranston, Synergistic Stabilization of Emulsions and
Emulsion Gels with Water-Soluble Polymers and Cellulose Nanocrystals, ACS Sustain.
Chem. Eng. 3 (2015) 10231031. doi:10.1021/acssuschemeng.5b00194.
[36] J.O. Zoppe, R.A. Venditti, O.J. Rojas, Pickering emulsions stabilized by cellulose
nanocrystals grafted with thermo-responsive polymer brushes, J. Colloid Interface Sci.
369 (2012) 202209. doi:10.1016/j.jcis.2011.12.011.
[37] J. Tang, M.F.X. Lee, W. Zhang, B. Zhao, R.M. Berry, K.C. Tam, Dual responsive
pickering emulsion stabilized by poly[2-(dimethylamino) ethyl methacrylate] grafted
cellulose nanocrystals, Biomacromolecules. 15 (2014) 30523060.
doi:10.1021/bm500663w.
[38] J. Tang, R.M. Berry, K.C. Tam, Stimuli-Responsive Cellulose Nanocrystals for
Surfactant-Free Oil Harvesting, Biomacromolecules. 17 (2016) 17481756.
doi:10.1021/acs.biomac.6b00144.
[39] M. Marquis, V. Alix, I. Capron, S. Cuenot, A. Zykwinska, Microfluidic Encapsulation of
Pickering Oil Microdroplets into Alginate Microgels for Lipophilic Compound Delivery,
ACS Biomater. Sci. Eng. 2 (2016) 535543. doi:10.1021/acsbiomaterials.5b00522.
[40] Z. Hu, H.S. Marway, H. Kasem, R. Pelton, E.D. Cranston, Dried and Redispersible
Cellulose Nanocrystal Pickering Emulsions, ACS Macro Lett. 5 (2016) 185189.
doi:10.1021/acsmacrolett.5b00919.
[41] K.E. Shopsowitz, H. Qi, W.Y. Hamad, M.J. MacLachlan, Free-standing mesoporous
silica films with tunable chiral nematic structures, Nature. 468 (2010) 422425.
doi:10.1038/nature09540.
[42] J.A. Kelly, M. Giese, K.E. Shopsowitz, W.Y. Hamad, M.J. MacLachlan, The
development of chiral nematic mesoporous materials, Acc. Chem. Res. 47 (2014) 1088
1096. doi:10.1021/ar400243m.
[43] M. Giese, L.K. Blusch, M.K. Khan, M.J. MacLachlan, Functional Materials from
Cellulose-Derived Liquid-Crystal Templates, Angew. Chem. Int. Ed. 54 (2015) 2888
2910. doi:10.1002/anie.201407141.
[44] T.D. Nguyen, J.A. Kelly, W.Y. Hamad, M.J. Maclachlan, Magnesiothermic reduction of
thin films: Towards semiconducting chiral nematic mesoporous silicon carbide and silicon
structures, Adv. Funct. Mater. 25 (2015) 21752181. doi:10.1002/adfm.201404304.
[45] P. X. Wang, W.Y. Hamad, M.J. MacLachlan, Polymer and Mesoporous Silica
Microspheres with Chiral Nematic Order from Cellulose Nanocrystals, Angew. Chem. Int.
Ed. 55 (2016) 1246012464. doi:10.1002/anie.201606283.
[46] H. Qi, K.E. Shopsowitz, W.Y. Hamad, M.J. MacLachlan, Chiral nematic assemblies of
silver nanoparticles in mesoporous silica thin films, J. Am. Chem. Soc. 133 (2011) 3728
3731. doi:10.1021/ja110369d.
[47] K.E. Shopsowitz, W.Y. Hamad, M.J. MacLachlan, Chiral Nematic Mesoporous Carbon
Derived From Nanocrystalline Cellulose, Angew. Chem. Int. Ed. 50 (2011) 1099110995.
doi:10.1002/anie.201105479.
[48] K.E. Shopsowitz, A. Stahl, W.Y. Hamad, M.J. MacLachlan, Hard Templating of
Nanocrystalline Titanium Dioxide with Chiral Nematic Ordering, Angew. Chem. Int. Ed.
51 (2012) 68866890. doi:10.1002/anie.201201113.
[49] K.E. Shopsowitz, W.Y. Hamad, M.J. MacLachlan, Flexible and Iridescent Chiral
Nematic Mesoporous Organosilica Films, J. Am. Chem. Soc. 134 (2012) 867870.
doi:10.1021/ja210355v.
[50] J.A. Kelly, K.E. Shopsowitz, J.M. Ahn, W.Y. Hamad, M.J. MacLachlan, Chiral Nematic
Stained Glass: Controlling the Optical Properties of Nanocrystalline Cellulose-Templated
Materials, Langmuir. 28 (2012) 1725617262. doi:10.1021/la3041902.
[51] J.A. Kelly, A.M. Shukaliak, C.C.Y. Cheung, K.E. Shopsowitz, W.Y. Hamad, M.J.
MacLachlan, Responsive Photonic Hydrogels Based on Nanocrystalline Cellulose,
Angew. Chem. Int. Ed. 52 (2013) 89128916. doi:10.1002/anie.201302687.
[52] M.K. Khan, M. Giese, M. Yu, J.A. Kelly, W.Y. Hamad, M.J. MacLachlan, Flexible
Mesoporous Photonic Resins with Tunable Chiral Nematic Structures, Angew. Chem. Int.
Ed. 52 (2013) 89218924. doi:10.1002/anie.201303829.
[53] J.A. Kelly, M. Yu, W.Y. Hamad, M.J. MacLachlan, Large, Crack-Free Freestanding
Films with Chiral Nematic Structures, Adv. Opt. Mater. 1 (2013) 295299.
doi:10.1002/adom.201300015.
[54] K.E. Shopsowitz, J.A. Kelly, W.Y. Hamad, M.J. MacLachlan, Biopolymer templated
glass with a twist: Controlling the chirality, porosity, and photonic properties of silica with
cellulose nanocrystals, Adv. Funct. Mater. 24 (2014) 327338.
doi:10.1002/adfm.201301737.
[55] M.K. Khan, W.Y. Hamad, M.J. MacLachlan, Tunable Mesoporous Bilayer Photonic
Resins with Chiral Nematic Structures and Actuator Properties, Adv. Mater. 26 (2014)
23232328. doi:10.1002/adma.201304966.
[56] T.-D. Nguyen, W.Y. Hamad, M.J. MacLachlan, CdS Quantum Dots Encapsulated in
Chiral Nematic Mesoporous Silica: New Iridescent and Luminescent Materials, Adv.
Funct. Mater. 24 (2014) 777783. doi:10.1002/adfm.201302521.
[57] M. Giese, L.K. Blusch, M.K. Khan, W.Y. Hamad, M.J. MacLachlan, Responsive
Mesoporous Photonic Cellulose Films by Supramolecular Cotemplating, Angew. Chem.
Int. Ed. 53 (2014) 88808884. doi:10.1002/anie.201402214.
[58] J.-H. Zhang, S.-M. Xie, M. Zhang, M. Zi, P.-G. He, L.-M. Yuan, Novel Inorganic
Mesoporous Material with Chiral Nematic Structure Derived from Nanocrystalline
Cellulose for High-Resolution Gas Chromatographic Separations, Anal. Chem. 86 (2014)
95959602. doi:10.1021/ac502073g.
[59] A. Ivanova, D. Fattakhova-Rohlfing, B.E. Kayaalp, J. Rathousk, T. Bein, Tailoring the
morphology of mesoporous titania thin films through biotemplating with nanocrystalline
cellulose, J. Am. Chem. Soc. 136 (2014) 59305937. doi:10.1021/ja411292u.
[60] P.-X. Wang, V.M. Zamarion, W.Y. Hamad, M.J. MacLachlan, Hard-templating of
Prussian blue analogues in mesoporous silica and organosilica, Dalt. Trans. 44 (2015)
1472414731. doi:10.1039/C5DT02213D.
[61] M.K. Khan, A. Bsoul, K. Walus, W.Y. Hamad, M.J. MacLachlan, Photonic Patterns
Printed in Chiral Nematic Mesoporous Resins, Angew. Chem. Int. Ed. 54 (2015) 4304
4308. doi:10.1002/anie.201410411.
[62] M. Giese, L.K. Blusch, M. Schlesinger, G.R. Meseck, W.Y. Hamad, M. Arjmand, et al.,
Magnetic Mesoporous Photonic Cellulose Films, Langmuir. 32 (2016) 93299334.
doi:10.1021/acs.langmuir.6b02974.
[63] Y. Shin, I.T. Bae, B.W. Arey, G.J. Exarhos, Simple preparation and stabilization of nickel
nanocrystals on cellulose nanocrystal, Mater. Lett. 61 (2007) 32153217.
doi:10.1016/j.matlet.2006.11.036.
[64] M. Rezayat, R.K. Blundell, J.E. Camp, D. a. Walsh, W. Thielemans, Green One-Step
Synthesis of Catalytically Active Palladium Nanoparticles Supported on Cellulose
Nanocrystals, ACS Sustain. Chem. Eng. 2 (2014) 12411250. doi:10.1021/sc500079q.
[65] X. Wu, C. Lu, W. Zhang, G. Yuan, R. Xiong, X. Zhang, A novel reagentless approach for
synthesizing cellulose nanocrystal-supported palladium nanoparticles with enhanced
catalytic performance, J. Mater. Chem. A. 1 (2013) 8645. doi:10.1039/c3ta11236e.
[66] Y. Zhou, E.-Y. Ding, W.-D. Li, Synthesis of TiO2 nanocubes induced by cellulose
nanocrystal (CNC) at low temperature, Mater. Lett. 61 (2007) 50505052.
doi:10.1016/j.matlet.2007.04.001.
[67] Y. Shin, I.T. Bae, B.W. Arey, G.J. Exarhos, Facile stabilization of gold-silver alloy
nanoparticles on cellulose nanocrystal, J. Phys. Chem. C. 112 (2008) 48444848.
doi:10.1021/jp710767w.
[68] M. Kaushik, A. Moores, Review: nanocelluloses as versatile supports for metal
nanoparticles and their applications in catalysis, Green Chem. 18 (2016) 622637.
doi:10.1039/C5GC02500A.
[69] E. Lam, K.B. Male, J.H. Chong, A.C.W. Leung, J.H.T. Luong, Applications of
functionalized and nanoparticle-modified nanocrystalline cellulose, Trends Biotechnol. 30
(2012) 283290. doi:10.1016/j.tibtech.2012.02.001.
[70] J. Tang, Z. Shi, R.M. Berry, K.C. Tam, Mussel-inspired green metallization of silver
nanoparticles on cellulose nanocrystals and their enhanced catalytic reduction of 4-
nitrophenol in the presence of-cyclodextrin, Ind. Eng. Chem. Res. 54 (2015) 32993308.
doi:10.1021/acs.iecr.5b00177.
[71] L. Chen, W. Cao, P.J. Quinlan, R.M. Berry, K.C. Tam, Sustainable Catalysts from Gold-
Loaded Polyamidoamine Dendrimer-Cellulose Nanocrystals, ACS Sustain. Chem. Eng. 3
(2015) 978985. doi:10.1021/acssuschemeng.5b00110.
[72] X. Wu, Z. Shi, S. Fu, J. Chen, R.M. Berry, K.C. Tam, Strategy for Synthesizing Porous
Cellulose Nanocrystal Supported Metal Nanocatalysts, ACS Sustain. Chem. Eng. (2016)
acssuschemeng.6b00551. doi:10.1021/acssuschemeng.6b00551.
[73] L. Chen, R.M. Berry, K.C. Tam, Synthesis of -Cyclodextrin-modified cellulose
nanocrystals (CNCs)@Fe3O4@SiO2 superparamagnetic nanorods, ACS Sustain. Chem.
Eng. 2 (2014) 951958. doi:10.1021/sc400540f.
[74] N. Drogat, R. Granet, V. Sol, A. Memmi, N. Saad, C. Klein Koerkamp, et al.,
Antimicrobial silver nanoparticles generated on cellulose nanocrystals, J. Nanoparticle
Res. 13 (2011) 15571562. doi:10.1007/s11051-010-9995-1.
[75] Z. Shi, J. Tang, L. Chen, C. Yan, S. Tanvir, W.A. Anderson, et al., Enhanced colloidal
stability and antibacterial performance of silver nanoparticles/cellulose nanocrystal
hybrids, J. Mater. Chem. B. 3 (2015) 603611. doi:10.1039/C4TB01647E.
[76] H.Y. Yu, G.Y. Chen, Y.B. Wang, J.M. Yao, A facile one-pot route for preparing cellulose
nanocrystal/zinc oxide nanohybrids with high antibacterial and photocatalytic activity,
Cellulose. 22 (2015) 261273. doi:10.1007/s10570-014-0491-0.
[77] L. Chen, Y. Liu, C. Lai, R.M. Berry, K.C. Tam, Aqueous synthesis and biostabilization of
CdS@ZnS quantum dots for bioimaging applications, Mater. Res. Express. 2 (2015)
105401. doi:10.1088/2053-1591/2/10/105401.
[78] L. Chen, C. Lai, R. Marchewka, R.M. Berry, K.C. Tam, Use of CdS quantum dot-
functionalized cellulose nanocrystal films for anti-counterfeiting applications, Nanoscale.
8 (2016) 1328813296. doi:10.1039/C6NR03039D.
[79] X. He, K.B. Male, P.N. Nesterenko, D. Brabazon, B. Paull, J.H.T. Luong, Adsorption and
Desorption of Methylene Blue on Porous Carbon Monoliths and Nanocrystalline
Cellulose, ACS Appl. Mater. Interfaces. 5 (2013) 87968804. doi:10.1021/am403222u.
[80] R. Batmaz, N. Mohammed, M. Zaman, G. Minhas, R.M. Berry, K.C. Tam, Cellulose
nanocrystals as promising adsorbents for the removal of cationic dyes, Cellulose. 21
(2014) 16551665. doi:10.1007/s10570-014-0168-8.
[81] N. Mohammed, N. Grishkewich, R.M. Berry, K.C. Tam, Cellulose nanocrystalalginate
hydrogel beads as novel adsorbents for organic dyes in aqueous solutions, Cellulose. 22
(2015) 37253738. doi:10.1007/s10570-015-0747-3.
[82] N. Mohammed, N. Grishkewich, H.A. Waeijen, R.M. Berry, K.C. Tam, Continuous flow
adsorption of methylene blue by cellulose nanocrystal-alginate hydrogel beads in fixed
bed columns, Carbohydr. Polym. 136 (2016) 11941202.
doi:10.1016/j.carbpol.2015.09.099.
[83] X. Yang, E.D. Cranston, Chemically Cross-Linked Cellulose Nanocrystal Aerogels with
Shape Recovery and Superabsorbent Properties, Chem. Mater. 26 (2014) 60166025.
doi:10.1021/cm502873c.
[84] K.H.M. Kan, J. Li, K. Wijesekera, E.D. Cranston, Polymer-Grafted Cellulose
Nanocrystals as pH-Responsive Reversible Flocculants, Biomacromolecules. 14 (2013)
31303139. doi:10.1021/bm400752k.
[85] X. Wu, V.L. Chabot, B.K. Kim, A. Yu, R.M. Berry, K.C. Tam, Cost-effective and
Scalable Chemical Synthesis of Conductive Cellulose Nanocrystals for High-performance
Supercapacitors, Electrochim. Acta. 138 (2014) 139147.
doi:10.1016/j.electacta.2014.06.089.
[86] X. Wu, J. Tang, Y. Duan, A. Yu, R.M. Berry, K.C. Tam, Conductive cellulose
nanocrystals with high cycling stability for supercapacitor applications, J. Mater. Chem.
A. 2 (2014) 1926819274. doi:10.1039/C4TA04929B.
[87] Q. Pang, J. Tang, H. Huang, X. Liang, C. Hart, K.C. Tam, et al., A Nitrogen and Sulfur
Dual-Doped Carbon Derived from Polyrhodanine@Cellulose for Advanced Lithium-
Sulfur Batteries, Adv. Mater. 27 (2015) 60216028. doi:10.1002/adma.201502467.
[88] Q. Liu, C. Chen, F. Pan, J. Zhang, Highly efficient oxygen reduction on porous nitrogen-
doped nanocarbons directly synthesized from cellulose nanocrystals and urea,
Electrochim. Acta. 170 (2015) 234241. doi:10.1016/j.electacta.2015.04.094.
[89] R. Sunasee, U.D. Hemraz, K. Ckless, Cellulose nanocrystals: a versatile nanoplatform for
emerging biomedical applications, Expert Opin. Drug Deliv. 13 (2016) 12431256.
doi:10.1080/17425247.2016.1182491.
[90] S.P. Akhlaghi, R.C. Berry, K.C. Tam, Surface modification of cellulose nanocrystal with
chitosan oligosaccharide for drug delivery applications, Cellulose. 20 (2013) 17471764.
doi:10.1007/s10570-013-9954-y.
[91] S.P. Akhlaghi, D. Tiong, R.M. Berry, K.C. Tam, Comparative release studies of two
cationic model drugs from different cellulose nanocrystal derivatives, Eur. J. Pharm.
Biopharm. 88 (2014) 207215. doi:10.1016/j.ejpb.2014.04.012.
[92] H. Wang, J. He, M. Zhang, K.C. Tam, P. Ni, A new pathway towards polymer modified
cellulose nanocrystals via a grafting onto process for drug delivery, Polym. Chem. 6
(2015) 42064209. doi:10.1039/c5py00466g.
[93] J. Tang, Y. Song, S. Tanvir, W.A. Anderson, R.M. Berry, K.C. Tam, Polyrhodanine
Coated Cellulose Nanocrystals: A Sustainable Antimicrobial Agent, ACS Sustain. Chem.
Eng. 3 (2015) 18011809. doi:10.1021/acssuschemeng.5b00380.
[94] J. Tang, Y. Song, R.M. Berry, K.C. Tam, Polyrhodanine coated cellulose nanocrystals as
optical pH indicators, RSC Adv. 4 (2014) 6024960252. doi:10.1039/C4RA09043H.
[95] S. Dong, M. Roman, Fluorescently Labeled Cellulose Nanocrystals for Bioimaging
Applications, J. Am. Chem. Soc. 129 (2007) 1381013811. doi:10.1021/ja076196l.
[96] K.A. Mahmoud, J.A. Mena, K.B. Male, S. Hrapovic, A. Kamen, J.H.T. Luong, Effect of
Surface Charge on the Cellular Uptake and Cytotoxicity of Fluorescent Labeled Cellulose
Nanocrystals, ACS Appl. Mater. Interfaces. 2 (2010) 29242932.
doi:10.1021/am1006222.
Graphical Abstract

Anda mungkin juga menyukai