Nanocrystalline Cellulose

download Nanocrystalline Cellulose

of 22

Transcript of Nanocrystalline Cellulose

  • 8/14/2019 Nanocrystalline Cellulose

    1/22

    Cellulose Nanocrystals: Chemistry, Self-Assembly, and Applications

    Youssef Habibi, Lucian A. Lucia,*, and Orlando J. Rojas,

    Department of Forest Biomaterials, North Carolina State University, Box 8005, Raleigh, North Carolina 27695-8005, and Department of ForestProducts Technology, Faculty of Chemistry and Materials Sciences, Helsinki University of Technology, P.O. Box 3320,

    FIN-02015 TKK, Espoo, Finland

    Received October 12, 2009

    Contents

    1. Introduction and State of the Art A2. Structure and Morphology of Celluloses B3. Cellulose Nanocrystals E

    3.1. Preparation of Cellulose Nanocrystals E3.2. Morphology and Dimensions of Cellulose

    NanocrystalsG

    4. Chemical Modifications of Cellulose Nanocrystals H

    4.1. Noncovalent Surface Chemical Modifications H4.2. TEMPO-Mediated Oxidation I4.3. Cationization I4.4. Esterification, Silylation and Other Surface

    Chemical ModificationsI

    4.5. Polymer Grafting J5. Self-Assembly and -Organization of Cellulose

    NanocrystalsK

    5.1. Self-Assembly and -Organization of CNs inAqueous Medium

    L

    5.2. Self-Assembly and -Organization of CNs inOrganic Medium

    N

    5.3. Self-Assembly and -Organization of CNs under

    External Fields

    N

    5.4. Self-Assembly and -Organization of CNs inThin Solid Films

    N

    6. Applications of Cellulose Nanocrystals inNanocomposite Materials

    O

    6.1. Nanocomposite Processing O6.1.1. Casting-Evaporation Processing O6.1.2. Sol-Gel Processing P6.1.3. Other Processing Methods P

    6.2. Mechanical Properties of CN-BasedComposites

    Q

    6.2.1. Morphology and Dimensions of CNs R6.2.2. Processing Method R

    6.2.3. Interfacial Interactions R6.3. Thermal Properties of CN-Based Composites R

    7. Conclusions and Outlook S8. Acknowledgments S9. References S

    1. Introduction and State of the Art

    Cellulose constitutes the most abundant renewable polymerresource available today. As a chemical raw material, it isgenerally well-known that it has been used in the form of

    fibers or derivatives for nearly 150 years for a wide spectrumof products and materials in daily life. What has not beenknown until relatively recently is that when cellulose fibersare subjected to acid hydrolysis, the fibers yield defect-free,rod-like crystalline residues. Cellulose nanocrystals (CNs)have garnered in the materials community a tremendous levelof attention that does not appear to be relenting. Thesebiopolymeric assemblies warrant such attention not onlybecause of their unsurpassed quintessential physical andchemical properties (as will become evident in the review)but also because of their inherent renewability and sustain-ability in addition to their abundance. They have been thesubject of a wide array of research efforts as reinforcingagents in nanocomposites due to their low cost, availability,renewability, light weight, nanoscale dimension, and uniquemorphology. Indeed, CNs are the fundamental constitutivepolymeric motifs of macroscopic cellulosic-based fiberswhose sheer volume dwarfs any known natural or synthetic

    biomaterial. Biopolymers such as cellulose and lignin and

    North Carolina State University.

    Helsinki University of Technology.

    Dr. Youssef Habibi is a research assistant professor at the Departmentof Forest Biomaterials at North Carolina State University. He receivedhis Ph.D. in 2004 in organic chemistry from Joseph Fourier University(Grenoble, France) jointly with CERMAV (Centre de Recherche sur lesMacromolecules Vegetales) and Cadi Ayyad University (Marrakesh,Morocco). During his Ph.D., he worked on the structural characterizationof cell wall polysaccharides and also performed surface chemicalmodification, mainly TEMPO-mediated oxidation, of crystalline polysac-charides, as well as their nanocrystals. Prior to joining NCSU, he worked

    as assistant professor at the French Engineering School of Paper, Printingand Biomaterials (PAGORA, Grenoble Institute of Technology, France)on the development of biodegradable nanocomposites based on nanoc-rystalline polysaccharides. He also spent two years as postdoctoral fellowat the French Institute for Agricultural Research, INRA, where he developednew nanostructured thin films based on cellulose nanowiskers. Dr. Habibisresearch interests include the sustainable production of materials frombiomass, development of high performance nanocomposites from ligno-cellulosic materials, biomass conversion technologies, and the applicationof novel analytical tools in biomass research.

    Chem. Rev. XXXX, xxx, 000000 A

    10.1021/cr900339w XXXX American Chemical Society

    1

    2

    34

    5

    6

    7

    8

    9

    10

    11

    12

    13

    14

    1516

    17

    18

    19

    20

    21

    22

    23

    24

    25

    26

    27

    2829

    30

    31

    32

    33

    34

    35

    36

    37

    38

    39

    40

    4142

    43

    44

    45

    46

    47

    48

    49

    50

    51

    52

    53

    54

    55

    56

    57

    58

    59

    60

    61

    62

    63

    64

    65

    66

    67

    68

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

    PAGE EST: 21.2

  • 8/14/2019 Nanocrystalline Cellulose

    2/22

    in some cases heteropolysaccharides provide the hierarchicalconstructs to bioengineer biological factories that give riseto a variegated distribution of plants and organisms. Surpris-ingly, a focus on nanoscale phenomena involving thesematerials has not been realized until the past few years inwhich a virtual collection of information has becomeavailable.

    In the following review, the salient chemical and physicalfeatures of the most dominant fundamental building blockin the biosphere, cellulose nanocrystals, are discussed. Aftera brief introduction to cellulose, three general aspects of CNsare covered, namely, their morphology and chemistry includ-ing their preparation and chemical routes for functionaliza-tion, self-assembly in different media and under differentconditions, and finally their applications in the nanocom-

    posites field. While these aspects are by no means compre-hensively inclusive of the vast number of research resultsavailable, they may be regarded as perhaps the mostscientifically and technologically pertinent aspects of CNsthat deserve attention.

    2. Structure and Morphology of Celluloses

    Cellulose is the most abundant renewable organic materialproduced in the biosphere, having an annual production thatis estimated to be over 7.5 1010 tons.1 Cellulose is widelydistributed in higher plants, in several marine animals (forexample, tunicates), and to a lesser degree in algae, fungi,bacteria, invertebrates, and even amoeba (protozoa), forexample, Dictyostelium discoideum. In general, cellulose isa fibrous, tough, water-insoluble substance that plays anessential role in maintaining the structure of plant cell walls.

    It was first discovered and isolated by Anselme Payen in1838,2 and since then, multiple physical and chemical aspectsof cellulose have been extensively studied; indeed, discover-ies are constantly being made with respect to its biosynthesis,assembly, and structural features that have inspired a numberof research efforts among a broad number of disciplines.Several reviews have already been published reporting thestate of knowledge of this fascinating polymer.1,3-13

    Regardless of its source, cellulose can be characterizedas a high molecular weight homopolymer of-1,4-linkedanhydro-D-glucose units in which every unit is corkscrewed180with respect to its neighbors, and the repeat segmentis frequently taken to be a dimer of glucose, known ascellobiose (Figure 1). Each cellulose chain possesses adirectional chemical asymmetry with respect to the terminiof its molecular axis: one end is a chemically reducingfunctionality (i.e., a hemiacetal unit) and the other has apendant hydroxyl group, the nominal nonreducing end. Thenumber of glucose units or the degree of polymerization (DP)is up to 20 000, but shorter cellulose chains can occur andare mainly localized in the primary cell walls.

    All-D-glucopyranose rings adopt a 4C1chair conforma-tion, and as a consequence, the hydroxyl groups arepositioned in the ring (equatorial) plane, while the hydrogenatoms are in the vertical position (axial). This structure is

    Dr. Lucian A. Lucia is currently an associate professor who works in thearea of renewable materials chemistry and engineering in the Departmentof Forest Biomaterials (formerly Wood & Paper Science) at North CarolinaState University. He received his Ph.D. in 1996 under Prof. Kirk S. Schanzefrom the University of Florida (Gainesville, FL) in organic chemistry witha focus on modeling the fundamental electron transfer pathways activein plant photosynthesis. He then spent almost 2 years at the NSF Centerfor Photoinduced Charge Transfer at the University of Rochester(Rochester, NY, under Prof. David G. Whitten) where he explored thedesign of highly functional materials based on photoactive amphiphilesand their behavior in heterogeneous media. He began his independentacademic career as an assistant professor at the Institute of Paper Scienceand Technology at the Georgia Institute of Technology (Atlanta, GA) in1997 where he extensively studied the materials chemistry and photo-chemistry of wood biopolymers. He joined the faculty at North CarolinaState University in 2004 where he began to explore fundamental andapplied aspects of renewable material-based composites. His researchinterests include antibacterial materials, cellulose nanocrystal templates,scaffolds for tissue engineering, drug delivery devices, biomass pretreat-ments, and wood pulping chemistry.

    Dr. Orlando Rojas is Associate Professor in North Carolina State Universityand Finland Distinguished Professor in Helsinki University of Technology.He worked as a Senior Scientist in the Department of Chemistry, PhysicalChemistry of the Royal Institute of Technology, KTH, and in the Institutefor Surface Chemistry, YKI. Prior appointments include his tenure asprofessor in the Department of Chemical Engineering of Universidad deLos Andes (Venezuela), after obtaining his Ph.D. in Chemical Engineeringin Auburn University and other graduate degrees that included a Diplomain Paper Engineering from the ETSII of Universidad Politecnica de Cataluna(Spain). He leads the Colloids and Interfaces Group at NC StateUniversity with a research focus on surface and colloid chemistry andthe adsorption behaviors of surfactants and polymers at solid/liquidinterfaces. He has studied the viscoelasticity of adsorbed monolayers bysurface laser light scattering, surface plasmon resonance, and quartzcrystal microgravimetry methods. His work has also involved state-of-the-art interferometric and bimorph surface force techniques, as well asatomic force microscopy, to unveil basic phenomena and interactions atthe nanoscale. His group currently works on nanocellulose structures,the dynamics of lignocellulose degradation, biosensor development, andseparation, derivatization, and use of natural polymers and surfactants.He is Associate Editor of theJournal of Surfactants and Detergentsandmember of the advisory committee of several other journals. Dr. Rojas isthe 2009-2010 Chair of the Division of Cellulose and Renewable

    Materials of the American Chemical Society and is the recipient of the2009 American Chemical Society Divisional service Award. He wasappointed 2009-2014 Finland Distinguish Professor by Tekes, the FinnishFunding Agency for Technology and Innovation and the Academy ofFinland.

    B Chemical Reviews, XXXX, Vol. xxx, No. xx Habibi et al.

    69

    70

    71

    72

    73

    74

    75

    76

    77

    78

    79

    80

    81

    82

    83

    8485

    86

    87

    88

    89

    90

    91

    92

    93

    94

    95

    96

    97

    98

    99

    100

    101

    102

    103

    104

    105

    106

    107108

    109

    110

    111

    112

    113

    114

    115

    116

    117

    118

    119

    120

    121

    122

    123

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystalline Cellulose

    3/22

    stabilized by an intramolecular hydrogen bond network

    extending from the O(3

    )-H hydroxyl to the O(5) ring oxygenof the next unit across the glycosidic linkage and from theO(2)-H hydroxyl to the O(6) hydroxyl of the next residue(Figure 2).

    The three most probable rotational positions of thehydroxymethyl group are defined by ascertaining the place-ment of the O6-C6 bond with respect to the O5-C5 andC4-C5 bonds: if O6-C6 isgaucheto O5-C5 andtranstoC4-C5, then the conformation is called gt, while the othertwo conformations are referred to as ggand tg (Figure 3).

    In nature, cellulose does not occur as an isolated individualmolecule, but it is found as assemblies of individual cellulosechain-forming fibers. This is because cellulose is synthesizedas individual molecules, which undergo spinning in ahierarchical order at the site of biosynthesis. Typically,approximately 36 individual cellulose molecules assembleare brought together into larger units known as elementaryfibrils (protofibrils), which pack into larger units calledmicrofibrils, and these are in turn assembled into the familiarcellulose fibers. However, celluloses from different sourcesmay occur in different packing as dictated by the biosynthesisconditions. The combined actions of biopolymerization,spinning, and crystallization occur in a rosette-shaped plasmamembrane complex having a diameter of 30 nm (Figure 4)and are orchestrated by specific enzymatic terminal com-

    plexes(TCs) that act as biological spinnerets.14 Because allthe cellulose chains in one microfibril must be elongated by

    the complex at the same rate, crystallization during cellulosesynthesis follows very closely polymerization of the chainsby the TCs.15,16 TCs are thought to be cellulose synthasecomplexes that belong to the large GT-A family of glyco-syltransferases; however, the reaction mechanism involvedin cellulose synthesis and assembly is still conjectural. The

    structure of cellulose microfibrils implies that their synthesesand assembly involve the coordinate activity of approxi-mately 36 active sites.17 However, diverse cellulose structuresin various organisms imply that the enzyme complex ismodular.14,18 Recent evidence from live-cell imaging ofcellulose indicates that microtubules exert a direct effect onthe orientation of cellulose deposition under specific condi-tions, but microtubules are not required for oriented deposi-tion of cellulose under other conditions (Figure 4).

    During the biosynthesis, cellulose chains are aggregatedin microfibrils that display cross dimensions ranging from 2to 20 nm, depending on the source of celluloses. Theaggregation phenomenon occurs primarily via van der Waalsforces and both intra- (Figure 2) and inter-molecularhydrogen bonds. If the TCs are not perturbed, they cangenerate a limitless supply of microfibrils having only alimited number of defects or amorphous domains.14,18 Anumber of models for the microfibril hierarchy have beenproposed that attempt to describe the supramolecular struc-ture of cellulose, including the crystalline structure, crystallitedimensions and defects, structural indices of amorphous

    domains, dimensions of fibrillar formation, etc. These modelsdiffer mainly in the description of the organization and thedistribution of the amorphous or less ordered regions withinthe microfibril. After many years of controversy, it iscommon practice to acknowledge that the amorphous regionsare distributed as chain dislocations on segments along the

    Figure 1. Chemical structure of cellulose.

    Figure 2. Intramolecular hydrogen-bonding network in a repre-sentative cellulose structure.

    Figure 3. The three most probable rotational positions of the hydroxymethyl group.

    Figure 4. Orientation of microtubules controlling the orientationof cellulose in the cell wall where the microtubules act like tracksto guide the cellulose enzymes floating in the cell membrane.Reprinted with permission from Ref 17. Copyright 2002 Elsevier.

    Cellulose Nanocrystals Chemical Reviews, XXXX, Vol. xxx, No. xx C

    124

    125126

    127

    128

    129

    130

    131

    132

    133

    134

    135

    136

    137

    138

    139

    140

    141

    142

    143

    144

    145

    146

    147

    148

    149

    150

    151

    152153

    154

    155

    156

    157

    158

    159

    160

    161

    162

    163

    164

    165

    166

    167

    168

    169

    170

    171

    172

    173

    174

    175

    176

    177

    178

    179180

    181

    182

    183

    184

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystalline Cellulose

    4/22

    elementary fibril where the microfibrils are distorted byinternal strain in the fiber and proceed to tilt and twist (Figure5).19

    In the ordered regions, cellulose chains are tightly packedtogether in crystallites, which are stabilized by a strong andvery complex intra- and intermolecular hydrogen-bondnetwork. The hydrogen-bonding network and molecularorientation in cellulose can vary widely, which can give riseto cellulose polymorphs or allomorphs, depending on therespective source, method of extraction, or treament.1,20 Sixinterconvertible polymorphs of cellulose, namely, I, II, IIII,IIIII, IVI, and IVII, have been identified.

    Native cellulose has been thought to have one crystal

    structure, cellulose I, but evidence for the existence of twosuballomorphs of cellulose I, termed IR and I, was

    established in 1984 by cross-polarization magic angle spin-ning (CP-MAS).21,22 Depending on the origin of cellulose,these two polymorphs exist in different ratios: IR is prevalentin celluloses from algae and bacteria, and both IRand Imay be present in celluloses in higher plants. However, thelatter result is not without controversy. Solid-state NMRstudies reported by Atalla and VanderHart23 has demonstratedseveral anomalies within the spectra of higher plant celluloses

    compared with those from algae, bacteria, and tunicates. Theanomalies found in the NMR spectra seem to suggest thathigher plants may contain only cellulose I, instead ofcellulose IR, with a distorted form of Ithat resides belowthe surface.

    In both the I and IRstructures, cellulose chains adoptparallel configurations, but they differ in their hydrogen-bonding patterns, which implies a difference in the crystallinestructure (Figure 6). Indeed, IR corresponds to a triclinicP1unit cell (a )6.717 , b )5.962 , c )10.400 , R )118.08, ) 114.80, and ) 80.37) containing only onechain per unit cell,24 whereas Iexists in a monoclinic P21unit cell having two cellulose chains (a ) 7.784 , b )

    8.201 , c ) 10.38 , R ) ) 90, and ) 96.5).25 IR,a metastable phase, can be converted to the more thermo-dynamically stable Iphase by high-temperature annealingin various media.26 Cellulose I (R and) has sheets stackedin a parallel-up fashion, and the hydroxylmethyl groupsare oriented in atgconformation so that their O6 atom pointstoward the O2 hydroxyl groups of the neighboring residue,which engenders a second inter-residue hydrogen bond.27 Ifcellulose is bent in a plane orthogonal to the hydrogen-bonded sheets of chains, a horizontal displacement of thesheets with respect to one another is induced.28 A bendingangle of 39has been shown to be sufficient to induce IRand Iinterconversion especially when the curvature of the

    chain sheets in the microfibril is modeled as a group ofconcentric circular arcs. Sine qua non, all microfibrils

    Figure 6. Hydrogen-bonding patterns in cellulose IRand I: (top) the two alternative hydrogen-bond networks in cellulose IR; (bottom)the dominant hydrogen-bond network in cellulose I(left) chains at the origin of the unit cell and (right) chains at the center of the unitcell according to Sturcova et al.32 Reprinted with permission from ref 32. Copyright 2004 American Chemical Society.

    Figure 5. Schematic representation of the elementary fibrilillustrating the microstructure of the elementary fibril and strain-distorted regions (defects). Reprinted with permission from Ref 19.Copyright 1972 John Wiley and Sons.

    D Chemical Reviews, XXXX, Vol. xxx, No. xx Habibi et al.

    185

    186

    187

    188

    189

    190

    191

    192

    193

    194

    195

    196

    197

    198199

    200

    201

    202

    203

    204

    205

    206

    207

    208209

    210

    211

    212

    213

    214

    215

    216

    217

    218

    219

    220

    221222

    223

    224

    225

    226

    227

    228

    229

    230

    231

    232

    233

    234235

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystalline Cellulose

    5/22

    generated from the TCs must torque sharply before they canadopt a parallel configuration with respect to the face of theinner cell wall. Therefore, the crystal form is likely to bedrastically changed before cellulose is incorporated into thecell wall. Such polymorphic differences were first evidencedby IR spectroscopy29 and electron diffraction30 and have beenfurther confirmed more recently by solid-state CP/MAS 13CNMR.21 The crystal structure and hydrogen-bonding pattern

    of cellulose I (Rand ) were later studied more deeply bysynchrotron X-ray and neutron fiber diffraction,24,25,31 wherein the latter, the hydrogen atom positions involved inhydrogen bonding were determined from Fourier-differenceanalysis with respect to hydrogenated and deuterated samples.The definition of all atomic spatial coordinates in thecellulose crystal structure was only possible for the first timebecause of the availability of these singular methods.

    Cellulose II, the second most extensively studied allo-morph, can be obtained by two different processes:

    (i) By chemical regeneration, which consists of dissolvingcellulose I in a solvent, then reprecipitating it in water.Suitable solvents for cellulose include, among others,

    solutions of heavy metal-amine complexes, mainlycopper with ammonia or diamine such as cupric hy-droxide in aqueous ammonia (Scheweizers reagentcalled cuoxam)33 or cupriethylenediamine (cuen),34

    ammonia or amine/thiocyanate,35 hydrazine/thio-cyanate,36 lithium chloride/N,N-dimethylacetamide(LiCl/DMAc),37,38 and N-methylmorpholine-N-oxide(NMMO)/water39-41 systems.

    (ii) By mercerization, a universally recognized process thename of which is derived from its inventor JohnMercer (1844),42 which consists of swelling nativecellulose in concentrated sodium hydroxide solutionsand yielding cellulose II after removing the swelling

    agent. Other swelling agents, such as nitric acid (65%),are also able to convert native fibers to cellulose II.43

    Some atypical bacterial species are reported to bio-synthesize cellulose II.44

    Cellulose II exists in a monoclinic P21phase (a )8.10, b ) 9.03 , c ) 10.31 , R ) ) 90, and )117.10).27,45,46 During conversion (I to II), the hydroxylgroups rotate from the tg to the gtconformation, whichexplicitly requires a change in the hydrogen-bond network.47,48

    In contrast to cellulose I, which has a parallel up arrangement,the chains in cellulose II are in an antiparallel arrangementyielding a more stable structure, which makes it preferablefor various textiles and paper materials. The conversion ofcellulose I to cellulose II has been widely consideredirreversible, although (partial) regeneration of cellulose I fromcellulose II has been reported.49,50

    If cellulose I or II is exposed to ammonia (gas or liquefied)or various amines,51 cellulose III is formed upon removal ofthe swelling agent. The resulting form of cellulose IIIdepends on whether the starting form is I or II, giving riseto cellulose IIIIor IIIII. Their diffraction patterns are similarexcept for the meridional intensities. Cellulose IIIIexists ina monoclinic P21 form (a ) 4.450 , b ) 7.850 , c )10.31 , R ) ) 90, and ) 105.10) with one chain inthe unit cell, displaying parallel chains as observed incellulose I.52 However, the hydroxymethyl groups are in thegtconformation and the intersheet hydrogen bond networkis similar to cellulose II. The exact structure of cellulose IIIIIis not clearly established yet, but the crystallographic andspectroscopic studies reported recently by Wada et al.53

    indicate that cellulose IIIIIis a disordered phase of cellulose.This disordered structure mainly contains a crystalline formhaving a unit cell (space group P21; a )4.45 , b )7.64, c )10.36 , R ) )90, )106.96) occupied byone chain organized in antiparallel fashion, in addition to asecond structure, as revealed by CP/MAS 13C NMR (spacegroup P21; a )4.45 , b )14.64 , c )10.36 , R )) 90, ) 90.05). Furthermore, either of these forms IIII

    or IIIII reverts to its parent structure if placed in a high-temperature and humid environment.Polymorphs IVI and IVII may be prepared by heating

    cellulose IIIIor IIIII, respectively, up to 260 C in glycerol.54,55

    In a like manner to the case of cellulose III, these two formscan revert to the parent structures I or II. In addition to thenative cellulose I, it has been shown that cellulose IV existsin several plant primary cell walls.56,57

    3. Cellulose Nanocrystals

    In the 1950s, Ranby reported for the first time that colloidalsuspensions of cellulose can be obtained by controlled

    sulfuric acid-catalyzed degradation of cellulose fibers.

    58-60

    This work was inspired by the studies of Nickerson andHabrle61 who observed that the degradation induced byboiling cellulose fibers in acidic solution reached a limit aftera certain time of treatment. Transmission electron microscopy(TEM) images of dried suspensions revealed for the first timethe presence of aggregates of needle-shaped particles, whilefurther analyses of these rods with electron diffractiondemonstrated that they had the same crystalline structure asthe original fibers.62,63 Simultaneously, the development byBattista64,65 of the hydrochloric acid-assisted degradation ofcellulose fibers derived from high-quality wood pulps,followed by sonification treatment, led to the commercializa-tion of microcrystalline cellulose (MCC). Stable, chemically

    inactive, and physiologically inert with attractive bindingproperties, MCC offered a significant opportunity for multipleuses in pharmaceutical industry as a tablet binder, in foodapplications as a texturizing agent and fat replacer, and also,as an additive in paper and composites applications. Afterthe acid hydrolysis conditions were optimized, Marchessaultet al.66 demonstrated that colloidal suspensions of cellulosenanocrystals exhibited nematic liquid crystalline alignment.Since the discovery of spectacular improvements in themechanical properties of nanocomposites with cellulosenanocrystals,67,68 substantial research has been directed tocellulose nanocrystal composites because of the growinginterest in fabricating materials from renewable resources.

    Cellulose nanocrystals are often referred to as microcrys-tals, whiskers, nanocrystals, nanoparticles, microcrystallites,or nanofibers. Hereafter, they are called cellulose nano-crystals (CNs). In the coming sections, methods for separa-tion of CNs and their morphologies, characterization, modi-fication, self-assembly, and applications will be reviewed.

    3.1. Preparation of Cellulose Nanocrystals

    The main process for the isolation of CNs from cellulosefibers is based on acid hydrolysis. Disordered or para-crystalline regions of cellulose are preferentially hydrolyzed,whereas crystalline regions that have a higher resistance toacid attack remain intact.69,70 Thus, following an acidtreatment that hydrolyzes the cellulose (leading to removalof the microfibrils at the defects), cellulose rod-like nano-crystals are produced. The obtained CNs have a morphology

    Cellulose Nanocrystals Chemical Reviews, XXXX, Vol. xxx, No. xx E

    236

    237

    238

    239

    240

    241

    242

    243

    244245

    246

    247

    248

    249

    250

    251

    252

    253

    254

    255

    256

    257258

    259

    260

    261

    262

    263

    264

    265

    266

    267

    268

    269

    270

    271

    272

    273

    274

    275

    276

    277

    278

    279

    280

    281

    282

    283

    284

    285

    286

    287

    288

    289

    290

    291

    292

    293

    294

    295

    296

    297

    298

    299

    300

    301

    302

    303

    304

    305

    306

    307

    308309

    310

    311

    312

    313

    314

    315

    316

    317

    318

    319

    320

    321

    322

    323

    324

    325

    326

    327

    328

    329

    330

    331

    332

    333

    334

    335

    336

    337

    338

    339

    340

    341

    342

    343

    344

    345

    346347

    348

    349

    350

    351

    352

    353

    354

    355

    356

    357

    358

    359

    360

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystalline Cellulose

    6/22

    and crystallinity similar to the original cellulose fibers;examples of such elements are given in Figure 7.

    The actual occurrence of the acid cleavage event isattributed to differences in the kinetics of hydrolysis betweenamorphous and crystalline domains. In general, acid hy-drolysis of native cellulose induces a rapid decrease in itsdegree of polymerization (DP), to the so-called level-off DP(LODP). The DP subsequently decreases much more slowly,even during prolonged hydrolysis times.65,71-75 LODP has

    been thought to correlate with crystal sizes along thelongitudinal direction of cellulose chains present in theoriginal cellulose before the acid hydrolysis. This hypothesiswas based on the reasonable assumption that disordered orpara-crystalline domains are regularly distributed along themicrofibers and therefore they are more susceptible to acidattack (in contrast to crystalline regions that are moreimpervious to attack). Also, homogeneous crystallites weresupposed to be generated after acid hydrolysis. Theseassumptions were actually confirmed by X-ray crystaldiffraction,76 electron microscopy with iodine-staining,76

    small-angle X-ray diffraction,72 and neutron diffractionanalyses.77 It was shown that the LODP values obtained by

    acid hydrolysis of cellulose correlated well with the periodiccrystal sizes along cellulose chains. The value of LODP hasbeen shown to depend on the cellulose origin, with typicalvalues of 250 being recorded for hydrolyzed cotton,64 300for ramie fibers,77 140-200 for bleached wood pulp,65 andup to 6000 for the highly crystalline Valonia cellulose.78

    However, a wide distribution of DPs is typically observedfor different cellulose sources, even at the LODP. Such adisparity in the quoted distributions stimulates vigorousdiscussion to the present day. In fact, the acid hydrolysis ofbacterial, tunicate, Valonia, or cotton results in a higherpolydispersity in the molecular weight, without any evidenceof the LODP, probably because these cellulosic materialshave no regular distribution of the less-organized domains.

    Typical procedures currently employed for the productionof CNs consist of subjecting pure cellulosic material to strongacid hydrolysis under strictly controlled conditions of tem-

    perature, agitation, and time. The nature of the acid and theacid-to-cellulosic fibers ratio are also important parametersthat affect the preparation of CNs. A resulting suspension issubsequently diluted with water and washed with successivecentrifugations. Dialysis against distilled water is thenperformed to remove any free acid molecules from thedispersion. Additional steps such as filtration,79 differentialcentrifugation,83 or ultracentrifugation (using a saccharose

    gradient)84 have been also reported.Specific hydrolysis and separation protocols have been

    developed that depend on the origin of the cellulosic fibers.Most common sources include among others, cellulose fibersfrom cotton,85,86 ramie,81,87,88 hemp,89 flax,90,91 sisal,82,92 wheatstraw,93 palm,94 bleached softwood95 and hardwood96 pulps,cotton linters pulp,97,98 microcrystalline cellulose,99-102 sugarbeet pulp,103 bacterial cellulose,104-106 and Tunicates.69,84,107

    Sulfuric and hydrochloric acids have been extensively usedfor CN preparation, but phosphoric108-111 and hydrobromic112

    acids have also been reported for such purposes. If the CNsare prepared by hydrolysis in hydrochloric acid, their abilityto disperse is limited and their aqueous suspensions tend to

    flocculate.113 On the other hand, when sulfuric acid is usedas a hydrolyzing agent, it reacts with the surface hydroxylgroups of cellulose to yield charged surface sulfate estersthat promote dispersion of the CNs in water, resulting inimportant properties that will be discussed shortly.114 How-ever, the introduction of charged sulfate groups compromisesthe thermostability of the nanocrystals.80 Also, differencesin the rheological behavior have been shown betweensuspensions obtained from sulfuric acid hydrolysis and thoseobtained from hydrochloric acid. In fact, the sulfuric acid-treated suspension has shown no time-dependent viscosity,whereas the hydrochloric acid-treated suspension showed athixotropic behavior at concentrations above 0.5% (w/v) and

    antithixotropic behavior at concentrations below 0.3%.113

    Post-treatment of CNs generated by hydrochloric acidhydrolysis with sulfuric acid has been studied to introduce,in a controlled fashion, sulfate moieties on their surfaces.86,95

    CNs generated from hydrochloric acid hydrolysis and thentreated with sulfuric acid solution had the same particle sizeas those directly obtained from sulfuric acid hydrolysis;however, the surface charge density could be tuned to givenvalues by sulfuric acid hydrolysis. With respect to themorphology of the particles, a combination of both sulfuricand hydrochloric acids during hydrolysis steps appears togenerate spherical CNs instead of rod-like nanocrystals whencarried out under ultrasonic treatment.115,116 These sphericalCNs demonstrated better thermal stability mainly becausethey possess fewer sulfate groups on their surfaces.116

    The concentration of sulfuric acid in hydrolysis reactionsto obtain CNs does not vary much from a typical value ofca. 65% (wt); however, the temperature can range from roomtemperature up to 70 C and the corresponding hydrolysistime can be varied from 30 min to overnight depending onthe temperature. In the case of hydrochloric acid-catalyzedhydrolysis, the reaction is usually carried out at refluxtemperature and an acid concentration between 2.5 and 4 Nwith variable time of reaction depending on the source ofthe cellulosic material. Bondenson et al.99,102 investigatedoptimizing the hydrolysis conditions by an experimentalfactorial design matrix (response surface methodology) usingMCC that was derived from Norway spruce (Picea abies)as the cellulosic starting material. The factors that were variedduring the process were the concentrations of MCC and

    Figure 7. TEM images of dried dispersion of cellulose nanocrystalsderived from (a) tunicate79 (Reprinted with permission from ref79. Copyright 2008 American Chemical Society), (b) bacterial80

    (Reprinted with permission from ref 80. Copyright 2004 American

    Chemical Society), (c) ramie81

    (From ref 81, Reproduced bypermission of The Royal Society of Chemistry), and (d) sisal82

    (Reprinted with permission from ref 82. Copyright 2006 Springer).

    F Chemical Reviews, XXXX, Vol. xxx, No. xx Habibi et al.

    361

    362

    363

    364

    365

    366

    367

    368

    369

    370371

    372

    373

    374

    375

    376

    377

    378

    379

    380

    381

    382

    383384

    385

    386

    387

    388

    389

    390

    391

    392

    393

    394

    395

    396

    397

    398

    399

    400

    401

    402

    403

    404

    405

    406

    407

    408

    409

    410

    411

    412

    413

    414

    415

    416

    417

    418

    419

    420

    421

    422

    423

    424

    425

    426

    427

    428

    429

    430

    431

    432

    433

    434435

    436

    437

    438

    439

    440

    441

    442

    443

    444

    445

    446

    447

    448

    449

    450

    451

    452

    453

    454

    455

    456

    457

    458

    459

    460

    461

    462

    463

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystalline Cellulose

    7/22

    sulfuric acid, the hydrolysis time and temperature, and theultrasonic treatment time. The responses that were measuredwere the median size of the cellulose particles and the yieldof the reaction. The authors demonstrated that with a sulfuricacid concentration of 63.5% (w/w) over a time of ap-proximately 2 h, it was possible to obtain CNs having alength between 200 and 400 nm and a width less than 10nm with a yield of 30% (based on initial weight). Prolonga-

    tion of the hydrolysis time induced a decrease in nanocrystallength and an increase in surface charge.85 Reaction timeand acid-to-pulp ratio on nanocrystals obtained by sulfuricacid hydrolysis of bleached softwood (black spruce, Piceamariana) sulfite pulp was investigated by Beck-Candanedoet al.96 They reported that shorter nanoparticles with narrowsize polydispersity were produced at longer hydrolysis times.Recently, Elazzouzi-Hafraoui et al.79 studied the size distri-bution of CNs resulting from sulfuric acid hydrolysis ofcotton treated with 65% sulfuric acid over 30 min at differenttemperatures, ranging from 45 to 72 C. By increasing thetemperature, they demonstrated that shorter crystals wereobtained; however, no clear influence on the width of the

    crystal was revealed.

    3.2. Morphology and Dimensions of CelluloseNanocrystals

    The geometrical dimensions (length,L, and width,w) of CNsare found to vary widely, depending on the source of thecellulosic material and the conditions under which the hydrolysisis performed. Such variations are due, in part, to the diffusion-controlled nature of the acid hydrolysis. The heterogeneity insize in CNs obtained from hydrolysis, for a given source type,can be reduced by incorporating filtration,79 differential cen-trifugation,83 or ultracentrifugation (using a saccharose gradi-ent)84 steps. The precise morphological characteristics are

    usually studied by microscopy (TEM, AFM, E-SEM,117 etc.)or light scattering techniques, including small angle neutronscattering (SANS)118 and polarized and depolarized dynamiclight scattering (DLS, DDLS).119 TEM images of CNstypically show aggregation of the particles, mainly due tothe drying step for the preparation of the specimens afternegative staining. Besides aggregation, additional instru-mental artifacts usually lead to an overestimation of CNdimensions. To overcome these issues, Elazzouzi-Hafraouiet al.79 recently reported the use of TEM in cryogenic mode(cryo-TEM) to prevent aggregation.

    Atomic force microscopy (AFM) has been widely usedto provide valuable and rapid indication of surface topog-

    raphy of CNs under ambient conditions at length scales downto the angstrom level.117,120-122 However, AFM topographymay show rounded cross-sectional profiles in cases whereother shapes are expected; for example, AFM imaging ofValoniagives shapes different than the square shape crosssection observed under TEM. This is probably due to artifactsthat result from substrate shape perturbations induced byAFM tip and tip-broadening effects. Finally, AFM was alsoreported to be a valuable technique to measure CNs me-chanical properties and interactions, such as stiffness andadhesion or pull-off forces.123

    Typical geometrical characteristics for CNs originatingfrom different cellulose sources and obtained with a varietyof techniques are summarized in Table 1. The reported widthis generally approximately a few nanometers, but the lengthof CNs spans a larger window, from tens of nanometers toseveral micrometers. An arresting observation is that there

    is a direct correspondence between the length of the CNsand the LODP of the corresponding material because it isgenerally recognized that the rodlike CN consists of fullyextended cellulose chain segments in a perfectly crystallinearrangement.

    CNs from wood are 3-5 nm in width and 100-200 nmin length, while those for Valonia, a sea plant, are reported

    to be 20 nm in width and 1000-2000 nm in length. Likewise,cotton gives CNs 5-10 nm in width and 100-300 nm long,and tunicate, a sea animal, gives ca. 10-20 nm in widthand 500-2000 nm long.69 The aspect ratio, defined as thelength-to-width (L/w) spans a broad range and can varybetween 10 and 30 for cotton and ca. 70 for tunicate.

    The morphology of the cross section of CNs also dependson the origin of the cellulose fibers. The basis of themorphological shape in the cross section may be attributedto the action of the terminal complexes during cellulosebiosynthesis. In fact, depending on the biological origin ofthe cell wall, different arrangements of TCs have beenobserved, which generate cellulose crystals with different

    geometries.

    14

    Despite the fact that acid hydrolysis appearsto erode the crystal by preferentially peeling off angularcellulose sheets, as has been reported by Helbert et al.,124 anumber of analyses of cross sections of CNs have neverthe-less attempted to characterize the inherent CN geometry.Based on TEM observations, Revol125 reported that the crosssection of cellulose crystallites in Valonia Ventricosa wasalmost square, with an average lateral element length of 18nm. In contrast, CNs from tunicate that were analyzed byTEM126 and SANS were found to have a rectangular 8.8 nm18.2 nm cross-sectional shape.118

    The morphology of CNs along the axis of the crystal seemsto also present different features, depending on the sourceof the nanocrystal. CNs from bacterial cellulose121 andtunicate79 have been reported to have ribbon-like shapes withtwists having half-helical pitches of 600-800 nm (Micras-terias denticulata) and 1.2-1.6m, respectively. However,

    Table 1. Examples of the Length (L) and Width (w) of CNsfrom Various Sources Obtained by Different Techniques

    source L(nm) w(nm) technique ref

    bacterial 100-1000 10-50 TEM 105100-1000 5-10 30-50 TEM 80, 104

    cotton 100-150 5-10 TEM 12770-170 7 TEM 128200-300 8 TEM 129255 15 DDL 119

    150-210 5-11 AFM 117cotton linter 100-200 10-20 SEM-FEG 97

    25-320 6-70 TEM 79300-500 15-30 AFM 130

    MCC 35-265 3-48 TEM 79250-270 23 TEM 101500 10 AFM 100

    ramie 150-250 6-8 TEM 8150-150 5-10 TEM 131

    sisal 100-500 3-5 TEM 82150-280 3.5-6.5 TEM 92

    tunicate 8.8 18.2 SANS 1181160 16 DDL 119500-1000 10 TEM 1071000-3000 15-30 TEM 132100-1000 15 TEM 129

    1073 28 TEM 79Valonia >1000 10-20 TEM 125soft wood 100-200 3-4 TEM 95, 113

    100-150 4-5 AFM 96hard wood 140-150 4-5 AFM 96

    Cellulose Nanocrystals Chemical Reviews, XXXX, Vol. xxx, No. xx G

    464

    465

    466

    467

    468

    469

    470

    471

    472473

    474

    475

    476

    477

    478

    479

    480

    481

    482

    483

    484

    485

    486

    487

    488

    489

    490

    491

    492

    493

    494

    495

    496

    497498

    499

    500

    501

    502

    503

    504

    505

    506

    507

    508

    509

    510511

    512

    513

    514

    515

    516

    517

    518

    519

    520

    521

    522

    523

    524

    525

    526

    527

    528

    529

    530

    531

    532

    533

    534535

    536

    537

    538

    539

    540

    541

    542

    543

    544

    545

    546

    547

    548

    549

    550

    551

    552

    553

    554

    555

    556

    557

    558

    559

    560

    561

    562

    563

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystalline Cellulose

    8/22

    these twisted features have not been clearly evidenced inCNs extracted from higher plants, which are believed to beflat with uniplanar-axial orientation.

    4. Chemical Modifications of CelluloseNanocrystals

    Because of a natural advantage of an abundance of

    hydroxyl groups at the surface of CNs, different chemicalmodifications have been attempted, including esterification,etherification, oxidation, silylation, polymer grafting, etc.Noncovalent surface modification, including the use ofadsorbing surfactants and polymer coating, has been alsostudied. All chemical functionalizations have been mainlyconducted to (1) introduce stable negative or positiveelectrostatic charges on the surface of CNs to obtain betterdispersion (CNs obtained after sulfuric acid hydrolysisintroduce labile sulfate moieties that are readily removed

    under mild alkaline conditions) and (2) tune the surfaceenergy characteristics of CNs to improve compatibility,especially when used in conjunction with nonpolar orhydrophobic matrices in nanocomposites. The main challengefor the chemical functionalization of CNs is to conduct theprocess in such a way that it only changes the surface ofCNs, while preserving the original morphology to avoid anypolymorphic conversion and to maintain the integrity of the

    crystal.

    4.1. Noncovalent Surface Chemical Modifications

    Noncovalent surface modifications of CNs are typicallymade via adsorption of surfactants. This approach has beenintroduced by Heux et al.,129,133 who used surfactantsconsisting of the mono- and di-esters of phosphoric acidbearing alkylphenol tails. The obtained surfactant-coated CNsdispersed very well in nonpolar solvents.129 Detailed analyses

    Figure 8. Scheme of TEMPO-mediated oxidation mechanism of the hydroxymethyl groups of cellulose (top, reaction scheme) and cross-sectional representation of cellulose nanocrystal indicating the occurrence of the surface TEMPO-mediated oxidation of available hydroxylgroups (bottom, surface crystal representations).

    H Chemical Reviews, XXXX, Vol. xxx, No. xx Habibi et al.

    564

    565

    566

    567

    568

    569

    570571

    572

    573

    574

    575

    576

    577

    578

    579

    580

    581

    582

    583

    584

    585

    586

    587

    588

    589

    590

    591

    592

    593

    594

    595

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystalline Cellulose

    9/22

    of the data provided by SANS revealed that the surfactantmolecules formed a thin layer of about 15 at the surfaceof the CNs.134 When the surface-modified CNs were incor-porated into isotactic polypropylene, they showed very goodcompatibility and they acted as remarkable nucleating agentsto induce the formation of the rare crystalline form inaddition to the regular crystalline form of isotactic polypro-pylene R.135,136 An anionic surfactant was also used by

    Bondeson et al.137 to enhance the dispersion of CNs inpoly(lactic acid) (PLA). Kim et al.138 and Rojas et al.139 usednonionic surfactants to disperse CNs in polystyrene-basedcomposite fibers. Zhou et al.140 recently reported a new andelegant way of CN surface modification based on theadsorption of saccharide-based amphiphilic block copoly-mers. By mimicking lignin-carbohydrate copolymers, theyadsorbed xyloglucan oligosaccharide-poly(ethylene glycol)-polystyrene triblock copolymer onto the surface of CNs. Theresulting CNs showed excellent dispersion abilities in non-polar solvents.

    4.2. TEMPO-Mediated Oxidation

    (2,2,6,6-Tetramethylpiperidine-1-oxyl)-mediated (orTEMPO-mediated) oxidation of CNs has been used toconvert the hydroxylmethyl groups present on their surfaceto their carboxylic form. This oxidation reaction, whichis highly discriminative of primary hydroxyl groups, isalso green and simple to implement. It involves theapplication of a stable nitroxyl radical, the 2,2,6,6-tetramethylpiperidine-1-oxyl (TEMPO), in the presenceof NaBr and NaOCl (see Figure 8, top). The use of thistechnique has been the subject of a number of reports sinceit was first introduced by De Nooy et al.,141 who showedthat only the hydroxymethyl groups of polysaccharides wereoxidized, while the secondary hydroxyls remained unaffected.

    In fact, TEMPO-mediated oxidation of CNs involves atopologically confined reaction sequence, and as a conse-quence of the 2-fold screw axis of the cellulose chain, onlyhalf of the accessible hydroxymethyl groups are availableto react, whereas the other half are buried within thecrystalline particle (Figure 8, bottom).

    TEMPO-mediated oxidation of CNs, obtained from HClhydrolysis of cellulose fibers, was first reported by Araki etal.127 as an intermediate step to promote grafting of polymericchains. These authors demonstrated that after TEMPO-mediated oxidation, the CNs maintained their initial mor-phological integrity and formed a homogeneous suspensionwhen dispersed in water. The basis for these latter observa-

    tions was the presence of the newly installed carboxyl groupsthat imparted negative charges at the CN surface and thusinduced electrostatic stabilization. Similar observations werereported by Montanari et al.142 who also showed that duringexcessive TEMPO-mediated oxidation, a decrease of thecrystal size occurred resulting from the partial delaminationof cellulose chains that are extant on the surface.

    Habibi et al.143 performed TEMPO-mediated oxidation ofCNs obtained from HCl hydrolysis of cellulose fibers fromtunicate and showed that it did not compromise the mor-phological integrity of CNs or their native crystallinity. Onthe basis of the supramolecular structure, morphology, andcrystallographic parameters of the CNs, these authorsdemonstrated that various degrees of oxidation can bepredicted and achieved by using specific amounts of theprimary oxidizing agent, that is, NaOCl (see Figure 8,bottom). When dispersed in water, TEMPO-oxidized or

    carboxylated CN suspensions display birefringence patternsand do not flocculate or sediment owing to the polyanioniccharacter imparted by the negative charges on the CNssurfaces (see Figure 9).

    4.3. Cationization

    Positive charges can also be easily introduced on thesurface of CNs; for example, weak or strong ammonium-containing groups, such as epoxypropyltrimethylammoniumchloride (EPTMAC), can be grafted onto the CN surfaces.144

    Such surface cationization proceeds via a nucleophilicaddition of the alkali-activated cellulose hydroxyl groups tothe epoxy moiety of EPTMAC and leads to stable aqueoussuspensions of CNs with unexpected thixotropic gellingproperties. Shear birefringence was observed, but no liquidcrystalline chiral nematic phase separation was detected forthese cationic CNs, most likely owing to the high viscosityof the suspension.

    4.4. Esterification, Silylation and Other SurfaceChemical Modifications

    Homogeneous and heterogeneous acetylation of modelCNs extracted from Valoniaand tunicate has been studiedby Sassi and Chanzy by using acetic anhydride in aceticacid.126 Their ultrastructural study, carried out by TEMimaging and X-ray diffraction, showed that the reactionproceeded by a reduction of the diameters of the crystals,while only a limited reduction in CN length was observed.It has been suggested that the reaction involved a nonswellingmechanism that affected only the cellulose chains localizedat the crystal surface. In the case of homogeneous acetylation,the partially acetylated molecules immediately partitionedinto the acetylating medium as soon as they were sufficientlysoluble, while in heterogeneous conditions, the celluloseacetate remained insoluble and surrounded the crystallinecore of unreacted cellulose chains. The simultaneous occur-rence of cellulose hydrolysis and acetylation of hydroxyl

    Figure 9. Aqueous 0.53% (w/v) suspensions of cellulose nano-crystals observed between crossed polarizers (1) after productionby HCl-catalyzed hydrolysis (left) and (2) after their oxidation viaTEMPO-mediated reactions (right). Reprinted with permission fromref 143. Copyright 2006 Springer.

    Cellulose Nanocrystals Chemical Reviews, XXXX, Vol. xxx, No. xx I

    596

    597

    598

    599

    600

    601

    602

    603

    604605

    606

    607

    608

    609

    610

    611

    612

    613

    614

    615

    616

    617

    618

    619

    620

    621

    622

    623

    624

    625

    626

    627

    628

    629630

    631

    632

    633

    634

    635

    636

    637

    638

    639

    640

    641

    642643

    644

    645

    646

    647

    648

    649

    650

    651

    652

    653

    654

    655

    656

    657

    658

    659

    660

    661

    662

    663

    664

    665

    666

    667

    668

    669

    670

    671

    672

    673

    674

    675

    676

    677

    678

    679

    680

    681

    682

    683

    684

    685

    686

    687

    688

    689

    690

    691

    692

    693

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystalline Cellulose

    10/22

    groups has been also reported. Fischer esterification ofhydroxyl groups simultaneously with the hydrolysis ofamorphous cellulose chains has been introduced as a viableone-pot reaction methodology that allows isolation of acety-lated CNs in a single-step process (Figure 10).145,146

    An environmentally friendly CN surface acetylation routewas recently developed by Yuan et al.147 involving a lowreagent consumption and simple-to-apply procedure. Themethod used alkyenyl succinic anhydride (ASA) aqueous

    emulsions as a template. The emulsions were simply mixedwith CN suspensions and freeze-dried, and the resulting solidwas heated to 105 C. The obtained derivative conferred tothe acylated CNs a highly hydrophobic character that wasevident because they were easily dispersible in solvents withwidely different polarities as measured by the respectivedielectric constant,; for example, they were dispersible notonly in DMSO having a very high of 46.45 but also in1,4-dioxane that has a quite low of 2.21. Berlioz et al.148

    have reported recently a new and highly efficient syntheticmethod for an almost complete surface esterification of CNs,leading to highly substituted CN esters. The reaction of fattyacid chains was carried out on dried CNs via a gas-phase

    process. It has been shown by SEM and X-ray diffractionanalyses that the esterification proceeded from the surfaceof the substrate to the crystal core. Under moderate condi-tions, the surface was fully reacted, whereas the originalmorphology was maintained and the core of the crystalremained unmodified. Esterification of CNs by reactingorganic fatty acid chlorides, having different lengths of thealiphatic chain (C12 to C18), has also been reported with agrafting density high enough that the fatty acids withbackbones of 18 carbons were able to crystallize on thesurface of the CNs.131

    Cellulose whiskers resulting from the acid hydrolysis oftunicate have been partially silylated by a series of alkyl-dimethylchlorosilanes, with the carbon backbone of the alkylmoieties ranging from a short carbon length of isopropyl tolonger lengths represented by n-butyl, n-octyl, and n-dodecyl.149 It has been demonstrated that with a degree of

    substitution (DS) between 0.6 and 1, the whiskers becamereadily dispersible in solvents of low polarity (such as THF)leading to stable suspensions with birefringent behavior,while their morphological integrity was preserved. However,at high silylation (DS greater than 1), the chains in the coreof the crystals became silylated, resulting in the disintegrationof the crystals and consequently the loss of original morphol-ogy. Surface trimethyl silylation of CNs from bacterialcellulose and their resulting cellulose acetate butyrate104,150

    or polysiloxane151 based nanocomposites was also investi-gated by Roman and Winter. Finally, coupling CNs with

    N-octadecyl isocyanate, via a bulk reaction in toluene, hasalso been reported to enhance their dispersion in organicmedium and compatibility with polycaprolactone, whichsignificantly improved the stiffness and ductility of theresultant nanocomposites.92

    4.5. Polymer Grafting

    Polymer grafting on the surface of CNs has been carriedout using two main strategies, namely, the grafting-ontoand grafting-from.152 The grafting onto approach involvesattachment onto hydroxyl groups at the cellulose surface ofpresynthesized polymer chains by using a coupling agent.In the grafting from approach, the polymer chains areformed by in situ surface-initiated polymerization fromimmobilized initiators on the substrate.

    The grafting onto approach was used by Ljungberg etal.153 to graft maleated polypropylene (PPgMA) onto thesurface of tunicate-extracted CNs. The resulting graftednanocrystals showed very good compatibility and highadhesion when dispersed in atactic polypropylene. Araki etal.127 and Vignon et al.154 studied the grafting of amine-terminated polymers on the surface of TEMPO-mediatedoxidized CNs by using a peptide coupling process catalyzedby carbodiimide derivatives in water. The same approachhas been implemented by Mangalam et al.155 who graftedDNA oligomers on the surface of CNs. The grafting ofpolycaprolactone having different molecular weights on the

    Figure 10. Reaction scheme illustrating the one-pot (tandem) cellulose hydrolysis and esterification reactivity of hydroxyl groups. Reprintedwith permission from ref 145. Copyright 2009 American Chemical Society.

    J Chemical Reviews, XXXX, Vol. xxx, No. xx Habibi et al.

    694

    695

    696

    697

    698

    699

    700

    701

    702

    703704

    705

    706

    707

    708

    709

    710

    711

    712

    713

    714

    715

    716717

    718

    719

    720

    721

    722

    723

    724

    725

    726

    727

    728

    729

    730

    731

    732

    733

    734

    735

    736

    737

    738

    739

    740

    741

    742743

    744

    745

    746

    747

    748

    749

    750

    751

    752

    753

    754

    755

    756

    757

    758

    759

    760

    761

    762

    763

    764

    765

    766

    767

    768

    769

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystalline Cellulose

    11/22

    surface of CNs has been achieved by using isocyanate-mediated coupling.87 These authors reported reaching agrafting density that was high enough that the grafted PCLchains were able to crystallize at the surface of CNs. Similarefforts were made by Cao et al.98 who reported the isocy-anate-catalyzed grafting of presynthesized water-borne poly-urethane polymers via a one-pot process. Such crystallizationprovoked cocrystallizations of the free chains of the respec-

    tive polymer matrices during CN-based nanocompositeprocessing. Furthermore, this cocrystallization phenomenoninduced the formation of a co-continuous phase between thematrix and filler, which significantly enhanced the interfacialadhesion and consequently contributed to a highly improvedmechanical strength of the resulting nanocomposites.

    The grafting from approach applied to CNs was firstreported by Habibi et al.,81 who grafted polycaprolactoneonto the surface of CNs via ring-opening polymerization(ROP) using stannous octoate (Sn(Oct)2) as a grafting andpolymerization agent. Likewise, Chen et al.156 and Lin etal.157 conducted similar grafting reactions under microwaveirradiation to enhance the grafting efficiency. In situ poly-

    merization of furfuryl alcohol from the surface of cellulosewhiskers was studied by Pranger et al.100 In this case, thepolymerization was catalyzed by sulfonic acid residues fromthe CN surface. At elevated temperatures, the sulfonic acidgroups were de-esterified and consequently released into themedium to catalyze in situ the polymerization. Yi et al.158

    and Morandi et al.159 propagated polystyrene brushes viaatom transfer radical polymerization (ATRP) on the surfaceof CNs with ethyl 2-bromoisobutyrate as the initiator agent.Similarly, other vinyl monomers, mainly acrylic monomerssuch asN-isopropylacrylamide, were also polymerized fromthe surface of CNs to produce thermoresponsive substr-ates.160,161 Grafting of polyaniline from CNs was achievedbyin situpolymerization of aniline onto CNs in hydrochloricacid aqueous solution, via an oxidative polymerization usingammonium peroxydisulfate as the initiator.162

    5. Self-Assembly and -Organization of CelluloseNanocrystals

    When sulfuric acid is used as the hydrolyzing agent, italso chemically reacts with the surface hydroxyl groups ofCNs to yield negatively charged (surface) sulfate groups thatpromote a perfectly uniform dispersion of the whiskers inwater via electrostatic repulsions.114 By inference, continuous

    removal of the water phase should therefore tend to causethe nanocrystals to adopt configurations that minimize theexisting electrostatic interactions. Indeed, (homogeneous)concentrated suspensions self-organize into spectacular liquidcrystalline arrangements, a phenomenon similar to whatoccurs in nonflocculating suspensions of other rod-likeparticles, such as poly(tetrafluoroethylene) whiskers,163 to-bacco mosaic viruses (TMV),164 DNA fragments,165 or

    crystallites extracted from other polysaccharides such aschitin.166

    This self-organization phenomenon was revealed by theappearance of fingerprint patterns obtained from suspen-sions observed by polarized optical microscopy, indicativeof a chiral-nematic ordering.114 An even more striking findingis that this chiral nematic structure can be preserved aftercomplete water evaporation to provide iridescent films ofCNs. These solid films, in addition to allowing fundamentalstudies of their striking behavior, have numerous potentialapplications such as coating materials for decorative materialsand security papers (because the optical properties cannotbe reproduced by printing or photocopying).167

    An investigation into these systems reveals that CNs arerandomly oriented in the dilute regime (isotropic phase).Indeed, polarized optical microscopy demonstrates that atdilute concentrations, CNs appear as spheroids or ovaloidsand the initial ordered domains are similar to tactoids. Anematic liquid crystalline alignment is adopted when the CNconcentration increases because these tactoids coalesce toform an anisotropic phase, which is characterized by aunidirectional self-orientation of the CN rods. When thesuspension reaches a critical concentration of CNs, it formsa chiral nematic ordered phase displaying lines that are thesignature of cholesteric liquid crystals (see Figure 14). Abovethe critical concentration of chiral nematic phase formation,

    aqueous CN suspensions produce shear birefringence, andon standing, they can spontaneously separate into an upperisotropic and a lower anisotropic phase (Figure 11).

    These chiral nematic or cholesteric structures in theanisotropic phase consist of stacked planes of CN rodsaligned along a vector (director), with the orientation of eachdirector rotated about the perpendicular axis from one planeto the next as shown in Figure 11. The self-induced parallelalignment phenomenon of the CNs that occurs above acritical concentration is attributed to the well-known en-tropically driven self-orientation phenomenon of rod-like

    Figure 11. (left) Aqueous 0.63% (w/w) CN suspension observed between crossed polarizers. Immediately after shearing the suspensionshows many iridescent birefringence patterns; after 1 week, the suspension separates into the upper isotropic and the lower anisotropicphases (Reprinted with permission from ref 105. Copyright 2001 American Chemical Society). (right) Schematic representation of CNorientation in both the isotropic and anisotropic (chiral nematic) phases.

    Cellulose Nanocrystals Chemical Reviews, XXXX, Vol. xxx, No. xx K

    770

    771

    772

    773

    774

    775

    776

    777

    778779

    780

    781

    782

    783

    784

    785

    786

    787

    788

    789

    790

    791792

    793

    794

    795

    796

    797

    798

    799

    800

    801

    802

    803

    804

    805

    806

    807

    808

    809

    810

    811

    812

    813

    814

    815

    816

    817

    818

    819

    820

    821

    822823

    824

    825

    826

    827

    828

    829

    830

    831

    832

    833

    834

    835

    836

    837

    838

    839

    840

    841

    842

    843

    844

    845

    846

    847

    848849

    850

    851

    852

    853

    854

    855

    856

    857

    858

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystalline Cellulose

    12/22

  • 8/14/2019 Nanocrystalline Cellulose

    13/22

    display a pitch that decreases with increasing CN concentra-tion and can vary from 20 to 80 m.

    The isotropic-to-anisotropic (chiral nematic phase) equi-librium is sensitive to the presence of electrolytes and thespecific nature of the electrolyte counterions. Quantitativestudies of the changes in composition of the isotropic andanisotropic phases as a function of electrolyte concentrationwhere both phases coexist,128 as well as the effect of thetype of counterion,173 have been conducted. It has been foundfrom the latter study that increasing the amount of addedelectrolyte decreases anisotropic phase formation. Interest-ingly, the chiral nematic pitch was found to decrease, thatis, the phase became more highly twisted, as the electrolyteconcentration increased. Apparently, the decrease in pitchoccurred because the decrease in the electrical double layerthickness increased the chiral interactions between thecrystallites.128 As already indicated, the phase separation ofsulfated CN suspensions also depends strongly on the natureof their counterions.173 For inorganic counterions, the criticalconcentration for ordered phase formation increases ingeneral as a function of increasing van der Waals radii, in

    the order H+ < Na+ < K+

  • 8/14/2019 Nanocrystalline Cellulose

    14/22

    Carboxylated CNs, prepared by TEMPO-mediated oxida-tion, have been shown to form homogeneous dispersions inwater that are strongly birefringent. This shear birefringencewas not uniformly distributed throughout the system, butconsisted instead of domains of various sizes and colorsindicative of local domain orientation within the CNs thatnever reached the chiral nematic order in the form of eithertactoids or fingerprints. The lack of further organization was

    ascribed to the high polydispersity among the length of theCNs (that were obtained from tunicate by HCl acid prehy-drolysis) and to the high viscosity of the suspensions.143

    However, when carboxylated CNs were prepared from cottonfibers, a reduced CNs length polydispersity was observed,and thus, these suspensions reached a chiral nematic orderwith a pitch of 7 m at a concentration of 5% or more(w/w). Furthermore, when PEG was grafted on the surfaceof the CNs, the resulting PEG-grafted CNs gave rise to achiral nematic mesophase through a phase separation similarto that of the unmodified CNs, but with a reduced spacingof the fingerprint pattern (around 4.0 m).127 Moreover,unlike what has been previously found, PEG-grafted CNs

    showed drastically enhanced dispersion stability even at highsolid content and the ability to redisperse into either wateror chloroform from a freeze-dried state. They also demon-strated strong stability at high ionic strength because of areduced electroviscous effect, and no aggregations wereobserved upon addition of electrolyte up to 2 M NaCl.

    Shear birefringence was also observed for suspensions ofcationic epoxypropyltrimethylammonium chloride-graftedCNs, but no liquid crystalline chiral nematic phase separationwas detected, most likely because the phase was inhibitedas a result of the high viscosity of the suspension.144

    5.2. Self-Assembly and -Organization of CNs in

    Organic MediumHeux et al.129 provided the first description of a self-

    ordering phenomenon for CNs in apolar solvents. In theirpreliminary study, surfactant coating was used to disperseCNs and thereby obtain a chiral-nematic structure. However,the pitch of the chiral nematic structure was found to beapproximately 4m, a value that is too small compared withthe case of aqueous suspensions (pitch between 20 and 80m). In addition, higher CN concentrations, up to 36%, couldbe achieved. These results were attributed to the stericstabilization exerted by the surfactant coating. In fact, thisstabilization screened out the electrostatic repulsion andconsequently induced stronger chiral interactions between

    rods that ultimately allowed for higher packing. Detailedexamination of the structure of this chiral nematic phase hasbeen reported recently by the same group.177,178 They studiedthe correlation of the aspect ratio of CNs extracted fromcotton fibers and their dispersion in an apolar solvent suchas cyclohexane. The critical concentration in which sponta-neous phase separation into a chiral nematic mesophase wasobserved was higher than that in water. Correlation withOnsagers theory for these organophilic suspensions was notpossible because the experimental critical concentrationswere much lower than the predicted ones, probably due toan attractive interaction between the rods in the apolarmedium. These strong interactions also induced a decreasein the chiral nematic pitches that were of the order of 2 mor less. In addition, suspensions prepared with CNs havinghigh aspect ratios did not show any phase separation butinstead produced an anisotropic gel phase at a high concen-

    tration.177 Conversely, the chiral nematic pitch was muchlarger (17m) for suspensions of CNs in toluene that werestabilized by xyloglucan oligosaccharide-poly(ethyleneglycol)-polystyrene triblock copolymer.140

    5.3. Self-Assembly and -Organization of CNsunder External Fields

    Nonflocculated CNs in an aqueous suspension under anexternal field, whether magnetic or electric, can be oriented.They can orient when they are subjected to a magnetic fielddue to the negative diamagnetic anisotropy of cellulose.Although the diamagnetic anisotropy of cellulose is relativelyweak per molecular repeat unit, the cellulose rod-likenanocrystals are long and heavy, so their total diamagneticanisotropy is rather larger compared with other particles suchDNA or TMV.179,180 Sugiyama et al.180 first demonstratedthat dilute aqueous suspensions of crystalline celluloseextracted from tunicate were able to orient when subjectedto a magnetic field of 7 T. Films were obtained in which thecrystals were oriented with their long axes perpendicular tothe magnetic field. An overall gross CN orientation couldbe achieved where the cholesteric axis became parallel tothe magnetic field rather than unwinding the chiral nematicstructure. Fleming et al. demonstrated that the orientationof a liquid crystalline CN suspension in the magnetic fieldof a NMR spectrometer can assist in interpretation of theNMR spectra of proteins added to the suspension.181 Re-cently, Kimura et al.132 showed that the chiral nematicbehavior of CN suspensions could be unwound by applyinga slowly rotating strong magnetic field. Kvien and Oksmanattempted to align CNs in a polymer matrix (e.g., polyvinylalcohol) by using a strong magnetic field to obtain aunidirectional reinforced nanocomposite, and interestinglythe results showed that the dynamic modulus of the nano-

    composite was higher in the aligned direction compared withthe transverse direction.182

    Aqueous suspensions of CNs extracted from ramie fibersand tunicate have been allowed to dry under an AC electricfield and have showed a high degree of orientation alongthe field vector in the films.183 Colloidal suspensions of CNsfrom ramie fibers in cyclohexane have been also oriented inan AC electric field184 as revealed by TEM and electrondiffraction. Moreover, these suspensions demonstrated abirefringence when observed under polarized light with crossnicols whose magnitude could be increased with increasingfield strength. They also displayed interference Newtoncolors, similar to those obtained for thin films of thermotropic

    liquid crystals, an interference pattern that reached a satura-tion plateau at ca. 2 kV/cm. This spectacular result wasobtained at concentrations of CNs below the isotropic/nematic transition, which excluded any cooperative effectbetween the electric field and a possible anisotropic phase.184

    5.4. Self-Assembly and -Organization of CNs inThin Solid Films

    Revol et al.167,185 have taken advantage of the ability ofsuspensions of CNs to engage in lyotropic chiral nematicordering to give rise to iridescent solid cellulosic films withunique and tunable optical properties by simply controllingthe evaporation of suspending water on a flat surface. Theliquid crystalline order obtained from these suspensions waspreserved in solid films and the chiral nematic pitch in thedried films was on the same length scale as the wavelength

    N Chemical Reviews, XXXX, Vol. xxx, No. xx Habibi et al.

    1020

    1021

    1022

    1023

    1024

    1025

    1026

    1027

    10281029

    1030

    1031

    1032

    1033

    1034

    1035

    1036

    1037

    1038

    1039

    1040

    10411042

    1043

    1044

    1045

    1046

    1047

    1048

    1049

    1050

    1051

    1052

    1053

    1054

    1055

    1056

    1057

    1058

    1059

    1060

    1061

    1062

    1063

    1064

    1065

    10661067

    1068

    1069

    1070

    1071

    1072

    1073

    1074

    1075

    1076

    1077

    1078

    1079

    1080

    1081

    1082

    108

    1084

    108

    108

    1087

    108

    10891090

    109

    1092

    109

    1094

    109

    109

    1097

    109

    1099

    1100

    110

    1102

    110

    1104

    110

    110

    1107

    110

    1109

    1110

    111

    1112

    111

    1114

    111

    1111117

    111

    1119

    1120

    112

    1122

    112

    1124

    112

    112

    1127

    112

    11291130

    113

    1132

    113

    1134

    113

    113

    1137

    113

    1139

    1140

    114

    1142

    114

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystalline Cellulose

    15/22

    of visible light. This unique system was described as an

    interference device having the capacity to reflect circularlypolarized light over a specific wavelength range. Becausethe wavelength of reflected light determines its spectral orintrinsic color, the perceived color of the film depends onthe pitch of the cholesteric order and the angle of incidenceof the light. The microstructure of such films is very sensitiveto the drying conditions. Moreover, the perceived color ofthe reflected polarized light is tunable because the finalpitch can be varied depending on processing variables suchas the aspect ratio of CNs and the electrolyte content. Filmsthat are prepared at ambient conditions generally show apolydomain (polymorphic) structure with the helical axes ofdifferent chiral nematic domains pointing in different direc-tions. In fact, Roman and Gray186,187 reported for the firsttime compelling evidence that a parabolic focal conic (PFC)structure (a symmetrical form of focal conic defects in whichthe line defects from a pair of perpendicular, antiparallel,and confocal parabolas) was trapped in these self-assembledsolid films (see Figure 15).

    Finally, it has been shown that a magnetic field duringdrying increased the size of the chiral nematic domains andaffected the orientation of the helical axis with respect tothe plane of the film.179

    6. Applications of Cellulose Nanocrystals inNanocomposite Materials

    Since the first publication related to the use of CNs asreinforcing fillers in poly(styrene-co-butyl acrylate) (poly(S-co-BuA))-based nanocomposites by Favier et al.,68 CNs haveattracted a great deal of interest in the nanocomposites fielddue to their appealing intrinsic properties such as nanoscaledimensions, high surface area, unique morphology, lowdensity (which is estimated to be 1.61 g/cm3188 for purecrystalline cellulose I), and mechanical strength. In addition,they are easily (chemically) modified, readily available,renewable, and biodegradable. CNs have been incorporatedinto a wide range of polymer matrices, including polysilox-anes,151 polysulfonates,189 poly(caprolactone),81,87,190 styrene-butyl acrylate latex,191 poly(oxyethylene),192-195 poly(styrene-co-butyl acrylate) (poly(S-co-BuA)),68 cellulose acetatebutyrate,104,196 carboxymethyl cellulose,197 poly(vinyl alco-hol),198 poly(vinyl acetate),82,97 poly(ethylene-vinyl acetate)(EVA),199 epoxides,70 polyethylene,131 polypropylene,136 poly-

    (vinyl chloride),200-203 polyurethane,204 and water-bornepolyurethane.98 Their incorporation into biopolymers, suchas starch-based polymers,69,89,90,205-208 soy protein,209 chito-san,130 or regenerated cellulose,210 and biopolymer-likepoly(lactic acid),137,211-213 poly(hydroxyoctanoate),214,215 andpolyhydroxybutyrates216 have also been reported.

    6.1. Nanocomposite Processing

    Processing techniques have an important impact on thefinal properties of the composites. The techniques that areadopted should take into consideration the intrinsic propertiesof CNs, their interfacial characteristics (modified or not), thenature of the polymeric matrix (solubility, dispersibility, anddegradation), and the desired final properties such as geo-metrical shape.

    6.1.1. Casting-Evaporation Processing

    Casting evaporation has been the main technique totransfer cellulose whiskers from an aqueous dispersion intoan organic polymer matrix. Nanocomposite films are formed

    via solution casting, that is, simply allowing the solvent toevaporate. Good dispersibility of the CNs in the polymermatrix, as well as in the processing solvent, is a prerequisiteto create polymer/whisker nanocomposites that display asignificant mechanical reinforcement. CNs without surfacemodification have intrinsically strong interactions and havebeen reported as notoriously difficult to disperse. Moreover,this issue is exacerbated when the CN dispersions are driedbefore nanocomposite processing, which generally impliesthat drying and redispersion of CNs without aggregation ischallenging.

    Due to the hydrophilic character of CNs, the simplestpolymer systems that incorporate CNs are water-based.

    Never-dried aqueous dispersions of CNs are simply mixedwith aqueous polymer solutions or dispersions. However,these systems suffer from limited utility and are onlyappropriate for water-soluble or dispersible polymers suchas latexes. The combination of aqueous solutions of polymers(subject to moderately strong hydrogen bonding, in misciblecosolvents such as THF) with aqueous CN suspensions hasbeen reported.217 The use of polar solvents, most commonly

    N,N-dimethylformamide (DMF), with CNs with no surfacemodification has been explored.218 Solvents nonmiscible withwater and with low polarity, such toluene, have also beenwidely used. However, a drawback is that the processrequires solvent exchange steps. Owing to the fact that

    acetone is miscible with water; it usually serves as carrierto transfer CNs from water to organic solvents. Freeze-dryingand redispersion of CNs from tunicate in toluene were usedto integrate these fillers into atactic polypropylene, but strongaggregation occurred.153 However, when formic acid wasused to redisperse dried CNs, very satisfactory dispersionswere achieved.219 Freeze-dried CNs were successfully re-dispersed in dipolar aprotic solvents, such DMSO and DMF,containing small amounts of water (0.1%), and it has beenpossible to obtain films of these suspensions by the casting-evaporation technique.220 More recently van den Berg et al.221

    investigated the factors limiting the dispersibility of CNsextracted from tunicate via hydrochloric or sulfuric acidhydrolyses in a series of polar protic and aprotic organicsolvents. HCl-generated CNs, which typically have a pro-nounced tendency to aggregate, did not disperse in polaraprotic solvents after being dried. Only protic solvents such

    Figure 15. Square lattice in a solid film of cellulose nanocrystalsbetween crossed polarizers. Scale bar 40 m. Reprinted withpermission from ref 186. Copyright 2005 American ChemicalSociety.

    Cellulose Nanocrystals Chemical Reviews, XXXX, Vol. xxx, No. xx O

    1144

    11451146

    1147

    1148

    1149

    1150

    1151

    1152

    1153

    1154

    1155

    1156

    1157

    1158

    1159

    1160

    1161

    1162

    1163

    1164

    1165

    1166

    1167

    1168

    1169

    11701171

    1172

    1173

    1174

    1175

    1176

    1177

    1178

    1179

    1180

    1181

    1182

    1183

    1184

    1185

    1186

    1187

    118

    1189

    1190

    119

    1192

    119

    1194

    119

    119

    1197

    119

    1199

    1200

    120

    1202

    120

    1204

    120120

    1207

    120

    1209

    1210

    121

    1212

    121

    1214

    121

    121

    1217

    1211219

    1220

    122

    1222

    122

    1224

    122

    122

    1227

    122

    1229

    1230

    1231232

    123

    1234

    123

    123

    1237

    123

    1239

    1240

    124

    1242

    124

    1244

    124

    124

    1247

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystalline Cellulose

    16/22

    as formic acid andm-cresol were shown to effectively disruptthe hydrogen bonds in aggregated CNs, dispersing both typesof CNs generated from hydrochloric or sulfuric acid hy-drolyses (Figure 16).

    Another approach already considered in this review tochange interactions of CNs is via surface modifications. Suchapproach can break the percolating hydrogen-bonded networkand affect the macroscopic mechanical properties of theresulting nanocomposite. Chemical modifications (discussedabove) have been explored to improve dispersibility of CNs

    in wide range of organic solvents, from medium to lowpolarity. This approach also allows manipulation of driedCNs because it facilitates freeze-drying and redispersion.

    6.1.2. Sol-Gel Processing

    Capadona et al. have recently reported a versatile process-ing approach consisting of forming a three-dimensionaltemplate through self-assembly of well individualized CNsand then filling the template with a polymer of choice.222-224

    The first step (Figure 17 left, a and b) is the formation of aCN template through a sol/gel process involving the forma-tion of a homogeneous aqueous whisker dispersion that isfollowed by gelation through solvent exchange with a water-miscible solvent (routinely acetone). In the second step(Figure 17 left, d and e), the CN template is filled with amatrix polymer by immersing the gel into a polymer solution

    (Figure 17). It should be noted that the polymer solvent mustbe miscible with the gel solvent andde rigueurnot redispersethe CNs.

    6.1.3. Other Processing Methods

    The use of twin extrusion as a processing method toprepare CN-based nanocomposites has been attempted.225

    The process consists of pumping an aqueous dispersion ofCNs coated with surfactant226 or poly(vinyl alcohol)211 into

    a melt polymer (i.e., poly(lactic acid), PLA) during extrusion.However, such systems have unfortunately shown a lack ofcompatibility. Starch and CN nanocomposites were processedby extrusion.227 Moreover, the extrusion of modified CNswith long chained molecules seemed to be much easier andcould be processed in solvent free conditions, especiallywhen the grafted chains can melt at the processing temper-ature. An example has been reported recently where fattyacid-grafted CNs were successfully extruded with low-density polyethylene.131 PCL-grafted CNs with long chainsof PCL have shown good thermoformability when subjectedto compression and injection.156

    Electrostatic fiber spinning or electrospinning, a versatilemethod to manufacture fibers with diameters ranging fromseveral micrometers down to 100 nm or less through theaction of electrostatic forces, has emerged as an alternativeprocessing method for CNs in polymer matrices. However,

    Figure 16. Photographs of 5.0 mg/mL dispersions of H2SO4-generated CNs viewed through cross polarizers: from left to right, as-preparedin water, freeze-dried, freeze-dried and redispersed in water, DMF, DMSO, N-methyl pyrrolidone, formic acid, and m-cresol. Reprintedwith permission from ref 219. Copyright 2007 American Chemical Society.

    Figure 17. (left) Schematic representation of the template approach to obtain well-dispersed polymer/CN composites: (a) a nonsolvent isadded to a dispersion of CNs in the absence of any polymer, (b) solvent exchange promotes the self-assembly of a gel of CNs, (c) the gelled

    CNs scaffold is interpenetrated with a polymer by immersion in a polymer solution, before the nanocomposite is (d) dried and (e) compacted.(right) The obtained gel after completing step b above Reprinted with permission from ref 101. Copyright 2009 American Chemical Society.

    P Chemical Reviews, XXXX, Vol. xxx, No. xx Habibi et al.

    1248

    1249

    1250

    1251

    1252

    1253

    1254

    1255

    1256

    1257

    12581259

    1260

    1261

    1262

    1263

    1264

    1265

    1266

    1267

    1268

    1269

    1270

    1271

    1272

    127

    1274

    127

    127

    1277

    127

    1279

    1280

    1281282

    128

    1284

    128

    128

    1287

    128

    1289

    1290

    129

    1292

    129

    1294

    129

    129

    1297

    ohio2/ycr-ycr/ycr-ycr/ycr99907/ycr2637d07z xppws 23:ver.7 2/23/10 8:58 Msc: cr-2009-00339w TEID: lmh00 BATID: 00000

  • 8/14/2019 Nanocrystal