Вы находитесь на странице: 1из 17

REVIEW

pubs.acs.org/Biomac

Regioselective Esterification and Etherification of Cellulose: A Review


S. Carter Fox, Bin Li, Daiqiang Xu, and Kevin J. Edgar*
Macromolecules and Interfaces Institute, Virginia Tech, Blacksburg, Virginia 24061, United States

ABSTRACT: Deep understanding of the structureproperty


relationships of polysaccharide derivatives depends on the
ability to control the position of the substituents around the
monosaccharide ring and along the chain. Equally important is
the ability to analyze position of substitution. Historically, both
synthetic control and analysis of regiochemistry have been very
dicult for cellulose derivatives, as for most other polysacchar-
ide derivatives. With the advent of cellulose solvents that are
suitable for chemical transformations, it has become possible to carry out cellulose derivatization under conditions suciently mild
to permit increasingly complete regiochemical control, particularly with regard to the position of the substituents around the
anhydroglucose ring. In addition, new techniques for forming cellulose and its derivatives from monomers, either by enzyme-
catalyzed processes or chemical polymerization, permit us to address new frontiers in regiochemical control. We review these
exciting developments in regiocontrolled synthesis of cellulose derivatives and their implications for in-depth structureproperty
studies.

1. INTRODUCTION evidence of the monosaccharide content of commercial cellulose


Polysaccharides are among the most abundant natural poly- ethers like methylcellulose and ethylcellulose because the sub-
mers on earth, possess remarkable structural diversity, and per- stituents are stable under acid hydrolysis conditions. Therefore, it
form an incredible variety of biological functions. Our has been possible to hydrolyze these cellulose ethers to mono-
understanding of this functional and structural diversity is still saccharides, reduce the monosaccharides to alditols, and analyze
growing rapidly. Cellulose is one of the simplest polysaccharides, the alditol mixture by chromatography versus authentic standards.8
being a homopolymer (f4--D-Glcp-1f) of a single mono- The analysis of monosaccharide content of cellulose esters is far
saccharide, with no branching or substituents in nature (Figure 1). more dicult because esters are not stable under the conditions
Cellulose is often cited (without supporting evidence) as the most needed to hydrolyze the polysaccharide to monosaccharides.
abundant polymer on earth; certainly as a major constituent of all Reports of reductive cleavage processes for the conversion of
plant matter and of many other organisms (e.g., bacteria,1 cellulose esters into monosaccharides without loss of the ester
tunicates2), it is vastly abundant. Because of this abundance, its groups were intriguing but have not proven to be useful in
relatively low cost, and the fact that its properties and processa- application.9 In the simplest cases, such as cellulose acetate, it is
bility can be dramatically modied through substitution reactions possible to quantify the bulk monosaccharide content by NMR
of its hydroxyl groups, cellulose is easily the most commercially spectroscopy.10,11 These studies have indicated that commercial
signicant polysaccharide, with billions of kilograms of its deriva- cellulose acetate is indeed rather randomly substituted, as was
tives sold each year for a wide range of applications.3 predicted. It is clear that control and analysis of regiochemistry of
To develop new applications or to better predict the perfor- substitution, even in such common and relatively simple cellulose
mance of cellulose derivatives in current applications,4 it is critical derivatives, is complex and challenging indeed.
to understand structure property relationships in detail. This can Issues of monosaccharide sequence control in polysaccharide
be dicult. The most important cellulose derivatives from an synthesis are even more dicult. Standard methods of poly-
applications perspective are esters5 and ethers.6 The simplest saccharide derivative synthesis, that is, by modifying a natural
cellulose ethers and esters contain only one substituent type; polysaccharide, typically by reaction with electrophiles, oer no
cellulose acetate (CA) and methylcellulose (MC) are important possible expectation of monosaccharide sequence control. We
examples. For cellulose derivatives with one substituent type like have to talk about expectations because most often we really do
CA and MC, there are eight possible monosaccharides present in not know the sequence of synthetic polysaccharide derivatives. In
the polymer. In the rather common case in which two substituent the case of cellulose ethers, we can compare the individual
types are present, such as cellulose acetate propionate (CAP) or substituted anhydroglucose unit (AGU) quantities, determined
cellulose acetate butyrate (CAB), there are 27 possible mono- as alditols as described above, with those expected by modeling of
saccharides (see examples of MC and CAP, Table 1). Standard
commercial methods for preparation of these cellulose deriva- Received: February 24, 2011
tives begin as heterogeneous reactions, involving in the case of Revised: April 11, 2011
esters a subsequent back-hydrolysis reaction.7 There is good Published: April 27, 2011

r 2011 American Chemical Society 1956 dx.doi.org/10.1021/bm200260d | Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

Figure 1. Anhydroglucose unit (AHG) and numbering.


Figure 2. Preparation of 6-O-trityl cellulose.
Table 1. Cellulose Derivative Monomer Content: Examples
and regiospecic techniques developed for synthesis of cellulose
CA, CAPa
ethers and esters. It pays particular attention to developments
CA monomer symbols CAP monomer symbols since the excellent review by Klemm and coauthors in 1997.26
(Note also the informative mini-review by Heinze and coauthors
1 UN 1 UN 10 36A 19 6A3P
in 2006,27 which largely focused on their own work.) By denition,
2 2M 2 2A 11 23P 20 2A36P every cellulose derivative synthesis is regioselective to some extent,
3 3M 3 3A 12 26P 21 3A26P so we imposed an arbitrary limit of 90% selectivity for one
4 6M 4 6A 13 36P 22 6A23P regioisomer for inclusion in this review to keep the scope manage-
5 23M 5 2P 14 2A3P 23 36A2P able and the review useful. We include as well for their instructional
6 26M 6 3P 15 2A6P 24 26A3P value a few examples that do not meet this criterion. We have
7 36M 7 6P 16 3A6P 25 23A6P organized the review by the general types of methodology that
8 236M 8 23A 17 3A2P 26 236A have been used by investigators, and we include a nal perspective
9 26A 18 6A2P 27 236P that attempts to point out as yet unconquered frontiers.
a
Legend: M = Methyl, A = Acetate, P = Propionate, UN = unsubstituted.
For example, 2A6P means cellulose 2-acetate-6-propionate.
2. REACTION OF CELLULOSE WITH BULKY REAGENTS
The advent of cellulose solvents that are useful for derivatiza-
ideal distributions, like random or copolymeric sequences, and tion reactions, beginning with N,N-dimethylacetamide (DMAc)/
evaluate the extent to which the measured quantities match those LiCl in the 1980s,20,28,29 created the possibility of far more selective
expected from a particular model.12,13 However, detailed analysis of reactions of cellulose. Researchers have explored in detail the
monosaccharide sequence for high polysaccharide polymers is simply possibility of taking advantage of inherent reactivity dierences
not possible using currently available analytical methods. Only by between cellulose hydroxyls. Key reactivity dierences include
using synthetic techniques that involve the laborious building of the greater acidity of the 2-OH and the fact that the 6-OH is a
polysaccharide sequences, adding monosaccharides to a growing primary alcohol and thus signicantly less sterically hindered
chain one-by-one using glycosylation reactions, with the accompany- than the other hydroxyl groups. To date, researchers have
ing protection and deprotection steps, or by using isolated poly- focused most heavily on exploiting the higher reactivity of the
saccharide biosynthetic enzymes (in the cases where they will accept 6-OH; the following sections describe the results of their eorts.
monosaccharides appropriately modied to achieve the desired 2.1. Etherification. Etherification has proven to be a far more
substitution pattern, or where biosynthetic enzymes are available effective and general strategy for regioselective substitution than
that catalyze appending the desired substituents to the polysacchar- has esterification, due to the closer proximity of the bulky group
ide chain after it is synthesized14) can we hope to approach specic, to the carbon atom to be substituted in ethers. The following
biologically important, complex polysaccharide sequences such as sections describe the useful tritylation and silylation chemistries.
those that control the specic protein binding of the glycosamino- 2.1.1. Tritylation. The reaction of cellulose with triphenyl-
glycans heparin15,16 and chondroitin sulfate.1719 chloromethane, hereafter referred to as trityl chloride, is one of
The advent of modern cellulose solvents2022 that permit the oldest and most effective protecting group strategies for
synthesis of cellulose derivatives in solution2325 has provided an synthesizing regioselectively modified cellulose derivatives. It has
opportunity for polysaccharide chemists to attempt regioselec- long been known that because of steric demands, trityl chloride
tive or even regiospecic syntheses of cellulose derivatives. (For reacts preferentially with the primary hydroxyl group at the O-6
the purpose of this review, we consider reactions to be regiose- position on the cellulose backbone rather than with either of the
lective if they provide >90% selectivity for one regioisomer and to secondary hydroxyl groups at the O-2/3 positions3033 (Figure 2).
be regiospecic if they provide >99% selectivity.) Cellulose Furthermore, the trityl group can be quantitatively removed by
hydroxyls are much more reactive in solution because dissolution acid hydrolysis after derivatization of the secondary hydroxyl
breaks up the extensive cellulose H-bonding network, meaning groups with acid stable functionalities (alkyl ethers, for example)
that reactions can be carried out under milder conditions to regenerate the free hydroxyl group at C-6. This permits
(temperature, catalysts) that permit higher selectivity. Clever synthesis of a variety of different cellulose derivatives where
cellulose chemists have taken advantage of this fact and of the hydroxyl O-2/3 substituents differ from those at O-6.
developing techniques for synthesis of polysaccharides from Tritylation of cellulose with high levels of crystallinity, such as
monosaccharides to prepare regioselectively substituted cellu- cotton cellulose or most dissolving pulps, fails to produce a
lose derivatives and make the rst advances toward full struc- readily soluble material. Thus, to obtain a soluble product, the
tureproperty understanding. This review covers regioselective cellulose must be activated to disrupt its crystallinity prior to
1957 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

tritylation. A historically common method to activate cellulose (DSmeth = 2.0) aqueous solutions, and it was concluded that
for tritylation is the complete deacetylation of cellulose acetate in the dimethylated AGUs do not aggregate upon heating.
aqueous ammonium hydroxide.31 Mercerized cellulose is also Petzold-Welke et al.45 found that the water solubility of cellulose
sometimes used.34,35 With these methods, the activated cellulose methylated at the O-2 and O-3 positions with DS 0.90 was
is then transferred to dry pyridine to form a slurry that becomes dependent on the frequency of occurrence of disubstituted
homogeneous as the reaction progresses. More recently, how- AGUs along the cellulose backbone. Samples containing more
ever, the cellulose solvent DMAc/LiCl has been used as the disubstituted AGUs were water-soluble, whereas samples contain-
reaction medium because it is a less time- and labor-intensive ing more monosubstituted AGUs tended to be water-insoluble.
procedure.36,37 Tritylation has also been performed homoge- Synthesis of 6-O-alkyl cellulose ethers using tritylation chem-
neously on cellulose dissolved in the ionic liquids 1-butyl-3- istry is somewhat more problematic than the synthesis of 2,3-di-
methylimidazolium chloride ([Bmim]Cl)38 and 1-allyl-3-methy- O-alkyl-celluloses. After the 6-O-trityl cellulose has been made,
limidazolium chloride ([Amim]Cl).39 The degree of substitution appropriate protecting groups must be chosen for the O-2 and
(DS) of the trityl group can range from 0.93 to 1.07, depending O-3 positions. These protecting groups must be stable to acid
on the reaction temperature, duration, and the molar ratios of the hydrolysis during the detritylation step, and then they must be
reactants. Approximately 90% of the substituents are found at the stable to alkaline conditions during the alkylation of O-6. Finally,
O-6 position, whereas a small proportion of the secondary they must still be easily removed after the alkylation step.
hydroxyls at O-2 and O-3 is also tritylated.31,40 Kondo46 provided a solution to this problem by choosing allyl
Heinze et al.41 and Gomez et al.42 reported that the use of groups as the protecting group for O-2 and O-3. After detrityla-
methoxy-substituted trityl chlorides signicantly increased the tion, 2,3-O-allyl-cellulose was isomerized to 2,3-di-O-(1-propenyl)-
rate of hydroxyl substitution on cellulose. A DS of 0.96 could be cellulose using potassium t-butoxide. Then, after alkylation at
achieved within 4 h at 70 C using p-monomethoxytriphenyl- O-6, the 1-propenyl groups were removed by acid hydrolysis. It
methyl chloride as the tritylation reagent, whereas use of the was reported that this method provided better results than
unsubstituted trityl chloride required over 24 h to achieve the protecting O-2 and O-3 as benzyl ethers.
same DS under similar reaction conditions. The O-6 selectivity of Inter- and intramolecular hydrogen bonding in regioselec-
the p-monomethoxytriphenylmethyl chloride was found to be tively methylated cellulose samples was studied by FTIR and
similar to that of the unsubstituted trityl chloride. The observed solid-state 13C NMR.47 The analyses indicated that intermole-
rate acceleration was attributed to stabilization of the triaryl- cular hydrogen bonding was largely absent in 6-O-methylcellu-
methyl cation reactive intermediate by the electron-donating lose, although it was prominent in 2,3-di-O-methylcelluloses.
methoxy group. Additionally, acid-catalyzed removal of meth- This suggests that a free primary hydroxyl group is necessary for
oxy-substituted trityl groups requires only 5.5 h, whereas strong intermolecular hydrogen bonding in methylcelluloses.
complete removal of unsubstituted trityl groups requires 100 h. These studies were extended in the investigations of hydrogen
This is particularly benecial because exposure of the cellulose to bonding within blends of regioselectively methylated cellulose
acid during detritylation can result in reduction of the molecular with poly(ethylene oxide) or poly(vinyl alcohol)47,48 as well as
weight of the polymer.31 In fact, Kern et al.34 reported that the the analysis of hydrogen bonding in aqueous solutions of
molecular weight of 2,3-di-O-methyl cellulose ethers synthesized regioselectively methylated cellulose.49 It was found that the
using microcrystalline cellulose as a starting material had not specic interactions of the methylcellulose with the synthetic
signicantly decreased after acid hydrolysis of monomethoxy polymers are highly dependent on the sites of substitution on the
trityl groups from O-6. Additionally, the organic solubility of cellulose. The 6-O-methylcellulose had high solubility in a variety
cellulose derivatized with alkoxy-substituted trityl groups can be of organic solvents because of the lack of hydrogen bonding
modied by increasing the alkoxy chain length.43 between cellulose chains, whereas 2,3-di-O-methylcellulose was
Tritylation chemistry has made possible many structure/ soluble only in polar aprotic solvents.
property relationship studies on cellulose derivatives substituted Regioselectively synthesized analogues of other important
at the O-2 and O-3 positions.44 Among the rst of these commercial cellulose ethers have also been produced using
compounds synthesized and characterized were 2,3-di-O-methyl- tritylation chemistry. Hydroxyethyl and hydroxypropyl deriva-
and 2,3-di-O-ethyl-cellulose.40 Kern et al.34 studied the behavior tives of trityl cellulose were heterogeneously synthesized in a
of 2,3-di-O-methyl-cellulose in aqueous solutions. They reported mixture of isopropanol and water.50 With the aid of ionic and
that subtle variations in the DS and distribution of the methyl nonionic surfactants in the reaction mixture, a molar degree of
substituents can have dramatic eects on the phase separation substitution of 2.0 was obtained for both the hydroxyethyl and
behavior of the cellulose derivative. Highly methylated cellulose hydroxypropyl derivatives. NMR analysis of the products showed
samples (DSmeth > 2.0) with a signicant number of anhydroglucose that the ethylene oxide and propylene oxide reagents reacted
units (AGUs) containing three methyl substituents would phase preferentially with the free secondary hydroxyl groups on the
separate upon heating of the aqueous solution. This observation cellulose backbone rather than with any free hydroxyl groups on
lends support to the hypothesis that phase separation in methyl- the hydroxyalkyl side chains. After detritylation, the 2,3-O-
cellulose solutions is due to the aggregation of these trisubstituted hydroxylethyl- and 2,3-O-hydroxypropyl-cellulose were soluble
AGUs. However, phase separation from aqueous solution at in water at a molar degree of substitution of only 0.3 and 0.8,
elevated temperatures was also observed in under-methylated respectively. For comparison, conventionally synthesized hydro-
cellulose samples (DSmeth < 2.0) containing many monomethylated xypropyl cellulose derivatives require a molar degree of substitu-
AGUs but no trimethylated AGUs, suggesting that monomethy- tion of 4.0 for water solubility. Regioselectively substituted 2,3-O-
lated AGUs are also prone to aggregation. The phase separation carboxymethylcellulose was synthesized41,51 by reacting trityl
occurred at a higher temperature in the under-methylated cellulose with sodium monochloroacetate, followed by detrityla-
samples compared with the overmethylated samples. No phase tion with hydrogen chloride in methanol. The highest carboxymethyl
separation was observed in the pure 2,3-di-O-methyl-cellulose DS obtained was 1.91. It was reported that a DS of at least 0.6 was
1958 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

required for the polymer to be water-soluble, compared with DS 2.1.2. Regioselective Silylation. Silylation with thexyldimethyl-
0.4 for conventionally synthesized carboxymethylcellulose. How- chlorosilane ((CH3)2CHC(CH3)2Si(CH3)2Cl; TDMS-Cl) has
ever, Liu et al.52 reported that using the same synthetic proce- proven to be another effective protecting group strategy for
dures as the previous papers DS of only 0.3 was required for water regioselective synthesis of cellulose derivatives, and one that is in
solubility. some respects superior to tritylation because the specific sites
Tritylation chemistry has also facilitated the study of structure that are protected can be controlled by the dispersity of the
property relationships of regioselectively synthesized cel- cellulose in the reaction mixture. In addition, the regioselectivity
lulose esters. To synthesize these compounds, trityl cellulose of this reaction can be higher than that achieved by tritylation,
was esteried at the secondary hydroxyl groups by reaction with and it is a route not only to 6-O-protected derivatives but also to
an appropriate carboxylic acid anhydride. The trityl protecting 2,6-O-diprotected derivatives. Under heterogeneous conditions
group was then removed by hydrogen bromide in a mixture of with swollen cellulose, TDMS-Cl has been shown to react
acetic acid and chloroform, and the regenerated primary hydro- exclusively at the C-6 position forming TDMS ethers. Homo-
xyl group was esteried with a dierent carboxylic acid anhydride geneous solution reactions between cellulose and TDMS-Cl, on
to produce the desired cellulose-2,3-O-A-6-O-B type triester. 1H the other hand, are known to protect both C-2 and C-6 hydroxyl
NMR shifts for cellulose esters were assigned using 6-O-acetyl- groups (Figure 3). The first reported thexyldimethylsilation of
2,3-di-O-propanoyl-cellulose and 6-O-propanoyl-2,3-di-O-acet- cellulose was performed in dimethylformamide saturated with
yl-cellulose synthesized via trityl cellulose intermediates.53 Single ammonia at 15 C, resulting in 6-O-thexyldimethylsilylcellu-
crystals of cellulose 2,3-di-O-acetyl-6-O-propionyl cellulose, 6-O- lose with a DS of 0.99.70 The selectivity of the reaction for the C-6
acetyl-2,3-di-O-propionyl cellulose, 2,3-di-O-acetyl-6-O-butyryl hydroxyl group was shown by NMR analysis of the permethy-
cellulose, and 6-O-acetyl-2,3-di-O-butyryl cellulose were pre- lated samples. Subsequent heterogeneous reactions of cellulose
pared from solutions in dibenzyl ether and n-tetradecane.54 with TDMS-Cl were run at 25 C in ammonia-saturated
The crystal structures of both 6-O-acetyl-2,3-di-O-propionyl- N-methylpyrrolidone but resulted in products with a lower
cellulose55 and 6-O-propionyl-2,3-di-O-acetyl-cellulose56 were DS.7173 Homogeneous thexyldimethylsilation reactions were
determined by X-ray and electron diraction analyses, and the first run on cellulose dissolved in DMAc/LiCl with pyridine,
samples were further examined by atomic force microscopy57 resulting in products with a DS as high as 1.90.72 The DS and
and DSC.58 Regioselectively acetylated trityl cellulose was also distribution of the TDMS ethers along the cellulose backbone for
used to investigate the solution behavior of cellulose esters in products produced under both homogeneous and heteroge-
polar solvents.59,60 It was observed that the position of the ester neous reactions were confirmed by HPLC analysis of samples
substituents has a large inuence on the cellulose chain con- that were permethylated and then acid hydrolyzed. Imidazole
formations and associations in solution. Kasuya et al.61 investi- was subsequently found to be a more effective base than pyridine
gated 6-O-acetyl-2,3-di-O-benzoyl-cellulose and 2,3-di-O-acetyl- for homogeneous derivatization of cellulose with TDMS-Cl,
6-O-benzoyl-cellulose as stationary phases for chiral separations allowing the synthesis of 2,6-di-O-thexyldimethylsilyl cellulose
and found that the location of the specic substituents signi- with a DS of 2.0,74 with near-perfect 2,6-regiospecificity.
cantly aected chiral discrimination. Silyl protecting groups can be completely removed by treat-
Cellulose derivatives with unconventional side groups synthe- ment with tetrabutylammonium uoride, opening up a relatively
sized via tritylation chemistry have been studied to determine simple pathway to the synthesis of cellulose ethers substituted
their physical properties for use in a few advanced applications. exclusively at the C-3 position. Table 2 lists all 3-O-substituted
Several papers have described investigations of the liquid crystal- cellulose ethers whose syntheses have been reported in the
line properties of regioselectively substituted cellulose derivatives literature to date as well as their reported solubility in various
in organic solutions. Harkness and Gray62,63 found that 6-O- solvents. A series of 3-O-alkyl cellulose ethers (methyl, ethyl,
trityl-2,3-di-O-ethyl-cellulose, 6-O-trityl-2,3-di-O-benzyl-cellulose, n-propyl, n-butyl, n-pentyl, isopentyl, dodecyl) has been synthe-
and 6-O-R-(l-naphthylmethyl)-2,3-di-O-pentyl-cellulose formed sized by treatment of 2,6-di-O-thexyldimethylsilyl cellulose with
chiral nematic phases in several dierent organic solvents and the appropriate alkyl iodide or alkyl bromide in the presence of
that the optical properties could be controlled by varying the strong base, followed by deprotection of the two and six hydroxyl
structure and DS of the hydroxyl substituents. Cellulose deriva- groups.7478 One interesting nding within this series was that
tives with carbanilate64,65 and poly(ethylene oxide) substituents66 3-O-ethyl and 3-O-n-propyl cellulose were both soluble in water,
synthesized from trityl cellulose have also been studied for though they would phase separate at temperatures above 58.5
their chiroptical properties. 6-O- and 2,3-di-O-octadecyl-cellu- and 15 C, respectively. All other 3-O-alkyl ethers were not
lose were successfully synthesized, cast into LangmuirBlodgett soluble in water. The hydrogen bonding structure in 3-O-methyl
lms, and then studied by atomic force microscopy and X-ray cellulose was studied by Kondo et al.79 3-O-Allyl cellulose was
diractometry.67 Cellulose modied with carbazole side groups synthesized by the reaction of 2,6-di-O-thexyldimethylsilyl cellu-
was investigated for potential use in organic light emitting diode lose with allyl chloride followed by desilylation.74 This reaction
applications, and it was discovered that placement of the was later used to produce 3-O-hydroxypropyl cellulose with a
carbazole groups at the C-2/3 positions versus the C-6 position molar substitution of 1.0 by hydroboration of the allyl double
impacted electronic properties of the polymer.68 The thermal bond using 9-borabicyclo[3.3.1]nonane and subsequent alkaline
properties of poly(ethylene glycol) grafted cellulose were studied oxidation with hydrogen peroxide.80 3-O-Hydroxyethyl cellulose
for their potential usefulness in phase change materials for energy with a molar degree of substitution of 1.0 has also been
storage.35 Kondo et al.69 synthesized cellulose-2,3-O-dicinna- synthesized via 2,6-di-O-thexyldimethylsilyl cellulose.81 Both
mate, cellulose-6-O-monoacetate-2,3-O-dicinnamate, and cellu- 3-O-hydroxyethyl cellulose and 3-O-hydroxypropyl cellulose
lose 2,3-O-diacetate-6-O-monocinnamate using trityl protected were found to be water-soluble. As a gateway to producing
cellulose, introducing the cinnamate group to impart photosen- cellulose regioselectively substituted with dendritic substituents,
sitive and electroconductive properties to the polymer. 3-O-propargyl cellulose was synthesized.82 The terminal triple
1959 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

Figure 3. Cellulose heterogeneous and homogeneous thexyldimethylsilation reaction products.

Table 2. Solubility of 3-O-Substituted Cellulose Ethers with The 3-O-allyl cellulose was methylated at the two and six
Free 2- and 6-Hydroxyl Groupsa hydroxyl groups, and then the allyl group was removed by
reaction with palladium chloride. The resulting 2,6-di-O-methyl
3-O-substituentb soluble in ref
cellulose was insoluble in water and common organic solvents,
methyl insoluble in water and common organic solvents 74 contrasting with the isomeric 2,3-di-O-methyl cellulose prepared
ethyl water below 58.5 C, DMSO, DMAc, DMF 76 via tritylation that is soluble in water.34
n-propyl water below 15 C, DMF, DMSO 77 Silylation of cellulose with tert-butyldimethylchlorosilane has
n-butylc insoluble in water 75
also been studied with respect to its regioselectivity. However, it
has been shown to be somewhat less selective for the substitution
n-pentyl MeOH, EtOH, THF, DMSO, DMF, 78
of the 2- and 6-hydroxyl groups than TDMS-Cl, as some 3,6-di-O
NMP, DMAc
functionalized AGUs were also detected after derivatization.
isopentyl MeOH, EtOH, dioxane, 78 Additionally, at high reaction temperatures, some trifunctiona-
THF, DMSO, DMF, NMP, DMAc lized AGUs were formed.86
2.2. Esterification. For the preparation of regioselectively
allyl DMSO 74 substituted cellulose esters, the simplest route would of course be
hydroxyethyl water 81 direct solution esterification of cellulose. As befits the commer-
hydroxypropyl water, DMSO, DMAc, DMF 80 cial importance of cellulose esters, many methods have been
2-methoxyethyl water, DMSO, DMAc, NMP 76 developed that attempt to improve regioselectivity versus con-
propargyl DMSO 82 ventional heterogeneous esterification methods. The most im-
ethylene glycold insoluble in common organic solvents 83 portant approaches and their successes and limitations are
dodecyl THF 78
reported here.
a 2.2.1. Direct Reaction with Activated Acyl Moieties. Since the
DMSO = dimethyl sulfoxide; DMAc = N,N-dimethylacetamide;
DMF = N,N-dimethylformamide; NMP = N-methylpyrrolidone;
advent of cellulose solvents suited for chemical modification
THF = tetrahydrofuran; MeOH = methanol; EtOH = ethanol. b DS reactions of cellulose approximately 30 years ago, one of the first
1.0, except when otherwise noted. c DS = 0.8. d DS = 0.5. explorations of each solvent system has been its suitability and
selectivity for cellulose esterification reactions, especially for the
reaction of cellulose with activated acyl moieties (carboxylic acid
bond of the propargyl group could then undergo a copper- chlorides, acid anhydrides, and others) to make esters of cellulose
catalyzed Huisgen reaction (click chemistry) with azido-pro- with organic acids. These studies can be summarized by saying
pyl-polyamidoamine (PAMAM) dendrons to produce 3-O-(4- that they have for the most part either indicated modest posi-
methyl-1-N-propyl-PAMAM-[1,2,3-triazole]) cellulose (Figure 4). tional selectivity or have not addressed the issue of positional
Cellulose derivatives with ethylene glycol chains of varying selectivity at all. It is clear that reactions of cellulose in commonly
lengths selectively attached at the 3-hydroxyl group were synthe- used cellulose reaction solvents with the relatively nonbulky
sized with a DS of 0.5.83 These derivatives were of interest to acylating reagents that are of greatest interest (e.g., acid chlorides
Kadla84 for producing cellulosic materials with a honeycomb-like or anhydrides of acetic, propionic, butyric, and hexanoic acids)
structure. give mixtures of acylated cellulose regioisomers. The limitations
The rst synthesis of 2,6-di-O-methyl cellulose from natural of relying on the steric hindrance of the esterification reagents to
cellulose was achieved using 3-O-allyl cellulose as an intermediate.85 drive regioselectivity were explored in a recent paper from the
1960 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

Figure 4. Synthesis of dendritic cellulose derivatives from 3-O-propargyl cellulose.

authors laboratory.87 Direct esterification of cellulose with reported that uncatalyzed formylation of regenerated cellulose
several very sterically demanding acylating reagents was ex- and H2SO4-catalyzed formylation of crystalline cellulose af-
plored; pivaloyl chloride, adamantoyl chloride, and 2,4,6-tri- forded CF of DS 2.91 They examined these products by 13C
methylbenzoyl chloride, employing all three important classes NMR and concluded that the O-6 positions were fully substi-
of cellulose reaction solvents (DMAc/LiCl, DMSO/TBAF, tuted and that the formylation preference was O6 > O2 > O3.36
[Amim]Cl) at the lowest practical reaction temperatures. Con- More recently, formylation under relatively mild conditions has
ditions for synthesis of cellulose monoesters were identified in been investigated by Schnabelrauch and coworkers, for example,
many cases, but product analysis indicated that whereas O-6 by reaction with formic acid at room temperature with phos-
substitution was consistently preferred, no conditions could be phoric acid catalysis.92 The authors were able to obtain low DS
identified in which it was exclusive; O-2/3 substitution was CF products with DMSO solubility, for example, with DS 1.0
always observed at significant levels even below DS 1.0. It should after 10 h; previous workers had found poor solubility for low DS
be noted that in 2009, Zhang and coworkers reported synthesis CF products. The authors report that the regioselectivity of this
of cellulose benzoate with unusually high O-6 regioselectivity.88 reaction is O-6 > O-2/3, as determined by 13C NMR analysis of
Microcrystalline cellulose (MCC) was treated with benzoyl the CS products; in a DS 0.6 sample, they interpreted the 13C
chloride in an ionic liquid ([Amim]Cl) at 80 C for 2 h without NMR spectrum to indicate nearly no O-2 substitution because of
any base or catalyst to give a product with DS ca. 1. The 13C the absence of an upfield C-1 resonance due to the neighboring
NMR spectrum was interpreted to indicate reaction exclusively substituent. It must be noted that 13C NMR is one of the few
with the primary 6-OH group, within the 13C detection limits. It available tools for determining the position of substitution in
is difficult to rationalize such high selectivity with the lower cellulose esters but cannot be regarded as either quantitative or
selectivity results obtained in the same solvent using a bulkier, definitive because of potential signal overlap, limited sensitivity,
substituted benzoyl chloride (mesitoyl) in the work cited above.87 and nuclear Overhauser effects.93 In 1996, Liebert et al. reported
We believe that regioselective substitution strategies that rely only cellulose formate with DS 2.2 and complete substitution of the
on acylation with bulky activated carboxylic acid derivatives are primary OH groups using a convenient method consisting of the
unlikely to be successful for most common, practical ester types. acylation of cellulose with mixtures of formic acid and phos-
(Perhaps only a very large dendritic acyl moiety, for example, could phorus oxychloride within 4 h at room temperature.94 Reaction
afford more or even complete O-6 selectivity.) of this CF with NaOH powder in DMSO, then with chloroacetic
2.2.2. Cellulose Formates. Researchers have explored the acid, afforded carboxymethyl cellulose (CMC) with an unusual
synthesis of easily hydrolyzed cellulose esters, ideally under mild substitution pattern, with much higher O-6 substitution than is
conditions that might permit regioselective substitution, to sup- attained by conventional carboxymethylation of cellulose. Work
port a strategy of protection of one hydroxyl (typically O-6), by Philipp and coworkers, however, highlighted a limitation of
derivatization of the other hydroxyls (typically O-2 and O-3), and the formate group as a protecting group for cellulose regioselec-
selective removal of the labile ester group. Cellulose formate (CF) tive substitution.95,96 Sulfonation of cellulose formate of DS 2.5,
has some attractive properties for this role. Reaction of cellulose followed by hydrolysis of the formate groups, affords cellulose
with formic acid to prepare CFs has been known for some sulfates of considerably higher DS than that of the DS(OH) in
time.89,90 It has also been shown that the formate groups are quite the starting ester. Clearly, deformylation is occurring; formate is
labile; for example, they are easily removed in boiling water. (A DS not a stable protecting group under sulfonation conditions.
2.4 CF was completely deformylated within 10 h in boiling water.)91 2.2.3. Cellulose Trifluoroacetates. Cellulose trifluoroacetate
With regard to formylation regioselectivity, Fujimoto and coworkers (CTFA) is an interesting intermediate for the synthesis of
1961 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

selectively substituted cellulose ether and ester derivatives, in


part because the lability of the trifluoroacetate groups toward
mild alkaline hydrolysis makes them potentially useful protecting
groups. The use of trifluoroacetate groups as protecting groups
became more promising upon reports of regioselectivity in the
trifluoroacetylation. Liebert and coworkers reported97 that so
long as moisture is excluded and the trifluoroacetic anhydride
(TFAA) content is kept below 50%, acylation of cellulose with a
mixture of trifluoroacetic acid/trifluoroacetic anhydride is suc-
cessful, affording CTFA with DS 1.5. They found that at DS 1.5
trifluoroacetate (the maximum achievable without cosolvents),
complete substitution at O-6 was observed, in addition to partial
substitution at the secondary hydroxyls. It must be noted that this
conclusion was partially the result of 13C NMR spectral analysis
of the product, which was not quantitative; the published
spectrum also contains a number of unassigned resonances.
The researchers also attempted to quantify position of substitu-
tion by chemical degradative means. Mild methylation of
the CTFA (DS 1.5) with methyl triflate in trimethylphosphate,
deacylation with base, hydrolysis to monosaccharides, and
analysis of the monosaccharides by HPLC gave indications of
O-6 and O-2 trifluoroacetylation selectivity.13,98,99 CTFA can be
prepared under quite mild conditions, suitable for maximum
regioselectivity, for example, by dissolving cellulose in trifluor-
oacetic acid, affording CTFA with DS 1.21 to 1.34 after 25 Figure 5. Partial DS of cellulose tosylate groups at C-6 (square
days.100102 The degree of O-6 selectivity is not entirely clear, symbols) and at C-2/C-3 (round symbols) versus total DS tosylate.
(Reprinted with permission from ref 110. Copyright 1996 Wiley-VCH
and the hydrolytic instability of the CTFAs probably contributes Verlag GmbH & Co. KGaA.)
to the difficulty in this determination.97 Other workers used the
DS 1.5 CTFA as a protected intermediate for the regioselective
synthesis of cellulose-3-O-sulfate,103 by sulfonation with SO3/ Cellulose tosylates with DS from 0.4 to 2.3 were synthesized
DMF, followed by alkaline saponification of the trifluoroacetates. by increasing the equivalents of tosyl chloride added to the
No detailed evidence was presented to indicate the degree of reaction. Those with a DS of 0.9 or above were soluble in polar
regioselectivity obtained. Liebert and coworkers used CTFA as a aprotic solvents such as DMSO, DMAc, and DMF. Tosylates
reaction intermediate for CMC, using alkylation in DMSO by with DS to 1.4 are also soluble in acetone, dioxane, and
similar methods to those described in the same paper for CF.94 tetrahydrofuran. Further increasing the DS(tosyl) above 1.8
As observed with CF, unconventional CMC substitution pat- expanded the solubility range to halogenated hydrocarbons like
terns were also observed using CTFA as starting material. chloroform and dichloromethane. Evaluation of the selectivity of
2.2.4. Tosylation. The synthesis of sulfonic acid esters of the tosylation reaction was performed by displacing the tosyl
cellulose, such as methanesulfonate and benzenesulfonate deri- groups at C-6 with iodide. The investigators calculated tosylation
vatives, has long been used in synthetic cellulose chemistry, but at the secondary OH groups by assuming that the primary
the reaction with p-toluenesulfonyl (tosyl) chloride to form tosyl tosylates were completely substituted by iodide and that no
esters of cellulose has been the most frequently studied of the secondary tosylates were displaced by iodide. They then sub-
group. Recently, tosylation has been widely used to synthesize tracted DS(I) in the products from DS(Ts) in the starting
new cellulose derivatives substituted at the C-6 position.104106 cellulose tosylates and assumed that the dierence corresponded
The benefit to this approach is that in addition to acting as a to the DS of tosylates at O-2 and O-3. This analysis indicated that
protecting group, the tosyl moiety is also a good leaving group, tosylation occurred at C-6 almost exclusively on derivatives with
enabling the synthesis of a variety of 6-deoxy-cellulose derivatives a low DS (0.46) but that increasing the DS resulted in substitu-
via nucleophilic displacement reactions. There are reports of tion occurring also at the secondary hydroxyl groups of cellulose.
cellulose tosylation going back several decades32 in the literature, Tosylation of C-6 was nearly complete in derivatives with a DS of
but use of the reaction in synthetic cellulose chemistry increased 1.4. The graph in Figure 5, from Rahn,110 shows the relation-
substantially after it was effectively demonstrated with a homo- ship between the partial DS values (at O-6 and at O-2/3) and the
geneous solution of cellulose in DMAc/LiCl.107 A subsequent total DS. However, it must be noted that these investigators
investigation found that using triethylamine as an organic base in found by 13C analysis of the iodotosylates that the displacement
the reaction rather than pyridine resulted in a higher DS for the of tosyl by iodide was not completely selective for C-6 either
tosyl group and helped avoid side reactions (Cl displacement of because the product contained a small amount of iodide at C-2
tosylate) that formed chlorodeoxycellulose.108 Recently, homo- and C-3 as well. Therefore, whereas substitution reactions of
geneous tosylation of cellulose was also demonstrated in the tosyl cellulose occur primarily by an SN2 mechanism at C-6, they
ionic liquid [Amim]Cl.109 can occur at C-2/3 under the right conditions; whether by SN2 or
Rahn et al.110 published an in depth study of the homogeneous SN1 mechanisms is unclear. Increased selectivity for tosylation at
tosylation of cellulose in DMAc/LiCl investigating how to O-6 was reported for reactions carried out in [Amim]Cl.109 The
control the total DS(tosyl), the solubilities of cellulose tosylates regioselectivity analysis, performed by 13C NMR, indicated
of varying DS, and the selectivity of the reaction for O-6. that tosyl cellulose with DS up to 0.84 was substituted almost
1962 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

Figure 6. Synthesis of 6-aminocellulose via cellulose tosylate.

exclusively at O-6, although substitution at O-2 became evident biomaterials.125 This methodology is useful for cases in which
at higher DS. Therefore tosylation is not as selective for O-6 as the desired 6-substituents are stable to the conditions required to
are the reactions with sterically demanding alkyl or silyl halides reduce o the 2,3-tosylate groups and is a clever way to circumvent
that we have discussed, such as tritylation. Despite the lack of the selectivity limitations of tosylation. Liebert et al.126 also
complete regioselectivity of tosylation for O-6 in cellulose, its synthesized 6-azido-6-deoxy-cellulose from tosyl cellulose, but
frequent use in the synthesis of regioselectively derivatized instead of reducing it to an amine, the azido groups were used
cellulose compounds over the past decade and a half warrants its as sites for Huisgen reactions. Using this chemistry, 1,4-disubsti-
inclusion in this review. tuted 1,2,3-triazoles can be used as linkers to the C-6 position for a
The thermal stability of tosyl cellulose with DS from 0.4 to 2.3 wide variety of functional groups. Other nitrogen nucleophiles that
was studied by Heinze et al.,111 and it was found that the have been reacted with tosyl cellulose include methylamine,127
compounds began to degrade between 169 and 196 C. Tosyl butylamine,128 and pyridine.128 In 2008, the Heinze group reported
cellulose was also investigated by dielectric spectroscopy to reaction of 6-O-tosyl cellulose with DS = 0.58 (and with high O-6
determine the inuence of DS on the polymer chain mobility.112 regioselectivity at that low DS(Ts)) with propargyl amine. This
The free OH groups of tosyl celluloses were reacted with a variety aorded an alkyne precursor for click reactions that were employed
of aliphatic, aromatic, and unsaturated acid anhydrides as well as to create dendritic structures.129
with isocyanates to form new compounds with a wide range of 2.2.5. Cellulose Esters via Silyl-Protected Cellulose. Silyl ethers
solubilities.113 of cellulose have been exploited for the regioselective synthesis of
The most widely studied applications for tosyl cellulose have cellulose esters partially because of the ready organic solubility of
been in displacement reactions with nitrogen-containing nucleo- these silyl cellulose derivatives, sometimes even in nonpolar
philes. A series of publications have examined the displacement organic solvents.130 Two strategies have been employed, both
of tosyl groups at the C-6 position of cellulose with aliphatic and starting with regioselectively silyl-protected cellulose ethers:
aromatic diamines for the purpose of creating lms capable of reaction of the free OH groups with an acylating reagent under
enzyme immobilization for biosensor applications.114119 Mais mild conditions in the presence of an organic base or re-
et al.120 synthesized cellulose-graft-poly(N-acetylethylenimine) placement of the silyl groups with acyl groups by reaction with
from tosyl cellulose by reaction with 2-methyl-1,3-oxazoline. acyl halides in the absence of base, often at higher temperatures
Bieser et al. recently published121 a study on the ecacy of using (Figure 7).
similar cellulose graft copolymers synthesized from tosyl cellu- Most of the published work has involved trimethylsilyl cellulose,
lose as antimicrobial coatings. Chirality has been introduced to even though trimethylsilyl chloride is not suciently sterically
the C-6 position of cellulose by the substitution of tosyl groups demanding to give highly 6-OH-selective etherication.131,132
with single enantiomers of 1-phenylethylamine.122 Water-solu- Therefore, the silylcellulose ester and cellulose ester products
ble cationic cellulose derivatives were made by reacting tosyl synthesized from these trimethylsilylcellulose ethers are of only
cellulose with dierent trialkylamines.123 A group in Aachen modest regioselectivity. The DS of the cellulose esters obtained
developed an interesting synthetic approach to truly regiospeci- was either roughly equal to the DS of the trimethylsilyl cellulose
cally substituted cellulose derivatives from tosyl cellulose starting materials, via direct replacement of the silyl groups, or
(Figure 6);124 they rst reacted cellulose that had been fully proportional to the DS of free OH groups of these trimethylsilyl
tosylated at C-6 with sodium azide. The resulting azido group at cellulose derivatives, by way of acylation of the remaining OH
C-6 was then reduced with lithium aluminum hydride to a groups; as detailed below, which pathway is taken can be
primary amine, which in the process reductively removed any manipulated by choice of reaction conditions.106
tosyl esters at the C-2 and C-3 positions on cellulose to generate Klemm et al.133 used trimethylsilyl cellulose containing several
free hydroxyl groups. The resulting 6-amino-6-deoxy-cellulose levels of trimethylsilyl substitution (DS 1.99, 2.46, 2.62) to make
was then used to make both 6-N-sulfonated and 6-N-carboxy- cellulose esters by direct displacement of the silyl groups using
methylated cellulose. These derivatives were tested for their acid chlorides (e.g., acetyl, propionyl, and stearoyl chlorides)
platelet adhesion properties for potential use as coatings for under acidic conditions. The advantage for such reactions is that
1963 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

Figure 7. Acylation of cellulose silyl ethers.

no solvent or catalyst was used, which is potentially economical film-forming materials, viscosity regulators, anionic polyelectro-
and aords a rather pure product after a simple work up lytes, and their biological activity.125,135137 Particularly for
procedure; more pertinent to this review is that acylation biological applications such as use in killing viruses138140 or
occurred exclusively at the trimethylsilyl-substituted cellulose preventing blood clotting,141,142 the ability to synthesize CS
hydroxyl groups, whereas the unsubstituted hydroxyl groups regioselectively may be essential for optimum activity and for
surprisingly did not react under these reaction conditions. This regulatory approval. The regioselectivity of CS synthesis has
was true even in the presence of excess acid chloride. The result is been reviewed several times26,125,135,143 in the past two decades.
that the product cellulose ester substituents are in the same A variety of reagents and solvent systems have been developed in
positions as the trimethylsilyl groups of the starting material. search of regioselective substitution; these are detailed in this
Similar substitution reactions were demonstrated with nitro- or section. The Mischnick laboratory has been a leader in develop-
bromobenzoyl chlorides.96 Clearly, if one can place the trialk- ing analytical techniques for determining the position of sub-
ylsilyl groups regioselectively, this could be an interesting route stitution of polysaccharide derivatives, especially polysaccharide
to regioselectively substituted cellulose esters. Remarkably, to ethers.144 Mischnick and coworkers have developed a method for
our knowledge, no publications have appeared describing appli- determination of the CS substitution pattern.145,146 This method
cation of these techniques to more regioselectively substituted is based on methylation of the CS (MeI, LiCH2SOCH3, DMSO,
silyl ethers of cellulose, such as the 6-thexyldimethylsilyl and 2,6- tetramethylurea), replacement of sulfate groups by acetate
bis(thexyldimethylsilyl) ethers mentioned previously. No evi- (Ac2O),147 replacement of acetate groups by deuteromethyl
dence has been published to indicate whether the thexyldi- ethers (NaOH/CD3I), and hydrolysis of the methylated CS to
methylsilyl group can be directly replaced by acyl groups under monomers (TFA/H2O/N-methylmorpholine-borane complex).
acidic conditions, as can the trimethylsilyl group. Their results from cellulose sulfates synthesized in various ways
An interesting example of reaction of regioselectively pro- are listed in Table 3. The ability to analyze the position of
tected silyl ethers of cellulose with acylating reagents was substitution of CS quantitatively, as with cellulose ethers, is
published recently by Luning and Schultz.134 They described extremely useful, and it is surprising that it has not been
reaction of 6-thexyldimethylsilylcellulose (and also 6-tritylcellulose) frequently employed to date by researchers in the CS field. Most
with diphenylketene. Ketenes are highly reactive toward alcohols, researchers report only 1H and 13C NMR spectra to support
aording esters without the generation of low-molecular-weight assignments of position of substitution; these techniques are
byproduct.25 These workers observed clean reaction with diphe- limited in the precision and accuracy of their ability to quantify
nylketene to aord cellulose diphenylacetate esters of DS 1 and position of substitution generally and especially so for cellulose
also observed that this DS did not increase even if reaction time sulfates in which the substituent is not directly measured but only
was extended, or a larger excess of reagent was used. They inferred from the shifts in nearby carbon and proton resonances.148
concluded that reaction was proceeding at either the 2- or 3-OH CS can be synthesized by direct cellulose sulfation hetero-
group, with high regioselectivity. Unfortunately, they relied geneously in pyridine or homogeneously in ionic liquids. In 2000,
heavily on IR spectroscopy to characterize the products, did Baumann et al. reported the synthesis of cellulose-2,6-sulfate
not report NMR analysis of the derivatives produced, and with complete regioselectivity (DSS = 2.0, Figure 8) by the
therefore could draw no conclusions about the actual position reaction of preswelled cellulose with a complex of pyridine and
of substitution. Still, this is a promising general approach that chlorosulfonic acid (ClSO3H).103 Such high regioselectivity as a
deserves more attention; the reactive ketenes may permit the use result of direct reaction of cellulose with a sterically undemanding
of low reaction temperatures, enhancing regioselectivity. reagent is surprising; the only analyses reported by the authors
2.2.6. Cellulose Sulfate. Cellulose sulfates (CS) have been that could help elucidate position of substitution were 1H and
13
of increasing interest in recent years because of their utility as C NMR spectra. In 2009, Wang et al. reported homogeneous
1964 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

Table 3. Cellulose Sulfate Monomer Composition versus Synthesis Method

weight % monomer by sulfate group position


entry starting material synth. conditions none 2 3 6 23 26 36 236 DS

1 cellulose N2O4/DMF/SO3 51.3 2.2 1.4 42.0 0.2 1.8 0.7 0.4 0.5
2 cellulose propanol/H2SO4a 27.9 2.3 1.0 61.7 0.1 5.1 1.6 0.4 0.8
3 cellulose DMF/Ac2O/ClSO3Hb 19.9 6.0 1.2 50.8 0.3 20.4 0.9 0.4 1.0
4 CA DS 1.5 DMF/NH2SO3H 28.9 13.5 6.9 6.1 25.5 5.0 1.9 12.1 1.3
5 CA DS 2.5 DMF/ClSO3H 69.1 5.3 4.4 17.3 0.4 1.9 1.4 0.3 0.4
6 CA DS 2.0 DMF/ClSO3H 42.0 7.4 6.1 29.5 0.5 7.6 6.2 0.8 0.7
a
Heterogeneous (all others homogeneous). b After solvolysis of acetyl groups.

Figure 8. Regioselective cellulose sulfonation.

sulfation of bagasse cellulose in the ionic liquid [Bmim]Cl at unquantied) regioselectivity and ascribed this to activation of
30 C using ClSO3H/DMF complex as sulfating agent, with modest the O-3 position by the neighboring triuoroacetyl group.103
regioselectivity as determined by 13C NMR (sulfate distribution Cellulose formate and acetate also can be used as intermediates
C-6 > C-2 > C-3).149 to synthesize cellulose sulfate.95 Partially regioselective hydro-
A widely investigated and direct pathway starting from cellu- lysis at O-2/3 positions of the commercial CTA DS 2.4 to 2.6 is
lose is acetosulfation, followed by deacetylation.150152 In 1995, reported to aord a 6-O-acetate intermediate, which can be sulfated
Philipp et al. rst reported the synthesis of CS by the reaction of using NH2SO3H, followed by deacetylation to form cellulose-
cellulose in system of ClSO3H/Ac2O/DMF at 50 C for 324 h, (2/3)-sulfate, where the regioselectivity is O-2 > O-3.150,156
followed by deacetylation in NaOH solution.150 They report that Direct, acid-catalyzed replacement of silyl ether groups, ana-
the regioselectivity of this reaction is C-6 > C-2 . C-3 (entry 3, logous to the direct replacement by acyl groups previously
Table 3). Cellulose-6-sulfate with DS around 0.8 was reported. described, can also lead to CS.150,157,158 The regioselectivity of
In 2008, Hettrich et al. studied the eect of various solvents sulfation of trimethylsilylcellulose with SO3 or ClSO3H in DMF
and sulfating agents on this reaction and reported that at DS < 1, is O-6 > O-2 > O-3.150,159 The literature examples so far involve
C-6 substitution is favored within a range of solvents (DMF, the use of trimethylsilylcellulose, which, as we have noted,
DMAc, NMP) and reagents (ClSO3H, H2SO4, SO2Cl2, SO3- typically contains a mixture of regioisomerically substituted
DMF), as measured by 13C NMR.153 All of these routes that AGUs. Sulfation of trimethylsilylcellulose (DS 1.55) with
either begin with cellulose acetates or create cellulose acetates ClSO3H in DMF, followed by treatment with NaOH/EtOH, is
in situ (acetosulfations) are awed in that they depend on the reported to aord a nearly completely regioselectively substi-
regioselectivity of acid-catalyzed acetylation, which the best litera- tuted cellulose-6-O-sulfate (DS 0.95); only NMR results are
ture evidence indicates is rather random.10,11 Product distributions reported to indicate position of substitution. At higher DS values,
are generally quantied by 13C NMR techniques whose aws we C-2 and to a lesser extent C-3 are also substituted. In 2003,
have already mentioned. The modest selectivity shown by Mis- Richter and Klemm studied the regioselectivity of sulfation of
chnick and coworkers146 (Table 3) for samples generated by nearly fully substituted trimethylsilylcellulose (DS 2.9) using
acetosulfation reactions is cautionary evidence that NMR methods dierent SO3-complexes.157 The complexing agent employed
may lack accuracy for quantifying regioselectivity of CS samples. inuences sulfation regioselectivity to a moderate extent. Using
CS synthesis via organo-soluble intermediates may enhance SO3-DMF, CS with higher DS at O-6 could be obtained; when
regioselectivity by permitting milder reaction conditions. Using using SO3-triethylamine, a slight preference for O-2 sulfation was
bulky protecting groups to block selectively one or two hydroxyl observed. Cellulose trinitrite prepared from cellulose in situ has
groups (usually C-6 or C-2/6) is an eective way to obtain CS been carefully investigated as an intermediate for CS regioselec-
with high regioselectivity. For example, in 2000, Heinze et al. tive synthesis. Cellulose trinitrite formed in situ by treatment of
employed 6-O-(4-methoxy-triphenylmethyl)cellulose to synthe- cellulose with N2O4/DMF is reacted with various sulfating
size cellulose-2,3-sulfate by sulfation (SO3/pyridine), followed agents.150,158 Whereas product regioselectivity is generally quite
by deprotection.154 As previously noted in this Review, tritylation modest (by 13C NMR), it is an interesting example in that
of cellulose is quite selective for O-6; however, the sensitivity of position of substitution depends strongly on the sulfating reagent
trityl groups toward acid may complicate this approach. In 1995, used and on temperature. Regioselectivity was attributed to two
Klemm et al. reported synthesis of cellulose-3-sulfate (DS 0.98) counteracting eects: the high spatial accessibility of the O-6
using cellulose 2,6-triuoroacetate (DS 1.5) by reaction with nitrite group and the high intrinsic reactivity of the O-2 nitrite
SO3/pyridine/DMF at 15 C for 3 h.155 They reported high (but group.135 This synthetic route may therefore provide products
1965 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

with diering regioselectivity depending upon reaction condi-


tions. For example, with SO3 as sulfating agent, the regioselec-
tivity at room temperature (20 C) is O-6 > O-2, but it reverses at
low temperature (20 C) to O-2 > O-6. Room-temperature
sulfation with SO2 or NOSO4H gave high O-6 regioselectivity at
low total DS (0.310.36), much more selective than with SO3
under similar conditions.150,158
In summary, a number of routes for partially regioselective
synthesis of cellulose sulfates have been reported, and clearly
such regioselectively substituted cellulose derivatives would be
useful for investigations of properties and applications, particu-
larly biological eects like blood clotting and antiviral properties.
Unfortunately, there are no reports as yet of entirely regioselec-
tive CS synthesis, backed up by quantitative analytical methods
such as degradation and monomer analysis. It will be important
in the future to bring together the most powerful synthetic and
analytical techniques to enable truly regioselective syntheses of
these important cellulose derivatives. Figure 9. Mechanism for regioselective cellulose bromination from
Furuhata et al.160

3. DIRECT REGIOSELECTIVE HALOGENATION


Table 4. Cellulose Derivatives Synthesized from 6-Deoxy-6-
In 1992, Furuhata et al.160 published a method for regioselec- bromocellulose
tive halogenation of cellulose at the C-6 position without going
through an isolated intermediate like tosyl cellulose. This reac- C-6 substituent DSa reference
tion required the dissolution of cellulose in DMAc/lithium
sulfur nucleophiles and derivatives
bromide, followed by the addition of the reagents triphenylpho-
methanethiol 0.55 163
sphine (Ph3P) and N-bromosuccinimide (NBS). The 6-deoxy-6-
benzenethiol 0.41 163
bromo-cellulose product is analogous to the 6-O-tosylate in that
it can act both as a protecting group under certain conditions and mercaptoacetic acid 0.10 163
as a good leaving group for substitution chemistry at the C-6 L-cysteine 0.56 164
position. The advantage versus tosylation is that the bromination 2-mercaptoethanol 0.71 164
is completely selective for C-6. That is, no measurable amount of 3-mercaptopropanoic acid 0.92 164
bromination takes place at the secondary hydroxyl groups on 2-mercaptobutanedioic acid 0.51 164
cellulose. In addition, a maximum bromide DS of 0.98 has been 4-aminobenzenethiol 0.84 164
reported using this procedure, indicating that virtually every 2-aminobenzenethiol 0.76 164
primary hydroxyl group on the cellulose backbone has been 2-mercaptobenzoic acid 0.64 164
replaced by a bromide.161 Chlorination of cellulose dissolved in thiocyanate 0.88 167
DMAc/LiCl using Ph3P and N-chlorosuccinimide has also been
2-aminoethanethiol 0.79 165
reported,162 although it was found that chlorination also takes
isothiourea 0.80 165
place at the secondary hydroxyl sites for a maximum DS of 1.86.
The remarkable selectivity of the C-6 bromination is believed to mercaptan 0.79 165
be due to a combination of factors: the steric bulk of the sodium sulfonate 0.75 168
triphenylphoshonium intermediate, which leads it to prefer nitrogen nucleophiles and derivatives
reaction with the less hindered primary OH and the SN2 2-cyanoethylamine 0.86 166
mechanism of the subsequent bromide displacement reaction azide 0.96 161
(Figure 9). In the event that oxophosphonium intermediates do primary amine 0.96 161
result from reaction of the secondary OH groups, their SN2 a
Highest DS for the derivative reported in the literature is the DS
displacement by bromide will be disfavored both by the need for recorded in this Table.
the nucleophile to approach from the interior of the glucopyr-
anose ring and by the inversion of conguration at the 2- or
3-carbon that would result, causing the entire cellulose chain to Reactions involving various thiols were shown to be eective in
move in space. All of these energetic costs work against bromina- regioselectively attaching dierent pendent functional groups,
tion at the 2- and 3-positions. such as carboxylic acids, dicarboxylic acids, and amines, to the
The main disadvantage to regioselective bromination com- cellulose backbone. These included reactions carried out both
pared with tosylation is that the 6-deoxy-6-bromo-cellulose is not heterogeneously in aqueous alkali163 and homogeneously in
soluble in typical aqueous or organic solvents, although it can be DMAc/LiBr.164 Several of these derivatives were tested for
redissolved in DMAc/LiBr. Nevertheless, several highly regiose- their metal ion sorption capacities.165 6-Deoxy-6-(2-cyanoethy-
lectively substituted cellulose derivatives with substituents at C-6 lamino)-cellulose and 6-deoxy-6-thiocyanato-6-cellulose were
have been synthesized via nucleophilic displacement with synthesized beginning with a heterogeneous suspension of
6-deoxy-6-bromo-cellulose starting under heterogeneous condi- 6-deoxy-6-bromo-cellulose in DMSO and DMF, respectively.
tions. Table 4 provides a list of all compounds synthesized The reactions became homogeneous as they progressed, and
from 6-deoxy-6-bromocellulose, as reported in the literature. the products were used to study -relaxations in cellulose by
1966 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

Figure 10. Cellulose by cellulase-catalyzed polymerization.

dielectric spectroscopy.166,167 Water-soluble sulfonate deriva- been demonstrated that CTA oligomers have spectral properties
tives were synthesized by reacting the brominated cellulose in equivalent to the corresponding CTA high polymer and thermal
aqueous solutions of sodium sulte.168 A regioselectively sub- transitions (Tg, Tm) somewhat lower than those of the high
stituted cationic cellulose derivative was produced by synthesiz- polymer at a DP between 7 and 9.174 Note that throughout this
ing 6-deoxy-6-azido cellulose from the brominated cellulose, article we define DP as the number of glucose (or substituted
followed by reduction of the azido group to a primary amine of glucose) units because glucose is the chemical repeat unit of
DS 0.96.161 cellulose; cellobiose is the spatial repeat unit, but in this account
In summary, phosphine-catalyzed bromination at O-6 is a very we are concerned primarily with cellulose substitution chemistry,
attractive method for the regioselective, nearly regiospecic so the glucose repeat unit is the appropriate one. Readers of
modication of cellulose, and we expect that its utility as a the cellulose literature must be careful because both glucose
gateway method to enable access to other interesting, functional and cellobiose are used as repeat units when DP numbers are
cellulose derivatives will be the subject of further innovation in presented in different articles.
the coming years. The glycosylase-catalyzed synthesis of polysaccharides is a
promising and interesting development, which has been used for
the synthesis of a variety of natural (chitin,175 xylan,176 amylose,177
4. DE NOVO SYNTHESIS OF CELLULOSE DERIVATIVES mixed-linkage -glucans178) polysaccharides. It has even been
Whereas polymerization of monomers is the standard method used to make mixed linkage polysaccharides, such as a glucan
for preparation of synthetic polymers, all polysaccharides and containing alternating -1f4 and -1f3 linkages.179 It has the
their derivatives that are used in commerce are derived from advantage of requiring only a relatively simple activation of the
natural polysaccharides, which are isolated and puried from anomeric position (conversion to glycosyl uoride) and not
plant or animal sources.169 Therefore, it was a major break- requiring protection of the other hydroxyl groups. By proper
through in polysaccharide chemistry when Kobayashi and enzyme selection, the stereochemical and positional selectivities
coworkers170 synthesized cellulose from monomers for the rst are excellent, providing -1f4 linkages in the case of cellulase and
time, polymerizing cellobiosyl uoride using a cellulase enzyme R-1f4 linkages in the case of amylase, for example.
catalyst. Armed with this vision of what was possible, Nakatsubo The major limitation of enzyme-catalyzed polymerization for
and coworkers 5 years later171 succeeded in the rst purely synthesis of substituted cellulose derivatives is the ability of the
chemical synthesis of cellulose by cationic polymerization of a cellulase enzymes to accept substituted monomers. In the case of
glucose orthopivalate ester, later extending this method to cellulose, there are only two cases of which we are aware in which
include cellulose derivatives. These de novo synthesis methods unnatural substituents may be present; one is the synthesis180
have expanded the palette of regioselective synthesis methodol- of 1-deoxy-1-thio analogues of cellulose, which is not relevant to
ogy, broadening the variety of regioselectively substituted deri- our account, and the other is the synthesis of cellulose derivatives
vatives that can be synthesized.172 We will illustrate that variety containing methyl ethers, which is quite relevant. Kobayashi and
and give several important examples while also pointing out some coworkers found181 that by working with selectively methylated
apparent limitations to the current state of the art. -cellobiosyl uoride, they could by cellulase catalysis synthesize
4.1. Enzyme-Catalyzed Polymerization. Kobayashis pio- selectively methylated cellulose ethers. Using protective group
neering synthesis of cellulose from -D-cellobiosyl fluoride by chemistry well-known to carbohydrate chemists, they were able
enzymatic polymerization employing a cellulase enzyme has led to synthesize 6-O-methyl--cellobiosyl uoride. Polymerization
to a considerable exploration of this general method for the of this disaccharide using cellulase in acetonitrile and pH 5.0
synthesis of natural and unnatural polysaccharides.173 Manipula- acetate buer aorded a cellulose derivative in which alternating
tion of the monomer concentration and solvent system enabled monomers were methylated at O-6 (Figure 11).
the Tokyo group to successfully employ cellulase enzymes Note that by polymerizing a disaccharide substituted selec-
(whose natural function is to hydrolyze the intermonomer tively in only one of the two rings, the authors were able to obtain
anomeric linkages of cellulose as part of its biodegradation) to a product containing not only regiospecic substitution but also
form anomeric bonds in selective fashion in the reverse of the completely controlled monosaccharide sequence (a perfectly alternat-
natural process. The original synthesis (Figure 10) was carried ing AB copolymer). The product was a mixture of oligomers (DP
out in acetonitrile and pH 5 buffer (using a high percentage of ca. 814), and conversion to the cellulose acetates allowed
organic solvent minimizes hydrolysis of the product cellulose by conrmation by NMR that the linkages were uniformly -1f4.
the cellulase, whereas the presence of some water affords adequate In contrast, the closely related cellobiosyl uoride derivative
enzyme activity) and gave a reasonable yield (up to 65%) of diering only in that the methyl group is on the 6-OH of the
cellulose. The authors confirmed structure unequivocally by other ring gave much more modest polymerization in the
conversion to cellulose triacetate (CTA), whose NMR spectra presence of the cellulase, aording DP of 4 (although -1f4
were identical to those of CTA from natural cellulose. The selectivity was observed).
product was also a substrate for natural cellulase enzymes, giving Unfortunately, these are the only examples of which we are
the expected degradation products. The synthetic cellulose had aware in which tolerance by the cellulase for ether or ester
degree of polymerization (DP) > 22, as measured by SEC; it has substituents has been demonstrated, and of course the rather low
1967 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

Figure 11. Cellulase-catalyzed synthesis of alternating glucose-6-O-methylglucose copolymer.

Figure 12. Monomethylcelluloses via orthopivalate polymerization.

DP attained is also a drawback. It is to be hoped that future can give cellulose in which all three hydroxyls have substituents
investigations will show the way to broader substrate tolerance that are dierentiated by reactivity.
and higher DP products, permitting the use of the excellent 3 Because the starting monomer is substituted, the polymeric
selectivity and eciency of cellulase-catalyzed polymerization as products are substituted. This prevents excessive hydrogen
a general method for synthesizing regioselectively substituted bonding and the resulting crystallinity in the product,
cellulose derivatives. greatly enhancing product solubility versus the cellulose
4.2. Acid-Catalyzed Orthopivalate Polymerization. The oligomers produced by cellulase-catalyzed polymerization
Nakatsubo group demonstrated the first chemical synthesis of of -cellobiosyl uoride. As a result, product precipitation is
cellulose, attaining a DP of ca. 15, by polymerization of an much less likely to severely limit attainable product molec-
orthopivalate derivative. The Nakatsubo and Kamitakahara ular weight.
groups have demonstrated the value of this general synthetic These features of the Nakatsubo/Kamitakahara de novo
approach for a considerable number of regioselectively substi- synthesis method have permitted the remarkable synthesis of
tuted cellulose derivatives. The methodology has several distinct all six signicant possible homopolymers related to methylcellu-
and valuable features: lose. (Two of the eight related homopolymers are synthetically
1 Under the right conditions of catalyst, temperature, solvent, trivial; the homopolymer of the unsubstituted AGU is just
and substituent, the reaction is remarkably selective for cellulose, whereas the homopolymer of the trisubstituted AGU
-anomeric stereochemistry. is 2,4,6-trimethylcellulose, which does not require regioselective
2 Because the starting material is a monosaccharide, one can synthesis.) For example, Figure 12 shows the synthetic metho-
take full advantage of the sophisticated chemistries available dology used182 to prepare the key monomers for synthesis of the
for regioselective substitution of monosaccharides to eciently monosubstituted homopolymers 2-methyl-, 3-methyl- and
synthesize monomers with desired (and dierentiated) sub- 6-methylcellulose.
stituents at O-3 and O-6. Along with the O-2 pivalate Kamitakahara and coworkers were able either to methylate
substituent that arises from the polymerization reaction, this the monomer regiospecically prior to polymerization (3M, 6M)
1968 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

Table 5. Solubilities of 2,6-O-Methyl/ethylcelluloses190 a,b,c,d attractive possibility of attaching dierent ester substituents to
O-3 and O-6 in the monomer stage and then polymerizing to
solvent 2E6E 2E6M 6E2M
aord a regiospecically substituted cellulose ester is not a
H2O I I I practical approach. Because Nakatsubo and Kamitakahara have
MeOH PS I I already demonstrated the synthesis of cellulose ethers with fully
EtOH PS I I dierentiated substitution at the three hydroxyls (for example,
acetone PS I I
3-O-allyl-6-O-benzylcellulose-2-O-pivalate193), there is reason to
believe that this chemistry could be adapted to the regioselective
CHCl3 S S PS
synthesis of cellulose esters. The two drawbacks will be the need
MeOH/CH2Cl2 (1/4, v/v) S S PS
a
to carry out highly selective deprotection and esterication
Reprinted with permission from ref 190. Copyright 2009 Springer reactions on the polymer, which will be less conducive to
ScienceBusiness Media B.V. b Solubility was evaluated at a concentra-
reactivity and selectivity than the orthopivalate monomer and
tion of 5 wt %. c S: soluble (clear solution); PS: partially soluble (cloudy
solution); I: insoluble. d 2E6E: 2,6-di-O-ethyl-cellulose; 2E6M: 2-O- the relatively restricted attainable DP. To date, the aforemen-
ethyl-6-O-methyl-cellulose; 6E2M: 6-O-ethyl-2-O-methyl-cellulose. tioned mixed cellulose ether esters (like the 3,6-di-O-benzyl-2-
pivalate) have been the only reported syntheses of regioselec-
tively substituted cellulose esters by the orthopivalate method,
or append protective groups such that selective deprotection but we can anticipate such studies in the future.
and monomethylation of the product cellulose derivative was
straightforward (2M). This same group has also made the
regiospecically substituted dimethylcelluloses by analogous 5. CONCLUSIONS AND FUTURE PROSPECTS
routes: starting from 3-O-benzyl-6-O-pivaloylglucose-1,2,4- Cellulose researchers now have available to them some
orthopivalate, polymerization with BF3 3 Et2O, followed by depi- remarkable synthetic tools that did not exist 20 years ago.
valoylation (NaOMe/THF/MeOH) at O-2 and O-6, methyla- The protecting group chemistry embodied in the thexyldimethyl-
tion (MeI/NaH/DMSO), and nally debenzylation at O-3 (H2/ silyl and trityl groups has made possible the synthesis of all
Pd(OH)2) aorded 2,6-di-O-methylcellulose, isolated as the possible methylcellulose homopolymers and some novel cellu-
3-acetate;85 starting from 3,6-di-O-methylglucose-1,2,4-orthopi- lose ester homopolymers. Protecting group chemistry has in-
valate, polymerization with BF3 3 Et2O, followed by depivaloyla- herent drawbacks such as extra synthetic steps, lost yield at each
tion (NaOMe/THF/MeOH) at O-2 aorded 36M, isolated as step, the potential for molecular weight degradation, and the
its 2-O-acetate;183 and polymerization of 6-O-allyl-3-O-methyl- possibility that the reactivity of some protected derivatives may
glucose-1,2,4-orthopivalate, followed by depivaloylation at O-2 not be as desired. Nonetheless, we can expect protecting group
(NaOMe/THF/MeOH), methylation at O-2 (MeI/NaOH/ chemistry to play an important role and to be further improved
DMF), and deallylation at O-6 (PdCl2/MeOH/CHCl3) af- methodologically, as researchers attempt to synthesize more
forded 23M, isolated as its 6-O-acetate.183 cellulose ester derivatives and derivatives containing two or more
These syntheses of the dierent methylcelluloses are elegant substituent types.
and useful; they have already permitted NMR studies183 of the The ability to synthesize cellulose derivatives substituted with
individual regiospecically substituted homopolymers, charac- halogens regiospecically at C-6 is a very important advance,
terization of lm properties,184 and the utilization of these bringing to bear not only greater steric accessibility at C-6 but
homopolymers as blocks in subsequent syntheses of blocky also the selectivity advantages of an SN2 substitution reaction, as
cellulose ether derivatives,185188 just to name a few representa- previously mentioned. This family of reactions aords even more
tive studies. Kamitakahara, Nakatsubo, and Rosenau have de- perfect regioselectivity than do the best ether protecting group
monstrated that these methods are rather general for cellulose reactions and has already served as a gateway reaction to regios-
ethers;189 indeed, their only drawbacks for the synthesis of pecically substituted cellulose derivatives such as azides, den-
regiospecically substituted cellulose ethers appear to be their drites, and suldes. We can expect further exploitation of this sort
multistep nature and the fact that the attainable DP of the of chemistry as well to include other types of derivatives and other
products is limited (depending at least in part on the types of chemistries, taking advantage of its very high regioselectivity.
3,6-substituents on the glucose-1,2,4-orthopivalate monomer; De novo syntheses are also notable for their near-perfect
3,6-di-O-benzyl monomer gives DP approaching 20,190 mixed regioselectivity. There is no question that researchers will further
3,6-O-(pivalate ester/benzyl ether) derivatives give DPs in the exploit the potential of this chemistry, making new cellulose
510 range,191 and some simple 3,6-dialkyl ether monomers ether structures, combining de novo synthesis with chemical
have aorded alkylcelluloses of DP 2070192). An example of synthesis of specic oligosaccharides to make blocky structures,
the interesting structureproperty studies enabled by these and in ways not yet imaginable. So far, these chemistries (and
synthetic methods is the synthesis of regiospecically substituted biochemistries) have not proven generally valuable for the
cellulose ethers containing methyl and ethyl substituents; the synthesis of regiospecically substituted cellulose esters; break-
impact of regiochemistry on solubility is highlighted in Table 5, throughs will be needed to open up this important additional
adapted from the Kamitakahara article.190 dimension.
The nature of these substituents is restricted by the require- The work cited in this Review has helped the eld of cellulose
ment that the monomer be reactive toward cationic polymeriza- derivative research move into a new era, in which we will begin to
tion. Electron-withdrawing substituents at O-3 and O-6 of the attain a much more complete understanding of structureprop-
orthopivalate reduce the stability of the intermediate carboca- erty relationships, and can begin to modify commercial synthetic
tion, dramatically reducing reactivity. From a practical point of processes to take advantage of the most desirable structures and
view, at least one ether substituent is essential for polymerization properties. We still have only the most modest tools to explore the
success, and two are preferred.191 One implication is that the dimension of monosaccharide sequence of cellulose derivatives. We
1969 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972
Biomacromolecules REVIEW

can look to the great advances in the biochemistry of other important (29) Turbak, A. F.; El-Kafrawy, A.; Snyder, F. W.; Auerbach, A. B.
polysaccharides, most notably in alginate biosynthesis194196 Solvent System for Cellulose 1981, 4, 252.
and in vitro synthesis of highly controlled monosaccharide (30) Helferich, B.; Koster, H. Ber. Dtsch. Chem. Ges. 1924, 57, 587591.
sequences197199 to see the next potential dimension of (31) Hearon, W. M.; Hiatt, G. D.; Fordyce, C. R. J. Am. Chem. Soc.
cellulose derivative regioselective synthesis. There is little 1943, 65, 24492452.
(32) Honeyman, J. J. Chem. Soc. 1947, 168173.
doubt that interdisciplinary advances combining chemistry (33) Hall, D. M.; Horne, J. R. J. Appl. Polym. Sci. 1973, 17, 28912896.
and biochemistry200 will be required to enable important progress (34) Kern, H.; Choi, S.; Wenz, G.; Heinrich, J.; Ehrhardt, L.;
toward this next dimension of regiochemical control. Mischnick, P.; Garidel, P.; Blume, A. Carbohydr. Res. 2000, 326, 6779.
(35) Li, Y.; Liu, R.; Huang, Y. J. Appl. Polym. Sci. 2008, 110, 17971803.
AUTHOR INFORMATION (36) Takahashi, S. I.; Fujimoto, T.; Barua, B. M.; Miyamoto, T.;
Inagaki, H. J. Polym. Sci., Part A: Polym. Chem. 1986, 24, 29812993.
Corresponding Author (37) Erler, U.; Klemm, D.; Nehls, I. Macromol. Rapid Commun.
*E-mail: kjedgar@vt.edu. 1992, 13, 195201.
(38) Erdmenger, T.; Haensch, C.; Hoogenboom, R.; Schubert, U. S.
Macromol. Biosci. 2007, 7, 440445.
(39) Granstr om, M.; Majoinena, J.; Kavakka, J.; Heikkila, M.;
REFERENCES Kemell, M.; Kilpelainen, I. Mater. Lett. 2009, 63, 473476.
(1) Romling, U. Res. Microbiol. 2002, 153, 205212. (40) Kondo, T.; Gray, D. G. Carbohydr. Res. 1991, 220, 173183.
(2) Sturcova, A.; Davies, G. R.; Eichhorn, S. J. Biomacromolecules (41) Heinze, T.; Rottig, K.; Nehls, I. Macromol. Rapid Commun.
2005, 6, 10551061. 1994, 15, 311317.
(3) Glasser, W. G. Macromol. Symp. 2004, 208, 371394. (42) Gomez, J. A. C.; Erler, U.; Klemm, D. Macromol. Chem. Phys.
(4) Edgar, K. J.; Buchanan, C. M.; Debenham, J. S.; Rundquist, P. A.; 1996, 197, 953964.
Seiler, B. D.; Shelton, M. C.; Tindall, D. Prog. Polym. Sci. 2001, (43) Ifuku, S.; Kamitakahara, H.; Takano, T.; Tanaka, F.; Nakatsubo,
26, 16051688. F. Org. Biomol. Chem. 2004, 2, 402407.
(5) Heinze, T.; Liebert, T.; Koschella, A. Esterication of Polysacchar- (44) Philipp, B.; Klemm, D.; Stein, A. Das Papier 1995, 49, 102108.
ides; Springer: Berlin, 2006; p 232. (45) Petzold-Welcke, K.; Kotteritzsch, M.; Heinze, T. Cellulose
(6) Archer, W. L. Drug Dev. Ind. Pharm. 1992, 18, 599616. 2010, 17, 449457.
(7) Malm, C. J.; Tanghe, L. J. Ind. Eng. Chem. 1955, 47, 995999. (46) Kondo, T. Carbohydr. Res. 1993, 238, 231240.
(8) Adden, R.; Muller, R.; Brinkmalm, G.; Ehrler, R.; Mischnick, P. (47) Kondo, T. J. Polym. Sci., Part B: Polym. Phys. 1994, 32, 12291236.
Macromol. Biosci. 2006, 6, 43544. (48) Kondo, T.; Sawatari, C.; Manley, R. S. J.; Gray, D. G. Macro-
(9) Yu, N.; Gray, G. R. Carbohydr. Res. 1998, 312, 225231. molecules 1994, 27, 210215.
(10) Buchanan, C. M.; Edgar, K. J.; Hyatt, J. A.; Wilson, A. K. (49) Kondo, T. J. Polym. Sci., Part B: Polym. Phys. 1997, 35, 717723.
Macromolecules 1991, 24, 30503059. (50) Schaller, J.; Heinze, T. Macromol. Biosci. 2005, 5, 5863.
(11) Buchanan, C. M.; Edgar, K. J.; Wilson, A. K. Macromolecules (51) Heinze, U.; Heinze, T.; Klemm, D. Macromol. Chem. Phys.
1991, 24, 30603064. 1999, 200, 896902.
(12) Adden, R.; Niedner, W.; Muller, R.; Mischnick, P. Anal. Chem. (52) Liu, H. Q.; Zhang, L. N.; Takaragi, A.; Miyamoto, T. Macromol.
2006, 78, 114657. Rapid Commun. 1997, 18, 921925.
(13) Mischnick, P.; Hennig, C. Biomacromolecules 2001, 2, 180184. (53) Iwata, T.; Azuma, J.; Okamura, K.; Muramoto, M.; Chun, B.
(14) Mu~noz, E.; Xu, D.; Avci, F.; Kemp, M.; Liu, J.; Linhardt, R. J. Carbohydr. Res. 1992, 224, 277283.
Biochem. Biophys. Res. Commun. 2006, 339, 597602. (54) Iwata, T.; Okamura, K.; Azuma, J.; Chanzy, H.; Tanaka, F.
(15) Lindahl, U.; Backstrom, G.; Thunberg, L.; Leder, I. G. Proc. Cellulose 1994, 1, 6776.
Natl. Acad. Sci. U.S.A. 1980, 77, 65515. (55) Iwata, T.; Okamura, K.; Azuma, J.; Tanaka, F. Cellulose 1996,
(16) Zilka, A.; Landau, G.; Hershkovitz, O.; Bloushtain, N.; Bar-Ilan, 3, 107124.
A.; Benchetrit, F.; Fima, E.; van Kuppevelt, T. H.; Gallagher, J. T.; (56) Iwata, T.; Okamura, K.; Azuma, J.; Tanaka, F. Cellulose 1996,
Elgavish, S.; Porgador, A. Biochemistry 2005, 44, 1447714485. 3, 91106.
(17) Gama, C. I.; Tully, S. E.; Sotogaku, N.; Clark, P. M.; Rawat, M.; (57) Iwata, T.; Doi, Y.; Azuma, J.-i. Macromolecules 1997, 30, 66836684.
Vaidehi, N.; Goddard, W. A., 3rd; Nishi, A.; Hsieh-Wilson, L. C. Nat. (58) Iwata, T.; Fukushima, A.; Okamura, K.; Azuma, J. J. Appl. Polym.
Chem. Biol. 2006, 2, 467473. Sci. 1997, 65, 15111515.
(18) Achur, R. N.; Valiyaveettil, M.; Alkhalil, A.; Ockenhouse, C. F.; (59) Tsunashima, Y.; Hattori, K. J. Colloid Interface Sci. 2000,
Gowda, D. C. J. Biol. Chem. 2000, 275, 4034440356. 228, 279286.
(19) Rawat, M.; Gama, C. I.; Matson, J. B.; Hsieh-Wilson, L. C. J. Am. (60) Tsunashima, Y.; Hattori, K.; Kawanishi, H.; Horii, F. Bioma-
Chem. Soc. 2008, 130, 29592961. cromolecules 2001, 2, 9911000.
(20) Dawsey, T. R.; McCormick, C. L. Polym. Rev. 1990, (61) Kasuya, N.; Nakashima, J.; Kubo, T.; Sawatari, A.; Habu, N.
30, 405440. Chirality 2000, 12, 670674.
(21) Kohler, S.; Heinze, T. Macromol. Biosci. 2007, 7, 307314. (62) Harkness, B. R.; Gray, D. G. Macromolecules 1990, 23, 14521457.
(22) Swatloski, R. P.; Spear, S. K.; Holbrey, J. D.; Rogers, R. D. J. Am. (63) Harkness, B. R.; Gray, D. G. Macromolecules 1991, 24, 18001805.
Chem. Soc. 2002, 124, 49744975. (64) Aust, N.; Derleth, C.; Zugenmaier, P. Macromol. Chem. Phys.
(23) El Seoud, O. A.; Koschella, A.; Fidale, L. C.; Dorn, S.; Heinze, T. 1997, 198, 13631374.
Biomacromolecules 2007, 8, 262947. (65) Derleth, C.; Zugenmaier, P. Macromol. Chem. Phys. 1997,
(24) Hasani, M. M.; Westman, G. Cellulose 2007, 14, 347356. 198, 37993814.
(25) Edgar, K. J.; Arnold, K. M.; Blount, W. W.; Lawniczak, J. E.; (66) Yue, Z.; Cowie, J. M. G. Macromolecules 2002, 35, 65726577.
Lowman, D. W. Macromolecules 1995, 28, 41224128. (67) Kasai, W.; Kuga, S.; Magoshi, J.; Kondo, T. Langmuir 2005,
(26) Klemm, D.; Heinze, T.; Philipp, B.; Wagenknecht, W. Acta 21, 23232329.
Polym. 1997, 48, 277297. (68) Karakawa, M.; Chikamatsu, M.; Nakamoto, C.; Maeda, Y.;
(27) Koschella, A.; Fenn, D.; Illy, N.; Heinze, T. Macromol. Symp. Kubota, S.; Yase, K. Macromol. Chem. Phys. 2007, 208, 20002006.
2006, 244, 5973. (69) Kondo, T.; Yamamoto, M.; Kasai, W.; Morita, M. Synthesis and
(28) McCormick, C. L. Novel Cellulose Solutions 1981, 4, 790. Properties of Regioselectively Substituted Cellulose Cinnamates. In

1970 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972


Biomacromolecules REVIEW

Polysaccharide Materials: Performance by Design; American Chemical (105) Belyakova, M. K.; Galbraikh, L. S.; Rogovin, Z. A. Cellul.
Society: Washington, D.C., 2009; Vol. 1017, pp 231241. Chem. Technol. 1971, 5, 405415.
(70) Klemm, D.; Stein, A. J. Macromol. Sci., Pure Appl. Chem. 1995, (106) Heinze, T.; Liebert, T. Prog. Polym. Sci. 2001, 26, 16891762.
32, 899904. (107) McCormick, C. L.; Callais, P. A. Polymer 1987, 28, 23172323.
(71) Koschella, A.; Haucke, G.; Heinze, T. Polym. Bull. 1997, (108) McCormick, C. L.; Dawsey, T. R.; Newman, J. K. Carbohydr.
39, 597604. Res. 1990, 208, 183191.
(72) Koschella, A.; Klemm, D. Macromol. Symp. 1997, 120, 115125. (109) Granstrom, M.; Kavakka, J.; King, A.; Majoinen, J.; Makela, V.;
(73) Petzold, K.; Koschella, A.; Klemm, D.; Heublein, B. Cellulose Helaja, J.; Hietala, S.; Virtanen, T.; Maunu, S.-L.; Argyropoulos, D.;
2003, 10, 251269. Kilpelainen, I. Cellulose 2008, 15, 481488.
(74) Koschella, A.; Heinze, T.; Klemm, D. Macromol. Biosci. 2001, (110) Rahn, K.; Diamantoglou, M.; Klemm, D.; Berghmans, H.;
1, 4954. Heinze, T. Angew. Makromol. Chem. 1996, 238, 143163.
(75) Yin, X.; Koschella, A.; Heinze, T. React. Funct. Polym. 2009, (111) Heinze, T.; Rahn, K.; Jaspers, M.; Berghmans, H. J. Appl.
69, 341. Polym. Sci. 1996, 60, 18911900.
(76) Koschella, A.; Fenn, D.; Heinze, T. Polym. Bull. 2006, 57, 3341. (112) Einfeldt, J.; Heinze, T.; Liebert, T.; Kwasniewski, A. Carbo-
(77) Heinze, T.; Pfeifer, A.; Sarbova, V.; Koschella, A. Polym. Bull. hydr. Polym. 2002, 49, 357365.
2010, 66, 12191229. (113) Heinze, T.; Rahn, K.; Jaspers, M.; Berghmans, H. Macromol.
(78) Petzold, K.; Klemm, D.; Heublein, B.; Burchard, W.; Savin, G. Chem. Phys. 1996, 197, 42074224.
Cellulose 2004, 11, 177193. (114) Tiller, J.; Berlin, P.; Klemm, D. Macromol. Chem. Phys. 1999,
(79) Kondo, T.; Koschella, A.; Heublein, B.; Klemm, D.; Heinze, T. 200, 19.
Carbohydr. Res. 2008, 343, 26004. (115) Berlin, P.; Klemm, D.; Tiller, J.; Rieseler, R. Macromol. Chem.
(80) Schumann, K.; Pfeifer, A.; Heinze, T. Macromol. Symp. 2009, Phys. 2000, 201, 20702082.
280, 8694. (116) Tiller, J.; Berlin, P.; Klemm, D. J. Appl. Polym. Sci. 2000, 75,
(81) Fenn, D.; Heinze, T. Cellulose 2009, 16, 853861. 904915.
(82) Fenn, D.; Pohl, M.; Heinze, T. React. Funct. Polym. 2009, (117) Tiller, J.; Klemm, D.; Berlin, P. Des. Monomers Polym. 2001,
69, 347352. 4, 315328.
(83) Ben-Aroya Bar-Nir, B.; Kadla, J. F. Carbohydr. Polym. 2009, (118) Berlin, P.; Klemm, D.; Jung, A.; Liebegott, H.; Rieseler, R.;
76, 6067. Tiller, J. Cellulose 2003, 10, 343367.
(84) Kadla, J. F.; Asfour, F. H.; Bar-Nir, B. Biomacromolecules 2007, (119) Becher, J.; Liebegott, H.; Berlin, P.; Klemm, D. Cellulose 2004,
8, 161165. 11, 119126.
(85) Kamitakahara, H.; Koschella, A.; Mikawa, Y.; Nakatsubo, F.; (120) Mais, U.; Binder, W. H.; Knaus, S.; Gruber, H. Macromol.
Heinze, T.; Klemm, D. Macromol. Biosci. 2008, 8, 690700. Chem. Phys. 2000, 201, 21152122.
(86) Heinze, T.; Pfeifer, A.; Petzold, K. BioResources 2008, 3, 7990. (121) Bieser, A. M.; Thomann, Y.; Tiller, J. C. Macromol. Biosci.
(87) Xu, D.; Li, B.; Tate, C.; Edgar, K. J. Cellulose 2011, 18, 405419. 2011, 11, 111121.
(88) Zhang, J.; Wu, J.; Cao, Y.; Sang, S.; Zhang, J.; He, J. Cellulose (122) Heinze, T.; Koschella, A.; Magdaleno-Maiza, L.; Ulrich, A. S.
2009, 16, 299308. Polym. Bull. 2001, 46, 713.
(89) Malm, C. J.; Hiatt, G. D. Cellulose Formate. In Cellulose and (123) Koschella, A.; Heinze, T. Macromol. Biosci. 2001, 1, 178184.
Cellulose Derivatives, 2nd ed.; Ott, E., Spurlin, H. M., Grain, M. W., (124) Liu, C.; Baumann, H. Carbohydr. Res. 2002, 337, 12971307.
Eds.; Interscience: New York, 1954; Vol. 2, pp 766767. (125) Baumann, H.; Liu, C.; Faust, V. Cellulose 2003, 10, 6574.
(90) Heath, J. E.; Jeries, R. J. Polym. Sci., Part A: Polym. Chem. 1967, (126) Liebert, T.; Hansch, C.; Heinze, T. Macromol. Rapid Commun.
5, 24652479. 2006, 27, 208213.
(91) Fujimoto, T.; Takahashi, S.; Tsuji, M.; Miyamoto, T.; Inagaki, (127) Knaus, S.; Mais, U.; Binder, W. H. Cellulose 2003, 10, 139150.
H. J. Polym. Sci., Part C: Polym. Lett. 1986, 24, 495501. (128) Liu, C.; Baumann, H. Carbohydr. Res. 2005, 340, 22292235.
(92) Schnabelrauch, M.; Vogt, S.; Klemm, D.; Nehls, I.; Philipp, B. (129) Pohl, M.; Heinze, T. Macromol. Rapid Commun. 2008,
Angew. Makromol. Chem. 1992, 198, 155164. 29, 17391745.
(93) Freeman, R.; Pachler, K. G. R.; La Mar, G. N. J. Chem. Phys. (130) Klebe, J. F.; Finkbeiner, H. L. J. Polym. Sci., Part A: Polym.
1971, 55, 45864593. Chem. 1969, 7, 19471958.
(94) Liebert, T.; Klemm, D.; Heinze, T. J. Macromol. Sci., Pure Appl. (131) Cooper, G. K.; Sandberg, K. R.; Hinck, J. F. J. Appl. Polym. Sci.
Chem. 1996, 33, 613626. 1981, 26, 38273836.
(95) Philipp, B.; Wagenknecht, W.; Nehls, I.; Ludwig, J.; (132) Schempp, W.; Krause, T.; Seifried, U.; Koura, A. Das Papier
Schnabelrauch, M.; Rim, K. H.; Klemm, D. Cellul. Chem. Technol. 1984, 38, 607610.
1990, 24, 667678. (133) Stein, A.; Klemm, D. Macromol. Rapid Commun. 1988, 9,
(96) Klemm, D.; Schnabelrauch, M.; Stein, A.; Philipp, B.; 569573.
Wagenknecht, W.; Nehls, I. Das Papier 1990, 44, 624632. (134) Schultz, P.; Luning, U. Macromol. Chem. Phys. 2002, 203, 961967.
(97) Liebert, T.; Schnabelrauch, M.; Klemm, D.; Erler, U. Cellulose (135) Klemm, D.; Philipp, B.; Heinze, T.; Heinze, U.; Wagenknecht,
1994, 1, 249258. W. Comprehensive Cellulose Chemistry; Wiley VCH: Weinheim,
(98) Mischnick, P. Cellulose 2001, 8, 245257. Germany, 1998; Vol. 2, pp 115133.
(99) Heinrich, J.; Mischnick, P. J. Polym. Sci., Part A: Polym. Chem. (136) Schwartz-Albiez, R.; Adams, Y.; von der Lieth, C. W.; Mischnick,
1999, 37, 30113016. P.; Andrews, K. T.; Kirschnk, M. Glycoconjugate J. 2007, 24, 5765.
(100) Geddes, A. L. J. Polym. Sci. 1956, 22, 3139. (137) Zhang, K.; Peschel, D.; Brendler, E.; Groth, T.; Fischer, S.
(101) Klenkova, N. I.; Buldova, O. S.; Matveeva, N. A.; Volkova, Macromol. Symp. 2009, 280, 2835.
L. A. Cellul. Chem. Technol. 1982, 16, 615629. (138) Yamamoto, I.; Takayama, K.; Honma, K.; Gonda, T.; Matsu-
(102) Hasegawa, M.; Isogai, A.; Onabe, F.; Usuda, M. J. Appl. Polym. zaki, K. Carbohydr. Polym. 1991, 14, 5363.
Sci. 1992, 45, 18571863. (139) El-Sadr, W. M.; Mayer, K. H.; Maslankowski, L.; Hoesley, C.;
(103) Baumann, H.; Richter, D.; Klemm, D.; Faust, V. Macromol. Justman, J.; Gai, F.; Mauck, C.; Absalon, J.; Morrow, K.; Masse, B.; Soto-
Chem. Phys. 2000, 201, 19501962. Torres, L.; Kwiecien, A. AIDS 2006, 20, 110916.
(104) Hon, D. N.-S. Chemical Modication of Cellulose. In Chemi- (140) Anderson, R. A.; Feathergill, K. A.; Diao, X. H.; Cooper, M. D.;
cal Modication of Lignocellulosic Materials; Hon, D. N.-S., Ed.; Marcel Kirkpatrick, R.; Herold, B. C.; Doncel, G. F.; Chany, C. J.; Waller, D. P.;
Dekker: New York, 1996; pp 97128. Rencher, W. F.; Zaneveld, L. J. J. Androl. 2002, 23, 42638.

1971 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972


Biomacromolecules REVIEW

(141) Groth, T.; Wagenknecht, W. Biomaterials 2001, 22, 271929. (178) Perez, X.; Faijes, M.; Planas, A. Biomacromolecules 2011,
(142) Wang, Z. M.; Li, L.; Zheng, B. S.; Normakhamatov, N.; Guo, 12, 494501.
S. Y. Int. J. Biol. Macromol. 2007, 41, 376382. (179) Viladot, J. L.; Moreau, V.; Planas, A.; Driguez, H. J. Chem. Soc.,
(143) Philipp, B.; Wagenknecht, W.; Nehls, I.; Klemm, D.; Stein, A.; Perkin Trans. 1 1997, 23832387.
Heinze, T. Polymer News 1996, 21, 155161. (180) Moreau, V.; Driguez, H. J. Chem. Soc., Perkin Trans. 1 1996,
(144) Vollmer, A.; Voiges, K.; Bork, C.; Fiege, K.; Cuber, K.; 525527.
Mischnick, P. Anal. Bioanal. Chem. 2009, 395, 17491768. (181) Okamoto, E.; Kiyosada, T.; Shoda, S.-I.; Kobayashi, S. Cellu-
(145) Gohdes, M.; Mischnick, P.; Wagenknecht, W. Carbohydr. lose 1997, 4, 161172.
Polym. 1997, 33, 163168. (182) Karrasch, A.; Jager, C.; Karakawa, M.; Nakatsubo, F.; Potthast,
(146) Gohdes, M.; Mischnick, P. Carbohydr. Res. 1998, 309, 109115. A.; Rosenau, T. Cellulose 2009, 16, 129137.
(147) Hyatt, J. A.; Tindall, G. W. Heterocycles 1993, 35, 227234. (183) Karakawa, M.; Mikawa, Y.; Kamitakahara, H.; Nakatsubo, F.
(148) Nehls, I.; Philipp, B.; Wagenknecht, M.; Klemm, D.; J. Polym. Sci., Part A: Polym. Chem. 2002, 40, 41674179.
Schnabelrauch, M.; Stein, A.; Heinze, T. Das Papier 1990, 44, 633639. (184) Ifuku, S.; Nakai, S.; Kamitakahara, H.; Takano, T.; Tsujii, Y.;
(149) Wang, Z.; Li, L.; Xiao, K.; Wu, J. Bioresour. Technol. 2009, 100, Nakatsubo, F. Biomacromolecules 2005, 6, 20672073.
16871690. (185) Kamitakahara, H.; Nakatsubo, F.; Klemm, D. Cellulose 2007,
(150) Philipp, B.; Klemm, D.; Wagenknecht, W. Das Papier 1995, 14, 513528.
49, 5864. (186) Kamitakahara, H.; Nakatsubo, F.; Klemm, D. Cellulose 2006,
(151) Zhang, K.; Brendler, E.; Geissler, A.; Fischer, S. Polymer 2011, 13, 375392.
52, 2632. (187) Enomoto, Y.; Kamitakahara, H.; Takano, T.; Nakatsubo, F.
(152) Zhang, K.; Peschel, D.; Baucker, E.; Groth, T.; Fischer, S. Cellulose 2006, 13, 437448.
Carbohydr. Polym. 2011, 83, 16591664. (188) Kamitakahara, H.; Nakatsubo, F.; Klemm, D. Synthesis of
(153) Hettrich, K.; Wagenknecht, W.; Volkert, B.; Fischer, S. Methylated Cello-Oligosaccharides: Synthesis Strategy for Blockwise
Macromol. Symp. 2008, 262, 162169. Methylated Cello-Oligosaccharides. In Polysaccharide Materials: Perfor-
(154) Heinze, T.; Vieira, M.; Heinze, U. Lenzinger Ber. 2000, 79, 3944. mance by Design; Edgar, K. J., Buchanan, C. M., Heinze, T., Eds.;
(155) Klemm, D.; Heinze, T.; Stein, A.; Liebert, T. Macromol. Symp. American Chemical Society: Washington, D.C., 2009; pp 199211.
1995, 99, 129140. (189) Kamitakahara, H.; Funakoshi, T.; Nakai, S.; Takano, T.;
(156) Philipp, B.; Klemm, D.; Wagenknecht, W.; Wagenknecht, M.; Nakatsubo, F. Macromol. Biosci. 2010, 10, 638647.
Nehls, I.; Stein, A.; Heinze, T.; Heinze, U.; Helbig, K.; Camacho, J. Das (190) Kamitakahara, H.; Funakoshi, T.; Takano, T.; Nakatsubo, F.
Papier 1995, 49, 37. Cellulose 2009, 16, 11671178.
(157) Richter, A.; Klemm, D. Cellulose 2003, 10, 133138. (191) Kamitakahara, H.; Hori, M.; Nakatsubo, F. Macromolecules
(158) Wagenknecht, W.; Nehls, I.; Philipp, B. Carbohydr. Res. 1993, 1996, 29, 61266131.
240, 245252. (192) Kamitakahara, H.; Funakoshi, T.; Nakai, S.; Takano, T.;
(159) Wagenknecht, W.; Nehls, I.; Stein, A.; Klemm, D.; Philipp, B. Nakatsubo, F. Cellulose 2009, 16, 11791185.
Acta Polym. 1992, 43, 266269. (193) Karakawa, K.; Kamitakahara, H.; Takano, T.; Nakatsubo, F.
(160) Furuhata, K.; Koganei, K.; Chang, H.-S.; Aoki, N.; Sakamoto, Biomacromolecules 2002, 3, 538546.
M. Carbohydr. Res. 1992, 230, 165177. (194) Draget, K. I.; Smidsrod, O.; Skjak-Braek, G., Alginates
(161) Matsui, Y.; Ishikawa, J.; Kamitakahara, H.; Takano, T.; from Algae. In Polysaccharides and Polyamides in the Food Industry;
Nakatsubo, F. Carbohydr. Res. 2005, 340, 14031406. Steinbuchel, A., Rhee, S. K., Eds.; Wiley-VCH: Weinheim, Germany,
(162) Furuhata, K.-i.; Chang, H.-S.; Aoki, N.; Sakamoto, M. Carbo- 2005; pp 130.
hydr. Res. 1992, 230, 151164. (195) Deretic, V.; Gill, J. F.; Chakrabarty, A. M. Nat. Biotechnol.
(163) Aoki, N.; Koganei, K.; Chang, H.-S.; Furuhata, K.-i.; Sakamoto, 1987, 5, 469477.
M. Carbohydr. Polym. 1995, 27, 1321. (196) Gacesa, P. Microbiology 1998, 144, 113343.
(164) Aoki, N.; Furuhata, K. I.; Saegusa, Y.; Nakamura, S.; Sakamoto, (197) Crescenzi, V.; Hartmann, M.; de Nooy, A. E.; Rori, V.; Masci,
M. J. Appl. Polym. Sci. 1996, 61, 11731185. G.; Skjak-Braek, G. Biomacromolecules 2000, 1, 3604.
(165) Aoki, N.; Fukushima, K.; Kurakata, H.; Sakamoto, M.; (198) Morch, Y. A.; Donati, I.; Strand, B. L.; Skjak-Braek, G.
Furuhata, K.-i. React. Funct. Polym. 1999, 42, 223233. Biomacromolecules 2007, 8, 280914.
(166) Saad, G. R.; Sakamoto, M.; Furuhata, K. Polym. Int. 1996, (199) Draget, K. I.; Skjak-Braek, G.; Smidsrod, O. Int. J. Biol.
41, 293299. Macromol. 1997, 21, 4755.
(167) Saad, G. R.; Furuhata, K. Polym. Int. 1997, 42, 356362. (200) Gremos, S.; Zarafeta, D.; Kekos, D.; Kolisis, F. Bioresour.
(168) Furuhata, K.-i.; Ikeda, H. React. Funct. Polym. 1999, Technol. 2011, 102, 137882.
42, 103109.
(169) Dumitriu, S. Polysaccharides: Structural Diversity and Func-
tional Versatility, 2nd ed.; Marcel Dekker: New York, 2005.
(170) Kobayashi, S.; Kashiwa, K.; Kawasaki, T.; Shoda, S.-I. J. Am.
Chem. Soc. 1991, 113, 30793084.
(171) Nakatsubo, F.; Kamitakahara, H.; Hori, M. J. Am. Chem. Soc.
1996, 118, 16771681.
(172) Wang, Q.; Dordick, J. S.; Linhardt, R. J. Chem. Mater. 2002,
14, 32323244.
(173) Kobayashi, S. Proc. Jpn. Acad., Ser. B 2007, 83, 215247.
(174) Buchanan, C. M.; Hyatt, J. A.; Kelley, S. S.; Little, J. L.
Macromolecules 1990, 23, 37473755.
(175) Kobayashi, S.; Kiyosada, T.; Shoda, S.-I. J. Am. Chem. Soc.
1996, 118, 1311313114.
(176) Fujita, M.; Shoda, S.-I.; Kobayashi, S. J. Am. Chem. Soc. 1998,
120, 64116412.
(177) Kobayashi, S.; Shimada, J.; Kashiwa, K.; Shoda, S.-I. Macro-
molecules 1992, 25, 32373241.

1972 dx.doi.org/10.1021/bm200260d |Biomacromolecules 2011, 12, 19561972

Вам также может понравиться