Sie sind auf Seite 1von 14

Cellulose 10: 283296, 2003.

2003 Kluwer Academic Publishers. Printed in the Netherlands.

283

Unconventional cellulose esters: synthesis, characterization


and structureproperty relations
Thomas Heinze1,2,3 , Tim F. Liebert1 , Katy S. Pfeiffer1 & Muhammad A. Hussain2
1 Institut

fr Organische Chemie und Makromolekulare Chemie, Friedrich-Schiller-Universitt Jena, Lessingstrasse 8, D-07743 Jena, Germany
2 Fachbereich 9 (Chemie), Bergische Universitt Wuppertal, Gau Strasse 20, D-42097 Wuppertal, Germany
3 Author for correspondence (E-mail: thomas.heinze@uni-jena.de)
Received 26 November 2002; accepted 30 March 2003

Key words: Cellulose ester, Cellulose solvents, In situ activation, Structure, Synthesis
Abstract
This paper summarizes selected results obtained during a two-year research project in the framework of the
focus program Cellulose and cellulose derivatives (SPP 1011), sponsored by the German Science Foundation
(DFG). New synthesis paths for the preparation of the most important cellulose ester, cellulose acetate, were
investigated. In contrast to conventional methods, cellulose was converted in a homogeneous phase reaction with
acetyl chloride in the presence of different bases, including polyvinyl pyridine and cross-linked polyvinyl pyridine.
Moreover, results of the conversion in the new solvent dimethyl sulfoxide/tetrabutylammonium fluoride trihydrate
are discussed. The structures obtained were analyzed both on the level of the anhydroglucose unit (AGU) and
along the polymer chain. It was found that the addition of a base can significantly change the selectivity of the
reaction and thereby the properties of the products (e.g., solubility). No signs of a non-statistical distribution of
the acetyl groups along the polymer chains were observed. Furthermore, reactivity and selectivity of the acylation
reactions, using in situ activation with p-toluenesulfonyl chloride (Tos-Cl), were studied for different long-chain
carboxylic acids (capric-, caprylic-, decanoic-, lauric-, palmitic-, stearic acid). The thermogravimetric analysis
of these derivatives showed that the decomposition temperature increased with an increasing number of carbon
atoms, starting from 292 C (cellulose caprate) to 318 C (cellulose stearate). New cellulose derivatives were
synthesized, for example, cellulose adamantoyl ester. For this purpose cellulose was converted homogeneously
in N,N-dimethylacetamide/LiCl with free acids in the presence of activating reagents, for example, Tos-Cl or
1,1 -carbonyldiimidazol.
Introduction
The development of new reaction paths for polymer
analogous modification is one of the most important tools for the design of cellulosics with tailored
properties. In recent years we have investigated alternative paths for the carboxymethylation of cellulose
(Liebert et al. 1996; Liebert and Heinze 1998a,b;
Heinze et al. 1999). A new concept was established,
including the conversion of cellulose dissolved in N,Ndimethylacetamide (DMA)/LiCl with sodium monochloroacetate in the presence of solid NaOH particles
(Liebert and Heinze 1998a,b). This path yielded de-

rivatives with high degrees of substitution (DS) and


a completely new distribution of substituents on the
level of the repeating units and along the polymer
chain, compared with carboxymethyl cellulose (CMC)
prepared in the industrially applied slurry process.
Thus, it was found that the alternative CMCs showed
a preferred functionalization at position 6 compared to
commercial samples. SEC analysis after polymer fragmentation with endoglucanase indicated a block-like
distribution of the carboxymethyl functions along the
polymer backbone (Saake et al. 2000). Atomic force
microscopy (AFM) revealed a new superstructure.
The alternatively prepared CMC forms a network-like

284
system in solution, while commercially prepared
samples show fringed micelles (Liebert and Heinze
2001). These molecular and supermolecular features
resulted in a number of amazing new macroscopic
properties, for example, different rheological and colloidal behavior (Ktz et al. 2001).
In one of our basic research projects, interest is
focused on the search for new tools for the preparation
of cellulose esters, including the application of cellulose solvents, the in situ activation of carboxylic acids
and the use of polymeric bases. The esterification
of cellulose in DMA/LiCl has been extensively studied during the last decade (Dawsey 1994; El Seoud
et al. 2000). The conversion of the polymer with
free acids after in situ activation was applied besides
acetylation with acid chlorides and acid anhydrides.
In situ activation of the carboxylic acids is possible
with p-toluenesulfonyl chloride (Tos-Cl). It was first
applied for the preparation of cellulose acetates
(Shimizu and Hayashi 1988). The extension of this
path on the homogeneous derivatization of cellulose with waxy carboxylic acids was studied. It was
shown that cellulose esters, having alkyl substituents in the range from C12 (laurylic acid) to C20
(eicosanoic acid), could be obtained with almost
complete functionalization of the OH groups (DS
values 2.82.9; Sealey et al. 1996). It was extended to the preparation of water-soluble oxocarboxylic
acid esters of cellulose (Heinze and Schaller 2000).
Moreover, the very powerful condensation agent N,Ndicyclohexylcarbodiimide (DCC) in combination with
4-pyrrolidinopyridine (PP) was exploited for the synthesis of cellulose esters, starting from free carboxylic
acids (Samaranayake and Glasser 1993a,b). This approach can be used to efficiently prepare derivatives
with low DS.
This paper deals with results concerning different new paths for the esterification of cellulose, including the application of both alternative cellulose
solvents and polymeric bases, as well as the exploitation of 1,1 -carbonyldiimidazole (CDA) as activating
agent. The influence of the esterification path on the
properties of the esters obtained was also studied.
Experimental
Materials
Avicel (Fluka, Avicel PH-101, degree of polymerization DP = 260) was used as a starting polymer. Non-crosslinked polyvinyl pyridine had a Mw

of 200,000 g/mol. LiCl was dried for 6 h at 105 C


in vacuum prior to use. Cross-linked polyvinylpyridine, tetrabutylammonium fluoride trihydrate (TBAF),
acetyl chloride, adamantoyl chloride (AdCl), CDA,
Tos-Cl, dimethyl sulfoxide (DMSO), DMA, and the
carboxylic acids, supplied by Fluka, were used as
received.
Methods
Dissolution of cellulose in DMA/LiCl (solution S1)
For a typical preparation, 1.0 g (6.2 mmol) of dried
cellulose and 40 mL DMA were kept at 130 C for 2 h
under stirring. After the slurry had been allowed to
cool to 100 C, 3 g of anhydrous LiCl were added. The
cellulose was completely dissolved by cooling down
to room temperature under stirring.
Acetylation of cellulose with acetyl chloride
A solution S1 (see above) was kept in an ice bath for
15 min. To this cooled solution was carefully added
2.2 mL acetyl chloride (5 mol/mol AGU). The system
was heated to 80 C for 2 h and kept at room temperature for 24 h. Isolation was carried out by precipitation
into 200 mL ethanol, washing with ethanol and drying
in vacuum at 50 C (sample A4).
Yield: 1.5 g (84.9%).
DSAcetate = 2.96 (determined by means of 1 H NMR
spectroscopy after perpropionylation).
FTIR (KBr): 3502 (OH), 2890 (CH), 1750
(C=
=OEster ) cm1 .
13 C NMR (DMSO-d ): 169.2169.9 ppm (C=
=O),
6
60.3102.5 ppm (cellulose backbone).
If a base was applied, it was added before the addition of the acetyl chloride. If polyvinyl pyridine was
used as a base the products were reprecipitated from
DMSO.
Acetylation of cellulose in DMSO/TBAF
For a typical conversion, a solution of 1 g (6.2 mmol)
cellulose in 33 mL DMSO and 6.6 mL TBAF was
treated with 1.14 mL (14.2 mmol) of vinyl acetate for
70 h at 40 C (for other reagents see Table 1). The
product was isolated by precipitation into 200 mL
isopropyl alcohol, adding 50 mL water (removal of
inorganic impurities, no signals for TBAF in NMR
spectra) and filtration. After washing with 200 mL isopropyl alcohol, the product was dried in vacuum at
50 C (sample B2).
Yield: 1.0 g (80.3%).

285
Table 1. Summary of reaction conditions and results of acetylation of cellulose dissolved in DMA/LiCl with
acetyl chloride.
No.

Molar ratio
(acetyl chloride/AGU)

Partial DSAc a in position


6

A1
A2
A3
A4

1.0
3.0
4.5
5.0

0.77
0.90
1.00
1.00

"

2,3
0.44
1.95
1.94
1.94

1.21d
2.85
2.94
2.96

Solubilityb
DMSOc

Acetone

CHCl3

+
+
+
+

+
e
+

a DS of the ester obtained, determined via 1 H NMR spectroscopy after perpropionylation.


b + Soluble; insoluble.
c Dimethylsulfoxide.
d This stoichiometrically impossible value may result from fractionation during work up.
e The insolubility cannot be explained by structural features.

DSAcetat = 1.04 (determined by means of 1 H NMR


spectroscopy after perpropionylation).
FTIR (KBr): 3490 (OH), 2905 (CH), 1752
(C=
=OEster) cm1 .
13 C NMR (DMSO-d ): 169.1169.9 ppm (C=
=O),
6
60.3102.5 ppm (cellulose backbone).

FTIR (KBr): 3458 (OH), 2910, 2855 (CH), 1728


(C=
=OEster ) cm1 .
13 C NMR (DMSO-d ): = 176.4 (CO), 102.6
6
(C-1), 99.5 (C-1 ), 78.8 (C-4), 73.4 (C-3, C-5, C-2),
62.9 (C-6s), 61.6 (C-6), 40.1 (-C), 38.8 (-CH2 ),
36.4 (-CH2 ), 27.8 ( -CH) ppm.

Reaction of cellulose with AdCl in DMA/LiCl


AdCl (3.7 g, 18.6 mmol) and 1.8 mL (22.3 mmol)
pyridine were added to a solution S1 and stirred for
24 h at 80 C. The homogeneous reaction mixture was
poured into 250 mL of ethanol. After filtration, the
polymer was washed with ethanol and dried in vacuum
at room temperature, product D13.
Yield: 2.3 g (78.6%).
DSAd = 1.92 (determined by means of 1 H NMR spectroscopy after perpropionylation).
FTIR (KBr): 3457 (OH), 2909, 2854 (CH), 1720
(C=
=OEster) cm1 .
13 C NMR (CDCl ): = 176.5 (CO), 103.0 (C-1),
3
100.9 (C-1 ), 81.3 (C-2,3s, C-4), 77.0 (C-3, C-5), 73.6
(C-2), 61.2 (C-6s), 40.9 (-C), 39.0 (-CH2 ), 36.4
(-CH2 ), 27.9 ( -CH) ppm.

Esterification of cellulose with lauric acid/Tos-Cl


Tos-Cl (35 g, 12.5 mmol) was added to a solution S1
followed by 2.47 g (12.5 mmol) of lauric acid under
stirring. The reaction mixture was stirred for 24 h
at 80 C under N2 . The homogeneous reaction mixture was precipitated in 800 mL buffer solution (7.14 g
K2 HPO4 and 3.54 g KH2 PO4 per liter of H2 O) and
the polymer was collected by filtration. After washing
the polymer with 800 mL water three times, Soxhlet
extraction with ethanol was carried out for 24 h. The
polymer was dried at 50 C under vacuum to yield
product C4.
Yield: 2.1 g (77.0%).
DSLaur = 1.55 (determined by means of 1 H NMR
spectroscopy after peracetylation).
FTIR (KBr): 3486 (OH), 2925, 2855 (CH), 1238
(COCEster ), 1753 (=COEster ) cm1 .
13 C NMR (CDCl ): = 173.8 (CO), 104.0 (C-1),
3
102.6 (C-1 ), 72.3 (C-2), 73.3 (C-3), 82.0 (C-4), 75.1
(C-5), 20.634.0 (CMethylene), 13.9 (CMethyl) ppm.

Reaction of cellulose with adamantane carboxylic


acid (AdOH)/CDA
AdOH (3.4 g, 18.6 mmol) was dissolved in 20 mL
DMA and 3.0 g (18.6 mmol) CDA was added. This
mixture was combined with a solution S1 and stirred
for 24 h at 80 C. The mixture was precipitated in
300 mL of ethanol, filtered off, washed with ethanol
and dried in vacuum at room temperature (product
D26).
Yield: 1.8 g (76.7%).
DSAd = 1.31 (determined by means of 1 H NMR spectroscopy after perpropionylation).

Typical example for perpropionylation


of a cellulose ester for DS determination
A mixture of 6 mL pyridine, 6 mL propionic acid
anhydride and 50 mg 4-(dimethylamino)pyridine was
added to 0.3 g of the adamantoyl cellulose D13. After
24 h at 80 C, the reaction mixture was cooled to
room temperature and precipitated in 50 mL ethanol.

286
For purification, the isolated product was reprecipitated from chloroform into 50 mL ethanol, filtered off,
washed with ethanol and dried in vacuum at room temperature.
Yield: 0.7 g (67%).
DSAd = 1.92, DSProp = 1.08 (both determined by
means of 1 H NMR spectroscopy).
FTIR (KBr): no (OH), 2910, 2854 (CH), 1758,
1737 (C=
=OEster ) cm1 .
13 C NMR (DMSO-d ): =177.0173.1 (CO), 100.2
6
62.6 (C atoms of the modified anhydroglucose unit
(AGU)), 41.1 (CH2 -propionate), 39.4 (-C), 39.0
(-CH2 ), 36.8 (-CH2 ), 28.2 ( -CH), 9.4 (CH3 propionate) ppm.
1 H NMR (CDCl ): = 5.10 (H-3), 4.68 (H-2), 4.38
3
(H-1, 6), 3.99 (H-6 ), 3.56 (H-4, 5), 2.18 (CH2 -2,3propionate), 2.03, 1.95, 1.88, 1.73 (H-adamantane),
1.03 (CH3 -2, 3-propionate) ppm.

The HPLC analysis of the Sisal cellulose samples


was carried out as described for cellulose derivatives
(Liebert and Heinze 2001).
For the methylation, 0.5 g of the starting cellulose ester was dissolved in 30 mL trimethylphosphate.
Methyl trifluoromethane sulfonate (4 mol/mol remaining hydroxyl group) and 2,6-di-tert-butylpyridine
(3 mol/mol hydroxyl group) were added. This mixture
was stirred for 4 h at 60 C and 16 h at room temperature using argon as protective gas. Isolation was carried out by precipitation into ethanol. For complete
depolymerization the methyl cellulose ester was treated with 2 N TFA for 4 h at 120 C. The acid and the
water were removed by distillation. The HPLC experiments were carried out as described (Erler et al. 1992).
The Karl Fischer titration was carried out with a
Mettler-Toledo Coulometer DL 37 using Hydranal A
and Hydranal C (Sigma-Aldrich) as reagents.

Measurements

Results and discussion

13 C

NMR spectra were acquired on a Bruker AMX


400 MHz spectrometer. The cellulose esters were
measured in DMSO-d6 , CDCl3 and THF-d8 at 40 and
70 C, respectively. The number of scans was in the
range from 5000 to 20,000.
1 H NMR spectra of the esters were acquired
in CDCl3 after perpropionylation of the unmodified
hydroxyl groups (Heinze and Schaller 2000) to determine the DS-values. FTIR spectra were measured
on a Bio-Rad FTS 25 PC, using the KBr pellet
technique.
Thermal decomposition temperatures (Td ) of the
cellulose esters were determined by thermogravimetric analysis (TGA) on a Mettler Toledo TC 15 Mettler
TG 50 Thermo balance. The Td was reported as the
onset of significant weight loss from the heated sample
(Sealey et al. 1996). Samples (10 mg) were measured
under air with a temperature increase of 10 C/min
from 35 C up to 600 C.
Elemental analyses were performed by CHNS 932
Analyzer (Leco).
For GPC analysis, JASCO equipment was used
including degasser (DG-980-50), pump (PU-980), RIdetector (RI-930) and UV-detector (UV-975) working
at 254 nm. THF was used as eluent (30 C, 1 mL/min).
The separation was carried out using columns from
polymer standards service (Mainz, Germany) with
1000, 10,000 and 1,000,000 . Polystyrene standards
were used for calibration.

Cellulose acetate influence of bases


on the reaction
Different paths for the homogeneous synthesis of cellulose acetates are known. Thus, cellulose was acetylated in DMA/LiCl using acetic anhydride (Marson
et al. 1999; El Seoud et al. 2000). We studied the
acetylation of cellulose dissolved in DMA/LiCl with
acetyl chloride without an additional base and in the
presence of different pyridine derivatives. In a preliminary set of experiments, cellulose dissolved in
DMA/LiCl was converted homogenously with acetyl
chloride. The experimental details and the values of
the DS of the products are summarized in Table 1.
The reaction succeeds with almost complete conversion of the reagent, that is, it can be controlled by
stoichiometry. 1 H NMR experiments of the perpropionylated samples show a preferred functionalization
of the primary hydroxyl group.
In addition to the NMR spectroscopic experiments,
the structure of sample A1 was studied by HPLC after
permethylation and depolymerization. For this purpose, the product A1 was permethylated with methyl
trifluoromethane sulfonate in trimethyl phosphate in
the presence of 2,6-di-tert-butylpyridine, to convert
the pattern of substitution of the acetate into an inverse methyl ether pattern (Figure 1). After complete
saponification of the ester functions and degradation
of the polymer with aqueous trifluoroacetic acid, the

287

Figure 1. Analytical path for the determination of the functionalization pattern of cellulose esters by means of HPLC after permethylation and
degradation.
Table 2. Summary of reaction conditions and results of the acetylation of cellulose dissolved in DMA/LiCl with acetyl
chloride in the presence of pyridine.
No.

A5
A6
A7
A8

Partial DSa in position

Molar ratio
Acetyl chloride/AGU

Pyridine/AGU

2,3

1.0
3.0
5.0
5.0

1.2
3.6
6.0
10.0

0.63
0.94
0.71
0.46

0.37
1.62
2.0
2.0

"

1.00
2.56
2.71
2.46

Solubilityb
DMSOc

Acetone

CHCl3

+
+
+
+

+
+

+
+
+

a DS of the ester obtained determined via 1 H NMR spectroscopy after perpropionylation.


b + Soluble; insoluble.
c Dimethylsulfoxide.

mixture of methyl glucoses obtained can be separated


by means of HPLC (Erler et al. 1992). A DSAcetate of
1.16 was calculated from the chromatogram, which is
in good agreement with the DS obtained by 1 H-NMR
spectroscopy (DS = 1.21). No evidence for ester group
migration during the procedure was found, but it cannot be completely excluded. A comparison of the
results obtained using this analytical strategy with statistical calculations was performed in the same way as
for the analysis of CMC (Heinze et al. 1999). No significantly increased amounts of glucose or trimethyl
glucose were found by means of HPLC. Thus, glucose was determined to be 3% (calculated 5.7%) and
21% trimethyl glucose was found (calculated 23.1%).
Consequently, cellulose acetates prepared via this path
have a statistically even distribution of substituents
along the polymer chain.
In another set of experiments the influence of a
base on the course of the reaction and on the distribution of substituents was studied. An amazing result
was that the application of pyridine as base leads to
products of a decreased DS (Table 2). Comparison
of samples A2 and A6 or samples A4 and A7 shows
a decrease of DS of about 0.3. It is even more pronounced if the amount of base is increased (see sample
A8). Moreover, 1 H NMR spectroscopy of the products
reveals less preferred substitution in position 6. This
selectivity is diminished by an increased concentra-

tion of the base. Thus, sample A8 shows a partial


DSO-6 of 0.46 versus an overall DS of 2.46, that
is, all the secondary OH groups are acetylated. This
could be a first hint for a preferred deacetylation at the
6-O-position.
GPC was applied to investigate hydrolytic degradation of the polymer chain during the reaction.
It was found that the depolymerization was rather
small without a base. All derivatives were prepared
with Avicel as starting polymer, having a DP of 260.
Product A7 possesses a DP of 256. However, the DP
decreases to 103 during the reaction under comparable
conditions but using pyridine as base (sample A8).
One possible explanation for the degradation might
be the formation of the acidic pyridinium hydrochloride in the case of the base-catalyzed reaction. Most
of the HCl formed is liberated from the system if no
additional base is applied. It needs to be mentioned
that the influence of the acidic pyridinium hydrochloride yields a product with a different solubility. Thus,
sample A8 dissolves completely in acetone in contrast
to sample A4 (prepared with no base, see Table 1).
Permethylation, degradation and HPLC as described
above did not show any hints for a non-statistical distribution of the substituents along the polymer chain.
Consequently, the different solubility is only due to the
different distribution of substituents on the level of the
AGU.

288
Table 3. Conditions and results of the acetylation of cellulose dissolved in DMA/LiCl with acetyl chloride in the presence
of crosslinked polyvinyl pyridine.
No.

A9
A10
A11
A12
A13
A14d
A15e

Partial DSa in position

Molar ratio
Acetyl chloride/AGU

Base/AGU

2,3

1.0
2.0
3.0
4.5
5.0
5.0
5.0

1.2
2.4
3.6
4.5
6.0
10.0
10.0

0.35
0.82
0.91
1.0
1.0

0.97

0.13
0.51
0.65
1.24
1.62

1.31

"

0.48
1.33
1.56
2.24
2.62

2.28

Solubilityb
DMSOc

Acetone

CHCl3

+
+
+
+
+

a DS of the ester obtained determined via 1 H NMR spectroscopy after perpropionylation.


b + Soluble; insoluble.
c Dimethylsulfoxide.
d Product isolation not possible because insoluble polymer is fixed on base surface.
e Non-crosslinked base was used.

For the first time, polymer-bound bases like crosslinked polyvinyl pyridine were applied for the preparation of cellulose carboxylic acid esters. Table 3
summarizes the reaction conditions and results. Again,
a significant decrease of overall DS values can be recognized in comparison to reactions applying no base.
Thus, for sample A2 (see Table 1) an almost complete
substitution (DS = 2.85) was found if a molar ratio of
3 mol acetyl chloride per mol AGU is used. The DS
reached was 1.56 in a comparable experiment (sample
A11) applying polyvinyl pyridine. High selectivity of
the acetylation at position 6 was observed, in contrast
to the acetylation reactions with pyridine as base. A
drastic decrease of both the DS values and the yield as
well as a different solubility of the product is observed
if a large surplus of polymer-bound base is used.
Thus, sample A13 is soluble in DMSO only, even
with a high DS of 2.62. If the molar ratio base/AGU
is in the range >10, product isolation is almost impossible (sample A14). Extraction of the precipitate,
which consists mainly of polyvinylpyridine and cellulose acetate, as can be confirmed by FTIR (signals
at 1595 and 3060 cm1 for the polyvinylpyridine and
signals at 1019 and 1740 cm1 for the cellulose acetate), yields only traces of the product, in the range
of 1%. If the extraction was carried out with DMSO,
9 g cellulose acetate was recovered for an experiment with 1 g of cellulose as starting material. If
THF was used, 11 g were isolated. 1 H-NMR spectroscopy was applied for structure determination but
no DS calculation was possible because of the poor
yield. The alternative solubility of these cellulose acetates is comparable to p-toluenesulfonic acid esters

of cellulose (cellulose tosylate), with a non-statistical


distribution of substituents along the polymer chain
prepared in a reactive microstructure, that is, by conversion of cellulose regenerated from solution on solid
NaOH particles (Einfeldt et al. 2002).
An acetylation experiment was carried out using
soluble, non-cross-linked polyvinyl pyridine (sample
A15; Table 3) to obtain a cellulose acetate that can
be isolated from the polymeric base. No regeneration
or precipitation of the polymers occurred during the
completely homogeneous reaction. A cellulose acetate
was obtained with a DS of 2.28, determined by 1 H
NMR spectroscopy, which is easily soluble in acetone.
HPLC analysis gave a DS of 2.30 and showed no increased values for non- or fully-substituted repeating
units. It may be assumed that during the conversion
the cellulose is not permanently fixed to the dissolved
polymeric base and an even distribution of substituents
resulted from an equilibrium reaction.
It should be mentioned that in the framework of
this study, acetylation experiments were carried out
using in situ activation of acetic acid with CDA (for
the mechanism see below). DSAcetate values of 0.7, 1.5
and 2.1, respectively, were achieved if molar ratios of
1:2:2, 1:5:5 and 1:10:10 (AGU/acid/CDA) were applied. However, this new path did not yield polymers
with a new pattern of functionalization.
Acylation in the new cellulose solvent
DMSO/TBAF
A mixture of DMSO/TBAF represents an efficient cellulose solvent. It dissolves cellulose completely with a

289
Table 4. Esterification of different types of cellulose in DMSO/TBAF. Summary of reaction conditions and results.
No.

Cellulose
type

Acetylating agent

Molar
ratioa

%TBAF in
DMSO

Time (h)

Temp. ( C)

DSb

Solubility

B1
B2
B3
B4
B5
B6
B7
B8
B9
B10
B11

Avicel
Avicel
Avicel
Avicel
Avicel
Avicel
Sisal
Sisal
Sisal
Sisal
Sisal

Acetic anhydride
Vinyl acetate
Vinyl acetate
Vinyl butyrate
Vinyl laurate
Vinyl benzoate
Vinyl laurate
Acetic anhydride
Acetic anhydride
Acetic anhydride
Acetic anhydride

1:2.3
1:2.3
1:10.0
1:2.3
1:10.0
1:2.3
1:2.3
1:2.3
1:2.3
1:2.3
1:2.3

16
16
16
16
16
16
11
11
8
7
6

70
70
70
70
70
70
3
3
3
3
3

40
40
40
40
40
40
60
60
60
60
60

0.83
1.04
2.72
0.86
2.60
0.95
1.24
0.3
0.96
1.07
1.29

Insoluble
DMSO
DMSO
Insoluble
Pyridine, THF, CHCl3
Insoluble
Insoluble
Insoluble
DMSO, pyridine
DMSO, pyridine
DMSO, DMF, pyridine

a Mol AGU/mol acylation reagent.


b DS of the ester obtained determined via 1 H-NMR.

DP of up to 650 without pretreatment within 15 min.


In 13 C NMR spectra signals appear only at 102.7
(C-1), 78.4 (C-4), 75.6 (C-5), 75 (C-3), 73.5 (C-2)
and 59.9 ppm (C-6) (Heinze et al. 2000). This clearly
shows that the cellulose is dissolved without covalent
interactions, as can be concluded from the comparison of the chemical shifts with values of cellulose
dissolved in DMA/LiCl, which is a typical so-called
non-derivatizing cellulose solvent. This new solvent
was exploited for a number of acylation reactions. A
summary of reaction conditions and DS values of the
products obtained is given in Table 4.
The dissolved cellulose was treated with acetic anhydride for 70 h at 40 C. A cellulose acetate with a
DS value of 0.83 was obtained if a molar ratio of
2.3:1.0 (acylation reagent/AGU) was applied. Comparable conditions were used for the reaction with
vinyl acetate as acylating reagent. In case of the same
molar ratio, a DS of 1.04 can be achieved, which is
due to the formation of acetaldehyde during this conversion, shifting the equilibrium towards the product
side. On the other hand, the lower DS in the case of the
application of acetic anhydride is caused by the comparably fast hydrolysis of the reagent, due to the water
content of the solvent. A variety of vinyl carboxylic
acid esters can be exploited for this type of conversion
(see Table 4). The DS values can be controlled via the
amount of reagent added. A remarkable result was a
DS as high as 2.6 for cellulose laurate, indicating that
this homogeneous esterification path is highly efficient
for the preparation of fatty acid esters of cellulose.
Experiments were carried out with Sisal cellulose,
which represents fast-growing lignocellulosic material

comparable to sugarcane bagasse and linters (El Seoud


et al. 2000; Marson et al. 2000; Ass and Frollini 2001;
Sun et al. 2001). The starting material had a DP of 650,
a crystallinity index (Ic ) of 77%, and contained about
14% hemicellulose, as confirmed by 13 C NMR spectroscopy (Figure 2) and HPLC analysis after complete
depolymerization (Figure 3). The conditions suitable
for dissolution of cellulose materials (Avicel, wood
pulp) in DMSO/TBAF discussed above did not yield
optically clear solutions in the case of Sisal cellulose.
This is obviously due to the presence of hemicellulose
and the fibrous structure of Sisal cellulose. However, it
was found that Sisal cellulose dissolves completely in
the mixture DMSO/TBAF after 30 min at room temperature and 60 min at 60 C (Ciacco et al. 2000).
Nevertheless, static light-scattering experiments of the
solution showed a fairly high amount of aggregation.
The values of the molecular weights determined were
in the range of 20 to 50 106 g/mol (Figure 4).
Sisal celluloses were esterified homogeneously in
DMSO/TBAF using acetic anhydride and vinyl laurate
as acylating reagents. Reaction conditions, results and
solubility of the esters obtained are listed in Table 4.
Transesterification with vinyl laurate yields the corresponding ester with a DS of 1.24, while conversion
with acetic anhydride gave an acetate with DS of 0.30.
The concentration of TBAF in the solution was varied
from 6 to 11% to study this influence on the dissolution and the product features. All mixtures gave clear
solutions. The DS values of the cellulose acetates prepared in these different solvent mixtures decrease with
increasing TBAF concentration (see Table 4). As the
salt is hydrated, the amount of water in the medium

290

Figure 2. 13 C-NMR spectrum of Sisal cellulose (cell) in DMSO/TBAF, also showing peaks due to the presence of xylose (xyl).

Figure 4. Berry plot of Sisal cellulose and alkali treated Sisal


cellulose in DMSO/TBAF.

Figure 3. HPL-chromatogram of a completely depolymerized Sisal


cellulose sample. (A chiral detector signal, B refraction index
detector (RI) signal, dr detector response, RT retention time, 1
inorganic salts, 2 glucose, 3 xylose).

increases with the salt concentration. This in turn increases the rate of hydrolysis both of the anhydride and
probably of the ester moieties formed as well. Furthermore, the interactions of the water with the cellulosic
OH groups may hinder the access of acetic anhydride,
resulting in a lower DS.

Experiments directed towards removal of the water


in the solvent were carried out. The water content was
analyzed by means of Karl Fischer titration. The addition of molecular sieves did not significantly influence
the water content. Because the treatment of the solvent
with strong dehydrating reagents like sodium hydride
would produce the undesired dimsyl ions as a byproduct, we studied the dewatering of DMSO/TBAF,
DMSO/TBAF/Sisal and DMSO/TBAF/Avicel by vacuum distillation.
A mixture of 60 ml DMSO and 6.6 g TBAF and
comparable mixtures containing cellulose (between
1.4 and 2.6%, w/w) were distilled stepwise (steps
of 0.6 mL). The first sample obtained for pure

291
DMSO/TBAF contained 55% water. In the case of
a solution of Avicel/DMSO/TBAF 22% water was
found and 5% for a solution of Sisal/DMSO/TBAF,
after the first distillation step. These data lead to the
assumption that the water is strongly involved in the
solution complex of cellulose. After distillation of a
total amount of about 6 mL of water (in the case of the
cellulose-containing solution) a drastic increase in the
viscosity occurred. Therefore, it is useful to prepare
a mixture of DMSO/TBAF, distill the majority of the
water by removing about 30% (v/v) of the mixture
in vacuum and dissolve the cellulose in the resulting
solvent mixture. Solutions prepared in this manner still
gave optically clear systems after the heat treatment
described above, which was also used for the acetylation of cellulose. The reactions in the solvent with a
reduced water content lead to products with a significantly higher DS under comparable conditions. The
DSAcetate increases from 0.30 (see Table 4, sample B8)
to 1.15. Thus, this path is more efficient for the acetylation of Sisal, applying acetic anhydride. 1 H NMR
analysis of the perpropionylated product shows a distribution of the acetyl groups at the reactive sites in the
order C-6 > C-2 > C-3. No hints for a non-statistical
distribution of substituents along the polymer chain
were found. A reaction of cellulose dissolved in anhydrous DMA/LiCl, applying the same molar ratio of
acetic anhydride, yields a product with a DS of 1.0
(Ciacco et al. 2000).
Studies on the acylation of cellulose with carboxylic
acids in situ activated with Tos-Cl
An interesting new path for cellulose ester preparation
is homogeneous acylation after in situ activation of
carboxylic acids with Tos-Cl (Figure 5). It was shown
that cellulose esters, having alkyl substituents in the

Figure 5. Schematic plot of the conversion of cellulose with


carboxylic acid applying in situ activation with Tos-Cl.

range from C12 to C20, could be obtained with almost complete functionalization of the accessible OH
groups (Sealey et al. 1996). A variety of different
cellulose esters was successfully synthesized via this
path, however, without the use of an additional base
(Koschella et al. 1997; Heinze and Schaller 2000).
Considering these results, the question arises if the
reaction conditions (time, molar ratio of the reagents)
and the application of an additional base, for example,
pyridine, influence the DS, the molecular weight and
other structural features of the products. These studies
were performed with long-chain fatty acids because
the efficiency of this particular system for the preparation of the corresponding esters had been shown
(Heinze and Liebert 2001).
Thus, cellulose dissolved in DMA/LiCl was allowed to react with two equivalents, carboxylic acid
(capric-, caprylic-, decanoic-, lauric-, palmitic- and
stearic acid, respectively) and Tos-Cl without an additional base for 24 h at 80 C. The corresponding esters
(polymers C16, Table 5) show two characteristic
peaks in FTIR spectra typical for the ester moieties at
about 1240 cm1 (COCEster) and about 1750 cm1
(C=
=OEster ). Elemental analysis reveals the absence of
sulfur in the samples, showing that there is no remarkable introduction of tosylate groups, either covalently
bounded or as impurity.
The 13 C NMR spectrum of C4, for example, recorded in CDCl3 , shows the characteristic signals at
= 173.8 (CO), 104.0 (C-1), 102.6 (C-1 ), 72.3 (C-2),
73.3 (C-3), 82.0 (C-4), 75.1 (C-5), 62.5 (C-6), 13.9
(CH3 ) ppm. The signals of the methylene groups of
the lauric acid appear in the range of 22.634.0 ppm
(Figure 6). The peak for C-6 bearing an ester group appears at = 62.5 ppm. The acylated primary OH group
exhibits a downfield shift of about 3 ppm compared
with the corresponding carbon of the CH2 OH function. Again DS values were determined by means of
1 H-NMR spectroscopy after peracetylation of the remaining OH groups. A representative 1 H-NMR spectrum of cellulose acetate laurate (synthesized from
sample C4) recorded in CDCl3 is shown in Figure 7.
The protons of the laurate moiety appear at 2.3 (H-8),
1.21.6 (H-10-17) and 0.8 (H-18) ppm. The acetate
methyl group leads to the signal at 1.9 (H-20) ppm.
These results are in very good agreement with values
reported for a cellulose acetate laurate synthesized in
the new solvent DMSO/TBAF, applying vinyl laurate
and acetic anhydride (see above). It was found that
the DS increased with the increasing carbon number
of the carboxylic acid. Thus, a DS of 0.6 was found

292
Table 5. Conditions and results of esterification of cellulose dissolved in DMA/LiCl mediated with Tos-Cl with different carboxylic acids.
No.

Carboxylic acid

Molar ratioa

Time (h)

DSb

Solubility

C1
C2
C3
C4
C5
C6
C7
C8
C9
C10
C11
C12
C13
C14
C15
C16
C17
C18
C19
C20
C21
C22

Capric
Caprylic
Decanoic
Lauric
Palmitic
Stearic
Caprylic
Lauric
Palmitic
Stearic
Caprylic
Lauric
Palmitic
Caprylic
Lauric
Palmitic
Caprylic
Lauric
Palmitic
Caprylic
Lauric
Palmitic

1:2:2:0
1:2:2:0
1:2:2:0
1:2:2:0
1:2:2:0
1:2:2:0
1:2:2:4
1:2:2:4
1:2:2:4
1:2:2:4
1:1:1:0
1:1:1:0
1:1:1:0
1:4:4:0
1:4:4:0
1:4:4:0
1:2:2:0
1:2:2:0
1:2:2:0
1:2:2:0
1:2:2:0
1:2:2:0

24
24
24
24
24
24
24
24
24
24
24
24
24
24
24
24
4
4
4
1
1
1

1.31
1.40
1.48
1.55
1.60
1.76
1.76
1.79
1.71
1.92
0.63
0.36
0.46
2.56
2.56
2.54
1.27
1.55
1.50
1.25
1.36
1.36

DMF
DMSOc , DMF, CHCl3
DMF, CHCl3 , toluene
Toluene, CHCl3
Toluene, CHCl3
Toluene, CHCl3
DMF, CHCl3
CHCl3
CHCl3
CHCl3
DMSO, DMF
Insoluble
Insoluble
Toluene, CHCl3
Toluene, CHCl3
Toluene, CHCl3
DMF, CHCl3
CHCl3
CHCl3
DMSO, DMF
Insoluble
CHCl3

a Mole AGU/mol carboxylic acid/mol Tos-Cl/mol pyridine.


b DS calculated by 1 H NMR spectroscopy after peracetylation.
c Dimethylsulfoxide.

Figure 6. 13 C-NMR spectrum of cellulose laurate C4 (DS = 1.55) recorded in CDCl3 at 40 C, index means influenced by a functionalization
of the neighbor position (number of scans 11,000).

for the cellulose caprate C1, while cellulose caprylate


C2 possesses a DS of 1.4. Under comparable conditions, a cellulose stearate C6 with a DS of 2.0 was
accessible.

The cellulose esters possess a different solubility depending on their DS and the chain length of
the carboxylic acid (Table 5). In general, cellulose
fatty acid esters having DS values higher than 1.4 are

293

Figure 7. 1 H-NMR spectrum of cellulose acetate laurate (starting polymer C4) recorded in CDCl3 at 40 C (16 scans were accumulated).

soluble in CHCl3 , independently of the chain length of


the carboxylic acid. Polymers with DS values higher
than 2.3 are soluble in toluene.
In another series of experiments the influence
of an additional base was investigated. Cellulose
was reacted with two equivalents of carboxylic acid
and Tos-Cl and four equivalents with pyridine as
base. Thus, polymers C710 were obtained bearing
caprylic- (C7), lauric- (C8), palmitic- (C9) and stearic
ester (C10) functions. It was found that the DS values
were higher than the samples prepared without base
(C16). For instance, a DS of 1.55 was found for the
cellulose laurate C4, synthesized without base. The
addition of base increases the DS to 1.79 (C8) (see
Figure 8). Elemental analysis revealed the absence of
sulfur. Therefore, it can be concluded that Tos-Cl acts
only as activating reagent. No tosylation occurs.
GPC was applied to investigate hydrolytic degradation of the polymer chain during the reaction.
Cellulose palmitate C5, synthesized in the absence of
base, yielded a polymer with DP 41, whereas cellulose
palmitate C9 obtained in the presence of base, yielded a DP value of 69. Similar results were obtained
for cellulose stearate C6 (without base, DP = 45) and
C10 (with base, DP = 61). Compared with the DP of
the starting cellulose Avicel (DP 260), a fairly drastic
degradation occurred in every case.
Td were obtained by TGA for cellulose caprate
(292 C), caprylate (300 C), decanoate (301 C),
laurate (302 C), palmitate (306 C) and stearate
(318 C). Cellulose esters C16 showed increasing

Figure 8. DS of cellulose esters synthesized in N,N-dimethyl acetamide/LiCl using in situ activation with the Tos-Cl dependent on
the carboxylic acid and the addition of pyridine (!) and without
pyridine (").

stability with an increase in chain length from C-6 to


C-18. The effect of the chain length on that behavior
is more pronounced than the influence of the rather
small increase in DS of the samples discussed. The
minimum Td value of cellulose laurate C4 was 292 C.
The maximum Td value for cellulose stearate C6 was
318 C. The results of TGA were comparable with the
reported behavior of long-chain fatty acid esters of
cellulose (Sealey et al. 1996).
Synthesis and characterization of adamantoyl
cellulose prepared via different paths
The esterification of cellulose with AdOH was studied because this ester moiety has found considerable

294
Table 6. Summary of reaction conditions and results of the homogeneous reaction of cellulose with AdCl in DMA/LiCl.
No.

Molar ratioa

Time (h)

Temperature ( C)

DSAd b

Solubility

D2
D3
D4
D5
D6
D7
D8
D9
D10
D11
D12
D13
D14
D15
D16
D17
D18

1.0
1.3
1.5
1.7
2.0
2.0
3.0
1.0
1.5
2.0
2.0
3.0
4.0
5.0
2.0
2.0
2.0

24
24
24
24
5
24
24
24
24
5
24
24
24
24
24
24
24

20
20
20
20
20
20
20
80
80
80
80
80
80
80
35
50
65

0.24
0.51
0.65
0.76
0.08
0.90
1.37
0.51
0.87
1.21
1.71
1.92
1.94
2.12
1.19
1.38
1.57

Insoluble
DMSOc , pyridine
DMSO, pyridine
DMSO, pyridine
DMSO, pyridine
DMSO, pyridine
DMSO, pyridine
DMSO, pyridine
DMSO, pyridine
DMSO, pyridine, THFd
Pyridine, THF, CHCl3
Pyridine, THF, CHCl3
Pyridine, THF, CHCl3
Pyridine, THF, CHCl3
DMSO, pyridine, THF
DMSO, pyridine, THF
DMSO, pyridine, THF

a Mole AdCl per mol AGU.


b DS of adamantoyl functions determined after perpropionylation by 1 H NMR spectroscopy.
c Dimethylsulfoxide.
d Tetrahydrofurane.

interest as a base-sensitive protecting group in


deoxyribonucleoside chemistry (Greene and Wuts
1991). On the other hand, incorporation of adamantoyl
functions leads to products with interesting biological activities, such as antimicrobial and antibacterial
activity (Orzeszko et al. 2000; Perrakis et al. 1999) as
well as antitumor activity (Gerzon and Kau 1967).
For the activation of the carboxylic acid, different
methods were studied, namely acid chloride as well
as in situ activation with Tos-Cl and CDA were applied. All reactions were carried out homogeneously
in DMA/LiCl. Thus, cellulose dissolved in DMA/LiCl
was allowed to react with AdCl in the presence of
pyridine (Grbner et al. 2002).
The conversion of Avicel with AdCl leads to the
adamantoylated products D218 (Table 6), which
show the typical IR spectra. Absorption bands of
the cellulose backbone were found and, additionally,
=OEster ), indicating the
a signal at 1758 cm1 , (C=
presence of the ester moiety. As can be seen from
Table 6, DS values up to 2.12 were accessible via this
path. With regard to the reaction efficiency, that is,
the amount of carboxylic acid bound to the polymer
related to the molar ratio, the use of 2 mol AdCl per
mol AGU is most effective. About 85% of the AdCl
reacts with the cellulose backbone (sample D12).

The reaction of the dissolved cellulose with AdOH


in the presence of Tos-Cl leads to corresponding
carboxylic acid esters with a rather high DS of 1.50
and 1.75 at a molar ratio of AGU/AdOH/TosCl of
1/2/2 and 1/3/3, respectively within 24 h at 80 C reaction temperature (samples D20 and D21, Table 7).
Surprisingly, conversion at a molar ratio of 1/1/1
(sample D19) at 80 C as well as at room temperature
does not yield sufficient DS.
Comparable results were obtained by the conversion of cellulose with AdOH in the presence of CDA.
This method is especially suitable for cellulose modification because the pH is not drastically changed
during the conversion and consequently diminished
chain degradation can be guaranteed. The by-products
formed during the reaction (Figure 9) are only CO2
and imidazole, and none of the reagents and byproducts are toxic. At room temperature, even at a
molar ratio of 1/3/3 (AGU/AdOH/CDA), no cellulose
ester was formed. On the other hand, at the comparable molar ratio of 1/3/3, however at 80 C, product
D26 with a DS of 1.31 is obtained, which is rather low
compared to product D13 (DS = 1.92), synthesized
with AdCl under the same conditions. It is interesting
that the increase in the amount of cellulose from 1 to
10 g of starting material results in higher DS values

295
Table 7. Conditions and results of the reaction of cellulose dissolved in DMA/LiCl with AdOH after in situ
activation with Tos-Cl (method A) or CDA (method B).
No.

Method

Molar ratioa

Time (h)

Temperature ( C)

DSb

D20
D21
D22
D23
D24
D25
D26
D27

A
A
Bc
Bc
B
B
B
Bc

1:2:2
1:3:3
1:3:3
1:3:3
1:1:1
1:2:2
1:3:3
1:3:3

24
24
6
8
24
24
24
24

80
80
80
80
80
80
80
80

1.50
1.75
0.62
0.90
0.54
0.98
1.31
1.42

a Mole AGU/mol AdOH/mol Tos-Cl or CDA.


b DS of adamantoyl functions determined after perpropionylation by 1 H NMR spectroscopy.
c Amount of cellulose 10 g.

polymer D10 and D17 are not active at all. Further


studies about the molecular weight, its distribution
as well as about the biological dependence of the
adamantoyl cellulose activity on the physico-chemical
properties are needed to gain clear structureproperty
understanding.

Conclusions and outlook


Figure 9. Schematic plot of the conversion of cellulose with
carboxylic acid applying in situ activation with CDA.

under comparable conditions. Thus, a product with a


DS of 1.42 was synthesized (sample D27, Table 7).
The structures of the synthesized adamantoyl celluloses were confirmed by means of 13 C NMR- and 1 H
NMR spectroscopy, including two-dimensional methods (Grbner et al. 2002). Adamantoyl celluloses D7,
D10, D11, D17 and D21, of different DSAd , were selected for screening for biological activity. It is known
that adamantoylated nucleosides are biologically active because the adamantoyl moiety binds precisely to
a complementary hydrophobic receptor region of protein molecules (Gerzon and Kau 1967). Thus, the antimicrobial, antioxidant and anti-inflammatory effects
of the cellulose derivative were tested. All samples investigated did not show antioxidant or anti-microbial
effects. However, polymers D7, D11 and D21 possessed a strong anti-inflammatory effect while D10
and D17 showed no effect. Obviously, the effect does
not only depend on the DSAd , because samples D7
and D10 or D11 and D17 have similar DS, however,

The results of our work show that polymeric bases,


such as polyvinyl pyridine and the new cellulose
solvent DMSO/TBAF, are new tools for the preparation of cellulose esters. Furthermore, the efficiency
of the in situ activation of the carboxylic acids with
CDA and Tos-Cl, respectively, has been shown. Thus,
a broad variety of cellulose esters with tailored properties, for example, solubility, biological activity or
thermal behavior can be synthesized. These features
are adjustable by the type of substituent introduced
and by the pattern of substitution, which can be modified by application of one of the methods described.
Unfortunately, the synthesis of cellulose esters
with an unconventional (non-statistical) distribution
of substituents along the polymer chain has still not
been achieved. Therefore, we continue our work on
the application of polymeric reagents for ester synthesis. On the other hand, the analysis of the pattern
of substitution on this level is not yet satisfactory.
The HPLC analysis, via methylation and depolymerization, is connected to a number of errors, mainly
undermethylation, migration of the substituents during
the second substitution and incomplete depolymerization. Thus, at present, a comparable method is under

296
investigation, applying percarbanilation of the cellulose ester as the second derivatization step.

Acknowledgements
Financial support for this study (project HE2054/5-3)
was provided as part of the focus research program
(Schwerpunktprogramm) on Cellulose and cellulose
derivatives molecular and supramolecular structural design by the Deutsche Forschungsgemeinschaft
(DFG) [described in the editorial commentary of
Prof G. Wenz in Cellulose 10-1] and by the Fonds
der Chemischen Industrie. The authors would like to
thank former co-worker D. Grbner, and PhD student
G.T. Ciacco (Instituto de Quimica des Sao Carlos,
Universidade de Sao Paulo, Brazil) having stayed
for six months in my group; they were included in
the basic research program on the acylation of cellulose. Furthermore, we thank Dr W. Radosta and
Dr W. Vorwerg (Fraunhofer Institut fr Angewandte
Polymerforschung, Golm, Germany) for GPC studies.
Moreover, the authors wish to thank M. Ktteritzsch
for technical assistance.

References
Ass B.A.P. and Frollini E. 2001. Aggregation of cellulose during
dissolution and acetylation in N,N-dimethylacetamide/lithium
chloride: An introductory study. An. Assoc. Bras. Qum. 50:
7682.
Ciacco G.T., Ass B.A.P., Ramos L.A. and Frollini E. 2000. Acetylation of cellulose under homogeneous reaction conditions.
In: Mattoso L.H.C., Leo A.L. and Frollini E. (eds), Natural
Polymers and Composites. So Carlos, pp. 139145.
Dawsey T.R. 1994. Applications and limitations of LiCl/N,Ndimethylacetamide in the homogeneous derivatization of cellulose. In: Gilbert R.D. (ed), Cellulosic Polymers, Blends and
Composites. Hanser Publishers, Mnchen, Wien, New York, pp.
157171.
Einfeldt J., Heinze T., Liebert T. and Kwasniewski A. 2002. Influence of the p-toluenesulfonylation of cellulose on the polymer dynamics investigated by dielectric spectroscopy. Carbohyd.
Polym. 49: 357365.
El Seoud O.A., Regiani A. and Frollini E. 2000. Derivatization of cellulose in homogeneous conditions. A Brief Review.
In: Frollini E., Leo A.L. and Mattoso L.H.C. (eds), Natural
Polymers and Agrofibers Composites. So Carlos, pp. 7390.
Erler U., Mischnick P., Stein A. and Klemm D. 1992. Determination
of the substitution pattern of cellulose methyl ethers by HPLC
and GLC comparison of methods. Polym. Bull. 29: 349356.
Gerzon K. and Kau D. 1967. The adamantoyl group in medicinal agents. III. Nucleoside 5 -adamantoates. The adamantoyl
function as a protecting group. J. Med. Chem. 10: 189198.
Grbner D., Liebert T. and Heinze Th. 2002. Synthesis of novel
adamantoyl cellulose using differently activated carboxylic acid
derivatives. Cellulose 9: 193201.
Greene T.W. and Wuts P.G.M. 1991. Protective Groups in Organic
Synthesis. John Wiley & Sons, Inc., New York, p. 100.

Heinze Th. and Liebert T. 2001. Unconventional methods in cellulose functionalization. Prog. Polym. Sci. 26: 16891762.
Heinze Th., Dicke R., Koschella A., Kull A.H., Klohr E.-A. and
Koch W. 2000. Effective preparation of cellulose derivatives in
a new simple cellulose solvent. Macromol. Chem. Phys. 201:
627631.
Heinze T. and Schaller J. 2000. New water soluble cellulose ester synthesized by an effective acylation procedure. Macromol.
Chem. Phys. 201: 12141218.
Heinze T., Liebert T., Klfers P. and Meister F. 1999. Carboxymethylation of cellulose in unconventional media. Cellulose 6:
153165.
Koschella A., Haucke G. and Heinze Th. 1997. New fluorescence
active cellulosics prepared by a convenient acylation procedure.
Polym. Bull. 39: 597604.
Ktz J., Bogen I., Heinze Th., Heinze U., Kulicke W.-M. and
Lange S. 2001. Peculiarities in the physico-chemical behavior
of non-statistically substituted carboxymethylcelluloses. Colloid
Surface. A 183185: 621633.
Liebert T. and Heinze Th. 1998a. Induced phase separation: A
new synthesis concept in cellulose chemistry. In: Heinze Th.J.
and Glasser W.G. (eds), Cellulose Derivatives: Synthesis, Characterization and Nanostructures. ACS Symposium Series 688,
American Chemical Society, Washington, DC, p. 61.
Liebert T. and Heinze T. 1998b. Synthesis path versus distribution
of functional groups in cellulose ethers. Macromol. Symp. 130:
271283.
Liebert T. and Heinze T. 2001. Exploitation of reactivity and
selectivity in cellulose functionalization using unconventional
media for the design of products showing new superstructures.
Biomacromolecules 2: 11241132.
Liebert T., Heinze Th. and Klemm D. 1996. Synthesis and carboxymethylation of organo-soluble trifluoroacetates and formates of
cellulose. J. Macromol. Sci. Pure A33: 613626.
Marson G., Regiani A., Frollini E. and El Seoud O.A. 1999. Cellulose esterification in homogeneous medium. J. Polym. Sci. Pol.
Chem. 37: 13571363.
Marson G., Ciacco G.T., Frollini E. and El Seoud O.A. 2000.
An efficient one pot acylation of cellulose under homogeneous
reaction conditions. Macromol. Chem. Phys. 201: 882889.
Orzeszko A., Gralewska R., Starosciak B.J. and Kazimierczuk Z.
2000. Synthesis and antimicrobial activity of new adamantane
derivatives I. Acta Biochim. Pol. 47: 8794.
Perrakis A., Antoniadou-Vyza E., Tsitsa P., Lamzin V.S., Wilson
K.S. and Hamodrakas S.J. 1999. Molecular, crystal and solution
structure of a beta-cyclodextrin complex with the bromide salt
of 2-(3-dimethylaminopropyl)tricyclo[3.3.1.1(3,7)]decan-2-ol, a
potent antimicrobial drug. Carbohyd. Res. 317: 1928.
Saake B., Horner S., Kruse Th., Puls J., Liebert T. and Heinze
Th. 2000. Detailed investigation on the molecular structure of
carboxymethyl cellulose with unusual substitution pattern by
means of enzyme-supported analysis. Macromol. Chem. Phys.
201: 19962002.
Samaranayake G. and Glasser W.G. 1993a. Cellulose derivatives
with low DS. 1. A novel acylation system. Carbohyd. Polym. 22:
17.
Samaranayake G. and Glasser W.G. 1993b. Cellulose derivatives
with low DS. 2. Analysis of alkanoates. Carbohyd. Polym. 22:
7986.
Sealey J., Samaranayake G., Todd J. and Glasser W.G. 1996. Novel
cellulose derivatives. IV. Preparation and thermal analysis of
waxy ester of cellulose. J. Polym. Sci. 34: 16131620.
Shimizu Y. and Hayashi J. 1988. A new method for cellulose
acetylation with acetic acid. Sen-I Gakkaishi 44: 451456.
Sun R.C., Fang J.M., Tomkinson J., Geng Z.C. and Liu J.C. 2001.
Fractional isolation, physico-chemical characterization and homogeneous esterification of hemicelluloses from fast-growing
poplar wood. Carbohyd. Polym. 44: 2939.

Das könnte Ihnen auch gefallen