You are on page 1of 7

International Journal of Biological Macromolecules 51 (2012) 228–234

Contents lists available at SciVerse ScienceDirect

International Journal of Biological Macromolecules


journal homepage: www.elsevier.com/locate/ijbiomac

Cellulose nanofibrils prepared from softwood cellulose by TEMPO/NaClO/NaClO2


systems in water at pH 4.8 or 6.8
Reina Tanaka, Tsuguyuki Saito, Akira Isogai ∗
Department of Biomaterials Sciences, Graduate School of Agricultural and Life Sciences, The University of Tokyo, Tokyo 113-8657, Japan

a r t i c l e i n f o a b s t r a c t

Article history: Catalytic oxidation of softwood cellulose using NaClO and either 2,2,6,6-tetramethylpiperidine-1-oxyl (4-
Received 5 April 2012 H-TEMPO) or 4-acetamido-TEMPO (4-AcNH-TEMPO) was applied with NaClO2 used as a primary oxidant
Received in revised form 8 May 2012 in an aqueous buffer at pH 4.8 or 6.8. When the 4-AcNH-TEMPO-mediated oxidation was applied to
Accepted 14 May 2012
softwood cellulose in water at pH 4.8 and 40 ◦ C, the carboxylate content rose to ∼1.3 mmol/g after reaction
Available online 19 May 2012
for 48 h and the DPv value was more than 1100. This 4-AcNH-TEMPO-oxidized softwood cellulose was
mostly converted to individual nanofibrils by mechanical disintegration in water, with uniform widths
Keywords:
of 3–4 nm and lengths greater than 1 ␮m.
Cellulose nanofibril
4-Acetamido-TEMPO © 2012 Elsevier B.V. All rights reserved.
Oxidation

1. Introduction been utilized for catalytic oxidation to selectively convert primary


hydroxyls of polysaccharides into carboxylates [12–14]. It has
Cellulose is the most abundant biomass on earth, and as such has previously been reported that 4-H-TEMPO-mediated oxidation of
recently attracted intense interest for use as a renewable carbon wood celluloses provides oxidized celluloses with C6 carboxylate
resource. Wood celluloses are synthesized in cell walls as highly groups that are selectively formed on the surfaces of cellulose
crystalline nanosized fibrils that consist of dozens of direction- nanofibrils [15–17]. These oxidized celluloses can be disintegrated
ally aligned molecular chains [1,2]. These cellulose nanofibrils have to individual cellulose nanofibrils, i.e., TEMPO-oxidized cellulose
high aspect ratios (∼4 nm in width and more than 1 ␮m in length), nanofibrils (TOCNs), by mild mechanical treatment in water.
high stiffness values [3], and low thermal expansion coefficients This is because the high density of carboxylate groups uniformly
[4], while simultaneously possessing low densities. Owing to these introduced on the nanofibril surfaces possess anionic charges in
characteristics, utilization of cellulose nanofibrils as new nanoma- water, and thus the repulsive forces effectively work between
terials has attracted much attention. the cellulose nanofibrils in water to aid separation [8,15,18].
Separation or individualization of cellulose nanofibrils from The TOCN/water dispersions can be converted to cast films with
wood cellulose fibers or pulps is problematic because they are extremely high oxygen-barrier properties [19,20], stiff hydrogels
firmly assembled to one another in the cell walls of the wood via and aerogels with large surface areas [21].
networks of both hydrogen bonds and van der Waals (or hydropho- For the 4-H-TEMPO-mediated oxidation of wood celluloses, the
bic) interactions [5]. Over the past few decades, various procedures TEMPO/NaBr/NaClO system in water at pH 10 has been used in most
to fibrillate wood cellulose fibers have been investigated; for exam- reported cases [8,15]. However, remarkable depolymerization of
ple, harsh mechanical disintegration treatment in water [6], acid cellulose chains is inevitable in this oxidation system at pH 10, and
hydrolysis [7], and various pretreatments of cellulose fibers used thus the TEMPO-oxidized celluloses have viscosity-average degrees
in the hope of increasing the preparation efficiency of cellulose of polymerization (DPv ) that are lower than 600 [22]. Such remark-
nanofibrils [8–11]. able depolymerization may result from ␤-elimination of glycosidic
Catalytic oxidation of cellulose using 2,2,6,6- bonds at the C6 aldehyde groups that are formed as intermedi-
tetramethylpiperidine-1-oxyl (4-H-TEMPO) was found to be a ates in the oxidized celluloses, which takes place under alkaline
useful method that enables wood cellulose fibers to be completely conditions [23], and/or from depolymerization due to some radi-
converted to individual nanofibrils when dispersed in water. cal species formed in situ during the TEMPO-mediated oxidation
4-H-TEMPO is a water-soluble and stable nitroxyl radical that has [24]. In addition, the C6 aldehyde groups are likely to disturb the
nano-dispersibility of the oxidized celluloses by the formation of
hemiacetal linkages between nanofibrils [25].
∗ Corresponding author. Tel.: +81 3 5841 5538; fax: +81 3 5841 5269. On the other hand, it has been recently reported that a
E-mail address: aisogai@mail.ecc.u-tokyo.ac.jp (A. Isogai). TEMPO/NaClO/NaClO2 system under neutral or weakly acidic

0141-8130/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ijbiomac.2012.05.016
R. Tanaka et al. / International Journal of Biological Macromolecules 51 (2012) 228–234 229

R R = H or NHCOCH3

N
O
TEMPO

½ NaClO

CH2OH
N O
O O OH
N-oxoammonium ion

Hydroxylamine
NaCl OH
R

CHO
N
O
OH O OH
NaClO

OH

COOH COONa
NaClO2 O Buffer at O
pH 4.8 or 6.8
O OH O OH

OH OH

Fig. 1. Oxidation of C6 primary hydroxyls to carboxylates by the TEMPO/NaClO/NaClO2 system under neutral or weakly acidic conditions.

conditions [26] is effective for the preparation of TEMPO-oxidized 2. Materials and methods
hardwood celluloses with high DP values [27]. In this system,
sodium hypochlorite oxidizes TEMPO molecules into the corre- 2.1. Materials
sponding N-oxoammonium ions, which then rapidly oxidize the
primary hydroxyls to aldehydes. Sodium chlorite is used as the The softwood and hardwood bleached kraft pulps produced by
primary oxidant and this immediately oxidizes the formed alde- ECF (elementary chlorine-free) sequences were supplied by Nippon
hyde groups into carboxylates in the buffer solutions (Fig. 1). Thus, Paper Industries Co., Ltd., Japan in a never-dried state with water
this system provides oxidized celluloses with no aldehyde groups contents of ∼80%, and used as wood cellulose samples. For dem-
(at least under conditions of pH 4.8) and high DPv values. In our ineralization, the pulp was soaked in dilute HCl at pH 2.5 and room
recent study we found that oxidation using 4-acetamido-TEMPO temperature for 0.5 h, and then washed repeatedly with water by
(4-AcNH-TEMPO) as a catalyst under neutral or weakly acidic con- filtration. The 4-H-TEMPO, 4-AcNH-TEMPO, sodium chlorite, 2 M
ditions is more suitable than using 4-H-TEMPO for the preparation sodium hypochlorite solution, and all other chemicals and solvents
of water-soluble ␤-(1 → 4)-linked polyglucuronic acids, i.e., cel- were of laboratory grade (Wako Pure Chemicals, Tokyo, Japan) and
louronic acids with high DP values in high yields, from regenerated used without further purification.
celluloses [28].
In the present study, the 4-AcNH-TEMPO/NaClO/NaClO2 sys- 2.2. TEMPO-mediated oxidation of wood cellulose
tem was applied to a softwood cellulose at pH 4.8 or 6.8,
and the 4-AcNH-TEMPO-oxidized celluloses were compared with The wood cellulose (1 g) was suspended in 0.1 M sodium phos-
4-H-TEMPO-oxidized ones in terms of carboxylate content, phate buffer (100 mL, pH 6.8) or 0.1 M acetate buffer (100 mL,
DPv , crystallinity, and crystal width of cellulose I. In addition, pH 4.8) dissolving 4-H-TEMPO or 4-AcNH-TEMPO (0.1 mmol) and
softwood and hardwood celluloses oxidized by the 4-AcNH- sodium chlorite (10 mmol) in an airtight flask. A 2 M sodium
TEMPO/NaClO/NaClO2 system were disintegrated in water under hypochlorite solution (0.5 mL, 1.0 mmol) was added to the flask
the same conditions, and their degrees of nanofibrillation, and in a single step. The flask was immediately stoppered and the
lengths and widths of the obtained nanofibrils evaluated by light suspension stirred at 40 or 60 ◦ C for a designated period of
transmittance spectroscopy and transmission electron microscopy time (6–72 h). After cooling the suspension to room tempera-
(TEM), respectively. ture, the oxidized cellulose was thoroughly washed with water by
230 R. Tanaka et al. / International Journal of Biological Macromolecules 51 (2012) 228–234

filtration and freeze-dried. The weight recovery ratios of the 1.5 1500
oxidized celluloses were more than 95% for all of the samples pH 6.8

Carboxylate content (mmol/g)


examined in this study. The carboxylate content of each oxidized
cellulose sample was determined using the electrical conducto-
metric titration method [29]. Some of the oxidized celluloses were 1.0 1000
post-oxidized with 1% NaClO2 in water at pH 4.5 and room tempera-

DPv
ture for 2 days to convert C6 aldehydes, if present, to C6 carboxylate
groups.
Carboxylate content
0.5 500
2.3. Mechanical fibrillation of TEMPO-oxidized celluloses DPv

4-AcNH-TEMPO
The oxidized cellulose was suspended in water (25 mL) to pro-
4-H-TEMPO
vide a solid content of 0.1%. The suspension was homogenized
0.0 0
for 2 min at 7500 rpm with a double-cylinder-type homogenizer 0 20 40 60 80
(Physcotron NS-56, Microtec) and then sonicated for 4 min using an
Reaction time (h)
ultrasonic homogenizer equipped with a 7 mm probe tip (US-300T,
Nihon Seiki) at 300 W of output power. The unfibrillated fraction
1.5 1500
was removed by centrifugation at 12,000 × g over 20 min and the pH 4.8

Carboxylate content (mmol/g)


supernatant subjected to light transmittance and TEM analyses. The
nanofibrillation ratio was calculated from the following equation:
C 
s
Nanofibrillation ratio (%) = 100 × (1) 1.0 1000
Co

DPv
where Cs (g) is the dry weight of TOCN in the supernatant and Co
(g) is that of TEMPO-oxidized cellulose in the original dispersion
before the centrifugation was carried out. 0.5 Carboxylate content 500
DPv

2.4. Analyses
4-AcNH-TEMPO
4-H-TEMPO
TEMPO-oxidized and then freeze-dried celluloses (0.04 g each) 0.0 0
were dissolved in 0.5 M cupriethylene diamine and their intrin- 0 20 40 60 80
sic viscosities measured using a capillary viscometer. These values Reaction time (h)
were converted to viscosity-average DP (DPv ) values by using
a previously reported equation, [] = 0.61(DPv )1.11 [22,30]. Each Fig. 2. Carboxylate content and viscosity-average degree of polymerization
(DPv ) of oxidized celluloses prepared from softwood cellulose by the 4-H-
freeze-dried sample (0.1 g) was converted to a pellet 1 cm in diam-
TEMPO/NaClO/NaClO2 or 4-AcNH-TEMPO/NaClO/NaClO2 system at 40 ◦ C, and at
eter by applying a pressure of about 600 MPa for 1 min. X-ray either pH 6.8 or 4.8 for 6–72 h.
diffraction patterns were recorded for the pellet samples for 2
diffraction angles from 10◦ to 30◦ using the reflection method by
means of a Rigaku RINT 2000 with monochromator-filtered Cu K␣ to C6 carboxylates by post-oxidation with NaClO2 in water at pH
radiation ( = 0.15418 nm) at 40 kV and 40 mA. The crystallinity 4.5 and room temperature for 2 days [22,29]. Thus, according to
and (2 0 0) crystal width of cellulose I were calculated accord- the scheme in Fig. 1, the 4-H-TEMPO- or 4-AcNH-TEMPO-oxidized
ing to conventional methods from the X-ray diffraction patterns celluloses prepared at pH 4.8 and 40–60 ◦ C for 48–72 h are likely to
[31,32]. The TOCN dispersion was put into a disposable poly(methyl have no C6 aldehyde groups.
methacrylate) cuvette and the light transmittance spectrum of the However, the oxidized celluloses prepared at pH 6.8 may have
dispersion measured from 300 to 800 nm with a spectrophotome- had some amounts of C6 aldehydes present because this pH con-
ter (JASCO V-670). A 10 ␮L aliquot of 0.005% TOCN dispersion was dition is not optimum for the conversion of aldehydes to carboxyls
mounted onto a glow-discharged carbon-coated Cu grid. The excess by NaClO2 oxidation. Hence, some of the 4-H-TEMPO- and 4-
liquid was absorbed by filter paper and one drop of 1% uranyl AcNH-TEMPO-oxidized celluloses prepared at pH 6.8 were further
acetate was added for negative staining of the TOCN. Excess solu- oxidized with NaClO2 in water at pH 4.5 and room temperature for
tion was blotted with filter paper and allowed to stand for drying 2 days. However, no increase in carboxylate content was observed
by natural evaporation. The sample grid was observed with a JEOL for the post-oxidized products, showing that the TEMPO-mediated
JEM-2000EX microscope operated at an accelerating voltage of oxidation of wood cellulose mostly proceeds according to the
200 kV. scheme in Fig. 1, even at pH 6.8, and with almost no C6 aldehydes
remaining in the oxidized celluloses.
3. Results and discussion
3.2. Carboxylate content and DPv of the TEMPO-oxidized
3.1. Possibility of the presence of C6 aldehydes in the celluloses
TEMPO-oxidized celluloses
The carboxylate content and DPv values of the oxidized
According to the scheme in Fig. 1, all C6 aldehyde groups cellulose prepared by the 4-H-TEMPO/NaClO/NaClO2 or 4-AcNH-
formed from C6 primary hydroxyls of cellulose by 4-H-TEMPO- or TEMPO/NaClO/NaClO2 system at pH 4.8 or 6.8 for 6–72 h are shown
4-AcNH-TEMPO-mediated oxidation ought to be converted to C6 in Fig. 2. For the oxidation at pH 6.8, more than half of the total
carboxylates through in situ oxidation by the NaClO2 present in the C6 carboxyls were rapidly formed within 6 h, before the carboxy-
reaction medium. It is well known that almost all C6 aldehydes, that late content in both oxidation systems gradually increased with
are formed by the TEMPO/NaBr/NaClO system at pH 10 and remain the increase in oxidation time. However, there was slight dif-
to some extent in the TEMPO-oxidized celluloses, can be converted ferences in carboxylate content between the 4-H-TEMPO- and
R. Tanaka et al. / International Journal of Biological Macromolecules 51 (2012) 228–234 231

4-AcNH-TEMPO-oxidized celluloses prepared at the same oxida- 1500


tion time; 4-AcNH-TEMPO had a little higher efficiency in the
formation of carboxylate groups. The carboxylate contents of the
Original softwood cellulose
4-H-TEMPO- and 4-AcNH-TEMPO-oxidized celluloses increased to
1.1 and 1.2 mmol/g, respectively, after oxidation for 72 h. 4-AcNH-
1000
TEMPO also showed a slightly higher efficiency for the formation
of carboxylate groups than the 4-H-TEMPO system. The DPv values

DPv
of the oxidized celluloses clearly decreased with an increase in the
oxidation time, decreasing from 1500 to 440 and 760 after oxida-
tion for 72 h with 4-H-TEMPO and 4-AcNH-TEMPO, respectively. 500 pH 6.8
pH 4.8
Thus, depolymerization of the cellulose chains is inevitable during
the oxidation in both systems, while 4-AcNH-TEMPO gave oxidized
4-AcNH-TEMPO
celluloses with higher DPv values than 4-H-TEMPO at the oxidation
4-H-TEMPO
times of 48 and 72 h.
0
When the oxidation was carried out at pH 4.8, the formation 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
patterns of carboxylate groups were different from those prepared
at pH 6.8. The carboxylate content of the 4-H-TEMPO-oxidized Carboxylate content (mmol/g)
cellulose was only 0.76 mmol/g even after the reaction for 72 h,
Fig. 3. Relationship between carboxylate content and DPv of the oxidized celluloses.
while 4-AcNH-TEMPO gave oxidized cellulose with carboxylate The original softwood cellulose had a carboxylate content and DPv of 0.05 mmol/g
content as high as 1.3 mmol/g under the same oxidation conditions. and 1500, respectively.
Thus, 4-AcNH-TEMPO is much more effective for the formation of
carboxylate groups in oxidized celluloses prepared at pH 4.8. It
has been reported that the maximum carboxylate content of 4-H- the 4-AcNH-TEMPO/NaClO/NaClO2 system in water at pH 4.8 is
TEMPO-oxidized hardwood celluloses prepared at pH 4.8 and 6.8 further validated as being better for the preparation of oxidized
are 0.6 and 0.8 mmol/g, respectively [27]. Thus, the use of softwood celluloses with higher carboxylate contents and higher DPv values
cellulose and 4-AcNH-TEMPO as the starting wood cellulose and when softwood cellulose is used as the starting material.
the TEMPO catalyst, respectively, provides an improved means of X-ray diffraction patterns of some of the oxidized celluloses are
preparing oxidized wood celluloses with higher carboxylate con- depicted in Fig. 4, together with their crystallinities and (2 0 0)
tent. Although the DPv values of the oxidized celluloses decreased
with the increase in oxidation time in both systems, those prepared
after reaction for 72 h still gave DPv values greater than 1100. Crystallinity Crystal width
The high oxidation efficiency of primary hydroxyls to carboxyls (%) (nm)
with 4-AcNH-TEMPO in comparison with 4-H-TEMPO, particularly
under neutral-weakly acidic conditions, has already been reported e: 4-AcNH-TEMPO at pH 6.8 66 3.6
for cellulose and other saccharides [27,28,33]. Formation of a cova- d: 4-AcNH-TEMPO at pH 4.8 69 3.4
c: 4-H-TEMPO at pH 6.8 68 3.6
lent bond between the N-oxoammonium ion and dissociated C6
b: 4-H-TEMPO at pH 4.8 74 3.8
primary alcohol of cellulose is the initial step required for the
a: Original softwood cellulose 73 3.9
formation of the C6 carboxylate groups [15,34]. Thus, the results
in Fig. 2 indicate that the covalent bond formation may have
occurred in a similar manner at pH 6.8 between 4-AcNH-TEMPO
and 4-H-TEMPO, but that its occurrence was favored for 4-AcNH-
TEMPO over 4-H-TEMPO at pH 4.8. The difference in redox potential
between the two TEMPOs is not directly related to the difference
in oxidation efficiency [35], and thus a detailed mechanism for the
different oxidation efficiency between the two TEMPOs remains
unknown.
The higher DPv values for the oxidized celluloses prepared at
pH 4.8 than those at pH 6.8 may be explained in terms of the
higher stability of ␤-(1 → 4)-glycoside bonds of polyuronates under e
weakly acidic conditions than when under neutral conditions [36].
d
Depolymerization due to ␤-elimination, which easily takes place
on polysaccharides possessing C6 aldehydes under alkaline condi-
tions, can, to some extent, occur even under neutral conditions at
c
pH 6.8 [23]. Thus, the TEMPO-mediated oxidation at pH 4.8 may
somewhat have restricted depolymerization due to ␤-elimination
b
compared to that at pH 6.8, resulting in the preparation of oxidized
celluloses with higher DPv values (Fig. 2).
When the oxidation temperature was increased from 40 to 60 ◦ C,
a
the carboxylate contents of the 4-AcNH-TEMPO-oxidized cellu-
loses prepared at pH 4.8 and 6.8 for 72 h increased to 1.37 and
1.34 mmol/g, respectively. However, the DPv values decreased in 10 15 20 25 30
turn to 450 and 550, respectively, and thus oxidation at 40 ◦ C is
better than at 60 ◦ C for the preparation of oxidized celluloses with Diffraction angle 2θ ( ° )
higher DPv values [28,37].
Fig. 4. X-ray diffraction patterns of 4-H-TEMPO- and 4-AcNH-TEMPO-oxidized cel-
The relationships between the carboxylate content and DPv of luloses prepared at pH 4.8 or 6.8. Their crystallinities and (2 0 0) crystal widths of
the oxidized celluloses shown in Fig. 2 are replotted in Fig. 3. Again, cellulose I are also shown.
232 R. Tanaka et al. / International Journal of Biological Macromolecules 51 (2012) 228–234

Fig. 5. UV–Vis transmittance spectra (A) of 0.1% hardwood and softwood TEMPO-oxidized cellulose nanofibrils (TOCNs) dispersed in water, and their photographic images
(B) before centrifugation. Photographic images (C) show the dispersions taken between cross-polarizers after removal of unfibrillated fractions by centrifugation at 12,000 × g
for 20 min.

crystal widths of cellulose I calculated from the diffraction pat- can be explained in terms of the different carboxylate content
terns. Slight decreases in both crystallinity and crystal width were in these samples. The nanofibrillation ratios were 88% and 44%
observed for the oxidized celluloses compared with the original for the softwood and hardwood oxidized celluloses, respectively.
softwood cellulose. However, the decreased levels are not partic- The carboxylate content of the TEMPO-oxidized cellulose strongly
ularly remarkable, and it can thus be concluded that almost all influences the nano-dispersibility observed when mechanical dis-
carboxyl groups formed by the oxidation are present on the surfaces integration treatment in water is applied to the oxidized celluloses
of crystalline wood cellulose microfibrils. under equivalent conditions [18]. This is evidenced by the fact that
the higher the carboxylate content, the higher the nanofibrillation
3.3. Nano-dispersibility of TEMPO-oxidized celluloses in water ratio obtained. The results in Fig. 5 indicate that crystalline cel-
lulose microfibrils in hardwood kraft pulp are partly surrounded
The softwood and hardwood celluloses were both by xylan molecules that are insusceptible to the TEMPO-mediated
oxidized under the same conditions using the 4-AcNH- oxidation.
TEMPO/NaClO/NaClO2 system at pH 4.8 for 48 h. The oxidized The softwood TOCN/water dispersion at 0.088% solid content
softwood and hardwood celluloses have carboxylate contents of shows clear birefringence when observed between crossed polar-
1.3 and 0.7 mmol/g, respectively. Both the softwood and hardwood izers, even at rest (right hand side of Fig. 5C). This results from the
bleached kraft pulps used in this study had ␣-cellulose contents of presence of individual TOCNs dispersed in water [8,15,18,21,27].
approximately 90%, while the remaining samples were hemicellu- Because the solid content of the hardwood TOCN/water disper-
loses, i.e., primarily glucomannan and xylan for the softwood and sion after removal of the unfibrillated fraction decreased to 0.044%,
hardwood, respectively. Thus, the difference in chemical structure almost no birefringence was observed at such a low TOCN concen-
of hemicelluloses between the two wood celluloses is likely to have tration (left hand side of Fig. 5C).
resulted in the large difference in carboxylate content between the Fig. 6 shows TEM images of the hardwood and softwood TOCNs,
two oxidized celluloses prepared. Glucomannan in the softwood which were prepared from the supernatants after centrifugation
cellulose has C6 hydroxyls that are oxidized to C6 carboxylates of the dispersions. Both hardwood and softwood TOCNs show uni-
upon oxidation, while xylan in the hardwood cellulose has no C6 form nanofibril widths of 3–4 nm. The few kinks observed along the
hydroxyls present. long TOCNs are likely to have been formed during the mechanical
To prepared TOCN dispersions, the oxidized softwood and hard- disintegration treatment. The TEM images in Fig. 6 indicate that
wood celluloses were then mechanically disintegrated in water the TOCNs prepared in this study had average lengths greater than
under the same conditions. Fig. 5A and B shows UV–Vis trans- 1 ␮m. However, accurate lengths and length distributions of the
mittance spectra and photographic images, respectively, of the TOCNs could not be measured by TEM because the TOCN widths are
0.1% hardwood and softwood TOCN dispersions before centrifu- too small and large in width and length, respectively, such that no
gation. The softwood TOCN/water dispersion clearly shows higher nanofibrils could not be observed at low magnification to determine
optical transmittance than the hardwood TOCN dispersion, which the lengths of the TOCNs.
R. Tanaka et al. / International Journal of Biological Macromolecules 51 (2012) 228–234 233

Fig. 6. TEM images of hardwood and softwood TOCNs.

If the relationship between the DPv and average length of the Acknowledgments
TOCNs prepared by the TEMPO/NaBr/NaClO oxidation at pH 10 [22]
is applicable to the softwood TOCNs prepared in this study by the This study was supported by Scientific Research S (21228007)
TEMPO/NaClO/NaClO2 system at pH 4.8 and 6.8, then the length- and Encouragement of Young Scientists A (23688020) from the
weighted average lengths of the TOCNs with DPv > 1100 in Fig. 2 Japan Society for the Promotion of Science (JSPS).
can be estimated to be more than 3.5 ␮m. Thus, mostly individ-
ual TOCNs with approximate carboxylate contents of 1.3 mmol/g,
and with width of 3–4 nm and high aspect ratios (at least >250), References
can be prepared in high yields from softwood cellulose by the 4-
AcNH-TEMPO/NaClO/NaClO2 oxidation in water at pH 4.8 and 40 ◦ C [1] I.M. Saxena, M. Brown Jr., Annals of Botany 96 (2005) 9–21.
[2] H.F. Jacob, D. Fengel, S. Tschegg, P. Fratzl, Macromolecules 28 (1995)
for 48–72 h and upon subsequent mechanical disintegration of the
8782–8787.
oxidized celluloses in water. For TOCNs to be used as nanofillers [3] I. Sakurada, Y. Nukushima, T. Ito, Journal of Polymer Science 57 (1962) 651–660.
in high-strength composite materials, those prepared in this study [4] R. Hori, M. Wada, Cellulose 12 (2005) 479–484.
with high aspect ratios and high DPv values would be advantageous. [5] Y.-S. Liu, J.O. Baker, Y. Zeng, M.E. Himmel, T. Hass, S.-Y. Ding, Journal of Biological
Chemistry 286 (2011) 11195–11201.
[6] A.F. Turbak, F.W. Snyder, K.R. Sandberg, Journal of Applied Polymer Science.
Applied Polymer Symposium 37 (1983) 815–827.
4. Conclusions [7] R.H. Marchessault, F.F. Morehead, N.M. Walter, Nature 184 (1959) 632–633.
[8] T. Saito, Y. Nishiyama, J.L. Putaux, M. Vignon, A. Isogai, Biomacromolecules 7
(2006) 1687–1691.
In the case of the oxidation at pH 6.8, the carboxylate con- [9] L. Wågberg, G. Decher, M. Norgren, T. Lindström, M. Ankerfors, K. Axnäs, Lang-
tent of the 4-H-TEMPO- and 4-AcNH-TEMPO-oxidized celluloses muir 24 (2008) 784–795.
[10] M. Henriksson, G. Henriksson, L.A. Berglund, T. Lindström, European Polymer
both increased to approximately 1.1 mmol/g after reaction for 72 h,
Journal 43 (2007) 3434–3441.
whereas the DPv values decreased to 760 or lower. In the oxidation [11] M. Pääkkö, M. Ankerfors, H. Kosonen, A. Nykänen, S. Ahola, M. Österberg, J.
at pH 4.8, on the other hand, 4-AcNH-TEMPO was more effective Ruokolainen, J. Laine, P.T. Larsson, O. Ikkala, T. Lindström, Biomacromolecules
in the formation of carboxylate groups in the oxidized celluloses 8 (2007) 1934–1941.
[12] A.E.J. de Nooy, A.C. Besemer, H. van Bekkum, Carbohydrate Polymers 49 (1995)
than was 4-H-TEMPO, and the 4-AcNH-TEMPO-oxidized celluloses 397–406.
with carboxylate contents of approximately 1.3 mmol/g had high [13] P.L. Bragd, A.C. Besemer, H. van Bekkum, Carbohydrate Research 328 (2002)
DPv values of more than 1100. Crystallinities and crystal widths 355–363.
[14] P.L. Bragd, H. van Bekkum, A.C. Besemer, Topics in Catalysis 27 (2004) 49–66.
of cellulose I in the oxidized celluloses were slightly lower than [15] A. Isogai, T. Saito, H. Fukuzumi, Nanoscale 3 (2011) 71–85.
those of the original wood cellulose. TEM observations showed [16] Y. Okita, T. Saito, A. Isogai, Biomacromolecules 11 (2010) 1696–1700.
that the TOCNs prepared in this study have uniform widths of [17] M. Hirota, K. Furihata, T. Saito, T. Kawada, A. Isogai, Angewandte Chemie Inter-
national Edition 49 (2010) 7670–7672.
3–4 nm and average lengths greater than 1 ␮m. Thus, the 4-AcNH- [18] T. Saito, S. Kimura, Y. Nishiyama, A. Isogai, Biomacromolecules 8 (2007)
TEMPO/NaClO/NaClO2 oxidation of softwood cellulose in water at 2492–2496.
40 ◦ C and pH 4.8 is advantageous for the preparation of TOCNs with [19] H. Fukuzumi, T. Saito, T. Iwata, Y. Kumamoto, A. Isogai, Biomacromolecules 10
(2009) 162–165.
high aspect ratios and high DPv values, and in high nanofibrillation
[20] H. Fukuzumi, T. Saito, S. Iwamoto, Y. Kumamoto, T. Ohdaira, R. Suzuki, A. Isogai,
yields. Biomacromolecules 12 (2011) 4057–4062.
234 R. Tanaka et al. / International Journal of Biological Macromolecules 51 (2012) 228–234

[21] T. Saito, T. Uematsu, S. Kimura, T. Enomae, A. Isogai, Soft Matter 7 (2011) [31] L. Segal, J.J. Creely, A.E. Martin Jr., C.M. Conrad, Textile Research Journal 29
8804–8809. (1959) 786–794.
[22] R. Shinoda, T. Saito, Y. Okita, A. Isogai, Biomacromolecules 13 (2012) 842–849. [32] L.E. Alexander, X-ray Diffraction Methods in Polymer Science, Kreiger, New
[23] A. Potthast, S. Schiehser, T. Rosenau, M. Kostic, Holzforschung 63 (2009) 12–17. York, 1969, p. 423.
[24] I. Shibata, A. Isogai, Cellulose 10 (2003) 151–158. [33] P.L. Bragd, A.C. Besemer, H. van Bekkum, Journal of Molecular Catalysis A:
[25] T. Saito, A. Isogai, Colloids and Surfaces A 289 (2006) 219–225. Chemical 170 (2001) 35–42.
[26] M. Zhao, J. Li, E. Mano, Z. Song, D.M. Tschaen, E.J.J. Grabowski, P.J. Reider, Journal [34] W.F. Bailey, J.M. Bobbitt, K.B. Wiberg, Journal of Organic Chemistry 72 (2007)
of Organic Chemistry 64 (1999) 2564–2566. 4504–4509.
[27] T. Saito, M. Hirota, N. Tamura, S. Kimura, H. Fukuzumi, L. Heux, A. Isogai, [35] S. Iwamoto, W. Kai, T. Isogai, T. Saito, A. Isogai, T. Iwata, Polymer Degradation
Biomacromolecules 10 (2009) 1992–1996. and Stability 95 (2010) 1394–1398.
[28] M. Hirota, N. Tamura, T. Saito, A. Isogai, Carbohydrate Polymers 78 (2009) [36] S. Fujisawa, T. Isogai, A. Isogai, Cellulose 17 (2010) 607–615.
330–335. [37] A. Kantouch, A. Hebeish, M.H. El-Rafie, Textile Research Journal 40 (1970)
[29] T. Saito, A. Isogai, Biomacromolecules 5 (2004) 1983–1989. 178–184.
[30] R. Evans, F.A.A. Wallis, Journal of Applied Polymer Science 37 (1989)
2331–2340.

You might also like