You are on page 1of 14

Cellulose (2016) 23:451–464

DOI 10.1007/s10570-015-0813-x

ORIGINAL PAPER

The thermal stability of nanocellulose and its acetates


with different degree of polymerization
Melissa B. Agustin . Fumiaki Nakatsubo .
Hiroyuki Yano

Received: 20 August 2015 / Accepted: 2 November 2015 / Published online: 6 November 2015
Ó Springer Science+Business Media Dordrecht 2015

Abstract Geared towards reinforcing thermoplas- for nanocellulose with high DPv, the thermal stabi-
tics of high melting points with nanocellulose, this lization caused by acetylation is mainly due to
study evaluated the factors affecting the thermal protection of the surface OH.
properties of and the thermal stabilizing effect of
acetylation on nanocellulose with different average Keywords Nanocellulose  Thermal stability 
degree of polymerization (DPv) from bacterial cellu- Degree of polymerization  Acetylation  Reducing
lose (BC). Cellulose nanocrystals with DPv values of ends
300 and 500 were prepared by hydrolyzing BC
nanofibers with a DPv of 1100 using hydrochloric
acid. The thermal stability decreased after acid
hydrolysis and showed a decreasing trend with Introduction
decreasing DPv. The decrease in thermal stability is
attributed to the increase in the number of reducing There has been significant growth in nanocellulose
ends (REs) with decreasing DPv. Heterogeneous research in the past decade, which is evident from the
acetylation to an average degree of substitution of numerous publications and patents involving this
0.38 improved the thermal stability, and the degree of material. The worldwide attention paid to nanocellu-
improvement increased with decreasing DPv. The lose can be attributed to its several unique properties.
dependence of the degree of improvement on the DPv Nanocellulose exhibits large surface area, low density,
is attributed to possible protection of the REs by more high mechanical strength, low coefficient of thermal
stable acetyl groups. The influence of protecting the expansion, and wide chemical-modification capacity
REs on the degree of improvement in thermal stability (Klemm et al. 2011).
was further confirmed by sodium borohydride Nanocellulose, which generally refers to all types
(NaBH4) reduction. The findings suggest that the of nanometric cellulosic substrates can be classified
thermal stabilization caused by acetylation to nanocel- into three main subcategories on the basis of their
lulose with small DPv is a combined effect of dimensions, functions and preparation methods
protecting both the surface OH and the REs; while (Klemm et al. 2011; Habibi 2014). (1) Cellulose
nanofibers (CNFs) produced by mechanically disinte-
grating cellulose fiber bundles, typically from wood,
M. B. Agustin (&)  F. Nakatsubo  H. Yano
into nanoscale fibrils with diameters ranging from 10
Research Institute for Sustainable Humanosphere, Kyoto
University, Gokasho, Uji, Kyoto 611-0011, Japan to 100 nm and lengths varying by several microme-
e-mail: agustin.melissa.78a@st.kyoto-u.ac.jp ters. (2) Cellulose nanocrystals (CNCs) obtained by

123
452 Cellulose (2016) 23:451–464

hydrolyzing the amorphous part of the cellulose with therefore needs only low processing temperature. To
strong acid. CNCs usually consist of elongated highly expand the use of nanocellulose to reinforce thermo-
crystalline rod-like nanoparticles with lengths typi- plastic matrices of high melting point such as
cally shorter than those of CNFs. CNCs also have polyamide 6, there is a need to produce thermally
lower degree of polymerization (DP) as a result of the stable nanocellulose. It is also viewed that having
cleavage of glycosidic bonds during hydrolysis. (3) thermally stable nanocellulose can assure long term
Bacterial cellulose (BC), synthesized by several durability and performance reliability of the end
bacterial species in the form of pellicles consisting product. To realize such application and vision,
of a network of randomly assembled ribbon shaped enhancing the thermal stability of nanocellulose is of
fibrils less than 100 nm wide. BC with relatively utmost importance.
straight, continuous, and dimensionally uniform The thermal stability of nanocellulose is influenced
(Habibi 2014) nanofibers can also be acid hydrolyzed by several factors, from the source to the processing
to produce CNCs. methods used to isolate the nanocellulose to the post-
Nanocellulose in various forms has applications as treatment methods used for purification or modifica-
reinforcing fillers in polymer composite preparation. tion, which can have profound effects on its thermal
Nanocellulose has been reported to improve mechan- properties. There have been studies that reported the
ical properties, enhance optical transparency and thermal stability of nanocellulose as affected by
reduce thermal expansion of the resulting composites. drying techniques (Quiévy et al. 2010; Peng et al.
One of the pioneering works in this field was the report 2013; Uetani et al. 2014), mechanical processing
by Favier et al. (1995) on the reinforcement of (Quiévy et al. 2010; Jacquet et al. 2011), bleaching
poly(styrene-co-n-butyl acrylate) (PBA) latex films (Jonoobi et al. 2009), delignification (Okahisa et al.
with cellulose whiskers from tunicates. A PBA film 2011), hydrolysis conditions (Martı́nez-Sanz et al.
with 6 % whiskers prepared by casting showed an 2011; Camarero Espinosa et al. 2013; Yu et al. 2013),
almost two orders of magnitude higher shear modulus and chemical treatments (Kabir et al. 2013). Improv-
than the non-reinforced PBA film. Yano et al. (2005) ing the thermal stability of nanocellulose, especially
pioneered the reinforcement of transparent thermoset- that of the CNCs which have been obtained by acid
ting resins by BC nanofibers. The BC nanofiber- hydrolysis, however, is still a challenge that needs
reinforced composites were optically transparent, further investigation.
showed low coefficient of thermal expansion, and One of the strategies that can be used to improve
exhibited mechanical strength five times that of thermal stability of nanocellulose is chemical modi-
engineered plastic. Intensive and informative reviews fication. There were already numerous chemical
on the application of nanocelluloses as reinforcing modifications reported for nanocellulose but acetyla-
fillers have also been published (Eichhorn et al. 2009; tion is considered the simplest and most widely used.
Siqueira et al. 2010; Klemm et al. 2011; Abdul Khalil Most of the reports that used acetylation to modify the
et al. 2012; Brinchi et al. 2013). properties of nanocellulose aimed, however, to
Despite remarkable developments in the past improve compatibility of the nanocellulose to the
decades, the processing technology of polymer rein- matrix (Ashori et al. 2014; Fahma et al. 2014), render
forced with nanocellulose is restricted to about 200 °C the surface of the nanocellulose more hydrophobic
(Azizi Samir et al. 2005). This limitation is attributed (Nogi et al. 2006; Tomé et al. 2011; Ashori et al. 2014;
to the poor thermal stability of cellulose which starts to Cunha et al. 2014; Fahma et al. 2014), or propose
degrade at about 230 °C (Klemm et al. 2011). During alternative methods of acetylation (Çetin et al. 2009;
compounding, the high shearing force needed to attain Li et al. 2009; Nishino et al. 2011; Hu et al. 2011;
a good dispersion of the filler into the matrix generates Ávila Ramı́rez et al. 2014; Božič et al. 2015). Only a
heat that eventually causes the actual temperature of limited number of papers focused on the thermal
the melt to be higher than the set processing temper- stability of acetylated nanocellulose before incorpo-
ature. Thus, to avoid the actual temperature of the melt ration to the polymer matrix. Moreover, studies that
to increase to as high as the degradation temperature of assessed thermal stability usually dealt with only one
cellulose, the polymer matrix for nanocellulose has type of nanocellulose, CNFs or CNCs alone (Hu et al.
been limited to those with low melting points, and 2011; Lee et al. 2011; Ávila Ramı́rez et al. 2014). So

123
Cellulose (2016) 23:451–464 453

far, the thermal stability of nanocellulose with differ- purchased from Alfa Aesar (Ward Hill, MA) and Kanto
ent DP values before and after acetylation has not been Chemical Co. (Tokyo, Japan), respectively. All chem-
reported. Our interest to investigate the effect of DP on icals were of reagent grade and used without prior
thermal stability lies from the fact that the DP is purification.
directly related to the number of cellulose reducing
ends (REs), which was reported to play a catalytic role Preparation of nanocellulose samples
during pyrolysis (Matsuoka et al. 2014). Moreover, DP
which is one of the properties that differentiate CNCs BC nanofibers
from CNFs may play an important role during
chemical modifications because of the reactivity of The preparation of BC nanofibers followed the
the REs. It is our view that understanding the role of scheme presented in Fig. 1. Approximately 5 kg of
DP in the thermal stabilization of nanocellulose by BC pellicles (2 9 292 cm3) with fiber content of
acetylation can be helpful in designing chemical *0.8 % were washed twice with 1 L of 0.05 M
modification procedure that will be effective in NaOH to partially remove the acetic acid present in the
thermally stabilizing each type of nanocelluloses. absorbed culture media, and then washed repeatedly
In this study, nanocellulose in the form of with distilled water. Every 300 g of BC pellicles was
nanofibers and nanocrystals with different DP values then disintegrated using a Waring blender operating at
were prepared from BC by mechanical disintegration 37,000 rpm (maximum speed) for 3 min. The disin-
and acid hydrolysis, respectively. BC was chosen as tegrated BC was then neutralized by 0.5 M NaOH and
the source of nanocellulose to eliminate the effect of centrifuged at 7500 rpm for 6 min. The collected BC
lignin and hemicellulose. The changes in morphology, residue was then heated in 1 L of 2 % sodium lauryl
crystallinity, and the number of reducing ends (REs) sulfate at 80 °C for 1 h with constant stirring using a
because of acid hydrolysis and their effect on thermal mechanical stirrer to remove residual proteins and
stability were investigated. As a preliminary chemical lipids from the culture media. The mixture was then
modification step, acetylation was carried out and centrifuged, heated in 0.5 M NaOH at 80 °C for 1 h to
controlled in order to obtain acetylated nanocellulose
with similar degree of substitution (DS). Finally, the
effect of acetylation on thermal stability and the BC pellicles
influence of REs to the degree of improvement in
thermal stability of acetylated nanocellulose with Washing with 0.05M NaOH to
different DP values were investigated in detail. partially remove acetic acid

Disintegration
Experimental
Neutralization
Materials

Unpurified BC pellicles were obtained from Fujicco Co. Sodium lauryl sulfate treatment
to remove residual lipids and
Ltd (Kobe, Japan). Acetyl chloride and bis(ethylenedi- proteins
amine) copper(II) hydroxide (CED) solution were
purchased from Sigma-Aldrich Co. (St Louis, MO).
NaOH treatment to further
N-methyl-2-pyrrolidone (NMP), hydrochloric acid,
remove bacterial cells
sodium hydroxide, dehydrated pyridine, ethanol, t-
butanol, L-serine, Na2CO3, and NaHCO3 were supplied
Washing until neutral
by Wako Pure Chemicals Industries (Osaka, Japan).
Sodium lauryl sulfate, sodium borohydride (NaBH4),
and glucose were purchased from Nacalai Tesque Inc. BC nanofibers
(Kyoto, Japan). The 2,20 -bicinchoninic acid (BCA)
disodium salt and copper sulfate pentahydrate were Fig. 1 Scheme for the preparation of BC nanofibers

123
454 Cellulose (2016) 23:451–464

further remove bacterial cells, and finally washed until the different conditions and the resulting DPv values
pH neutral by repeated centrifugation. The neutralized presented in Fig. 2, hydrolysis for 4 h at 50 °C was
residue was further suspended in water and vacuum found optimum to produce BC500. There were several
filtered to obtain a BC mat. The percentage of BC in conditions that can produce BC300 but hydrolysis at
the mat was obtained after subtracting the percent 70 °C using the same acid concentration (2 M) and
moisture determined by infrared moisture determina- duration (4 h) as the preparation of BC500 was
tion balance (FD720, Kett US, Villa Park, CA). selected. In such a way, only the temperature was
Viscosity measurement showed that the disintegrated the differentiating factor in their preparation. More-
BC had an average degree of polymerization (DPv) of over, visual inspection of hydrolyzed BC at shorter
1100 and these BC nanofibers will be referred to as duration still showed long and agglomerated fibers.
BC1100 in succeeding sections. Thus, to ensure the liberation of rod like nanocrystals,
longer hydrolysis time was chosen.
BC nanocrystals Briefly, the preparation of BC300 and BC500 using
the selected conditions is described. A 6 g (dry
CNCs were then prepared from BC1100 by acid weight) sample of BC1100 was placed in a three-
hydrolysis using hydrochloric acid to prevent the necked flask and was suspended in water to give a
introduction of sulfonate esters on the surface of the final concentration of 1 g BC/25 mL H2O. The
crystals. Initially, time series experiments with differ- suspension was then stirred and 300 mL 3 M HCl
ent hydrolysis conditions were conducted to monitor was added to give a final concentration of 2 M HCl in
the change in the DPv with time and to determine the the mixture. The hydrolysis setup was then con-
leveling off of DP (LODP). The LODP is the relatively structed by attaching a thermometer, condenser, and
constant DP value obtained after continued hydrolysis. N2-containing balloon to the openings of the three-
Figure 2 shows the different conditions used and the necked flask. The flask containing the suspension was
DPv values obtained. The LODP for BC was found to then placed in an oil bath preheated to 50 °C for
be *300. Based from the results, it was decided to BC500 or 70 °C for BC300, and the suspension was
prepare BC nanocrystals that reached the LODP and stirred continuously. After 4 h, the hydrolyzed BC
one with a DPv value of 500, which is between 1100 was then diluted with cold distilled water, centrifuged
and 300. For easy reference in the following discus- at 10,000 rpm for 10 min twice to remove most of the
sion, these samples will be referred to as BC500 and HCl, neutralized with 0.1 M NaOH, and washed again
BC300, where the numbers represent the DPv. From with distilled water by repeated centrifugation. The
residue from the last centrifugation was collected,
suspended in water, and finally filtered to form a
1200 nanocellulose mat.

1000 2M, room temperature DPv measurement


2M, 50 °C
2M, 70 °C
800 4M, 70 °C The DPv of pretreated BC before and after hydrolysis
was determined by viscosity measurements. In a 50 mL
DPv

600 vial, 100 mg of the freeze dried nanocellulose sample


was suspended in 10 mL of distilled water and stirred
400
for 30 min under N2 gas. To dissolve the nanocellulose,
10 mL of 1 M CED was added and stirred for 30 min.
200
The resulting solution was then transferred to a Cannon–
0
Fenske capillary viscometer to determine the intrinsic
0 1 2 3 4 5 viscosities of each sample at 25 °C, and the values were
Time (h) then converted to the DPv using the Mark–Houwink–
Sakurada equation [g] = 0.57 9 DPv, where [g] is the
Fig. 2 Change in the average degree of polymerization (DPv)
of bacterial cellulose with hydrolysis time for different acid intrinsic viscosity and 0.57 is the solvent–polymer
concentrations and temperatures constant between cellulose and CED.

123
Cellulose (2016) 23:451–464 455

Acetylation of 4 cm-1 after background correction using a Perkin


Elmer Spectrum One spectrometer equipped with a
Before acetylation, water in the nanocellulose was universal attenuated total reflectance accessory. The
removed by distillation with NMP (Nakatsubo et al. derived spectra were baseline corrected and normal-
2010). To 200 mg BC (dry weight, 1.23 mmol) in a ized before integrating the peak at around 1730 cm-1,
round bottom flask, 60 mL of NMP was added and the which corresponds to the absorption band of carbonyl
suspension was stirred to completely disperse the BC. groups. The area (A) of the carbonyl group for each
Distillation was carried out at 150 °C and 65 hPa. The sample and the calculated DS in the range of 0.3–0.7
resulting water-free suspension (50 mL) of nanocel- from the back-titration method were then plotted, and
lulose in NMP was then cooled to room temperature. the equation DS = 0.0095A ? 0.0193 (r2 = 0.9993)
Pyridine (1.4 mL, 17.8 mmol) and acetyl chloride was determined from the graph. The equation was then
(1.0 mL, 14.8 mmol) were then added and acetylation used to calculate the DS of other acetylated nanocel-
was carried out with continuous stirring at room luloses. It must be noted however that the given
temperature or 50 °C depending on the reactivity of equation maybe applicable only for the samples used
the sample. A 5 mL aliquot of the reaction mixture herein. Constructing a calibration curve for each type
was drawn at various times, diluted with ethanol, of starting material and types of ester is recommended.
filtered, washed repeatedly with ethanol, solvent
exchanged in t-butanol, and then freeze-dried. The Microscopy analysis
collected samples were used to monitor the progress of
the reaction by determining the degree of substitution The morphological changes in BC caused by acid
(DS) using the equation derived from the calibration hydrolysis and acetylation were observed with a field-
curve, which will be discussed in the next section. emission scanning electron microscope (JSM-6700F,
JEOL Ltd., Tokyo, Japan). One drop of 0.01 %
DS determination nanocellulose (in water or in dichloromethane for
non-acetylated and acetylated nanocellulose, respec-
A calibration curve that relates DS to the area of the tively) was placed on a glass plate attached to the
carbonyl group in the Fourier transform infrared sample holder. The sample was then allowed to dry at
(FTIR) spectra of acetylated BC1100 was first con- 40 °C, sputter-coated with platinum, and viewed at an
structed to facilitate the rapid determination of DS accelerating voltage of 1.5 kV.
when monitoring the progress of the reaction during
acetylation. Crystallinity
The DS was first determined by the back-titration
method. For this purpose, time-series acetylation was X-ray diffraction (XRD) analysis was performed to
carried out in 1.50 g BC1100, and about 300 mg was determine the relative crystallinity of cellulose with
collected each time and freeze-dried. To 100 mg different DPv values before and after acetylation.
acetylated BC in an Erlenmeyer flask, 15 mL of ethanol Acetylated nanocellulose (100 mg) was first hydro-
was added and stirred until the BC was uniformly lyzed in 0.2 M NaOH for 1 h at 60 °C to remove the
dispersed. Ten milliliters of 0.5 M NaOH was then acetyl groups, cooled, washed with water and freeze
added, tightly capped, and heated in an oil bath at 70 °C dried. Pellets of freeze-dried nanocellulose samples
for 30 min. The mixture was cooled and the excess prepared by pressing at 1.4 MPa were mounted into the
NaOH was titrated with 0.2 M HCl to a phenolphthalein sample holder and irradiated with nickel filtered CuKa
endpoint. The amount of acetyl groups per gram of (k = 0.154 nm) radiation generated by UltraX 18HF
sample (mol/g) was then calculated and converted to the (Rigaku Corp., Tokyo, Japan) operating at 40 kV and
DS, which is the number of moles of ester groups (acetyl 300 mA. The scattered radiation was detected in the
groups in this study) per mole of anhydroglucose unit. range 5°–40° using an X-ray goniometer that collected
Each acetylated sample was analyzed in duplicate. data at a scan rate of 1°/min with increments of 0.02° in
To construct the calibration curve, each acetylated reflection mode. The crystallinity index was deter-
BC was subjected to FTIR analysis. A portion of the mined from the normalized diffraction profiles using
freeze-dried BC was scanned 16 times at a resolution the equation developed by Segal et al. (1959).

123
456 Cellulose (2016) 23:451–464

Reducing end (RE) analysis The sample was allowed to equilibrate at 110 °C for
10 min to remove bound moisture. Dynamic TGA was
The number of REs expressed as glucose equivalent in done by increasing the temperature from 110 to
lmol/g was analyzed based on the BCA assay reported 500 °C at a rate of 10 °C/min. Isothermal TGA was
by Johnston et al. (1998). The BCA working reagent carried out by increasing the temperature to 250 °C
was prepared similar to the method of Garcia et al. and keeping it isothermally for 1 h. The flow rate of
(1993). The nanocellulose suspension was prepared by nitrogen in both cases was maintained at 100 mL/min
adding 2 mL distilled deionized water into 2 mg throughout the test. Each sample was run in duplicate.
nanocellulose contained in a glass tube. A small
magnetic bar was dropped into the solution and the
mixture was stirred for 30 min to disperse the Results and discussion
nanocellulose. BCA working reagent (2 mL) was then
added, the tube was capped, and the mixture was Acetylation
heated at 80 °C in a water bath with continuous
stirring for 30 min. The mixture was cooled to room To compare the thermal stability of acetylated
temperature, transferred to centrifuge tubes, and nanocelluloses, it is important that each material has
centrifuged at 7500 rpm for 6 min. The absorbance similar DS value to eliminate the effect of unequal
of the supernatant was then read at 560 nm using a amounts of acetyl groups in a sample. For the purpose
U-2910 Hitachi UV–Vis spectrophotometer (Tokyo, of obtaining similar DS, time series acetylation was
Japan). The number of REs was determined from the first carried out and the change in DS was monitored
calibration curve obtained from standard glucose with time.
solutions (0–50 lM) treated as above. Figure 3 shows the increase in DS of each sample
as acetylation progressed. At room temperature, the
NaBH4 reduction reactivity of BC1100 and BC500 were nearly the same
and slightly higher than that of BC300. In of the first
The REs of each nanocellulose samples were reduced 4 h, the DS of BC1100 and BC500 increased to 0.38
by treatment with NaBH4 following the procedure while that of BC300 only reached 0.26. From the
reported by Matsuoka et al. (2011a). To 200 mg of observed DS values at room temperature, the highest
freeze dried nanocellulose dispersed in 8 mL of 0.1 M value of 0.38 was selected to prepare acetylated
Na2CO3 buffer (pH 10) was added 10 mg NaBH4. The nanocelluloses with different DPv values. We did not
mixture was purged with N2, and the reaction was
carried out for 5 h at 80 °C with occasional stirring.
0.6
The reaction was terminated by the addition of 1 N
HCl dropwise until pH 4. The terminated reaction was
Degree of Substitution (DS)

0.5
further heated at 80 °C for 30 min before neutralizing
with saturated NaHCO3 solution. The NaBH4-reduced
0.4
nanocellulose was washed repeatedly with 0.001 N
HCl; then with water; solvent-exchanged to ethanol, 0.3
acetone and t-butanol; and freeze-dried. The reducing
ends remaining in the sample after reduction was 0.2
analyzed following the same BCA method mentioned BC1100 (rt)
BC500 (rt)
in the previous section. 0.1 BC300 (rt)
BC300 (50 °C)
Thermogravimetric analysis (TGA) 0
0 1 2 3 4 5

TGA was carried out using a TGA Q50 (TA Instru- Time (h)
ments, Tokyo, Japan). Dynamic and isothermal TGA
Fig. 3 Change in the degree of substitution (DS) as a function
were performed, each utilizing approximately 5 mg of of time during acetylation of BC1100, BC500 and BC300 at
freeze-dried nanocellulose placed on a platinum pan. room temperature (rt) and of BC300 at 50 °C

123
Cellulose (2016) 23:451–464 457

attempt to obtain a DS value higher than 0.38 to avoid crystallinity indexes increased to 95.7 and 95.5 for
possible collapse of the crystalline core of the BC500 and BC300, respectively. During acid hydrol-
cellulose associated with higher DS. As reported in ysis, the amorphous domains of cellulose are digested,
the work of Ifuku et al., (2007) the XRD profile of thereby increasing the crystallinity of BC500 and
acetylated BC nanofiber at high DS (0.95) showed a BC300. The different hydrolysis conditions for the
diffraction peak of cellulose triacetate and an amor- preparation of BC500 and BC300 did not seem to
phous halo, while at low DS (0.45) the crystallinity of affect the crystallinity of the hydrolyzed nanocellu-
the original BC nanofiber was maintained. Thus, a DS lose. Both had almost the same crystallinity index.
of 0.38 was a reasonable choice with the assumption This result suggests that hydrolysis of BC at similar
that at this value only the surface of the nanocellulose acid concentration and duration but at different
was acetylated and the crystalline core was intact. temperatures can digest the amorphous part of cellu-
BC300, however, did not reach a DS of 0.38 at room lose to the same extent but the cleavage of glycosidic
temperature. It was therefore decided to increase the bonds occurs at different rates thereby producing
reaction temperature to 50 °C. As shown in Fig. 3, the hydrolyzed bacterial nanocellulose with similar crys-
reaction rate of BC300 significantly increased and tallinity but different DPv values.
reached a DS of 0.38 after about 1 h. After acetylation, the XRD profiles (Fig. 5b) still
showed the same distinct peaks as the original
Morphology nanocellulose. The result indicates that the acetylation
conditions used to reach a DS of 0.38 did not affect the
The morphologies of bacterial nanocelluloses with crystalline core of the nanocellulose. The crystallinity
different DPv values before and after acetylation are indexes for BC1100, BC500 and BC300 after acety-
shown in Fig. 4. BC1100 (Fig. 4a) shows long inter- lation were 94.1, 96.1 and 95.7, respectively.
twined nanofibrils, while BC300 (Fig. 4c) shows short
rigid rods characteristic of CNCs. Interestingly, BC Reducing ends
500 (Fig. 4b) did not show nanocrystal characteristics
but retained the ribbon-like structure of the unhy- The RE is the terminal end of cellulose that contains
drolyzed BC1100. The hydrolysis conditions for the anomeric carbon bearing the hemiacetal hydroxyl
BC500 must have not been sufficient to cut the (OH), which can be converted to a carbonyl group in
nanofibrils into shorter crystalline rods. After acety- the form of an aldehyde and is thus susceptible to
lation, no change in morphology was observed except reduction and oxidation. Since analysis of REs
for BC300 which showed shorter crystalline rods than involves the reduction of a carbonyl group to form
the unacetylated ones. OH group, the number of REs was only determined in
the nanocellulose before acetylation and after NaBH4
Crystallinity reduction.
With breaking of the b1–4 glycosidic bonds during
Crystallinity is an important property that affects acid hydrolysis, the decrease in the DPv is accompanied
physical, mechanical, and chemical properties of by production of more REs. As shown in Fig. 6, the
cellulose. Tensile strength, dimensional stability, and number of REs increases as DPv decreases. The
density are known to increase with increasing cellu- unhydrolyzed BC1100 consisting of long intertwined
lose crystallinity, while chemical reactivity and nanofibrils had an average number of REs of 4.6 lmol/g.
swelling decrease (Agarwal et al. 2010). The crys- Using the equation RE = 106/[(162)DP ? 18] and
tallinity of cellulosic materials can vary depending on assuming an ideal condition that the nanocellulose is
the source and chemomechanical treatment used. monodispersed, the theoretical number of REs were
The X-ray diffraction profiles of nanocelluloses calculated by substituting different DP values. The
with different DPv values before and after acetylation calculated theoretical number of REs agreed well with
are presented in Fig. 5. BC has a typical cellulose I the experimental values, except for BC300. The dis-
structure with three well-defined diffraction peaks at crepancy observed for BC300 can be attributed to the
2h = 14.2°, 16.6°, and 22.4°. Initially, BC1100 had a longer hydrolysis time used in the preparation of BC300.
crystallinity index of 93.6. After acid hydrolysis, the As mentioned earlier, several hydrolysis conditions can

123
458 Cellulose (2016) 23:451–464

BC1100 BC500 BC300


a b c

d e f

Fig. 4 Scanning electron micrographs of bacterial nanocelluloses with varying average degree of polymerization before (a–c) and
after (d–f) acetylation

002 35
101

101 28
Reducing ends (µmol/g)

(a) 040 Experimental RE


21
Intensity

Theoretical RE

BC1100 14
BC500
BC300
(b) 7

5 10 15 20 25 30 35 40 0
1100 800 500 200
Diffraction angle (2θ)
DPv
Fig. 5 X-ray diffraction profiles of bacterial nanocelluloses
with varying average degree of polymerization (DPv) before Fig. 6 Average number of reducing ends (REs) of bacterial
(a) and after (b) acetylation nanocelluloses with varying average degree of polymerization
(DPv) with error bars represented by standard deviation from
three measurements and the theoretical number of REs
calculated for the different DP values
attain the LODP. As Fig. 1 showed, LODP can also be
reached after an hour of hydrolysis at 70 °C in 4 M HCl
but the 4 h hydrolysis time was selected to ensure the Daruwalla and Nabar (1956). It was hypothesized that
liberation of nanocrystalline rods. The continued hydrol- continued treatment with acid after the LODP was
ysis therefore increased the number of REs without reached caused the formation of smaller chain segments
changing the DPv. An increase in the number of REs that did not contribute to changes in viscosity but were
without a change in the DPv was also reported by capable of increasing the number of REs.

123
Cellulose (2016) 23:451–464 459

After NaBH4 reduction, the number of REs (BC1100) showed the highest thermal stability. After
decreased and leveled off to an average value of 1.8 acid hydrolysis, the decrease in the DPv was accom-
(±0.6) lmol/g for all the nanocelluloses. The number panied by a corresponding decrease in 10 %WLT and
of REs decreased by 73.8, 83.2 and 91.8 % for Tmax, resulting to a decreasing trend in thermal stability
BC1100, BC500, and BC300; respectively. In our with decreasing DPv.
attempt to completely reduce all the REs in each A similar trend in thermal stability was also observed
sample, the NaBH4-reduced nanocelluloses were by Gurgel et al. (2012) who studied the thermal
again subjected to a second NaBH4 treatment. The characteristics of depolymerized residues from extre-
number of REs, however, decreased only to an average mely low acid hydrolysis of sugarcane bagasse, and by
value of 1.3(±0.4) lmol/g and the percent reduction Calahorra et al. (1989) who reported the effect of
also did not significantly change in each of the molecular weight on the thermal decomposition of
nanocelluloses. The result suggests that the number cellulose obtained by fractionation from a solution of
of REs not accessible for reduction under the reaction eucalyptus pulp in cadoxen. Calahorra et al. attributed
conditions used in this study is relatively the same in the decrease in thermal stability to the decrease in the
each of the nanocelluloses. crystallinity index with decreasing DPv. In this study,
however, the crystallinity index increased after acid
Thermal stability hydrolysis. Moreover, BC500 and BC300 which had the
same crystallinity index showed different thermal
TGA was used to evaluate the thermal stability of stability. Thus, we cannot attribute the differences in
nanocellulose before and after acetylation. Figure 7 thermal stability to the change in crystallinity. In this
shows the thermogravimetric (TG) and derivative TG case, the decrease in thermal stability with decreasing
(DTG) curves of the different nanocelluloses before DPv was attributed to the increase in the number of REs.
acetylation. All of the nanocelluloses showed typical As mentioned in the previous section, the number of
single-step thermal degradation. Taking the 10 % REs increased with decreasing DPv. By plotting Tmax
weight loss temperature (10 %WLT), considered here and 10 % WLT against the number of REs (Fig. 8), it is
as the onset of degradation and the temperatures at the clear that there is an inverse relationship between
maximum weight loss rate (Tmax) as measures of thermal stability and the number of REs. Even though
thermal stability, the unhydrolyzed BC nanofibers the number of reducing ends is far less than the number
of surface OH groups in cellulose, the reactivity of the
REs can affect the properties of cellulose. The reducing
a 100
ends have been reported to be the starting point of
Weight (%)

75 various chemical and biochemical reactions, and ther-


moactivated reactions such as pyrolysis are not an
50
BC1100
exception. Matsuoka et al. (2011a, b, 2014) suggested
25 BC500 that the RE is the reactive site in cellulose during low-
BC300 temperature pyrolysis (\300 °C). From Fig. 8, with
0 increasing number of REs, the 10 %WLT decreased.
b BC300, which has the highest number of REs, showed
4.5
an early stage of degradation at around 250 °C (See
DTG (% / °C)

Fig. 7) and already lost 10 % of its initial weight at


3
around 300 °C. At this initial stage of decomposition,
1.5
the REs must play a catalytic role similar to the
mechanism proposed by Matsuoka et al. (2014).
0 Accordingly, during the onset of thermal decomposition
100 200 300 400 500 before significant weight loss occurs, the REs that are
Temperature (°C) initially present in the cellulose or produced by
depolymerization activate the crystalline cellulose dur-
Fig. 7 Thermogravimetric (a) and first derivative thermogravi-
metric (DTG) (b) curves of bacterial nanocelluloses with ing pyrolysis and the activation mechanism is very
varying average degree of polymerization effective at low temperature.

123
460 Cellulose (2016) 23:451–464

380 chloride and Tomé et al. (2011) who acetylated BC


nanofibers with acetic anhydride in a non-swelling
Temperature (°C) 360 ionic liquid. The differences in the thermal stability of
acetylated BC can be ascribed to different preparation
340 conditions. Pre-treatment processes used to purify the
BC or to remove the water in the BC, as well as the
320
reaction conditions during acetylation may have
caused improvement or deterioration in the thermal
Tmax property of acetylated BC.
300
10%WLT The increase in thermal stability after acetylation
can be attributed to the possible inhibition of inter- and
280
0 10 20 30 intramolecular hydrogen bond formation when the OH
Reducing ends (μmol/g) groups are substituted with acetyl groups. Matsuoka
et al. (2014) proposed that proton donation to the
Fig. 8 Inverse relationship between thermal stability repre- oxygen at carbon 1 of the glucose unit in cellulose
sented by temperatures at maximum weight loss rate (Tmax) and through hydrogen-bond rearrangement or hydrogen
10 % weight loss (10 %WLT) and the number of reducing ends
bonding with other molecules acts as an acid catalyst
that promotes the depolymerization reaction during
The effect of acetylation on the thermal stability of pyrolysis.
nanocelluloses with different DPv values is shown in It should also be noted that the Tmax value of each
Fig. 9. Acetylation did not change the single-step acetylated nanocellulose did not show much variation
thermal degradation of nanocellulose but only delayed despite the differences in the DPv. Tmax values for
the decomposition, which is reflected by the shift of acetylated nanocelluloses were in the range of
Tmax and 10 %WLT to higher temperatures after 365–371 °C, while a wider range of 330–356 °C was
acetylation. This indicates that acetylation improved recorded for non-acetylated nanocelluloses (Fig. 10).
the thermal stability of the nanocellulose, which The small range in Tmax values of acetylated nanocel-
agrees with the previous findings of Hu et al. (2011) luloses indicates that for acetylated nanocelluloses
who acetylated BC nanofibers under solvent free with similar DS the thermal decomposition tempera-
conditions and Ávila Ramı́rez et al. (2014) who also tures do not show significant variation despite the
acetylated BC nanofibers by organocatalytic route. differences in the DPv.
The results, however, contradict the report of Lee et al. Although the Tmax values did not greatly vary for
(2011) who acetylated BC nanofibers using acetic acid the acetylated nanocelluloses despite the differences
in the presence of pyridine and p-toluene sulfonyl in the DPv values, the degree of improvement in

Fig. 9 Thermogravimetric 100


and first derivative
Weight (%)

thermogravimetric (DTG)
curves of bacterial
nanocelluloses with 50
different average degree of
polymerization values
BC1100 BC500 BC300
before (dashed lines) and 0
after (solid lines) acetylation 4
DTG (%/°C)

1.5

-1
100 200 300 400 500 100 200 300 400 500 100 200 300 400 500
Temperature (°C) Temperature (°C) Temperature (°C)

123
Cellulose (2016) 23:451–464 461

thermal stability caused by acetylation varied in each reduction. In this manner only the hemiacetal OH was
of the nanocelluloses. We define here the degree of modified leaving the surface OH groups unchanged.
improvement in thermal stability as the difference in NaBH4 reduction has been reported to inhibit ther-
Tmax and 10 %WLT before and after acetylation. To mally-induced discoloration of cellulose (Avicel PH-
clearly show the degree of improvement in relation to 101) at relatively low temperatures, \300 °C (Mat-
the DPv, comparison of Tmax and 10 %WLT before suoka et al. 2011a) and also prolonged the induction
and after acetylation are shown in Fig. 10. As the DPv period of cellulose during pyrolysis at slow heating
decreased, the difference in the values of Tmax and conditions (Matsuoka et al. 2014). The induction
10 %WLT before and after acetylation increased, and period is the initial stage during TGA where the
it is reflected by the increase in the height of the white cellulose is said to be converted to an active interme-
shaded region of each bar in Fig. 10. The results diate before significant weight loss takes place (Brad-
indicate that introducing acetyl groups at the same DS bury et al. 1979). Thus, to observe the effect of NaBH4
to nanocelluloses with different DPv values does not reduction, we used isothermal heating at the 1 %
result in the same degree of improvement in thermal weight loss temperature (1 %WLT) of BC300, which
stability. The degree of improvement was found to is 250 °C. The 1 %WLT was determined from the TG
increase with decreasing DPv. This dependence on the curves obtained under dynamic heating shown in
DPv is possibly because of protection of the REs by Fig. 7. Within an hour of isothermal heating at 250 °C,
more stable acetyl groups. differences in the thermal behavior of the untreated
To verify the effect of protecting the reducing ends nanocelluloses with different DPv values were
to the degree of improvement in thermal stability, the observed and the effect of NaBH4 reduction was
REs were converted to its glucitol moiety by NaBH4 already evident as shown in Fig. 11. Before reduction,
the percent weight loss increased with decreasing DPv
after acetylation which again supported the effect of REs to thermal
380 before acetylation
stability, i.e. decreasing thermal stability with increas-
ing number of REs. After NaBH4 reduction, weight
355
loss was inhibited, but only slightly for BC1100 as
Tmax (°C)

shown in the small difference in the percent weight loss


330
before and after the treatment. The effect of NaBH4
reduction was more evident for BC300 which has the
305
highest number of REs (Fig. 11d). Similar result was
also obtained by Matsuoka et al. (2014) who also
280
1100 500 300 compared the effect of NaBH4 reduction to the weight-
DPv loss behavior of Avicel PH-101 and Whatman No. 42.
380 Considering the difference between the percent
weight loss before and after NaBH4 reduction as a
355 measure of the degree of improvement in thermal
10% WLT (°C)

stability because of NaBH4 reduction, the degree of


330 improvement also followed the same trend as that of
the effect of acetylation. The degree of improvement
305 in thermal stability because of NaBH4 reduction also
increased with decreasing DPv. This result confirms
280 that the improvement in thermal stability is affected by
1100 500 300
DPv
protecting the REs. The thermal stabilization because
of acetylation therefore can be ascribed to possible
Fig. 10 The temperature at maximum weight loss rate (Tmax) protection of the surface OH and of the REs. Although
and 10 % weight loss (10 %WLT) before and after acetylation it has been widely reported that acetylation begins on
of bacterial nanocelluloses with different average degree of
the accessible surface hydroxyl groups of cellulose,
polymerization (DPv) values and the degree of improvement in
thermal stability represented by the height of the white shaded which are in the C2, C3, and C6 positions, it can be
portion of each bar deduced that the two OH groups at each terminal end

123
462 Cellulose (2016) 23:451–464

a 100 thermoplastics with high melting points, this study

Weight (%)
investigated the thermal stabilization of nanocellu-
96
loses with different DPv values by acetylation.
92 BC 1100 In summary, acid hydrolysis of BC nanofibers
88 (DPv = 1100) at different temperatures but at the
same acid concentration and duration yielded nanocel-
b 100 lulose with different DPv values (500 and 300) and
Weight (%)

96 morphologies but similar crystallinity indexes. Of the


92
different characteristics of each of the nanocelluloses
BC500
prepared, the DPv was found to significantly affect the
88
thermal properties. Bacterial nanocelluloses with
c 100 varying DPv values showed decreasing thermal sta-
Weight (%)

96
bility with decreasing DPv. The decrease in thermal
stability as DPv decreased is attributed to the increase
92
BC300
in the number of REs. Improved thermal stability was
88 achieved by acetylating the nanocelluloses and the
0 20 40 60
degree of improvement increased as the DPv
Time (min)
decreased. The highest degree of improvement in
d acid-hydrolyzed CNC was with a DPv of 300, which
Weight loss (%)

12
recovered the thermal stability of unhydrolyzed BC
before reduction
8 nanofibers. The higher stabilization caused by acety-
after reduction
4 lation to CNCs with low DPv was attributed to possible
0
protection of the hemiacetal OH groups at the REs by
BC1100 BC500 BC300 more stable acetyl groups. The effect of protecting the
REs to the degree of improvement in thermal stability
Fig. 11 The weight loss-curve (a–c) and the weight loss values
was also confirmed by NaBH4 reduction. The degree
in percentage (d) before (dashed lines) and after (solid lines)
NaBH4 reduction of bacterial nanocelluloses with varying of improvement in thermal stability of NaBH4-
average degree of polymerization during isothermal heating at reduced nanocelluloses also increased with decreasing
250 °C for 1 h DPv. The recent findings suggest the strong effect of
REs to the thermal stability of nanocellulose. For
of the cellulose chains can also participate in the CNCs characterized by small DP, protection of the
reaction. This is analogous to heterogeneous acetyla- REs, apart from surface OH modification, can greatly
tion of cellobiose in a separate model experiment, contribute to its thermal stabilization. For CNFs with
where acetylation took place at the six OH groups high DP and low number of REs, protection of the
around the two glucopyranose rings and also at the two surface OH improves thermal stability while processes
terminal OH groups to produce cellobiose octaacetate. that generate more REs such as acid hydrolysis causes
The model experiment indicated that, if the terminal thermal destabilization.
OH groups of cellulose are accessible, they can also be
acetylated. Therefore, during acetylation, protection Acknowledgments The scholarship grant from the Ministry
of Education, Culture, Sports, Science and Technology, Japan
of the surface OH and the REs can take place.
(Monbukagakusho) to Melissa B. Agustin is gratefully
Furthermore, for nanocellulose with low DPv values, acknowledged. Ms. Yoko Homma is also thanked for her
acetylation of the REs can greatly contribute to valuable help in the calibration and maintenance of the TG
improvement in the thermal stability. Analyzer.

Conclusions References

Abdul Khalil HPS, Bhat AH, Ireana Yusra AF (2012) Green


With the goal of expanding the application of composites from sustainable cellulose nanofibrils: a
nanocellulose to high processing temperatures of review. Carbohydr Polym 87:963–979

123
Cellulose (2016) 23:451–464 463

Agarwal UP, Reiner RS, Ralph SA (2010) Cellulose I crys- Habibi Y (2014) Key advances in the chemical modification of
tallinity determination using FT–Raman spectroscopy: nanocelluloses. Chem Soc Rev 43:1519–1542
univariate and multivariate methods. Cellulose 17:721–733 Hu W, Chen S, Xu Q, Wang H (2011) Solvent-free acetylation
Ashori A, Babaee M, Jonoobi M, Hamzeh Y (2014) Solvent-free of bacterial cellulose under moderate conditions. Carbo-
acetylation of cellulose nanofibers for improving compat- hydr Polym 83:1575–1581
ibility and dispersion. Carbohydr Polym 102:369–375 Ifuku S, Nogi M, Abe K et al (2007) Surface modification of
Ávila Ramı́rez JA, Suriano CJ, Cerrutti P, Foresti ML (2014) bacterial cellulose nanofibers for property enhancement of
Surface esterification of cellulose nanofibers by a simple optically transparent composites: dependence on acetyl-
organocatalytic methodology. Carbohydr Polym group DS. Biomacromolecules 8:1973–1978
114:416–423 Jacquet N, Quiévy N, Vanderghem C et al (2011) Influence of
Azizi Samir MAS, Alloin F, Dufresne A (2005) Review of steam explosion on the thermal stability of cellulose fibres.
recent research into cellulosic whiskers, their properties Polym Degrad Stab 96:1582–1588
and their application in nanocomposite field. Biomacro- Johnston DB, Shoemaker SP, Smith GM, Whitaker JR (1998)
molecules 6:612–626 Kinetic measurements of cellulase activity on insoluble
Božič M, Vivod V, Kavčič S et al (2015) New findings about the substrates using disodium 2,20 bicinchoninate. J Food
lipase acetylation of nanofibrillated cellulose using acetic Biochem 22:301–319
anhydride as acyl donor. Carbohydr Polym 125:340–351 Jonoobi M, Harun J, Shakeri A, Misra M (2009) Chemical
Bradbury AGW, Sakai Y, Shafizadeh F (1979) A kinetic model composition, crystallinity, and thermal degradation of
for pyrolysis of cellulose. J Appl Polym Sci 23:3271–3280 bleached and unbleached kenaf bast (Hibiscus cannabinus)
Brinchi L, Cotana F, Fortunati E, Kenny JM (2013) Production pulp and nanofibers. BioResources 4:626–639
of nanocrystalline cellulose from lignocellulosic biomass: Kabir MM, Wang H, Lau KT, Cardona F (2013) Effects of
technology and applications. Carbohydr Polym chemical treatments on hemp fibre structure. Appl Surf Sci
94:154–169 276:13–23
Calahorra ME, Cortazar M, Eguiazabal JI (1989) Thermo- Klemm D, Kramer F, Moritz S et al (2011) Nanocelluloses: a
gravimetric analysis of cellulose: effect of the molecular new family of nature-based materials. Angew Chem Int Ed
weight on thermal decomposition. J Appl Polym Sci Engl 50:5438–5466
37:3305–3314 Lee K-Y, Quero F, Blaker JJ et al (2011) Surface only modifi-
Camarero Espinosa S, Kuhnt T, Foster EJ, Weder C (2013) cation of bacterial cellulose nanofibres with organic acids.
Isolation of thermally stable cellulose nanocrystals by Cellulose 18:595–605
phosphoric acid hydrolysis. Biomacromolecules Li J, Zhang LP, Peng F et al (2009) Microwave-assisted solvent-
14:1223–1230 free acetylation of cellulose with acetic anhydride in the
Çetin NS, Tingaut P, Özmen N et al (2009) Acetylation of presence of iodine as a catalyst. Molecules 14:3551–3566
cellulose nanowhiskers with vinyl acetate under moderate Martı́nez-Sanz M, Lopez-Rubio A, Lagaron JM (2011) Opti-
conditions. Macromol Biosci 9:997–1003 mization of the nanofabrication by acid hydrolysis of
Cunha AG, Zhou Q, Larsson PT, Berglund LA (2014) bacterial cellulose nanowhiskers. Carbohydr Polym
Topochemical acetylation of cellulose nanopaper struc- 85:228–236
tures for biocomposites: mechanisms for reduced water Matsuoka S, Kawamoto H, Saka S (2011a) Reducing end-group
vapour sorption. Cellulose 21:2773–2787 of cellulose as a reactive site for thermal discoloration.
Daruwalla EH, Nabar GM (1956) Acid hydrolysis of cellulose. Polym Degrad Stab 96:1242–1247
J Polym Sci XX:205–208 Matsuoka S, Kawamoto H, Saka S (2011b) Thermal glycosy-
Eichhorn SJ, Dufresne A, Aranguren M et al (2009) Review: lation and degradation reactions occurring at the reducing
current international research into cellulose nanofibres and ends of cellulose during low-temperature pyrolysis. Car-
nanocomposites. J Mater Sci 45:1–33 bohydr Res 346:272–279
Fahma F, Takemura A, Saito Y (2014) Acetylation and stepwise Matsuoka S, Kawamoto H, Saka S (2014) What is active cel-
solvent-exchange to modify hydrophilic cellulose whiskers lulose in pyrolysis? An approach based on reactivity of
to polychloroprene-compatible nanofiller. Cellulose cellulose reducing end. J Anal Appl Pyrolysis 106:138–146
21:2519–2527 Nakatsubo F, Yoshida N, Abe K, Yano H (2010) Chemical
Favier V, Canova GR, Cavaille JY et al (1995) Nanocomposite surface-modification of cellulose nanofibers in cellulose-
materials from latex and cellulose whiskers. Polym Adv compatible solvents. In: 239th ACS national meeting
Technol 6:351–355 technical program archive
Garcia E, Johnston D, Whitaker JR, Shoemaker SP (1993) Nishino T, Kotera M, Suetsugu M et al (2011) Acetylation of
Assessment of endo- 1,4-beta-d-glucanase activity by a plant cellulose fiber in supercritical carbon dioxide. Poly-
rapid colorimetric assay using disodium 2,20 -bicinchoni- mer (Guildf) 52:830–836
nate. J Food Biochem 17:135–145 Nogi M, Abe K, Handa K et al (2006) Property enhancement of
Gurgel LVA, Marabezi K, Ramos LA, Curvelo AADS (2012) optically transparent bionanofiber composites by acetyla-
Characterization of depolymerized residues from extre- tion. Appl Phys Lett 89:233123
mely low acid hydrolysis (ELA) of sugarcane bagasse Okahisa Y, Abe K, Nogi M et al (2011) Effects of delignification
cellulose: effects of degree of polymerization, crystallinity in the production of plant-based cellulose nanofibers for
and crystallite size on thermal decomposition. Ind Crops optically transparent nanocomposites. Compos Sci Tech-
Prod 36:560–571 nol 71:1342–1347

123
464 Cellulose (2016) 23:451–464

Peng Y, Gardner DJ, Han Y et al (2013) Influence of drying fibers using ionic liquids as solvent media and catalysts.
method on the material properties of nanocellulose I: Green Chem 13:2464
thermostability and crystallinity. Cellulose 20:2379–2392 Uetani K, Watanabe Y, Abe K, Yano H (2014) Influence of
Quiévy N, Jacquet N, Sclavons M et al (2010) Influence of drying method and precipitated salts on pyrolysis for
homogenization and drying on the thermal stability of nanocelluloses. Cellulose 21:1631–1639
microfibrillated cellulose. Polym Degrad Stab 95:306–314 Yano H, Sugiyama J, Nakagaito AN et al (2005) Optically
Segal L, Creely JJ, Martin AE, Conrad CM (1959) An empirical transparent composites reinforced with networks of bac-
method for estimating the degree of crystallinity of native terial nanofibers. Adv Mater 17:153–155
cellulose using the X-ray diffractometer. Text Res J Yu H, Qin Z, Liang B et al (2013) Facile extraction of thermally
29:786–794 stable cellulose nanocrystals with a high yield of 93 %
Siqueira G, Bras J, Dufresne A (2010) Cellulosic bio- through hydrochloric acid hydrolysis under hydrothermal
nanocomposites: a review of preparation, properties and conditions. J Mater Chem A 1:3938
applications. Polymers (Basel) 2:728–765
Tomé LC, Freire MG, Rebelo LPN et al (2011) Surface
hydrophobization of bacterial and vegetable cellulose

123

You might also like