You are on page 1of 8

Biomacromolecules 2010, 11, 1217–1224 1217

Nanostructural Reorganization of Bacterial Cellulose by


Ultrasonic Treatment
Paula C. S. Faria Tischer,*,† Maria Rita Sierakowski,† Harry Westfahl, Jr.,*,‡ and
Cesar Augusto Tischer†
Laboratório de Biopolı́meros, Universidade Federal do Paraná (UFPR), Caixa Postal 19081,
81531-990, Curitiba, Paraná, Brazil, and Laboratório Nacional de Luz Sı́ncrotron (LNLS), CEP 13083-970,
Caixa Postal. 6192, Campinas, SP, Brazil
Received December 4, 2009; Revised Manuscript Received March 16, 2010

In this work, bacterial cellulose was subjected to a high-power ultrasonic treatment for different time intervals.
The morphological analysis, scanning electron microscopy, and atomic force microscopy revealed that this treatment
changed the width and height of the microfibrillar ribbons and roughness of their surface, originating films with
new nanostructures. Differential thermal analysis showed a higher thermal stability for ultrasonicated samples
with a pyrolysis onset temperature of 208 °C for native bacterial cellulose and 250 and 268 °C for the modified
samples. The small-angle X-ray scattering experiments demonstrated that the treatment with ultrasound increased
the thickness of the ribbons, while wide-angle X-ray scattering experiments demonstrated that the average crystallite
dimension and the degree of crystallinity also increased. A model is proposed where the thicker ribbons and
crystallites result from the fusion of neighboring ribbons due to cavitation effects.

Introduction The high crystallinity of bacterial cellulose is the result of a


high number of inter- and intrahydrogen bonds between adjacent
Cellulose is one of the most important biopolymers and is chains of glucan. These bonds create a regular crystalline
widely used based on its availability, biocompatibility, biological arrangement between the glucan molecules,16 resulting in the
degradability, and sustainable production.1 In particular, bacterial distinct diffraction pattern, swelling, and reactivity of cellulose.
cellulose is a pure biopolymer, without collateral biogenic
One of the methods used to obtain microcrystals of cellulose
compounds like lignin, hemicelluloses, and pectin, and it is an
is to submit cellulose to controlled acid (H2SO4) hydrolysis
attractive biopolymer due to its facile production and purifica-
conditions. The hydronium ions can penetrate the cellulose
tion.2 The structure of bacterial cellulose (BC) was described chains in the amorphous domains, promoting the hydrolytic
more than a century ago,3,4 and the material has been extensively cleavage of the glycosidic bonds and releasing individual
analyzed and characterized.5-9 crystallites. In this process, sulfate groups are randomly
Bacterial cellulose has been produced from Acetobacter distributed on the cellulose microfibril surface, inducing a
xylinum by fermentation of different substrates as glucose from negative electrostatic layer covering the microfibers.20-22 The
corn syrup,10-12 and more recently, rice bark supplemented with cellulose microcrystals, because of their stiffness, thickness, and
glucose was used as a substrate to produce cellulose nano- length, are commonly called “whiskers”.17-19
spheres.13 For pursuing applications that require extensively entangled
Cellulose molecules form long chains in polycrystalline networks and higher strength, chemically less aggressive hy-
fibrous bundles, which contain crystalline and amorphous drolysis concepts have been found to maintain a high aspect
regions, and the degree of crystallinity and crystal dimensions ratio of the cellulose I fibrils or fibril aggregates, potentially
are dependent on the origin of the cellulose. The biosynthetic allowing strong or even permanent junction points for the
process is the same in all organisms, but there are some networks. A classic example is provided by omitting the
differences in the cellulose synthase complexes that determine hydrolysis step and by solely imposing high shearing forces
the size and thickness of the cellulose microfibrils, and the great for disintegration. This process yields a highly entangled
interest in cellulose macromolecules is due to their crystalline network, which typically consists of elements with a wide size
orientation.14 distribution down to the nanoscale. The resulting material has
The microstructures formed by the ultrafine microfibrils of been denoted microfibrillated cellulose (MFC), originally in-
bacterial cellulose have lengths varying from 1 to 9 µm and troduced by Turback et al.23 and Herrick et al.24
create a dense reticulated structure stabilized by various By this method, the cellulose forms a highly entangled
hydrogen bonds. These networks show a high index of crystal- network with a low degree of crystallinity, and some of the
linity (above 60%) and a higher degree of polymerization, noncrystalline domains remain intact. This method has the same
normally around 16.000 or 20.000,15 in comparison with vegetal problems, and the increased energy of tension necessary to cause
cellulose. the disintegration of fibers requires several steps of treatment.
Recently, a novel route was demonstrated that combines
* To whom correspondence should be addressed. Phone: +55 41 3361- enzymatic hydrolysis and mechanical shearing to form a long
3260 (P.C.S.F.); +55 19 3512-1034 (H.W.). E-mail: paula.tischer@pq.cnpq.br and highly entangled network of cellulose I elements.25
(P.C.S.F.); westfahl@lnls.br (H.W.).

Universidade Federal do Paraná (UFPR). Another method cited to prepare spherical nanoparticles of

Laboratório Nacional de Luz Sı́ncrotron (LNLS). cellulose is based on a hard pretreatment of cellulose: heating
10.1021/bm901383a  2010 American Chemical Society
Published on Web 04/06/2010
1218 Biomacromolecules, Vol. 11, No. 5, 2010 Tischer et al.

at 80 °C in 5.0 M NaOH for 3 h, followed by the resuspension radiation wavelength (λ ) 1.488 Å) at the Brazilian Synchrotron
of fibers in DMSO in a water bath at 80 °C for 3 h. Next, the Light Laboratory (LNLS) also in the transmission geometry. The
pretreated fibrils are suspended in an acid solution and sonicated samples were thin films, supported in the sample holder and kept
with heating at 80 °C for 8 h. By this process, nanospheres of in a 10-2 mbar vacuum. Scattering patterns were recorded during
60-570 nm composed of cellulose II are obtained.26 300 s exposures on a marCCD 165 detector (8 × 8 binning), placed
All these methods were realized with vegetal cellulose, and first at 2507 mm and then at 1484 mm away from the samples.
hard conditions for hydrolysis were used, which caused intense Calibration of the distances was achieved using a silver Behenate
pattern.44 The SAXS profiles from the samples consisted of isotropic
degradation of the polymer and a consequent reduction in
scattering patterns that were subtracted from the CCD bias and dark-
crystallinity.
noise according to a reference.45 The SAXS chamber parasitic
The effect of ultrasound in degrading polysaccharide linkages
scattering was also recorded (with bias and dark-noise subtraction)
is well described. Studies with chitosan,27-30 dextran,31-33 and subtracted from the sample pattern after sample attenuation
agarose and carrageenans,34 xyloglucan,35 cellulose derivatives, correction. The final patterns were normalized by the phase space
and carboxymethylcellulose36-38 have been described, and in volume and azimuthally averaged for a sequence of wave-vectors q
the majority of cases, the treatment with ultrasound has been (q ) (4π/λ) sin(θ); 2θ ) scattering angle) ranging from 0.04 to
used in acid solutions with relatively low temperature and in 1.14 nm-1 when the detector was at the longer distance and 0.1 to
alkaline39 or acidic conditions.40 2.3 nm-1 for the shorter one.
The energy of ultrasound is transferred to the polymer chains Wide-Angle X-ray Diffraction (WAXS). The WAXS patterns were
through a process called cavitation, which is the formation, used to determine the relative crystallinity. WAXS diagrams were
growth, and violent collapse of cavities in the water. The energy recorded on the D03B beamline using a fixed radiation wavelength (λ
provided by cavitation in this so-called sonochemistry is ) 1.433 Å) at the LNLS in the transmission geometry. The samples
approximately 10-100 kJ/mol,41 which is within the hydrogen were thin films supported in 1 cm aluminum rings aligned perpendicu-
bond energy scale.42 larly to the beam. Diffraction patterns were recorded during 120 s
The main goal of this work was to analyze the effect of exposures on a MarCCD 165 detector (2 × 2 binning), placed 100.0
ultrasonic treatment on the reorganization of bacterial cellulose mm away from the samples. Calibration of the distances was achieved
microfibrils using suitable conditions (aqueous medium, without using a LaB6 pattern. The WAXS profiles from the samples consisted
acid or heating). Such results are needed to develop an easy of circular Debye-Scherrer rings and were subtracted from the CCD
method to obtain films with different nanostructures and bias and dark-noise according to a reference.45 The air scattering was
characteristics (porosity, roughness, and crystallinity) and to also recorded (with bias and dark-noise subtraction) and subtracted from
the sample pattern after sample attenuation correction. The final patterns
develop a process in nanotechnology that permits the formation
were normalized by the phase space volume and azimuthally averaged
of new scaffolds for tissue engineering.
for a sequence of wave-vectors q (q ) (4π/λ) sin(θ); 2θ ) scattering
angle) ranging from 1.4 to 28.6 nm-1.
Scanning Electron Microscopy (SEM) Analysis. Scanning electron
Materials and Methods microscopy (SEM) was conducted to observe the surface of the bacterial
Microorganism, Culture Media, and Growth Conditions. The cellulose films before and after ultrasonic treatment. Samples were
bacterial strain Acetobacter xylinum ATCC 23769 (reclassified as the sputter coated with gold and examined using a JEOL JSM-6360LV in
genus Gluconacetobacter), supplied by Foundation André Tosello from the CEM (Centre of Electron Microscopy) at the UFPR (Universidade
Campinas, São Paulo, was grown in a glucose medium based on the Federal do Parana).
Hestrin-Schramm’s medium culture.43 All media were autoclaved at Atomic Force Microscopy (AFM) Analysis. The topography and
121 °C and 1.02 atm for 20 min. surface roughness of the cellulose films were determined using a
The inoculum was prepared in a nonagitated 200-mL Erlenmeyer Shimadzu FPM-9500J3 operating in air. Small pieces were cut from
flask containing about 5 mL of the thawed culture and 40 mL of the each membrane, glued onto metal disks and attached to a magnetic
above medium. The flask was incubated at 28 ( 1 °C for 48 h. The sample holder located on top of the scanner tube. The membrane surface
obtained cell suspension was used as the inoculum. The static was scanned in tapping mode. All of the AFM images were taken at
fermentation was carried out for 10 days at 28 °C, and the pH was 25 °C.
adjusted to 5.25 with citric acid (5%, w/v). Differential Thermal Analysis. This analysis was carried out to
Treatment of Bacterial Cellulose. The bacterial cellulose (BC) determine the thermal stability of our materials by employing a Mettler
pellicle that formed at the liquid-air interface of the fermentation Toledo TGA/SDTA851. Samples were heated from ambient temperature
medium was removed, rinsed thoroughly with deionized water, and to 1000 °C in an oxygen current of 50 mL/min at a heating rate of
purified by immersion in an aqueous solution of 0.1 M NaOH for one 10 °C/min.
day. The films were washed repeatedly with deionized water and then
vacuum-dried at 40 °C. This treatment was controlled to avoid the
mercerization of the BC. Results and Discussion
Ultrasonic Process. The pellicles of the BC after vacuum drying at
SEM Analysis. After ultrasonic treatment for different lengths
40 °C were cut with scissors. The cut films (200 mg, immersed in 200
of time in an aqueous medium, the cellulose was vacuum-dried
mL Mili-Q water) were shaken vigorously for 15 min. This material
at 40 °C, and the morphology of new films was initially observed
was submitted to ultrasonic treatment for 15, 30, 60, and 75 min using
an ultrasonic processor, Ultrasound SONICS (200 W, 20 kHz). The
by scanning electron microscopy (SEM). Figure 1a shows the
ultrasonic treatment was carried out in an ice bath, and the ice was 3-D reticulated structure of the cellulose fibrils with intercon-
maintained throughout the entire sonication time. After this process, nected pores of different sizes, a typical structure of cellulose
the ultrasonicated BC was reconstituted in the form of films and produced by A. xylinum in a glucose medium by static
vacuum-dried at 40 °C. In all of the analysis techniques, the stable fermentation.16 The bacteria Acetobacter xylinum synthesize
films formed after ultrasonication were used in the solid state as whole primary nanofibrils with lateral sizes in the range of 7-13 nm,
films or cut films. which aggregate into thin and flat bands or ribbons having
Small-Angle X-ray Diffraction (SAXS). The SAXS patterns were widths of 70-150 nm.46
used to determine the dimensions of the microfibrils in the cellulose. The SEM micrograph of the native bacterial cellulose showed
SAXS diagrams were recorded on the D02A beamline using a fixed bands of microfibrils (ribbons) with widths of 90-140 nm. After
Nanostructural Reorganization of Bacterial Cellulose Biomacromolecules, Vol. 11, No. 5, 2010 1219

Figure 1. Scanning electron micrographs of bacterial cellulose, native (a) and ultrasonicated for 15 (b), 30 (c), 60 (d), and 75 min (e).

the shorter ultrasonication time, we can observe that there is a Apparently, the ultrasonic treatment changed the microfibrillar
reduction in the width of the ribbons and a decrease in the arrangement, leading to a film with a different nanostructure.
number/density of pores (Figure. 1b). Apparently, a change in From the height obtained with AFM, we observed that after 15
the morphology causes the surface to become more compact min of ultrasonic processing, there is an increase in the ribbons’
with fewer pores. Increasing the length of the treatment reduces thickness (Table 1). Apparently, increasing the time of ultrasonic
the ribbons even further, and the surfaces become more uniform application to 60 min or more caused a reduction of thickness
(Figure 1c-e). The width of the ribbons cannot be determined of the ribbons.
by SEM analysis because of their reduced size, so it was The image presented in Figure 2c shows the surface of the
determined by AFM analysis and by SAXS instead. film that was treated for 60 min with ultrasound, showing
AFM Analysis. The surface morphology and surface char- changes in the morphology of the ribbons. The surface of the
acteristics, such as roughness, are important factors in interfacial BC film that was ultrasonicated for 30 min has similar features
interactions; thus, the surface changes in bacterial cellulose, after
(data not shown). After 75 min, individual ribbons with widths
ultrasonic treatment, were determined. The tapping-mode height
of 40-70 nm could be observed. The AFM analysis allowed
images of the ultrasonic cellulose surfaces are shown in Figure
us to observe that on the surface of the cellulose films that were
2. The width of the ribbons in the native BC was determined to
ultrasonicated for 30 and 60 min the shape of this nanomaterial
be between 110-140 nm, values near those observed by SEM.
In the BC that was ultrasonicated for 15 min, there was, was irregular.
apparently, a reduction in the width (70-110 nm) and an Determining the thickness of the ribbons from AFM is prone
increase in the thickness (39-42 nm) of the ribbons; there seems to imprecision due to the arrangement of the ribbons on a flat
to be a rather open fibrillar network on the surface of this film. surface. The determination of the variation in the thickness and
1220 Biomacromolecules, Vol. 11, No. 5, 2010 Tischer et al.

Figure 2. AFM tapping-mode height images in air of a bacterial cellulose thin film: native (a) and ultrasonicated for 15 (b), 60 (c), and 75 min
(d).

Table 1. Width, Height, and Roughness (rms) of the Native BC


and the Ultrasonicated BC at Different Times, Analyzed by Atomic
Force Microscopy (AFM)a
rms (nm) width (nm) height (nm)
BC 38.5 110.0 25.2
120.0 20.6
140.0 20.4
BC15 37.9 51.0 39.4
64.6 42.3
70.4 40.3
BC60 29.3 30.9 12.6
49.9 12.0
52.9 22.1
BC75 17.3 26.4 22.0
36.2 26.4
39.5 20.0
a
Area: 2 × 2 µm.

Figure 3. Differential thermal analysis of bacterial cellulose, native


width of the ribbons after ultrasonic treatment is performed by and ultrasonicated.
SAXS analyses in the next section.
The corresponding mean roughness values (rms) were dif-
ferent for all of the films reconstituted after the ultrasonication
Differential Thermal Analysis. Ciolacu and Popa47 showed
process, varying from 38.5 to 17.3 nm (Table 1).
that the lower the crystallinity of the polymer, the lower is its
At least two films of the same composition were analyzed at thermal stability. Differential thermal analysis of native bacterial
different areas of the surface. It is difficult to compare the cellulose and ultrasonicated samples is shown in Figure 3.
different roughness values reported in the literature because of
a lack of uniformity in the methods applied to determine surface The thermal decomposition pattern of cellulose can be
roughness; moreover, the size of the scan area is highly variable. explained mainly by two different mechanisms.48 The first one
After the ultrasonication process, we can see evidence of involves the degradation due to the breaking of internal bonds,
homogeneous profiles and low roughness values due to more dehydration or formation of free radicals, and reactive carbona-
homogeneous substrates, opening the possibility for a future ceous char. The second mechanism involves the cleavage of
study of biomolecular adsorption. secondary bonds and the formation of intermediate products,
Nanostructural Reorganization of Bacterial Cellulose Biomacromolecules, Vol. 11, No. 5, 2010 1221

Figure 5. SAXS plots from the three samples along with the
extrapolated areas (light dots).

Table 2. Dimensions of the Ribbons Obtained from SAXS


a (nm) b (nm)
BC 7.6 75
BC15 21 86
BC30 15 81

b of the ribbons byrc ) (a2 + b2)/(12). In this region, which in


our experiment belongs to the lowest wave vectors, the cross
sectional radius of gyration is determined53 by the linear
coefficient of the curve ln(qI(q)) × q2 as in Figure 4.
It is clear from Figure 4 that rc increases with ultrasonic
treatment, indicating that one or both cross sectional dimensions
increased with ultrasonic treatment. However, this does not
necessarily imply a thickening of the ribbon. In fact, because
the cross sectional radius of gyration alone cannot determine a
and b independently, we also need to evaluate the Porod cross
sectional area of the ribbons,S ) ab, given by the SAXS curves
as52,53 S ) 2π limqf0(qI(q))/Q. The scattering invariant Q can
be determined by the total area of the scattering curve,53
Figure 4. Guinier plots for native cellulose (a), cellulose treated for extrapolated into the Guinier and Porod regions (see Supporting
15 min (b), and cellulose treated for 30 min (c). Information for details of this calculation). From Figure 5, we
can see that the normalized scattering curve, I(q)/Q, is shifted
upward with ultrasonic treatment, indicating that the cross
such as anhydromonosaccharides, which are converted into low sectional area of the ribbons has also increased.
molecular weight polysaccharides and finally carbonized prod- Finally, combining the data for S and rc, one obtains the cross
ucts.49 sectional dimensions a and b of the ribbons, as presented in
The results obtained from differential thermal analysis showed Table 2. The dimensions of the native microfibril bands obtained
us the onset temperature of the pyrolysis was 208, 250, and here are similar to the ones reported previously;52 however, as
268 °C for the native BC and the BC ultrasonicated for 15 and suggested by the SAXS data analysis, the treatment with
30 min, respectively. In addition, after comparing the three ultrasound seems to increase their thickness, resulting in a
samples, it was observed that the weight losses of the two modification of the nanostructure of the BC. Note, however,
degradation events were similar, indicating that the ultrasonic that this observation does not necessarily imply an increase in
process did not change the cellulose chemically; apparently, only crystallinity because the ribbons are composed of crystalline
structural changes occurred. More data from differential thermal microfibrils and amorphous cellulose. The crystallite dimensions
analysis can be seen in the Supporting Information. and crystalline fractions of the samples are determined from
SAXS Analysis. Previous X-ray scattering studies on BC50,51 the WAXS data analysis in the next section.
have already evidenced an ultrastructure composed of nanofibrils WAXS Analysis. The great interest in the mechanical and
(in the nm range) aggregated into microfibrils (in the 10 nm medical properties of bacterial cellulose is due to its crystalline
range), attached together through interconnecting cellulose orientation. The physical properties, as well as the accessibility
chains, to produce ribbons (in the 100 nm range) that are the in the chemical reactions, swelling and adsorption, are strongly
basic units of the 3D network observed in bacterial cellulose. influenced by the crystallinity.54
The BC ribbons are the basic units that contribute most for Due to the linearity of the cellulose backbone, adjacent chains
the X-ray scattering at small angles. However, due to their of cellulose form a framework of water-insoluble aggregates
extensive length (L), the full radius of gyration of the ribbons of varying length and width. These microfibrils consist of both
is way beyond the detection limit of conventional SAXS. highly ordered (crystalline) and less ordered (amorphous)
Nevertheless, it is still possible to observe an intermediate regions. Depending on the source and the conditions used in
scattering region where a cross sectional radius of gyration,rc, cellulose treatment, different degrees of crystallinity can be
can be identified and related52 to the thickness a and the width obtained.55
1222 Biomacromolecules, Vol. 11, No. 5, 2010 Tischer et al.

Table 3. Crystallinity (xc) and Widths of the Crystallites in the


Directions Normal to the Corresponding Bragg Planes
xc D010 D001 D011 D-411
native 0.44 3.4 10.1 7.5 10.9
BC 15 0.55 6.4 9.6 6.5 19.4
BC 30 0.63 5.4 8.5 6.0 17.6

where K is a constant that depends on the shape of the


crystallites (K ≈ 0.94 for cubic and K ≈ 1.1 for a spherical
shape), and ∆qhkl is the fwhm of the hkl reflection.
Although subtle, the differences in the crystalline structure
of plant and bacterial cellulose have been a subject of great
debate.54 In bacterial cellulose, the major fraction is usually
ascribed to IR crystallinity. For the present data, a Rietveld or
linked atoms least squares (LALS) fit would have to be
performed to determine precisely the relative amounts of IR and
Iβ components.55,56 However, our main concern here is only
the relative amount of crystallinity between the native and the
ultrasonically treated samples. We, therefore, index the reflec-
tions according to IR crystallinity. In Table 3, we present the
average crystallite lengths along selected directions, according
to the Scherrer relation.
As we can see, due to the ultrasonic treatment, the average
crystallite dimension increases in the (010) direction but shrinks
in the other directions. These changes in crystalline dimensions
are compatible with the increase in the ribbon dimensions
obtained from the SAXS analysis.
The crystallinity of the native bacterial cellulose and the
samples ultrasonicated for 15 and 30 min was calculated from
the invariant integral of the crystalline and total scattering
contributions as

Figure 6. Lorenzian fits to the WAXS reflections in the Debye-Scherrer


∫ Icr(q)q2dq
mode for the native BC (a), BC treated for 15 min (b), and BC treated xc )
for 30 min (c).
∫ I(q)q2dq
The dimensions of the crystalline regions can also be where Icr(q) is the contribution to the scattering curve due to
obtained from the line widths of the crystalline reflections only the crystalline peaks and I(q) is the full (amorphous +
using the Scherrer method. Although not accurate, as a result crystalline) WAXS curve.
of neglecting the contributions due to paracrystallinity The crystallinity of these samples, native and ultrasonically
defects, this method can be useful for comparing the treated for 15 and 30 min, is 0.44, 0.55, and 0.63, respectively.
crystallite sizes of the three samples. Thus, we fit Lorenzians Although this method is not as reliable for obtaining the
to the main crystalline reflections with different positions crystallinity57 as the Rietveld method, the relative increase in
and widths, along with a smooth amorphous background, as xc is clear and appears to be due to the conversion of the
presented in Figure 6. amorphous halo background directly into the (010) facets of
The positions of the crystalline peaks are fixed for the three the new crystallites.
samples, using the indexing of the main IR reflections according The transformation of amorphous cellulose into cellulose I
to Nishiyama et al.56 Note also that the Miller indices of the is expected because polymerization and crystallization are
reflections are from a unit cell with the chains aligned along coupled in Acetobacter xylinum. It has been shown by Benziman
the a-axis. The widths and heights of the Lorenzian peaks are et al.58 that by preventing crystallization using Calcofluor, the
fitted simultaneously to the smooth amorphous background, amorphous cellulose is formed as bundles of parallel glucan
which we represent by four broad Lorenzians spread along the chains that crystallize into cellulose IR after drying. This suggests
entire range of measured wave-vectors and centered at ap- that, in normal biogenic conditions, the amorphous parts from
proximately q ) 0, 7, 12, and 26 nm-1. bacterial cellulose are cell-directed into parallel bundles ready
The fwhm (full width at half maximum) of the Lorenzian to crystallize into cellulose I. The activation energy required
peaks are determined by the widths of the crystallites in the for such a process is provided by cavitation from the ultrasonic
direction normal to the Bragg planes given by the Scherrer treatment. This process occurs primarily in the amorphous
relation as follows: regions, where water can more effectively penetrate.
It is clear that part of the amorphous background correspond-
ing to the broad features around q ) 7 and 12 nm-1 (the blue
2πK curve in Figure 6), which may come from the intermicrofibrillar
Dhkl ) cellulose chains, is being converted into crystallites. However,
∆qhkl
if the conversion of amorphous material into crystalline material
Nanostructural Reorganization of Bacterial Cellulose Biomacromolecules, Vol. 11, No. 5, 2010 1223

thank Mateus B. Cardoso for insightful discussions and help


with Figure 7.

Supporting Information Available. Details of the SAXS


data analysis and additional data obtained by differential thermal
analysis. This material is available free of charge via the Internet
at http://pubs.acs.org.

References and Notes


(1) Saxena, I. M.; Kudlicka, K.; Okuda, K.; Brown, R. M., Jr. J. Bacteriol.
1994, 176, 5735–5752.
(2) Hornung, M.; Ludwig, M.; Schmauder, H. P.; Gerrard, A. M. Eng.
Life Sci. 2006, 6, 546–551.
(3) Gardner, D. J.; Oporto, G. S.; Mills, R.; Samir.; Azizi, M. A. S. J.
Adhes. Sci. Technol. 2008, 22, 545–567.
(4) Brown, A. J. J. Chem. Soc. 1886, 49, 172–187.
(5) Brown, A. J. J. Chem. Soc. 1886, 49, 432–439.
(6) Iguchi, M.; Yamanaka, S.; Budhiono, A. J. Mater. Sci. 2000, 35, 261–
270.
(7) Brown, R. M., Jr.; Willison, J. H. M.; Richardson, C. L. Proc. Natl.
Acad. Sci. U.S.A. 1976, 73, 4565–4569.
(8) Yamanaka, S.; Watanabe, K.; Kitamura, N. J. Mater. Sci. 1989, 24,
3141–3145.
(9) Cousins, S. K.; Brown, R. M., Jr. Polymer 1997, 38, 897–902.
(10) Son, H. J.; Heo, M. S.; Kim, Y. G.; Lee, S. J. Biotechnol. Appl.
Biochem. 2001, 33, 1–5.
(11) Othmer, K. Encyclopedia of Chemical Technology: Cellulose; John
Wiley & Sons, New York, 1993.
(12) Okiyami, A. M.; Motoki, M.; Yamanaka, S. Food Hydrocolloids 1993,
Figure 7. Scheme illustrating the changes after ultrasonic treatment 6, 503–511.
and the creation of new crystallites. (13) Goelzer, F. D. E.; Faria-Tischer, P. C. S.; Vitorino, J. C.; Sierakowski,
M.-R.; Tischer, C. A. Mater. Sci. Eng., C 2009, 29, 546–551.
(14) Sarko, A.; Muggli, R. Macromolecules 1974, 7, 486–494.
occurred in each ribbon independently, we would expect a (15) Watanabe, K.; Tabuchi, M.; Ishikawa, A.; Takemura, H.; Tsuchida,
decrease in their thickness rather than an increase because the T.; Morinaga, Y.; Yoshinaga, F. Biosci., Biotechnol., Biochem. 1998,
62, 1290–1292.
density of crystalline cellulose is higher than that of amorphous
(16) Gardner, K. H.; Blackwell, J. Biopolymer 1974, 13, 1975–2001.
cellulose. (17) Lima, M. M. D.; Borsali, R. Macromol. Rapid Commun. 2004, 25,
Therefore, we propose that the thicker crystallites result from 771–787.
the fusion of neighboring ribbons, as shown in Figure 7. (18) Araki, J.; Wada, M.; Kuga, S.; Okano, T. Colloids Surf., A 1998, 142,
Indeed, it is expected that the surface of the ribbons will be 75–82.
(19) Battista, O. A. Ind. Eng. Chem. Res. 1950, 42, 502–507.
much more susceptible to the effects of cavitation than the (20) Marchessault, R. H.; Morehead, F. F.; Walter, N. M. Nature 1959,
crystalline core. This scenario supports the conclusions obtained 184, 632–633.
from SEM, AFM, SAXS, and WAXS. (21) Marchessault, R. H.; Morehead, F. F.; Joan, K. M. J. Colloid Sci.
1961, 16, 327–344.
(22) Dong, X. M.; Kimura, T.; Revol, J. F.; Gray, D. G. Langmuir 1996,
12, 2076–2082.
Conclusions (23) Turbak, A. F.; Snyder, F. W.; Sandberg, K. R. J. Appl. Polym. Sci.
1983, 37, 815–827.
In the present work, we have demonstrated that ultrasonic
(24) Herrick, F. W.; Casebier, R. L.; Hamilton, J. K.; Sandberg, K. R.
processing in mild conditions was effective in changing the J. Appl. Polym. Sci. 1983, 37, 797–813.
microfibrillar structure of bacterial cellulose. The ultrasound (25) Paakko, M.; Ankerfors, M.; Kosonen, H.; Nykanen, A.; Ahola, S.;
energy is transferred through shearing and cavitation to the Osterberg, M.; Ruokolainen, J.; Laine, J.; Larsson, P. T.; Ikkala, O.;
glucan chains, promoting the conversion of amorphous material Lindostrom, T. Biomacromolecules 2007, 8, 1934–1941.
(26) Zhang, J.; Elder, T. J.; Pu, Y.; Ragauskas, A. J. Carbohydr. Polym.
into crystalline material. The energy scale of the cavitation 2007, 69, 607–611.
processes is within the hydrogen bond energy scale and occurs (27) Liu, H.; Bao, J.; Du, Y.; Zhou, X.; Kennedy, J. F. Carbohydr. Polym.
primarily in the amorphous regions because these are the regions 2006, 64, 553–559.
where water can more effectively penetrate. Moreover, due to (28) Liu, H.; Du, Y.-M.; Kennedy, J. F. Carbohydr. Polym. 2007, 68, 598–
600.
the intrinsic nature of polymerization and crystallization pro-
(29) Wu, T.; Zivanovic, S.; Hayes, D. G.; Weiss, J. J. Agric. Food Chem.
moted by Acetobacter xylinum, the crystalline material formed 2008, 56, 5112–5119.
is of type I and not of type II, as would be expected by random (30) Kasaai, M. R.; Arul, J.; Charlet, G. Ultrason. Sonochem. 2008, 15,
crystallization of the amorphous phase. A scheme illustrating 1001–1008.
this effect is depicted in Figure 7. The shape of these crystallites (31) Cote, G. L.; Willet, J. L. Carbohydr. Polym. 1999, 39, 119–126.
(32) Lorimer, J. P.; Mason, T. J.; Cuthbert, T. C.; Brookfield, E. A.
changed, and this process created films with higher crystallinity Ultrason. Sonochem. 1995, 2, 55–57.
and lower surface roughness. (33) Portenlanger, G.; Heusinger, H. Ultrason. Sonochem. 1997, 4, 127–
130.
Acknowledgment. We thank CNPq for a postdoctoral (34) Lii, C.-Y.; Chen, C.-H.; Yeh, A.-I.; Lai, V. M. F. Food Hydrocolloids
fellowship (P.C.S.F.T.), CAPES/Brazil for financial support, 1999, 13, 477–481.
FINEP for financial support for the use of electronic microscopy (35) Vodenicarova, M.; Driı́malová, G.; Hromádková, Z.; Malovı́ková, A.;
Ebringerová, A. Ultrason. Sonochem. 2006, 13, 157–164.
equipment (CEM), Prof. Paulo Cesar Camargo for the use of (36) Imai, M.; Ikari, K.; Suzuki, I. Biochem. Eng. J. 2004, 17, 79–83.
AFM microscopy, and LAMIR (Laboratório de Análises de (37) Gronroos, A.; Pirkonen, P.; Ruppert, O. Ultrason. Sonochem. 2004,
Minerais e Rochas) for Differential Thermal Analysis. We also 11, 9–12.
1224 Biomacromolecules, Vol. 11, No. 5, 2010 Tischer et al.

(38) Aliyu, M.; Hepher, M. J. Ultrason. Sonochem. 2000, 7, 265–268. (49) Kawamoto, H.; Murayama, M.; Saka, S. J. Wood Sci. 2003, 49, 469–
(39) Mislovičová, D.; Masarova, J.; Bebdzalova, K.; Soltes, L.; Machova, 473.
E. Ultrason. Sonochem. 2000, 7, 63–68. (50) Klemm, D.; Heublein, B.; Fink, H.-P.; Bohn, A. Angew. Chem., Int.
(40) Chen, R. H.; Chang, J. R.; Shyur, J. S. Carbohydr. Res. 1997, 299, Ed. 2005, 44, 3358–3393.
287–294. (51) Nishiyama, Y. J. Wood Sci. 2009, 55, 241–249.
(41) Suslick, K. S. Science 1990, 247, 1439–1445. (52) Astley, O. M.; Chanliaud, E.; Donald, A. M.; Gidley, M. J. Int. J. Biol.
(42) Li, S. Y.; Liu, X. F.; Zhuang, X. P.; Guan, Y. L.; Cheng, G. X. Chin. Macromol. 2001, 29, 193–202.
J. Appl. Chem. 2003, 20, 1030–1035. (53) Glatter, O. In Small Angle X-ray Scattering; Glatter, O., Kratky, Eds.;
(43) Hestrin, S.; Schramm, M. Biochem. J. 1954, 58, 345–352. Academic Press: London, 1982; p 167.
(44) Blanton, T. N.; Barnes, C. L.; Lelental, M. J. Appl. Crystallogr. 2000, (54) Klemm, D.; Schumann, D.; Udhardt, U.; Marsch, S. Prog. Polym.
33, 172–173. Sci. 2001, 26, 1561–1603.
(45) Narayanan, T.; Diat, O.; Bosecke, P. Nucl. Instrum. Methods Phys. (55) Iwata, T.; Indrarti, L.; Azuma, J. Cellulose 1998, 5, 215–228.
Res.,Sect. A 2001, 467, 1005–1009. (56) Nishiyama, Y.; Sugiyama, J.; Chanzy, H.; Langan, P. J. Am. Chem.
(46) Klemm, D.; Schumann, D.; Kramer, F.; Hessler, N.; Hornung, M.; Soc. 2003, 125, 14300–14306.
Schmauder, H.-P.; Marsch, S. AdV. Polym. Sci. 2006, 205, 49–96. (57) Thygesen, A.; Oddershede, J.; Lilholt, H.; Belinda, A. T.; Stahl, K.
(47) Ciolacu, D; Popa, V. On the thermal degradation of cellulose Cellulose 2005, 12, 563–576.
allomorphs. Cellul. Chem. Technol. 2006, 40, 445–449. (58) Benziman, M.; Haigler, C. H.; Brown, M. R.; White, R. A.; Cooper,
(48) Rowell, R. M., Le Van-Green, S. L. In Handbook of Wood Chemistry M. K. Proc. Natl. Acad. Sci. U.S.A. 2007, 77, 6678–6682.
and Wood Composites; Rowell, R. M., Ed.; CRC Press: Gainesville,
FL, 2005; p 121. BM901383A

You might also like