Green Chemistry Volume Issue 2013

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Green Chemistry

View Article Online


PAPER View Journal
Published on 06 August 2013. Downloaded by University of Missouri at Columbia on 22/08/2013 17:39:04.

Mechanical depolymerisation of acidulated cellulose:


Cite this: DOI: 10.1039/c3gc40945g understanding the solubility of high molecular
weight oligomers†
Abhijit Shrotri,a Lynette Kay Lambert,b Akshat Tanksalec and Jorge Beltramini*a

Water soluble oligomers of cellulose were produced by milling acidulated microcrystalline cellulose. Acids
with low pKa were found to be more effective for the treatment. The yield of water soluble fraction was
proportional to the increasing severity of milling or the acid amount. Soluble oligomers were found to
have an average degree of polymerisation of ∼7 monomer units. High resolution NMR spectroscopy was
used to determine the chemical structures of soluble oligomers. It was found that branched oligomers
with α (1 → 6) linkages were formed, which increased their solubility in water and reduced the gene-
Received 21st May 2013,
ration of monomers and dimers which may degrade during milling or subsequent hydrolysis–hydro-
Accepted 5th August 2013
genation. The soluble oligomers showed excellent reactivity towards hydrolysis–hydrogenation in the
DOI: 10.1039/c3gc40945g
presence of bi-metallic Ni–Pt/alumina catalyst. High yields (∼90%) of sorbitol and mannitol were
www.rsc.org/greenchem obtained with only 1 h of reaction time.

1. Introduction in a slow and sequential cleavage of the cellulose chain. It is


also important to note that all research in this field has been
Among all forms of biomass available, lignocellulosic biomass performed exclusively on batch reactor systems.
is considered the most suitable for biofuel and green chemical To facilitate large-scale operation it is imperative to develop
production. It is primarily composed of polymers cellulose, a continuous process for the production of fuels and chemi-
hemicellulose, and lignin.1 Cellulose is one of the most abun- cals from cellulose. A possible solution to overcome these
dant sources of organic carbon. In its pure state it exists as a issues is to solubilise cellulose in water prior to the reaction.
large polymer composed of repeating units of anhydro glucose Dissolution of cellulose is known to increase the rate of hydro-
monomer units linked via β (1 → 4) glycoside bonds. Aqueous lysis. Handling a solution would be much easier in an indus-
phase hydrolysis of cellulose selectively produces glucose trial scenario and it would also allow the reaction to be
which can then be reacted to form key platform chemicals.2 performed in a continuous fixed bed reactor. However, cellu-
Despite worldwide focus, commercial application of this lose is known to be insoluble in most conventional solvents.
technology is still not viable. One of the primary roadblocks is Cellulose has an ordered crystalline structure, which is
the slow rate of the hydrolysis reaction, which leads to low the result of an intermolecular network of hydrogen bonds
yields and long reaction times. Cellulose hydrolysis proceeds between hydroxyl groups of monomer units. This structural
via a heterogeneous mechanism where the reaction rates are rigidity makes cellulose insoluble in water. Nonetheless, in-
influenced by the physical state of cellulose. Insolubility of cel- solubility of native cellulose in water is not yet fully under-
lulose limits the access to β (1 → 4) glycoside bonds, resulting stood.3,4 It is commonly accepted that dissolution of cellulose
in aqueous solutions is achieved by disruption of the hydrogen
bond network through decrystallisation and depolymerisation
a
ARC Centre of Excellence for Functional Nanomaterials, University of Queensland, of cellulose.5 Recently, aqueous solvent systems such as alkali-
Brisbane, QLD 4072, Australia. E-mail: j.beltramini@uq.edu.au;
urea6 and supercritical water7 were proposed for the dissol-
Fax: +61 7 3346 3973; Tel: +61 7 3346 3803
b
Centre for Advanced Imaging, University of Queensland, Brisbane, QLD 4072,
ution of cellulose. However, a large-scale application of these
Australia. E-mail: lynette.lambert@cai.uq.edu.au; Fax: +61 7 3365 3833; processes is questionable due to their high-energy requirement
Tel: +61 7 3365 9737 and the use of large amounts of concentrated alkali solutions.
c
Department of Chemical Engineering, Monash University, Clayton, VIC 3800, Modification of cellulose structure via pre-treatment is
Australia. E-mail: akshat.tanksale@monash.edu; Fax: +613 9905 5686;
widely practised to improve its solubility and reactivity with
Tel: +61 3 9902 4388
† Electronic supplementary information (ESI) available. See DOI: 10.1039/ heterogeneous catalysis. All pre-treatment processes primarily
c3gc40945g aim to decrystallise cellulose to produce amorphous cellulose.

This journal is © The Royal Society of Chemistry 2013 Green Chem.


View Article Online

Paper Green Chemistry

This can be achieved easily by milling crystalline cellulose in a catalyst supported on alumina and beta-zeolite can also be
ball mill.8 Along with reducing the crystalline structure, used for this reaction.24 The catalyst showed a comparable
milling also reduces particle size and increases the effective yield and selectivity towards hexitols during hydrolysis of
surface area. Milling efficiency is a function of various para- microcrystalline cellulose. However, depolymerisation of cellu-
meters including milling time, ball size and the sample to ball lose into glucose monomer units is the rate limited step due to
ratio. Irrespective of the conditions used, amorphous cellulose low solid–solid interaction.23 Therefore, it is expected that
thus produced exhibits similar properties. Amorphous cellu- using soluble oligomers will enhance the reaction rate, result-
Published on 06 August 2013. Downloaded by University of Missouri at Columbia on 22/08/2013 17:39:04.

lose has fewer hydrogen bonds per repeating monomer unit ing in a high yield of hexitols.
than crystalline cellulose.9 β (1 → 4) Glycoside bonds in amor-
phous cellulose are more easily accessed by water and other
reactants for hydrolysis.10–12 Amorphous cellulose is also 2. Experimental
known to undergo hydrolysis at a lower temperature.13 Despite
2.1 Cellulose treatment
various advantages amorphous cellulose is insoluble in water
and reaction rates are governed by heterogeneous pathways. The required amount of acid was diluted in water to a total
Very recently catalytic depolymerisation of cellulose to oligo- volume of 25 ml. 10 g of Sigmacell 20 μm micro-crystalline cel-
saccharides by milling of acid treated cellulose was proposed lulose (MCC) was added to this solution and stirred for a few
by Meine et al.14 However, there is a lack of understanding of minutes. The resulting slurry was air dried at 50 °C for
the mechanism of depolymerisation during the treatment. 48 hours to remove excess water to produce acidulated MCC. It
Moreover, the significance of the organic solvent used during was then milled in a planetary ball mill with cellulose to a ball
their acid impregnation is unclear. weight ratio of 1 : 10. The mill was operated at 300 rpm, with a
The focus of this research is on accomplishing complete 20 minute pause after 15 minutes of continuous milling. The
solubility of cellulose in water without the use of organic sol- pause allowed dissipation of the heat produced during
vents. We also aim to determine the factors influencing the milling, maintaining the milling temperature below 40 °C
cellulose treatment and propose a mechanism of oligomer for- throughout the treatment. The milling time reported here
mation. Here we present a pre-treatment method involving a refers only to the active milling time.
combination of acid impregnation in aqueous media and ball Solubility was determined by dissolving 0.5 g of the treated
milling to depolymerise cellulose into water soluble oligomers cellulose in 15 ml of water under ambient conditions. Undis-
at room temperature. Parameters affecting the solubility of cel- solved solids were separated via centrifugation and weighed
lulose were analysed to optimise the process with an aim to after drying. Solubility was calculated by subtracting the
achieve 100% solubility, while maintaining a high degree of weight of undissolved solids from the weight of total cellulose
polymerisation to prevent degradation of monomers. Struc- added and reported in weight percentage.
tural changes in cellulose were investigated by X-ray diffraction
and Scanning Electron Microscopy (SEM). Nuclear Magnetic 2.2 Characterisation and analysis
Resonance (NMR) spectroscopy was used to resolve repeat unit The average degree of polymerisation (DP) of cellulose was
structures of soluble fractions of cellulose. A combination of determined by methods described elsewhere.25 Typically, the
1D and 2D NMR techniques, including 1H, 13C NMR, gradient number of reducing groups in a sample was determined by
selected correlation spectroscopy (gCOSY),15 one-dimensional assaying the sample using 2,2′-bicinchoninate method (BCA).
total correlation spectroscopy (1D-TOCSY),16 multiplicity- A BCA solution was prepared fresh by mixing equal volumes of
edited heteronuclear single quantum coherence (HSQC-edit),17 solutions A and B. Solution A was prepared by dissolving
heteronuclear 2 bond correlation (H2BC),18 heteronuclear 0.097 g of disodium 2,2′-bicinchoninate, 2.71 g of Na2CO3, and
single quantum coherence total correlation spectroscopy 1.21 g of NaHCO3 in 50 ml of water. Solution B was prepared
(HSQC-TOCSY),17 and heteronuclear multiple bond correlation by dissolving 0.062 g of CuSO4·5H2O and 0.063 g of L-serine in
(HMBC),19 was used to elucidate the molecular bonding struc- 50 ml of water. 2 ml of a BCA solution was mixed with less
ture of the soluble oligomers. To the best of our knowledge, a than 1 mg of cellulose dispersed in 2 ml of water. The result-
detailed NMR study of cellulose oligomers produced via ball ing mixture was incubated at 75 °C for 30 min with constant
milling has not been reported up to now. stirring using a magnetic stirrer. After cooling to room temp-
To validate the positive effects of soluble oligomers when erature the sample was centrifuged to remove cellulose, the
compared to insoluble cellulose, we tested the catalytic conver- supernatant was then analysed for absorbance at 560 nm wave-
sion of cellulose and soluble oligomers. The cellulose sub- length using a Shimadzu UV-Vis spectrophotometer (UV-2450).
strates were reacted over a supported metal catalyst to A calibration curve was prepared using glucose solutions with
selectively produce sorbitol and mannitol (hexitols). Sorbitol concentrations of 0–150 μM. Total glucosyl monomer units in
has been identified as a key platform chemical produced from the sample were determined by dissolving cellulose in 72%
cellulose,20 which can be utilised to produce hydrogen and sulphuric acid and diluting to obtain 0.2 g L−1 cellulose con-
liquid hydrocarbon fuels.21,22 After Fukuoka et al.23 first centration. 0.10 ml of a 40 wt% phenol solution was then
reported that sorbitol can be produced from cellulose using added to 2 ml of a cellulose solution. Then 5 ml of 95% sul-
Ru/C catalysts, we recently showed that bi-metallic Ni–Pt phuric acid was added rapidly and the mixture was allowed to

Green Chem. This journal is © The Royal Society of Chemistry 2013


View Article Online

Green Chemistry Paper

stand for 10 min prior to shaking for 20 min at 25 °C. The Table 1 Effect of various acids on cellulose solubility after milling of acidulated
resulting mixture was transferred to a cuvette and tested for cellulose
absorbance at 460 nm. The average DP was calculated by divid-
Concentrationa Milling Solubility
ing the number of glycosyl monomer units by the number of Entry Acid pKa (mmol g−1) time (h) (%)
reducing ends for each sample.
Powder X-ray diffraction (XRD) of cellulose was carried out 1 None — — 5 1.5
2 C2H4O2 4.76 0.16 5 1.6
using a Rigaku Miniflex X-ray diffractometer with Fe-filtered 3 H2C2O4 1.27 0.16 5 3.6
Published on 06 August 2013. Downloaded by University of Missouri at Columbia on 22/08/2013 17:39:04.

Co radiation. Measurements were made from 10 to 50 degree 4 H3PO4 2.12 0.16 5 2.5
2θ with a step size of 0.02 degree 2θ. Samples for the Scanning 5 HCl −8.0 0.16 5 3.4
6 HNO3 −1.3 0.16 5 12.3
Electron Micrograph (SEM) were prepared by loading them 7 H2SO4 −3.0 0.16 5 34.6
onto a carbon stub and coated with a platinum or carbon a
Millimoles of acid per gram of dry cellulose.
Sputter Coater. Images were taken using a JEOL 6300 SEM
operating at 5–10 kV.
High resolution NMR spectra of completely soluble
samples were obtained using a Bruker Avance III 900 spectro- acid and phosphoric acid had little effect on solubility,
meter operating at 900.13 MHz and 226.34 MHz for 1H and whereas high strength ( pKa < 0) acids improved the solubility
13
C, respectively, using a 5 mm triple-resonance inverse (TCI) significantly. H2SO4 was the most active catalyst resulting in
cryoprobe with the sample temperature of 25 °C. 15 or 30 mg 34.5% solubility, followed by HNO3, which resulted in 12.3%
of the treated cellulose sample was dissolved in 650 μL of solubility. HCl was the exception with low pKa and low solubi-
deuterium oxide; the spectra were referenced internally to DSS lity. This could be due to loss of HCl in vapour form during
(sodium 2,2-dimethyl-2-silapentane-5-sulfonate) at 0 ppm for the drying step. The amount of acid on MCC before milling
both 1H and 13C. One-dimensional TOCSY experiments using was measured by observing the change in pH when 1 g of
gradients and a selective refocusing pulse were performed acidulated MCC (0.25 mmol of acid) was dispersed in 20 ml
using mixing times of 30–80 ms. Two-dimensional experi- of water. The observed pH value of 2.95 corresponds to
ments performed included gradient-selected COSY, multi- 0.0224 mmol of HCl which suggests that only 10% HCl was
plicity-edited HSQC, HSQC-TOCSY and H2BC. The HMBC left on the surface after drying. In contrast, dispersion of
spectrum, phase-sensitive in the indirect dimension, was H2SO4 impregnated cellulose resulted in pH of 1.83 which is
acquired with the long-range coupling delay set for 4 Hz. close to the expected value of 1.77 for 0.25 mmol of acid dis-
persed in 20 ml of water. This supports our observation of
2.3 Catalytic test higher solubility when H2SO4 was used even though its pKa is
Catalytic conversion of cellulose and soluble oligomers was higher than that of HCl. We also observed that when sulphuric
performed in a Parr high pressure batch reactor at 50 bar H2 acid was used, solubility was proportional to an increase in
pressure (at room temperature) and 200 °C for 1 h. 20 g L−1 of acid concentration (Fig. 1A). Similarly milling for a longer dur-
a cellulose solution was prepared in 100 ml of deionised water, ation also increased the solubility of MCC. It is important to
and 0.5 g of reduced Ni–Pt/alumina catalyst (see ESI†) was note that neither acid impregnation nor milling was effective
added to the reactor before applying pressure and heat. After on its own because we did not observe any change in solubility
the reaction was complete, the resulting solution was filtered under such conditions. We also found that complete solubility
using a 0.45 mm filter to recover the catalyst and unconverted can be achieved under different conditions by varying the
cellulose. The product solution was analysed using a Shi- severity of the acid amount and milling time.
madzu Prominence HPLC system equipped with a Rezex RCM MCC used in these experiments has a DP of ∼200 monomer
Monosaccharide column (7.8 × 300 mm) with 8% Ca cross units.25 Dissolution of cellulose in water can only take place
linking with an oven temperature of 65 °C and a water (mobile via depolymerisation into smaller oligomer fractions. DP was
phase) flow rate of 0.6 ml min−1. The products from the reac- analysed at each stage to analyse its effect on overall solubility
tion were detected using a Shimadzu low temperature ELSDII (Fig. 1B). Acidulation of MCC reduced its degree of polymeri-
detector at 30 °C and 370 kPa N2 pressure. sation without milling, suggesting some hydrolysis of cellulose
before the milling. Milling without acid impregnation also
resulted in depolymerisation of cellulose. However, rapid
3. Results and discussion depolymerisation of the cellulose chain into soluble oligomers
was observed only when milling was performed on acidulated
3.1 Effect of acid strength and milling time on the solubility MCC. The resulting oligomers had a DP of less than 15. We
and degree of depolymerisation found that higher milling time and acid loading improved
MCC was acidulated with various acids and milled to study the solubility without significantly changing the DP. Samples with
effect of acid strength on solubility (Table 1). In the absence of 100% solubility exhibited an average DP of 6–9 monomer units
acid, milled MCC was sparingly soluble in water (entry 1). Solu- which did not decrease further by increasing the severity of
bility was influenced by the strength of acid present during the treatment. HPLC analysis of the completely soluble oligo-
milling (entries 2–7). Low strength ( pKa > 0) acetic acid, oxalic mer solutions showed only 1.0 wt% glucose concentration.

This journal is © The Royal Society of Chemistry 2013 Green Chem.


View Article Online

Paper Green Chemistry


Published on 06 August 2013. Downloaded by University of Missouri at Columbia on 22/08/2013 17:39:04.

Fig. 1 (A) Solubility and (B) degree of polymerisation of cellulose after milling in the presence of increasing acid concentration. Milling time: (●) 0 h, (○) 2.5 h,
(▼) 5 h, (△) 7.5 h, (■) 10 h.

This confirmed that within the limits of our experiments,


depolymerisation of cellulose did not lead to the formation of
large amounts of glucose. This is advantageous since glucose
formation may lead to formation of glucose degradation
products.

3.2 Structural and chemical changes due to acid catalysed


milling
We observed that for cellulose samples with DP < 10, an
increase in solubility was unrelated to the change in DP,
suggesting other contributing factors. All further characteris- Fig. 2 SEM micrograph of cellulose: (a) C-0/0, (b) C-0/10 and (c) C-0.25/10.
ations to determine these factors were performed on the as-
received Sigmacell cellulose (C-0/0), cellulose milled for 10 h
without acid impregnation (C-0/10) and cellulose milled for The absence of acid in C-0/10 caused agglomeration of milled
10 h with 0.25 mmol g−1 H2SO4 (C-0.25/10). The last sample cellulose flakes to form spherical particles of 10 μm size after
was the only one of the three to be completely soluble in water milling (Fig. 2b). Despite being disordered, amorphous cellu-
under ambient conditions. lose exhibits hydrogen bonding between molecules which pro-
Milling MCC in a ball mill resulted in the formation of motes agglomeration of particles.9 It is evident from the SEM
amorphous cellulose. XRD peaks characteristic of microcrystal- image of C-0.25/10 (Fig. 2c) that the acid present on the cellu-
line cellulose (C-0/0) centred at 26.2° and 18.1° 2θ were lose surface interacts with the oligomers, inhibiting the
replaced by a broader peak with a maximum at 22.0° 2θ for formation of hydrogen bonds, and reduces the tendency of
C-0.25/10 (see ESI Fig. S1†), suggesting complete conversion of agglomeration. Rapid dispersion of smaller particles and
cellulose into amorphous form. Although the degree of crystal- reduced intermolecular bonding are likely to promote the solu-
lisation in C-0/10 was also very low, the sample was sparsely bility of oligomers formed during acid catalysed milling.
soluble in water. Therefore, we conclude that cellulose crystal- Authors believe that changes in the physical structure alone
linity is not a determining factor in solubility. may not result in 100% solubility of cellulose because we know
To test whether the impregnated H2SO4 was chemisorbed that oligomers with more than six glucopyranose monomer
on the cellulose surface or physisorbed in the pores, we dis- units (GMUs) are insoluble in water under ambient con-
persed never-milled acidulated cellulose powder into water. ditions.26,27 Using a solution state high resolution NMR study
We observed that the change in pH of water was instantaneous of the dissolved C-0.25/10 sample, it was possible to determine
due to the acid leaching into water. The lower pH was main- the configuration of oligomers present. The 1H NMR spectrum
tained even after filtering out the cellulose powder. SEM EDS of C-0.25/10 dissolved in D2O is presented in Fig. 3. Peaks
of acidulated MCC was performed to study surface distribution between δH 4.4 and 5.6 (δH is 1H chemical shift in ppm) are
of acid. Observation of the relative intensity of sulphur on the indicative of anomeric protons. Apart from the water peak at
cellulose surface suggested a uniform distribution of sulphuric δH 4.78, several peaks for anomeric protons were detected in
acid on the smooth surface and pits of cellulose particles (see this region pertaining to GMUs with β (1 → 4) internal linkages
ESI Fig. S2†). These findings suggest that acid is loosely (i), β (1 → 4) linked sugars with non-reducing ends (nr) and
bonded to cellulose and is evenly distributed on the surface. reducing ends in β and α configurations.28 The dominant

Green Chem. This journal is © The Royal Society of Chemistry 2013


View Article Online

Green Chemistry Paper

linked monomers with a non-reducing end (Table 2; entries


1 and 2).29 A cluster of doublets was observed between δH 4.6
and 4.7, all of which were identified as GMUs with β reducing
ends. The most abundant of these species were GMUs with β
reducing ends linked at C4 (Table 2; entry 3), followed by
monomeric β-glucose. Similarly, overlapping species of GMUs
with α reducing ends (J ∼ 3.4 Hz) were identified between δH
Published on 06 August 2013. Downloaded by University of Missouri at Columbia on 22/08/2013 17:39:04.

5.18 and 5.25. GMUs with an α reducing end in the oligomer


chain comprised the dominant species (Table 2; entry 4), fol-
lowed by monomeric α-glucose.30 These findings suggest that
monomers are primarily linked with β (1 → 4) bonds as
1
Fig. 3 The anomeric region of the H NMR spectrum of C-0.25/10 expected from cellulose based oligomers.
(30 mg ml−1) at 298 K. Peaks between δH 4.93 and 5.02 are not characteristic of
anomeric protons of any GMU present in cellulose. It was
doublet (J = 8 Hz) at δH 4.52 in our experiment was character- difficult to determine the coupling constant for these
istic of internal β (1 → 4) linked glucose monomers. Likewise, H1 multiplets in the one-dimensional proton spectrum because
the doublet (J = 8 Hz) at δH 4.50 was assigned to the β (1 → 4) of the overlap of different oligomer peaks. By examining the

Table 2 13
C and 1H chemical shifts and coupling constants for sugar residues present in C-0.25/10 dissolved in D2Oa

Proton shift δH Carbon shift δC Coupling constant


Anomeric configuration Illustration (ppm) (ppm) (Hz)

β (1 → 4) Linked (i) H1 4.52 C1 105.0 JH1,H2 8.0


H2 3.35 C2 75.6
H3 3.65 C3 76.7
H4 3.67 C4 80.9
H5 3.62 C5 77.5
H6 3.82 C6 62.5
H6′ 3.97

β (1 → 4) Linked (nr) H1 4.50 C1 105.2 JH1,H2 8.0


H2 3.31 C2 75.8
H3 3.50 C3 78.1
H4 3.41 C4 72.1
H5 3.49 C5 78.6
H6 3.73 C6 63.2
H6′ 3.91

β Reducing end H1 4.65 C1 98.4 JH1,H2 8.0


H2 3.27 C2 76.6
H3 x C3 76.9
H4 x C4 81.2
H5 x C5 77.4
H6 x C6 62.7
H6′ x

α Reducing end H1 5.21 C1 94.5 JH1,H2 3.4


H2 3.56 C2 73.9
H3 3.70 C3 x
H4 3.94 C4 81.4
H5 3.86 C5 72.8
H6 3.81 C6 62.6
H6′ x

α (1 → 6) Linked H1 4.98 C1 101.3 JH1,H2 4.0


H2 3.55 C2 74.1
H3 3.79 C3 75.6
H4 3.41 C4 72.3
H5 x C5 x
H6 x C6 63.2
H6′ x
a
δH and δC are 1H and 13C chemical shifts in ppm. (i) Internal sugar residue in an oligomer chain linked at both the C1 and C4 positions.
(nr) Sugar residue at the non-reducing end of an oligomer chain linked only at C1 position. (x) Assignment not possible due to overlap of
multiple peaks.

This journal is © The Royal Society of Chemistry 2013 Green Chem.


View Article Online

Paper Green Chemistry


Published on 06 August 2013. Downloaded by University of Missouri at Columbia on 22/08/2013 17:39:04.

Fig. 4 (A) HSQC-TOCSY and (B) HMBC spectra of C-0.25/10 (30 mg ml−1) in D2O at 298 K (13C shifts – vertical axis; 1H shifts – horizontal axis).

cross-peak to H2 in a gradient-COSY experiment it was noted peaks between δH 4.4 and 5.5 ppm in the 1H spectrum. We
that the coupling to H2 was approximately 4 Hz, which is diag- found that 68% of GMUs were linked via β (1 → 4) bonds,
nostic of H1 being in the α configuration. In a HSQC-TOCSY while 12% and 7% of GMUs had β and α reducing ends,
experiment, a proton signal in a glucose ring shows correlations respectively. This ratio is in accordance with the known equili-
to all carbons in a sugar residue due to the magnetisation trans- brium ratio of 64 : 36 for β- and α-glucose in aqueous solu-
fer between the coupled protons in the ring. In this spectrum tion.32 Approximately 9% of GMUs were linked via α (1 → 6)
(Fig. 4A) the signal at δH 4.98 shows correlations to a number of bonds and approximately 4% of sugar residues corresponded
carbon signals. In particular, the cross-peak at δH 4.98, δC 63.2 to small and broad peaks which could not be identified.
was assigned as the C6 methylene. In an HMBC experiment, a Concluding from the above results we propose that depoly-
proton signal will show cross-peaks to long-range coupled merisation of cellulose is catalysed by the presence of acids
carbon atoms (2-bond and 3-bond). Therefore, an anomeric impregnated on glycosidic bonds and the required energy for
proton will potentially exhibit correlations to H2, H3 and H5 in bond cleavage is supplied by the impact of balls on the cellu-
the sugar ring (intra-residue couplings). It can also display a lose powder during milling. Furthermore, the appearance of a
three-bond correlation across the anomeric bond to the linked high percentage of sugar units with the α (1 → 6) linkage is
sugar residue (inter-residue coupling). In the expansion of the interesting as these linkages are not a part of a cellulosic struc-
HMBC experiment (Fig. 4B), the proton at δH 4.98 exhibits a pro- ture and are not formed during depolymerisation of cellulose
minent correlation at δC 68.6 which must be an inter-residue using conventional techniques. We believe that this was
cross-peak as it does not appear in the HSQC-TOCSY spectrum. caused due to repolymerisation of monomers and low mole-
In the multiplicity-edited HSQC spectrum (Fig. S3†), the cross- cular weight oligomers during the milling process which is cat-
peak at δC 68.6 is clearly due to a CH2 group because it is the alysed by sulphuric acid. These new linkages create branching,
opposite phase to the CH sugar cross-peaks. Hence it was deter- which promotes the solubility of high molecular weight oligo-
mined that these GMUs are α (1 → 6) linked sugars. Chemical mers which are not completely soluble under ambient con-
shifts were assigned for the major component present in this ditions. The repolymerisation phenomenon also prevents
cluster. The chemical shifts were found to be in agreement with formation of glucose, which would otherwise undergo degra-
those of the α (1 → 6) linked glucose unit of the disaccharide, dation to form undesirable compounds.
isomaltose31 (see ESI†). The C4 of this major component had a
chemical shift of 72.3 ppm, which is indicative of a non-redu- 3.3 Comparison of catalytic conversion of dissolved
cing end. HSQC-TOCSY showed a low intensity component, oligosaccharides versus solid cellulose into hexitols
which had a carbon signal at 81.3 ppm, indicative of a C-4 Bi-metallic Ni–Pt catalyst supported on γ-alumina was used for
linkage. This suggests that the α (1 → 6) branching consisted of catalytic conversion of dissolved oligosaccharides and solid
one or more GMUs. One monomer unit branch was the most MCC in water suspension. The BET surface area of alumina
abundant species. was 220 m2 g−1 with an average pore size of 6 nm. Ni and Pt
The approximate composition of sugar residues based on were co-impregnated on the support with 10 wt% Ni and
the anomeric proton shifts was calculated by integrating all 1 wt% Pt, which gave an atom ratio of 22 : 1. The average metal

Green Chem. This journal is © The Royal Society of Chemistry 2013


View Article Online

Green Chemistry Paper

experiment. Rapid depolymerisation of cellulose was observed


during treatment which was proportional to acid concentration
and milling time. Soluble fractions were found to have a DP of
less than 10 monomer units. These units were primarily linked
via β (1 → 4) glycosidic bonds along with α (1 → 6) linkages
which were formed during the treatment. Complete decrystalli-
sation was also observed along with the formation of highly
Published on 06 August 2013. Downloaded by University of Missouri at Columbia on 22/08/2013 17:39:04.

dispersed particles of size less than 1 μm. Formation of


α (1 → 6) linkage was found to be responsible for enhancing
the solubility of larger oligosaccharides and reducing the
degradation of glucose. Soluble oligomers were highly reactive
towards the hydrolysis and hydrogenation reactions. High
yields of hexitols were obtained using Ni–Pt bimetallic catalyst
supported on alumina. Authors believe that this process is a
possible solution for increasing cellulose reactivity without the
Fig. 5 Yield of sugar alcohol (bars) and conversion (▲) after catalytic hydrolysis
use of hazardous conditions. It has been argued that the
and hydrogenation of cellulose. energy required for this process is comparable to other
methods on an industrial scale14,34 and understanding the
mechanism of depolymerisation will result in improving the
surface area of the catalyst as measured by CO-TPD was efficiency, making it feasible for commercial application.
4.2 m2 g−1. Although the hydrogenation activity of Ni catalyst
is lower than that of Ru and Pt, it is significantly economical.
Here the hydrogenation activity of Ni is promoted by addition
of a small amount of Pt. The relative yield of hexitol from cellu- Acknowledgements
lose is similar to noble catalyst with 3–5% metal loading.24,33 This work was supported by the National and International
Using soluble oligosaccharides allows access to all glycoside Research Alliances Program (NIRAP) funded by Queensland
bonds present, increasing the rate of hydrolysis, and therefore State Government, Australia. The authors would like to
higher polyol yield is achieved at lower reaction times. Partial acknowledge the work of Ms Jun Jhang from the ARC Centre of
hydrolysis due to depolymerisation also reduces the reaction Excellence for Functional Nanomaterials at the University of
time during the catalytic test. ∼80% yield of hexitols was Queensland, who helped with SEM-EDS experiments.
obtained with the dissolved C-0.25/10 sample after only 1 h of
reaction (Fig. 5, C-0.25/10(A)).
Glucose produced as a result of hydrolysis undergoes rapid
hydrogenation to form sorbitol. Sulphuric acid present on the References
surface of the oligosaccharides also catalyses the hydrolysis 1 A. Wiselogel, S. Tyson and D. Johnson, in Handbook on
reaction, thereby increasing the rate of conversion substan- Bioethanol: Production and Utilization, ed. C. E. Wyman,
tially. However, sulphuric acid also catalysed dehydration of Taylor & Francis, 1996, pp. 105–118.
sorbitol to form 1–4 anhydro sorbitol. To avoid the 2 G. W. Huber, S. Iborra and A. Corma, Chem. Rev., 2006,
dehydration reaction, the C-0.25/10 sample was neutralised 106, 4044–4098.
with Ba(OH)2 and the resulting BaSO4 precipitate was separ- 3 B. Medronho, A. Romano, M. Miguel, L. Stigsson and
ated by centrifugation. Yields of sorbitol and mannitol B. Lindman, Cellulose, 2012, 19, 581–587.
increased to ∼90% (Fig. 5, C-0.25/10 (B)) after the removal of 4 W. Glasser, R. Atalla, J. Blackwell, R. Malcolm Brown Jr.,
sulphuric acid, whereas 1–4 anhydro sorbitol yields decreased W. Burchard, et al., Cellulose, 2012, 19, 599–599.
from 10.8% to 1.4%. The results prove that 1 h reaction time 5 A. Pinkert, K. N. Marsh and S. Pang, Ind. Eng. Chem. Res.,
was sufficient for complete conversion of oligosaccharides in 2010, 49, 11121–11130.
the absence of sulphuric acid, while increasing the hexitol 6 J. Cai and L. Zhang, Macromol. Biosci., 2005, 5, 539–548.
yield by avoiding the dehydration reaction. In comparison, the 7 M. Sasaki, Z. Fang, Y. Fukushima, T. Adschiri and K. Arai,
yield of polyols from untreated cellulose (C-0/0) and milled Ind. Eng. Chem. Res., 2000, 39, 2883–2890.
cellulose (C-0/10) was substantially lower. 8 T. Paakkari, R. Serimaa and H. P. Fink, Acta Polym., 1989,
40, 731–734.
9 K. Mazeau and L. Heux, J. Phys. Chem. B, 2003, 107, 2394–
4. Conclusions 2403.
10 M. S. Bertran and B. E. Dale, J. Appl. Polym. Sci., 1986, 32,
Water soluble oligomers were produced by milling acidulated 4241–4253.
cellulose. Sulphuric acid was found to be the most active 11 T. Jeoh, C. I. Ishizawa, M. F. Davis, M. E. Himmel,
towards producing the water soluble fraction in our W. S. Adney, et al., Biotechnol. Bioeng., 2007, 98, 112–122.

This journal is © The Royal Society of Chemistry 2013 Green Chem.


View Article Online

Paper Green Chemistry

12 H. Zhao, J. H. Kwak, Y. Wang, J. A. Franz, J. M. White, 22 G. W. Huber, R. D. Cortright and J. A. Dumesic, Angew.
et al., Energy Fuels, 2005, 20, 807–811. Chem., Int. Ed., 2004, 116, 1575–1577.
13 Y. Yu and H. Wu, Ind. Eng. Chem. Res., 2010, 49, 3902– 23 A. Fukuoka and P. L. Dhepe, Angew. Chem., Int. Ed., 2006,
3909. 45, 5161–5163.
14 N. Meine, R. Rinaldi and F. Schüth, ChemSusChem, 2012, 5, 24 A. Shrotri, A. Tanksale, J. N. Beltramini, H. Gurav and
1449–1454. S. V. Chilukuri, Catal. Sci. Technol., 2012, 2, 1852–1858.
15 S. Braun, H.-O. Kalinowski and S. Berger, 150 and more 25 Y. H. P. Zhang and L. R. Lynd, Biomacromolecules, 2005, 6,
Published on 06 August 2013. Downloaded by University of Missouri at Columbia on 22/08/2013 17:39:04.

basic NMR experiments: a practical course, Wiley-VCH, Wein- 1510–1515.


heim, New York, 1998. 26 J. B. Taylor, Trans. Faraday Soc., 1957, 53, 1198–1203.
16 G. Xu and J. S. Evans, J. Magn. Reson., Ser. B, 1996, 111, 27 G. L. Miller, Methods Carbohydr. Chem., 1963, 3, 134–139.
183–185. 28 H. Sugiyama, K. Hisamichi, T. Usui, K. Sakai and
17 T. D. W. Claridge, High-resolution NMR techniques in organic J. i. Ishiyama, J. Mol. Struct., 2000, 556, 173–177.
chemistry, Pergamon, Amsterdam, New York, 1999. 29 L. A. Flugge, J. T. Blank and P. A. Petillo, J. Am. Chem. Soc.,
18 N. T. Nyberg, J. Ø. Duus and O. W. Sørensen, J. Am. Chem. 1999, 121, 7228–7238.
Soc., 2005, 127, 6154–6155. 30 A. Teleman, K. Kruus, E. Ämmälahti, J. Buchert and
19 D. O. Cicero, G. Barbato and R. Bazzo, J. Magn. Reson., K. Nurmi, Carbohydr. Res., 1999, 315, 286–292.
2001, 148, 209–213. 31 D. S. Wishart, D. Tzur, C. Knox, R. Eisner, A. C. Guo, et al.,
20 T. Werpy and G. Petersen, Top Value Added Chemicals from Nucleic Acids Res., 2007, 35, D521–D526.
Biomass Volume I-Results of Screening for Potential Candi- 32 S. J. Angyal, Angew. Chem., Int. Ed. Engl., 1969, 8, 157–166.
dates from Sugars and Synthesis Gas, National Renewable 33 P. L. Dhepe and A. Fukuoka, Catal. Surv. Asia, 2007, 11,
Energy Laboratory (NREL), 2004. 186–191.
21 D. M. Alonso, J. Q. Bond and J. A. Dumesic, Green Chem., 34 S. M. Hick, C. Griebel, D. T. Restrepo, J. H. Truitt,
2010, 12, 1493–1513. E. J. Buker, et al., Green Chem., 2010, 12, 468–474.

Green Chem. This journal is © The Royal Society of Chemistry 2013

You might also like