You are on page 1of 6

Catalysis Communications 10 (2009) 15581563

Contents lists available at ScienceDirect

Catalysis Communications
journal homepage: www.elsevier.com/locate/catcom

Catalytic conversion of glucose to 5-hydroxymethylfurfural over SO42/ZrO2


and SO42/ZrO2Al2O3 solid acid catalysts
Hongpeng Yan, Yu Yang, Dongmei Tong, Xi Xiang, Changwei Hu *
Key Laboratory of Green Chemistry and Technology, Ministry of Education, College of Chemistry, Sichuan University, Chengdu, Sichuan 610064, PR China

a r t i c l e

i n f o

Article history:
Received 14 December 2008
Received in revised form 3 April 2009
Accepted 14 April 2009
Available online 24 April 2009
Keywords:
Glucose
5-Hydroxymethylfurfural
SO42/ZrO2
SO42/ZrO2Al2O3

a b s t r a c t
SO42/ZrO2 and SO42/ZrO2Al2O3 catalysts were prepared by impregnation of Zr(OH)4 and Zr(OH)4
Al(OH)3 with ethylene dichloride solution of chlorosulfonic acid, and characterized by ICP-AES, BET,
XRD, NH3-TPD, CO2-TPD and FTIR of adsorbed pyridine. The catalysts were used in the catalytic conversion of glucose to 5-hydroxymethylfurfural. An optimized 5-hydroxymethylfurfural yield of 47.6% was
obtained within 4 h at 403 K over SO42/ZrO2Al2O3 with ZrAl mole ratio of 1:1. The catalyst with higher
acidity and moderate basicity was more favorable for the formation of the target product.
2009 Published by Elsevier B.V.

1. Introduction
Increasing interest has been devoted in recent years to the
selective dehydration of hexose to 5-hydroxymethylfurfural
(HMF) [14], as HMF and its derivatives have potential applications
in the production of ne chemicals, pharmaceuticals, plastics [3]
and liquid alkanes [5]. HMF could be formed with high yield by
the acid-catalyzed dehydration of fructose [1,69]. However, one
of the key problems for the use of fructose is its high cost. Another
hexose, glucose, which is the unit compound of cellulose, is one of
the most important starting chemicals from biomass. With its low
cost and wide supply, the conversion of glucose to HMF has attracted the interests of researchers [2,1013]. For gaining a high
HMF yield, the choice of catalysts is very important. Several kinds
of catalysts, such as CrCl2 in the 1-alkyl-3-methylimidazolium
chloride [2], H2SO4 [9], and TiO2 (anatase) [10,12,13] were employed. However, the catalysts mentioned above have some disadvantages, such as pollution, low yield, separation and recycling
problems. Therefore, the development of environmentally benign
catalysts with high activity is expected.
SO42/ZrO2 and modied SO42/ZrO2 solid acid catalysts have
attracted considerable interests due to their strong acidity and catalytic activity, and have been widely used to catalyze dehydration,
isomerization, acylation, esterication, alkylation reactions and so
on [1417]. H2SO4, (NH4)2SO4 and (NH4)2S2O8 solutions were usually employed as sulfating agents [18,19]. Recently, Yadav et al. reported a novel SO42/ZrO2 catalyst by the introduction of ethylene
* Corresponding author. Tel./fax: +86 28 85411105.
E-mail addresses: gchem@scu.edu.cn, chwehu@mail.sc.cninfo.net (C. Hu).
1566-7367/$ - see front matter 2009 Published by Elsevier B.V.
doi:10.1016/j.catcom.2009.04.020

dichloride solution of chlorosulfonic acid as sulfating agent. The


catalyst showed optimal performance in hydrocarbon isomerization, benzylation and alkylation [20,21]. In the present work,
SO42/ZrO2 and SO42/ZrO2Al2O3 catalysts bearing different acidity and basicity were prepared, and their catalytic performance in
the conversion of glucose to HMF was studied.
2. Experimental
2.1. Material
5-Hydroxymethylfurfural used for calibration was obtained
from Aldrich (99% purity). Other reagents and chemicals were all
of analytical grade from Kelong Chemical Company (Chengdu, China) and used without further purication.
2.2. Catalyst preparation
SO42/ZrO2 and SO42/ZrO2Al2O3 solid acid catalysts were synthesized according to the procedures reported in the literatures
[19,20]. The impregnation solution was prepared by dissolving
0.1 mol chlorosulfonic acid in 100 mL ethylene dichloride. Powered
Zr(OH)4 and Al(OH)3Zr(OH)4 were immersed in the above ethylene dichloride solution with a solution/solid ratio of 7.5 ml/g for
30 min. The prepared materials were calcined at 823 K for 4 h.
The SO42/ZrO2 solid acid catalyst was labeled as CSZ, and SO42/
ZrO2Al2O3 samples as CSZA-1, CSZA-2, CSZA-3, CSZA-4, and
CSZA-5 according to the ZrAl mol ratio of 9:1, 7:3, 1:1, 3:7, and
1:9, respectively. The catalyst particle size used for the catalytic
tests was between 100 and 120 mesh.

1559

H. Yan et al. / Catalysis Communications 10 (2009) 15581563

2.3. Catalysts characterization


The BET surface areas and pore volume distributions of the
catalysts were measured on a Quantachrome NOVA 1000e surface
area analyzer by N2 adsorption at 77 K. Prior to analysis, each
sample was degassed at 573 K for 3 h in vacuum under 103 Torr.
BET nitrogen porosimetry analysis was done with multipoint
calibration curves. The pore volume was calculated from the
adsorption isotherms using BJH (BarrettJoynerHalenda) method. ICP-AES was performed on an IRIS Advantage ER/S instrument
for determining the actual contents of Zr and Al elements in the
catalysts.
X-ray diffraction (XRD) patterns of the solid catalysts were recorded on a DX-1000 diffractometer, operating at 40 kV and
25 mA and using nickel-ltered Cu Ka radiation. Scanning speed
was 3.6 min1 and scanning range was over 1070o. FTIR of adsorbed pyridine on the catalysts were recorded on a NEXUS
670FTIR spectrometer. Before measurement, the catalyst sheets
were evacuated at 573 K for 2 h under vacuum of 102 Pa and
then cooled to ambient temperature. After pyridine adsorption
for 30 min and evacuation at 473 K for 1 h, IR spectra were
recorded.
NH3-TPD was performed in a xed-bed reactor at atmosphere
pressure. The catalysts were swept by N2 ow at 573 K for 1 h
and then cooled to 353 K. Pure NH3 was introduced for about
20 min, followed by purging with N2 for 30 min at 353 K; then
the samples were purged further with He ow (50 ml/min) for
2 h at 373 K. The samples were heated at a ramp of 10 K/min up
to 973 K. NH3 desorbed was analyzed by GC with a thermal conductivity detector (TCD). CO2-TPD was performed similarly to
NH3-TPD with adsorption at ambient temperature instead of
353 K, and the TPD curve was recorded from 303 to 773 K.
2.4. Activity tests and analytical procedures
Batch catalytic experiments were carried out under nitrogen
atmosphere in a 5 ml glass vessel equipped with a reex condenser. In a typical reaction procedure, a dimethylsulfoxide
(DMSO) solution of glucose (420 wt.%) was poured into the
reactor along with 18 mg solid catalysts. DMSO was used as solvent for it could suppress some undesirable parallel reactions
and the rehydration of HMF [1,8]. The reaction mixtures were
heated with oil bath and stirred magnetically at 400 rpm. The
reaction endured 4 h starting from the mixing of all the components and it took about 22 min to heat the system to the desired
temperature.
HMF was identied qualitatively by GCMS (Agilent, 5973 Network 6890N). When the reaction was over, solid catalyst and the
reaction mixture were separated by centrifugation. Saturated sodium chloride solution was poured into the mixture and the products were extracted by diethyl ether. No rehydration by-products
such as levulinic acid and formic acid were detected in the diethyl
ether extracted samples. The quantitative analysis of HMF was performed by HPLC using a Waters 1525 pump, a SB-C18 reverse
phase column (Aglient) and Waters 2487 UV detector (284 nm).
A mixture of acetonitrile and water (1:9 v/v, pH 2) was used as
the mobile phase at a ow rate of 1.2 ml/min and a column temperature of 303 K. The appearance of fructose and residual glucose
were determined by GC analysis through a silylation step, and
hexadecane was used as internal standard, according to the literature [22]. Thus the by-products were ascribed to fructose and humins formed via condensation reactions [23]. A SE-54 capillary
column (30 m  0.25 mm), a hydrogen ame-ionization detector
(FID) and ZB-2020 integrator were employed in GC analysis. The
analytical error in HMF yield and glucose conversion was evaluated
to be in the 1% range.

Turnover number was calculated as follows:

Turnover number



moles of HMF formed
1
mmolg1  h
mass of catalyst  retention time

3. Results and discussion


3.1. Actual Zr/Al mole ratio, BET surface areas and pore volume of the
catalysts
The ICP-AES results revealed that only small difference between
the actual and controlled mole ratio of ZrAl was observed in CSZA
catalysts. The BET surface areas and pore volume of various samples were shown in Table 1. The BET surface areas of all the catalysts were much lower than those reported in the literatures
[20,24] and this may be due to the formation of Zr(SO4)2 in CSZ catalyst or Al2(SO4)3 salt in CSZA catalysts, which can be observed in
XRD patterns, and will be discussed later. This may be attributed to
the inuence of several factors in the preparation such as the calcination temperature, sulfating agents and their concentration
[25,26]. As can be seen from Table 1, the BET surface areas increased with the addition of Al, and changed slightly when the
ZrAl mole ratio changing from 7:3 to 3:7. When the amount of
Al further increased to give a ZrAl mole ratio of 1:9, both the
BET specic area and pore volume arrived at their maxima of
66 m2/g and 0.09 cm3/g, respectively. This trend was similar to that
reported by Hua et al. [18].
3.2. XRD
XRD patterns of the samples were shown in Fig. 1. The formation of Zr(SO4)2 and monoclinic ZrO2 phase were observed in CSZ,
and Al2(SO4)3 crystalline phase emerged on CSZA series in addition
to zirconium species. Similar phenomenon was also observed in
the SO42/ZrO2TiO2 sample, where both Ti2(SO4)3 and Zr(SO4)2
crystalline phase were observed [27]. When small amount of Al
was added into CSZ, a new ZrO phase was formed. And ZrS phase
formed when the amount of Al was further increased. When the
ZrAl mole ratio was 1:1, the formation of the tetragonal ZrO2
phase was observed. It might be due to the interaction of Al, Zr
and sulfating agents resulting in the formation of AlOZr bands
[18]. Besides, among all the catalysts, no diffraction patterns for
Al2O3 were observed, indicating that the Al2O3 was sufciently
homogeneously mixed with ZrO2 [19]. SEM characterization revealed that the surface of the samples became more inhomogeneous with increasing amount of Al.
3.3. TPD measurement
The amounts of acid and base sites on the catalyst surface, and
the peak temperatures of the TPD spectra were listed in Table 2.
The amount of acid sites on CSZ was 1.83 mmol/g, which was the

Table 1
The BET surface areas, pore volume, and ZrAl mole ratio of various catalysts.
Catalyst

Zr:Al controlled
(mol ratio)

Zr:Al (mol
ratio)a

Surface area
(m2/g)

Pore volume
(cm3/g)

CSZ
CSZA-1
CSZA-2
CSZA-3
CSZA-4
CSZA-5

9:1
7:3
1:1
3:7
1:9

9:0.9
7:2.5
1:1
3:7
1:8.7

6.87
11.1
28.2
27.3
26.6
66.0

0.008
0.015
0.034
0.038
0.042
0.090

Determined by ICP-AES analysis.

1560

H. Yan et al. / Catalysis Communications 10 (2009) 15581563

Fig. 1. X-ray powder diffraction patterns of SO42/ZrO2 and SO42/ZrO2Al2O3


catalysts calcined at 823 K. (A) CSZ, (B) CSZA-1, (C) CSZA-2, (D) CSZA-3, (E) CSZA-4
and (F) CSZA-5.

highest among all the catalysts. The total amounts of acid sites on
CSZA catalysts were lower than that on CSZ catalyst due to the
addition of Al, and decreased gradually with increasing amount
of Al, varying from 1.79 mmol/g for CSZA-1 to 0.95 mmol/g for
CSZA-5. The NH3-TPD proles were shown in Fig. 2a. It is shown
that all the catalysts afforded a broad desorption proles over
the range from 373 to 973 K, suggesting a broad distribution of
heterogeneous acid sites. The acid sites existed on the samples,
were articially divided into three categories which has weak-,
middle-, and strong acidity, respectively, according to the classication in the literatures [2829]. The amount of acid sites on the
CSZ sample was the highest; however, CSZA-1 and CSZA-3 catalysts exhibited stronger acidity than CSZ catalyst. This was also observed by Hua et al. [18] when adding small amount of Al2O3 into
SO42/ZrO2. It was reported that the increase of acid strength was
caused by the formation of AlOZr generating additional acidic
active sites. However, when the amount of Al increased further,
the acid strength of the CSZA catalysts decreased gradually. There
were several strong acid sites on CSZA-3, which gave NH3-TPD
peaks at 693, 743 and 823 K, respectively. Combined with the results of XRD, the enhanced acid strength on the CSZA-3 catalyst
might be related to the formation of tetragonal ZrO2 phase [20].
It was also reported that the presence of weak basic sites on the
SO42/ZrO2 catalysts was due to the basicity of ZrO2 [24]. As can be
seen from Fig. 2b, the highest CO2-TPD peak of all the samples was
over the range of 620670 K, which indicated the presence of
mainly middle strength base on the catalysts. The amount of base
sites of CSZ was 0.32 mmol/g, which was the lowest among all the
catalysts. Compared with CSZ catalyst, the amounts of base sites on
CSZA samples were higher, which increased from 0.35 to
0.98 mmol/g with the ZrAl mole ratio changing from 9:1 to 1:9.
This change was ascribed to the presence of aluminum species,

Fig. 2. NH3-TPD and CO2-TPD proles of the catalysts. (A) CSZ, (B) CSZA-1, (C) CSZA2, (D) CSZA-3, (E) CSZA-4 and (F) CSZA-5.

showing stronger basicity than ZrO2 [30]. Part of the base sites
based on aluminum species might not be affected by chlorosulfonic acid in the impregnation process.
3.4. FTIR studies
The results of FTIR of adsorbed pyridine were shown in Fig. 3.
Three peaks due to CC stretching vibrations of pyridine were
observed. The peak at 15401541 cm1 was assigned to pyridine
adsorbed on Brnsted acid sites, and that at 14891491 cm1

Table 2
NH3- and CO2-TPD data of the catalysts.
Catalyst

Total NH3 desorbed (mmol/g)

Peak temp. of NH3-TPD (acid sites) (K)

Total CO2 desorbed (mmol/g)

Peaks temp. of CO2-TPD (base sites) (K)

CSZ
CSZA-1
CSZA-2
CSZA-3
CSZA-4
CSZA-5

1.83
1.79
1.60
1.55
1.49
0.95

463,
633,
473,
693,
593,
643,

0.32
0.35
0.49
0.52
0.84
0.98

623
653
653
573, 613
623, 663
623

743, 853
813
603, 763
743, 823
703
683

1561

H. Yan et al. / Catalysis Communications 10 (2009) 15581563

Fig. 3. FTIR spectra recorded after the chemisorption of pyridine on CSZ and CSZA
catalysts and evacuation at 473 K. (A) CSZ, (B) CSZA-1, (C) CSZA-2, (D) CSZA-3, (E)
CSZA-4 and (F) CSZA-5.

was characteristic of pyridine adsorbed on both Brnsted and Lewis acid sites while the peak around 1640 cm1 was related to pyridine adsorbed on Brnsted acid sites [31]. From the spectra we
can see that the amounts of acid sites decreased with increasing
amount of Al. This was consistent with the observation of NH3TPD. The quantity of Brnsted acids over SCZ sample was higher
than those over CSZA, while CSZA-5 had the lowest amount of
Brnsted acids. It was notable that no bands indicative of pyridine
adsorbed on Lewis acid sites were observed around 1450 cm1,
which was different from that obtained over SO42/ZrO2 when taking H2SO4 solution as sulfating agent. The absence of peak around
1450 cm1 was also found when SF4 was used as sulfating agent
[32].
3.5. Catalytic activity and catalyst stability
As seen from Table 3, in the absence of catalysts, the yield of
HMF was only 4.3%. By the use of CSZ, HMF yield was enhanced
signicantly with slight increase of glucose conversion. Vogel
et al. [9] has reported the increase of HMF yield from glucose when
H2SO4 was used as the catalyst, but the improvement was less evident than that when fructose was used as starting material. Compared with CSZ, the yield of HMF was further increased by the use
of CSZA catalysts. The CSZA-3 sample with the ZrAl mole ratio of
1:1 was found to be the most effective, leading to a yield of 47.6%

for HMF within 4 h reaction at 403 K, and a TON value of 3.31.


When the concentration of glucose increased from 7.6% to 20%, a
progressive increase of the TON to 9.03 was achieved. Early researches [2,13,33] indicated that in situ glucosefructose isomerization was one of the way for the conversion of glucose to HMF,
and the isomerization could improve the yield of HMF. As shown
in Fig. 4, fructose was observed in the process. Its concentration
achieved the highest at 22 min then disappeared gradually with
increasing yield of HMF. The yield of fructose varied almost linearly
with the amounts of base sites on the catalysts. This evidenced that
glucosefructose isomerization occurred in the present system. To
investigate the effect of isomerization on the formation of HMF, a
control experiment using fructose as the reactant was carried out
and the results were listed in the Table 4. As seen from Table 4,
the yield of HMF was about 72% without catalyst and decreased
monotonically with the basicity of the catalysts. This evidenced
that DMSO was an excellent solvent for the fructose dehydration
to HMF [34]. Early researchers also reported that the existence of
base in some extent suppressed the dehydration of fructose to
HMF [10,13]. This suggested that the production of HMF in the
present system may not be mainly via isomerizationdehydration
process, though it might contribute to HMF formation. The acid
sites on the CSZA catalysts exhibited no catalytic performance for
the conversion of fructose to HMF under the conditions tested.
There were more base sites on the surface of CSZA-4 and CSZA-5
catalysts, which made the glucosefructose isomerization more
obviously; however, the yields of HMF were lower than CSZA-3,
indicating the possibility of HMF formation via non-isomerization
process in the present system. The yield of HMF might be due to
an integrated effect of suitable amounts of acid and base sites on
the surface of the catalysts. The CSZA-3 sample may just bear the
compromised amounts of acid and base sites, and show the best
catalytic performance.
The catalyst stabilities and reusability were tested using CSZA-3
sample. After reaction, the catalyst was separated by centrifugation, washed with diethyl ether, dried at 403 K and calcined at
823 K for 2 h to remove adsorbed by-products prior to reuse in
the next run. As seen from Table 5, although a decrease of the
HMF yield from 47.6% to 35.2% was observed in the second run,
no further decrease in activity was observed in the subsequent
runs, that is to say, a stable HMF yield of about 35% could be obtained. The decrease of activity in the second run may be attributed
to partially loss of surfur [35]. However, it should be pointed out
that the deactivation of the catalyst was observed in successive
runs when the catalyst was reused without calcination, and the
yield of HMF decreased from 47.6% to 5.53%. Fortunately, the activity of catalyst could be recovered by calcination and a HMF yield of
33.6% (close to the stable value) was regained. It indicated that the
deactivation of the catalyst was mainly caused by the adsorbed hu-

Table 3
Conversion of glucose to HMF in the presence of CSZ and CSZA catalystsa.
Catalyst

Substrate

Rb (w/w)

Time (h)

Conversion (%)

HMF yield (mol%)

Tonc

No
CSZ
CSZA-1
CSZA-2
CSZA-3

Glucose
Glucose
Glucose
Glucose
Glucose
Glucose
Glucose
Glucose
Glucose
Glucose
Glucose

5
5
5
2.5
5
5
5
16.7
5
5

4
4
4
4
4
4
6
15
4
4
4

94.0
95.2
97.3
98.1
99.1
97.2
100
100
98.1
97.8
95.2

4.3
19.2
27.3
41.5
48.0
47.6
47.9
47.6
39.2
43.1
37.0

1.32
1.87
2.85
1.67
3.31
2.22
0.88
9.03
2.98
2.57

CSZA-4
CSZA-5
a
b
c

(7.6 wt%)
(7.6 wt%)
(7.6 wt%)
(7.6 wt%)
(3.9 wt%)
(7.6 wt%)
(7.6 wt%)
(7.6 wt%)
(20 wt%)
(7.6 wt%)
(7.6 wt%)

Reaction conditions, T = 403 K; Solvent, DMSO, under N2 protection.


Substrate to catalyst weight ratio (g/g).
Turnover number expressed as: mmol of HMF/(g of catalyst  h).

1562

H. Yan et al. / Catalysis Communications 10 (2009) 15581563

Fig. 4. (a) Effect of the retention time on the yield of fructose and HMF at 403 K over CSZA-3. (b) Effect of the amounts of base sites over different catalysts on the yield of
fructose at 22 min.

Table 4
Dehydration of fructose to HMF in the presence of CSZ and CSZA catalystsa.
Catalyst

Substrate

Rb (w/w)

Time (h)

Conversion (%)

HMF yield (mol%)

Tonc

No
CSZ

Fructose
Fructose
Fructose
Fructose
Fructose
Fructose
Fructose
Fructose
Fructose

5
2.5
16.7
5
5
5
5
5

4
4
4
4
4
4
4
4
4

99.6
99.8
100
100
99.6
100
99.4
99.7
99.1

71.9
67.7
68.2
61.6
64.2
58.1
56.6
52.2
47.2

4.71
2.36
14.1
4.46
4.02
3.93
3.63
3.28

CSZA-1
CSZA-2
CSZA-3
CSZA-4
CSZA-5
a
b
c

(7.6 wt%)
(7.6 wt%)
(3.9 wt%)
(20 wt%)
(7.6 wt%)
(7.6 wt%)
(7.6 wt%)
(7.6 wt%)
(7.6 wt%)

Reaction conditions, T = 403 K; Solvent, DMSO, under N2 protection.


Substrate to catalyst weight ratio (g/g).
Turnover number expressed as: mmol of HMF/(g of catalyst  h).

mins and the removal of these species made the catalyst reactivated. How to improve the stability of the catalyst remained a
problem for future study.
4. Conclusions
The CSZ catalyst prepared contained monoclinic ZrO2 phase and
Zr(SO4)2 phase. The crystalline structure varied with the amount of
aluminum species on the series of CSZA. Both Brnsted and Lewis

acid sites existed on the catalysts and the former was stronger. The
amounts of acid sites decreased, while base sites increased with
increasing amount of Al. Basicity promoted the glucosefructose
isomerization. CSZA series catalysts were more active than CSZ
for the catalytic conversion of glucose to HMF. In particular, over
CSZA-3 with ZrAl mole ratio of 1:1, a yield of about 47.6% to
HMF was obtained with a turnover number (TON) of about 3.31.
The activity of catalyst could be recovered, and a stable HMF yield
of about 35% could be obtained.

H. Yan et al. / Catalysis Communications 10 (2009) 15581563


Table 5
Yield of HMF and conversion of glucose in the recycling runs.
Run

Catalyst

Conversion (%)

HMF yield (mol%)

1
2
3
4
5

CSZA-3
CSZA-3
CSZA-3
CSZA-3
CSZA-3b

97.2
92.3 (95.3)a
92.8 (94.8)
92.0 (93.9)
94.0

47.6
35.2 (21.6)a
35.8 (8.62)
35.5 (5.53)
33.6

[7]
[8]
[9]
[10]
[11]
[12]

Conversion of glucose and yield of HMF when the catalyst was reused without
calcination prior to the next run.
b
The catalyst successively used for four times without calcination (almost
completely deactivated) was calcined at 823 K for 2 h and reused.

[13]
[14]
[15]
[16]
[17]
[18]
[19]

Acknowledgements

[20]
[21]
[22]

This work was nancially supported by National Basic Research


Program of China (973 program, No. 2007CB210203) and the Special Research Fund for the Doctoral Program of Higher Education of
China (No. 20050610013).
Appendix A. Supplementary material

[23]
[24]
[25]
[26]
[27]

Supplementary data associated with this article can be found, in


the online version, at doi:10.1016/j.catcom.2009.04.020.

[28]
[29]

References

[30]
[31]

[1]
[2]
[3]
[4]
[5]
[6]

Y. Roman-Leshkov, J.N. Chheda, J.A. Dumesic, Science 312 (2006) 1933.


H.B. Zhao, J.E. Holladay, H. Brown, Z.C. Zhang, Science 316 (2007) 1597.
J. Lewkowski, Arkivoc 2 (2001) 17.
A. Corma, S. lborra, A. Velty, Chem. Rev. 107 (2007) 2411.
G.W. Huber, J.N. Chheda, C.J. Barrett, J.A. Dumesic, Science 308 (2005) 1446.
C. Lansalot-Matras, C. Moreau, Catal. Commun. 4 (2003) 517.

[32]
[33]
[34]
[35]

1563

C. Moreau, A. Finiels, L. Vanoye, J. Mol. Catal. A: Chem. 253 (2006) 165.


Q. Bao, K. Qiao, D. Tomida, C. Yokoyama, Catal. Commun. 9 (2008) 1383.
M. Bicker, J. Hirth, H. Vogel, Green Chem. 5 (2003) 280.
M. Watanable, Y. Aizawa, T. Iida, T.M. Aida, C. Levy, H. Inomata, Carbohydr. Res.
340 (2005) 1925.
T.M. Aida, Y. Sato, M. Watanabe, K. Tajima, T. Nonaka, J. Supercrit. Fluids 40
(2007) 381.
M. Watanabe, Y. Aizawa, T. Iida, R. Nishimura, H. Inomata, Appl. Catal. A: Gen.
295 (2005) 150.
X.H. Qi, M. Watanabe, T.M. Aida, R.L. Smith Jr., Catal. Commun. 9 (2008) 2244.
A. Corma, Chem. Rev. 95 (1995) 559.
P.J. Skrdla, C. Lindemann, Appl. Catal. A: Gen. 246 (2003) 227.
J.H. Wang, C.Y. Mou, Micropor. Mesopor. Mater. 110 (2008) 260.
D.E. Lopez, J.G. Goodwin Jr., D.A. Bruce, S. Fruta, Catal. Commun. 339 (2008) 76.
W.M. Hua, Y.D. Xia, Y.H. Yue, Z. Gao, J. Catal. 196 (2000) 104.
B.M. Reddy, P.M. Sreekanth, Y. Yamada, T. Kobayashi, J. Mol. Catal. A: Chem.
227 (2005) 81.
G.D. Yadav, A.D. Murkute, J. Catal. 224 (2004) 218.
G.D. Yadav, G.S. Pathre, Appl. Catal. A: Gen. 297 (2006) 237.
C.C. Sweeley, R. Bentley, M. Makita, W.W. Wells, J. Am. Chem. Soc. 85 (1963)
2497.
B.M.F. Kuster, H.S. van der Baan, L.M. Tebbens, Carbohydr. Res. 54 (1977)
159.
V. Vishwanathan, G. Balakrishna, B. Rajesh, V. Jayasri, N.J. Coville, Catal.
Commun. 9 (2008) 2422.
R. Srinivasan, C.R. Hubbard, O.B. Cavin, B. Davis, Materials 5 (1993) 27.
A. Corma, V. Fornes, M.I. Juan-Rajadell, J.M. Lopez-Nieto, Appl. Catal. A: Gen.
116 (1994) 151.
B.M. Reddya, P.M. Sreekanth, Y. Yamada, Q. Xub, T. Kobayashi, Appl. Catal. A:
Gen. 228 (2002) 269.
G.D. Yadav, A.D. Murkute, Adv. Synth. Catal. 346 (2004) 389.
S.R. Kirumakki, B.G. Shpeizer, G.V. Sagar, K.V.R. Chary, A. Cleareld, J. Catal. 242
(2006) 319.
K.I. Shimizu, Y. Saito, N. Miyoshi, A. Satsuma, Catal. Today 139 (2008) 24.
B.H. Davis, R.A. Keogh, S. Alerasool, D.J. Zalewski, D.E. Day, J. Catal. 183 (1999)
45.
V. Quaschning, J. Deutsch, P. Druska, H.J. Niclas, E. Kemnitzi, J. Catal. 177
(1998) 164.
B.F.M. Kuster, Starch 42 (1990) 314.
R.M. Musau, R.M. Munavu, Biomass 13 (1987) 67.
F.T.T. Ng, N. Hovat, Appl. Catal. A: Gen. 123 (1995) 197.

You might also like