Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

Molecular Catalysis 438 (2017) 167–172

Contents lists available at ScienceDirect

Molecular Catalysis
journal homepage: www.elsevier.com/locate/mcat

Application of sulfonated carbon-based catalyst for the furfural


production from d-xylose and xylan in a microwave-assisted biphasic
reaction
Yantao Wang a , Frederic Delbecq b , Witold Kwapinski c , Christophe Len a,∗
a
Sorbonne Universités, Université de Technologie de Compiègne (UTC), CS 60319, 60203 Compiègne Cedex, France
b
Ecole Supérieure de Chimie Organique et Minérale (ESCOM), 1 allée du Réseau Jean-Marie Buckmaster, 60200 Compiègne, France
c
Chemical Sciences Department, Bernal Institute, Faculty of Science and Engineering, University of Limerick, Limerick, Ireland

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

Article history: Sulfonated carbon-based catalyst has been prepared, characterized by elemental analysis, SEM, EDX and
Received 19 March 2017 FTIR and tested as a heterogeneous catalyst in the dehydration of d-xylose and xylan to produce furfural as
Received in revised form 30 May 2017 target compound. With the use of the catalyst the appropriate reaction temperature and time in a mixture
Accepted 31 May 2017
of water-CPME (1:3, v/v) were 190 ◦ C and 60 min, in both cases. In these conditions, dehydration starting
Available online 7 June 2017
from d-xylose and xylan gave 60 and 42% of furfural yields, respectively. The influence of temperature,
the residence time and the substrate to catalyst ratios on the d-xylose conversion and furfural selectivity
Keywords:
were investigated.
Furfural
Biphasic system © 2017 Elsevier B.V. All rights reserved.
Heterogeneous catalyst
Microwave-assisted dehydration
Sulfonated char

1. Introduction ern organic solvents such as 2-methyltetrahydrofuran (MeTHF) [6],


tetrahydrofuran (THF) and also for the elaboration of oil additives.
Due to the severe decreasing of petroleum resources, many Industrially, furfural is produced by batch and continuous acid
academic and industrial researchers focuse on the development hydrolysis of lignocellulosic biomass in a narrow range of temper-
of alternative pathway to access to the twelve top-value added ature found between 170 and 185 ◦ C under conventional heating.
chemicals issued from the transformation of agricultural wastes Immersed in water, the woody precursor is converted using strong
and woody biomass. Thus, the catalytic conversion of cellulose and mineral acid essentially HCl [7,8], H2 SO4 [9], and H3 PO4 [10]. Unfor-
hemicellulose rich of lignocellulosic biomass is efficient to generate tunately, these homogeneous acid catalysts are the main cause of
various organic platform molecules such as furan derivatives and equipment corrosion, high waste streams and increasing treatment
many other compounds. Because hemicellulose and xylan are both costs.
polysaccharides, mainly composed of d-xylose sub-units, hydroly- To recover the furfural, only distillation was efficient despite the
sis of the polymer backbone is source of the pentose release in the formation of a black carbonaceous side-product called humin, gen-
reaction medium [1]. Especially in acidic condition, this hydrolysis erally observed in an aqueous monophasic system. Humin is overall
is generally followed by the rapid dehydration of the d-xylose into generated by the cross-polymerization between the just formed
furfural [2] by loss of three molecules of water. furfural and free d-xylose present in solution. Recently, instead of
Furfural or “2-furancarboxaldehyde” is useful for the prepara- the homogeneous acids, a large variety of heterogeneous catalysts
tion of many small commercial available intermediates such as such as sulfonated materials [11–20], HY zeolite [21] or heteropoly-
furoic acid [3], furfuryl alcohol [4] or 2-furonitrile [5] employed acids [22] have demonstrated their potential to dehydrate d-xylose.
for the synthesis of polymeric materials or bioactive compounds. Humin formation is at the origin of furfural selectivity drop.
Furfural is also the starting material used for the synthesis of mod- Two recent strategies are possible to remedy to this problem.
First, furfural could be isolated from the previous undesired cross
polymerization and from its resinification by adding a non water
miscible organic solvent such as cyclopentylmethyl ether (CPME)
∗ Corresponding author. [23] or methylisobutylketone (MIBK). Our group has already stud-
E-mail address: christophe.len@utc.fr (C. Len). ied different water-CPME biphasic systems including non miscible

http://dx.doi.org/10.1016/j.mcat.2017.05.031
2468-8231/© 2017 Elsevier B.V. All rights reserved.
168 Y. Wang et al. / Molecular Catalysis 438 (2017) 167–172

sulfonated materials such as Nafion NR50 [17] or sulfonated Brunauer-Emmett-Teller (BET) nitrogen adsorption measure-
sporopollenin, a very resistant chemically modified biopolymer ments were carried out at 77 K using an Micrometrics Gemini
[18] in presence of NaCl excess. In both cases, catalysts are active 236 volumetric adsorption analyzer. The samples were outgassed
at temperature found between 170 and 190 ◦ C, and the salt pro- for 12 h at 200 ◦ C under vacuum (po 10−2 Pa) and subsequently
moted the transfer of the furfural into the organic layer via a salt-out analysed. The linear part of the BET equation (relative pressure
effect. In order to accelerate the d-xylose dehydration, in paral- between 0.05 and 0.30) was used for the determination of the
lel to the furfural selectivity enhancement, many studies consider specific surface area.
microwave irradiation as a great technique to reach this objective
and our results obtained in biphasic systems corroborate this idea. 2.4. General procedure for the synthesis of furfural in
In this work, a new sulfonated carbon-based catalyst has suc- water-CPME as biphasic media from d-xylose and xylan
cessfully demonstrated its capacity to convert d-xylose into furfural
in a water-CPME biphasic system under different conditions of time In a typical experiment, 0.15 g (1.0 mmol) of D(+)-xylose and
and temperature. The catalyst reusability was proven and its mor- 15.0 mg of sulfonated char based catalyst (SCh) were introduced in
phology was studied by SEM, EDX and FTIR at different stages of a 10 mL vial closed with a septum, followed by the addition of water
this study. (1 mL) and CPME (3 mL). Biphasic batch reactions were carried out
by way of microwave heating apparatus (AntonPaar Monowave
2. Materials and methods 300) and stirred under magnetic stirring (600 rpm) for the desired
time. Temperature in the reaction vessel was measured by means
2.1. Materials of an IR sensor and the vial was pressurized due to the normal vapor
pressure of the solution at defined reaction temperature. At the end
Furfural (99% minimum) and D(+)-xylose were purchased from of the reaction, each sample was diluted with acetonitrile to obtain
Acros Organic. Cyclopentylmethyl ether was supplied by Sigma 100 mL of diluted solution and filtered prior to analysis through a
Aldrich. Xylan (from benchwood >90% (HPLC)) was obtained from syringe filter (PTFE, 0.45 ␮m, VWR).
Sigma Life Sciences. The water used in all experiments was a Milli- Identical procedure was done starting from 0.15 g of xylan.
pore Milli-Q grade.
2.5. Product analysis
2.2. Catalyst preparation
Each sample of the reaction mixture was analyzed separately
Miscanthus x giganteus was the source of biomass for the pro- by means of a Shimadzu Prominence HPLC. D-Xylose was detected
duction of char by carbonization in the process of slow pyrolysis. with a low temperature evaporative light scattering detector
This form of pyrolysis utilizes a long vapor residence time in order (ELSD-LTII) and products were detected with a UV–vis detec-
to maximize char yields. Char was produced in tubular furnace at tor (SPD-M20A) at a wavelength of 275 nm. The column used
temperature of 600 ◦ C. To achieve anaerobic conditions, a nitro- was a Grace C18 column (250 × 4.6 mm 5 ␮m). The mobile phase
gen pressure of 1.6 atm was maintained in the reactor. Once the was MeOH–H2 O (90:10) solution flowing at rate of 0.5 mL min−1 .
char were collected, they were ground and sieved between 180 and The column oven was set at 40 ◦ C. Regularly, the calibration was
600 mm and underwent sulfonation. In order to produce the cat- checked out in order to avoid eventual experimental errors associ-
alytic acid materials, 20.0 g of char was placed in a 500 mL flask with ated with all measurements reported below.
200 mL of concentrated (95–97%) sulfuric acid. The mixtures were d-Xylose conversion (X), furfural yield (Yi) and furfural selectiv-
heated at 150 ◦ C for 24 h on magnetic hotplate stirrers, fitted with ity (Si) were calculated by the following equations:
J-type PTFE coated thermocouples. The slurries were then filtered
and washed several times with distilled water until the pH became (Initial xylose amount (mol) − Final xylose amount (mol))
X= × 100
neutral, and sulfonated char based catalyst (SCh) was obtained [24]. Initial xylose amount (mol)

2.3. Catalyst characterization Final furfural amount (mol)


Yi = × 100
Initial xylose amount (mol)
Carbon, hydrogen, nitrogen and sulphur were analysed with an
Elemental Vario EL Cube. The analysis was conducted according Furfural yield
to BS EN 15104 (2011). Sulphanilamide was used as a standard. Si = × 100
Conversion of xylose
Oxygen was calculated by difference, subtracting the percentage of
C, H, N, S and ash from 100%.
SEM (Scanning Electron Microscopy)-EDX (Energy Dispersive 3. Results and discussion
X-ray Diffraction) analysis of native and sulfonated char (SCh) at
different stage of the experiments was performed on a Hitachi 3.1. Preparation, structure and morphology analysis of the
SU-70 equipped with a microanalysis detector for EDX. SEM micro- catalyst
graphs acquired in secondary electron mode were obtained at low
vacuum, 20 kV of accelerating voltage with a 10 mm working dis- Elemental composition and ash content for Miscanthus based
tance. EDX spectra were collected at 30◦ angle, 15 kV accelerating. char made at 600 ◦ C and for the sulfonated char are presented in
INCA software was used to examine the energy spectrum in order Table 1.
to determine the abundance of specific elements and to give an The direct sulfonation of char by the concentrated sulfuric acid
overall view of the composition of the samples. leads to the production of its sulfonated counterparts proved by
FTIR spectra were recorded on a Agilent Technologies Cary 630 both increases sulfur atom and oxygen atom percentages recorded
instrument in the range of 4000–650 cm−1 . during the elementar analysis. BET experiment showed that sur-
XRD experiments were recorded on a PanAnalytic/Philips X’pert face for the char was only 7 m2 g−1 , and sulfonation process led to
MRD diffractometer (40 kV, 35 mA) using Cu Ka (␭= 0.15418 nm) increase of the surface area to 317 m2 g−1 , probably due to the harsh
radiation. Scans were performed over a 2␪ range from 10 to 80, at conditions capable of breaking exfolating the successive carbon
step size of 0.018◦ with a counting time per step of 5 s. layers.
Y. Wang et al. / Molecular Catalysis 438 (2017) 167–172 169

Table 1
Surface area, elemental composition and ash content of char and the catalyst.

Sample BET N C H S O Ash Atomic ratio

m2 g−1 Wt% H/C O/C S/C

char 7 79.7 79.7 1.91 0.05 6.5 11.0 0.024 0.082 0.001
SCh 317 62.3 62.3 1.72 2.95 31.1 1.34 0.028 0.500 0.047

Fig. 1. SEM images of Miscanthus-based material carbonized at 600 ◦ C (a) and then sulfonated (b).

Visual observation was performed by SEM to assess the mate- also a great influence on the produced furfural yield. To confirm
rials morphology throughout the production process. There was this hypothesis, the influence of water-CPME volume ratio on the
no visible difference between untreated chars and the SCh catalyst. production of furfural was determined at 190 ◦ C for 60 min under
Catalyst material presented the same inner, well developed, honey- microwave irradiation using different volume ratios, and the ini-
comb structure with a highly complex network of pores, channels tial SCh loading was set as 10 wt% of d-xylose (1.0 mmol). d-Xylose
and fibrous ridged surfaces (Fig. 1). conversion and furfural yield reached 87% and 37% respectively
All materials collected at each stage of the catalyst prepara- in sole water (Fig. 2). In a biphasic system, d-xylose conversion,
tion were also analysed using the energy-dispersive (EDX) detector furfural yield and selectivity increased with the addition of CPME
integrated into the SEM. Spectra of the catalyst showed the pres- to the detriment of water content, and the maximum values of
ence of sulfur due to sulfonation in association withslight traces of 60% and 62% were reached for the furfural yield and its selectiv-
silica, (see ESI, Fig. S1). ity respectively for a system composed of 1 mL water and 3 mL of
Besides, FT-IR measurements were carried out to obtain an CPME. By increasing the volume of CPME, after the reaction, the
understanding of chemical changes of the surface during the cat- d-xylose almost disappeared, however, furfural yield and selectiv-
alyst production. When carbonization occured, the biomass is ity slightly decreased together. Simultaneously, more humin was
degraded and rearranged producing phenolic and polyaromatic observed after the reaction meaning that the second type of side-
rings. Some of these functional groups such as aromatic rings will reactions or resinification of furfural itself was promoted in the
then undergo sulfonation reactions with sulfuric acid introduc- organic layer. It is also possible to imagine that furfural formed in
ing sulfonic groups on the surface of the catalyst. Both spectra water phase can react with d-xylose to form hydroxyl groups hold-
of Miscanthus based materials (see ESI Fig. S2) show the pres- ing carbonaceous material and water-soluble side-products in the
ence of aromatic combination band at around 2117 cm−1 persisting remaining small volume of water. These experiments showed that
throughout sulfonation. The C O stretching found at 1644 cm−1 is the best water-CPME ratio (1:3, v/v) was kept for this catalyst and
attributable to the carbonyl group, while the peak at 1600 cm−1 this ratio was set up for the dehydration of d-xylose to furfural in
is characteristic of the C C aromatic ring stretch of large polyaro- the whole study.
matic structures. A signal located at 917 cm−1 also confirmed the
existence of simple C H bonds inside the material. Vibration bands 3.3. Effect of reaction temperature for the production of furfural
of SCh observed at 1169 and 1022 cm−1 could be assigned to the
O S O stretching vibration group and indicated the presence of It is well-known that reaction temperature has effectively a
SO3 H groups on the catalyst [25]. crucial effect on the dehydration process. Therefore, reaction tem-
Then, XRD patterns of Miscanthus-based char was rich in crys- perature ranged from 150 ◦ C to 210 ◦ C for 60 min under microwave
talline structures as presented in Fig. S3 of the ESI proven by broad C irradiation in batch was investigated (Fig. 3). At a temperature
(002) diffraction peaks in native char sample between 2␪ = 15–30◦ lower than 180 ◦ C, d-xylose conversion can become superior to 77%,
and another broad and weak C (101) diffraction peaks between however the furfural yield and selectivity were both limited. In fact,
2␪ = 40–50◦ indicating the presence of an amorphous solid carbon at the lower temperatures, the catalyst has the ability to dehydrate
structure. However, through the sulfonation process, any crys- partially the d-xylose into various intermediates, thus the reaction
talline structures remained intact and the SCh became amorphous is surely not complete. With the increase of reaction temperature
confirmed by the suppression of the previous signal. from 150 ◦ C to 190 ◦ C, the furfural yield, the selectivity and the d-
xylose conversion increased to 60%, 62% and 97% respectively. At
3.2. Effect of volume ratio for the production of furfural 210 ◦ C, furfural yield and selectivity were similar (59%) but at higher
temperature furfural has a poor stability and will quickly react to
CPME is considered as a promising green co-solvent to extract furnish the dark brown humin. In our case, the more relevant reac-
furfural formed in aqueous phase in order to prevent this mate- tion temperature was detected at 190 ◦ C. This condition afforded
rial of its undesired degradation. Thus, the volume of CPME has the best furfural yield and selectivity.
170 Y. Wang et al. / Molecular Catalysis 438 (2017) 167–172

Fig. 2. Effect of water-CPME volume ratio on dehydration of d-xylose to furfural. Reaction conditions: d-xylose (1.0 mmol), SCh (15.0 mg), water-CPME, MW, 190 ◦ C, 60 min.

Fig. 3. Effect of reaction temperature on dehydration of d-xylose to furfural. Reaction conditions: d-xylose (1.0 mmol), SCh (15.0 mg), water-CPME (1:3, mL/mL), MW, 60 min.

invariable. Here, the maximum of furfural yield and selectivity at


190 ◦ C gave the values of 60% and 62%, respectively when the reac-
tion was carried out for 60 min. It should be important to notice
that furfural content decreased by 7% without catalyst at 190 ◦ C
for 60 min even taken in the best current water-CPME (1:3, v/v)
biphasic system, due to the induced acidic character of water in sub-
critical condition, mainly caused by the resinification of produced
furfural.

3.5. Effect of catalyst and substrate loading for the production of


furfural

The effect of the initial catalyst (from 0 to 25 mg) and d-xylose


(varying from 0.05 g to 0.25 g) loadings in a mixture of water-CPME

(1:3, v/v) was also studied at 190 C for 60 min (Fig. 5). The concen-
tration of catalyst can significantly affect d-xylose conversion and
furfural yield (Fig. 5a). d-Xylose cannot be sufficiently consumed
when catalyst concentration is lower than 10% but higher cata-
Fig. 4. Effect of reaction time on dehydration of d-xylose to furfural. Reaction con- lyst concentration did not cause higher furfural yield, the medium
ditions: d-xylose (1.0 mmol), SCh (15.0 mg), water-CPME (1:3, mL/mL), MW, 190 ◦ C. become too acidic. To investigate the behavior of the catalyst, one
experiment with furfural as substrate was tested at 190 ◦ C for
60 min in presence of 15.0 mg of the catalyst immersed in the most
3.4. Effect of reaction time for the production of furfural suitable volume ratio. The results proved that 15% of furfural con-
tent decreased, but only 7% (two times lower) of the compound
The variation of residence time from 10 to 60 min at 190 ◦ C disappeared when the catalyst was not added to the medium. This
was investigated in order to identify the most favorable residence result indicated that an excess of catalyst loading could be the main
time. d-Xylose conversion, furfural yield and selectivity gradually cause of furfural loss. The degradation occurring here could be also
increased with the time extension (Fig. 4). d-Xylose could be con- explained by the nature and three-dimensional structure of this
verted in a very short time. Only 10 min are enough to obtain 28% of catalyst. Once the furfural is adsorbed on the surface, it could eas-
selectivity which rapidly increased after 30 min, and it tended to be ily react to form the humin on the catalyst periphery. D-Xylose
Y. Wang et al. / Molecular Catalysis 438 (2017) 167–172 171

Fig. 5. Effect of catalyst (a) and d-xylose (b) loadings on dehydration of d-xylose to furfural. Reaction conditions: d-xylose (0.15 g for a) and SCh loading (0.02 g for b),
water-CPME (1:3, mL/mL), MW, 190 ◦ C, 60 min.

Fig. 6. Reusability of the catalyst. Reaction conditions: d-xylose (0.33 mmol), SCh (20 mg), water-CPME (1:3, mL/mL), MW, 190 ◦ C, 60 min.

conversion was quite complete when its loading was inferior to


150 mg (Fig. 5b). In that situation, furfural yield and selectivity
slightly increased to 64% in parallel to the d-xylose loading drop to
50 mg. Therefore, the optimized initial loading was set up at 20 mg
of catalyst and 50 mg (0.33 mmol) of d-xylose respectively.

3.6. Effect of catalyst recovery for the production of furfural

Recycling performance is a significant index to evaluate the per-


formance of a catalyst. In our case, ten cycles of experiments were
studied under optimized conditions to estimate the properties of
recovered catalyst (as shown in Fig. 6). In this work, the organic
CPME phase containing the furfural was separated from the reac-
tion medium after each catalytic run. Then, d-xylose (50.0 mg) and
fresh CPME (3 mL) were added to the recycled aqueous phase, con-
taining the original catalyst without being subjected to further
treatment before starting the next cycle. Interestingly, after one
cycle, only 1.2% of sulfate traces were detected by Inductivity Cou-
pled Plasma (ICP) analysis in the reaction medium, especially in the Fig. 7. Effect of reaction temperature on furfural and d-xylose yield from xylan.
aqueous layer. Up to ten cycles, furfural yield containing in CPME Reaction conditions: Xylan (0.33 mmol based on d-xylose units), SCh (20 mg), water-
phase gave an average value of 60% for all cycles, which means that CPME (1:3, mL/mL), MW, 60 min.
the catalyst has stable and excellent capacity to dehydrate d-xylose
into furfural in biphasic system under microwave irradiation.
(0.33 mmol based on d-xylose units), sulfonated catalyst (20 mg),
water-CPME (1:3, mL/mL), MW, 190 ◦ C, 60 min, and furfural yield
3.7. Furfural production from xylan reached 37%. This poorer furfural yield was attributed to the two-
steps reaction mechanism: a) hydrolysis of xylan into d-xylose
To extend this study into lignocellulosic biomass for the pro- (1 step in several steps); b) dehydration of d-xylose into furfural
duction of furfural, experiments with xylan as a suitable model (3 steps). For this reason, variation of reaction temperature from
were carried out under the optimized method as follow: xylan 160 ◦ C to 210 ◦ C for 60 min (Fig. 7) and reaction time from 10 min
172 Y. Wang et al. / Molecular Catalysis 438 (2017) 167–172

Acknowledgements

YW would like to thank the China Scholarship Council (CSC) for


the financial support. This work was performed, in partnership with
the COST Action FP 1306 Valorisation of ligno-cellulosic biomass
strams for sustainable production of chemicals, materials and fuels
using low environmental impact technologies.

Appendix A. Supplementary data

Supplementary data associated with this article can be found,


in the online version, at http://dx.doi.org/10.1016/j.mcat.2017.05.
031.

References

[1] S.R. Kamireddy, E.I. Kozliak, M. Tucker, Y. Ji, Int. Agric. Biol. Eng. 7 (2014)
86–98.
[2] R. Mariscal, P. Maireles-Torres, M. Ojeda, I. Sadaba, M. Lopez Granados, Energy
Fig. 8. Effect of reaction time on furfural and d-xylose yield from xylan. Reaction Environ. Sci. 9 (2016) 1144–1189.
[3] S.K. Nalli, C.S. Vimal, B. Suddhasatwa, Indian Chem. Eng. 55 (2013) 153–164.
conditions: xylan (0.33 mmol based on d-xylose units), SCh (20 mg), water-CPME
[4] B. Zeynizadeh, D. Setamdideh, J. Chin. Chem. Soc. 52 (2005) 1179–1184.
(1:3, mL/mL), MW, 60 min.
[5] P. Data, A. Kurowska, S. Pluczyk, P. Zassowski, P. Pander, R. Jedryslak, M.
Czwartosz, L. Otulakowski, J. Suwinski, M. Lapkowski, A.P. Monkman, J. Phys.
to 90 min at 190 ◦ C (Fig. 7) have been developed to optimize fur- Chem. C 120 (2016) 2070–2078.
[6] N.S. Birudar, A.A. Hengne, S.N. Birajdar, R. Swami, C.V. Rode, Org. Process Res.
fural yield from xylan. At 170 ◦ C, 22% of d-xylose was yielded after Dev. 18 (2014) 1434–1442.
60 min under microwave irradiation, however, only 13% of furfural [7] G. Marcotullio, W. De Jong, Green Chem. 12 (2010) 1739–1746.
was produced, which means that the temperature is not sufficient [8] E.I. Fulmer, L.M. Christensen, R.H. Hixon, R.L. Foster, J. Phys. Chem. 4 (1935)
133–141.
for the formation of furfural but enough to hydrolyze xylan into [9] C. Rong, X. Ding, Y. Zhu, Y. Li, L. Wang, Y. Qu, X. Ma, Z. Wang, Carbohydr. Res.
monomers and oligomers sub-unit (Fig. 7). The higher furfural yield 350 (2012) 77–80.
of 37% was obtained at 190 ◦ C for 60 min, and 8% of d-xylose was [10] O. Yemis, G. Mazza, Bioresour. Technol. 102 (2011) 7371–7378.
[11] J. Takagaki, M. Ohara, S. Nishimura, K. Ebitani, Chem. Lett. 39 (2010) 838–840.
remained after reaction. Despite of reaction temperature increas- [12] A. Tellaria, A. Larreategui, J. Requies, M.B. Guenez, P.L. Arias, Bioresour.
ing to become superior to 200 ◦ C, furfural yield was not improved. Technol. 102 (2011) 7478–7485.
In fact, even if SCh expressed high porosity, due to its high poly- [13] C. Aellig, D. Scholz, P.Y. Dapsens, C. Mondelli, J. Perez Ramirez, J. Catal. Sci.
Technol. 5 (2015) 142–149.
merization degree of the biopolymer, the reaction was limited to
[14] P.P. Upare, D.W. Hwang, Y.K. Hwang, V.H. Lee, D.Y. Hong, J.S. Chang, Green
the accessible sulfonated surface of SCh and cannot occur inside the Chem. 17 (2015) 3310–3313.
cavities of the catalyst. However, furfural yield slightly decreased [15] J. Tureja, S. Nishimura, K. Ebitani, Bull. Chem. Soc. Jpn. 85 (2012) 275–281.
owing to the occurrence of side-reactions. At 190 ◦ C, furfural was [16] E. Lam, E. Hajid, A.C.W. Leung, J.H. Chong, K.A. Hahmoud, J.H.T. Luong,
ChemSusChem 4 (2011) 535–541.
quickly formed, simultaneously, d-xylose formed from xylan was [17] S. Le Guenic, D. Gergela, C. Ceballos, F. Delbecq, C. Len, Molecules 21 (2016)
rapidly consumed in 40 min, and a maximal furfural yield of 42% 1102–1112.
was obtained after heating for 80 min and then remained almost [18] Y. Wang, T. Len, Y. Huang, A. Diego-Taboada, A.N. Boa, C. Ceballos, F. Delbecq,
G. Mackenzie, C. Len, ACS Sustainable Chem. Eng. 5 (2017) 392–398.
constant for up to 90 min (Fig. 8). In regards to these results, the [19] P.K. Khatri, N. Karanwal, S. Kaul, S.L. Jain, Tetrahedron Lett. 56 (2015)
optimized formation of furfural from xylan was obtained at 190 ◦ C 1203–1206.
for 80 min upon microwave irradiation in the presence of 20 wt% [20] A. Deng, Q. Lin, Y. Yan, H. Li, J. Ren, C. Liu, R. Sun, Bioresour. Technol. 216
(2016) 754–760.
of catalyst in water-CPME (1:3, v/v) biphasic system. [21] R. Weingarten, G.A. Tompsett, W.C. Conner Jr., G.W. Huber, J. Catal. 279 (2011)
174–182.
4. Conclusion [22] S. Dias, S. Lima, M. Pillinger, A.A. Valente, Carbohydr. Res. 341 (2006)
2946–2953.
[23] S. Le Guenic, F. Delbecq, C. Ceballos, C. Len, J. Mol. Cat. A 410 (2015) 1–7.
In summary, this work reports a new efficient process using [24] F. Severini, J.J. Leahy, W. Kwapinski, Waste Biomass Valorization 7 (2016)
heterogeneous sulfonated catalyst coming from Miscanthus x 1–11.
[25] B.V.S.K. Rao, K. Chandra Mouli, N. Rambabu, A.K. Dalai, R.B.N. Prasad, Catal.
giganteus for the production of furfural from d-xylose in 60% yield. Commun. 14 (2011) 20–26.
This new promising bio-based acidic catalyst was fully character-
ized using elemental analysis, SEM, EDX and FTIR spectroscopy. Its
activity and recyclability in a water-CPME biphasic system under
microwave irradiation in batch put it as a potent candidates to
replace the conventional homogeneous acid catalysts such as sul-
furic acid. This material is currently studied in our laboratory for
realizing other chemical transformations.

You might also like