You are on page 1of 13

Journal of Catalysis 305 (2013) 204216

Contents lists available at SciVerse ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Synthesis and characterization of acidic ordered mesoporous


organosilica SBA-15: Application to the hydrolysis of cellobiose
and insight into the stability of the acidic functions
Mariam Karaki a,b,c, Ali Karout a, Joumana Toufaily a, Franck Rataboul c, Nadine Essayem c,,
Bndicte Lebeau b
a
Laboratory of Materials, Catalysis, Environment and Analytical Methods, Faculty of Sciences, Lebanese University, Campus Rac Hariri, Beirut, Lebanon
b
Equipe Matriaux Porosit Contrle, IS2M, LRC 7228 CNRS, UHA, 3 bis rue Alfred Werner, 68093 Mulhouse Cedex, France
c
Institut de Recherche sur la Catalyse et lEnvironnement de Lyon, IRCELYON, CNRS, University of Lyon 1, 2 Avenue Albert Einstein, 69626 Villeurbanne Cedex, France

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

Article history: The present work aimed to investigate the potential of acidic ordered mesoporous organosilica SBA-15
Received 26 February 2013 with a controlled local environment of the acid sites for applications in acid-catalyzed reactions in hot
Revised 25 April 2013 water, such as cellobiose hydrolysis. The hybrid silica was prepared by condensation of 1,4-bis(triethoxy-
Accepted 29 April 2013
silyl)benzene. The material was sulfonated using chlorosulfonic acid or 3-mercaptopropyltrimethoxysi-
Available online 19 June 2013
lane and further oxidized with H2O2 to give Brnsted solid acids, which were fully characterized. Their
acidic properties were studied by calorimetry of NH3 adsorption and in the model reaction of gas-phase
Keywords:
isopropanol dehydration and compared with the reference acidic sulfonated resin, Amberlyst 15. The
Sulfonated periodic mesoporous
organosilicas
introduction of mercaptopropyl organic arms, oxidized by H2O2, did not change the structure of the mate-
Cellobiose hydrolysis rial, while sulfonation with chlorosulfonic acid led to a less organized solid. In both cases, calorimetry of
Isopropanol dehydration NH3 adsorption has evidenced the heterogeneity of the acid strength distribution, suggesting the pres-
Solid hybrid acid catalysts ence of distinct sites of sulfonation, contrary to our expectation. This was supported by XPS analysis.
Water-tolerant solid acids For gas-phase isopropanol dehydration, the solids sulfonated with chlorosulfonic acid exhibited activity
equivalent to that of the sulfonated resin, Amberlyst 15, but were less stable due to sulfur species release,
assumed to be sulfonated silanols. The acidic organosilicas SBA-15 obtained via H2O2 oxidation of the
mercaptopropyl group are less acidic catalysts, showing low activity for gas-phase isopropanol dehydra-
tion. However, no sulfur release was observed. The hybrid solid acids were evaluated in a reaction of great
interest in the context of the biomass valorization: cellobiose hydrolysis in hot water. The solids were
active at 150 C; however, sulfur leaching analysis showed that the reaction proceeds mainly homoge-
neously, especially for the material acidied with chlorosulfonic acid. A hot washing pretreatment
applied to the catalysts containing the sulfonated propyl groups led to a decrease in their hydrolysis
activity, but along with an increase in stability, allowing recycling.
2013 Elsevier Inc. All rights reserved.

1. Introduction types of solid acid catalysts, those bearing sulfonated functions


have been introduced with the ultimate goal of their employment
Catalysis with solid acids has been the subject of intense in processes currently using liquid sulfuric acid. The hydrolysis of
research activities for a long time, especially in the petrochemical cellulose into glucose is a reaction of this last kind. These processes
industry [1]. Nowadays, solid acid catalysts are increasingly stud- are known not to be selective in glucose [10] and suffer from the
ied in other areas and particularly in those related to green chem- usual drawbacks associated with the use of homogeneous strong
istry and bioresources valorization, such as biodiesel synthesis acids: corrosion and production of large volume of acidic wastewa-
[2,3] and now polysaccharide and carbohydrate transformation ter. Nowadays, the necessity of upgrading renewable carbon re-
[48]. Zeolites and related materials still represent the most stud- sources in a sustainable way makes the substitution of solid
ied systems; however, for applications in liquid water, their lack of acids for liquid mineral acids of high interest. Sulfonated materials
hydrothermal stability could preclude their use [7,9]. Among other are the logical candidates due to their acid sites chemical structure.
Many studies have already reported the preparation and the use of
sulfonated catalysts for direct cellulose hydrolysis [11,12]; the
Corresponding author.
E-mail address: nadine.essayem@ircelyon.univ-lyon1.fr (N. Essayem).

0021-9517/$ - see front matter 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.jcat.2013.04.024
M. Karaki et al. / Journal of Catalysis 305 (2013) 204216 205

Scheme 1. Hydrolysis of cellobiose into glucose.

model reaction consisting of cellobiose hydrolysis is mainly stud- is also expected that the sulfonated group covalently bonded to the
ied (Scheme 1). organic entities would be more stable in water as well as in the gas
Early studies by Moreau and co-workers involved the use of phase. Note that to the best of our knowledge, this kind of silica has
zeolites [13]. Later, other systems were investigated [1419], prin- not been previously applied to the hydrolysis of cellobiose, but
cipally sulfonated catalysts, mainly based on silica [2024]. Shanks only to the hydrolysis of starch and sucrose [26].
and co-workers reported the use of sulfonated mesoporous silicas For that, we prepared SBA-15 mesoporous materials with or-
such as SBA-15. They observed a cellobiose conversion of 90% after ganic groups incorporated within the silica framework [2729].
30 min at 175 C, but the glucose yield was not clearly given, and Three different ways of sulfonation were investigated, including
no catalyst stability tests were performed [20,21]. Wang and co- an in situ formation of the acidic function, and two postsynthesis
workers introduced CoFe2O4silica nanoparticles containing sur- methods [28,30,31]. Here, we compared the inuence of the mode
face sulfonic groups for this reaction. They obtained conversions of sulfonation on the characteristics of the materials and on the
up to 70% with 45% glucose yield after 1 h at 175 C. Recycling catalytic efciency for cellobiose hydrolysis in aqueous liquid
experiments showed a signicant loss of activity due to the loss phase and for gas-phase isopropanol dehydration with a focus on
of sulfonic groups during the reaction [24]. Similar materials were the stability of the materials during the reaction, especially the
used by Ebitani and co-workers showing a glucose yield of 60% acidic sulfonated function.
after 9 h (88% after 18 h) at 120 C, well above the performances
of Amberlyst-15 resin. Recyclability was obtained three times
without loss of activity [23]. Degirmenci et al. reported the use of 2. Experimental
sulfonated zirconia-SBA-15 materials giving up to 80% of conver-
sion with 60% glucose yield after 1 h 30 min at 160 C, which is 2.1. Materials synthesis
greater than with conventional sulfated zirconia. No recyclability
studies were performed, but the authors evoked the possibility of 2.1.1. Chemicals
sulfur leaching during the reaction [22]. 1,4-Bis(triethoxysilyl)benzene (BTEB) and 3-mercaptopropyltri-
From the literature data, one can state that the potential of sul- methoxysilane (MPTMS) were purchased from Gelest. Pluronic
fonatedsilica-based catalysts for cellobiose hydrolysis is evident. P123, chlorosulfonic acid, and H2O2 (30 wt%) were obtained from
However, little information on their effective stability in hot water Aldrich. HCl (37 wt%) was obtained from Riedel de Haen. All chem-
is known or reported. In addition, one cannot ensure that they al- icals were used as received.
ways work as pure heterogeneous catalysts. Indeed, the problem of
the stability of sulfonated silica-based catalysts in liquid hot water 2.1.2. Synthesis of the mesoporous ordered silica C-SBA-15
is the worst situation due to the polarity, the solvating power of In a polypropylene bottle, P123 (3.9 g) was dissolved in 0.87 ml
this medium, and accordingly the associated inhibition of the acid of HCl and 140 ml of H2O. The solution was heated to 40 C and
sites and even their leaching. Also, hot water can be considered as a maintained under vigorous stirring for 3 h. Then, BTEB (4.0 g)
reactive solvent because of the improved water self-protolysis fa- was added. A white precipitate was obtained after the mixture
vored by the temperature increase [25] leading to higher concen- was stirred for 3 h at 40 C and was then aged under static condi-
trations of hydroxonium ions, which might favor secondary tions for 24 h at 100 C. The solid product was recovered by ltra-
homogeneously acid-catalyzed reactions. tion, washed with ethanol, and dried at 70 C for 24 h. The
For gas-phase alcohol dehydration, another important class of template was removed by treating the material (1.0 g) in reuxing
reactions in the general context of biomass valorization, the ques- ethanol (200 ml) and HCl (4 ml) overnight, to give C-SBA-15.
tion of catalyst deactivation is a priori less problematic, since the
solvation of the active function in a polar liquid medium will not
2.1.3. Synthesis of grafted-C-SBA-15-SO3H
intervene. The deactivation process will mainly concern the ther-
C-SBA-15 (0.5 g) was added to a solution of MPTMS (1.0 g) in
mal stability of the active functions and their possible poisoning.
toluene (25 ml). The suspension was reuxing at 110 C for 8 h.
In this paper, we report our study on the preparation, character-
The resulting solid was recovered by ltration, washed with tolu-
ization, and use of sulfonated mesoporous organosilicas for cellobi-
ene, and dried at 70 C for 24 h to give grafted-Ph-SBA-15-SH.
ose hydrolysis in water and also for gas-phase isopropanol
The oxidation was carried out by treatment of 0.5 g of the solid
dehydration. Our goals were to evaluate the potential of these
in H2O2 (50 g) at room temperature overnight.
types of materials for acid-catalyzed transformations of carbohy-
drates or alcohol, in water phase or in gas phase, respectively.
Rather than the silica materials cited above, we studied periodic 2.1.4. Synthesis of C-SBA-15-SO3H
ordered mesoporous organosilica materials (PMOS). Indeed, the C-SBA-15 (1.0 g) was added to a solution of chlorosulfonic acid
mesoporous structure of this kind of catalysts would be favorable (0.3 ml) in dichloromethane (75 ml), and the suspension was stir-
for converting bulky reactants, and moreover, the introduction of red at room temperature overnight. The solid was ltered off and
hydrophobic moieties into the walls of the material could improve washed with dichloromethane (30 ml), acetone (30 ml), and meth-
the stability of the catalyst by limiting the deactivation by water. It yltertiobutyl ether (10 ml) before drying at 70 C for 24 h.
206 M. Karaki et al. / Journal of Catalysis 305 (2013) 204216

2.1.5. Synthesis of in situ-C-SBA-15-SO3H 2.3. Characterization of the acidic properties


The same procedure as for C-SBA-15 was followed, except that
H2O2 (3.35 g) and MPTMS (0.51 g) were added to the starting solu- 2.3.1. Calorimetry of NH3 adsorption
tion just after BTEB. The acidic strength of the solids was characterized by NH3
adsorption at 80 C, using a Tian Calvet calorimeter coupled with
a volumetric ramp. The catalyst samples (0.05 g) were rst evacu-
2.2. Material characterizations: structure, morphology, and texture ated under vacuum (105 Torr) at 120 C for 2 h and then placed
into the calorimeter up to the stabilization of the temperature
2.2.1. Powder X-ray diffraction (one night). Then, the solid was contacted with small amounts of
Powder X-ray diffraction was performed on a PANalytical XPert gas up to equilibrium and the differential enthalpy of adsorption
PRO diffractometer with Cu Ka radiation (k = 1.5418 ). Measure- was recorded, together with the amount of adsorbed ammonia.
ments were achieved for 2h values in the 110 range with a step
size of 0.010 and a step time of 4 s. The unit cell parameters ahex 2.3.2. Gas-phase dehydration of isopropanol (IPA)
p
was determined from the d100 (ahex = 2d100/ 3) reection for Gas-phase dehydration of isopropanol (IPA) was carried out in a
SBA-15-type materials. ow-type reactor with a xed catalyst bed operating at atmo-
spheric pressure. The catalyst (20100 mg) was introduced into a
Pyrex glass tube reactor and pretreated under owing nitrogen
2.2.2. 1H13C solid-state CPMAS NMR spectra for 2 h at 120 C. The catalytic reaction was carried out at 120 C
1
H13C solid-state CPMAS NMR spectra were recorded on an using a total ow rate of 5 l h1 of IPA (9 Torr) diluted in nitrogen
Avance II-400 WB (Bruker) spectrometer (B0 = 9.4 T) operating at as a carrier gas. The reaction products are analyzed on line by
a Larmor frequency of 100.2 MHz with a p/2 1H pulse duration of means of a FID-GC for propene, di-isopropyl ether, and IPA.
5.2 ls, a contact time of 1.5 ms, a recycle delay of 4 s, and a spin-
ning frequency of 10 kHz. 1H decoupled 29Si solid-state MAS
2.4. Cellobiose hydrolysis in water
NMR spectra were recorded on an Avance II-300 WB (Bruker) spec-
trometer operating at a Larmor frequency of 59.62 MHz with a
Cellobiose (0.20 g, 0.6 mmol), the catalyst (0.02 g), and H2O
pulse duration of 2 ls corresponding to a ip angle of p/6, a recycle
(20 ml) were introduced into a batch-type autoclave reactor
delay of 80 s, an acquisition duration of 24 h, and a spinning fre-
equipped with a magnetic stirrer. Then, the system was purged
quency of 4 kHz.
with helium. After the reaction, the solid catalyst was separated
by ltration, and the liquid phase was analyzed on a Shimadzu
HPLC (COREGEL 87C column, 0.5 ml min1 of water, 80 C, RID).
2.2.3. SEM
Yields are molar yields and were calculated as the molar ratio of
The morphology of the particles was characterized by scanning
the product i and the initial cellobiose, corrected by the number of
electron microscopy (SEM) using a Philips XL30 microscope oper-
carbon atoms:
ating at a voltage of 9 kV. Before analysis, samples were coated
with gold (thickness about 10 nm) by cathodic sputtering. yield % ni =ninitialcellobiose  nCi =12  100;

nCi is the number of carbon atoms in the product i, and ni is the


2.2.4. Nitrogen adsorption/desorption measurements number of moles of the product i.
Nitrogen adsorption/desorption measurements were performed
on a Micromeritics TRISTAR at 196 C. Prior to the measurements, 3. Results and discussion
the samples were outgassed at 90 C overnight. The specic surface
area was calculated using the BrunauerEmmettTeller (BET) 3.1. Materials synthesis and characterization
method in the range of relative pressure from 0.03 to 0.2. The mes-
opore-size distribution was evaluated by the BarrettJoynerHal- 3.1.1. Materials synthesis
enda method from the desorption branch of the isotherm and In this study, we choose to work with SBA-15 type materials
using a cylindrical model. The average pore diameters were evalu- with C6H6 organic moieties incorporated into the walls. These
ated by the Broekhoffde Boer (BdB) method applied to the desorp- materials were prepared following a procedure suggested by Cho
tion branch. et al. [30] with the use of BTEB as silica source and Pluronic P123
triblock copolymer as structure-directing agent, to give the parent
ordered mesoporous material C-SBA-15.
2.2.5. XPS analysis Three types of sulfonation were realized. The rst two involved
XPS was performed on as-synthesized materials and on the postsynthesis methods:
materials recovered after isopropanol (IPA) dehydration. The XPS
analyses were carried out on a Kratos Ultra DLD spectrometer  reacting C-SBA-15 with MTPMS to produce grafted-C-SBA-15-
using monochromated Al Ka radiation (10 mA, 15 kV). All spectra SH followed by oxidation with H2O2 to give grafted-C-SBA-15-
were taken in the hybrid (combined electrostatic and magnetic SO3H;
lens) mode. All spectra were referenced to the C1s line at binding  reacting C-SBA-15 with chlorosulfonic acid to give C-SBA-15-
energy 284.6 eV, characteristic of ever-present adventitious carbon SO3H.
(CC and CH). This peak position was obtained after Shirley back-
ground subtraction and decomposition of the C1s peak envelope The last involved the co-condensation of BTEB with MPTMS in
using a GaussianLorentzian (70%30%) curve t. All other photo- the presence of H2O2 to give in situ-C-SBA-15-SO3H.
electron peaks were background-subtracted and curve t in the These material preparations with MTPMS as sulfonated reagent
same manner. Quantication was carried out with the VISION soft- are well known from the literature. However, the sulfonation of
ware supplied. The relative sensitivity factors (RSF) applied here periodic mesoporous silicas with chlorosulfonic acid was only re-
are inherent in this software and incorporate Wagner photoelec- ported for a couple of examples for organic-phase reactions
tron cross-sections and analyzer transmission correction. [32,33], while other reports concern the sulfonation of organic
M. Karaki et al. / Journal of Catalysis 305 (2013) 204216 207

100 reduced intensity are indicative of the lower degree of organization


of the material. For C-SBA-15-SO3H, the d100 peak is much broader,
and the d110 and d200 are hardly observable; this shows a poor de-
gree of organization. Therefore, in contrast to previous reports [32],
our present data indicate that sulfonation by chlorosulfonic acid
has disordered the organization, resulting in a lower ahex, as indi-
110
200 cated in Table 1.
a C-SBA-15 Fig. 2 shows nitrogen adsorption/desorption isotherms of the
materials. All isotherms are of type IV according to the IUPAC clas-
sication, with an H1 hysteresis loop as expected for such meso-
porous materials. The BJH pore diameter distribution is also
reported. The porous characteristics determined from these iso-
therms are reported in Table 1.
The grafted-C-SBA-15-SO3H material has textural characteris-
b C-SBA-15-SO3H tics (SBET, Vp, and DBdB) very similar to those of its parent material
C-SBA-15. This indicates that the grafting occurred at the outside
wall surface and pore entrance and that this grafting was quite lim-
ited, which is in agreement with the low sulfur content (Table 1).
For the C-SBA-15-SO3H material, a large decrease in the SBET, Vp,
and pore diameter is observed as a consequence of the disordering
c grafted-C-SBA-15-SO3H upon sulfonation already observed by XRD. However, this sample
possesses a higher sulfur amount than grafted-C-SBA-15-SO3H,
4.2 wt% against 0.8 wt%.
The formation of in situ-C-SBA-15-SO3H led to a material with a
lower SBET, 440 m2 g1, but a larger total pore volume, 0.62 cm3 -
g1. A quite large sulfur amount was also determined for this sam-
ple. The decrease in the specic area and the pore size of the in situ
d in situ -C-SBA-15-SO3H material can be explained by the presence of strong hydrophobic
interactions between the functional groups and the surfactant tails
0 1 2 3 4 in the precursor solution. Such interactions would draw the organ-
2 theta () ic precursors further into the micelles, leading to shrinkage of the
pores [37].
Fig. 1. XRD patterns of (a) C-SBA-15, (b) C-SBA-15-SO3H, (c) grafted-C-SBA-15- The SEM micrographs presented in Fig. 3 show that the postsyn-
SO3H, and (d) in situ-C-SBA-15-SO3H. thesis sulfonated species have a rod-type hexagonal morphology
similar to that of the parent material. However, in situ-C-SBA-15-
SO3H is mostly constituted of spherical particles of ca. 12 lm,
polymers [34,35]. Therefore, little is known of these sulfonated
as previously noted by Cho et al. [30]. This indicates that the pres-
materials, and we wanted here to assess their potential for reaction
ence of MPTMS during the synthesis process, even if the arrange-
under hydrothermal conditions as well as in the gas phase.
ment is unchanged, clearly had an inuence on particle
formation and aggregation, a phenomenon remaining difcult to
3.1.2. Structure, morphology, and texture explain.
XRD patterns of the materials are displayed in Fig. 1. They exhi- The 29Si MAS NMR spectra of the materials are reported in
bit three peaks that can be indexed as d100, d110, and d220 reec- Fig. 4. Whatever the sulfonation method, they all present the char-
tions. These data are consistent with a well-ordered 2-D acteristic resonances between 60 and 80 ppm of an organ-
hexagonal (P6 mm) pore arrangement and characteristics of SBA- osiloxane network [Tn = RSi(OSi)n(OH)3n, n = 13] from BTEB
15 type materials [36,30]. However, if the intensity of the XRD [38]. The absence of resonance in the 90/110 ppm range that
peaks of in situ-C-SBA-15-SO3H was similar to that for C-SBA-15- is characteristics of Qn units proves the exclusive presence of Si
SO3H, indicating that the introduction of mercaptopropyl groups, C bonds. For in situ-C-SBA-15-SO3H, signals are slightly broader
and its in situ H2O2 oxidation had no signicant inuence in the due to additional peaks, which are attributed to trifunctional silox-
arrangement of the material, this is not true for the two other ane units T 0n from sulfonate/thiol propylsilyl groups and over-
materials. For grafted-C-SBA-15-SO3H, the diffraction peaks of lapped with the Tn of BTEB. The presence of these T 0n resonances

Table 1
Physicochemical features of the sulfonated organic mesoporous silicas.

Sample SBETa (m2 g1) DBdBb () Vpc (cm3 g1) Pore sized () te () af () Sulfur contentg (wt%)
C-SBA-15 768 82 0.56 53 31 109
C-SBA-15-SO3H 406 68 0.47 49 36 101 4.2
Grafted-C-SBA-15-SO3H 742 78 0.58 52 34 108 0.8
In situ-C-SBA-15-SO3H 440 68 0.62 51 43 107 3.1
a
BET specic area.
b
Pore diameter determined by Broekhoffde Boer method from the desorption branch.
c
Total pore volume.
d
Calculated from the desorption branch using the BJH method.
e
Wall thickness (t = a  0.95DBdB).
f
Unit cell parameter.
g
Sulfur content obtained by elemental analysis.
208 M. Karaki et al. / Journal of Catalysis 305 (2013) 204216

dV/dD (cm.g-1-1)
0,06

0,04

Volume adsorbed (cm3.g-1)


0,02

0
20 40 60 80 100
Pore size ()

a b

c d
0
0 0.5 1
P/P0

Fig. 2. Nitrogen adsorption/desorption isotherms of (a) in situ-C-SBA-15-SO3H, (b) C-SBA-15, (c) C-SBA-15-SO3H, and (d) grafted-C-SBA-15-SO3H. Inset shows the BJH pore
size distributions obtained from the desorption branch of the isotherm.

a b

2 m 2 m

c d

2 m 2 m

Fig. 3. Scanning electron microscopy images of (a) in situ-C-SBA-15-SO3H, (b) C-SBA-15, (c) C-SBA-15-SO3H, and (d) grafted-C-SBA-15-SO3H.

is not clearly observed for grafted-C-SBA-15-SO3H, mostly due to


the low grafting rate as mentioned before. T2
13
C MAS NMR experiments were run to check the integrity of
T3
the organic moieties constituting the materials (Fig. 5). The most
intense signal, at 130 ppm, is attributed to the benzene ring, and T1
a C-SBA-15-SO3H
the other resonances observed at 1220 (1C, 2C) and 5560
(3C) ppm in grafted- and in situ-C-SBA-15-SO3H samples are attrib- T2, T3
uted to the carbons of Si1CH22CH23CH2SO3H. Note that these
T3
results are well in agreement with reported data [26,30]. The broad
resonance at 3238 ppm is characteristic of the carbon from propyl T1, T2
b grafted-C-SBA-15-SO3H
chains linked to SH and indicates the presence of unoxidized thiol T2, T3
groups (Si1CH22CH23CH2SH). The spectrum of C-SBA-15-SO3H
does not exhibit differences from its parent. This indicates here the T3

difculty of certifying that sulfonation with ClSO3H actually oc- T1, T2

curred on the benzene moieties. Note that the complexity of reso- c insitu-C-SBA-15-SO3H
nances is probably due to the presence of two additional
overlapping resonances at 18 and 58 ppm, which are attributed
-10 -20 -30 -40 -50 -60 -70 -80 -90 -100
to CH3 and CH2 (4C and 5C) of ethoxy groups formed from the reac-
tion of residual silanols and ethanol during the extraction process (ppm)
[39]. Signals due to trace amount of remaining P123 are also pres- Fig. 4. 29Si solid-state NMR spectra of (a) C-SBA-15-SO3H, (b) grafted-C-SBA-15-
ent at 7078 ppm. SO3H, and (c) in situ-C-SBA-15-SO3H.
M. Karaki et al. / Journal of Catalysis 305 (2013) 204216 209

13
Fig. 5. C solid-state NMR spectra of (a) C-SBA-15-SO3H, (b) grafted-C-SBA-15-SO3H, and (c) in situ-C-SBA-15-SO3H.

200 C-SBA-15-SO3H
grafted-C-SBA-15-SO3H
insitu-C-SBA-15-SO3H

150
Q (kJ.mol )
-1

100

50

0
0 100 200 300 400 500 600 700
n NH3 adsorbed (mol.g-1)

Fig. 6. Differential heat of NH3 adsorption as a function of NH3 coverage.

SO3H

O OH

HO Si Si O

O O

C-SBA-15-SO3H

O O OH

HO3S Si O Si Si O

O
OH OH

grafted-C-SBA-15-SO3H or insitu-C-SBA-15-SO3H
Scheme 2. Molecular structure of the expected acid sites for C-SBA-15-SO3H, grafted-C-SBA-15-SO3H, and in situ-C-SBA-SO3H.
210 M. Karaki et al. / Journal of Catalysis 305 (2013) 204216

Table 2
Acidic features of the acid hybrid materials: sulfur content, number of acid sites, differential heat of ammonia adsorption drawn from NH3 adsorption monitored by calorimetry.

Acid hybrid solids Sulfur contenta Gas-phase NH3 adsorption monitored by calorimetry
(wt%)
Qdiff NH3 adsorptionb Total acidity (tot. H+) Qdiff > 80 kJ mol1 Strong, medium acidity(s,mH+)
(kJ mol1) Qdiff > 100 kJ mol1
tot. H+ amount Molar ratio s,m H+ amount Molar ratio s,m
(mmol g1) tot.H+/S (mmol g1) H+/S
Grafted-C-SBA-15- 0.8 130 0.35 1.4 0.10 0.4
SO3H
In situ-C-SBA-15- 3.1 130 0.40 0.4 0.20 0.2
SO3H
C-SBA-15-SO3H 4.2 150 0.65 0.5 0.50 0.4
a
Obtained by elemental analysis.
b
Mean differential heat of ammonia adsorption at half NH3 coverage.

Table 3
Comparison of the acidic hybrid SBA-15 solids with the acidic resin Amberlyst-15 by
means of turnover frequencies deduced from the initial activity for IPA dehydration.

Total number Initial rate IPA TOF


of acid sites dehydration (h1)
(mmol g1) (mmol h1 g1)
Grafted-C-SBA-15-SO3H 0.35a 4.4 13
In situ-C-SBA-15-SO3H 0.40a 2.2 5.5
C-SBA-15-SO3H 0.65a 70 107
Amberlyst-15 4.70b 55 12
a
Measured by NH3 adsorption isotherms coupled with the calorimetry
experiments.
b
Capacity exchange of A-15 provided by Rohm and Haas Company.

Fig. 7. Evolution of the rate of isopropanol dehydration with time on stream for
grafted-C-SBA-15-SO3H and in situ-C-SBA-15-SO3H. Conditions: Treaction = 120 C, Table 4
PIPA = 8.7 Torr diluted in N2, total ow rate = 2 l h1, mcata = 100 mg, WHSVIPA = 0.5 - Supercial atomic composition obtained by XPS analysis of C-SBA-15-SO3H, in situ-C-
h1. Catalyst pretreatment: 2 h under owing N2 at 120 C. SBA-15-SO3H, grafted-C-SBA-15-SO3H, and Amberlyst-15.

S Si O C
(atomic%) (atomic%) (atomic%) (atomic%)
C-SBA-15-SO3H 3.3 15.2 30.2 50.6
C-SBA-15-SO3H 2.3 15.8 32.4 48.8
After IPA dehydration
In situ-C-SBA-15-SO3H 1.9 15.2 27.8 54.5
In situ-C-SBA-15-SO3H 1.8 15.7 27.4 54.6
After IPA dehydration
Grafted-C-SBA-15-SO3H 0.4 17.3 26.3 55.8
Grafted-C-SBA-15-SO3H 0.5 15.3 28.1 55.5
After IPA dehydration
Amberlyst-15 6.8 19.8 72.5

sites. These results were not the expected ones, since the formed
acidic sites should exhibit a given intrinsic acidic strength corre-
lated with their molecular structure: a sulfonated aromatic ring
for C-SBA-15-SO3H and a sulfonated propyl group for grafted-C-
Fig. 8. Evolution of the rate of isopropanol dehydration with time on stream.
SBA-15-SO3H and in situ-C-SBA-SO3H (Scheme 2). Note that the
Comparison of C-SBA-15-SO3H with a commercial sulfonated resin, Amberlyst A-15. calorimetric curve recorded with Amberlyst 15 exhibits a perfect
Conditions: Treaction = 120 C, PIPA = 8.7 Torr diluted in N2, total ow rate = 5 l h1, plateau characterized by a QdiffNH3 of 135140 kJ mol1 [40]. How-
mcata = 20 mg, WHSVIPA = 6.7 h1. Catalyst pretreatment: 2 h under owing N2 at ever, it is observed that the mean heat of NH3 adsorption at half
120 C.
NH3 coverage clearly has a higher value for the solid obtained by
sulfonation with ClSO3H (C-SBA-15-SO3H) than that for the solids
obtained via MPTMS/H2O2 (grafted or in situ-C-SBA-15-SO3H)
3.1.3. Acidic properties (Table 2). The total number of acid sites is in the same order of
3.1.3.1. Calorimetry of NH3 adsorption. Calorimetry of ammonia magnitude, from 0.35 to 0.65 mmol H+ g1. Note that this is not
adsorption was performed to determine the acid strength of these correlated with the bulk S content. In particular, grafted-C-SBA-
materials. Fig. 6 shows the differential enthalpy of ammonia 15-SO3H, which has the lowest S bulk content, 0.8%, presents a
adsorption plotted against the amount of adsorbed ammonia for supercial H+ density equivalent to those of the other two solids,
all catalysts. The calorimetric curves show a progressive decrease in situ-C-SBA-15-SO3H and C-SBA-15-SO3H, with 3.1% and 4.2% of
in the differential heat of ammonia adsorption with the coverage. sulfur in the bulk. As regards the medium and strongest acid sites,
This is indicative of the heterogeneity of the acid strength of the the solid prepared via the in situ method shows the smallest
M. Karaki et al. / Journal of Catalysis 305 (2013) 204216 211

Fig. 9. XPS analysis, photopeak S2p of C-SBA-15-SO3H and Amberlyst 15.

number of sites, while the two other catalysts are equivalent. These initial rate, we have calculated the initial activity per active site
results suggest that the accessibility of the sulfonated moieties is (Table 3). Based on the TOF values, the following ranking is ob-
strongly dependent on the method of synthesis: the organo-SBA- served: C-SBA-15-SO3H  grafted-C-SBA-15-SO3H Amberlyst-
15 acidied via a post-treatment method leads logically to the 15 > in situ-C-SBA-15-SO3H. The discrepancy observed between
highest accessibility of the acid sites, while for the solid prepared the sulfonated ordered mesoporous organosilicas and the Amber-
via the in situ method, part of the acid sites remain inaccessible. lyst 15 accounts for the higher protonic density of the resins. Note
The H+/S molar ratio gave information on the origin of the acidic that this result is not the expected one, since the molecular struc-
properties. When the total number of acids sites is taken into ac- ture of the acid sites of the ordered mesoporous organosilica, C-
count, i.e., sites with QdiffNH3 > 80 kJ mol1, the H+/S molar ratio SBA-15-SO3H, should be a sulfonated aromatic ring, similar to that
overreaches the expected value of 1 for grafted-C-SBA-15-SO3H, of Amberlyst-15. This difference could be due to the differences in
while signicantly lower values are obtained for in situ-C-SBA- the supercial polarities of the two types of organic acid catalysts;
SO3H and C-SBA-15-SO3H, 0.4 and 0.65 respectively (Table 2). otherwise, this might indicate the presence of acid sites of a differ-
When the molar H+/S ratio is determined with the stronger acid ent nature on C-SBA-15-SO3H, in agreement with the calorimetric
site numbers only (QdiffNH3 > 100 kJ mol1), the number of acid curves. The two other ordered mesoporous organosilicas, which
sites per sulfur atoms is always lower than 0.5. content a sulfonated alkyl group, grafted-C-SBA-15-SO3H and
in situ-C-SBA-15-SO3H, present TOF values one order of magnitude
lower. The results seem logical according to the higher electronic
3.1.3.2. Isopropanol dehydration in gas phase as model reaction. The
density of the SO3 of the propyl group than to the one linked to
conversion of isopropanol (IPA) is used as a model reaction to eval-
the aromatic ring, leading to a lower intrinsic acidity.
uate the acidic properties of the sulfonated SBA-15 organosilica
The deactivation in the gas phase can be caused by the poison-
catalysts in the gas phase. Indeed, the reaction products depend
ing of the acid sites and/or by a structural modication of the acid
on the nature of the catalytic sites: formation of propene and di-
sites. To understand the deactivation process, XPS studies were
isopropyl ether (DIPE) by dehydration on acid sites; formation of
performed on the as-synthesized catalysts, and those recovered
acetone by dehydrogenation on basic sites or by oxidation on redox
after the IPA reaction.
sites when the reaction is performed in the presence of molecular
oxygen [41]. Here, the dehydration of isopropanol is performed un-
der standard conditions, in N2 as carrier gas, at 120 C, with IPA- 3.1.3.3. XPS analysis of the fresh and used ordered mesoporous
WHSV values chosen to get an initial conversion between 10% organosilicas. The supercial atomic compositions are reported in
and 60%. Table 4. The supercial S contents are roughly in agreement with
The acid-functionalized hybrid SBA-15 show quite different the bulk ones, except for in situ-C-SBA-15-SO3H, where the S spe-
activities depending on the method of acid functionalization. The cies were introduced during the solid synthesis and not via a
use of MPTMS/H2O2 gives a low activity compared to the hybrid post-treatment. Thus, some S atom might be entrapped in the bulk.
SBA-15 prepared via sulfonation with ClSO3H. Fig. 7 shows the evo- The photopeaks S2p of C-SBA-15-SO3H and of Amberlyst-15 are
lution of the rate of IPA dehydration with TOS on grafted-C-SBA-15- compared in Fig. 9.
SO3H and on in situ-C-SBA-SO3H, prepared with MPTMS/H2O2. The For Amberlyst-15, the S2p line shows two distinct signals due to
former solid exhibits very fast initial deactivation down to a low the S2p3/2 and S2p1/2 levels. For Amberlyst-15, the S2p3/2 bind-
rate of dehydration. The latter solid is initially less active, but the ing energy of 168.9 eV is the expected value for a sulfonated aro-
initial deactivation is less pronounced, and then the rate of dehy- matic compound. The S2p3/2 peak for C-SBA-15-SO3H is
dration decreases slowly with time on stream. The results for C- observed at higher binding energy and shows two components.
SBA-15-SO3H are presented in Fig. 8. Its initial rate of isopropanol The main component is observed at 170.1 eV; the minor one has
dehydration is 10 times higher than that of the previous catalysts. a lower binding energy, close to that measured on Amberlyst 15.
The catalyst also shows deactivation with time on stream. Note This indicates the presence of S species distinct from a sulfonated
that the activity level of the hybrid SBA-15 sulfonated with ClSO3H aromatic group. For in situ-C-SBA-15-SO3H, the S2p3/2 signal also
is comparable to that of the reference catalyst, Amberlyst 15. How- presents two components, a main component at a quite high bind-
ever, the latter is stable with TOS. ing energy level, 169.2 eV, and one at 168 eV (Fig. 10). The intensity
Although the synthesized catalysts exhibited fast initial deacti- of the S2p signal is very low on grafted-C-SBA-15-SO3H, but the
vation, which may have hindered reliable determination of the maximum of the S2p3/2 peak is also at a quite high binding energy
212 M. Karaki et al. / Journal of Catalysis 305 (2013) 204216

Fig. 10. XPS analysis, photopeak S2p for C-SBA-15-SO3H, in situ-C-SBA-15-SO3H, grafted-C-SBA-15-SO3H. Fresh and used samples recovered after IPA reaction.

level, 169.0 eV. Note that for the two previous ordered mesopor- revealed no changes, showing the relatively higher thermal stabil-
ous organosilicas, an additional signal is observed at 163.4 eV, ity of the S species in the sulfonated propyl group. This result
ascribed to the SH group. This means, as already observed by rather supports a deactivation process by acid site poisoning on
13
CP-MAS NMR, that the oxidation of the MPTMS was not these two solids.
completed.
After reaction, the XPS analysis indicated a decrease in the 3.2. Cellobiose hydrolysis in hot water and hydrothermal stability of
supercial S content for the C-SBA-15-SO3H, from 3.3 down to sulfonated organosilica SBA-15 catalysts
2.3 atomic%. Since the supercial C concentration did not change
signicantly, most likely, the deactivation is caused by the release The sulfonated organosilica SBA-15 materials were applied to
of some S species. The S2p3/2 photopeak of the used catalyst seems cellobiose hydrolysis in hot water in order to compare the mode
to indicate a decrease in the component at higher binding energy, of sulfonation for this acidic demanding reaction. In Fig. 11 are pre-
which could characterize weakly bonded S species such as sented the results obtained at 150 C with the same mass of cata-
sulfonated silanols. In contrast, the XPS analysis of the in situ-C- lysts for 2 h of reaction. Under these conditions, C-SBA-15-SO3H,
SBA-15-SO3H and grafted-C-SBA-15-SO3H recovered after reaction the acid catalyst obtained after sulfonation by ClSO3H, gave the
M. Karaki et al. / Journal of Catalysis 305 (2013) 204216 213

100 conversion 2 hours 4 hours


yield 150 C 150 C
80 76 74

64
60 52
% 45
40 33 33 33
24
17 17 19
20 13 14
6 8

0
H
H H H O3
t t -S
O3 SO
3
ys H H O 3 ys -S 15
tal SO
3
-S
O3 -S tal 15 -1
5-
A-
ca 5- 15 15 ca - A
no 1 A-
- no SB
A
SB SB
A- BA C- C-
SB - SB -S C- d-
C- C -C it u- te
- d af
itu fte ins g r
ins g ra

Fig. 11. Cellobiose conversion and glucose yield for the different solid catalysts. Conditions: batch, cellobiose 0.2 g, catalyst 0.02 g, H2O 20 ml, He 10 bar.

grafted-C-SBA-15-SO3H, gave lower cellobiose conversion together


Table 5
Sulfur leaching after reaction at 150 C for 4 h. with lower glucose selectivity, the latter being the less effective
catalyst.
Catalyst S in solid S in liquid S leaching
A priori, the superiority of C-SBA-15-SO3H is in agreement with
before reaction after (%)
(wt%)a reaction its much higher acidic properties in terms of strength and number
(mg l1)a of acid sites (see Table 2). If we deduce from these results the activ-
C-SBA-15-SO3H 4.2 49 100 ity per acid site, then the most active catalyst is C-SBA-15-SO3H,
Grafted-C-SBA-15-SO3H 0.8 5.7 75 followed by in situ-C-SBA-15-SO3H and grafted-C-SBA-15-SO3H.
In situ-C-SBA-15-SO3H 3.1 19 52 We then investigated in detail the stability of these sulfonated
a
Obtained by elemental analysis. organosilica SBA-15. In Table 5 are reported the amounts of sulfur
detected in the liquid phase after 4 h of reaction at 150 C. For all
the catalysts, a large leaching occurred. C-SBA-15-SO3H was clearly
the least stable, with an impressive 100% loss of sulfur. Grafted-C-
Table 6
SBA-15-SO3H lost 75% of sulfur. Interestingly, in situ-C-SBA-15-
Sulfur leaching for C-SBA-15-SO3H after cellobiose conversion at various T for 2 h.
SO3H was able to maintain half of its sulfur amount.
Temperature Cellobiose conversion Glucose yield S leaching Most likely, the reaction operated mainly via the sulfur-based
(C) (%) (%) (%)
acidic species formed in the liquid phase. To conrm this, we trea-
150 64 52 100 ted C-SBA-15-SO3H in water at 150 C for 1 h. Here also, we ob-
140 30 27 98 served a 100% leaching of sulfur. Then, we used separately the
130 32 12 84
120 27 3 95
two phases for cellobiose hydrolysis at 150 C. The washed solid
phase has a conversion of 28% for a glucose yield of 6%, while the
solid-free liquid phase alone gave a conversion of 88% for a glu-
cose yield of 67%. This denitively conrmed that the cellobiose
higher conversion, yield, and also selectivities. After 4 h, up to 74% hydrolysis at 150 C proceeds homogeneously.
glucose yield was obtained, with a selectivity of 97%. Note that Then, we decided to decrease the reaction temperature to
there was not a strong increase in conversion between 2 and 4 h 120 C to see whether high activity could still be obtained, while
of reaction. The two other catalysts, in situ-C-SBA-15-SO3H and the sulfur leaching was minimized. Results are presented in Table 6

SO3H

O OH

HO Si Si O

O O
1
O OH
ClSO3H
HO Si Si O

O O
O OSO3H

HO3SO Si Si O

O O
2
Scheme 3. Some possible sulfur species formed during the sulfonation of C-SBA-15 with ClSO3H.
214 M. Karaki et al. / Journal of Catalysis 305 (2013) 204216

50 conversion 50
45
yield
40 40
33 33
30 29
30 28 30
26
% 23 % 24
22
20 19 19 20 19
17 17 18

10 10

0 0
before after run 2 run 3
before after run 2 run 3
washing washing
washing washing
run 1
run 1

insitu -C-SBA-15-SO3H grafted -C-SBA-15-SO3H

Fig. 12. Cellobiose conversion and glucose yield after catalyst washing and recycling over in situ-C-SBA-15-SO3H and grafted-C-SBA-15-SO3H. Conditions: cellobiose 0.2 g,
catalyst 0.02 g, H2O 20 ml, He 10 bar, 150 C, 4 h.

200
insitu-C-SBA-15-SO3H
insitu-C-SBA-15-SO3H after washing

150
Q (kJ.mol-1)

100

50

0
0 100 200 300 400 500 600 700
n NH3 adsorbed (mol.g-1)

Fig. 13. NH3 adsorption calorimetry for in situ-C-SBA-15-SO3H, as-synthesized and hot water washed samples.

and show that decreasing the temperature led to a loss of activity, order to favor the sulfonation of the aromatic ring, which would
while a high level of leaching was maintained. This indicates that provide more water-tolerant acid sites than sulfonated silica sites.
the acid sites of this material were not stable in water even at a With in situ-C-SBA-15-SO3H and grafted-C-SBA-15-SO3H, we
quite low temperature. This proves also that a temperature lower performed recycling studies. The catalysts were initially washed
than 150 C does not give a signicant cellobiose conversion even with water (150 C, 1 h) to try to remove most of the unstable sul-
in the presence of a large amount of solubilized acid. fur species. Then, after a rst catalytic run, the solids were ltered
We attempted to stabilize the sulfur species on the C-SBA-15- off, dried, and reused for further runs. Results are presented in
SO3H catalyst by applying thermal treatments. Basically, the solid Fig. 12.
was heated under owing N2 at a temperature of 200 C for 3 h. We note different behavior for the two catalysts. For in
This led to a slight decrease in the S amount in the solid (3.6 wt% situ-C-SBA-15-SO3H, a decrease of 30% in conversion and yield is
instead of 4.2 wt%), but leaching studies have shown sulfur loss observed after the rst hot water washing. For grafted-C-SBA-15-
identical to that above. SO3H, a lower conversion decrease is observed after the hot water
C-SBA-15-SO3H, the acid organosilica SBA-15 prepared by sulfo- washing, of 20%. Then, a slow deactivation is obtained for the suc-
nation of the parent acid with ClSO3H, is denitively not a stable cessive runs in both cases.
acid catalyst for water-phase reactions, at least at such tempera- Note that for both washed catalysts, these performances are
tures. These results are in contrast to what was reported by Esquivel just better than that obtained without any catalyst at 150 C for
et al. [32], who claimed the stability of acid species of similar mate- 4 h, giving a cellobiose conversion of 17% (Fig. 11), but a gain is still
rials after reactions in water but at 80 C. In fact, the exact place of observed. Nevertheless, even if the present solids gave only a slight
the sulfonation by chlorosulfonic acid remains unclear. Neverthe- added activity to the reaction, it seems that the remaining species
less, the XPS analysis reported above clearly showed the presence are stable enough to allow recycling of the catalysts.
of S species distinct from the sulfonated aromatic groups. These Obviously, the sulfur species still present on the solid present
species are predominant, and most likely, they might correspond too weak acidity and/or are in too limited amount to perform this
to any sulfonated silica sites. These species would not be stable reaction efciently, as checked by calorimetry. This was conrmed
and would rapidly hydrolyze when heated in water (Scheme 3). by calorimetry of ammonia adsorption (Fig. 13). After hot water
The improvement of the method of synthesis might be an issue in washing at 150 C for 1 h, the in situ-C-SBA-15-SO3H shows drastic
M. Karaki et al. / Journal of Catalysis 305 (2013) 204216 215

13
Fig. 14. C solid-state NMR spectra of (a) grafted-C-SBA-15-SO3H and (b) in situ-C-SBA-15-SO3H after washing.

solids were characterized by the calorimetry of NH3 adsorption


Table 7 and by mean of the model reaction of gas-phase isopropanol dehy-
Total organic carbon (TOC) in liquid phase and S leaching after washing for 1 h at dration. The most acidic material was that obtained by sulfonation
150 C.
with chlorosulfonic acid, in terms of number and strength of the
Catalyst S in solid TOC S mol C/mol S acid sites. However, contrary to our expectation, we evidenced
(wt%)a (g l1) leaching in solution by calorimetry the heterogeneity of the acid strength distribution,
(%)a
suggesting the presence of distinct sites of sulfonation. This is
Grafted-C-SBA-15-SO3H 0.8 0.05 37 39 strengthened by XPS analysis of the mesoporous organosilicas
In situ-C-SBA-15-SO3H 3.1 0.02 22 6
structure in comparison with that of the reference, Amberlyst-15.
a
Obtained by elemental analysis. The organosilica SBA-15, sulfonated with ClSO3H, presents a higher
TOF than the reference Amberlyst-15 for isopropanol dehydration;
reduction in its numbers of acid sites: 75% of the acid sites with however, the former solid is far less stable. XPS analysis of the used
heat of NH3 adsorption higher than 100 kJ mol1 were lost. catalyst indicated that the deactivation in the gas-phase IPA dehy-
We performed solid-state 13C NMR on the washed materials to dration resulted from the release of the most weakly bonded S spe-
try to evaluate if any structural changes occurred during the hot cies, assumed to be sulfonated silanols. The organosilicas, SBA-15,
water treatment. Fig. 14 shows that the spectra of the materials ob- obtained via H2O2 oxidation of the thiol functions are less acidic
tained after washing are very close to those before washing. materials and have a low activity for IPA dehydration, one order
All the initial signals are present, still indicating the presence of of magnitude lower. However, no sulfur release was observed for
Si(CH2)3SO3H and Si(CH2)3SH moieties on the material. How- gas-phase reaction.
ever, the corresponding signals are roughly and homogeneously The acidic materials were used for cellobiose hydrolysis into
less intense. This could indicate the cleavage of the SiC bond. To glucose in water. Good performances were obtained, with glucose
clarify the mechanism of sulfur species release, TOC and sulfur ele- yields between 19% and 74% after 4 h at 150 C. Acid stability stud-
mentary analyses were performed on the water recovered from the ies showed that the reaction proceeded homogeneously due to
washing step. The results in Table 7 show a large excess of the car- large sulfur leaching, in particular in the case of the acid solid
bon molar amount over the sulfur amount. Although the aromatic formed with chlorosulfonic acid. We particularly showed that this
group might contribute to the TOC analysis, this result rather sup- catalyst is not stable under hydrothermal conditions, with total
ports a mechanism of S leaching via SiC bond cleavage. leaching of its sulfur content. Concerning the catalyst prepared
with MTPMS, a washing pretreatment led to a decrease in the ef-
4. Conclusions ciency for cellobiose hydrolysis, but a gain of stability allowing
recycling was obtained. The C and S analysis in the water recovered
In this paper, we reported the preparation and full characteriza- from a hot water washing step suggests the partial loss of some
tion of sulfonated periodic mesoporous organosilicas with SBA-15 sulfonated functions via the SiC bond cleavage of the functional
structure. The materials were prepared from the condensation of group Si(CH2)2CH2SO3H via hydrolysis.
1,4-bis(triethoxysilyl)benzene. The material was acidied follow- As seen in Section 1, some studies report the use and recycling
ing different routes using 3-mercaptopropyltrimethoxysilane/ of such sulfonated catalysts for reactions under hydrothermal con-
H2O2 or chlorosulfonic acid as a sulfonating agent. We showed par- ditions without deactivation. However, it seems that these systems
ticularly that the introduction of mercaptopropyl groups and their are actually not stable enough if used in hot water, even in the
further oxidation with H2O2 did not change the structure of the presence of hydrophobic organic moieties within the framework,
material, while the sulfonation with chlorosulfonic acid led to a assumed to improve their water tolerance and to bring peculiar
less organized solid. The acidic properties of the as-synthesized performances for water-phase reaction.
216 M. Karaki et al. / Journal of Catalysis 305 (2013) 204216

Further studies are still needed to improve the synthesis of [13] C. Moreau, R. Durand, J. Duhamet, P. Rivalier, J. Carbohydr. Chem. 16 (1997)
709714.
acidic ordered mesoporous organosilica SBA-15 and achieve a con-
[14] S. Suganuma, K. Nakajima, M. Kitano, S. Hayashi, M. Hara, ChemSusChem 5
trolled local environment of the acid site. This is probably the key (2012) 18411846.
to solid materials that would exceed the performances of the refer- [15] M. Marzo, A. Gervasini, P. Carniti, Carbohydr. Res. 347 (2012) 2331.
ence acidic sulfonated resins. [16] D. Takahashi, S. Kobayashi, K. Tsumura, K. Toshima, Chem. Lett. 38 (2009) 728
729.
The search for strong Brnsted acid catalysts stable under [17] W. Deng, R. Lobo, W. Setthapun, S.T. Christensen, J.W. Elam, C.L. Marshall,
hydrothermal conditions, active, and recyclable is a challenge for Catal. Lett. 141 (2011) 498506.
the development of clean processes, especially those involving bio- [18] A. Charmot, A. Katz, J. Catal. 276 (2010) 15.
[19] C. Tagusagawa, A. Takagaki, A. Iguchi, K. Takanabe, J.N. Kondo, K. Ebitani, T.
mass transformations in water. Tatsumi, K. Domen, Chem. Mater. 22 (2010) 30723078.
[20] J.A. Bootsma, B.H. Shanks, Appl. Catal. A Gen. 327 (2007) 4451.
[21] J.A. Bootsma, M. Entorf, J. Eder, B.H. Shanks, Bioresour. Technol. 99 (2008)
Acknowledgments 52265231.
[22] V. Degirmenci, D. Uner, B. Cinlar, B.H. Shanks, A. Yilmaz, R.A. van Santen, E.J.M.
The authors thank the CNRS and the Universities of Haute-Als- Hensen, Catal. Lett. 141 (2011) 3342.
[23] A. Takagaki, M. Nishimura, S. Nishimura, K. Ebitani, Chem. Lett. 40 (2011)
ace and Lyon and the Lebanese University for funding. The authors 11951197.
thank the technical services of IS2M and IRCELYON, particularly S. [24] L. Pena, M. Ikenberry, B. Ware, K.L. Hohn, D. Boyle, X.S. Sun, D. Wang,
Prakash for XPS analysis. Biotechnol. Bioprocess Eng. 16 (2011) 12141222.
[25] M. Chaplin, Water Structure and Science (2009).
[26] P.L. Dhepe, M. Ohashi, S. Inagaki, M. Ichikawa, A. Fukuoka, Catal. Lett. 102
(2005) 163169.
References
[27] D. Margolese, J.A. Melero, S.C. Christiansen, B.F. Chmelka, G.D. Stucky, Chem.
Mater. 12 (2000) 24482459.
[1] C. Marcilly, Catalyse Acido-basique, Ed. Technip, 2003. [28] K. Nakajima, I. Tomita, M. Hara, S. Hayashi, K. Domen, J.N. Kondo, Catal. Today
[2] D.E. Lopez, J.G. Goodwin Jr., D.A. Bruce, J. Catal. 245 (2007) 381. 116 (2006) 151156.
[3] B. Hamad, R.O. Lopes de Souza, G. Sapaly, M.G. Carneiro Rocha, P.G. Pries de [29] Y. Goto, S. Inagaki, Chem. Commun. (2002) 24102411.
Oliveira, W.A. Gonzalez, E. Andrade Sales, N. Essayem, Catal. Commun. 10 [30] E.-B. Cho, D. Kim, J. Phys. Chem. Solids 69 (2008) 11421146.
(2008) 9297. [31] J.A. Melero, G.D. Stucky, R. van Grieken, G. Morales, J. Mater. Chem. 12 (2002)
[4] R. Weingarten, G.A. Tompsett, W. Curtis Conner Jr., G.W. Huber, J. Catal. 279 16641670.
(2011) 174182. [32] D. Esquivel, C. Jimenez-Sanchidrian, F.J. Romero-Salguero, J. Mater. Chem. 21
[5] I. Agirrezabal-Telleria, J. Requies, M.B. Gmez, P.L. Ariaz, Appl. Catal. B Environ. (2011) 724733.
115-116 (2012) 169178. [33] M.P. Kapoor, Y. Kasama, M. Yanagi, T. Yokoyama, S. Inagaki, T. Shimada, H.
[6] C. Carlini, M. Giuttari, A.M. Raspolli Galletti, G. Sbrana, T. Armaroli, G. Busca, Nanbu, L.R. Juneja, Micropor. Mesopor. Mater. 101 (2007) 31239.
Appl. Catal. A Gen. 183 (1999) 295302. [34] G.-J. Sohn, H.-J. Choi, I.-Y. Jeon, D.W. Chang, L. Dai, J.-B. Baek, ACS Nano 6
[7] F. Chambon, F. Rataboul, A. Cabiac, C. Pinel, E. Guillon, N. Essayem, Appl. Catal. (2012) 63456355.
B Environ 105 (2011) 171181. [35] F. Liu, X. Meng, Y. Zhang, L. Ren, F. Nawaz, F.-S. Xiao, J. Catal. 271 (2010) 5258.
[8] Rodrigo Lopes de Souza, Franck Rataboul, Nadine Essayem, Challenges 3 (2) [36] Y. Goto, S. Inagaki, Chem. Commun. (2002) 24102411.
(2012) 212232, http://dx.doi.org/10.3390/challe3020212. [37] M.H. Lim, A. Stein, Chem. Mater. 11 (1999) 3285.
[9] R.M. Ravenelle, F. Schbler, A. DAmico, N. Danillina, J.A. Van Bokhoven, J.A. [38] W. Guo, F. Kleitz, K. Cho, R. Ryoo, J. Mater. Chem. 20 (2010) 82578265.
Lercher, C.W. Jones, C. Sievers, J. Phys. Chem. C 114 (2010) 1958219595. [39] V. Goletto, M. Imperor, F. Babonneau, Stud. Surf. Sci. Catal. 135 (2001) 1129.
[10] B. Kamm, P.R. Gruber, M. Kamm, Bioreneries Industrial Processes and [40] T.H.T. Vu, H.T.Au, T.H.T Nguyen, M.H. Do, N.Q. Bui, N. Essayem, Catal.Lett.
Processes and Products, Ed. Wiley-VCH, 2006. submitted for publication.
[11] S. Van de Vyver, L. Peng, J. Geboers, H. Schepers, F. de Clippel, C.J. Gommes, B. [41] B. Chelighem, S. Launay, N. Essayem, G. Coudurier, M. Fournier, J. Chim. Phys.
Goderis, P.A. Jacobs, B.F. Sels, Green Chem. 12 (2010) 15601563. 94 (1997) 18311837.
[12] A. Onda, T. Ochi, K. Yanagisawa, Green Chem. 10 (2008) 10331037.

You might also like