Vous êtes sur la page 1sur 7

Direct synthesis, characterization and catalytic application of SBA-15

mesoporous silica with heteropolyacid incorporated into their


framework
Xiaoli Sheng, Jie Kong, Yuming Zhou

, Yiwei Zhang, Zewu Zhang, Shijian Zhou


School of Chemistry and Chemical Engineering, Southeast University, Nanjing 211189, PR China
a r t i c l e i n f o
Article history:
Received 26 May 2013
Received in revised form 9 September 2013
Accepted 9 December 2013
Available online 13 December 2013
Keywords:
Heteropolyacid
Mesoporous SBA-15
Acidic catalysis
Catalysis stability
a b s t r a c t
The Keggin phosphotungstic acid, H
3
PW
12
O
40
(HPW) incorporated into SBA-15 ordered mesoporous silica
were synthesized via a method involving the introduction of HPW in an acidied solution of P123 triblock
copolymer (EO
20
PO
70
EO
20
), the SBA-15 mesostructuring agent (direct synthesis). Samples with similar
HPW loadings were also prepared by impregnation of SBA-15. A comparison between direct incorpora-
tion of HPW into mesoporous silica and impregnation of HPW on mesoporous silica was done. Character-
ization by elemental analysis, XRD, N
2
adsorption, TEM, DRS-UV and FTIR spectroscopy showed that after
calcination HPW in the direct synthesized samples was better dispersed or may even be partially embed-
ded in the pore walls. Moreover, their catalytic behaviors were investigated in the alkylation of o-xylene
with styrene. Results show that direct synthesized sample has the better catalytic performances in terms
of yield and stability. This behavior may be owing to high dispersion of the HPW species on the SBA-15
and the strong interaction between the HPW and the support, thus prevent HPW leaching from the
support.
2013 Elsevier Inc. All rights reserved.
1. Introduction
Hetropoly acids (HPA) have witnessed rapid growth in the last
decade as solid acid catalyst [13]. Polyoxometalates with Keggin
structure have been chosen as catalyst because of their easy
availability and extreme stability in solution as well as in solid
state. 12-Tungstophosphoric acid (HPW), in particular, has been
the target catalyst among the Keggin series in many earlier reports
because of the strongest acidity [45]. However, the main
disadvantage is their very low surface area (<10 m
2
g
1
) and hence
it becomes necessary to disperse HPW on supports that possess
large surface area. The use of SBA-15 mesoporous molecular sieves
is attractive, because it possesses well-ordered pore structures,
high thermal stability and high surface area [6]. One possible
route to obtain these supported HPW catalysts is the direct
impregnation of the support with a heteropoly acid solution
followed by evaporation of the solvent [79]. Although the
conventional wet impregnation method is easy to increase its
surface area by supporting HPW onto various carriers. However,
weak interaction between the HPW and the support resulted in
its leaching in polar media [10]. The reaction stabilities of the
catalysts are still not satisfactory. Thus, it is necessary to further
improve the catalytic performances of the catalysts.
Leaching of HPW from ordered mesoporous silica in polar reac-
tion media can be prevented by surface modication of the sup-
port. In our previous work, the deposition of basic alumina
clusters; doping of the silicate with La atoms; and functionaliza-
tion of the silicate walls with aminosilane groups for anchoring
the HPW molecules were reported to be successful [1113]. Incor-
poration of HPA into the pores of a mesoporous material can also
be achieved by encapsulating HPA during the synthesis of the silica
material itself [1420]. Indeed, Yang et al. incorporated HPW into
SBA-15 support by adding HPW to the SBA-15 synthesis mixture
[15,17]. Toufaily et al. reported a similar approach to incorporate
HPW into MSU type ordered mesoporous silica [18]. Shi et al. ob-
tained SBA-15-supported HPW catalysts by adding P and W
sources into the initial solgel system during hydrolysis of tetra-
ethyl orthosilicate to form the Keggin-type HPA in situ [14,16].
In this work, The Keggin phosphotungstic acid (HPW) incorpo-
rated into SBA-15 ordered mesoporous silica were synthesized via
a method involving the introduction of HPWin an acidied solution
of P123 triblock copolymer (EO
20
PO
70
EO
20
), the SBA-15 mesostruc-
turing agent (direct synthesis). There are some important differ-
ences between other synthesis procedures and ours. The main
difference is related to the timing of the introduction of HPW.
We introduced HPW before the hydrolysis of TEOS, whereas in
References [15,17,18] TEOS was already present and had reacted
1387-1811/$ - see front matter 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.micromeso.2013.12.007

Corresponding author. Tel./fax: +86 25 52090617.


E-mail address: ymzhou@seu.edu.cn (Y. Zhou).
Microporous and Mesoporous Materials 187 (2014) 713
Contents lists available at ScienceDirect
Microporous and Mesoporous Materials
j our nal homepage: www. el sevi er . com/ l ocat e/ mi cr omeso
prior to HPW addition. Shi et al. introduced P and W sources
(Na
2
WO
4
and Na
2
HPO
4
) instead of HPW in the synthesis mixture.
The aim of this paper is to compare the differences between
direct incorporation of HPW into mesoporous silica and impregna-
tion of HPW on mesoporous silica. The effect of the incorporation/
impregnation of HPW on the structure of mesoporous solids was
investigated by different techniques. The catalytic properties of
the catalysts were assessed in the alkylation of o-xylene with
styrene. Special attention was paid to catalyst stability and
reusability.
2. Experimental
2.1. Catalyst preparation
SBA-15 samples were synthesized following a previously pub-
lished method using the P123 tri-block copolymer (EO
20
PO
70
EO
20
)
as a structure directing agent [6]. After the hydrothermal step,
samples were thoroughly washed with distilled water and were
dried at 120 C for 24 h. Calcination was performed in air at
540 C for 6 h (heating rate: 2 C min
1
). The Keggin phosphotung-
stic acid (H
3
PW
12
O
40
-HPW) was dispersed on the silica support
either using an aqueous incipient wetness impregnation technique
or via a direct synthesis method.
The aqueous incipient wetness impregnation technique was
previously employed by many authors for the immobilization and
dispersion of HPA on high surface area supports [79]. We inserted
12-Tungstophosphoric Acid (HPW) by stirring 2 g of freshly cal-
cined SBA-15 in a 10 mL aqueous solution containing 0.27 mmol
of H
3
PW
12
O
40
6H
2
O for 6 h at 353 K. The impregnated powders
are dried at 120 C overnight and calcined in air at 300 C for 4 h.
The resulting HPW/SBA-15 samples prepared via impregnation
are denoted as HPW/SBA-15-PS (PS refers to the impregnation syn-
thesis method).
Incorporationof HPWvia the direct synthesis route was achieved
via the following adaptation of the SBA-15 synthesis procedure
(shown in Fig. 1). P123 polymer, 4.00 g; 30 g of water; and 120 g of
HCl (2 M) were mixed following the standard method for SBA-15
synthesis. This solution was stirred for over 3 h at 40 C and then
added to a second solution containing 0.27 mmol of H
3
PW
12
O
40
6H
2
O in a 10 mL aqueous solution. The mixture was stirred for
24 h before addition of 9.4 g tetraethyl orthosilicate (TEOS). During
hydrolysis of TEOS a white precipitate was formed. After stirring for
another 30 min, the mixture then transferred into a Teon-lined
autoclave and aged for 48 h at 80 C. The resulting solid was ltered,
washed with deionized water and dried at 100 C for 24 h. The calci-
nation step was performed in air at 540 C for 6 h (heating rate:
2 C min
1
). The resulting HPW/SBA-15 samples were labeled as
HPW/SBA-15-DS (DS refers to the direct synthesis method).
2.2. Characterization
Elemental analysis of samples was performed by means of
X-ray uorescenece (XRF) analysis on a SWITZERLAND ARL9800
XRF. The corresponding weight ratio of anhydrous HPW on dry
SBA-15 of the different catalysts is exhibited in Table 1.
Powder X-ray diffraction (XRD) patterns were obtained with a
Rigaku D/max-rC Siemens diffractometer using nickel-ltered Cu
Ka as monochromatic X-ray radiation. The scattering intensities
were measured over an angle range of 0.58 < 2h < 40 with a step
size D(2h) = 0.028 and a step time of 8 s.
The nitrogen adsorption and desorption isotherms were mea-
sured at 196 C on an ASAP-2020 (Micromertics USA). The spe-
cic surface area, A
BET
, was determined from the linear part of
the BET equation (P/P
0
= 0.050.25). The pore size distribution
was derived from the desorption branch of the N
2
isotherm using
the BarrettJoynerHalenda (BJH) method. The total pore volume
was estimated from the amount of nitrogen adsorbed at a relative
pressure (P/P
0
) of ca. 0.995. Pore structures of the samples were
examined by TEM (Jeol, JEM-2000EXII).
Infrared spectra were recorded on a Bruker Tensor 27 (German)
using DRIFT techniques, scanned from 4000 to 400 cm
1
. The sam-
ple was ground with KBr and pressed into a thin wafer. The sam-
ples were evacuated at 300 C for 4 h before the measurement.
The diffuse reectance UVvis spectra were collected using a
SHIMADZU UV3600 (Japan) scanning spectrophotometer. The
powder sample was loaded into a quartz cell, and the spectra were
collected over the range of 200800 nm reference to BaSO
4
.
2.3. Catalytic tests
The alkylation reactions were carried out in a continuously
stirred batch reactor under reux conditions using a three-neck
100-mL round-bottomaskequippedwitha condenser. Preliminary
runs were conducted with 7.50 g (0.0721 mol) of styrene, 57.35 g
(0.5402 mol) of o-xylene (mole ratio of o-xylene to styrene, 7.5:1)
and 1.50 g of catalyst (20% w/w of styrene) at 120 C for 180 min.
The required amount of o-xylene was initially added to the reactor
at the reaction temperature, followed by the desired amount of
catalyst, a known amount of styrene was then added to the reaction
mixture at the same temperature. After the reaction, unreacted
o-xylene was distilled out under atmospheric pressure and then a
collected part was called as crude product. The crude product was
analyzed with GC-9890A gas chromatograph equipped with OV-1
capillary column and a ame ionization detector (FID). The yield of
PXE was dened as follows:
yield of PXE%
actual product weight
all theoretical product weight
100
actual product weight crude product weight
PXEchromatography%
1
Surfactant (S
0
) P123: (EO)
70
(PO)
20
(EO)
70
S
0
+ HCl(aq.) (S
0
/H
3
O
+
)(Cl
-
) (S
0
/H
3
O
+
)(Cl
-
/H
5
SiO
4
+
) SBA-15
TEOS
T
(S
0
/H
3
O
+
)(Cl
-
;PW
12
3-
) (S
0
/H
3
O
+
)(Cl
-
;PW
12
3-
/H
5
SiO
4
+
)
TEOS
HPW/SBA-15-DS
T
Surfactant (S
0
) P123: (EO)
70
(PO)
20
(EO)
70
S
0
+ HCl(aq.) (S
0
/H
3
O
+
)(Cl
-
) (S
0
/H
3
O
+
)(Cl
-
/H
5
SiO
4
+
) SBA-15
TEOS
T
(S
0
/H
3
O
+
)(Cl
-
;PW
12
3-
) (S
0
/H
3
O
+
)(Cl
-
;PW
12
3-
/H
5
SiO
4
+
)
TEOS
HPW/SBA-15-DS
T
Fig. 1. Proposed synthesis mechanism for HPW/SBA-15-DS samples.
8 X. Sheng et al. / Microporous and Mesoporous Materials 187 (2014) 713
3. Results and discussion
3.1. Characterization of the HPW/SBA-15 catalysts
The differences in HPW aggregation on the SBA-15 supports fol-
lowing the two synthesis methods were evidenced by XRD (Fig. 2).
All materials showed a sharp peak indexed as (100) and a smaller
peaks indexed as (110), which are typical of the 2-D hexagonal
(p6 mm) SBA-15 material [6]. As for the HPW/SBA-15-DS sample,
the peak height of the main (100) reection is nearly one second
as the initial SBA-15 and the reection of (200) is little intense.
The diffraction peak of impregnated HPW appears to have almost
the same intensity. Generally, the appearance of the main (100)
reection means the existence of mesopore in all HPW/SBA-15
samples. Therefore, this nding suggests that the mesoporous
structure of SBA-15 is not destroyed after the introduction of
HPW irrespective of the synthesis methods. The actual concentra-
tion of HPW in the samples was estimated by means of XRF anal-
ysis. XRF results are listed in Table 1. The amount of HPW found in
HPW/SBA-15-PS sample is less than that in HPW/SBA-15-DS sam-
ple possibly due to the reaction between surfactant and HPW dur-
ing the synthesis of the HPW/SBA-15-DS sample. In addtion, as
shown in Fig. 1(b), the high-angle region did not show the charac-
teristic diffraction pattern of crystalline HPW phase for the HPW/
SBA-15-DS sample, which is an evidence of the dispersed nature
of HPW in HPW/SBA-15-DS samples. On the contrary, samples pre-
pared by the impregnation of the same HPW loading onto SBA-15
presented the characteristic XRD pattern of crystalline HPW.
Fig. 3a shows the N
2
adsorptiondesorption isotherms of SBA-
15, HPW/SBA-15-PS, and HPW/SBA-15-DS. As can be seen, all sam-
ples exhibit typical IV type isotherms and H1 type hysteresis loops
at high relative pressures. This indicates that SBA-15 with a fairly
uniform pore size distribution was successfully prepared by the
two synthesis methods. However, the volumes adsorbed inected
sharply at relative pressure (P/P
0
) 0.63 for SBA-15, 0.50 for the
HPW/SBA-15-PS and HPW/SBA-15-DS samples. The desorption
branch extended to a lower relative pressure suggesting a partial
loss of structural organization and the formation of some narrower
pores. The BJH pore size distribution (calculated from the analysis
of the desorption branch of the isotherms) is presented in Fig. 3b. It
can be seen that SBA-15 samples exhibit a fairly uniform pore size
distribution (57 nm). It is well worth mentioning that that the
HPW/SBA-15-DS sample has pore sizes around 57 nm, similar to
the SBA-15. The result is consistent with the reported literature
[19].
In the DS synthesis method, the size of the majority of mesop-
ores is not affected by the presence of HPW, which is in agreement
with the proposed synthesis model (Fig. 1) in which the pore size is
imposed by the P123 micelles, which have the same diameter in
the presence and absence of HPW. On the contrary, impregnation
of the calcined SBA-15 sample with HPW resulted in a reduction
of the mesopore diameter, which shows that HPW inside the mes-
opores occupies space and decreases the pore width. On the other
hand, the results obtained fromthe textural analysis of the samples
based on nitrogen adsorption are presented in Table 1. The HPW/
SBA-15-DS samples have a specic surface area of 430 m
2
/g, which
is signicantly lower than that of the unmodied SBA-15 (810 m
2
/
g). The HPW/SBA-15-DS samples have a higher specic surface
area of 622 m
2
/g. It is clear that the surface area and pore volume
decrease with the introduction of the HPW. The loss of surface area
upon HPW loading using impregnation can be related to the
agglomeration of HPW molecules on the external surface of the
material resulting in pore blockage. Interestingly, the HPW/SBA-
15-DS sample shows very similar pore size to that of SBA-15, in
Table 1
Texture properties of various catalysts (inside parentheses, the data after reaction).
Entry Sample S
BET
a
(m
2
g
1
) V
total
(cm
3
g
1
) Pore size
b
(nm) HPW (wt.%) (calculated) HPW
c
(wt.%) (found)
1 SBA-15 810 1.27 6.45
2 HPW/SBA-15-PS 622 (720) 0.92 (1.07) 6.02 (5.96) 40 39.5
3 HPW/SBA-15-DS 430 (456) 0.82 (0.87) 6.45 (6.45) 40 41.9
a
BET method.
b
BJH model applied to the desorption branch of the isotherm.
c
From XRF analysis.
1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
HPW/SBA-15-DS
HPW/SBA-15-PS
(110)
(200)
(110)
(100)

i
n
t
e
n
s
i
t
y
(
a
.
u
.
)
2 Theta(degree)
(a)
SBA-15
5 10 15 20 25 30 35 40
i
n
t
e
n
s
i
t
y
(
a
.
u
.
)
2 Theta(degree)
(b)
HPW
HPW/SBA-15-DS
HPW/SBA-15-PS
Fig. 2. Low-angle (a) and high-angle (b) XRD patterns of pure SBA-15 and SBA-15
materials modied with HPW via the PS and DS methods.
X. Sheng et al. / Microporous and Mesoporous Materials 187 (2014) 713 9
good agreement with the pore size distribution calculated from the
BJH isotherm model (Fig. 3b). This result clearly indicates that
HPW in the direct synthesis samples was better dispersed or
may even be partially embedded in the pore walls, compared to
that of HPW/SBA-15-PS sample.
The TEM images for HPW/SBA-15-DS and HPW/SBA-15-PS sam-
ples are presented in Fig. 4. Obviously, the pores of the impreg-
nated sample are hexagonal analogous to those of pure silica
SBA-15 [6]. The DS sample has a different morphology with irreg-
ular particle size conrming that the HPW had an impact on the
formation process of the mesoporous material. This result suggests
that the ordering of the HPW/SBA-15-DS sample is less than that of
the HPW/SBA-15-PS sample, which is in agreement with the re-
sults of XRD.
The IR spectra of the pure HPW, HPW/SBA-15-PS and HPW/SBA-
15-DS are shown in Fig. 5. Pure HPW shows IR bands approxi-
mately at 1080 (P-O in the central tetrahedron), 980 (terminal
W = O) and 890 and 800 (W-O-W) cm
1
corresponding to asym-
metric vibration associated with Keggin ion [21]. For the HPW/
SBA-15-PS and HPW/SBA-15-DS samples, although the absence of
vibration band at 1080 cm
1
could be probably owing to the
concealment by the strong and broad background of SBA-15, the
vibration bands at approximately 980, 890 and 800 cm
1
can be
clearly observed. This unambiguously demonstrates that the
primary structure of HPW Keggin anions is preserved after the
modied via the two synthesis methods. On the other hand, it is
noteworthy that the W-O and W-O-W bands of [PW
12
O
40
]
3
in
the HPW/SBA-15-PS and HPW/SBA-15-DS samples appear at
slightly red-shift positions compared to those of the pure
HPW, indicating the presence of a strong interaction between
[PW
12
O
40
]
3
and SBA-15 supports [22]. As can be seen, the band
of 980 cm
1
in the HPW/SBA-15-PS and HPW/SBA-15-DS samples
appear at the 976 cm
1
and the 950 cm
1
, respectively. It could
be considered that interaction between [PW
12
O
40
]
3
and SBA-15
supports of the HPW/SBA-15-DS sample is more than that of the
HPW/SBA-15-PS sample.
The interaction between HPW and SBA-15 can also be con-
rmed by DRS-UV analyses as shown in Fig. 6. It is known that
the diffuse reectance UVvis spectroscopy is a sensitive probe
for the identication and characterization of metal ion coordina-
tion, as well as its existence in the framework or in the extraframe-
work position of metal-containing zeolites [23]. As can be seen, a
strong signal in the UVvisible spectra at k = 265 nm is observed
for the HPW/SBA-15-PS(PS-1) catalyst, which is assigned to
the oxygen-metal charge transfer of tungstophosphate anion
PW
12
O
40
3
[24]. However, for the HPW/SBA-15-DS (DS-2) sample,
the signal at the k = 265 nm is shifted slightly at the k = 240 nm,
implying that in this case PW
12
O
40
3
Keggin type structure of the
ligand environment has been affected. This suggests that phospho-
tungstic acid in the direct synthesis method may part into the hole
wall of mesoporous material, which changed the interaction
between the carrier and phosphotungstic acid, thus affecting its
characteristic absorption peak of the peak position.
3.2. Catalytic activity
The catalytic activity of different catalysts is showed in Table 2.
The detailed reaction scheme is shown in Scheme 1, reaction (1) is
the PXE formation reaction whereas reaction (2) and (3) represent
the formation of styrene oligomers and more substitutes,
respectively.
As listed in Table 2, the homogeneous HPW show very high cat-
alytic performances for the reaction, however, it is difcult to sep-
arate the HPW from the product mixture. Furthermore, the SBA-15
support itself shows no catalytic performance. It is noteworthy
that the HPW/SBA-15-DS catalyst exhibits the much higher con-
version (100%) and higher product yield (91.5%) than those of
HPW/SBA-15-PS catalyst. In this case a signicant difference is
found between the two catalysts as a function of the different
preparation methods. Combined with the data of Table 1, it can
be seen that although the catalyst in the impregnation synthesis
method has a larger specic surface area, but its pore size is small,
is not conducive to the macromolecular reaction. In contrast, the
HPW/SBA-15-DS catalyst has a larger pore size, is benecial to
the reaction. Possibly, this behavior may be explained by the high-
er dispersion as revealed in Fig. 5. From these reasons, it can be de-
duced that high dispersion of HPW onto SBA-15 and the large
surface area with suitable pore size might account for the high cat-
alytic activity.
The important questions that must be addressed while studying
alkylation processes over a solid catalyst relate to the stability of
the catalyst to leaching of the active component and the possibility
of catalyst recycling. The catalytic reusability of the HPW/SBA-15-
PS and HPW/SBA-15-DS catalysts was evaluated by carrying out
the reaction with used catalyst under the optimized conditions.
After each run, the catalyst was recovered by ltration, then
washed with ethanol, dried and used again. The data obtained
are shown in Fig. 7. It can be seen that only 7% reduction in the
0.0 0.2 0.4 0.6 0.8 1.0
0
200
400
600
800
1000
HPW/SBA-15-DS
HPW/SBA-15-PS
(a)
V
o
l
u
m
e

a
d
s
o
r
b
e
d

(
c
m
3

S
T
P
/
g
)
Relative Pressure(P/P
0
)
SBA-15
0 5 10 15 20
0
5
10
15
20
25
HPW/SBA-15-DS
HPW/SBA-15-PS
SBA-15
P
o
r
e

V
o
l
u
m
e
(
c
m
3
/
g
)
Pore Diameter(nm)
(b)
Fig. 3. (a) Low-temperature nitrogen adsorptiondesorption isotherms and (b) Pore
size distribution patter of purely siliceous SBA-15 HPW/SBA-15-PS and HPW/SBA-
15-DS samples.
10 X. Sheng et al. / Microporous and Mesoporous Materials 187 (2014) 713
activity is observed after 4 runs on the HPW/SBA-15-DS catalyst. In
contrast, the deactivation of the HPW/SBA-15-PS catalyst is much
faster and the PXE yield drops to a very low level of 40% after
the fourth reaction cycle. The poor catalytic stability of HPW/
SBA-15-PS may be due to the possibility that HPW leaching from
the catalyst support into the liquid solvent, which may result in
the low conversion. On the other hand, the decrease of yield arising
from catalysts lost during separation and transfer of catalysts into
the next reaction cycle cannot be excluded. This observation re-
veals satised reusability for HPW/SBA-15-DS, which means that
HPW have only a slight tendency to leach from the SBA-15 carrier
in reaction. This result is also implying that a strong interaction
between the HPW and the SBA-15 in the direct synthesis method,
thus prevent HPW leaching from the catalyst support.
In order to further conrm the interaction between the HPW
and the SBA-15, the catalysts after reaction have been studied by
the N
2
adsorptiondesorption isotherms. The results obtained from
the textural analysis of the samples based on nitrogen adsorption
are presented in Table 1 (inside parentheses). It is clear that the
surface area and pore volume increase after the reaction, this
may be due to the blockage of the micropores with grafted HPW
species is exposed, led to the pore volume and specic surface area
increase. On the other hand, it is also possible that at the end of the
reaction, HPW leaching from the catalyst support into the liquid
b
d
a
b
c
c
b
d d
aa
b
c
c
c
c c
Fig. 4. TEM images of various samples: (a) HPW/SBA-15-DS in [100] direction, (b) HPW/SBA-15-DS in [110] direction, (c) HPW/SBA-15-PS in [100] direction, and (d) HPW/
SBA-15-PS in [110] direction.
1200 1100 1000 900 800 700 600
(c)
(b)
T
r
a
n
s
m
i
t
t
a
n
c
e
(
a
.
u
)
Wavenumber cm
-1
1080
980
890 800
976
950
(a)
Fig. 5. FT-IR spectra of pure HPW (a), HPW/SBA-15-PS (b) and HPW/SBA-15-DS (c).
200 300 400 500 600
0
1
2
DS-2
PS-1
A
b
s
o
r
b
a
n
c
e

Wavelength / nm
Fig. 6. UVvis diffuse reectance spectra of HPW/SBA-15-PS(PS-1) and HPW/SBA-
15-DS(DS-2) catalysts.
X. Sheng et al. / Microporous and Mesoporous Materials 187 (2014) 713 11
solvent may lead to a number of pores on the surface of the carrier
increases, which could contribute to the surface area of the cata-
lysts. In addition, from the data (see Table 1), it can be seen that
only 6% increase in the specic surface area is observed after 4 runs
on the HPW/SBA-15-DS catalyst. In contrast, in the case of the
HPW/SBA-15-PS catalyst the specic surface area was increased
by 16% after the fourth reaction cycle. These data can also conrm
much HPW leaching from the HPW/SBA-15-PS catalyst, but little
HPW leaching from the HPW/SBA-15-DS catalyst after four cata-
lytic cycles. Therefore, it is reasonable that a strong interaction be-
tween the HPW and the SBA-15 in the direct synthesis method
results in efcient immobilization of HPW, which maintains its
high activity in acid-catalyzed reactions.
4. Conclusions
In this work, a procedure is described for direct incorporation of
Keggin-type heteropolyacids into ordered mesoporous silica by
using a mixture of the introduction HPW in an acidied solution of
P123 triblock copolymer. After incorporation or impregnation, the
heteropolyacid anions preserved their Keggin structure on the sur-
face of mesoporous SBA-15.
Characterization results from XRF, XRD, N
2
adsorption, TEM,
DRS-UV and FTIR spectroscopy indicate that HPW was better dis-
persed or may even be partially embedded in the pore walls in
the direct synthetic method. The HPW/ SBA-15-DS catalyst is
highly efcient in the alkylation of o-xylene with styrene. It had
the best catalytic performances with styrene conversion up to
100% and PXE yield up to 91.5%. The catalyst could be used for
more than four times without any signicant loss of activity and
leaching of tungsten species in the reaction mixture. The good
stability can be attributed to the strong interaction between the
SBA-15 and HPW molecules.
Acknowledgments
The authors are grateful to the nancial supports of National
Natural Science Foundation of China (Grant No. 21306023,
21376051, 21106017 and 51077013), Fund Project for Transforma-
tion of Scientic and Technological Achievements of Jiangsu Prov-
ince of China (Grant No. BA2011086), Specialized Research Fund
for the Doctoral Program of Higher Education of China (Grant
No.20100092120047), KeyProgramfor the Scientic ResearchGuid-
ing Found of Basic Scientic Research Operation Expenditure of
Southeast University (Grant No. 3207043101) and Instrumental
Analysis Fund of Southeast University.
References
[1] F.E. Celik, H. Lawrence, A.T. Bell, J. Mol. Catal. A Chem. 288 (2008) 8796.
[2] A. Heydari, S. Khaksar, M. Sheykhan, M. Tajbaksh, J. Mol. Catal. A Chem. 287
(2008) 58.
[3] A. Kumar, P. Singh, S. Kumar, R. Chandra, S. Mozumdar, J. Mol. Catal. A Chem.
276 (2007) 95101.
[4] J. Kaur, I.V. Kozhevnikov, Chem. Commun. 19 (2002) 25082509.
[5] T. Rajkumar, G. Ranga Rao, J. Mol. Catal. A Chem. 295 (2008) 19.
[6] D. Zhao, J. Feng, Q. Huo, N. Melosh, G.H. Fredrickson, B.F. Chmelka, G.D. Stucky,
Science 279 (1998) 548552.
[7] J. Wang, H.O. Zhu, Catal. Lett. 93 (2004) 209212.
[8] Q.Y. Liu, W.L. Wu, J. Wang, X.Q. Ren, Y.R. Wang, Micropor. Mesopor. Mater. 76
(2004) 5160.
[9] H.X. Jin, Q.Y. Wu, W.Q. Pang, Mater. Lett. 58 (2004) 36573660.
[10] A. Popa, V. Sasca, J. Halasz, Appl. Surf. Sci. 255 (2008) 18301835.
[11] X.L. Sheng, Y.M. Zhou, Y.W. Zhang, Y.Z. Duan, M.W. Xue, Chem. Eng. J. 179
(2012) 295301.
[12] X.L. Sheng, Y.M. Zhou, Y.W. Zhang, Y.Z. Duan, Z.W. Zhang, Micropor. Mesopor.
Mater. 161 (2012) 2532.
[13] X.L. Sheng, Y.M. Zhou, Y.W. Zhang, Y.Z. Duan, M.W. Xue, Catal. Lett. 142 (2012)
360367.
[14] C. Shi, R. Wang, G. Zhu, Eur. J. Inorg. Chem. 23 (2005) 48014807.
[15] L. Yang, Y. Qi, X. Yuan, J. Shen, J. Kim, J. Mol. Catal. A Gen. 229 (2005) 199205.
Table 2
Activity of various catalysts a in alkylation of o-xylene with styrene.
a
Catalyst Styrene conversion (%) PXE Yield
c
(%) PXE Selectivity
d
(%)
SBA-15
HPW
b
100 97.9 9:1
HPW /SBA-15-PS 91 68.3 9:1
HPW /SBA-15-DS 100 91.5 9:1
a
Reaction conditions: o-xylene: styrene = 7.5:1, reaction temperature = 120 C,
reaction time = 3.0 h, catalyst loading = 20% (w/w of styrene).
b
Homogeneous catalyst, 0.30 g.
c
Isolated yield based on the amount of styrene.
d
Ratio of para-to-ortho product.
+
Cat
O-Xylene Styrene Phenylxylyl ethane(PXE)
(1)
n
Cat
Styrene Styrene Oligomer
(2)
O-Xylene Styrene
Cat
more substitutes
(3)
Scheme 1. Reaction scheme of alkylation of o-xylene with styrene over a
heterogeneous catalyst.
1 2 3 4
0
20
40
60
80
100 HPW/SBA-15-DS
HPW/SBA-15-PS
P
X
E

Y
i
e
l
d
(
%
)
Reaction cycle
Fig. 7. Catalytic stability of the HPW/SBA-15-DS and HPW/SBA-15-PS catalysts in
the alkylation of o-xylene with styrene (Reaction conditions: o-xylene:sty-
rene = 7.5:1, reaction temperature = 120 C, reaction time = 3.0 h, catalyst load-
ing = 20% (w/w of styrene).
12 X. Sheng et al. / Microporous and Mesoporous Materials 187 (2014) 713
[16] C. Shi, R. Wang, G. Zhu, Eur. J. Inorg. Chem. 15 (2006) 30543061.
[17] L. Yang, J. Li, X. Yuan, J. Mol. Catal. A Gen. 262 (2007) 114118.
[18] J. Toufaily, M. Soulard, J.L. Guth, Colloid Surf. A Physicochem. Eng. Aspects 312
(2008) 285291.
[19] B.C. Gagea, Y. Lorgouilloux, Y. Altintas, J. Catal. 265 (2009) 99108.
[20] A. Popa, V. Sasca, E.E. Kiss, Mater. Res. Bull. 46 (2011) 1925.
[21] P. Staiti, S. Freni, S. Hocevar, J. Power Sources 79 (1999) 250255.
[22] S.S. Wu, J. Wang, W.H. Zhang, X.Q. Ren, Catal. Lett. 125 (2008) 308314.
[23] S.C. Laha, P. Mukherjee, S.R. Sainkar, J. Catal. 207 (2002) 213223.
[24] J.C. Hu, Y.D. Wang, L.F. Chen, Micropor. Mesopor. Mater. 93 (2006) 158163.
X. Sheng et al. / Microporous and Mesoporous Materials 187 (2014) 713 13

Vous aimerez peut-être aussi