Vous êtes sur la page 1sur 9

Applied Catalysis A: General 400 (2011) 176–184

Contents lists available at ScienceDirect

Applied Catalysis A: General


journal homepage: www.elsevier.com/locate/apcata

Direct synthesis of well-ordered mesoporous Al-SBA-15 and its correlation with


the catalytic activity
Pallavi Bhange a,∗ , Deu S. Bhange a,b , Sivaram Pradhan a,c , Veda Ramaswamy d
a
Catalysis Division, National Chemical Laboratory, Pune 411008, India
b
Department of Chemistry, Shivaji University, Kolhapur 416004, MS, India
c
Department of Chemistry, University of Liverpool, UK
d
Chemical Physics Laboratory, Central Leather Research Institute, Chennai 600020, India

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

Article history: The synthesis of Al-SBA-15 under milder acidic conditions were made through an adjusting the molar
Received 4 March 2011 H2 O/HCl ratio which indicate the formation of Si–O–Al linkages that lead to isomorphous substitution
Received in revised form 23 April 2011 of Si4+ by some Al3+ ions. By adjusting the H2 O/HCl molar ratio, Al gets incorporated into the lattice of
Accepted 25 April 2011
SBA-15, which is evidenced by XRD, N2 adsorption, TEM, 29 Si and 27 Al MAS NMR spectroscopic data.
Available online 30 April 2011
Aluminium incorporated SBA-15 samples retained both structural and textural properties of SBA-15. The
27
Al MAS NMR confirms that a large proportion of the Al is inserted into tetrahedral positions within the
Keywords:
framework. The Al3+ ions could assume a tetrahedral coordination and be part of the hexagonal structure
Al-SBA-15
Esterification of acetic acid
of silica in Al-SBA-15. In the present study, we have examined the effectiveness of Al-SBA-15, as an
Benzylation of anisole acid catalyst for the esterification of acetic acid and in benzylation of anisole. The high activity of Al-
20 sample in esterification reaction has been attributed to isolated, generally tetrahedrally coordinated,
framework Al species. While in benzylation of anisole, the best result was achieved with Al-5 catalyst.
The selectivity for benzylanisole increases and that for dibenzylether decreases with increase in the Si/Al
ratio of Al-SBA-15.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction pH-adjusting method in their synthesis gel. Yue et al. [1] reported
the direct synthesis of Al-SBA-15 and found that catalytic activity of
The incorporation of aluminium into SBA-15 by post-synthetic Al-SBA-15 in cumene cracking is higher as compared to AlMCM-41.
and direct methods has been reported [1–7]. During materials However, the highly acidic synthesis gel required for the formation
preparation via post-synthetic methods often metal oxides are of SBA-15 limits the direct incorporation of high amounts of triva-
formed in the channels or on the external surface. Metal oxides lent metal ions into the neutral silica framework and requires a
formed in the mesopores will block the pores partially or fully, post-synthetic treatment to remove octahedral aluminium. A dif-
thereby reducing surface area, pore volume, and pore diameter, ferent strategy is the transformation of amorphous SBA-15 walls
or play a negative role in catalysis. Hartmann et al. [8] have how- into crystalline aluminosilicates [10,11]. Even though the materi-
ever reported the successful in situ incorporation of Al in SBA-1, but als formed are good cracking catalysts, the synthesis procedure is
the uptake of Al in the framework under highly acidic conditions somewhat tedious and requires two steps. Therefore, it is still a
of synthesis of SBA-1 or similar mesoporous materials is random challenge to find a one-step route to SBA-15 materials with high Al
with a huge difference between the input and output ratios of Si/Al. content in order to increase the acidity without changing its struc-
Mesoporous Al-SBA-15 synthesized under highly acidic condition tural order or increasing the complexity of the synthesis. Recently,
(pH < 1) has a low amount of Al incorporated into SBA-15, possibly Vinu et al. [12] have reported an optimized procedure for the syn-
due to the high solubility of Al precursors, which hinder their incor- thesis of high Al content SBA-15 molecular sieves including an
poration into the silica walls of SBA-15. Wu et al. [9] reported, for adjustment of the nH2 O /nHCl ratio in order to lower the pH of the
the first time, that Al3+ species were highly substituted into meso- synthesis medium. Recent reports suggest that structurally inte-
porous SBA-15 molecular sieves by using NH4 F acidic method and gral SBA-15 and Al-SBA-15 samples can be synthesized at slightly
higher pH (2–5) conditions, provided the rate of hydrolysis of tetra
alkyl-orthosilicate (TMOS or TEOS used as the silica source) and
∗ Corresponding author. Present address: Department of Chemical Engineering, condensation in presence of other metal cations are controlled, for
Pohang, University of Science & Technology, South Korea. Tel.: +82 054 279 8276. example, by adjusting the nH2 O to nHCl molar ratio of the synthesis
E-mail address: bhangepallavi@yahoo.in (P. Bhange). gel [13–17].

0926-860X/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2011.04.031
P. Bhange et al. / Applied Catalysis A: General 400 (2011) 176–184 177

Esterification of acetic acid with n-butyl alcohol is commercially diffractometer using Ni filter, Cu K␣ radiation ( = 1.5406) and pro-
important as the product n-butyl acetate having applications in portional counter as detector. The BET surface area of the samples
the manufacture of lacquer, artificial perfume, flavouring extract, was determined by N2 adsorption at 77 K by using Autosorb-1
leather, photographic films, plastics and safety glass. Usually esteri- instrument. The TEM of the samples were recorded on a JEOL
fication reaction carried out by use of various conventional mineral Model 1200EX microscope operating at 100 kV. The 27 Al and 29 Si
acids such as H2 SO4 , HF, H3 PO4 , HCl and p-toluene sulphonic acids MAS NMR spectra were recorded at a Bruker MSL 300 NMR spec-
[18–20]. The replacement of these conventional hazardous and trometer. Tetraethyl orthosilicate (ı = 82.4 ppm from TMS) and
corrosive liquid acid catalysts by solid acid catalyst is one of the Al(H2 O)6 3+ was used as the reference for 29 Si and 27 Al, respec-
demands of the society. Many solid catalysts, e.g., ion-exchange tively. The total amount of acidity present in the catalyst was
resins [21,22], zeolites [23,24] and acidic clay catalysts [25] used estimated using temperature-programmed desorption (TPD) of
in reactions have been reported in the literature. Many other het- NH3 on a Micromeritics AutoChem 2910 instrument. To evaluate
erogeneous catalysts reported for esterification reaction include the nature of acid sites (Brönsted and Lewis) of the Al-SBA-15 sam-
supported heteropolyacids [26,27], niobic acid [28], sulphated zir- ples with different Al loading, the pyridine adsorption spectra in
conia in SBA-15 [29] and sulphonic acid functionalized SBA-15 the region 1650–1400 cm−1 were measured with subsequent ther-
[30]. mal treatment at temperature between 323 and 473 K by in situ
Benzylation of anisole is an important reaction as these com- Fourier transform infrared (FTIR) spectroscopy in drift mode on
pounds are constituents of lubricants with improved properties, an FTIR-8300 Shimadzu SSU-8000 instrument with 4 cm−1 reso-
like thermal oxidation and hydrolytic stabilities [31]. Use of benzyl lution and averaged over 500 scans. These studies were performed
alcohol instead of benzyl chloride is advantageous due to the for- by heating precalcined powder samples in situ from room temper-
mation of water as only side product. Lachter et al. have reported ature to 673 K with a heating rate of 5 K min−1 in a flowing stream
niobium phosphate giving high yield and selectivity in the benzyla- (40 ml min−1 ) of pure N2 . The samples were kept at 673 K for 3 h
tion of anisole with benzyl alcohol [32]. Niobium oxide supported and then cooled to 323 K; then pyridine vapors (20 ␮l) were intro-
on alumina also showed the high activity for anisole alkylation [33]. duced under N2 flow and then IR spectra were recorded at various
Heteropoly acid catalyst, H3 PO4 –WO3 –Nb2 O5 presented higher stages of pyridine desorption, which was continued by evacuation
activity than the pure niobium oxide and a facile reusability for at progressively higher temperatures (323–473 K). A resolution of
the same reaction [34]. 4 cm−1 was attained after averaging over 500 scans for all the IR
In the present report, we describe our attempt to prepare Al- spectra recorded here.
SBA-15 by in situ synthesis method. The introduction of Al into the
mesoporous SBA-15 was done with different nSi /nAl ratios without
2.3. Catalytic activity
changing the structural integrity of the parent SBA-15 materials by
simply adjusting the nH2 O /nHCl molar ratio to 796. The structural
The esterification of acetic acid reaction was carried out by tak-
variations in these are characterized using powder XRD, N2 adsorp-
ing 0.025 g of the calcined sample (dried at 393 K for 6 h in an oven)
tion, TEM, TPD, 29 Si and 27 Al MAS NMR spectroscopic techniques.
with 2 mmol of acetic acid (Merck, 99.8%) and 4 mmol of alcohol
The intrinsic activity of aluminium is investigated in the esterifica-
(Merck, 98%) at 353 K in a 100 ml round-bottomed flask equipped
tion of acetic acid with n-butanol and in the benzylation of anisole
with a reflux condenser. The liquid phase benzylation of anisole
over Al-SBA-15 samples. The influence of various reaction parame-
was carried out using Al-SBA-15 catalyst in a two-necked round-
ters such as temperature, mole ratio, catalyst amount and different
bottomed flask in nitrogen atmosphere using anisole and benzyl
Al loading is also presented.
alcohol in a 10:1 mole ratio with 0.1 g of freshly prepared catalyst in
the temperature range of 353–393 K. Aliquots of the reaction mix-
2. Experimental ture were collected at different time intervals and analyzed by GC
(Agilent Technologies, model 6890N, capillary column HP-5, 30 m,
2.1. Preparation of Al-SBA-15 by direct synthesis route containing 5% methyl + 95% phenyl siloxane). Identification of the
products was done by GC–MS.
The Al-SBA-15 samples were prepared by using a fixed water
nH2 O /nHCl molar ratio of 796 (70 ml of 0.07 M HCl) while varying the
3. Results and discussion
initial nSi /nAl ratios of 80, 60, 40, 20, 10 and 5 (pH < 1.0). The molar
gel composition was 1TEOS:0.01–0.2 Al2 O3 :0.016 P123:0.16–0.46
3.1. Structure of Al-SBA-15 samples
HCl:127 H2 O. In a typical synthesis procedure of Al-SBA-15 mate-
rials, 4 g of P123 was added to 30 ml of water. After stirring for
Typical powder XRD patterns of calcined Al-SBA-15 samples in
few hours, a clear solution was obtained. Thereafter, the required
the range from 0.5◦ < 2 < 3◦ were shown in Fig. 1A. They exhibit
amount of HCl was added, and the solution was stirred for another
very similar patterns with well-resolved diffraction peaks at 0.8◦
2 h. Then, 9 g of tetraethyl orthosilicate and required amount of alu-
(2) and two weak peaks at 1.6 and 1.7◦ (2) due to (1 0 0),
minium chloride were added and then the resulting mixture was
(1 1 0) and (2 0 0) Bragg reflections, respectively. The characteris-
continuously stirred at 313 K for 24 h, and finally crystallized in a
tic hexagonal features of the parent SBA-15 are maintained in all
Teflon-lined autoclave at 373 K for 2 days. The crystallized product
the Al-SBA-15 samples. In Fig. 1A, we notice that the intensity is
was filtered off, washed with deionized water, dried and calcined
maximum for Al-20 and Al-5 samples. Considering the intensity
in air at 823 K for 6 h. The Al-SBA-15 samples were designated as
(counts) of (1 0 0) reflection for all samples, we notice particularly
Al-X, where X denotes the input nSi /nAl ratio.
a higher intensity for some of the Al-SBA-15 samples (viz., Al-5 and
Al-20) which can be correlated to an increase in the wall thick-
2.2. Catalyst characterization ness at higher loading of Al (Table 1). Imperor-Clerc et al. [35] have
noted an increase in the intensity of the reflections (1 0 0), (1 1 0)
The chemical compositions of the samples were estimated by and (2 0 0), on annealing these micropores by hydrothermal treat-
atomic absorption spectrophotometer (Varian spectra AA 220). ment, which increases the pore diameter slightly (viz., Al-5). Sauer
The powder XRD patterns of calcined samples in the 2 range et al. [15] attribute such an increase in intensity to the X-ray scatter-
from 0.5◦ to 3◦ were recorded on X’Pert Pro (M/s Panalytical) ing of the nano-guest species that dilute the Si-SBA-15. The better
178 P. Bhange et al. / Applied Catalysis A: General 400 (2011) 176–184

Table 1
Physicochemical characteristic of Al-SBA-15 samples.

Samples Si/Al Total S.A. (m2 g−1 ) Meso area (m2 g−1 ) Micro area (m2 g−1 ) Total PVa (cm3 g−1 ) Micro vol. (cm3 g−1 ) ‘a’b (nm) PDc (nm) WTd (nm)

In put Out put AAS

Al-80 80 95 656 376 280 0.83 0.126 11.5 7.8 4.8


Al-60 60 78 766 448 318 0.976 0.136 11.3 6.6 5.2
Al-40 40 54 816 397 419 0.92 0.181 10.7 5.7 5.9
Al-20 20 31 826 451 375 1.032 0.173 11.8 6.2 5.0
Al-10 10 22 792 622 170 1.36 0.069 11.3 7.8 3.5
Al-5 5 14 1107 800 307 1.73 0.131 12.0 7.9 4.1
Al-40(I) 40 56 904 513 392 1.26 0.180 10.2 5.6 4.6
Al-10(I) 10 16 873 560 313 1.62 0.139 10.5 7.4 3.1
a
PV = pore volume.

b
a = 2d100 / 3.
c
PD = pore diameter.
d
WT = wall thickness.

ordering in the corona region induces both an increase in intensity peak at 2 = 23◦ , which is of amorphous silica. It is seen that there
and a slight modification of the channel diameter. are no diffraction lines due to crystalline Al2 O3 reflections even at
From a summary of physico-chemical data of all samples given higher Al loadings. As XRD of the samples did not detect any alu-
in Table 1, it is seen that there is a lattice expansion in Al-SBA- mina phase on the surface these species must be XRD amorphous
15 samples as the ionic radii of Al3+ (0.039 nm) is larger than that micro-domains.
of Si4+ (0.026 nm). Isomorphous substitution of Si4+ by Al3+ ions
generally results in an increase in d100 that is consistent with the 3.2. Porosity and surface area
presence of Al3+ ions in the framework or in the silica pore walls.
This should result in well-dispersed Al3+ ions in the lattice with The adsorption–desorption isotherms and pore size distribu-
no isolated Al2 O3 clusters. The powder XRD patterns in the high tions of Al-SBA-15 samples prepared with different amounts of
angle region of 10◦ < 2 < 80◦ for the Al-40 and Al-10 samples are Al content were shown in Fig. S1 (Supplementary Information).
given in Fig. 1B. The wide-angle patterns show a broad diffraction All isotherms are of type IV according to the IUPAC classification
and exhibited a H1-type broad hysteresis loop, which is typical
of large-pore mesoporous solids. A sharp increase in the volume
of N2 adsorbed is observed with increasing Al content above a
A p/po of 0.6 which is characteristic of the capillary condensation
5000
f
d
step. The total surface area increases (from 656 to 1107 m2 g−1 )
with increasing Al content from Si/Al = 80–5, which is essentially
4000 due to an increase in the mesopore area (from 376 to 800 m2 g−1 )
c
a (Table 1).
Intensity, a.u.

The total pore volume also increases with an increase in Al con-


3000 tent (from 0.83 cm3 g−1 to 1.73 cm3 g−1 ). A small decrease in the
e
wall thickness with increase in Al content in Al-SBA-15 is also
2000 noticed. Based on these observations, the localization of Al in the
b
SBA-15 channels may be visualized to consist of two parts, one
in the walls of SBA-15, where Si4+ ions are isomorphously sub-
1000 stituted by Al3+ ions and the other in the corona region, where
the silanol groups during condensation interact with Al3+ ions
0 present in the gel and end up with a decrease in the microp-
0.5 1.0 1.5 2.0 2.5 3.0 ores on calcination. Unit cell size and pore diameter increase with
increasing Al content but the trend is not linear (Table 1). Gener-

ally, in the crystalline zeolites, Al-incorporation slightly increases
the pore size. However, there is no regular rule in SBA-15 as it has
B 500
an amorphous structure where both bond length and angle may
change.
Intensity, counts

400
It appears, therefore, that the Al-ions in SBA-15 exist in dif-
ferent configurations; as Al-associated with Si-ions in the pore
300 b walls (both Td and Oh species) and on the surface as well as dis-
persed alumina like domains containing Td and Oh-types of Al-ions.
200 SBA-15 has thicker pore walls relative to zeolites so that the incor-
porated Al-ions cannot be substituted into the silica framework
100 a
completely. That is, a part of the Al-ions will be exposed on the
pore wall surface so that it might have properties similar to impreg-
0 nated Al-ions on the SBA-15 walls. The incorporated metal-ion
10 20 30 40 50 60 70 80 may interact with surface hydroxyl groups and may contract the
2θ pore wall when combined with two or three hydroxyl groups, so
that the pore size decreases. If these Al-ions are deeply incorpo-
Fig. 1. A powder X-ray diffraction patterns in the region 0.5–3◦ of Al-SBA-15 sam-
rated within the silica mesoporous walls, the pore size might be
ples: (a) Al-80, (b) Al-60, (c) Al-40, (d) Al-20, (e) Al-10 and (f) Al-5. (B) Powder X-ray
diffraction patterns in the region 10–70◦ of Al-SBA-15 samples: (a) Al-40 and (b)
increased as for zeolites. In this study, the pore size increases after
Al-10. Al-ion incorporation as shown in Table 1. This strongly suggests
P. Bhange et al. / Applied Catalysis A: General 400 (2011) 176–184 179

1200 1.8
(b)

1000
Adsorbed Volume (cm3g-1)

Normalized surface area


1.5
800

(a)
600
1.2
Vmeso
400

0.9
200
Vmicro

0
1 2 3 4 5 6 7 0.6
beta parameter 3 6 9 12 15 18 21 24 27

Fig. 2. ␤-Plots of (a) Al-20 and (b) Al-5 samples. Al2O3 (wt%)

Fig. 3. Normalized surface area of Al-SBA-15 samples as a function of Al2 O3 con-


centration for localization of metal oxide particles within SBA-15 channels.
that Al-ions are incorporated into mesoporous silica walls accord-
ing to the above hypothesis. Moreover, it is shown that the pore
size enlargement occurs without any ascertainable change in the 25 wt% Al2 O3 ), the data can be correlated to Al2 O3 existing as a thin
structural order of the materials with increasing Al-ions in the syn- film anchored inside the mesopores of SBA-15.
thesis gel. While in impregnated Al-SBA-15 samples the location of
Al is predominantly in the outer surface of the corona region, as a
3.3. Transmission electron microscopy
thin film or as clusters of Al2 O3 depending upon the concentration
of Al.
The homogeneity of the distribution of Al and the ordering
A significant increase in the BET surface area with an increase
of the hexagonal array of mesopores are examined by transmis-
in Al content is further observed from the ␤-plot where
sion electron microscopy. The transmission electron micrographs
 ln(0.4) 1/2.7 of Al-SBA-15 sample shows the hexagonal array of uniform chan-
␤= nels where Al2 O3 nanoparticles are not observed as dark objects
ln(p/po )
between the walls of SBA-15 (Fig. 4a and b). From the well-ordered
Three different regions can be seen on these plots (Fig. 2): (i) a lin- hexagonal array of mesopores, the two dimensional hexagonal
ear region due to multilayer adsorption in mesopores; (ii) a steep structure can be confirmed. Even at high Al loading for Al-5 sample,
region due to capillary condensation within these mesopores; and small and well-dispersed Al2 O3 -oxide particles are formed, which
(iii) a last linear region due to multilayer formation onto the exter- is not detectable by XRD.
nal surface of the grains.
Micropore and mesopore volumes can be obtained by the 3.4. MAS-NMR characterization
intercept between the adsorbed amount (y-axis) and the lin-
ear segments (i) and (iii). These two extrapolations give Vmicro 29 Si MAS-NMR spectrum of Al-X (X = 40, 20 and 10) samples are

and (Vmicro + Vmeso ), respectively. The ␤-plots of Al-SBA-15 yield shown in Fig. 5a–c, respectively. All the Al-SBA-15 samples con-
nonzero Vmicro values, thus indicating the presence of micropores. sist of two well resolved lines with detectable maxima at −102 to
The present method of preparation leads to an increase in meso- −112 ppm, which have been attributed to the formation of Q3 and
porous surface area (Vmeso ) with increase in Al content that may Q4 species and a low intensity line in the range of −93 to −94 ppm,
indicate the presence of some Al3+ ions as a part of the mesopores which is due to the presence of Si in Q2 (2Si, 2Al) environments. The
formed during the hydrothermal synthesis. presence of Al atoms in the SBA-15 network may generate Si (3Si,
We would still like to see how Al2 O3 species are localized in the Al) and Si (2Si, 2Al) environments, which contribute to the reso-
mesoporous structure of SBA-15. This picture of the location of Al nance peaks at ∼−93 ppm. Broadening of all signals in the spectra
is supported by a further analysis of the data and the application of has been attributed to the large distribution of the T–O–T angles.
the concept of the normalized surface area (NSA). Fig. 3 shows the The Q4 /Q3 ratio decreases with an increase in nSi /nAl ratio from 80
normalized surface area of the samples as a function of Al2 O3 con- to 40. 27 Al MAS NMR spectra of calcined Al-SBA-15 with Si/Al ratio
tent in Al-SBA-15 samples which is calculated using the equation: of 40, 20, 10 and 5 are presented in Fig. 6a–d. The major peak with a
chemical shift of 52–54 ppm is assigned to Al in a tetrahedral envi-
ronment (AlO4 structural unit, Al (tet)) presumably present in the
SAAl-SBA-15
NSA = × SASBA-15 Al-SBA-15 framework.
1−x
The peaks corresponding to ∼35 and 1 ppm are attributed to
where SAAl-SBA-15 and SASBA-15 are specific surface areas of Al-SBA- penta- and hexacoordinated (non-framework) aluminium which
15 and SBA-15 samples, respectively and x is the weight fraction is present in the material. The spectrum shows that most of the Al
of Al2 O3 in the samples. The trend for the Al-SBA-15 samples show in Al-SBA-15 is tetrahedrally coordinated even at a high Al content
that the NSA is significantly greater than unity and increases with Al (Si/Al = 5), and only a small amount of non-framework Al is present
concentration (from Al-80 to Al-10 samples). This indicates absence in the material. The tetrahedral to octahedral aluminium ratio is
of pore blocking and presence of nanoparticles of Al2 O3 embedded found to increase with higher aluminium incorporation as shown
in silica matrix. Whereas at higher Al loading (in Al-5 sample with in Fig. 6a–d.
180 P. Bhange et al. / Applied Catalysis A: General 400 (2011) 176–184

Fig. 4. Transmission electron micrographs: (a) Al-20 and (b) Al-5 samples.

29
Fig. 5. Si MAS NMR spectra of (a) Al-40, (b) Al-20 and (c) Al-10 samples.

3.5. Temperature-programmed desorption of ammonia shown in Fig. S2 (Supplementary information). The highest acidity
in these samples is attributed to the high dispersion of Al atoms at
The temperature-programmed desorption of ammonia (NH3 - the atomic level in the SiO2 lattice. It is noticed, that the total num-
TPD) was performed to determine the total amount of acidity of the ber of acid sites decrease at high Al loading as expected. Strong
catalysts. The TPD of NH3 for Al-SBA-15 samples with different Si/Al acid sites (Brönsted) are generated by tetrahedrally coordinated Al
ratios are presented in Table 2. The broad desorption pattern indi- atoms forming Al-O(H)-Si bridges. Assuming that the strong acid
cates a large distribution of different types of acid sites which are sites are directly related to the Al ions in the samples, the ratio
NH3 /Al (mole NH3 desorbed per mole Al) was calculated which is
given in Table 2. The results indicate a decrease in dispersion of Al
Table 2
at higher Al loading.
Acidity of Al-SBA-15 with different Si/Al ratios by TPD.

Sample Total Acidity (mmol/g) NH3 /Ala (mmol/g)


3.6. FTIR of pyridine adsorption
Al-80 0.26 2.954
Al-40 0.28 1.823
Al-20 0.44 1.673 The spectra of pyridine desorbed on a typical Al-5 sample at
Al-10 0.15 0.495 different temperatures in the range 1700–1400 cm−1 is shown in
Al-5 0.10 0.118
Fig. S3 (Supplementary information). The sample gives the IR bands
a
Moles of NH3 desorbed /mol of Al based on strong acid sites. due to hydrogen-bonded pyridine (1446 and 1596 cm−1 ). The Al-5
P. Bhange et al. / Applied Catalysis A: General 400 (2011) 176–184 181

27
Fig. 6. Al MAS NMR spectra of Al-SBA-15 samples: (a) Al-40, (b) Al-20, (c) Al-10 and (d) Al-5.

sample shows the band at 1456 cm−1 which is attributed to pyri- with Si/Al ∼ 40 and 80 does not fall in the above trend due to the
dine adsorbed on Lewis acid sites and the band at 1547 cm−1 are due non-accessibility of the acid sites to the reactant molecules at low
to pyridine adsorbed on Brönsted acid sites. The band at 1491 cm−1 Al contents.
attributed to pyridine associated with both Lewis and Brönsted acid
sites which are also present in Al-5 sample which is due to the 3.7.2. Effect of reaction temperature
large agglomerates of surface Al2 O3 . The intensity of the bands at Fig. 8 illustrates the effect of reaction temperature (298–383 K)
1456 and 1542 cm−1 which is due to Lewis and Brönsted acid sites on the esterification of acetic acid with n-butanol over Al-20
decreases with increase in temperature. The relative intensity of sample. Catalytic activity increased with increasing reaction tem-
the band due to weakly held hydrogen-bonded pyridine is more perature. For Al-20, the reaction is faster and almost 95% conversion
compared to the other characteristics bands.

100
3.7. Catalytic reactions: esterification of acetic acid Al-5
Al-10
3.7.1. Effect of Si/Al ratio 90 Al-20
The variation of catalytic performance of Al-SBA-15 samples Al-40
Conversion (%)

with different Si/Al ratio has been compared in Fig. 7. The data on
80
conversion for esterification of acetic acid over Al-20 sample shows
highest catalytic activity with 95% conversion and 100% selectivity
towards n-butyl acetate. The acetic acid conversion decreases with 70
increasing Al content, which may be due to loss of Brønsted acid
sites that are responsible for the esterification reaction. Indeed, in
60
the case of Al-10 and Al-5 samples, the presence of an excess of
extra-framework Al2 O3 oxide particles that are formed is respon-
sible for the decrease of activity. The high activity of Al-20 sample 50
has been attributed to isolated, generally tetrahedrally coordinated,
framework Al species. Al-5 catalysts show negligible activity, which
40
had attributed to the large agglomerates of surface Al2 O3 which 1 2 3 4 5 6
known to inhibit esterification reactions by decreasing the num-
Run Time (h)
ber of accessible catalytic sites. From the study, it can be concluded
that not only the high acidity but also the dispersion of Al2 O3 and Fig. 7. The effect of Si/Al molar ratio on the esterification reaction of acetic acid with
its nature/coordination favors the reaction. The data of the sample n-butanol at 353 K.
182 P. Bhange et al. / Applied Catalysis A: General 400 (2011) 176–184

100 60

Al-5
50 Al-10
80

Conversion (%)
Al-20
Conversion (%)

Al-40
40

60 30

20
40
10

20 2 4 6 8 10 12
300 320 340 360 380 400
Run Time (h)
Temperature, K
Fig. 9. Variation of catalytic activity with different Si/Al molar ratio on the benzy-
Fig. 8. The effect of temperature on the esterification reaction of acetic acid with lation of anisole, at 373 K and 0.05 g of catalyst amount.
n-butanol over Al-20 catalyst.

is achieved within 1 h at 353 K. The catalytic activity does not affect sion at 1 h, was nearly the same from 83 to 79% indicating that these
with further increase in temperature to 383 K. The analysis of reac- catalyst can be easily recycled.
tion products showed that n-butyl acetate was obtained as a major
product with almost 100% selectivity. The conversion of acetic acid
was nearly the same with further increase in the reaction temper- 3.8. Benzylation of anisole
ature having nearly 100% selectivity.
3.8.1. Effect of Si/Al ratio
3.7.3. Effect of molar ratio of reactants The effects of different Si/Al ratios on the catalytic perfor-
Mole ratios of acetic acid to n-butanol were varied from 1:2, mance of the Al-SBA-15 are shown in Fig. 9. It can be seen that
1:4, 1:10 to 1:16 and the results are shown in Table 3. The con- increasing this ratio leads to improved conversion and selectivity
version increased from 93% to 98% with a change in mole ratio of in the benzylation of anisole. The best result was achieved with
acetic acid to n-butanol from 1:2 to 1:10. With further increase in Al-5 catalyst hence this catalyst was used for the further stud-
mole ratio of acetic acid to n-butanol, a decrease in conversion was ies. The Al content increases the conversion and also yield due to
observed. Decrease in conversion in higher molar ratio of acid to an increase in the number of active centers. It can be seen that
alcohol may be due to the saturation of the catalytic surface with after about 3 h (initial reaction time), the conversion of anisole
the alcohol. This confirms Eley–Rideal mechanism with chemisorp- increased from about 6 to 17% with an increase in Si/Al ratio from
tion of alcohol on the Brønsted acid sites of the catalyst. This result 40 to 5. This is due to the availability of large surface area and acid
is very similar to Kirumakki et al. [36,37] who used zeolite as the sites, which favors the dispersion of more active species. There-
catalyst over esterification of benzyl alcohol with acetic acid and fore, the accessibility of the large number of molecules of the
Jermy and Pandurangan [38], who used Al-MCM-41 catalyst over reactants to the catalyst surface is favored. The conversion and
esterification of acetic acid with n-butyl alcohol. selectivity of benzylanisole increases with temperature from 353 K
to 373 K over Al-5 catalyst. Therefore, increasing the temperature
is apparently favorable for the acceleration of the reaction. Fig. 9
3.7.4. Effect of various alcohols
shows that a maximum conversion of up to 58% can be obtained
The Effect of various alcohols on the esterification of acetic acid
with Al-5 at 373 K at the end of a 12 h reaction. On this cata-
using Al-20 catalyst was summarized in Table 3. The conversion of
lyst, ortho- and para-substituted benzylanisole were obtained as
acetic acid decreases with increase in the chain length of the alco-
products with 54% selectivity to p-benzylanisole, whereas meta-
hol. Recycling of the catalyst is an important aspect of any industrial
benzylanisole was not detected. The selectivity for benzylanisole
process. The resulting solid was then reused as catalyst for a new
increases and that for dibenzylether decreases with increase in the
reaction under the same conditions. After two cycles, the conver-
Si/Al ratio of Al-SBA-15. The formation of by-products like ben-
zyl ether was observed due to the auto-etherification of benzyl
Table 3 alcohol.
Conversion of acetic acid under various conditions over Al-20 catalysts.

Sr. no. Acetic acid:alcohol Alcohols Conv. (%)


3.8.2. Influence of run duration
1 1:2 Butanol 94.0
As shown in Fig. 10, the conversion and the selectivity of ortho-
2 1:4 Butanol 97.3
3 1:10 Butanol 98.0 and para-substituted benzylanisole increase with run time over the
4 1:16 Butanol 95.0 Al-5 catalyst. Conversion and yield both increase steadily up to
5 1:4 Butanol 97.3
12 h. Further increase in the reaction time, the conversion remains
6 1:4 Hexanol 98.4 constant, but the selectivity of benzylanisole decreased because of
7 1:4 Heptanol 98.0 the formation of consecutive alkylation products due to very high
8 1:4 Octanol 95.0 contact time. In this case besides the monobenzylation product,
9 1:4 Dodecenol 90.0
benzylether was observed but its concentration decreases with run
Procedure: 25 mg catalyst, time = 1 h, and temperature = 353 K. time.
P. Bhange et al. / Applied Catalysis A: General 400 (2011) 176–184 183

Table 4
Conversion of anisole to benzyl alcohol at various mole ratios over Al-5 catalysts.

Sr. no. Catalyst Anisol:benzyl alcohol Conv. (%) Selectivity

o-Benzylanisole p-Benzylanisole Dibenzyl ether

1. Al-5 10:1 32 50.2 49.6 0.26


2. 15:1 12 49.1 49.0 2.02
3. 20:1 9.0 49.6 49.0 1.76
4. Recycle-1 10:1 31 50.0 49.3 0.7
5. Recycle-2 10:1 29 50.1 49.4 0.5

Procedure: 10 mmol of anisole, 1 mmol benzyl alcohol, 0.1 g catalyst, and time = 3 h.

alytic sites for the reaction. The number of acid sites increased,
Total conversion
60 leading to the formation of more carbenium ions per unit of time
ortho-benzylanisole
para-benzylanisole which, in turn, increased the reaction rate.
Conversion and selectivity (%)

50 dibenzylether
3.8.4. Effect of mole ratio
The effect of mole ratio of anisole to benzyl alcohol on conver-
40 sion of anisole is shown in Table 4. The yield of the benzylanisole
increased with increase in the concentration of anisole upto 10:1
30 ratio. Further increase in anisole concentration does not affect
the conversion and selectivity of anisole. Recycling of Al-5 was
attempted at 373 K by simply separating the catalyst by filtration
20
followed by washing with acetone to remove products remaining
after the first run. As shown in Table 4, the conversion of benzyl
10 alcohol was maintained after recycling two times. No further reac-
tion occurred in the filtrated solution in the presence of residual
benzyl alcohol, indicating that the catalytic reaction truly occurred
0
2 4 6 8 10 12
over the solid catalyst.

Run Time (h)


4. Conclusions
Fig. 10. Influence of reaction time on ( ) the conversion of anisole, selectivity of
(䊉) o-benzylanisole, () p-benzylanisole and () dibenzylether plotted as a function The Al-SBA-15 samples were prepared under milder acidic con-
of time over Al-5 catalyst at 373 K. ditions by direct hydrothermal synthesis method. By adjusting the
H2 O/HCl molar ratio, Al gets incorporated into the lattice of SBA-15,
3.8.3. Effect of catalyst amount which is evidenced by XRD, TEM, 29 Si and 27 Al MAS NMR spec-
Fig. 11 shows the effect of catalysts amount on anisole conver- troscopic data. Aluminium incorporated SBA-15 samples retained
sion as a function of time over Al-SBA-15 samples. It can be seen both structural and textural properties of SBA-15. The 27 Al MAS
that the rate of conversion of anisole increases with an increase in NMR confirms that a large proportion of the Al is inserted into
catalyst loading for Al-5 catalyst. The highest activity was recorded tetrahedral positions within the framework. The Al-SBA-15 showed
when the concentration of Al-5 reached 0.1 g. Comparing 3 h data, good catalytic activity and selectivity in the esterification of acetic
the conversion and yield increase from 16% to 32% when the catalyst acid and in the benzylation of anisole. The high activity of Al-20
amount is increased from 0.05 g to 0.1 g on Al-5 catalyst. The higher sample in esterification reaction has been attributed to isolated,
the catalyst loading, the faster the rate of reaction was observed generally tetrahedrally coordinated, framework Al species. Higher
because of the increase in the total number of available active cat- conversion of acetic acid was achieved at moderate temperature
and with sufficient amount of catalyst. While in the case of benzyla-
tion reaction, the Al content increases the conversion and also yield
70
due to an increase in the number of active centers. The selectivity
for benzylanisole increases and that for dibenzylether decreases
60 with increase in the Si/Al ratio of Al-SBA-15 and run time.
0.05 g
Conversion (%)

0.1 g Appendix A. Supplementary data


50

Supplementary data associated with this article can be found, in


40 the online version, at doi:10.1016/j.apcata.2011.04.031.

30 References

[1] Y. Yue, A. Geı̌deı̌on, J.-L. Bonardet, N. Melosh, J.-B. D’Espinose, J. Fraissard, Chem.
20 Commun. (1999) 1967–1968.
[2] Z. Luan, M. Hartmann, D. Zhao, W. Zhou, L. Kevan, Chem. Mater. 11 (1999)
1621–1627.
[3] W.-H. Zhang, J. Lu, B. Han, M. Li, J. Xiu, P. Ying, C. Li, Chem. Mater. 14 (2002)
10
3413–3421.
2 4 6 8 10 12 [4] M. Cheng, Z. Wang, K. Sakurai, F. Kumata, T. Saito, T. Komatsu, T. Yashima, Chem.
Run Time (h) Lett. (1999) 131–132.
[5] Z. Luan, E.M. Maes, P.A.W. van der Heide, D. Zhao, R.S. Czernuszewicz, L. Kevan,
Fig. 11. Effect of catalyst loading on the benzylation of anisole over Al-5 catalyst at Chem. Mater. 11 (1999) 3680–3686.
373 K. [6] Z. Luan, J.Y. Bae, L. Kevan, Chem. Mater. 12 (2000) 3202–3207.
184 P. Bhange et al. / Applied Catalysis A: General 400 (2011) 176–184

[7] Y. Du, S. Liu, Y. Zhang, D. Li, F.-S. Xiao, in: R. Xu, Z. Gao, J. Chen, W. Yan (Eds.), [22] M.M. Sharma, React. Funct. Polym. 26 (1995) 3–23.
Studies in Surface Science and Catalysis, vol. 170B, 2007, pp. 734–1739. [23] T.L. Marker, G.A. Funck, T. Barger, U. Hammershaimb, US Patent 5,504,258,
[8] M. Hartmann, A. Vinu, S.P. Elangovan, V. Murugesan, W. Bohlmann, J. Chem. 1996.
Soc. Chem. Commun. (2002) 1238–1239. [24] D.E. Hendriksen, J.R. Lattner, M.J.G. Janssen, US Patent 6,002,057, 1999.
[9] S. Wu, Y. Han, Y. Zou, J. Song, L. Zhao, Y. Di, S. Liu, F.-S. Xiao, Chem. Mater. 16 [25] S.R. Chitnis, M.M. Sharma, React. Funct. Polym. 32 (1997) 93–115.
(2004) 486–492. [26] W. Chu, X. Yang, X. Ye, Y. Wu, Appl. Catal. A: Gen. 145 (1996) 125–129.
[10] Y. Han, Y. Sun, D. Li, F.-S. Xiao, J. Liu, X. Zhang, Chem. Mater. 14 (2002) [27] J. Michael Verhoef, J. Patrica Kooyman, A. Joop Peters, F.H. van Bekkum, Micro-
1144–1148. por. Mesopor. Mater. 27 (1999) 365–371.
[11] Y. Han, F.-S. Xiao, S. Wu, Y. Sun, X. Meng, D. Li, S. Lin, F. Deng, X. Ai, J. Phys. [28] A. Corma, H. Garcia, S. Iborra, J. Primo, J. Catal. 120 (1989) 78–87.
Chem. B 105 (2001) 7963–7966. [29] S. Garg, K. Soni, G. Muthu Kumaran, R. Bal, K. Gora-Marek, J.K. Gupta, L.D.
[12] A. Vinu, V. Murugesan, W. Böhlmann, M. Hartmann, J. Phys. Chem. B 108 (2004) Sharma, G. Murali Dhar, Catal. Today 141 (2009) 125–129.
11496–11505. [30] D. Srinivas, L. Saikia, Catal. Surv. Asia 12 (2008) 114–130.
[13] R. Ryoo, C.H. Ko, M. Kruk, V. Antochshuck, M. Jaroniec, J. Phys. Chem. B 104 [31] S. Chu, M.M.S. Wu, Y. Xiong, L.B. Yang (Mobil Oil Corporation, USA) Jpn, Kokai
(2000) 11465–11471. Tokyo Koho, 1999, p. 9. CODEN: JKXXAF JP 11181456 A 2 19990706, Applica-
[14] M. Kruk, M. Jaroniec, M.C.H. Ko, R. Ryoo, Chem. Mater. 12 (2000) 1961–1968. tion: JP 1997-304443 19971106.
[15] J. Sauer, F. Marlow, F. Schuth, Phys. Chem. Chem. Phys. 3 (2001) 5579–5584. [32] M.H.C. de la Cruz, J.F.C. da Silva, E.R. Lachter, Catal. Today 118 (2006) 379–384.
[16] A.H. Janssen, C.M. Yang, Y. Wang, F. Schüth, A.J. Koster, K.P. De Jong, J. Phys. [33] M.H.C. de la Cruz, M.A. Abdel-Rehim, A.S. Rocha, J.F.C. da Silva, A.C. Faro Jr., E.R.
Chem. B 107 (2003) 10552–10556. Lachter, Catal. Commun. 8 (2007) 1650–1654.
[17] J. Parmentier, S. Saadhallah, M. Reda, P. Gibot, M. Roux, L. Vidal, C. Vix-Guterl, [34] K. Okumura, K. Yamashita, M. Hirano, M. Niwa, J. Catal. 234 (2005) 300–307.
J. Patarin, J. Phys. Chem. Solids 65 (2004) 139–146. [35] M. Imperor-Clerc, P. Davidson, A. Davidson, J. Am. Chem. Soc. 122 (2000)
[18] J. Lilja, D.Y. Murzin, T. Salmi, J. Aumo, P. M¨aki-Arvela, M. Sundell, J. Mol. Catal. 11925–11933.
A: Chem. 182–183 (2002) 555–563. [36] S.R. Kirumakki, N. Nagaraju, K.V.V.S.B.S.R. Murthy, S. Narayanan, Appl. Catal. A:
[19] N. Sanchez, M. Martinez, J. Aracil, A. Corma, J. Am. Oil Chem. Soc. 69 (1992) Gen. 226 (2002) 175–182.
1150–1153. [37] S.R. Kirumakki, N. Nagaraju, K.V.R. Chary, S. Narayanan, Appl. Catal. A: Gen. 248
[20] W.T. Liu, C.S. Tan, Ind. Eng. Chem. Res. 40 (2001) 3281–3286. (2003) 161–167.
[21] M.R. Altiokka, A. Citak, Appl. Catal. A: Gen. 239 (2003) 141–148. [38] B. Rabindran Jermy, A. Pandurangan, J. Mol. Catal. A: Chem. 237 (2005) 146–154.

Vous aimerez peut-être aussi