Вы находитесь на странице: 1из 6

Journal of Industrial and Engineering Chemistry 20 (2014) 222227

Contents lists available at SciVerse ScienceDirect

Journal of Industrial and Engineering Chemistry


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

Catalytic dehydration of methanol to dimethyl ether (DME) over


Al-HMS catalysts
Behrouz Sabour a, Mohammad Hassan Peyrovi a,*, Touba Hamoule a, Mehdi Rashidzadeh b
a
b

Department of Chemistry, Faculty of Science, Shahid Beheshti University, G.C., P.O. Box 19396-4716, Tehran, Iran
Research Institute of Petroleum Industry (RIPI), Tehran 1485733111, Iran

A R T I C L E I N F O

Article history:
Received 16 January 2013
Accepted 30 March 2013
Available online 18 April 2013
Keywords:
Al-HMS
Acid catalyst
Methanol dehydration
Dimethyl ether

A B S T R A C T

A series of Al-HMS with different Si/Al ratio was used as a solid acid catalyst for methanol dehydration to
dimethyl ether (DME). The effect of temperature, feed composition, space velocity, and the catalyst Si/Al
ratio on the catalytic dehydration of methanol was investigated. By decreasing Si/Al, the temperature
required to reach equilibrium conversion of methanol decreased due to the increased number of acidic
sites. Compared to commercial g-Al2O3, Al-HMS-5 and Al-HMS-10, catalysts exhibited a high yield of
DME. Among all Al-HMS catalysts, Al-HMS-10 exhibited an optimum yield of 89% with 100% selectivity
and excellent stability for methanol dehydration to DME.
2013 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

1. Introduction
Dimethyl ether (DME) has received considerable attention due
to its potential uses as an alternative to diesel oil and LPG because
of similarity of its thermal efciency to that of traditional fuels,
low NOx emission, negligible smoke amounts, and less engine noise
[14]. It has already been used as an ozone-friendly aerosol
propellant to replace ozone destructive CFCs [5]. In addition, it is an
important intermediate for producing highly valuable chemicals
such as lower olens, dimethyl sulfate and methyl acetate [6,7].
DME can be produced directly from syngas over bifunctional
Cu-based catalysts or synthesized by methanol dehydration over
solid acid catalysts at 200450 8C in a pressure of up to 18 atm [8
12]. Among different kinds of solid acids used for methanol
dehydration, H-ZSM-5 and g-Al2O3 have been investigated
intensively both in laboratory and commercial scales. H-ZSM-5
has been reported to be more active than g-Al2O3, but rapid
deactivation occurs on its strong acidic sites due to the generation
of undesirable hydrocarbons [13,14]. On the other hand, g-Al2O3 is
more selective to DME, but it has some disadvantages including
low activity and rapid deactivation in the presence of water.
Water is both produced in signicant amounts through direct
synthesis of DME from synthesis gas, and it is also present in large
quantities (2030%) in the crude methanol. It is reported that
replacing pure methanol with crude methanol would bring about

* Corresponding author. Tel.: +98 21 29902892; fax: +98 21 22431663.


E-mail address: m-peyrovi@sbu.ac.ir (M.H. Peyrovi).

great industrial benets. Therefore, researchers are trying to


develop effective catalysts to optimize the DME production from
crud methanol and improve the catalyst stability [14,15].
Recently, ordered mesoporous materials have been the subject
of a large number of studies because of their high surface areas,
regular frameworks, and large pore size with narrow distribution
[16]. Pure mesoporous silica does not have enough acidity, but
acidity can be improved through the insertion of foreign metal ions
into its structure during the synthesis [1719]. Among such
materials, Al-incorporated mesoporous molecular sieves which
possess the acidic sites and good hydrothermal stability are more
favorable. HMS is a hexagonal mesoporous silica with a particular
wormlike pore structure. It has a simple preparation method using
cheap primary alkyl amines which can be extracted without
pollution [20]. Surprisingly, despite its wide uses for catalytic
reactions, to the best of our knowledge, aluminated hexagonal
mesoporous material (Al-HMS) has not been employed as an acidic
catalyst to the dehydration of methanol up to now.
In continuation of previous studies on the application of AlHMS mesoporous material as a solid acid catalyst for various
reactions [21,22], here we report catalytic behavior of this
material in the reaction of methanol dehydration to DME. The
main objective of this study is to investigate the effects of the
incorporation of Al into the HMS framework on the activity,
selectivity, and durability in the methanol dehydration. Also, the
effects of temperature, and feed composition on catalytic
performance were studied. BET, XRD, XRF, NH3-TPD, FT-IR
pyridine, and TG/DTA techniques were employed for the material
characterization.

1226-086X/$ see front matter 2013 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jiec.2013.03.044

B. Sabour et al. / Journal of Industrial and Engineering Chemistry 20 (2014) 222227

223

2. Experimental
2.1. Materials and methods
Al-HMS mesoporous materials with four different Si/Al ratios of
5, 10, 20, and 35 were synthesized via neutral templating pathway
similar to the procedure reported in previous studies [23] using
tetraethylorthosilicate (TEOS) as the silica source, aluminum
isopropoxide (Al(OPri)3) as the aluminum source and dodecylamine (DDA) as the surfactant. The molar composition of nal gel
was 1.0 SiO2:x Al(OPri)3:0.25 DDA:10 isopropyl alcohol:100 H2O,
where the value of x is dependent on various Si/Al ratios. Each solid
product was separated by ltration and dried at 110 8C overnight
and calcined at 540 8C for 6 h in the owing air. The prepared
samples were named Al-HMS(x), where x is the Si/Al ratio.
2.2. Characterization of catalysts
An X-ray diffraction (XRD) analysis of the calcined samples was
performed in the 2u range of 1108, by using an X-PERT
diffractometer employing Ni-ltered Cu Ka radiation at 45 kV
and 50 mA.
The specic surface area and pore volume of the samples were
measured in an ASAP-2010 Micromeritics (USA) using low
temperature N2 physisorption isotherms. Before the analysis,
the sample was evacuated at 350 8C under vacuum conditions.
The acidity of Al-HMS was measured by TPD of ammonia in a
TPD/TPR analyzer (2900 Micromeritics) with a thermal conductivity detector. To determine and analyze the type of acidic sites,
pyridine adsorption on the samples was performed on a Fouriertransform infrared spectrometer (170-SX). The Si/Al ratio of AlHMS was determined by XRF (XRF-8410 Rh 60 kV). The content of
coke laid down on the surface of used catalysts was measured by
thermogravimetric analyzer using a STA503M TG/DTA instrument.

Fig. 1. XRD patterns of Al-HMS catalysts with different Si/Al ratio.

2.3. Catalytic evaluation


Vapor phase dehydration of methanol was carried out at the
temperature range of 250400 8C and atmospheric pressure in a
continuous xed-bed micro-reactor packed with 0.5 g of the
catalyst. Prior to each experiment, the catalyst was pretreated for
1 h at 300 8C in an N2 ow. Nitrogen saturated by pure methanol
(11% MeOH in N2) was used as feed with the space velocity (WHSV)
1.0 h 1.
Moreover,
methanolwater
mixture
(methanol
80 mol% + water 20 mol%) was introduced to the reactor under
aforementioned conditions to investigate the capability of Al-HMS
as a dehydration catalyst for the crude methanol dehydration. The
performance of the catalysts was measured after 0.5 h time on
stream (TOS) at noted temperatures for each experiment. The
analysis of the reaction products was carried out by on-line gas
chromatography using a gas chromatograph (Shimadzu-8A)
equipped with a thermal detector.
3. Results and discussion
3.1. Characterization
Fig. 1 shows that the XRD patterns of the Al-HMS materials are
similar to those reported in literature [20,24]. There is a single
broad reection that can be assigned to a lattice with the shortrange hexagonal symmetry. The increase of Al content in the
samples results in the broadening of the peak, indicating that
the incorporation of Al is associated with increasing the lattice
disorder.
The chemical compositions, BET surface area, and pore volume
of calcined Al-HMS materials are given in Table 1. It can be seen

Fig. 2. NH3-TPD on Al-HMS catalysts with different Si/Al ratio.

that the actual Si/Al ratios are very similar to added metal amounts
in the gel compositions, suggesting that most of the added
aluminum heteroatoms are embedded into the HMS bulk. It can be
inferred from Table 1 that the surface area and pore volume
decrease with the increase of Al amounts. This may be attributed to
decrease in the structural order of the samples as a result of Al
incorporation as proved by XRD.
Fig. 2 shows NH3-TPD proles of all Al-HMS materials with
different Si/Al ratios. The asymmetric shapes of the desorption
proles indicate the presence of different surface acid sites in the
range of 150500 8C, corresponding to the distribution of acid sites
from weak to strong. The maximum of desorption peak is in the
range of 250300 8C corresponding to the medium acid sites
responsible for the selective formation of DME [25,26]. The TPD
prole of pure HMS shows no evident peak due to the absence of
acidic sites on the HMS (not shown here). Table 1 shows that the
concentration of acidic sites of Al-HMS increases with the decrease
of Si/Al ratios. It can be seen that the maximum of the TPD diagram
shifts to the high temperatures with the decreasing Si/Al ratios. It is

B. Sabour et al. / Journal of Industrial and Engineering Chemistry 20 (2014) 222227

224

Table 1
Physicochemical characteristics of the Al-HMS catalysts.
Sample

Si/Al

Surface area (m2/g)

Pore volume
(cm3/g)

Peak temperature
of TPD (8C)

Acidity
(mmol NH3/g)

B/L

nsted acidity
BrO
(mmol NH3/g)

Lewis acidity
(mmol NH3/g)

Al-HMS-35
Al-HMS-20
Al-HMS-10
Al-HMS-5

36.3
18.7
9.6
4.5

1370
1289
1109
870

1.7
2.1
2.0
0.79

258.9
270.9
278.4
287.1

0.426
0.845
1.398
1.556

1.8
1.37
1.06
0.86

0.277
0.489
0.720
0.719

0.153
0.355
0.677
0.837

obvious that the strength of the acidic sites enhances in high Al


content [21].
In order to investigate the Bronsted and Lewis acid sites, the FTIR spectra of pyridine adsorption were recorded at room
temperature (Table 1). As it can be seen, the B/L acid site ratio
declines with increasing Al content. It is in good agreement with
the results reported previously; i.e. the tetrahedrally coordinated
framework aluminum (potential Bronsted acid site) decreases with
increasing Al content while octahedrally extra framework aluminum (potential Lewis acid site) increases [27]. The number of
Bronsted and Lewis acid sites was calculated from the NH3-TPD
results by using the B/L ratios (Table 1).
3.2. Catalytic performance
3.2.1. Activity
According to the literature, methanol dehydration is an acid
catalyzed reaction and both Lewis and Bronsted acid sites are
active in this reaction [2831]. Fig. 3 illustrates the catalytic
performance of a series of Al-HMS on methanol vapor phase
dehydration at atmospheric pressure in the temperature range of
250400 8C and WHSV = 1 h 1. From Fig. 3a it can be seen that
methanol conversion increases with the increase in reaction
temperatures, and after reaching maximum conversion, it shows a
slight decrease due to thermodynamic limitations as the
dehydration of methanol is slightly an exothermic reaction
[14,29]. With the decrease in Si/Al ratios, the temperature
required to achieve equilibrium conversion decreases due to
the increased number of acidic sites which are favorable for
methanol dehydration. It is apparent from Fig. 3a that the
conversion over Al-HMS-20 and Al-HMS-35 did not reach
equilibrium conversion at reaction conditions since these
catalysts possessed low number of acid sites as conrmed by
NH3-TPD. However, the conversion of methanol on Al-HMS-5 and
Al-HMS-10 exceeded the equilibrium values predicted by
thermodynamics for dehydration reaction in temperature range
of 275325 8C. Similar results were reported by other authors
previously using molecular sieves, e.g. HZSM5 and H-Mordenite
as an acid catalyst for methanol dehydration [26,32]. In the case of
our catalysts, easily separation of product molecules over Al-HMS
molecular sieve may explains the reason why conversion of
methanol exceeded the equilibrium values predicted by thermodynamics for reaction in this temperature range. However, the
conversions at higher temperature follow the values predicted by
thermodynamics since dehydration of methanol is an exothermic
reaction and increasing in temperature results in a shift of the
equilibrium composition toward the left direction.
Fig. 3b shows the calculated reaction rates per surface area as a
function of reaction temperature. The rate of the methanol
dehydration reaction depends strongly on the acidity of catalyst
employed. It can be seen that the order of reaction rates is Al-HMS5 > Al-HMS-10 > Al-HMS-20 > Al-HMS-35 which is in the same
order as the increase in acidic sites density.
Despite the fact that the mesoporosity of Al-HMS samples
decreased slightly with increasing Al content, these catalysts
showed higher surface area and pore volume compared to
commercial acid catalysts applied for methanol dehydration [9,14].

Fig. 3. Catalytic performance of Al-HMS catalysts versus temperature (0.5 g of


catalyst and WHSV = 1 h 1). (a) Methanol conversion proles; (b) activity per
surface area proles; (c) DME selectivity proles; (d) yield of DME proles.

B. Sabour et al. / Journal of Industrial and Engineering Chemistry 20 (2014) 222227


30

Table 2
Catalytic performance of synthesized catalysts as compared to commercial g-Al2O3
catalyst at 300 8C.
Methanol conversion (%)

DME selectivity (%)

DME yield (%)

g-Al2O3
Al-HMS-10
Al-HMS-5

76.2
89
91.2

100
100
94.1

76.2
89
85.8

Al-HMS-10 (P)

Al-HMS-10 (C)

Al-HMS-20 (P)

28

Reaction rate *10 -3

Sample

Al-HMS-5 (P)

225

26

24

22

As compared to a commercial g-Al2O3 with BET surface area of


202 m2/g and surface acidity of 0.728 mmol NH3/g, the catalytic
activity of our synthesized catalysts was much higher (Table 2).
The lower activity of g-Al2O3 can be attributed to the lower
number of acid sites as well as lower surface area.
It seems that a combination of high porosity and high surface
area increases the accessibility of methanol to active sites and
consequently enhances the catalyst activity. It is noteworthy that
pure HMS showed no noticeable methanol conversion even at
higher temperatures. Obviously, this incompetence of HMS for this
process is due to its lower acidity according to NH3-TPD results.
3.2.2. Selectivity and yield
The effects of temperature and Si/Al ratios of the catalysts on
the selectivity and yield of DME were also investigated. As
presented in Fig. 3c, while the selectivity of DME is almost 100% on
Al-HMS-35 and Al-HMS-20 in the temperature range of 250
400 8C, Al-HMS-10 and Al-HMS-5 follow different trends. The
selectivity of Al-HMS-10 decreases slightly from 100% to 92% with
the increase in temperature up to 400 8C, whereas the increase in
temperatures causes drastic change in Al-HMS-5 selectivity,
decreasing from 100% to 45% toward DME as the temperature
rises from 250 8C to 400 8C. Illustrated in Fig. 3d, although a
relatively similar yield of DME was obtained over Al-HMS-5 and
Al-HMS-10, keeping approximately 100% selectivity of DME up to
higher temperatures in the case of Al-HMS-10 can be considered its
main advantage over Al-HMS-5. The results revealed that higher
yields (>80%) have been preserved in a wide temperature range of
275350 8C over Al-HMS-10, which is included the operating
temperature range for the direct synthesis of DME from synthesis
gas [33]. The by-products were CH4 and light olens mostly. This
can be interpreted by the fact that acid sites with relatively weak
and intermediate strengths are favorable in methanol dehydration
reaction while strong acid sites catalyze methanol to hydrocarbon
reaction, especially at higher temperatures [9,14,34]. This trend is
in good accord with the increase in the acidity strength shown in
Table 1.

20
0

10

20

30

40

50

60

70

Time on stream (h)


Fig. 4. Long-term test of catalyst samples at 300 8C; (P): tested for pure methanol
feed; (C): tested for crude methanol feed.

3.2.4. Effect of water


Replacing pure methanol by crude methanol (containing
20 mol% water) as a feed for the methanol dehydration to DME
would reduce the process cost [14]. Thus, crude methanol was
introduced to the reactor in the same conditions to investigate the
capability of our catalysts for crude methanol dehydration. Fig. 5
shows that the changes in methanol conversion for pure methanol
feed and crude methanol feed after 72 h time on stream. As seen,
the less Si/Al ratio catalysts have, the more decline in activity they
show for crude methanol dehydration.
The XRD patterns and BET data of the used catalysts for crude
methanol after 72 h reaction at 300 8C were taken (not shown here)
and proved that the structural and textural properties of all used
catalysts had been preserved, and it showed no remarkable
changes in the structural characteristics. As a result, Al-HMS
catalysts have a good hydrothermal stability, and a decrease in
catalyst activity and durability in the crude methanol dehydration
are not imposed by the changes of catalyst structure. The slight
decrease in Al-HMS conversion in the crude methanol feed
indicated the role of different types of acidic sites. It is wellknown that water is adsorbed preferentially on Lewis acid sites
rather than Bronsted sites. Based on FT-IR results given in Table 1,
lower Si/Al ratio leads to an increase in Lewis acid sites, and this
explains why Al-HMS-5 activity declines during the crude
methanol dehydration. A similar trend was reported when gAl2O3 and modied H-ZSM-5 were used for dehydration of crude
methanol. It was proved that water had more negative effects on
the performance of g-Al2O3 than on that of ZSM-5 since ZSM-5 has

100

Pure Methanol

crude methanol

90

Methanol conversion (%)

3.2.3. Stability
Because dehydration catalysts cannot be regenerated in situ,
keeping high stability during a long-term reaction is very
important; therefore, based on the experimental results, AlHMS-5, Al-HMS-10 and Al-HMS-20, which showed a high activity,
were chosen to be surveyed for durability tests. As it is indicated in
Fig. 4, Al-HMS-10 and Al-HMS-20 virtually keep their activity with
time on stream and show no remarkable changes during the 72 h
time on stream. However, the activity of Al-HMS-5 showed a slight
decrease with time-on-stream, which might be due to coke
formation on its strong acid sites and/or the blocking of strong acid
sites by water produced during the reaction [8,33].
Generally speaking, even though Al-HMS-10 showed a slightly
lower activity (89%) in comparison with Al-HMS-5 (91.2%) at
300 8C, combining its high activity, excellent selectivity (100%) to
DME in the wide range of operating temperature, and high
resistance to deactivation makes it the best catalyst for methanol
dehydration.

80
70
60
50
40
30
20
10
0

Al-HMS-20

Al-HMS-10

Al-HMS-5

Fig. 5. Methanol conversion variations for pure methanol and crude methanol
(reaction conditions were 300 8C, 0.5 g of catalyst, reaction time = 72 h, and
WHSV = 1 h 1).

B. Sabour et al. / Journal of Industrial and Engineering Chemistry 20 (2014) 222227

226

100

(a)

-20

TGA

-4

DTA

-6

-40

-8
-10

rel. DTA

Weight (%)

-2

-60

-12

Methanol conversion (%)

90
80
Al-HMS-10

70
-Al2O3

60

-80

-14
0

100

200

300

400

500

600

700

50

800

Temperature (o C)
2

60

(b)

Fig. 7. Effect of WHSV on methanol conversion over Al-HMS-10 and commercial gAl2O3 at 300 8C.

40

-2

20

-4

TGA
DTA

-6

-20

-8

-40

-10

-60

-12
0

100

200

300

400

500

600

700

rel. DTA (%)

Weight (%)

WHSV h-1

-80
800

Temperature (o C)
Fig. 6. TG/DTA proles of used Al-HMS-10 for dehydration of (a) pure methanol and
(b) crude methanol at 300 8C after 72 h.

higher resistance due to its hydrophobic properties resulting from


its high SiO2/Al2O3 ratio [15]. Based on the above discussion, AlHMS-10 appears to be a promising catalyst for the crude methanol
dehydration. Additionally, coke laid down over Al-HMS-10 catalyst
used for dehydration of both pure and crude methanol at 300 8C
after 72 h time on stream, was measured by TG/DTA technique and
results were shown in Fig. 6.
Both used catalysts showed two weight losses in TGA curves.
The weight losses at <200 8C corresponding to endothermic peaks
in DTA curves, were attributed to desorption of physically
adsorbed water. The weight losses at >400 8C, corresponding to
exothermic peaks in DTA curves, were ascribed to the combustion
of carbonaceous material deposited inside used Al-HMS-10
catalysts. The amount of coke analyzed by TGA technique is also
listed in Table 3. It is well-known that catalysts with less stability
possess relatively high amount of carbon deposit. As it can be seen,
Al-HMS-10 possess high level of coke during dehydration of pure
methanol. In other words using crude methanol instead of pure
methanol decrease mainly the total amount of coke on the surface
of the catalyst. As a result relatively more stable catalytic behavior
was seen using crude methanol owning to attenuation of coke
deposition on Al-HMS-10 (Fig. 4). It is clear from above discussion
that water inhibits coke formation and remove coke easily over
mesoporous molecular sieves [35].
3.2.5. Effect of WHSV
The methanol conversion to DME under various WHSV (1
4 h 1) over Al-HMS-10 catalyst, which showed optimum activity,
was investigated and compared to commercial g-Al2O3.
Table 3
Effect of different feed on coke formation over Al-HMS-10 after 72 h.
Sample

Temperature
range (8C)

Coke content
(wt.%)

Al-HMS-10 (used for pure methanol)


Al-HMS-10 (used for crude methanol)

400600
445520

3.14
0.9

As it is depicted in Fig. 7, an increase in WHSV causes no


signicant reduction in methanol conversion over Al-HMS-10,
while methanol conversion over commercial g-Al2O3 reduces
signicantly. This can be interpreted by the fact that Al-HMS-10
possesses a high surface area as well as high number of medium
strength acid sites responsible for selective dehydration of
methanol to DME. It should be noted that the selectivity of DME is
100% under reaction conditions.
4. Conclusion
Several Al-HMS materials with Si/Al ratio of 535 were
synthesized and studied as novel acid catalysts for methanol
dehydration reaction. According to the NH3-TPD results, Al
incorporation into HMS framework enhanced number of surface
acid sites as well as acidity strength. It was found that increasing Al
content is followed by increasing conversion of methanol and
decreasing selectivity to DME. The synthesized samples showed
good catalytic performance in the presence of water. Al-HMS-10,
as the best catalyst, exhibited optimum activity of 89% with the
DME selectivity of 100% and high resistance to deactivation.
Acknowledgments
The authors gratefully acknowledge nancial support from the
Research Council of Shahid Beheshti University and Catalyst Center
of Excellence (CCE).
References
[1] T.A. Semelsberger, R.L. Borup, H.L. Greene, Journal of Power Sources 156 (2006)
497.
[2] C. Arcoumanis, C. Bae, R. Crookes, E. Kinoshita, Fuel 87 (2008) 1014.
[3] C.-J. Yang, R.B. Jackson, Energy Policy 41 (2012) 878.
[4] T.H. Fleisch, A. Basu, R.A. Sills, Journal of Natural Gas Science and Engineering 9
(2012) 94.
[5] Y. Onaka, A. Miyara, K. Tsubaki, International Journal of Refrigeration 33 (2010)
1277.
[6] M. Bjrgen, S. Akyalcin, U. Olsbye, S. Benard, S. Kolboe, S. Svelle, Journal of
Catalysis 275 (2010) 170.
[7] R.B. Diemer, W.L. Luyben, Industrial and Engineering Chemistry Research 49
(2010) 12224.
[8] F. Raoof, M. Taghizadeh, A. Eliassi, F. Yaripour, Fuel 87 (2008) 2967.
[9] F. Yaripour, F. Baghaei, I. Schmidt, J. Perregaard, Catalysis Communications 6
(2005) 147.
[10] S. Alamolhoda, M. Kazemeini, A. Zaherian, M.R. Zakerinasab, Journal of Industrial
and Engineering Chemistry 18 (2012) 2059.
[11] L. Liu, W. Huang, Z.-H. Gao, L.-H. Yin, Journal of Industrial and Engineering
Chemistry 18 (2012) 123.
[12] J.W. Bae, S.-H. Kang, Y.-J. Lee, K.-W. Jun, Journal of Industrial and Engineering
Chemistry 15 (2009) 566.
[13] Y.-J. Lee, J.M. Kim, J.W. Bae, C.-H. Shin, K.-W. Jun, Fuel 88 (2009) 1915.
[14] S.D. Kim, S.C. Baek, Y.-J. Lee, K.-W. Jun, M.J. Kim, I.S. Yoo, Applied Catalysis A
General 309 (2006) 139.
[15] V. Vishwanathan, K.-W. Jun, J.-W. Kim, H.-S. Roh, Applied Catalysis A General
276 (2004) 251.

B. Sabour et al. / Journal of Industrial and Engineering Chemistry 20 (2014) 222227


[16] J. Cejka, S. Mintova, Catalysis Reviews 49 (2007) 457.
[17] K. Szczodrowski, B. Prelot, S. Lantenois, J.-M. Douillard, J. Zajac, Microporous and
Mesoporous Materials 124 (2009) 84.
[18] M.J. Gracia, E. Losada, R. Luque, J.M. Campelo, D. Luna, J.M. Marinas, A.A. Romero,
Applied Catalysis A General 349 (2008) 148.
[19] K.C. Tokay, T. Dogu, G. Dogu, Chemical Engineering Journal 184 (2012) 278.
[20] T. Chiranjeevi, G.M. Kumaran, J.K. Gupta, G.M. Dhar, Thermochimica Acta 443
(2006) 87.
[21] T. Hamoule, M.H. Peyrovi, M. Rashidzadeh, M.R. Toosi, Catalysis Communications
16 (2011) 234.
[22] M.H. Peyrovi, T. Hamoule, B. Sabour, M. Rashidzadeh, Journal of Industrial and
Engineering Chemistry 18 (2012) 986.
[23] R. Mokaya, W. Jones, Journal of Catalysis 172 (1997) 211.
[24] R. Mokaya, W. Jones, Chemical Communications 0 (6) (1996) 981.
[25] F. Yaripour, F. Baghaei, I. Schmidt, J. Perregaard, Catalysis Communications 6
(2005) 542.
[26] N. Khandan, M. Kazemeini, M. Aghaziarati, Applied Catalysis A General 349
(2008) 6.

227

[27] A. Yin, X. Guo, W.-L. Dai, K. Fan, Journal of Physical Chemistry C 114 (2010)
8523.
[28] A.I. Osman, J.K. Abu-Dahrieh, D.W. Rooney, S.A. Halawy, M.A. Mohamed, A.
Abdelkader, Applied Catalysis B Environmental 127 (2012) 307.
[29] Q. Tang, H. Xu, Y. Zheng, J. Wang, H. Li, J. Zhang, Applied Catalysis A General 413/
414 (2012) 36.
[30] Y. Fu, T. Hong, J. Chen, A. Auroux, J. Shen, Thermochimica Acta 434 (2005) 22.
[31] A. Ciftci, D. Varisli, K. Cem Tokay, N. Asl Sezgi, T. Dogu, Chemical Engineering
Journal 207/208 (2012) 85.
[32] S. Hassanpour, F. Yaripour, M. Taghizadeh, Fuel Processing Technology 91 (2010)
1212.
[33] M. Xu, J.H. Lunsford, D.W. Goodman, A. Bhattacharyya, Applied Catalysis A
General 149 (1997) 289.
[34] A.A. Rownaghi, F. Rezaei, M. Stante, J. Hedlund, Applied Catalysis B Environmental 119/120 (2012) 56.
[35] K.W. Jun, H.S. Lee, H.S. Roh, S.E. Park, Bulletin of the Korean Chemical Society 24
(2003) 106.

Вам также может понравиться