You are on page 1of 28

Accepted Manuscript

Title: Catalytic dehydration of methanol to dimethyl ether


(DME) over Al-HMS Catalysts

Author: Behrouz Sabour Mohammad Hassan Peyrovi Touba


Hamoule Mehdi Rashidzadeh

PII: S1226-086X(13)00150-0
DOI: http://dx.doi.org/doi:10.1016/j.jiec.2013.03.044
Reference: JIEC 1312

To appear in:

Received date: 16-1-2013


Revised date: 29-3-2013
Accepted date: 30-3-2013

Please cite this article as: B. Sabour, M.H. Peyrovi, T. Hamoule, M.


Rashidzadeh, Catalytic dehydration of methanol to dimethyl ether (DME) over
Al-HMS Catalysts, Journal of Industrial and Engineering Chemistry (2013),
http://dx.doi.org/10.1016/j.jiec.2013.03.044

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Catalytic dehydration of methanol to dimethyl ether (DME) over Al-HMS
Catalysts

Behrouz Saboura, Mohammad Hassan Peyrovi*a, Touba Hamoulea[b1], Mehdi Rashidzadehb


a
Department of Chemistry, Faculty of Science, Shahid Beheshti University, G. C., P. O.

t
Box 19396-4716, Tehran, Iran

ip
b
Research Institute of Petroleum Industry (RIPI), Tehran, 1485733111, Iran

cr
us
an
M
d
p te
ce
Ac

*Corresponding author; E-mail: m-peyrovi@sbu.ac.ir


Phone: +98 21 29902892;
Fax: +98 21 22431663

Page 1 of 27
Abstract

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

t
ip
methanol was investigated. By decreasing Si/Al, the temperature required to reach

equilibrium conversion of methanol also decreased due to the increased number of acidic

cr
sites. Compared to commercial γ-Al2O3, Al-HMS-5 and Al-HMS-10, catalysts exhibited a

us
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.

an
M
d
p te
ce
Ac

Keywords: Al-HMS, Acid catalyst, Methanol dehydration, Dimethyl ether

Page 2 of 27
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 efficiency to that of

traditional fuels, low NOx emission, negligible smoke amounts, and less engine noise [1-

t
ip
4][b2]. It has already been used as an ozone-friendly aerosol propellant to replace ozone

destructive CFCs [5][b3]. In addition, it is an important intermediate for producing highly

cr
valuable chemicals such as lower olefins, dimethyl sulfate and methyl acetate [6-7].

us
DME can be produced directly from syngas over bifunctional Cu-Based catalysts or

synthesized by methanol dehydration over solid acid catalysts at 200-450 oC in a pressure

an
of up to 18 atm [8-12][b4]. Among different kinds of solid acids used for methanol

dehydration, H-ZSM-5 and γ-Al2O3 have been investigated intensively both in laboratory
M
and commercial scales. H-ZSM-5 has been reported to be more active than γ-Al2O3,but

rapid deactivation occurs on its strong acidic sites due to the generation of undesirable
d

hydrocarbons [13-14]. On the other hand, γ -Al2O3 is more selective to DME, but it has
te

some disadvantages including low activity and rapid deactivation in the presence of water.

Water is both produced in significant amounts through direct synthesis of DME from
p

synthesis gas, and it is also present in large quantities (20-30%) in the crude methanol. It
ce

is reported that replacing pure methanol with crude methanol would bring about great
Ac

industrial benefits. 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

Page 3 of 27
improved through the insertion of foreign metal ions into its structure during the synthesis

[17-19][b5]. 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

t
preparation method using cheap primary alkyl amines which can be extracted without

ip
pollution [20]. Surprisingly, despite its wide uses for catalytic reactions, to the best of our

cr
knowledge, aluminated hexagonal mesoporous material (Al-HMS) has not been employed

as an acidic catalyst to the dehydration of methanol up to now.

us
In continuation of previous studies on the application of Al-HMS mesoporous material as a

solid acid catalyst for various reactions [21-22], here we report catalytic behavior of this

an
material in the reaction of methanol dehydration to DME. The main objective of this study
M
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
d

temperature, and feed composition on catalytic performance were studied. BET, XRD,
te

XRF, NH3-TPD, FT-IR Pyridine, and TG/DTA techniques were employed for the material

characterization.
p
ce

2. Experimental

2.1. Materials and method


Ac

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 final 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.
4

Page 4 of 27
Each solid product was separated by filtration and dried at 110 oC overnight and calcined

at 540 oC for 6 h in the flowing air. The prepared samples were named Al-HMS(x), where

x is the Si/Al ratio.

2.2. Characterization of catalysts

t
ip
An X-ray diffraction (XRD) analysis of the calcined samples was performed in the 2-

cr
range of 1-10°, by using an X-PERT Diffractometer employing Ni-filtered Cu kα radiation

at 45 kV and 50mA.

us
The specific surface area and pore volume of the samples were measured in an ASAP-

2010 Micromeritics (USA) using low temperature N2 physisorption isotherms. Before the

an
analysis, the sample was evacuated at 350 oC under vacuum conditions.

The acidity of Al-HMS was measured by TPD of ammonia in a TPD/TPR analyzer (2900
M
Micromeritics) with a thermal conductivity detector. To determine and analyze the type of

acidic sites, pyridine adsorption on the samples was performed on a Fourier-transform


d

infrared spectrometer (170-SX). The Si/Al ratio of Al-HMS was determined by XRF
te

(XRF-8410 Rh 60 kV). The content of coke laid down on the surface of used catalysts was
p

measured by thermogravimetric analyzer using a STA503M TG/DTA instrument.


ce

2.3. Catalytic evaluation


Ac

Vapor phase dehydration of methanol was carried out at the temperature range of 250-400
o
C and atmospheric pressure in a continuous fixed-bed micro-reactor packed with 0.5 g of

the catalyst. Prior to each experiment, the catalyst was pretreated for 1 h at 300 oC in an N2

flow. Nitrogen saturated by pure methanol (11% MeOH in N2) was used as feed with the

space velocity (WHSV) 1.0 h−1. Moreover, methanol-water mixture (methanol 80 mol% +

water 20 mol%) was introduced to the reactor under aforementioned conditions to

Page 5 of 27
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)

t
equipped with a thermal detector.

ip
3. Results and discussion

cr
3.1. Characterization

us
Figure 1 shows that the XRD patterns of the Al-HMS materials are similar to those

an
reported in literature [20, 24]. There is a single broad reflection that can be assigned to a

lattice with the short-range hexagonal symmetry. The increase of Al content in the samples
M
results in the broadening of the peak, indicating that the incorporation of Al is associated

with increasing the lattice disorder.


d
te

Fig. 1.
p
ce

The chemical compositions, BET surface area, and pore volume of calcined Al-HMS

materials are given in Table 1. It can be seen that the actual Si/Al ratios are very similar to
Ac

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.

Page 6 of 27
Table 1

Figure 2 shows NH3-TPD profiles of all Al-HMS materials with different Si/Al ratios. The

asymmetric shapes of the desorption profiles indicate the presence of different surface acid

t
sites in the range of 150-500 oC, corresponding to the distribution of acid sites from weak

ip
to strong. The maximum of desorption peak is in the range of 250-300 oC corresponding to

cr
the medium acid sites responsible for the selective formation of DME [25-26]. The TPD

profile of pure HMS shows no evident peak due to the absence of acidic sites on the HMS

us
(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

an
shifts to the high temperatures with the decreasing Si/Al ratios. It is obvious that the
M
strength of the acidic sites enhances in high Al content [21].
d

Fig. 2.
te

In order to investigate the Brönsted and Lewis acid sites, the FT-IR spectra of pyridine
p

adsorption were recorded at room temperature (Table 1). As it can be seen, the B/L acid
ce

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


Ac

Brönsted acid site) decreases with increasing Al content while octahedrally extra

framework aluminum (potential Lewis acid site) increases [27]. The number of Brönsted

and Lewis acid sites was calculated from the NH3-TPD results by using the B/L ratios

(Table 1).

Page 7 of 27
3.2. Catalytic performance

3.2.1. Activity

According to the literature, methanol dehydration is an acid catalyzed reaction and both

Lewis and Brönsted acid sites are active in this reaction [28-31][b6]. Figure 3 illustrates the

t
ip
catalytic performance of a series of Al-HMS on methanol vapor phase dehydration at

cr
atmospheric pressure in the temperature range of 250-400 oC and WHSV=1 h-1. From Fig.

3a it can be seen that methanol conversion increases with the increase in reaction

us
temperatures, and after reaching maximum conversion, it shows a slight decrease due to

thermodynamic limitations as the dehydration of methanol is slightly an exothermic

an
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
M
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
d

since these catalysts possessed low number of acid sites as confirmed by NH3-TPD.
te

However, the conversion of methanol on Al-HMS-5 and Al-HMS-10 exceeded the


p

equilibrium values predicted by thermodynamics for dehydration reaction in temperature


ce

range of 275-325 oC. Similar results were reported by other authors previously using

molecular sieves e.g. HZSM5 and H-Mordenite as an acid catalyst for methanol
Ac

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

Page 8 of 27
increasing in temperature results in a shift of the equilibrium composition toward the left

direction.

Fig. 3.

t
ip
Fig. 3b shows the calculated reaction rates per surface area as a function of reaction

cr
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-HMS-5  Al-

us
HMS-10  Al-HMS-20  Al-HMS-35 which is in the same order as the increase in acidic

sites density.

an
Despite the fact that the mesoporosity of Al-HMS samples decreased slightly with
M
increasing Al content, these catalysts showed higher surface area and pore volume

compared to commercial acid catalysts applied for methanol dehydration[9, 14].

As compared to a commercial γ-Al2O3 with BET surface area of 202 m2/g and surface
d
te

acidity of 0.728 mmol NH3/g, the catalytic activity of our synthesized catalysts was much

higher (Table 2). The lower activity of γ-Al2O3 can be attributed to the lower number of
p

acid sites as well as lower surface area.


ce
Ac

Table 2

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.


9

Page 9 of 27
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 oC, Al-

t
ip
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 oC, whereas the

cr
increase in temperatures causes drastic change in Al-HMS-5 selectivity, decreasing from

us
100% to 45% toward DME as the temperature rises from 250 oC to 400 oC. Illustrated in

Fig. 3d, although a relatively similar yield of DME was obtained over Al-HMS-5 and Al-

an
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
M
revealed that higher yields (>80%) have been preserved in a wide temperature range of 275
o
C -350 oC over Al-HMS-10, which is included the operating temperature range for the
d

direct synthesis of DME from synthesis gas[33]. The by-products were CH4 and light
te

olefins mostly. This can be interpreted by the fact that acid sites with relatively weak and
p

intermediate strengths are favorable in methanol dehydration reaction while strong acid
ce

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.
Ac

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, Al-

HMS-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

10

Page 10 of 27
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].

t
ip
Fig. 4.

cr
us
Generally speaking, even though Al-HMS-10 showed a slightly lower activity (89%) in

comparison with Al-HMS-5 (91.2%) at 300 oC, combining its high activity, excellent

an
selectivity (100%) to DME in the wide range of operating temperature, and high resistance

to deactivation makes it the best catalyst for methanol dehydration.


M
3.2.4. Effect of Water
d

Replacing pure methanol by crude methanol (containing 20 mol% water) as a feed for the
te

methanol dehydration to DME would reduce the process cost [14]. Thus, crude methanol
p

was introduced to the reactor in the same conditions to investigate the capability of our
ce

catalysts for crude methanol dehydration. Figure 5 shows that the changes in methanol

conversion for pure methanol feed and crude methanol feed after 72 h time on stream. As
Ac

seen, the less Si/Al ratio catalysts have, the more decline in activity they show for crude

methanol dehydration.

Fig. 5.

11

Page 11 of 27
The XRD patterns and BET data of the used catalysts for crude methanol after 72 h

reaction at 300 oC 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

t
stability, and a decrease in catalyst activity and durability in the crude methanol

ip
dehydration are not imposed by the changes of catalyst structure. The slight decrease in Al-

cr
HMS conversion in the crude methanol feed indicated the role of different types of acidic

sites. It is well-known that water is adsorbed preferentially on Lewis acid sites rather than

us
Brönsted sites. Based on FTIR results given in Table1, lower Si/Al ratio leads to an

increase in Lewis acid sites, and this explains why Al-HMS-5 activity declines during the

an
crude methanol dehydration. A similar trend was reported when γ-Al2O3 and modified H-
M
ZSM-5 were used for dehydration of crude methanol. It was proved that water had more

negative effects on the performance of γ-Al2O3 than on that of ZSM-5 since ZSM-5 has
d

higher resistance due to its hydrophobic properties resulting from its high SiO2/Al2O3 ratio
te

[15]. Based on the above discussion, Al-HMS-10 appears to be a promising catalyst for the

crude methanol dehydration. Additionally, coke laid down over Al-HMS-10 catalyst used
p

for dehydration of both pure and crude methanol at 300 oC after 72 h time on stream, was
ce

measured by TG/DTA technique and results were shown in Fig. 6.


Ac

Fig. 6.

Both used catalysts showed two weight losses in TGA curves. The weight losses at < 200
o
C corresponding to endothermic peaks in DTA curves, were attributed to desorption of

12

Page 12 of 27
physically adsorbed water . The weight losses at > 400 oC, 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

t
amount of carbon deposit. As it can be seen, Al-HMS-10 possess high level of coke during

ip
dehydration of pure methanol. In other words using crude methanol instead of pure

cr
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

us
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[b7] [35].
an
M
Table 3
d
te

3.2.5. Effect of WHSV


p

The methanol conversion to DME under various WHSV (1-4 h-1) over Al-HMS-10
ce

catalyst, which showed optimum activity, was investigated and compared to commercial γ-

Al2O3. As it is depicted in Fig. 7, an increase in WHSV causes no significant reduction in


Ac

methanol conversion over Al-HMS-10, while methanol conversion over commercial γ-

Al2O3 reduces significantly. 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.

13

Page 13 of 27
Fig. 7.

4. Conclusion

t
ip
Several Al-HMS materials with Si/Al ratio of 5-35 were synthesized and studied as novel

cr
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

us
acidity strength. It was found that increasing Al content is followed by increasing

conversion of methanol and decreasing selectivity to DME. The synthesized samples

an
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
M
resistance to deactivation.
d

Acknowledgment
te

The authors gratefully acknowledge financial support from the Research Council of Shahid
p

Beheshti University and Catalyst Centre of Excellence (CCE).


ce

References
Ac

[1] T.A. Semelsberger, R.L. Borup, H.L. Greene, J. Power Sources 156 (2006) 497-511.

[2] C. Arcoumanis, C. Bae, R. Crookes, E. Kinoshita, Fuel 87 (2008) 1014-1030.

[3] C.-J. Yang, R.B. Jackson, Energy Policy 41 (2012) 878-884.

[4] T.H. Fleisch, A. Basu, R.A. Sills, J. Nat. Gas Sci. Eng. 9 (2012) 94-107.

[5] Y. Onaka, A. Miyara, K. Tsubaki, Int. J. Refrig. 33 (2010) 1277-1291.

14

Page 14 of 27
[6] M. Bjørgen, S. Akyalcin, U. Olsbye, S. Benard, S. Kolboe, S. Svelle, J. Catal. 275

(2010) 170-180.

[7] R.B. Diemer, W.L. Luyben, Ind. Eng. Chem. Res. 49 (2010) 12224-12241.

[8] F. Raoof, M. Taghizadeh, A. Eliassi, F. Yaripour, Fuel 87 (2008) 2967-2971.

t
[9] F. Yaripour, F. Baghaei, I. Schmidt, J. Perregaard, Catal. Commun. 6 (2005) 147-152.

ip
[10] S. Alamolhoda, M. Kazemeini, A. Zaherian, M.R. Zakerinasab, J. Ind. Eng. Chem. 18

cr
(2012) 2059-2068.

[11] L. Liu, W. Huang, Z.-h. Gao, L.-h. Yin, J. Ind. Eng. Chem. 18 (2012) 123-127.

us
[12] J.W. Bae, S.-H. Kang, Y.-J. Lee, K.-W. Jun, J. Ind. Eng. Chem. 15 (2009) 566-572.

[13] Y.-J. Lee, J.M. Kim, J.W. Bae, C.-H. Shin, K.-W. Jun, Fuel 88 (2009) 1915-1921.

an
[14] S.D. Kim, S.C. Baek, Y.-J. Lee, K.-W. Jun, M.J. Kim, I.S. Yoo, Appl. Catal. A: Gen.
M
309 (2006) 139-143.

[15] V. Vishwanathan, K.-W. Jun, J.-W. Kim, H.-S. Roh, Appl. Catal.A: Gen. 276 (2004)
d

251-255.
te

[16] J. Čejka, S. Mintova, Catal. Rev. 49 (2007) 457-509.

[17] K. Szczodrowski, B. Prélot, S. Lantenois, J.-M. Douillard, J. Zajac, Micropor.


p

Mesopor. Mater. 124 (2009) 84-93.


ce

[18] M.J. Gracia, E. Losada, R. Luque, J.M. Campelo, D. Luna, J.M. Marinas, A.A.

Romero, Appl. Catal. A: Gen. 349 (2008) 148-155.


Ac

[19] K.C. Tokay, T. Dogu, G. Dogu, Chem. Eng. J. 184 (2012) 278-285.

[20] T. Chiranjeevi, G.M. Kumaran, J.K. Gupta, G.M. Dhar, Thermochim. Acta 443

(2006) 87-92.

[21] T. Hamoule, M.H. Peyrovi, M. Rashidzadeh, M.R. Toosi, Catal. Commun. 16 (2011)

234-239.

15

Page 15 of 27
[22] M.H. Peyrovi, T. Hamoule, B. Sabour, M. Rashidzadeh, J. Ind. Eng. Chem. 18 (2012)

986-992.

[23] R. Mokaya, W. Jones, J. Catal. 172 (1997) 211-221.

[24] R. Mokaya, W. Jones, Chem. Commun. (1996) 981-982.

t
[25] F. Yaripour, F. Baghaei, I. Schmidt, J. Perregaard, Catal. Commun. 6 (2005) 542-549.

ip
[26] N. Khandan, M. Kazemeini, M. Aghaziarati, Appl. Catal. A: Gen. 349 (2008) 6-12.

cr
[27] A. Yin, X. Guo, W.-L. Dai, K. Fan, J. Phys. Chem. C 114 (2010) 8523-8532.

[28] A.I. Osman, J.K. Abu-Dahrieh, D.W. Rooney, S.A. Halawy, M.A. Mohamed, A.

us
Abdelkader, Appl. Catal. B: Env. 127 (2012) 307-315.

[29] Q. Tang, H. Xu, Y. Zheng, J. Wang, H. Li, J. Zhang, Appl. Catal. A: Gen. 413–414

(2012) 36-42.
an
M
[30] Y. Fu, T. Hong, J. Chen, A. Auroux, J. Shen, Thermochim. Acta 434 (2005) 22-26.

[31] A. Ciftci, D. Varisli, K. Cem Tokay, N. Aslı Sezgi, T. Dogu, Chem. Eng. J. 207–208
d

(2012) 85-93.
te

[32] S. Hassanpour, F. Yaripour, M. Taghizadeh, Fuel Process. Technol. 91 (2010) 1212-

1221.
p

[33] M. Xu, J.H. Lunsford, D.W. Goodman, A. Bhattacharyya, Appl. Catal. A: Gen. 149
ce

(1997) 289-301.

[34] A.A. Rownaghi, F. Rezaei, M. Stante, J. Hedlund, Appl. Catal. B: Env. 119–120
Ac

(2012) 56-61.

[35] K. W. Jun, H. S. Lee, H. S. Roh, S. E. Park, Bull. Korean Chem. Soc. 24(2003) 106-

108.

16

Page 16 of 27
Table 1. Physicochemical characteristics of the Al-HMS catalysts

Sample Si/Al Surface Pore Peak Acidity B/L BrÖnsted Lewis acidity
Area Volume Temprature acidity
of TPD (°C) (mmol NH3/g) (mmol NH3/g)
(m2/g) (cm3/g) (mmol NH3/g)

t
Al-HMS-35 36.3 1370 1.7 258.9 0.426 1.8 0.277 0.153

ip
Al-HMS-20 18.7 1289 2.1 270.9 0.845 1.37 0.489 0.355

cr
Al-HMS-10 9.6 1109 2.0 278.4 1.398 1.06 0.720 0.677

Al-HMS-5 4.5 870 0.79 287.1 1.556 0.86 0.719 0.837

us
an
M
d
p te
ce
Ac

Page 17 of 27
Table 2. Catalytic performance of synthesized catalysts as compared to commercial γ-Al2O3

catalyst at 300 oC.

t
Sample Methanol DME DME

ip
conversion Slectivity Yield
(%) (%) (%)

cr
γ-Al2O3 76.2 100 76.2

us
Al-HMS-10 89 100 89

Al-HMS-5 91.2 94.1 85.8

an
M
d
p te
ce
Ac

Table 3. Effect of different feed on coke formation over Al-HMS-10

Page 18 of 27
sample temperature range (oC) coke content
(wt.%)

Al-HMS-10 (used for pure methanol) 400-600 3.14

Al-HMS-10 (used for crude methanol) 445-520 0.9

t
ip
cr
us
an
M
d
p te
ce
Ac

Page 19 of 27
Figure captions:

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

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

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

ip
and WHSV=1h-1). (a) Methanol conversion profiles; (b) Activity per surface area

cr
profiles (c) DME selectivity profiles; (d) Yield of DME profiles

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

us
(C): tested for crude methanol feed)

Fig. 5. Methanol conversion variations for pure methanol and crude methanol (reaction

an
conditions were 300 oC, 0.5 g of Catalyst, reaction time=72h, and WHSV=1h-1)

Fig. 6. TG/DTA profiles of used Al-HMS-10 for dehydration of (a) pure methanol and
M
(b) crude methanol at 300oC after 72h
d

Fig. 7. Effect of WHSV on methanol conversion over Al-HMS-10 and commercial γ-


te

Al2O3 at 300 oC
p
ce
Ac

Page 20 of 27
t
251658240

ip
E:\Behrouz\WORD\Data\XRD\Al-HMS\modi\modi\1edited.jpg

cr
us
an
M
d
p te
ce
Ac

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

Page 21 of 27
t
ip
cr
us
251658240
an
M
Fig. 2. NH3-TPD on Al-HMS catalysts with different Si/Al ratio
d
p te
ce
Ac

Page 22 of 27
t
ip
cr
251658240

us
an
M
d
251658240
p te
ce
Ac

251658240

Page 23 of 27
t
ip
cr
251658240

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

us
and WHSV=1h-1). (a) Methanol conversion; (b) Activity per surface area (c) DME

selectivity; (d) Yield of DME

an
M
d
p te

251658240
ce

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

for crude methanol)


Ac

Page 24 of 27
t
ip
cr
us
an
M
d
te

251658240

Fig. 5. Methanol conversion variations for pure methanol and crude methanol (reaction
p

conditions were 300 oC, 0.5 g of Catalyst, reaction time=72h, and WHSV=1h-1)
ce
Ac

Page 25 of 27
t
ip
251658240

cr
us
251658240
an
Fig. 6. TG/DTA profiles of used Al-HMS-10 for dehydration of (a) pure methanol and (b)
M
crude methanol at 300oC after 72h
d
p te
ce
Ac

Page 26 of 27
t
ip
251658240

cr
Fig. 6. Effect of WHSV on methanol conversion over Al-HMS-10 and commercial γ-

Al2O3 at 300 oC

us
an
M
d
p te
ce
Ac

Page 27 of 27

You might also like