You are on page 1of 17

I NTERNATIONAL J OURNAL OF C HEMICAL

R EACTOR E NGINEERING
Volume 8

2010

Article A45

Dimethyl Ether Synthesis over Novel


Silicotungstic Acid Incorporated
Nanostructured Catalysts
Aysegul Ciftci

Dilek Varisli

Timur Dogu

Middle East Technical University, aysegulciftci@gmail.com


Gazi University, dilekvarisli@gazi.edu.tr

Middle East Technical University, tdogu@metu.edu.tr


ISSN 1542-6580
c
Copyright 2010
The Berkeley Electronic Press. All rights reserved.

Dimethyl Ether Synthesis over Novel Silicotungstic


Acid Incorporated Nanostructured Catalysts
Aysegul Ciftci, Dilek Varisli, and Timur Dogu

Abstract
Dimethyl ether (DME), which is an excellent green diesel-fuel alternate with
excellent clean burning properties, is synthesized by dehydration of methanol over
novel solid acid catalysts, which are synthesized following a direct hydrothermal
route and using silicotungstic acid (STA) as the active compound. These mesoporous silicate structured catalysts have surface area values of 108-393 m2 /g, depending upon their W/Si ratio. These catalysts showed very high methanol dehydration activity and also very high DME selectivity values, approaching 100%.
The STA-SiO2 mesoporous nanocomposite catalyst having a W/Si atomic ratio of
0.33 showed the highest activity, with a DME selectivity over 99% and a methanol
conversion over 60%, at 250 C and at a space time of 0.27 s.g.cm3 . Effects of
W/Si atomic ratio, calcination temperature and the synthesis procedure on the
catalytic performance of these novel mesoporous catalytic materials were investigated.
KEYWORDS: DME, methanol, dehydration, silicotungstic acid, mesoporous
catalyst

METU Research Fund and TUBITAK grants are gratefully acknowledged. We thank Assist.
Prof. Dr. Emrah Ozensoy from Bilkent University Chemistry Department for the EDX analyses,
and Assoc. Prof. Dr. Naime A. Sezgi for her contributions.

Ciftci et al.: DME Synthesis over Silicotungstic Acid-SiO2 Mesoporous Catalysts

1. INTRODUCTION
Fast increase of the rate of oil consumption and related environmental problems
necessitated the development of sustainable alternative fuels. Being clean energy
carriers, alcohols and ethers are considered as potential transportation fuel
alternates. Dimethyl ether (DME) is an attractive transportation fuel substitute for
compression ignition engines, in the sense that it has higher cetane number (5560) than diesel fuel (40-55) and it does not produce black smoke. NOx emissions
of DME derived diesel engines are also very low. DME is considered as a nontoxic and environmentally benign fuel, which can be produced from nonpetroleum feedstocks (Dogu and Varisli, 2007; Olah et al., 2006).
DME can be synthesized by dehydration of methanol over solid acid
catalysts. Methanol dehydration activity of different acidic catalysts, such as Al2O3 (Yaripour et al, 2005; Raoof et al., 2008; Jun et al., 2002; Tokay, 2008), HZSM-5 (Fu et al., 2005), mesoporous aluminosilicates (Tokay, 2008; Varisli et al.,
2009) and Nafion-silica nanocomposites (Ciftci et al., 2010) were tested in the
literature. These catalysts were reported to show activity at different temperature
ranges. -alumina and aluminosilicate type catalysts were reported to show good
catalytic performance in a temperature range between 300-370oC. However, there
are also some results reported at higher (up to 400oC) and lower temperatures.
Our recent work with Nafion based nano-composite catalysts (Ciftci et al., 2010)
showed good methanol dehydration activity at temperatures lower than 300oC.
Green and sustainable aspects of heteropolyacid (HPA) catalysts are mentioned in
the work of Misono (2000), by referring to some of their features, such as noncorrosiveness, no-waste production and pseudo-liquid phase behavior. The most
important property of HPAs is their high acidity, due to the high concentration of
Brnsted acid sites in their structure (Okuhara et al., 1996). It was shown in our
recent study that (Varisli et al., 2007); activity of silicotungstic acid (STA) was
much higher than the activities of tungstophosphoric acid and molybdophosphoric
acid, in ethanol dehydration reaction to produce ethylene. Their very low surface
area values (1-5 m2/g) and high solubility in polar solvents limit the wide use of
HPAs as catalysts.
High surface area mesoporous materials are quite suitable for
immobilization of active species in their framework. Development of silicate
structured mesoporous solids (MCM-41, MCM-48, SBA-15 etc.) opened a new
area in catalysis research (Taguchi and Schth, 2005). Metals or metal oxides are
generally incorporated into such mesoporous materials to increase their catalytic
performance (Sener et al., 2006; Nalbant et al., 2008; Ozdogan et al., 2008). Said
et al. (2007) and Vazquez et al. (2000) synthesized HPA/SiO2 catalysts by
impregnation. A one-pot procedure was developed in our recent work (Varisli et
al., 2009), for the synthesis of STA incorporated silicate structured mesoporous

Published by The Berkeley Electronic Press, 2010

International Journal of Chemical Reactor Engineering

Vol. 8 [2010], Article A45

catalysts. These catalysts and STA impregnated MCM-41 showed excellent


catalytic performance in dehydration of ethanol (Varisli et al., 2008, 2009) to
produce ethylene. In the present study, STA incorporated silicate structured
catalysts were synthesized following a modified one-pot hydrothermal route and
the catalysts containing different W/Si molar ratios in their structure were tested
in the dehydration reaction of methanol to produce DME.
2. EXPERIMENTAL
2.1. Catalyst Synthesis and Characterization
In this study, novel silicotungstic acid (STA) incorporated nano-composite
catalysts were synthesized following a one-pot hydrothermal synthesis route. STA
incorporated silicate structured mesoporous catalysts TRC-62(L), TRC-82(L) and
TRC-92(L), containing different W/Si atomic ratios, were synthesized in our
earlier study (Varisli et al., 2009, 2010). Following a modified procedure, new
catalysts with a W/Si atomic ratio of 0.40, were synthesized in the present study.
In the synthesis of this material, cetyltrimethylammonium bromide was used as
the surfactant and TEOS (tetraethylorthosilicate) was used as the silica source.
Predetermined amount of STA was dissolved in deionized water and added to the
solution of TEOS and the surfactant. After mixing for 1 hour, the pH of the
mixture was measured as about 1.0 and this solution was transferred into a
Teflon-lined stainless steel autoclave, where it was kept at 120C for 96 hours.
Hydrothermal synthesis was carried out under autogenous pressure in the fully
sealed autoclave. The resulting material was washed successively with deionized
water, until the pH of the washing liquid became constant. The solid product was
dried at 40C and calcined in a tubular reactor in a continuous flow of dry air, at
350C (denoted as TRC-75(L)) or at 400C (denoted as TRC-75-400) for 8 hours.
It was shown in our earlier study of STA incorporated mesoporous catalysts
(TRC-62(L), TRC-82(L) and TRC-92(L)) that, calcination temperature had very
important effects on the structure and on the catalytic performance of these
materials. Significant loss of Brnsted acidity and consequently the activity was
reported at calcination temperatures higher than 400oC.
As an alternative technique, in the present work supercritical carbon
dioxide extraction was used for surfactant removal from the material, which was
synthesized by the one-pot hydrothermal route. Extraction was carried out at
100C, 350 bar and with a CO2 flow rate of 1mL/min, for 3 hours. Considering
that the maximum reaction temperature of the catalytic performance tests was
350C, the synthesized material was heated up to this temperature in the presence
of dry air, after the supercritical CO2 extraction step. This material was denoted as
TRC-75(L)-CO2.

http://www.bepress.com/ijcre/vol8/A45

Ciftci et al.: DME Synthesis over Silicotungstic Acid-SiO2 Mesoporous Catalysts

Scanning Electron Microscopy (SEM) analyses of the synthesized


catalysts were performed in Middle East Technical University (METU) Central
Laboratory, by a Quanta 400F Field Emission SEM instrument, and Energy
Dispersive Spectroscopy (EDS) analyses were done in METU Metallurgical and
Materials Engineering Department by a JSM-6400 (JEOL) instrument equipped
with NORAN System. Energy Dispersive X-Ray (EDX) elemental mapping
analyses were performed in the Bilkent University Chemistry Department. SEM
and EDX data were collected using a Zeiss EVO40 environmental SEM, that is
equipped with a LaB6 electron gun, a vacuum SE detector, an elevated pressure
SE detector, a backscattering electron detector (BSD), and a Bruker AXS XFlash
4010 detector. In order to get information about the relative intensities of
Brnsted and Lewis acid sites of the synthesized catalysts, diffuse reflectance FTIR spectroscopy (DRIFTS) analyses of pyridine adsorbed samples were
performed by using a Perkin Elmer Spectrum One instrument. X-Ray Diffraction
(XRD) analyses were made using the Rigaku D/MAX2200 diffractometer with a
CuK radiation source, within a 2 scanning range between 1-50. Fourier
transform infrared (FT-IR) spectroscopy analyses of the STA-silicate structured
catalysts in the present work were performed by using a Bruker FTIR-IFS66/S
instrument. Nitrogen adsorption-desorption analyses were carried out by a
Quantachrome Autosorb-1-C/MS instrument, in the METU Central Laboratory.
Multipoint BET surface area values, pore size distributions and pore volumes of
the samples were analyzed by this characterization technique. XPS analysis of the
used and the fresh catalysts was made with a SPECS instrument, also in the
central laboratory of Middle East Technical University. Supercritical CO2
extraction was carried out with a SFX 3560 extractor. The synthesized STA
incorporated nano-composite materials were also characterized by a ScanningTransition Electron Microscopy (STEM) and Energy-Filtered Transmission
Microscopy (EFTEM) analyses, by a Jeol 2100F instrument attached with Jeol
EDX & HAADF detectors and a GIF Tridem STEM Pack filter.
2.2. Methanol dehydration reactions
Methanol dehydration reactions were carried out in a flow system between 180350C, with the TRC-62(L), TRC-75(L), TRC-82(L) and TRC-92(L) catalysts.
Catalyst was placed into the stainless steel tubular reactor and heated up to the
reaction temperature under continuous flow of helium. Methanol was first fed to
an evaporator at a flow rate of 2.1mL/hr, by using a syringe pump. It was mixed
with He gas in the evaporator to adjust the composition of the reactor feed stream.
The volume fraction of methanol was adjusted as 0.48 in this stream, in the
experiments carried out in the present study. Online analysis of the product stream
was performed with a Varian CP 3800 gas chromatograph, equipped with a

Published by The Berkeley Electronic Press, 2010

International Journal of Chemical Reactor Engineering

Vol. 8 [2010], Article A45

Porapak T column. All the lines were heated to 150C, in order to avoid any
condensation of reactants or products. Catalytic activity test experiments were
carried out at different space times (at 0.14, 0.27 and 0.41 s.g.cm-3) by changing
the amount of catalyst placed into the tubular reactor (0.1-0.3 g).
3. RESULTS AND DISCUSSION
3.1. Catalyst Characterization Results
The mesoporous catalytic materials synthesized in this work are composed of a
porous silicate network, with STA being well dispersed in the structure as the
active acidic component. Characterization results of TRC-62(L), TRC-82(L) and
TRC-92(L) were reported elsewhere (Varisli et al., 2009, 2010). STEM and XRD
analysis had indicated the presence of WOx nanorods dispersed within the
mesoporous silicate matrix of the catalysts synthesized in our earlier study (Fig.
1). However, XRD analysis of the catalysts, which were synthesized in the
present work in a sealed autoclave, did not show any sharp peaks corresponding
to large WOx crystals (Fig. 2). This XRD spectrum indicated well dispersion of
tungsten in the silicate lattice of TRC-75(L) catalyst. This XRD spectrum is
different from the XRD spectra of former catalysts (TRC-62(L), TRC-82(L) and
TRC-92(L)), for which XRD analysis had indicated sharp peaks corresponding to
W20O58. This difference is thought to be due to the modification of the
hydrothermal synthesis procedure used in the present study. In this work,
synthesis of the new catalysts was achieved in a fully sealed autoclave, while in
the previous synthesis procedure; hydrothermal synthesis had been achieved in an
open vessel at atmospheric pressure.
300
Intensity (counts)

250
200
150
100
50
0
0

Figure 1. TEM image of TRC-82(L).

10

20
30
2-tetha (deg)

40

50

Figure 2. XRD pattern of TRC-75(L).

http://www.bepress.com/ijcre/vol8/A45

Ciftci et al.: DME Synthesis over Silicotungstic Acid-SiO2 Mesoporous Catalysts

Physical properties of the catalysts used in this study in DME synthesis are
given in Table 1. EDS analysis results indicated that STA was successfully
incorporated into the structure of TRC-75(L). From the elemental mapping
analysis of the sample (Fig. 3) it can be concluded that W and Si are very well
dispersed in the nano-structured material. According to the results obtained from
nitrogen physisorption analyses, pore volumes were in the range 0.47-0.37 cm3/g
and average pore diameters were 7.8 nm, for the TRC-75 type catalytic materials
(Table 1).
Table 1. Physical properties of STA incorporated silicate structured catalysts
Catalyst
TRC-75(L)
TRC-75(L)-CO2
TRC-75-400
TRC-62(L) (*)
TRC-82(L) (*)
TRC-92(L) (*)

W/Si
atomic
EDS
0.33
0.33
0.16
0.47
0.78

W/Si
atomic
Soln.
0.40
0.40
0.40
0.25
0.50
1.00

Multipoint BET
Surface Area
(m2/g)
252
187
241
393
179
108

Pore
volume
(cm3/g)
0.37
0.32
0.47
0.55
0.45
0.21

Avg pore
diameter
(nm)
7.8
7.8
7.8
5.5
10.0
7.8

(*) Adapted from Varisli et al. (2009).

Figure 3. EDX mapping of TRC-75(L).


Nitrogen adsorption-desorption isotherms (Fig. 4) are Type IV, implying a
mesoporous structure. Physical properties of the supercritical CO2 treated catalyst
(TRC-75(L)-CO2) were quite close to the physical properties of TRC-75(L).

Published by The Berkeley Electronic Press, 2010

Quantity Adsorbed [cc/g STP]

International Journal of Chemical Reactor Engineering

300

Adsorption

Vol. 8 [2010], Article A45

Desorption

250
200
150
100
50
0
0.0

0.2

0.4

0.6

0.8

Relative Pressure (P/Po)

1.0

Figure 4. Nitrogen adsorption-desorption isotherm of TRC-75(L).


STA incorporated catalysts synthesized in this work were calcined at
350C. At temperatures above 400C, protons of STA in the nano-composite
materials were reported to be lost (Varisli et al., 2009). Calcination of the
synthesized material at 400C caused an increase in the pore volume (Table 1).
Also, the pore size distribution of TRC-75-400 was broader than the pore size
distributions of TRC-75(L) and TRC-75(L)-CO2 (Fig. 5).

dV/dlog D (cc/g)

2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0

TRC-75(L)
TRC-75-400
TRC-75(L)-CO2

10

100
Pore Diameter ()

1000

Figure 5. Pore size distributions of TRC-75(L), TRC-75(L)-CO2 and TRC-75400.


Morphologies of the samples were observed by scanning electron
microscopy (SEM) images. SEM image of TRC-75(L) is given in Fig. 6. SEM
image of TRC-75-400 (Fig. 7) showed the formation of rod-like structures on the

http://www.bepress.com/ijcre/vol8/A45

Ciftci et al.: DME Synthesis over Silicotungstic Acid-SiO2 Mesoporous Catalysts

surface of the catalyst. Such a structure was also observed in our earlier work
(Varisli et al., 2009) for a STA/silica catalyst containing a W/Si ratio of 0.47.
Results of FT-IR analysis of pure STA showed characteristic IR peaks at
780 cm-1 (W-O-W), 876 cm-1 (W-Ocorner-W), 921 cm-1 (Si-O) and 977 cm-1
(W=O) (Fig. 8). FT-IR spectra of TRC-75(L) and TRC-75(L)-CO2 also showed
these bands (Fig. 8) at the same wave numbers. Although some decrease of
intensities of these bands was observed, they were still quite sharp in the FT-IR
spectra of the STA incorporated mesoporous materials. This decrease of intensity
of the FT-IR bands is essentially due to the low STA content of the synthesized
materials. Especially the band corresponding to W-Ocorner-W was quite weak in
the synthesized materials. In the case of TRC-75-400, which was calcined at
400oC, the bands corresponding to the characteristic STA structure were not as
sharp as the corresponding bands observed in the spectra of TRC-75(L) and TRC75(L)-CO2. This result is an indication of some distortion of the STA structure
during calcination at 400oC. Similar distortions were observed in our earlier
studies (Varisli et al., 2009, 2010) for TRC-62, TRC-82 and TRC-92 type
catalytic materials, which were calcined at temperatures higher than 350oC. FTIR results obtained in this study indicated that the STA structure was not distorted
within the lattice of the synthesized material which was calcined at 350oC and/or
treated with supercritical CO2.

Figure 6. SEM image of TRC-75(L).

Figure 7. SEM image of TRC-75-400.

Published by The Berkeley Electronic Press, 2010

International Journal of Chemical Reactor Engineering

TRC-75(L)
TRC-75-400

Vol. 8 [2010], Article A45

TRC-75(L)-CO2
STA

Transmittance

1450

1300

1150 1000 850


Wavenumber (cm-1)

700

550

Figure 8. FTIR spectra of TRC-75(L), TRC-75-400, TRC-75(L)-CO2 catalysts;


and pure STA adapted from Varisli et al. (2009).
DRIFTS analysis of the pyridine adsorbed samples gave IR absorption
bands at 1539, 1488 and 1445 cm-1 (Fig. 9). The first band at 1539 cm-1
corresponds to pyridinium ion and shows the presence of Brnsted acid sites on
the catalyst surface. The third band is associated with the adsorbed molecules on
the Lewis acid sites (Damyanova et al., 1999). The second band corresponds to
another form of adsorbed pyridine molecules. For TRC-75(L), all of these bands
were observed. However it is clear that Brnsted acid sites were stronger than
Lewis acid sites. As stated by Herrera et al. (2008) and also as reported in our
earlier publications (Ciftci et al., 2010; Varisli et al., 2009), presence of Brnsted
acid sites is the major indication of the activity of the catalysts in alcohol
dehydration reactions. As originally proposed by Bandiera and Naccache (1991)
and later supported in the theoretical analysis of Blaszkowski and Van Santen
(1996), DME formation on zeolite type catalytic materials was expected to take
place by the reaction of the [CH3.OH2]+ and [CH3O]- surface species, which were
adsorbed on the Brnsted acid and on the adjacent Lewis basic sites. Besides the
Brnsted acidity, surface area is also expected to contribute to the activity of such
solid acid catalysts in dehydration of alcohols. As shown in Figure 9, the
Brnsted acidity of the material prepared by the supercritical CO2 extraction
process (TRC-75(L)-CO2) was even stronger than the Brnsted acidity of TRC75(L).
The results of the characterization studies indicated that, by the synthesis
of the nanocomposite catalysts following a one-pot hydrothermal route, catalytic
properties of STA were enhanced by increasing its surface area to around 200

http://www.bepress.com/ijcre/vol8/A45

Ciftci et al.: DME Synthesis over Silicotungstic Acid-SiO2 Mesoporous Catalysts

m2/g. Acidic characteristics of STA were preserved. Also, these catalysts did not
lose their acidic character even after successive washing steps.
TRC-75(L)

Absorbance

TRC-75(L)-CO2

1400

1450

1500
1550
Wavenumber (cm-1)

1600

Figure 9. DRIFT results of pyridine adsorbed samples


3.2. Methanol Dehydration Results
Methanol dehydration reactions were carried out in vapour phase. The observed
products were dimethyl ether and formaldehyde, which were produced according
to the following dehydration and dehydrogenation reactions:
2CH 3 OH
CH 3 OCH 3 + H 2 O

(1)

CH 3 OH
CH 2 O + H 2

(2)

Typical methanol conversion data obtained with the catalysts containing


different W/Si ratios are shown in Fig. 10. Results reported in this figure
correspond to the average of at least three data points obtained during the first
hour of operation at a given temperature. Temperature dependence of DME
selectivity values obtained with these catalysts are given in Fig. 11. The main
superiority of the new STA incorporated mesoporous catalysts synthesized in this
work over the conventional alumina catalysts is their very high DME selectivity
(approaching to 100%) and quite high activity at relatively low temperatures.
These catalysts showed methanol dehydration activity at temperatures as low as
180oC. The catalyst containing a W/Si ratio of 0.16 gave the lowest activity in
methanol dehydration in a temperature range of 180-350C. The lowest activity of
TRC-62(L) is due to its less STA content. In fact, the Brnsted acidity of this
material was reported to be less than the Brnsted acidities of TRC-82(L) and
TRC-92(L) (Varisli et al., 2009). The highest DME yield obtained with the STA

Published by The Berkeley Electronic Press, 2010

International Journal of Chemical Reactor Engineering

10

Vol. 8 [2010], Article A45

1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0

TRC-62(L)
TRC-75(L)

180

230
280
Temperature (C)

DME Selectivity

MeOH Conversion

incorporated nano-structured catalysts was about 0.60, at a space time of 0.27


s.g.cm-3 and at 250C. TRC-82(L), TRC-92(L) and TRC-75(L) exhibited similar
trends in dehydration of methanol in the temperature range of 180-350C. As far
as DME selectivity was concerned, TRC-75(L) showed the best performance,
approaching to 100%. Relative intensity of the band (at 1539 cm-1) corresponding
to Brnsted acidity of TRC-75(L) (Fig. 9) was even higher than the relative
intensities of the bands corresponding to Brnsted acidities of TRC-82(L) and
TRC-92(L), which were reported in our earlier publications (Varisli et al., 2009,
2010). These results supported the very high DME selectivity values obtained
with this catalyst.

330

Figure 10. Conversion of methanol


obtained with hydrothermally
synthesized catalysts containing
different W/Si ratios at 0.27 s.g.cm-3

1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0

TRC-62(L)
TRC-82(L)
180

TRC-75(L)
TRC-92(L)

230
280
330
Temperature (C)

Figure 11. DME selectivity obtained


with hydrothermally synthesized
catalysts containing different W/Si
ratios at 0.27 s.g.cm-3

Surface area of the STA incorporated mesoporous catalysts decreased with


an increase in the STA content (Table 1) (Varisli et al., 2009). Decrease of surface
area and increase of Brnsted acidity have opposing effects on the activity of STA
incorporated catalytic materials synthesized in this work. Although it contains a
less W/Si ratio (0.33), TRC-75(L) is as active as TRC-82(L) and TRC-92(L),
which have W/Si ratios of 0.47 and 0.78, respectively. Better catalytic
performance of TRC-75(L) can be attributed to the well dispersion of the active
phase (STA) within the catalyst. Best operation temperature is between 200250C for this catalyst. Decrease of conversion at higher temperatures indicated
deactivation by coke formation. Some experiments were repeated for periods
longer than 6 hours at high temperatures. XPS analyses of the fresh and the used

http://www.bepress.com/ijcre/vol8/A45

Ciftci et al.: DME Synthesis over Silicotungstic Acid-SiO2 Mesoporous Catalysts

11

DME Yield

DME Yield

catalysts in these experiments


1.0
0.9
clearly indicated the formation of
0.8
coke on the catalyst surface. C/Si
0.7
atomic ratio was about 1.5 for the
0.6
0.5
used catalysts in these experiments.
0.4
Some activity decrease was
0.3
0.14
observed at long test periods,
0.2
0.27
0.1
especially at reaction temperatures
0.41
0.0
higher than of 250C. However, the
180
230
280
330
catalyst was easily reactivated to its
Temperature
(C)
original activity by its recalcination at 350C in the presence
of dry air. The tolerance of TRC- Figure 12. DME yields at different
-3
75(L) to coke deposition is space times (0.14, 0.27, 0.41 s.g.cm )
apparently higher than TRC-82(L) with TRC-75(L)
and TRC-92(L) (Fig. 10).
1.0
As it is expected, DME
0.9
yield values increased with an
0.8
0.7
increase in space time (80% at a
-3
0.6
space time of 0.41 s.g.cm ), at
0.5
about 250C (Fig. 12). At this
0.4
TRC-75(L)
temperature, DME selectivity was
0.3
TRC-75-400
0.2
100%.
0.1
TRC-75(L)-CO2
DME yield values obtained
0.0
with the catalysts containing a W/Si
180
230
280
330
ratio of 0.4 (in the solution) but
Temperature (C)
calcined at different conditions or
treated with supercritical CO2 are Figure 13. DME yields obtained with
shown in Fig. 13. The catalyst TRC-75(L), TRC-75-400 and TRCprepared
by
removing
the 75(L)-CO2
surfactant from the structure using
supercritical CO2 (TRC-75(L)-CO2) appears to be as active as TRC-75(L).
DRIFTS analysis of pyridine adsorbed materials had indicated that the strength of
the Brnsted acid sites of TRC-75(L)-CO2 was somewhat higher than the
Brnsted acidity of TRC-75(L). However, the surface area of this material was
lower than TRC-75(L) (Table 1). Apparently, disadvantage of lower surface area
was compensated by higher Brnsted acidity (Fig. 9), so that the catalytic
performance of TRC-75(L)-CO2 was quite close to the performance of TRC75(L). Calcination at 400C caused some decrease of the activity, indicating some
loss of protons of the catalyst. In fact, the FT-IR analysis results reported in

Published by The Berkeley Electronic Press, 2010

12

International Journal of Chemical Reactor Engineering

Vol. 8 [2010], Article A45

Figure 8 had also indicated some distortions of the STA structure when the
synthesized material was calcined at 400oC.
DME selectivities of TRC-75(L) and TRC-75(L)-CO2 were close,
approaching to 100 %. Formaldehyde forms at lower temperatures. Formaldehyde
selectivity of TRC-62(L) was the highest, approaching to 56 % at 180C. Trace
amount of ethylene formation was also observed with these catalysts at high
temperatures.
4. CONCLUSIONS
It was shown that, dimethyl ether, which is considered as a green transportation
fuel alternate, can be produced at very high yields, using the novel STA
incorporated mesoporous silicate structured catalysts synthesized in this work.
Especially, the catalysts containing a W/Si ratio of 0.4 in the synthesis solution
(TRC-75(L)) gave very high DME selectivity values approaching to 100 %. Best
operating temperature of methanol dehydration was between 200-250C, with the
new STA incorporated mesoporous catalysts. Low temperature activity and very
high DME selectivity are important superiorities of these new catalytic materials.
Formaldehyde formation was observed as the main by-product at lower
temperatures. Results indicated that the procedure used during the hydrothermal
synthesis and the calcination temperature of the synthesized materials are quite
important, as far as their catalytic performances were concerned. Supercritical
CO2 extraction caused no defect in the catalyst structure or activity loss. Increase
of calcination temperature from 350C to 400C caused some decrease in DME
yield, which was considered to be due to the loss of some Brnsted acidity and
partial deformation of the STA structure. The results obtained in this study
showed the possibility of synthesizing this non-petroleum fuel alternate (DME) at
very high yields starting from methanol and using the nanocomposite STAsilicate structured mesoporous catalysts synthesized here.

http://www.bepress.com/ijcre/vol8/A45

Ciftci et al.: DME Synthesis over Silicotungstic Acid-SiO2 Mesoporous Catalysts

13

REFERENCES
Bandiera J. and C. Naccache, Kinetics of Methanol Dehydration on
Dealuminated H-mordenite- Model with Acid and Basic Active Centers,
Appl. Catal., 1991, 69, 139-148.
Blaszkowski S. R. and R. A. Van Santen, The Mechanism of Dimethyl Ether
Formation from Methanol Catalyzed by Zeolite Protons, J. Am. Chem.
Soc. 1996, 118, 5152-5153.
Ciftci A., Sezgi N. A. and T. Dogu, Nafion-Incorporated Silicate Structured
Nanocomposite Mesoporous Catalysts for Dimethyl Ether Synthesis, Ind.
Eng. Chem. Res., 2010, (DOI: 10.1021/ie9015667) article in press.
Damyanova S., Fierro J.L., Sobrados I. and J. Sanz, Surface Behavior of
Supported 12-Heteropoly Acid As Revealed by Nuclear Magnetic
Resonance, X-ray Photoelectron Spectroscopy, and Fourier Transform
Infrared Techniques, Langmuir, 1999, 15(2), 469-476.
Dogu T. and D. Varisli, Alcohols as Alternatives to Petroleum for
Environmentally Clean Fuels and Petrochemicals, Turk. J. Chem., 2007,
31, 551-567.
Fu Y., Hong T., Chen J., Auroux A. and J. Shen, Surface Acidity and the
Dehydration of Methanol to Dimethyl Ether, 2005, Thermochim. Acta,
434, 22-26.
Herrera J.E., Kwak J.H., Hu J.Z., Wang Y. and C. H. F. Peden, Effects of Novel
Supports on the Physical and Catalytic Properties of Tungstophosphoric
Acid for Alcohol Dehydration Reactions, Topics on Catal., 2008, 49,
259267.
Jun K. W., Lee H. S., Roh H. S. and S. E. Park, Catalytic Dehydration of
Methanol to Dimethyl Ether (DME) over Solid-Acid Catalysts, Bull.
Korean Chem. Soc., 2002, 23, 803-806.
Misono M., Acid Catalysts for Clean Production. Green Aspects of
Heteropolyacid Catalysts, C. R. Acad. Sci. Paris, Srie Ilc,
Chimie/Chemistry, 2000, 3, 471475.

Published by The Berkeley Electronic Press, 2010

14

International Journal of Chemical Reactor Engineering

Vol. 8 [2010], Article A45

Nalbant A., Dogu T. and S. Balci, Ni and Cu Incorporated Mesoporous


Nanocomposite Catalytic Materials, J. Nanoscience Nanotechnology,
2008, 8, 549-556.
Okuhara T., Nishimura T. and M. Misono, Novel Microporous Solid
Superacids, Studies in Surface Science and Catalysis, 1996, 101, 581590.
Olah G.A., Goeppert A. and G. K. S. Prakash, Beyond Oil and Gas: The
Methanol Economy, 2006, Wiley-VCH, Weinheim.
Ozdogan E., Dogu T. and G. Dogu, Ni-MCM-41 Type Mesoporous Catalysts
Synthesized by One-pot Hydrothermal Procedure for Steam Reforming of
Ethanol, Int. J. Chem. Reactor Eng., 2008, 5, A111.
Raoof F., Taghizadeh M., Eliassi A. and F. Yaripour, Effects of Temperature and
Feed Composition on Catalytic Dehydration of Methanol to Dimethyl
Ether over -Alumina, Fuel, 2008, 87, 2967-2971.
Said A. E. A, El-Wahab M.M.M.A. and A. M. Alian, Perspective Catalytic
Performance of Brnsted Acid Sites During Esterification of Acetic Acid
with Ethyl Alcohol over Phosphotungestic Acid Supported on Silica, J.
Chem. Technol. Biotechnol., 2007, 82, 513523.
Sener C., Dogu T. and G. Dogu, Effects of Synthesis Conditions on the Structure
of Pd Incorporated MCM-41 Type Mesoporous Nanocomposite Catalytic
Materials, Microporous Mesoporous Mater., 2006, 94, 89-98.
Taguchi A. and F. Schth, Ordered Mesoporous Materials in Catalysis,
Microporous Mesoporous Mater., 2005, 77, 1-45.
Tokay K. C., Dimethyl Ether (DME) Synthesis over Novel Mesoporous
Catalysts, M.S. Thesis, Middle East Technical University, August 2008,
Ankara Turkey.
Varisli D., Dogu T. and G. Dogu, Ethylene and Diethyl-ether Production by
Dehydration Reaction of Ethanol over Different Heteropolyacid
Catalysts, Chem. Eng. Sci., 2007, 62, 5349-5352.

http://www.bepress.com/ijcre/vol8/A45

Ciftci et al.: DME Synthesis over Silicotungstic Acid-SiO2 Mesoporous Catalysts

15

Varisli D., Dogu T. and G. Dogu, Silicotungstic Acid Impregnated MCM-41-like


Mesoporous Solid Acid Catalysts for Dehydration of Ethanol, Ind. Eng.
Chem. Res., 2008, 48, 4071-4076.
Varisli D., Dogu T. and G. Dogu, Novel Mesoporous Nanocomposite WOxSilicate Acidic Catalysts: Ethylene and Diethylether from Ethanol, Ind.
Eng. Chem. Res., 2009, 48, 9394-9401.
Varisli D., Dogu T. and G. Dogu, Petrochemicals from Ethanol over a W-Si
based Nanocomposite Bidisperse Solid Acid Catalyst, Chem. Eng. Sci.,
2010, 65, 153-159.
Varisli D., Tokay K. C., Ciftci A., Dogu T. and G. Dogu, Methanol Dehydration
to Produce Clean Diesel Alternative Dimethyl Ether over Mesoporous
Aluminosilicate-Based Catalysts, Turk. J. Chem., 2009, 33, 355-366.
Vazquez P., Pizzio L., Caceres C., Blanco M., Thomas H., Alesso E.,
Finkielsztein L., Lantano B., Moltrasio G. and J. Aguirre, Silica
Supported Heteropolyacids as Catalysts in Alcohol Dehydration
Reactions, J. Mol. Catal. A:Chem., 2000, 161, 223232.
Yaripour F., Baghaei F., Schmidt I. and J. Perregaard, Catalytic Dehydration of
Methanol to Dimethyl Ether over Solid Acid Catalysts, Catal. Commun.,
2005, 6, 147-152.

Published by The Berkeley Electronic Press, 2010

You might also like